Sei sulla pagina 1di 325

This article was downloaded by: 10.3.98.

93
On: 17 Sep 2018
Access details: subscription number
Publisher:Routledge
Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered office: 5 Howick Place, London SW1P 1WG, UK

Handbook of Induction Heating

Valery Rudnev, Don Loveless, Raymond L. Cook

Heat Treatment by Induction

Publication details
https://www.routledgehandbooks.com/doi/10.1201/9781315117485-4
Valery Rudnev, Don Loveless, Raymond L. Cook
Published online on: 11 Jul 2017

How to cite :- Valery Rudnev, Don Loveless, Raymond L. Cook. 11 Jul 2017 ,Heat Treatment by
Induction from: Handbook of Induction Heating Routledge.
Accessed on: 17 Sep 2018
https://www.routledgehandbooks.com/doi/10.1201/9781315117485-4

PLEASE SCROLL DOWN FOR DOCUMENT

Full terms and conditions of use: https://www.routledgehandbooks.com/legal-notices/terms.

This Document PDF may be used for research, teaching and private study purposes. Any substantial or systematic reproductions,
re-distribution, re-selling, loan or sub-licensing, systematic supply or distribution in any form to anyone is expressly forbidden.

The publisher does not give any warranty express or implied or make any representation that the contents will be complete or
accurate or up to date. The publisher shall not be liable for an loss, actions, claims, proceedings, demand or costs or damages
whatsoever or howsoever caused arising directly or indirectly in connection with or arising out of the use of this material.
4
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

Heat Treatment by Induction

This chapter is devoted to the study of the metallurgical aspects, design approaches, the
selection of process parameters, equipment specifications, and physical phenomena that
are imperative to understand the manufacturing and operation of induction heat treating
(IHT) machinery.
Although the basic principles of heat treatment may be applicable to many different
types of metallic materials, the primary focus in this chapter will be with respect to steels,
cast irons, and powder metallurgy ferrous materials since these are, by far, the area of
greatest interest in the field of IHT.
We recognize that obtaining needed steel microstructures and improving certain indus­
trial properties via heat treatment is only one of several possible processes. Other methods
to alter structures and properties include, but are not limited to, the use of electrical fields
(for example, an application of pulsed electrical currents—work of R.S. Qin, A. Rahnama,
and others), magnetic field processing and the use of effect of ultrastrong magnetic fields on
phase transformation (for example, the work of G.M. Ludtka and others), and an application
of mechanical stresses (for example, work hardening in spring wire manufacturing industry),
just to name a few. Recognizing the presence of a number of existed processes and technolo­
gies under development, this chapter solely concentrates on IHT.
Before discussing the features of IHT, it is essential for readers to refresh their knowl­
edge of the basic metallurgical principles associated with the heat treatment of steels and
cast irons. Because of space limitations, only a brief introduction to the principles of heat
treatment are provided here, emphasizing the specifics of heat treating by induction. A
thorough discussion on heat treatment of metals and metallic alloys can be found in clas­
sical books written by E. Bain, M. Grossmann, J. Hollomon, L. Jaffe, G. Krauss, Ch. Brooks,
R.W.K. Honeycombe, H.K.D.H. Bhadeshia, G. Kurdyumov, K. Thelning, A. Gulyaev,
W. Smith, and others [2,28–30,33,144–171].

4.1 The Basics of Metallurgy and Principles of Heat


Treatment of Steels and Cast Irons
Metallurgy as an art has been practiced since the beginning of the history of mankind, but
as a science, it traces its origin to the early 1860s when light optical microscopy began to
be used to inspect the structure of metals and alloys. Metallurgy is a broad term that can
be defined as a domain of science that deals with the process of extracting metals from
the ores in which they are found, followed by its refining, alloying, developing desirable
structures, obtaining needed properties, and fabricating useful objects.
Physical metallurgy (which is a part of the science of metallurgy) deals with the physi­
cal, chemical, and mechanical characteristics of metals, various intermetallic compounds,
and mixtures referred to as alloys. It also focuses on the effects of chemical composition,
microstructure, metal working, thermal treatment, and some other factors affecting the
139
140 Handbook of Induction Heating

desired properties such as hardness, strength, ductility, toughness, corrosion resistance,


wear resistance and many others. Thermal treatment (also referred to as heat treatment) is
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

related to the effect of temperature, the rate of heating, holding time at elevated tempera­
ture, and cooling intensity of the material in order to arrive at a specific microstructure
and desired industrial characteristics.

4.1.1 Crystalline Structures and Critical Temperatures


Although our understanding of the matter is improving, regularly unveiling its subtleties,
this section provides only a simplified introduction regarding the structure of metallic
materials and critical temperatures.
An element is a pure substance or material that cannot be separated chemically or by
any other means to provide any other type of substance. There are more than 100 known
elements recognized in the Periodic Table of the Elements (see Appendix). The elements
are arranged in the table by increasing atomic weight and tend to be grouped accord­
ing to physical and chemical properties. These elements are composed of atoms that are
arranged in a specific combination to produce the element. If the atoms are separated, they
no longer maintain the characteristic properties of the element.
At room temperature, most elements are solid; however, some are liquids (i.e., mercury)
and some are gases (i.e., oxygen, nitrogen). Some elements belong to a group of metals that
exhibit certain metallic properties. These include but are not limited to being electrically
conductive, relatively hard, and not transparent [49].
Atoms are composed of a solid nucleus containing neutrons and protons. Around the
nucleus, there are various numbers of electrons circulating in specific orbits determined
by their energy level. An example would be a helium atom, which has two orbiting elec­
trons about its nucleus. The electrons carry a negative electrical charge that is balanced
by the protons in the nucleus that carry a positive electrical charge. A neutral atom would
contain an equal number of electrons and protons.
Two or more elements can combine to form a molecule. For example, in water (chemical
formula, H2O), two hydrogen atoms combine with a single oxygen atom to form a molecule
of water.
Chemically bonded elements can produce different compounds. A compound often has
quite different properties and even phase compared to the elements of which it is consti­
tuted (e.g., H2O or NaCl). In the example of water, which is composed of two gases, hydro­
gen and oxygen form the liquid phase of water at room temperature.
Combining or mixing two or more elements or compounds that are not chemically
joined together form a mixture. In contrast to a compound, a mixture can be relatively eas­
ily separated into its components. The characteristics of a mixture typically have similari­
ties with the characteristics of the components that it is made from.
A solution is a special kind of mixture in which one substance is dissolved in another.
In a solution, the material that is dissolved is referred to as the solute. The material in
which another material is dissolved is referred to as the solvent. It is possible to have a
mixture of two materials that are solid. A solid solution is a solution in which the solute(s)
and solvent are solids. Elevated or high temperature is usually required to form a solid
solution. This is the case with materials commonly referred to as alloys that may involve
the mixing of elements such as copper and tin to form the alloy called bronze or copper
and zinc to form brass.
When a group of atoms cluster together, they form a crystal. In a crystal, the atoms are all
oriented in a certain three-dimensional (3-D) orderly fashion (space lattice). For example,
Heat Treatment by Induction 141

with pure iron, cast irons, and steels, there is a specific crystalline structure depending on
different factors, including temperature, pressure, and so on.
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

At temperatures below approximately 912°C (1674°F), the crystalline structure of iron is


described as body-centered cubic (BCC) space lattice (referred to as α-iron or ferrite). The
atoms are arranged in the form of a cube with an atom at each corner and one at the center
of the cube. α-Iron exhibits strong ferromagnetic properties at room temperature. Besides
iron, some other metals including tungsten, molybdenum, and chromium have a BCC
structure at room temperature. BCC structure is relatively brittle.
At higher temperatures (below 1392°C [2538°F]), the atoms are still arranged in the shape
of a cube with an atom at each corner and six additional atoms located at the center of
each face of the cube and no atom at the center of the cube. This structure is referred to
as a face-centered cubic (FCC) space lattice that is called γ-iron (gamma iron) or austenite.
Austenite has a closer-packed structure compared to ferrite, meaning reduced volume and
greater density. Austenite is also noticeably more ductile and does not exhibit ferromag­
netic properties.
Upon further heating and after exceeding the temperature of approximately 1392°C
(2538°F), the FCC lattice of iron will transform back to the BCC structure known as δ-iron
(delta iron), which is not crystallographically different from α-iron, except the temperature
range it exists at.
It should be mentioned that it is not always necessary to heat metals to elevated tempera­
tures in order to obtain an FCC crystalline structure. Such metals as copper, aluminum,
gold, and nickel have an FCC structure at room temperature.
The changes occurring in the crystal structure (space lattice) are called allotropic trans­
formation. When the iron structure changes from one type to another, there is a thermal
effect called the latent heat of transformation. Table 4.1 shows experimental values of the
latent heat for pure iron [170,222].
The appearance of the latent heat is different depending on whether the iron is
being heated or cooled down. On heating, additional energy is absorbed to support
the processes taking place during crystalline transformation, but on cooling, energy
is released. On steady heating with relatively low heat intensity, the temperature rise
will be appreciably slowed down or even stopped (as in the case of equilibrium condi­
tions exhibiting a temperature plateau) when iron experiences a structural change. The
first interruption in the temperature rise during heating of iron takes place when the
pure iron loses its ferromagnetic properties, becoming paramagnetic (Figure 4.1). This
critical temperature is known as the A 2 temperature or the Curie point. Additional
energy is required to disorient the magnetic dipoles in the iron, resulting in the loss of
ferromagnetic properties [48].

TABLE 4.1
Latent Heats of Phase Transformation for Pure Iron
Transformation Temperature (°C) Latent Heat (kJ/kg)
Alpha ferrite to austenite 912 16
Austenite to delta ferrite 1394 15
Fusion (liquid to solid) 1538 247 ± 7
Source: J. Dossett and G. Totten (editors), ASM Handbook, Vol. 4A: Steel Heat Treating Fundamentals and
Processes, ASM Int., Materials Park, OH, 2013; Y.S. Touloukian, Thermodynamic Properties of High
Temperature Solid Materials, Vol. 1, MacMillan, 1967, p. 604.
142 Handbook of Induction Heating

Liquidus Solidus
1600
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

δ-Iron
BCC
1400 Ac4 A4 Ar4 (Ferrite)

(Austenite)
Nonmagnetic

γ-Iron
1200
FCC
Temperature, °C

1000 Ac3 Ar3


A3
800 Ac2 Ar2
A2
600

(Ferrite)
α-Iron
Magnetic
ting

Cool
400 A4 = 1394°C BCC
Hea

in
A3 = 912°C

g
200
A2 = 770°C
0
Time

FIGURE 4.1
Heating and cooling curves for pure iron.

In some publications, the paramagnetic form of α-iron is referred to as β-iron (beta iron).
However, it is broadly agreed by modern metallurgists to avoid using the term β-iron [33].
Since the paramagnetic state of α-iron still has a BCC structure, it is commonly accepted
to characterize the state of iron within the temperature range of 768°C (1414°F) to 912°C
(1674°F) as a nonmagnetic form of α-iron.
As can be seen in Figure 4.1, the second interruption in the temperature rise takes
place after reaching a temperature of approximately 912°C (1674°F), when the iron
structure undergoes a change from α-iron (ferrite) to γ-iron (austenite). The third and
fourth interruptions occur during the transition of the iron structure from γ-iron to
δ-iron (1392°C [2538°F]) and at the melting point (1528°C [2782°F]), correspondingly.
Table 4.2 shows the states of pure iron as a function of temperature at a pressure of
1 atm.
The α-iron and γ-iron are two of the most important forms of iron that are present in the
majority of induction applications. Since δ-iron exists only at temperatures above 1392°C
(2538°F), induction heating practitioners seldom come across this structure unless melting
or welding.
As one can notice from Figure 4.1, the heating and cooling curves are almost identical.
However, there are two principal differences:

TABLE 4.2
States of Pure Iron as a Function of Temperature
Form Crystalline
State of Iron of Iron Structure Name Temperature Range

Solid α-Iron BCC Ferrite Up to 912°C (1674°F)


Solid γ-Iron FCC Austenite 912°C (1674°F) to 1392°C (2538°F)
Solid δ-Iron BCC 1392°C (2538°F) to 1528°C (2782°F)
Liquid 1528°C (2782°F) to 2880°C (5216°F)
Gas Above 2880°C (5216°F)
Heat Treatment by Induction 143

• All critical temperatures on cooling are lower to some degree than critical tem­
peratures on heating.
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

• As stated above, contrary to heating, during cooling, when the iron undergoes a
crystalline transformation, additional energy is freed and heat is generated (latent
heat). This additional energy accounts for the interruptions seen on the cooling
curves.

Critical (transformation) temperatures are sometimes referred to as delay or arrest


points, and their existence represents an inherent resistance of the material to any struc­
tural transformation or modification.
The symbol “A” representing critical temperatures (i.e., A1, A2, A3, Acm) originates from the
French word arret (meaning arrest). Those transformations are diffusion-driven processes.
In order to distinguish critical temperatures that appear during the heating from similar
temperatures in the cooling cycle, the symbol “c” (which stands for the French chauffage,
“heating”) is used as a subscript of the symbol A to designate the heating cycle. The sym­
bol “r” (which stands for French refroidissement, “cooling”) is used to represent a cooling
cycle [33,149,151].
In addition to the phenomenon of interruption or delay of the heating/cooling
curves, several other phenomena may also occur at critical temperatures. Marked
changes in the behavior of several physical properties of metals and alloys includ­
ing electrothermal properties, volumetric changes, variation of density, and elongation
might be observed at certain critical temperatures. Figure 4.2 shows an example of how
the existence of the critical points affects the thermal expansion and density of steel
during the heating/cooling cycle. At the beginning of the heating cycle (Figure 4.2a),
steel expands proportionally to the temperature rise. Upon reaching the Ac1 critical
point, the expansion of steel stops and the steel begins to experience a contraction of
its volume, but upon reaching a critical temperature of Ac3 (upper critical temperature
that signifies complete austenitization during continuous heating cycle), the steel then
starts to expand again. When the steel undergoes a cooling cycle, the reverse changes
occur.

8000
ing

α − Fe, Density = 7876–0.297 T–5.6 10–5 T2


7900
at

γ − Fe, Density = 8100–0.506 T


g

Density of pure iron, kg/m3


Thermal expansion of steel

He

Ac1
olin

7800
Co

Ar1
Ac3 7700
ng
ati

Ar3 7600 α − Fe
He

7500
g
in
ol

γ − Fe
Co

7400

7300
Temperature 0 200 400 600 800 1000 1200 1400
(a) (b) Temperature, °C

FIGURE 4.2
Thermal expansion of steel (a) and iron density (b) versus temperatures. (From A. Jablonka, K. Harste,
K. Schwerdtfeger, Steel Research, Vol. 62, 1991, pp. 24–33.)
144 Handbook of Induction Heating

Density and volumetric changes of metallic materials taking place during the heating/
cooling cycle played an important role in the appearance and magnitude of transient and
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

residual stresses that might result in noticeable shape distortion of the heat-treated compo­
nents. The transient and residual stresses under certain conditions may also cause crack­
ing. Review of factors effecting crack initiation and propagation as well as their prevention
are discussed later in this chapter and in Refs. [2,4,7,15,149,151,172–186].
Idealistically speaking, in the case of sufficiently slow heating/cooling that represents
the equilibrium condition; transformation temperatures should be the same (practi­
cally speaking) on heating as well as on cooling, resulting in no considerable difference
between Ac and Ar critical temperatures. Under equilibrium conditions, phase transforma­
tion changes are reversible.
However, realistically, the equilibrium condition and even near-equilibrium conditions
simply do not exist in the “real world of commercial applications.” It is so, particularly in
the case of IHT.
The difference between Ac and Ar temperatures represents thermal hysteresis, which
is a function of several factors including the heating/cooling rate, the alloy chemical
composition, and microstructure. A greater rate of heating/cooling results in a greater
difference between the Ac and Ar temperatures. Ac1 and Ac3 critical temperatures will
exceed the corresponding A1 and A3 temperatures during rapid heating. However, Ar1
and Ar3 critical temperatures will respectfully lower the A1 and A3 temperatures during
intense cooling.
It is especially important to take the phenomenon of thermal hysteresis into consid­
eration for induction hardening applications where rapid heating (e.g., 100°C/s and
higher) and intense cooling are utilized when the hot steel is spray quenched in water
or polymer-based aqueous quenchants have temperatures near room temperature.
Tables 4.3 and 4.4 show examples of typical heat intensities and cooling severities for
common applications.

TABLE 4.3
Heat Intensities in Common IHT Applications (Assuming Heating from Ambient
Temperature to Temperatures of Ac3 Temperature Range)
Heat Intensities
Application °C/s °F/s
From ambient temperature to temperatures of A c3 temperature range
Contour hardening of small- and medium-size gears. 300–1500 572–2732
Surface hardening of shaft-like components. 150–800 302–1482
Through hardening or deep case surface hardening. 50–500 122–932
Normalizing of thin wires, ropes, rods, strips, etc. 250–400 482–752
Normalizing of “thick” workpieces and through heating before hot working. 2–60 36–142

From ambient temperature to temperatures below A c1 temperature


Subcritical annealing of “thin” workpieces. 50–350 122–662
Stress-relieving and high-temperature tempering (from ambient to 20–60 88–142
approximately 650°C).
Low-temperature tempering of medium-size components (from ambient to 4–10 39–50
approximately 300°C).
Heat Treatment by Induction 145

TABLE 4.4
Cooling Intensities
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

Quenchant Temperature Velocity Cooling Intensity


Quenchant °C °F m/s W/(m 2 °C)
Pure water 32 90 0 5000
0.51 12,000
55 130 0 1000
0.76 10,500
25% Polyvinyl pyrrolidone 43 110 0 3500
0.76 7500
Conventional oil 65 150 0.51 3000
Air 27 80 0 200
5.08 350
Source: G. Totten, C. Bates, N. Clinton, Handbook of Quenchants and Quenching Technology, ASM Int’l, Materials
Park, OH, 1993.

4.1.2 Fe–Fe3C Phase Transformation Diagram and Steel Classifications


Among all alloys, steel is the most commonly used in industry. Ideally, plain carbon steel
is a binary alloy of iron and carbon. It is a solid solution that often contains more than 98%
iron. The amount of carbon in steels varies from slightly above 0% to 2%. Most steels used
in induction hardening comprise from 0.2% to 1% carbon. Above 2% carbon, the mixture
is referred to as cast iron. Above 6% carbon, the mixture becomes so brittle that it has no
real practical use [29,158,163].
It should be stated at this point that although carbon is the major alloying element in
steel, depending on the specifics of the steelmaking practices and raw materials, a so-
called plain carbon steel may consist of a limited amount of other chemical elements,
including manganese (<1%), sulfur (<0.05%), phosphorus (<0.04%), and small traces of
others.
With the increased use of electric furnace steel, which is made from recycled scrap, other
residual elements (impurities) can be found in steels. Over the last several decades, the
amount of copper residual has increased in most common steel grades (including forging
grades). Copper as a residual element is not eliminated in the steelmaking and is often
present as residuals within a certain range, and its maximum value needs to be well con­
trolled [30,187,188].
It is commonly advised to keep the amount of sulfur and phosphorus in plain carbon
steel at a minimum since they introduce brittleness and crack sensitivity. A limited amount
of manganese plays an important role in plain carbon steels, since it prevents or minimizes
the formation of iron sulfides (FeS) that concentrate at the grain boundaries. In addition,
being very hard and brittle, FeS also has a relatively low melting temperature. A combi­
nation of these features can result in undesirable properties of steel such as sensitivity to
intergranular cracking. When an adequate amount of manganese is present in the steel,
sulfur combines with the manganese to form manganese sulfide (MnS), which is distrib­
uted within the grains, creating a less crack-prone structure. A manganese-to-sulfur ratio
of 4:1 to 5:1 is quite common for carbon steels.
The effect of the most commonly used chemical elements in steel is discussed in Refs.
[30,144–158,160–162,166] and later in this chapter.
146 Handbook of Induction Heating

4.1.2.1 Classifications of Steels
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

Carbon is the principal alloying element in steels. When iron is alloyed with a differ­
ent percentage of carbon, the critical temperatures upon slow heating/cooling can be
determined based on the iron–iron carbide (Fe–Fe3C) phase transformation diagram
(Figure 4.3). This diagram represents a graph of temperature versus carbon content of the
steel and shows the effect of sufficiently slow heating the steel or sufficiently slow cooling
that may cause a transformation in its crystalline structure [29,153,158,166,189]. It also can
be used as a rough guide to estimate the range of temperatures in which certain types of
heat treatment may be carried out for plain carbon steels.
Strictly speaking, the Fe–Fe3C phase transformation diagram is a metastable form of
the Fe–graphite diagram. Given a very long time frame, Fe3C will transform into graphite.
However, since the process of graphitization is not very common in conventional steels
and requires special alloying, the Fe–Fe3C diagram is more appropriate for steels.
The Fe–Fe3C diagram is valid only for the equilibrium condition of plain carbon steel
at a pressure of 1 atm. The existence of nonequilibrium conditions, significant amounts of
additional elements (including residuals and alloying elements), and pressure, in combina­
tion with certain prior treatment, can noticeably shift the critical temperatures.
Steels with carbon content below approximately 0.77% belong to the group of hypo­
eutectoid steels. If the steels contain more than 0.78% carbon, they are referred to as
hypereutectoid steels, and those with carbon content of approximately 0.77% to 0.78%
are called eutectoid steels (though, strictly speaking, eutectoid steel contains approxi­
mately 0.77% C).

1600 2912°F
A H D
1500 B 2732°F
(δ − Fe) J Liquid
1400 2552°F
N
1300 Liquid + 2372°F
Austenite Liquid +
Temperature, °C

1200 Cementite (Fe3C) 2192°F


Austenite E
(γ − Fe) F
1100 C 2012°F

1000 1832°F
Austenite Austenite + Cementite
G
900 + Ferrite 1652°F
A3
M 800 O Acm 1472°F
Ferrite
K
(α − Fe) 700 P S 1292°F
A1 Pearlite + Cementite Cementite (Fe3C)
600 1112°F
Ferrite + 0 1 2 3 4 5 6 7
Perlite Carbon content (wt. %)

Hypo- Hyper-
Cast irons
eutectoid eutectoid
Steels

FIGURE 4.3
Iron–iron carbide equilibrium diagram.
Heat Treatment by Induction 147

Besides categorizing steels as hypoeutectoid, eutectoid, and hypereutectoid, there are


several other systems used in industry to classify steels [30,33,157,158,169,223,224]. This
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

include the following categories:

• Steelmaking practices (e.g., blast furnace, multihearth furnace, electric arc furnace,
electroslag remelting, etc.)
• Finishing methods (hot rolled, cold rolled, as-cast, etc.)
• Deoxidation practice (killed, semikilled, rimmed, capped)
• Required minimum strength
• Prior heat treatment (spheroidized, annealed, normalized, quenched and tem­
pered, etc.) and others

Besides that, taking into consideration that in contrast to carburizing, nitriding, and
other thermochemical diffusion treatment techniques, induction hardening does not
change the chemical composition of steel (including its carbon content, unless decarburi­
zation takes place); the selected steel grade should be suitable for obtaining the required
hardness levels. This is the main reason why since the most carburized steel grades lack
sufficient carbon content, it is commonly required to change the steel grade when replac­
ing carburizing with induction hardening.
Steels can be classified depending on carbon content as ultralow-carbon (<0.15% C), low-
carbon (<0.3% C), medium-carbon (from 0.3 to 0.6% C), and high-carbon steels (>0.6% C).
This classification is not widely accepted, and the range of carbon concentration may be
shifted. Some heat treat practitioners use this classification to indicate certain features of
carbon steels, including the ability of steel to be hardened to certain hardness levels, how
easily the steel responds to induction hardening, whether the steel may exhibit a pro­
nounced tendency toward cracking during heating and quenching, and others.
Ultralow-carbon steels (e.g., SAE 1008, 1010) are used where high drawability and high form­
ability are needed (e.g., body panels for automotive industry, wires, sheets, etc.). Those steels
are considered to be noninduction hardenable because their inability to be quench hardened to
hardness levels required for the majority of applications. However, as always in life, there are
some exceptions from this general rule. For example, procedure used in induction hardening
(rapid heating and intense quenching) in combination with steel stretching can be applied in
some heat treating of strips, plates, wires, rods, and thin-walled tubes made of ultralow car­
bon steels to improve certain properties including steel strengthening, weldability, etc. Mixed
structures are typically obtained in such application. As stated above, those steels provide
good machinability, formability, and toughness rather than high hardness and are used in
structural sections, clutch plates, tubular products and others. Surface hardness and wear
resistance of components made from ultralow-carbon steels can be significantly increased by
applying heat-treating processes that are alternative to induction hardening.
Induction hardening can be effectively used for many low-carbon steels (and in particular
for those with a carbon content in the upper range that is specified for low-carbon steels).
Higher carbon contents are associated with the ability to achieve greater hardness levels.
Medium-carbon steels (e.g., SAE 1038, 1045, 15B45, 1055, 4140, 5152, etc.) are the most com­
mon steels that routinely undergo induction hardening and strengthening [1,17,30,160,190].
These steels exhibit sufficient carbon in solution in austenite to form martensitic structures
upon quenching and have been used for a variety of applications including transmission
components (i.e., input shafts, output shafts, gears, clutches, etc.), suspension and steering
148 Handbook of Induction Heating

(i.e., front wheel drive components—CV joints, steering racks, ball studs and sockets, etc.),
engine components (i.e., crankshaft, camshafts, connecting rods, rocker arms, etc.), and
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

fasteners (bolts, screws, studs, and many others). H-steels (e.g., 1050H, 4340H) are pre­
ferred in critical applications since they produce a more consistent response to hardening
as indicated by their specified hardenability [19].
In some cases, high-carbon steels (e.g., SAE 1065, 1070, 52100) are used, though the appli­
cation range of high-carbon steels in industry is limited by their low toughness, poor duc­
tility, reduced machinability, and higher cost compared to medium-carbon steels. At the
same time, there are a variety of applications including valve-spring wire, flat springs, drill
bits and other cutting tools, small washers, thin stamped components, farming parts (plow
components, scraper blades, mover knifes, etc.), grinding balls for the mining industry,
and others where high-carbon steels provide a noticeable advantage over medium- and
low-carbon steels because of their improved wear characteristics and strength levels. It is
imperative to remember that high-carbon steels might exhibit a tendency to cracking dur­
ing rapid heating or during intense quenching as well as postquench cracking. Therefore,
care should be taken in determining the hardening and tempering/stress-relieving
­recipes/protocols in order to avoid crack development when dealing with these steels.
Microalloying additions of typically 1000 parts per million or less are currently used in
a variety of forged and heat-treated steel products ensuring suitable mechanical and tech­
nological characteristics and allowing reduced production costs compared to alloy steels
[30,188,228–235,253,477]. For example, microalloying additions (including V, Nb, and Ti) in
medium-carbon steels allow improvement in mechanical properties in the as-forged con­
dition. As a result, these microalloyed forging steels can be economically advantageous as
compared to traditional quenched and tempered (Q&T) grades for a variety of applications
by reducing alloying additions (such as Cr, Ni, and Mo) and postforging processing oper­
ations. Although these microalloyed steels can exhibit hardness and fatigue properties
comparable to Q&T grades, they often have lower impact strengths than Q&T grades [188].
The lower impact properties reduce their comparative feasibility for some applications.
Another example of the economically advantageous application of microalloying is related
to their capability to enhance the grain size of Q&T steels (e.g., as is done using Nb micro­
alloying) and also increase the recrystallization temperature of austenite [30,235,236].
Microalloying techniques are also used in the production of many so-called high-strength
low-alloy steels (HSLA steels). The term HSLA steel designates a relatively large group of
low- and medium-carbon steels [30,169,227–236] that utilize the principle of microalloying
to attain certain properties (e.g., high yield strength, discontinuous yielding behavior, con­
trolling grain size, and others). HSLA steels can be categorized into several subgroups [230],
including control-rolled steels, low-carbon manganese steels, weathering steels, and so on.
Microalloying additions in steels modify mechanical properties predominantly through
the precipitation of carbides and carbonitrides during thermal or thermomechanical pro­
cessing. These carbonitrides can be used to increase strength (i.e., dispersive strengthen­
ing) or improve toughness (through microstructural refinement).
Although plain carbon, microalloyed, and low-alloy steels, being the least expensive
steels, are widely used in industry, there are applications where their properties are not
adequately suitable for meeting certain industrial requirements, which may include the
necessity to increase the depth of hardness (hardenability), to suppress the grain growth at
elevated temperatures, to increase the impact strength and toughness, and to reduce oxi­
dation. Often, it is required not only to achieve a certain property of the steel component
but also to satisfy complex criteria, including a combination of properties that are often
contradictory, for example, obtaining a combination of strength, toughness, and ductility,
Heat Treatment by Induction 149

as well as improving the mechanical properties of the steel at elevated and at low tempera­
tures and to optimize the manufacturing cost versus load-carrying capability. Therefore, a
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

reasonable compromise between often opposing requirements is needed, and this justifies
the need for using alloy steels, which may also be induction hardened.
Tool steels, martensitic and some PH grades of stainless steels, and some cast irons and
powder metallurgy materials can also be induction hardened in a similar manner to plain
carbon steels and low-alloy steels. At the same time, there are some specifics associated
with the hardening recipes/protocols when dealing with these materials.

4.1.2.2 Steel Identification Systems


A widely used system for designating carbon and alloy steel grades has been developed by
the American Iron and Steel Institute (AISI) and the Society of Automotive Engineers (SAE).
Since AISI does not write specifications, current SAE designations are primarily used in
North America. It is not realistic to expect that a certain grade of steel, having the exact chemi­
cal composition, will always be supplied. Thus, certain acceptable ranges are specified.
In recent years, steel has become a globally sourced commodity, which raises the need
for the international materials community to have a clear understanding of what kind
of deviations a heat treater might expect when dealing with different suppliers or using
certain steel grades in different applications; this accentuates the need to have the ability
to cross-reference the steels.
It should be noted that besides the steel identification system developed by the SAE,
there are several alternative systems that have been developed and administrated by other
societies, professional institutions, and industries. This includes but is not limited to the
American Petroleum Institute (API), the Steel Founders Society of America (SFSA), the
Aerospace Materials Specifications (AMS), the American Society for Testing and Materials
(ASTM), the American Society of Mechanical Engineers (ASME), the American Welding
Society (AWS), the Association of American Railroads (AAR), the Military Specifications
(MIL), the American National Standard Institute (ANSI), and many others [29,30].
In addition, many countries have their own national identification systems and steel
standards (e.g., Comité Européen de Normalisation—EN [the European standard] the
International Standards Organization—ISO, DIN [Germany], AFNOR [France], JIS [Japan],
GOST (Russia), GB [China], UNI [Italy], and others [29,30,257]). There are a number of
books and guides that provide cross-references that correlate different national steel des­
ignations and develop so-called equivalent steels based on chemical compositions and
mechanical properties or application specifics. It is imperative to remember that these
cross-references of what are supposed to be similar grades or “closest matches” might be
sufficiently accurate in some cases but could lead to noticeably different steel responses to
IHT in others. It should also be noted that certain steels used in some countries simply do
not have sufficiently close counterparts.
Dealing with so many identification systems and steel designations might be confusing
for heat treaters. As an attempt to simplify different identification systems for metals and
alloys, the Unified Numbering System (UNS) was created. UNS is gaining in popularity,
but still at the time of the writing of this book (2016), the SAE identification system is still
the most popular system in North America and its identification will be used here.
According to the SAE steel identification system, the number that is assigned to the
steel makes it easy to identify whether it is a plain carbon or alloy steel by recognizing
not only the mean value of the carbon content in a given steel but also the main alloying
element(s) (Table 4.5).
150 Handbook of Induction Heating

TABLE 4.5
General Classification of Steel Grades Based on Identification Number
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

Identification Number Type of Steel Grades Based on Principal Alloy Element


10xx Plain carbon (non-resulfurized)
11xx Plain carbon (resulfurized)
12xx Resulphurized and rephosphorized
13xx Manganese alloy
2xxx Nickel alloy
3xxx Nickel–chromium alloy
4xxx Molybdenum alloy (may contain some nickel, chromium, or both)
5xxx Chromium alloy
6xxx Chromium–vanadium alloy
7xxx Tungsten alloy
8xxx Nickel–chromium–molybdenum alloy
92xx Silicon–manganese alloy
94xxx/98xx Nickel–chromium–molybdenum alloy (may contain some manganese)
14Bxx Boron alloy
50Bxx/51Bxx Chromium–boron alloy
8xBxx Nickel–chromium–molybdenum–boron alloy

The SAE identification number of plain carbon steels starts with a number 10xx or 11xx.
The last two or three numbers represent the nominal percentage of carbon in the steel. For
example, 1045 steel would be a plain carbon steel with approximately 0.45% carbon; 1080
steel would be a plain carbon steel with approximately 0.8% carbon.

4.1.2.3 Phases of the Equilibrium Fe–Fe3C Diagram


As indicated earlier, the process that causes an atomic rearrangement of any material
(including iron or iron–carbon alloys) with a temperature increase or decrease is referred
to as phase transformation. Figure 4.1 demonstrates the phase transformations of pure
iron on sufficiently slow heating or cooling and the temperature plateaus associated with
that transformation.
The Fe–Fe3C phase transformation diagram of iron alloyed with carbon is shown in
Figure 4.3. As mentioned in Section 4.1.2.1, the equilibrium Fe–Fe3C phase transformation
diagram represents a graph of temperature versus carbon content of the steel indicating
temperatures and composition ranges where different phases are formed (iron carbide
Fe3C also referred to as cementite). Cementite is a compound with a particular “Fe-to-C”
ratio and contains 6.67% C by weight.
As can be seen from Figure 4.3, the maximum amount of carbon that can be dissolved
in α-iron (ferrite) is approximately 0.02%, taking place at a temperature of approximately
727°C (1341°F) (at point P, Figure 4.3), and with a temperature decrease on slow cooling to
room temperature, the carbon solubility in the ferrite became negligible.
In contrast to α-iron (ferrite), γ-iron (austenite), can dissolve a much greater amount of
carbon with the maximum of carbon solubility slightly exceeding 2% at a temperature of
1147°C (2097°F) (point E). A single phase of austenite occupies the area of GSEJNG, as seen
in Figure 4.3. Being an austenite stabilizer, carbon expands the austenite phase region.
Heat Treatment by Induction 151

Under equilibrium conditions, austenite does not exist in plain carbon steels at tempera­
tures below the A1 critical temperature.
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

Point S (Figure 4.3) of the intersection of lines representing the critical temperatures A1,
A2, A3, and Acm is called the eutectoid point (approximately 0.77% carbon content on the
equilibrium Fe–Fe3C diagram).
Figure 4.4 shows a simplified kinetics of the formation of austenite. In reality, when steel
is heated to the temperatures of the austenitic range, the process of grain refinement and
growth takes place. It is a diffusion-driven process being a function of temperature and
time. Below the Ac1 critical temperature, grain growth is insignificant. Even though fer­
ritic grains can grow to some extent below a low critical temperature, during an austenitic
formation, those grains are refined. Above the Ac3 critical temperature, noticeable grain
growth can be observed. ASTM E112 and ISO-643 standards provide guidelines and pro­
cedures for determining the average (apparent) grain size.
During slow heating of eutectoid steel, the formation of austenite is completed at tem­
peratures slightly above the Ac1 critical point. When heating hypoeutectoid or hypereutec­
toid plain carbon steels, it is necessary to reach higher temperatures in order to complete
the formation of sufficiently homogeneous austenite.
According to the Fe–Fe3C diagram, as a result of slow cooling at the eutectoid point,
a homogeneous austenite transforms into pearlite, which is a heterogeneous two-phase
structure. A two-dimensional (2-D) metallographic appearance of pearlite looks as a
lamellar mixture of α-iron (ferrite) and iron carbide Fe3C (cementite). A sudden change
in orientation of lamellae can be observed at the pearlite grain boundaries. Since cement­
ite contains much more carbon than ferrite, it is significantly harder but noticeably more
brittle than ferrite.

1200 2192°F
γ γ γ γ E
Austenite #4 Austenite #8 Austenite
γ γ γ γ
(γ-Fe) (γ-Fe) (γ-Fe)
1100 2012°F
γ γ γ γ
910°C #3 #7 γ
γ γ γ Fe3C
(1670°F)
1000 γ γ γ γ Acm 1832°F
α-Fe
Proeutectoid #2 γ γ
#6
γ γ Eutectoid
α Fe3C
Temperature, °C

G Fe3C
900 Austenite (γ-Fe) + 1652°F
#1 #5
Eutectoid Cementite (Fe3C)
α Proeutectoid
800 A3 #4 #8
1472°F
Austenite + Fe3C
Ferrite #3 S #7
#2 #6
A1 723°C (1333°F)
700 P #1 #5 1292°F
Eutectoid
Ferrite
Perlite

Ferrite (α-Fe) + Cementite (Fe3C) +


(α-Fe) 600 1112°F
Perlite Perlite

500 932°F
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
Carbon content (wt. %)

FIGURE 4.4
Lower left hand of the iron–iron carbide equilibrium diagram (compare with Figure 4.3).
152 Handbook of Induction Heating

Eutectoid transformation can be illustrated by Equation 4.1 [191].


Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

Cooling

γ (0.78 wt. %C) α -iron(0.022 wt. %C) + Fe3 C(6.7 wt. %C) (4.1)

Heating

Thus, eutectoid transformation represents itself as a certain type of solid state reaction,
in which one solid phase transforms into two other solid phases. This reaction is reversible
assuming equilibrium or near-equilibrium conditions.
Being a mixture of α-iron and Fe3C, the mechanical properties of the pearlite are interme­
diate between the soft ductile ferrite and the hard brittle cementite and may form over the
range of temperatures. Depending on the specific temperatures of pearlitic transformation
during cooling stage, interlamellar spacing can be different and the hardness of the pearlite
can vary measurably.
At higher temperatures, the eutectoid reaction is sluggish, requiring a relatively long
time and producing coarse grains with large interlamellar spacing. Increased undercool­
ing provides a driving force for intensifying the nucleation and growth rates of pearlitic
colonies, decreasing interlamellar spacing, and increasing its hardness.
As stated earlier, pearlite is a mixture of two phases: ferrite and cementite. This also
means that practically all of the carbon that is present in the ferritic–pearlitic structure is
contained in the pearlite.
It is difficult sometimes to reveal details of a lamellar structure of fine pearlite as it
appears in 2-D metallographic examination using light optical microscopy. Because of the
tight spacing of a mixture of α-iron and Fe3C lamellas, pearlite colonies appear as dark
areas. Figure 4.5a shows the 2-D metallographic appearance of the ferritic–pearlitic micro­
structure SAE 1538MV steel at 400× etched using 4% Nital (ferrite etches in light color,
but pearlite etches in dark colors). SEM or TEM microscopes provide better capability to
resolve the structures of pearlitic colonies.
The existence of the GSPG area must be understood since it plays an important role in
choosing the proper process parameters in a variety of steel heat-treating applications,
including hardening and intercritical annealing. The uniqueness of the area GSPG reveals

0.10 mm

0.05 mm

(a) (b)

FIGURE 4.5
(a) 2-D metallographic appearance of ferritic–pearlitic structure using light optical microscopy (pearlite, dark
gray; ferrite, light regions). (b) 2-D metallographic appearance of lath martensite. Etched with 4% Nital.
Heat Treatment by Induction 153

that regardless of appreciably different carbon contents, both α-iron (ferrite) and γ-iron
(austenite) can coexist.
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

The critical line GS (also called A3 or the upper transformation temperature line) limits
the GSPG area on the high-temperature side. An increase of carbon concentration causes
a decline of line GS. This means that the carbon lowers the temperature A3 for a complete
austenitization of hypoeutectoid steels and it also decreases the liquidus temperatures,
which has a marked effect on the selection of target temperatures for hot working and heat
treatment of steels and cast irons.
On the low-temperature side, the critical line A1, which is called the lower transforma­
tion temperature line, limits the GSPG region.
A similar phenomenon takes place in the area defined as SEFKS (Figure 4.3), where
γ-iron and iron carbide also coexist simultaneously. The critical line Acm (called cementite
solubility line) limits this area on the high-temperature side, determining the limit of solid
solubility of carbon in γ-iron.
It is imperative to emphasize at this point that all three microstructures discussed above
(ferrite–pearlite, pearlite, and pearlite–cementite) and obtained upon equilibrium cool­
ing (meaning a sufficiently slow cooling of homogeneous austenite) are produced by a
­diffusion-driven transformation [30].
From the standpoint of actually trying to predict the results of heat treatment, there is
at least one crucial factor that is missing on the Fe–Fe3C equilibrium diagram and that is
the factor of time. In order to estimate the effect of a certain heat treatment, it is necessary
to have a certain knowledge regarding the effect of different rates of heating and cooling
and conditions of austenite before cooling (homogeneous vs. heterogeneous austenite and
degree of its heterogeneity).

4.1.3 Time–Temperature Transformation Diagram


and Continuous- Cooling Transformation Diagram
The isothermal transformation (IT) diagram (also referred to as the time–temperature
transformation diagram or TTT diagram) helps develop some ideas regarding the for­
mation of the post–heat-treating microstructures based on the specifics of cooling. As an
example, Figure 4.6 shows an IT diagram (TTT diagram) for SAE 1080 plain carbon steel,
which was austenitized at a temperature of 900°C (1652°F) [4,164].
The isothermal transformation diagram plots time on the X-axis (a logarithmic scale)
versus temperature on the Y-axis. The characteristic C-shaped curves show the effect of
different cooling scenarios for a particular steel grade on the end products of homoge­
neous austenite upon its cooling by holding the steel at a fixed temperature below the A1
critical temperature. A different chart is required for each chemical composition of steel
and grain size.
As one can see from Figure 4.6, depending on the cooling conditions, a TTT diagram
consists of several well-defined regions representing different reactions:

• γ-Iron into pearlite (upper region)


• γ-Iron into bainite (middle region)
• γ-Iron into martensite (lower region)

Capital letters are also used on those diagrams to indicate the particular reaction/struc­
ture. For example, A stands for austenite, F for ferrite, P for pearlite, B for bainite, and M for
154 Handbook of Induction Heating

800
Eutectoid temperature
γ Pearlite
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

700
Ferrite
Transformation Coarse pearlite
600 ends
Fine pearlite
Temperature, °C α + Fe3C
500
Upper bainite
400 γ Bainite
Transformation
begins Low bainite
300 50%
Ms
200

100
M90% M50% γ Martensite
0
1 10 100 1000 10,000
Time, s

FIGURE 4.6
Isothermal TTT curves for plain carbon steel SAE 1080, austenitized at 900°C, grain size: 6. (From A. Troiano,
A. Greninger, Metal Progress, Vol. 50, 1946, pp. 303–307.)

martensite. In order to emphasize a certain feature of a particular structure, double capital


letters are sometimes utilized. For example, CP and FP stand for coarse and fine pearlite,
respectively, and UB and LB stand for upper and lower bainite, correspondingly.
Another distinguishing feature of the TTT diagrams is related to the presence of at least
two C-shaped curves (sometimes also referred to as S-shaped curves). The left C-shaped
curve (Figure 4.6) represents the beginning (the start) of the particular isothermal trans­
formation process. The right solid curve designates its end.
Sometimes, TTT diagrams show an additional curve positioned between the two
C-shaped curves discussed above. That additional curve represents the completion
of 50% of the transformation of the austenite and is usually shown as a dotted curve
located between the curves that designate the beginning and the end of the isothermal
transformation.
In addition to TTT diagrams, the continuous-cooling transformation diagrams (CCT
diagrams, Figure 4.7) were developed. CCT diagrams are more appropriate for induction
heat treaters because they take into consideration the continuous nature of the cooling of
austenite [29,158,164,165,170].
As has been shown by several authors [33,147,148,156,192–194], CCT diagrams are often
shifted to lower temperatures and longer times compared to TTT diagrams.
As stated above, a typical procedure for steel hardening involves heating the alloy in an
austenite phase temperature range. Then, some holding might be required at that tempera­
ture for a period long enough for the formation of a sufficiently homogeneous austenite.
Next, the steel is rapidly cooled until it is near ambient temperature. This rapid cooling
allows replacing the diffusion-dependent transformation by a shear-type (diffusion-less)
transformation of austenite creating a constituent called martensite. As a result of mar­
tensitic transformation in carbon steels, the crystallographic lattice changes from FCC aus­
tenite to body-centered tetragonal (BCT) martensite.
Besides carbon steels, the martensitic reaction is observed in other materials (e.g., uranium
and titanium alloys as well as polymorphic ceramics such as ZrO2 [260]) and the properties
of the martensite in other materials might be quite different compared to martensite in steels.
Heat Treatment by Induction 155

1600
Ac1 800°C Eutectoid temperature
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

1400
13% F #3 #1
15% F 700°C 50%
35% F #2
1200 #5
Temperature, °F

5% F F #4
P 600°C C1
65% P #5a
1000
C2

Temperature
85% P 500°C
10% B
40 15% P #6
800 B
F − Ferrite 400°C
P − Pearlite #7 #8
600 B − Bainite
M − Mart.
300°C #9
M
#10 C3
400 70% B 200°C
Hardness C60 C60 C58 C31 C24 B96
C5 C4
200 100°C
100 101 102 103 104 105 Time
Cooling time, s
(a) (b)

FIGURE 4.7
(a) CCT diagram: SAE 15B41 (0.42% C, 1.61% Mn, 0.29% Si, 0.006% P, 0.019% S, 0.004% B), grain size: 7–8. (From
Heat Treater’s Guide: Practices and Procedures of Irons and Steels, ASM Int’l, Materials Park, OH, 1999.) (b) Sketch of
examples of various cooling scenarios.

The causes for the martensitic reaction might also be quite different. For example, in
addition to being thermally induced upon cooling, the martensitic reaction can also be
caused by the presence of mechanical stresses. In cases where the martensite reaction is
thermally driven (e.g., because of intense cooling from austenite phase), the temperature
ranges where the martensitic reaction occurs and the characteristics of the martensitic
structures obtained (including hardness, strength, ductility, and toughness, to name a few)
can be substantially different for different materials. In this book, the discussion regarding
the martensitic reaction and the obtained martensitic structures will be primarily related
to carbon steels and cast irons. In these applications, martensite is commonly associated
with a constituent that is hard and strong but having a lack of ductility and toughness.
Although the TTT (IT) and CCT diagrams provide helpful general information, there are
several considerable limitations when trying to apply these diagrams to IHT.

• In addition to the assumptions discussed above, a condition of formation of a suf­


ficiently homogeneous austenite before quenching has been achieved in devel­
oping both the TTT and the CCT diagrams. This is not always the case in rapid
induction hardening.
• TTT diagrams are of limited use for heat treaters because IHT is far from being an
isothermal process.
• It is very important to remember that samples with small cross sections were used
to obtain these diagrams. Therefore, there will be some inherent errors in trying to
directly apply these curves to moderate- or large-sized parts and complex geom­
etry components.
• Although CCT diagrams take into consideration the continuous cooling nature
during quenching, an assumption has been made that the cooling curves have a
constant cooling rate, which is not always the case because of the presence of cold
heat sink(s) and other thermal phenomena that are specific for many IHT applica­
tions. In surface hardening, the existence of a cold core that acts as a heat sink has
a marked effect on the cooling rate during quenching, measurably affecting the
hardness and the hardened patterns.
156 Handbook of Induction Heating

• Finally, in IHT, the thermal heat exchange between the surface of the heated
workpiece and the quenchant, among other factors, is a function of the surface
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

temperature, which obviously might not be the same as used during testing. The
workpiece temperature (austenitizing temperature) before applying the quen­
chant is typically noticeably higher compared to those used in developing the
CCT and TTT diagrams.

As mentioned above, regardless of the obvious limitations of the TTT (IT) and CCT dia­
grams, they are useful tools that help the heat treat professionals to understand the basic
phenomena and principles of heat treatment.
In the case of using moderate quenching or interrupted quenching, the final micro­
structure of steels can combine different structures. Figure 4.7b shows several examples
of different cooling scenarios based on the severity of quenching. The time-cooling line
C1 represents a slow cooling. This line crosses the left C–shaped transformation curve in
the upper region of the isothermal transformation diagram. At point 1, austenite starts its
transformation. The phase transformation process continues until the time-cooling line C1
crosses the right solid curve that represents the end of the transformation process. At point
2, all of the austenite transforms into coarse pearlite.
Time-cooling curve C2 represents more intense cooling compared to line C1. It enters a
transformation region at point 3, meaning that austenite will start to transform into fine
pearlite. After crossing the curve that represents the end of transformation (point 4), all
austenite will complete its transformation, forming fine pearlite, which has closely spaced
alternating lamellae of ferrite and cementite. Fine pearlite is associated with greater hard­
ness compared to coarse pearlite.
The “austenite-to-pearlite” transformation is a diffusion-driven process that, in the
case of plain carbon steels, occurs within the temperature range of approximately 720°C
(1328°F) to 550°C (1022°F). A temperature range of approximately 720°C (1328°F) to 620°C
(1148°F) corresponds to coarse pearlite transformation temperatures and a temperature
range of approximately 620°C (1148°F) to 550°C (1022°F) corresponds to temperatures typi­
cal for a fine pearlite (formerly known as troostite) transformation. Besides higher hard­
ness, fine pearlite that is formed upon more rapid cooling has better wear resistance and
strength compared to coarse pearlite.
Upon cooling of austenite below the A1 critical temperature, colonies of pearlite occur
initially at the austenite grain boundaries. Thanks to the diffusion-driven nature of pearl­
itic transformation, the initially developed colonies of pearlite continue to grow, enlarging
in size and crossing the grain boundaries of the parent austenite. At the same time, newly
developed pearlitic colonies continue to occur during the transformation process.
Time-cooling curve C 3 crosses the left transformation curve in the area of fine pearl­
ite transformation (point 5). Approximately, 50% of the austenite will transform into fine
pearlite at point 5a. Then, curve C3 moves away from the dotted curve, representing a 50%
isothermal transformation of austenite. The transformation process practically stops when
curve C 3 moves back from the dotted curve. The remaining 50% of the austenite will con­
tinue its transformation after crossing the 50% dotted curve. At point 6, the rest of the 50%
of the austenite will complete its transformation into bainite.
In plain carbon and low-alloy steels, the upper bainitic isothermal transformation occurs
within the temperature range of approximately 550°C (1022°F) to 380°C (716°F), represent­
ing a nonuniform structure of ferrite and cementite. Figure 4.8 shows a 2-D metallographic
appearance of upper bainite (dark or outlined) and as-quenched martensite (gray or white)
in 5160 alloy steel (Fe–0.6% C–0.85% Mn–0.25% Si–0.8% Cr) that was austenitized at 830°C
Heat Treatment by Induction 157

2% Nital 4% Picral
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

10 µm
10 µm

(a) (b)

FIGURE 4.8
2-D metallographic appearance of upper bainite (dark or outlined) and as-quenched martensite (gray or
white) in 5160 alloy steel (Fe–0.6% C–0.85% Mn–0.25% Si–0.8% Cr) that was austenitized at 830°C (1525°F)
for 30 min, isothermally held at 538°C (1000°F) for 30 s to partially transform the austenite, and then water
quenched (untransformed austenite forms martensite). (a) 2% Nital etched. (b) 4% Picral etched. (Courtesy
of George F. Vander Voort.)

(1525°F) for 30 min., isothermally held at 538°C (1000°F) for 30 s to partially transform
the austenite, and then water quenched (untransformed austenite forms martensite).
Figure 4.8a is a 2% Nital etched image and Figure 4.8b is a 4% Picral etched image for a
comparison.
Time-cooling curve C 4 represents further increased quenching severity compared to
the previous three scenarios. An important feature of curve C 4 deals with the fact that it
“misses” the “nose” of the left transformation curve. This means that austenite will not
transform into any soft structures such as ferrite, pearlite, or upper bainite. As shown in
Figure 4.7b, just before reaching a region of martensitic transformation, the cooling process
has been interrupted or its severity has been drastically reduced, resulting in entering and
exiting a transformation region in the lower bainitic area. At point 7, the steel still consists
of 100% austenite. At point 8, all austenite will transform into lower bainite. An isothermal
reaction (“austenite-to-lower bainite”) occupies the temperature range of approximately
380°C (716°F) to 220°C (428°F). Lower bainite has an acicular morphology and appears as a
combination of plate-shaped entities looking somewhat similar to martensite [30].
As an example, Figure 4.9 illustrates a 2-D metallographic appearance of lower bainite
(dark) and as-quenched martensite (white/gray) in 5160 alloy steel (Fe–0.6% C–0.85%
Mn–0.25% Si–0.8% Cr) that was austenitized at 830°C (1525°F) for 30 min, isothermally held
at 343°C (650F°) for 20 min to partially transform the austenite, and then water quenched
(untransformed austenite forms martensite). Figure 4.9a is a 2% Nital etched image and
Figure 4.9b is a 4% Picral etched image.
SEM microscopy is better suited to reveal the details of the morphology of upper bainite
and lower bainite.
Among modern metallurgists, work by Professor Sir H.K.D.H. Bhadeshia and his col­
leagues [154,155] mark a substantial impact on systematizing accumulated knowledge and
developing the theory on bainitic transformation, exposing and explaining its intricacies.
Bainitic microstructures are noticeably harder than pearlite and, in particular, compared
to ferrite. Regardless of higher hardness, the lower bainite is both harder and tougher than
upper bainite. This is so because upper bainite consists of aggregates of ferrite plates (also
referred to as sheaves) with iron carbides situated between plates. This concentration of
iron carbides is responsible for the lack of toughness and exposes the brittleness of this
158 Handbook of Induction Heating

2% Nital 4% Picral
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

10 µm
10 µm

(a) (b)

FIGURE 4.9
2-D metallographic appearance of lower bainite (dark) and as-quenched martensite (white/gray) in 5160 alloy
steel (Fe–0.6% C–0.85% Mn–0.25% Si–0.8% Cr) that was austenitized at 830°C (1525°F) for 30 min, isothermally
held at 343°C (650°F) for 20 min to partially transform the austenite, and then water quenched (untransformed
austenite forms martensite). (a) 2% Nital etched. (b) 4% Picral etched. (Courtesy of George F. Vander Voort.)

structure. Since lower bainite is formed at lower temperatures, then there is less chance for
all cementite particles to segregate to areas between the bainitic ferrite plates and some of
the carbon would participate within the supersaturated ferritic plates, making this struc­
ture less brittle compared to the upper bainite [155].
In some steels, both the upper and lower bainite may be formed upon heat treating;
in others, only the upper or the lower bainite can be formed. For example, the results of
calculations conducted by M. Takahashi and H.K.D.H. Bhadeshia [356] indicate that when
heat treating the plain carbon steels with the carbon content less than 0.32% C, the lower
bainite was not formed (Figure 4.10).
In this case, the formation of upper bainite was followed by the creation of martensite,
skipping the formation of lower bainite. In contrast, in the plain carbon steels with >0.4% C,
it is expected that the formation of the upper bainite would be skipped and instead the
lower bainite and martensite (if cooling severity is sufficient) would be formed. Both the
upper bainite and lower bainite could be formed on quenching of the plain carbon steels
from an austenite phase if carbon content is within 0.32% C to 0.4% C.
700
Bs
600
Upper bainite
Temperature (°C)

500 Ms
Lower
400 bainite
Martensite
300

200 Martensite
LBs
100

0
0 0.1 0.2 0.3 0.4 0.5 0.6
Carbon content (wt. %)

FIGURE 4.10
Calculated lower bainite transformation start temperatures as well as martensite-start and bainite-start tem­
peratures for plain carbon steels as a function of temperature. (Data based on materials provided in H.K.D.H.
Bhadeshia, Bainite in Steels, Institute of Materials, UK, 2015; M. Takahashi, H.K.D.H. Bhadeshia, Model for transi­
tion from upper to lower bainite. Mater. Sci. Technol. 6: 592–603, 1990.)
Heat Treatment by Induction 159

The strength of lower bainite approaches the strength of martensite, but its toughness
can be superior compared to the toughness of tempered martensite.
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

There is an opinion that the formation of bainite is governed by a complex combination


of a diffusion-controlled and diffusion-less (shear-type) process. However, many metal­
lurgists consider that the formation of lower bainite is not a diffusion-driven process, but
rather a diffusion-less transformation. A detailed discussion on the features of bainitic
reactions, the variety of its morphologies, microstructures, and properties can be found in
Refs. [29,30,149,154,155,158,196].
Time-cooling line C5 represents an intense quenching that is typical for obtaining a mar­
tensitic structure. Line C5 crosses the transformation curves in a martensitic region only
(it enters and exits the region at points 9 and 10, correspondingly). It is imperative to stress
one more time that in most induction hardening applications, it is required to “miss” the
nose of the TTT curves preventing the formation of upper transformation products.
Unfortunately, the isothermal TTT diagrams of some non-eutectoid plain carbon steels are
shifted far to the left. This means that regardless of quench severity (practically speaking), the
cooling curve enters the region of the beginning of the transformation in the upper transfor­
mation area. This could indicate an inability to obtain an entirely martensitic microstructure
while isothermally transforming some plain carbon steels from the austenite state.
It should be mentioned at this point that there are much less frequent cases when form­
ing predominately bainitic or even fine-grain pearlitic structures are desired, instead of
forming martensitic structures [30,155,233]. For example, in contrast to the great major­
ity of induction hardening applications, when hardening of high-carbon steel rails for
railways, because of specifics of process requirements and safety concerns, forming any
martensite in the as-hardened structure is not permitted, but it is required to form fine
pearlitic structures. In another example, in European railways, forming martensitic struc­
tures of as-quenched rails is also prohibited; instead, forming bainitic structures or using
pearlitic steels.
Nevertheless, such cases are more of the exception than the rule, and for the great major­
ity of induction hardening applications, the goal is to develop fully or predominately mar­
tensitic structures, and an amount of martensite in the as-quenched structure is often the
measure of how successful the induction hardening was.
Martensite is a nonequilibrium phase of supersaturated solid solution of iron and car­
bon. When the austenite is rapidly cooled, the carbon is trapped in the crystal structure
[28,30,154,166–171]. Two dimensions of the unit cell of martensite are equal, but the third
one is enlarged as a result of the trapped carbon. This distortion of the martensitic crys­
talline structure is the reason for the high hardness that is developed when austenite is
transformed to martensite [28,30,161].
In a properly polished and etched 2-D section of specimen examined in a light opti­
cal microscope, martensite appears as an acicular or needlelike structure. As mentioned
above, the formation of martensite is governed by a shear-type (diffusion-less) transfor­
mation of austenite, meaning that the transformation takes place almost instantaneously
upon reaching a certain temperature. Since the transformation is diffusion-less, the ini­
tially created martensite plates do not grow in size with time. Instead, new plates continue
to form upon further cooling. The diffusion-less nature of transformation also means that
the chemical composition of martensite is the same as the parent austenite. Thus, higher
carbon austenite produces a martensite with a greater amount of carbon trapped in it and
with higher densities of crystal imperfections (also referred to as dislocations).
Martensitic transformation takes place over a temperature range from Ms (which stands for
“start”) to Mf (which stands for “finish”) that primarily depends on the chemical composition
160 Handbook of Induction Heating

of a given steel. If the quenching cycle is interrupted within the temperature range of martens­
itic transformation, no further transformation to martensite would occur. Only upon further
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

cooling to a lower temperature will it be possible to continue the martensitic transformation


again. Being a diffusion-less process, martensitic transformation is independent of time.
Figure 4.11 shows that for plain carbon steels, the Ms–Mf temperature range is directly
related to the steel’s carbon content.
The carbon content (Figure 4.12a), actual amount of martensite formed, and grain size
determine the maximum achievable hardness for a given steel. In the range of 0.25% to
0.50% carbon, the maximum hardness of the plain carbon steel assuming water spray
quenching can be approximated as a linear function of carbon content and can be roughly
estimated using a simple formula often applied in industry (Equation 4.2).

HRC = 50 * % C + 38 (4.2)

Shape of martensite 600


Lath + 500
Lath Plate
plate
600 400
Temperature (°C)

500 300
Temperature (°C)

Ms
400 200
300 Ms
100
200
0
Mf Mf
100
–100
0
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 –200
Carbon content (wt. %) 0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
Carbon content (wt. %)
(a) (b)

FIGURE 4.11
Ms and Mf temperature as a function of carbon content. (From A.R. Marder, G. Krauss, The morphology of
martensite in iron-carbon alloys, Trans. ASM, 60, 1967, pp. 651–660 (a) and A. Gulyaev, Metallurgy, Metallurgiya,
Moscow, Russia, 1977 (b).)

70
Maximum hardness 1100
100% martensite
60 A C 900
Hardness ( HRC)

Hardness (HV)

50
B
700
40 % martensite
A - 99%
D B - 90% 500
30 C - 80%
D - 50%
300
20 0 0.2 0.4 0.6 0.8 1 1.2
0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
Carbon content (wt. %) Carbon content (wt. %)

FIGURE 4.12
Relationship among hardness, carbon content, and amount of martensite. (From ASM Handbook, Vol. 4, Heat
Treating, ASM International, Materials Park, OH, 1991 (a) and L. Samuels, Light Microscopy of Carbon Steels, ASM
Int., Materials Park, OH, 1999 (b).)
Heat Treatment by Induction 161

Besides the Ms temperature, there are two additional lines (M50 and M90 lines) shown on
the TTT diagram (Figure 4.6), representing the temperatures at which 50% and 90% of the
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

austenite has been transformed to martensite, respectively.


For some steels (in particular those with high carbon content) and cast irons, the Mf tem­
peratures are below room temperature (Figure 4.11). This means that under conventional
heat-treating conditions, the existence of a certain amount of untransformed austenite (also
referred to as retained austenite [RA]) is unavoidable (Figure 4.13). In plain carbon steels
having carbon content below 0.5%, the RA is typically less than 2%. RA is located between
individual plates of lath martensite. It is often difficult to measure and even reveals a small
amount of RA using light microscopy.
As carbon content increases, so does RA, rising to around 6% in steels with 0.8% C and
to more than 30% for steels with 1.25% C [246]. The amount of RA also increases with
increasing hardening temperatures. Most alloying elements lower the Ms temperatures
and typically increase the amount of RA.
In case of the presence of an excessive amount of RA, the use of cryogenic treatment can
help transform it into martensite. The use of cryogenic treatment is associated with an
additional increase in hardness (shaded area shown in Figure 4.12b).
Besides carbon content and alloying elements, the temperature of the quenchant can also
affect the amount of RA. Several relationships—the Harris–Cohen correlation for one—
have been developed for calculating the amount of RA in a given steel based on the differ­
ence between the Ms and quench temperatures.
The Ms temperature of plain hypoeutectoid steels can be estimated using the following
formula [28,33,191]:

Ms (°C) = 512 − 453C + 217(C)2 − 71.5 (C)(Mn). (4.3)


It is more difficult to accurately calculate Mf temperatures since it is more difficult to


conduct experiments and measurements causing a wide scatter, particularly when there
is a considerable amount of alloying involved. Some investigators have suggested the

60
Volume percent retained austenite

50 [153]

40

30
[30]
20

10

0
0.2 0.4 0.6 0.8 1 1.2 1.4 1.6
Carbon content, wt. %

FIGURE 4.13
RA as a function of carbon content. (Solid curves from G. Krauss, Steels: Heat Treatment and Processing Principles, ASM
Int’l, Materials Park, OH, 2015; dotted curves from A. Gulyaev, Metallurgy, Metallurgiya, Moscow, Russia, 1977.)
162 Handbook of Induction Heating

following rough correlations between the particular stage of martensitic transformation


and Ms temperatures for plain hypoeutectoid steels.
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

M10% (°C) = Ms (°C) − 10 (error ± 3°C)


M50% (°C) = Ms (°C) − 477 (error ± 6°C)
(4.4)
M90% (°C) = Ms (°C) − 103 (error ± 12°C)
Mf (°C) = Ms (°C) − 215 (error ± 15°C).

However, other researchers [191] have found that the Mf-to-Ms temperature ratio is more
complex, and in many cases, Mf temperatures are noticeably lower than the values sug­
gested in the abovementioned formulas.
Martensitic transformation is accompanied by a volume increase. Table 4.6 shows a
change in volume and length of a plain carbon steel specimen [151]. The phenomenon of
volume change has a marked impact on the formation of transient and residual stresses
and it also affects shape/size distortion of a heat-treated component.
Depending on the carbon content (Figure 4.11a), there may be two forms of martensite:
lath martensite (in carbon steels with up to 0.6% C) and plate martensite (when carbon
content exceeds 1%), or there could be a combination of both morphologies [30]. Both mor­
phologies of martensite (lath and plate) reflect the shapes of the martensitic crystals and
appear in 2-D light optical microscopy as needle-like structures consisting of numerous
plates.
In the case of lath martensite, the martensitic plates are smaller in size and grouped
together in units having approximately the same orientation within a particular unit
(Figure 4.5b). Martensitic plates do not propagate beyond the austenitic grain boundaries
and indicate the grain size of parent austenite.
Neighboring plates of plate martensite are randomly oriented (forming “zig-zag”–type
arrangements), appearing noticeably larger in size than the plates of lath martensite.
Structures consisting of plate martensite are noticeably more brittle.
As-quenched martensite is hard but usually quite brittle, exhibiting low toughness and
ductility. This is particularly typical for freshly formed plate martensite. Numerous micro­
cracks might be developed in as-quenched plate martensite. This necessitates a subsequent
reheating (referred to as tempering) with minimum time delay. Upon tempering, instead
of “as-quenched” martensite, the heat-treated structure consists of tempered martensite.
Figure 4.14 shows the hardness variation of plain carbon steels as a function of the car­
bon content of the various phase transformation products [33].
Tempering relieves internal stresses caused by the hardening process and may reduce
to a certain degree the hardness, providing the required balance of hardness, strength,
ductility, toughness, and residual stress [30,31,154–158]. Higher tempering temperatures

TABLE 4.6
Incremental Changes in Volume and Length as a Result of Austenitic Transformation
Transformation Change in Volume (%) Change in Length (%)
γ-Iron into pearlite +2.4 +0.8
γ-Iron into lower bainite +3.2 +1.07
γ-Iron into martensite +4.2 +1.4
Source: K. Thelning, Steel and Its Heat Treatment, Butterworth, London, 1975.
Heat Treatment by Induction 163

1000
Martensite
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

800 (water quench)

Hardness (HV)
Bainite - 300°C
600

400 Bainite - 450°C


Pearlite - 650°C
200 Pearlite - 705°C

0
0 0.2 0.4 0.6 0.8 1
Carbon content (wt. %)

FIGURE 4.14
Hardness variations of plain carbon steel as a function of carbon content of the various phase transformation
products. (From L. Samuels, Light Microscopy of Carbon Steels, ASM Int., Materials Park, OH, 1999.)

applied to plain carbon steels lead to a greater hardness reduction and a corresponding
improvement in ductility, toughness, and elongation.

4.1.4 Steel’s Trace (Residual) Elements and Alloying Elements


Although the percentage of carbon by far provides the most important influence on the
properties of hardenable steels, there are a number of other chemical elements that, depend­
ing on their amount and mutual interactions, can also notably affect the steel properties.
All commercial grades of even plain carbon steels, in addition to iron and carbon, con­
tain limited amounts of other chemical elements. Some of these additional chemical ele­
ments “happened to be” in the plain carbon steel as traces or residual impurities (residuals
or contaminants) contained in the raw materials or were added to the melting pot for the
creation of certain conditions during the steelmaking process.
Because the amount of these trace elements is limited to a permissible range and closely
controlled, they allow considering plain carbon steels as a binary Fe–C alloy and typically
do not greatly affect the process of heat treating and obtained properties (assuming nor­
mal processing conditions without reaching excessive temperatures).
Another source of the presence of a considerable amount of additional elements
in steels deals with the fact that though plain carbon steels, being the least expen­
sive steels, are widely used in industry, there are many industrial applications where
their properties are not suitable for meeting certain engineering requirements. Such
requirements may include the necessity to increase the depth of hardness (hardenabil­
ity), to reduce the tendency for grain growth, to improve creep resistance, to increase
impact strength, and so on. Often, it is required to achieve not just a certain property
but a combination of properties.
Therefore, in order to provide steels with specific properties, a considerable amount of
certain chemical elements is purposely added to the steel. Chemical elements that typi­
cally serve as alloying elements include manganese, nickel, chromium, molybdenum, and
others. Alloy steels are more expensive to produce but are often worth the cost in applica­
tions demanding their use. Some examples of steels consisting of a substantial amount of
alloying elements are stainless steels, tool steels, maraging steels, bearing steels, spring
steels, and so on.
164 Handbook of Induction Heating

Because of the presence of a substantial amount of alloying elements, most alloy steels
should not be considered as binary alloys, but as at least ternary or quaternary alloys (e.g.,
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

Fe–Cr–Mn, Fe–Cr–Mo, and Fe–Cr–Ni alloy steels).


An increased amount of alloying elements leads to their potential complex interactions
that, in turn, provide a wide variety of microstructures and the needed combination of
mechanical, chemical, thermal, and electromagnetic properties.
Maraging steels can serve as an example of steel that provides a unique combination of
properties such as high strength, good ductility, and toughness, thanks to the presence of
such alloying elements as nickel (8%), molybdenum (5%), and titanium (0.4%).
When alloying elements are added to the mixture, they typically change certain heat
treat characteristics, including the position of critical temperatures, eutectoid carbon con­
tent, critical cooling rates, amount of RA, and so on.
The ability of some alloying elements to shift the nose of the C-shaped curves to the
right on the CCT diagram provides a significant practical benefit, as it makes it possible to
harden the steel using less severe quenching and improving hardenability that is associ­
ated with the capability to through harden larger sections (if required).
On many occasions, moderate and low quenching rates applied for alloyed steels reduce
the probability of crack initiation and growth as well as size/shape distortion of the heat-
treated component without sacrificing the ability to obtain the required martensitic struc­
ture, hardness levels, and depths.
It should be also mentioned here that some researchers (i.e., Dr. N. Kobasko [198,251] and
Dr. K. Shepelyakovskii [199]) believe that the effect of the cooling rate (within the martens­
itic transformation range) on crack initiation appears as a bell-shaped curve. This postu­
lation suggests that in cases when the cooling rate is extremely severe, which is several
times greater than conventional cooling, the probability of cracking can be reduced when
hardening even high-carbon steels. Though not all researchers share this point of view,
there are several experimental and computer modeling data that support this postulation.
A detailed description of alloy steels and the effect of alloying elements on microstruc­
ture and the properties of the heat-treated parts are discussed in Refs. [146,148,149,151–
162,191,195,197]. Only some factors will be outlined here.
Depending on the effect made by the alloying elements on microstructure and proper­
ties of the steel, these elements can be divided into several groups based on their interac­
tion with iron, carbon, and other elements and their ability to promote or stabilize a certain
phase. Such elements as Mn, Ni, and Co belong to a group of austenite stabilizers, which
lower the eutectoid temperature and expand the austenite phase region. For example,
because of a considerable amount of nickel (from 7% to 20%), austenitic stainless steels (i.e.,
AISI 301, 302, 304, 310) retain an austenitic structure even at room temperature.
Other elements, including Si, Cr, W, Nb, Mo, Ti, and Al, act as ferrite stabilizers. For
example, silicon, being a strong ferrite former, serves as the principal alloying element
(3% to 7% Si) for manufacturing ferritic steel sheets used as laminations for magnetic flux
concentrators, transformer cores, and shunts.
Some ferrite-forming elements can also promote the formation of carbides. Generally
speaking, metals that stand to the left of iron in the Periodic Table of the Elements
(Appendix) are capable of forming carbides. Such elements as Ti, Nb, V, Mo, W, and Cr are
called carbide formers. Under certain conditions, complex carbides can be formed in steel.
The majority of elements that do not form chemical compounds with carbon and iron are
presented in solid solution with iron [153].
When choosing the appropriate process parameters for IHT, it is particularly important
to take into account the existence of certain carbides in the steels and their size because of
Heat Treatment by Induction 165

the short heat time of the induction process. Some carbide-forming elements form stable
carbides; others form relatively unstable carbides. Experimental data show [153] that the
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

elements located further to the left of the row (see the Periodic Table of the Elements,
Appendix) form more stable carbides. Therefore, carbides formed by Ti are more stable
than carbides formed by Cr or Mn.
The austenitizing temperatures needed can be 70°C–100°C higher in alloys with strong
carbide-forming elements (e.g., Ti, Cr, Mo, V, Nb, and W) than for plain carbon steels. These
increases are attributed to the kinetics of alloy carbide dissolution in austenite, which can
be substantially slower compared to cementite dissolution, for example when NbC and VC
are involved. Such increases in induction hardening temperature and time are not gener­
ally deleterious from the viewpoint of austenite grain growth because of the effects of
alloying and rapid heating. However, there might be exceptions. For example, TiC (which
is a stable carbide) typically remains within austenite phase temperatures in the great
majority of hardening applications.
In addition to the abovementioned groups, alloying elements can also be subdivided
into subgroups, based on their effect on the position of critical temperatures (i.e., A1, A3,
and Acm), solubility in iron, the influence on the kinetics of austenitic transformation upon
cooling, and the susceptibility for grain growth, just to name a few.
As stated above, alloying elements affect the heat treatment process in many different
ways, and some of them are described in Refs. [146,148,149,151–162,191,195,197]. Table 4.7
provides a brief summary of the effect of commonly used alloying elements in steels.
Figure 4.15 illustrates the effect of alloying elements on eutectoid carbon content and
eutectoid transformation temperature [30,160]. Analysis of Figure 4.15 leads to the impor­
tant conclusion that the majority of alloying elements reduces eutectoid carbon content.
This means that a pearlitic structure may be developed with lower carbon content enhanc­
ing the steel’s ductility.
Such elements as nickel, which lowers the A1 and A3 critical temperatures, affect the
process of austenitization in a favorable way by lowering the austenitizing temperature.
Similar to nickel, silicon does not form carbides; however, it does slow the diffusion of
carbon into the iron, making it noticeably more sluggish. This requires an increased aus­
tenitizing temperature for steel alloyed with silicon and also necessitates having longer
times at austenitizing temperature compared to plain carbon steel.
It is also imperative to keep in mind that the effects of alloying elements shown in
Table 4.7 and Figure 4.15 illustrate their independent action without taking into consid­
eration the potential complex interactions among elements. The interactions between
such elements as nickel–manganese, nickel–copper, boron–titanium, and tungsten–
molybdenum–chromium can serve as good examples of the strong alloy interaction,
making noticeable impacts on the specifics of heat treating, process recipe, and the
structures obtained.
The majority of alloying elements have a pronounced effect on the transformation of
austenite during its cooling, which includes the following transformations: γ-iron into
pearlite, γ-iron into bainite, and γ-iron into martensite.
As stated earlier, one of the most significant practical benefits of using alloying elements
is the ability to shift the nose of the CCT curves to the right. Some of the alloying elements
increase the austenite precipitation rate (i.e., cobalt); however, the majority of the com­
monly used alloying elements slow the austenitic transformation.
Figure 4.16 shows two scenarios of commonly modifying the TTT and CCT curves
(“shifted nose” vs. “split nose”) based on whether the alloying element belongs to the
group of carbide formers (e.g., V, Mo, Cr, and W) or not.
166 Handbook of Induction Heating

TABLE 4.7
Some Individual Effects of Commonly Used Alloying and Residual Elements in Steels
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

Element Effect
Manganese (Mn) Lowers Ac1 and Ac3 temperatures. Increases strength and decreases critical cooling rate.
Improves hardenability. Lowers Ms temperature. Increases amount of retained
austenite. Reduces brittleness due to forming MnS instead of FeS and reduces
susceptibility to hot shortness.
Nickel (Ni) Does not form carbides. Austenite stabilizer. Ferrite strengthener. Lowers Ac1 and Ac3
temperatures. Improves hardenability and corrosion resistance. Increases impact strength
and toughness. Reduces eutectoid carbon content. Decreases critical cooling rate.
Silicon (Si) Deoxidizer in steelmaking. Strong ferrite stabilizer. Does not form carbides. Reduces
eutectoid carbon content. Slows the diffusion of carbon into ferrite, requiring longer
time or higher temperatures for austenitization (raises Ac1 and Ac3 temperatures).
Improves hardenability. Promotes graphitization. Increases strength of ferrite in
low-alloy steels but may increase the steel’s susceptibility to decarburization. Has
marked impact on electromagnetic properties (e.g., reduces magnetic hysteresis loss).
Chromium (Cr) Carbide former. Raises Ac1 and Ac3 temperatures. Requires longer time or higher
temperatures for austenitization. Improves hardenability. Resists softening on
tempering and may result in secondary hardening. Slightly decreases Ms temperature.
Increases wear resistance, corrosion resistance, toughness, and abrasion resistance.
Decreases oxidation and increases high temperature strength.
Molybdenum (Mo) Carbide and nitride former. Raises Ac1 and Ac3 temperatures. Decreases Ms
temperature. Improves hardenability particularly in high-carbon steels, hot strength,
and fatigue strength. Has a pronounced effect on secondary hardening. Suppresses
grain coarsening and tempering back.
Copper (Cu) Improves atmospheric corrosion resistance but can be undesirable in hot working.
Aluminum (Al) Strong deoxidizer used in steelmaking and grain refiner.
Boron (B) Powerful impact on hardenability (in very small amounts in low- and medium-carbon
steels).
Vanadium (V) Carbide and nitride former. Raises Ac1 and Ac3 temperatures. Improves strength and
hardenability (in additions up to 0.05%) but to a lesser extent than some other
elements. Has a noticeable effect on secondary hardening during tempering. Resists
grain growth. Increases strength and toughness. Popular microalloying element.
Titanium (Ti) Strong carbide former and carbide stabilizer. Nitride former. Noticeably raises Ac1 and
Ac3 temperatures. Improves hardenability (in combination with boron) and toughness.
Promotes globular sulfide inclusions instead of elongated inclusion. Grain growth
inhibitor.
Niobium (Nb) Carbide and nitride former. Raises Ac3 and Ar3 temperatures. Resists tempering back
(Columbium) and grain coarsening. Popular microalloying element.
Tungsten (Wolfram) Carbide and nitride former. Raises Ac1 and Ac3 temperatures. Increases hardness, wear
(W) resistance, and strength at elevated temperatures. Resists grain growth. Principal
alloying element in tool steels and also used in microalloy steels forming fine carbide
particle network.
Cobalt (Co) Austenite former. Increases strength of ferrite. Resists softening at elevated
temperatures. Increases the austenite precipitation rate. Reduces hardenability.
Sulfur (S) Improves machinability. Decreases ductility and impact strength. May lead to structural
segregation. Its presence is often considered to be undesirable (exception is free
cutting steels).
Phosphorus (P) Increases strength and hardness but decreases toughness and ductility. Introduces
noticeable brittleness. Promotes structural segregation. Its presence is often
considered to be undesirable.
Heat Treatment by Induction 167

0.8 1200
Molybdenum
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

0.7 1100
Eutectoid carbon content, wt. %

Nickel
Titanium

Eutectoid temperature, °C
0.6 Tungsten
1000
Chromium
0.5
900 Silicon
0.4
800
0.3 Chromium
Silicon
700
0.2 Manganese
0.1 600
Titanium Molybdenum Tungsten Nickel
0 500
0 2 4 6 8 10 12 14 16 18 0 2 4 6 8 10 12 14 16 18
Alloying element, wt. % Alloying element, wt. %

FIGURE 4.15
Effect of alloying elements on positioning of the eutectoid carbon content in steel and on the eutectoid transfor­
mation temperature. (From E. Bain and H. Paxton, Alloying Elements in Steel, ASM Int’l, Materials Park, OH, 1966.)

“Shifted nose” of steel “Split nose” of steel


alloyed with elements that alloyed with
do not form carbides carbide forming elements

A1 A1

P
Temperature

Temperature

B
Plain Plain
carbon steel carbon steel
Ms Ms

Time Time

FIGURE 4.16
Comparison of the “shifted nose” (a) versus “split nose” (b) isothermal transformation diagrams. (From
A. Sverdlin, A. Ness, The Effects of Alloying Elements on the Heat Treatment of Steel, Chapter 2 of Steel Heat
Treatment Handbook, edited by G. Totten and H. Howes, Marcel Dekker, New York, 1997.)

Figure 4.17 illustrates the effect of alloying elements on the Ms temperature [197] and on
the amount of RA remaining in the as-quenched steel with 1% C.
There have been several attempts to develop convenient mathematical expressions that
would allow the heat treater to calculate the critical temperatures of transformation for
different phases. Some of the formulas that can be used for a rough estimation of A1, A3, Bs,
and Ms temperatures of industrial carbon steels are shown below.
Austenitic critical temperatures (from Ref. [200]):

A3 (°C) = 910 − 203 C − 15.2 Ni + 44.7 Si + 104 V + 31.5 Mo + 13.1W


(4.5)
A1 (°C) = 723 − 10.7 Mn − 16.9 Ni + 29.1Si + 16.9 Cr + 290 As + 6.38W

168 Handbook of Induction Heating

300
Al
0.76% C 80
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

Co
250 0.9% C 70 Mn
0°C

Retained austenite, wt. %


60 125
ing
Ms temperature, °C

200 Si at Cr
50 He
40 Ni
150 Cu
Mn 30 Cu
100 Ni 20
V
V 10 Co
50 1% C Cr Al
0
Mn 0 1 2 3 4 5 6
0 Alloying element, wt. %
0 1 2 3 4 5 6
Alloying element, wt. % (b)
(a)

FIGURE 4.17
Effect of alloying elements on Ms temperature (a) and the amount of RA in quenched steel (1% C) (b). (From
A. Sverdlin, A. Ness, The Effects of Alloying Elements on the Heat Treatment of Steel, Chapter 2 of Steel Heat
Treatment Handbook, edited by G. Totten and H. Howes, Marcel Dekker, New York, 1997.)

Critical temperatures for beginning bainitic transformation (from Ref. [154]):

Bs (°C) = 830 − 270 C − 90 Mn − 37 Ni − 70 Cr − 83 Mo (4.6)

Formulas for an estimation of Ms transformation temperature:

(from Ref. [200])

Ms (°C) = 539 − 423 C − 30.4 Mn − 12.1Cr − 17.7 Ni − 7.5 Mo (4.7)


Ms (°C) = 512 − 453 C − 16.9 Ni + 15 Cr − 9.5 Mo


(4.8)
− 217 (C)2 − 71.5 (C)(Mn ) − 67.6 (C)(Cr)

(from Ref. [201])

Ms (°C) = 561 − 474 C − 33 Mn − 17 Ni − 17 Cr − 21 Mo (4.9)

(from Ref. [202])

Ms (°C) = 550 − 350 C − 40 Mn − 20 Cr − 17 Ni − 10 Mo


(4.10)
− 8 W − 35 V − 10 Cu + 15 Co + 30 Al

Heat Treatment by Induction 169

As opposed to other hardening processes including carburization and nitriding, induction


hardening is affected to a greater extent by variations of chemical composition. As discussed
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

earlier, the chemical composition of any steel may vary to some degree. For example, accord­
ing to SAE standard [29], a potential deviation of chemical composition of SAE 1055 steel can
be with the following range: 0.5%–0.6% C, 0.6%–0.9% Mn, 0.04% P (max), and 0.05% S (max),
whereas that for SAE 4340 steel can be as follows: 0.38%–0.43% C, 0.6%–0.8% Mn, 0.035% P
(max), 0.04% S (max), 0.15%–0.3% Si, 1.65%–2% Ni, 0.7%–0.9% Cr, and 0.2%–0.3% Mo.
Thus, the induction heat treaters should have a great deal of awareness regarding the
exact chemical composition of steel they are dealing with as well as its potential variations
and possible impact on the results of the heat treatment and part’s performance. Wide
compositional limits resulting from multiple steel suppliers may cause corresponding sur­
face hardness and case depth variations. For example, potential variation of carbon content
with 0.5%–0.6% C range represents ±0.05% variation, which not only could be sufficient to
affect achievable hardness level and case depth but also may even trigger crack develop­
ment under certain conditions.
In contrast, tight control of the composition dramatically reduces potential variations of
the heat treat pattern. If steel does not respond to heat treatment in an expected way, then
one of the first steps in finding the cause for such behavior is to check the chemical com­
position and assess the steel’s prior microstructures.
Besides that, some steels have the letter “M” placed at the end of their identification
number (e.g., 1038M), which means that it was modified according to corporate recipes that
are often confidential compared to conventional steels.

4.1.5 Hardenability
Hardenability is an important property of steels and cast irons. It defines the ability of
the alloy to be hardened to a certain depth and is measured as the distance from the sur­
face where certain hardness can be obtained or a specific percentage of martensite can be
formed (e.g., an achievement of 50 HRC hardness or 50% martensitic structure is often
used to define the hardenability of steel). The ability of an alloy to be hardened to a certain
depth is a function of chemical composition, grain size, the degree of homogenization of
the austenite, and the intensity of cooling during the quenching stage. High-hardenability
alloys can form martensite even if cooled relatively slowly. Low-hardenability alloys
require more severe cooling to avoid formation of upper transformation products. Thus,
high-hardenability steels are commonly affected to a lesser degree by a deviation of quen­
chant characteristics.
There are several experimental techniques that exist to quantify hardenability. Among
these techniques, the Grossmann’s hardenability test and the Jominy end-quench test are
two of the most popular.
When discussing the subject of hardenability, it is important to recognize certain subtle­
ties associated with this property when dealing with induction surface hardening com­
pared to through hardening.

4.1.5.1 Through Hardening
In order to harden the workpiece throughout its entire cross section, it is typically required
to heat the workpiece as uniformly as possible to the austenitizing temperature range and
then quench it to room temperature. Figure 4.18 [21] shows the dynamics of the radial tem­
perature distribution during through hardening SAE 4340 steel shaft (16 mm diameter)
in a normalized condition using a frequency of 10 kHz. The process cycle includes 8 s of
170 Handbook of Induction Heating

Dwell
Heating Quenching
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

1000
900 Surface

800
Temperature, °C 700
600 1 mm below

500 3 mm below
Core
400
300
200 16 mm diameter
10 kHz
100
0
0 2 4 6 8 10 12 14 16 18
Cycle time, s

FIGURE 4.18
Results of computer modeling of induction through hardening of a medium-carbon steel shaft using a fre­
quency of 10 kHz assuming constant coil current. (From V. Rudnev, G. Fett, A. Griebel, J. Tartaglia, Principles
of induction hardening and inspection, in ASM Handbook, Volume 4C: Induction Heating and Heat Treating,
V. Rudnev and G. Totten (editors), ASM Int., Materials Park, OH, 2014, pp. 58–86.)

heating and 0.5 s of dwell/soak. A short time delay (also called a dwell or soak time) is
sometimes used in through hardening to help improve the “surface-to-core” temperature
uniformity and reduce the thermal shock during the initial stage of quenching.
As can be seen, the time–temperature curves are substantially nonlinear owing to the
nonlinear nature of the electromagnetic and thermal physical properties. An impact of
physical properties on induction heating has been discussed in Chapter 3 and several ref­
erences [19,21,44,48]. The value of current penetration depth δ varies at different heating
stages. Table 4.8 shows the δ versus frequency and heating stage for typical conditions of
induction through hardening.
Upon approaching the Curie point, the intensity of heating noticeably declines. There
are four major factors responsible for this behavior [21]:

• Variation of μr and ρ during heat cycle


• Localized spike of the specific heat
• Surface heat losses
• Elimination of the heat generation as a result of magnetic hysteresis

TABLE 4.8
Current Penetration Depth (in mm) versus Frequency and Heating Stage for Typical Conditions
of Induction Hardening of Medium-Carbon Steel
Frequency (kHz)
Heating stage 0.5 3 10 30 70 200
Initial heating stage 3.6–3.9 1.4–1.6 0.7–0.85 0.42–0.5 0.3–0.38 0.15–0.22
(ambient temperature)
Final heating stage 17.5 10 5.6 3.2 2.1 1.2
(surface above Ac3)
Heat Treatment by Induction 171

As it has been discussed in Chapter 3, the core of the solid cylinder (its center) is always
heated because of thermal conduction when using solenoid-type inductors. This is so
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

because regardless of the frequency used, there is no heat generation in the core (see
Figure 3.13).
During quenching, the cooling rate at the surface of the through-heated workpiece is
more intense compared to the cooling rate of its internal areas and, particularly, at its core
(see Section 2.1.1.2). If the workpiece is thin enough, then the intensity of cooling at the core
might still be severe enough to “miss” the upper transformation region of the CCT dia­
gram and to form a sufficient amount of martensite in the core (Figure 4.19a). As a result, a
relatively uniform “surface-to-core” hardness distribution may take place.
Since the quenching intensity at the workpiece surface is greater than the cooling rate
at its core, the amount of martensite formed in the surface and subsurface areas might be
noticeably greater compared to the amount of martensite formed in the core. Therefore, the
surface hardness might be higher to some extent compared to the core hardness (assuming
that the surface has not been overheated and severe oxidation or decarburization does not
occur). Regardless of the “surface-to-core” hardness variation, the hardness pattern, being
within the specified range, may still be acceptable.
When the diameter or thickness of a through-heated workpiece increases, the core will
be positioned further away from the quenched surface, resulting in reduced quench sever­
ity there. At a certain point, with further increased workpiece diameter/thickness, the
hot core could be situated too far from the quenched surface, and the thermal conduction
might not provide sufficiently intense cooling of the core during surface quenching. The
core cooling curve will be shifted farther to the right and finally enter the CCT curve at
the upper transformation area (Figure 4.19b). As a result, a certain amount of upper trans­
formation products (e.g., ferrite or pearlite) would be formed within the core, making it
measurably softer compared to the surface.

Surface Core Surface Surface Core Surface


Hardness

Hardness

50% martensite

Surface Core Surface Core

Austenite Austenite

F F
Temperature

Temperature

P P

B B

Ms Ms

Time Time
(a) (b)

FIGURE 4.19
Effect of diameter on hardness profile and core cooling rate for through hardening (a) vs. surface hardening (b).
172 Handbook of Induction Heating

For a given steel grade, grain size, and quench intensity, and assuming that upon austen­
itization a homogeneous austenite has been formed, there would be a round bar specimen
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

of a particular diameter that would form 50% martensite at its core upon quenching. This
“particular diameter” is commonly referred to as the critical diameter Dcr. If the workpiece
diameter exceeds the critical diameter, then the amount of martensite formed in the core
of the cylinder would be less than 50% softening the core. The value of Dcr represents the
maximum diameter of a cylinder that can be through hardened.
The Grossmann’s hardenability test implies the concept of the critical diameter. A vari­
ety of specified round specimens having different diameters are cooled from a certain aus­
tenitic temperature to room temperature using a given quenchant and as identical cooling
conditions as possible (practically speaking).
More intense quenching provides the ability to through harden cylinders of bigger
diameters (as long as the material’s hardenability permits). Figure 4.20 illustrates the effect
of using different quenchants on the ability to through harden solid cylinders with differ­
ent diameters made from the same steel.
The main factors influencing the size of the critical diameter (chemical composition,
grain size, and homogenization of austenite) can be relatively easily determined. The
fourth factor (the quenching condition) is often the least-defined factor particularly when
spray quenching is used.
As discussed further in this chapter, spray quenching is the most popular quench method
used in induction hardening. The severity of cooling during spray quenching depends on
a combination of several factors including the surface temperature of the workpiece, its
surface conditions, type and purity of quenchant, quench device design, number and dis­
tribution of quench holes, size of orifices and angle of drilled quench holes (impingement
angle), quenchant temperature, flow rate, pressure, and others. The specifics of a particular
application and complexity of the thermo-hydro-dynamics involved in spray quenching
makes it challenging to determine the exact cooling conditions when quenching speci­
mens of various diameters.
In order to reduce the uncertainty and provide a more practical and universal quantita­
tive measure of hardenability of steels and cast irons, the term “ideal” critical diameter Di
has been accepted in industry. A value of Di represents the so-called ideal quenching that
is associated with thermal boundary condition when the surface of the round bar speci­
men would be cooled instantly down to the temperature of the quenchant (postulating

Water quench
ODcr

Oil quench
ODcr

FIGURE 4.20
Illustration of the effect of specimen diameter and quench media on hardenability (unhardened area shaded
in gray).
Heat Treatment by Induction 173

an infinite cooling capacity). A correlation between critical diameter Dcr and ideal critical
diameter Di has been established.
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

Original work done by M. Grossmann and E. Bain [147] suggests that the grain size has
marked impact on steel hardenability. Figure 4.21 shows the values of Di as a function of
carbon content and grain size for plain carbon steel. For example, the ideal critical diame­
ter that corresponds to a maximum through harden diameter of plain carbon steel bar con­
sisting of 0.5% C with a grain size of 6 would be equal to 0.26 in. (6.6 mm). Coarser-grain
steels could exhibit superior hardenability than finer-grain steels. This is one of only very
rare cases when coarser-grain structures may exhibit an advantage over fine-grain ones.
As expected, steel specimens with a greater value of Di can be easier to through harden or
hardened to deeper case depth.
Alloying elements have a marked effect on hardenability and originally have been
obtained by a number of researchers including M. Grossmann, E. Bain, H. Paxton [147,148],
W. Craft, J.L. Lamont, I.R. Kramer, and several others [221–226]. The individual effect of
commonly used alloying elements on the value of the Di can be calculated by using multi­
plying factors, which indicate an individual effect of certain element on forming martens­
itic structure at certain depth.
Earlier relationships indicating an effect of grain size and chemical composition on
the Di were re-examined by A. Moser, A. Legat, T. Kasuya, Ch. Brooks, and some other
researchers (including Climax Molybdenum Co.) and corrections have been summarized
in Refs. [149,151,157] as well as in several publications of ASM International. As an example,
Figure 4.22 shows multiplying factors for selected elements published in Ref. [157].
It can be easily shown using data from Figures 4.21 and 4.22 that for a particular steel
grade, the ideal critical diameters Dimin and Dimax (that are calculated based on the extreme
variation of the amounts of alloying elements within the allowable range though) can vary
dramatically. However, the probability of having the worst combinations of the amounts
of alloying elements (extreme combinations) is very low.
Our experience shows that in induction hardening (perhaps because of relative short aus­
tenization), we did not observe as dramatic an impact of grain size on steel’s hardenability

0.4
10 mm

0.36 Grain size 9 mm


ASTM
Ideal diameter, D1, inches

0.32 4
8 mm
5
0.28 7 mm
6
7
0.24 6 mm
8

0.2 5 mm

0.16
0.1 0.3 0.5 0.7 0.9
Carbon content, wt. %

FIGURE 4.21
Ideal critical diameter Di as a function of the carbon content and austenitic grain size. (From M. Grossmann and
E. Bain, Principles of Heat Treatment, ASM Int’l, Materials Park, OH, 1964.)
174 Handbook of Induction Heating

3.4
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

Mn
3 Mo

2.6
Cr

2.2

Si
1.8

Ni
1.4

1
0 0.2 0.4 0.6 0.8 1 1.2 1.4
Alloy content, %

FIGURE 4.22
Multiplying factors for different alloying elements for hardenability calculations. (From Heat Treater’s Guide:
Practices and Procedures of Irons and Steels, ASM Int’l, Materials Park, OH, 1999.)

as it was illustrated in Figure 4.21. Temperatures of grain coarsening range do affect the
hardenability because steels consisting of carbides and carbonitrides have a greater ten­
dency to be dissolved into austenite at higher temperatures, affecting a chemical composi­
tion of austenite.
It should be noted that an individual effect of alloying and residual elements on
steel’s hardenability can be altered to a noticeable degree in the presence of several
elements which could produce a combined impact. It is important to recognize that
depending on different steel suppliers or even different batches of steel produced by
a single supplier; it is still possible to have noticeably different steel hardenability.
This means that the hardened case depth as well as the hardness pattern could be
affected to some degree depending on the steelmaking practice, the raw materials, and
a number of other factors. Therefore, it might be beneficial to use H-steels that have
tighter control on hardenability limits (band) compared to bands of steels with normal
variation of composition. H-steels have basically the same identification number as
regular steels; however, the suffix “H” is added at the end of the H-steels’ identification
numbers (e.g., SAE 1045H, 4150H, 4340H). It should be noted here that not all steels are
available as H-steels.
Discussing the subject of the tight control of the hardness pattern, it is important to
notice an enormously powerful impact of very small amount of boron (less than 0.005% B)
on steel hardenability enhancement. However, this beneficial quality of boron may intro­
duce some challenges emphasizing the need of its tight control, because even seemingly
very small variations of boron content (e.g., ±0.0005% B) might have a significant impact on
hardenability and hardness pattern.
In order to conduct a Grossmann’s hardenability test, it is necessary to have a variety
of round specimens (cylinders) with different diameters that have identical initial prior
structure and chemical composition and are uniformly heated to the same temperatures
and placed under identical cooling conditions using a particular quenchant and an agita­
tion rate.
Heat Treatment by Induction 175

The results of the Grossmann’s hardenability test are more suitable for through-­
hardening  applications. Unfortunately, it is more difficult to apply those results for surface
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

hardening. Moreover, a quantification of an agitation rate (e.g., mild, moderate or severe) might
be challenging.
Besides a Grossmann’s hardenability test and its concept of ideal critical diameter, there
is another parameter termed “the carbon equivalent” (CE) that is used sometimes by heat
treaters to assess a steel’s hardenability and the potential impact of the most common
alloying elements in a combined state on the hardenability of the steel. For example, CE is
often used in welding applications to correlate the percentage of alloying elements other
than carbon to the equivalent carbon percentage of plain carbon steel [29,30,33,262,263].
It should be noted that CE represents a rough rule-of-thumb estimate based on empiri­
cal data fit, providing a general assessment of the steel’s weldability and the probability of
suffering brittle failures as a result of the high values of CE.
A number of empirical formulas have been developed in industry to estimate CE values
in steels. This includes but is not limited to Equations 4.11 through 4.14 [263].
The Dearden and O’Neill formula (Equation 4.11) was adopted by the International Institute
of Welding in 1967 and has been found to be appropriate for calculating CE values and assess­
ing the hardenability in a large range of commonly used plain carbon and carbon–manganese
steels, but not microalloyed high-strength steels or low-alloy Cr–Mo steels [263,264]

 % Mn   % Cr + % Mo + % V   % Cu + % Ni 
CE = %C +  +  +   . (4.11)
 6   5 15

Another available formula is Equation 4.12 [263,265]:

 % Mn + % Si   % Cr + % Mo + % V   % Cu + % Ni 
CE = % C +   +   +   . (4.12)
 6  5 155

N. Yurioka [263,266] developed a CE formula for low-carbon alloy steels (Equation 4.13):

% Mn % Cu % Ni % Cr % Mo
CE* = %C * + + + + + , (4.13)
3.6 20 9 5 4

where % C* = 5% C for % C ≤ 0.3%; % C* = % C/6 for % C ≥ 0.3%.


Lorenz and Duren suggested a formula (Equation 4.14) for low-carbon pipeline applica­
tions [169,267]:

% Si % Mn + % Cu % Cr % Ni % Mo % V
CE = %C + + + + + + . (4.14)
125 16 16 60 40 15

Again, the value of CE should be treated only as a rough ballpark estimator or an indica­
tor helping to make a decision whether a particular steel grade may exhibit brittleness and
preheating before rapid heating or welding is required or not and how critical it is to use
after heating (e.g., stress relieving or normalizing). In other cases, some companies specify
certain values of CE to develop an idea regarding obtaining the ferritic–pearlitic micro­
structure after controlled cooling of forged microalloyed steels.
176 Handbook of Induction Heating

4.1.5.2 Induction Surface (Case) Hardening


Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

As discussed in Section 2.1.1.1, some applications, instead of through hardening, require


the formation of martensitic layers at specific areas of the workpiece (e.g., at its surface)
while allowing the microstructure of the remainder of the part (i.e., core) to be unaffected
by the heat treatment process. Thanks to the skin effect and some other electromagnetic
phenomena (see Section 3.1), it is possible with induction heating to generate heat sources
in selected areas of the workpiece where the metallurgical changes are required.
Surface hardening is accomplished by raising the required depth of the hardenable alloy
above the Ac3 critical temperature to the point where sufficiently homogeneous austenite is
formed followed by cooling the workpiece rapidly to produce martensite.
As an example, Figure 4.23 shows the dynamics of temperatures during surface hard­
ening of the same shaft as in the previous case study shown in Figure 4.18. However,
the required hardness case depth range in this case study is only 1.2–2.7 mm. It is also
required that the total heat-affected zone (HAZ, the region where the phase transforma­
tion occurs) should not exceed 3.5 mm below the surface.
The comparison of these two figures (Figures 4.18 and 4.23) reveals three main features
of induction surface hardening versus through hardening:

• The radial temperature distribution before quenching is considerably nonuniform


in case of surface hardening.
• Surface hardening requires using higher frequencies compared to through hard­
ening (assuming similar size parts) because of the necessity to localize the heat
generation to the surface and near-surface regions.
• Heat times are noticeably shorter with surface hardening suppressing the ten­
dency of thermal conduction to spread the heat within the entire cross section.

Heating Quenching
1000
900
800 Surface
1 mm below
700
Temperature, ºC

3 mm below
600
Core
500
400
300
200 16 mm diameter
100 125 kHz

0
Cycle time, s

FIGURE 4.23
Results of computer modeling of the dynamics of the radial temperature distribution during single-shot
hardening of a steel shaft (16 mm diameter) using a frequency of 125 kHz assuming constant coil current.
Required case depth: 1.2 mm; SAE 4340; heating time: 2 s; quench time: 6 s. (From V. Rudnev, G. Fett, A. Griebel,
J. Tartaglia, Principles of induction hardening and inspection, in ASM Handbook, Volume 4C: Induction Heating
and Heat Treating, V. Rudnev and G. Totten (editors), ASM Int., Materials Park, OH, 2014, pp. 58–86.)
Heat Treatment by Induction 177

Figure 4.23 reveals that after 2 s of heating, the surface layer of the shaft required to be
hardened reaches the needed thermal condition for austenitization, taking into consider­
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

ation the nonequilibrium nature of the phase transformation associated with rapid heating
(see Section 4.1.6). Average heat intensity exceeds 450°C/s. Because of intense heating and
short heat time, the core temperature is approximately 520°C at the end of the heating cycle
regardless of the relatively small size of the shaft (the radius is only 8 mm). Spray quench­
ing begins immediately after the completion of the heating stage (though in other cases,
some dwell or soak time might be applied).
A dramatic decrease of surface and subsurface temperatures occurs practically instanta­
neously during quenching. At the same time, there is a time delay in cooling of the internal
regions (e.g., 3 mm below the surface) and particularly in the shaft’s core. Because of ther­
mal inertia, the core temperature continues to rise during 1 s of quenching. The internal
regions located at a distance greater than approximately 3 mm below the surface will not
be heated above the Ac1 critical temperature; thus, they will not be austenized or hardened.
This also means that the total HAZ will be slightly less than a distance of 3 mm below the
surface, which will satisfy the hardening requirements for a maximum permissible total
case depth of 3.5 mm.
It should be noted that it is typically specified to avoid forming a combination of mar­
tensitic and nonmartensitic products in as-quenched structures. However, as always in
life, there are some exceptions. For example, the user might specify upon quenching to
have a certain amount of RA in predominately martensitic structures as it is in the case of
quenching and partitioning (Q&P) processes proposed by J.G. Speer (Colorado School of
Mines) and associates.
Since the heating time in induction surface hardening is commonly quite short, the δ is
the major factor in obtaining a certain depth of austenization at the surface of the work­
piece (O.D. or I.D. surfaces).
In order to overcome the complexity of applying the results of the Grossmann’s harden­
ability test for surface hardening, the Jominy end-quench test was developed. ASTM A255
(Standard Test Methods for Determining Hardenability of Steels) provides the guidelines/
procedures for performing the Jominy test.
Being one of the simplest and the least expensive, a Jominy end-quench hardenability
test has become one of the most popular techniques used by induction heat treaters for the
determination of steel hardenability.
According to the Jominy end-quench test, a solid cylindrical specimen (25 mm diameter
and 100 mm long) is uniformly heated to ensure the condition of homogeneous austenite
and then spray quenched from one end (Figure 4.24). As a result, a longitudinal hardness
distribution as a function of the distance from the quenched end is obtained.
Obviously, the Jominy end-quench test is a much simpler technique compared to the
more cumbersome Grossmann’s hardenability test. It is also more precise than the CE
indicator and it is also easier to apply.
Similar to the steel’s ideal critical diameter variations (Dimin and Dimax ) used in the
Grossmann hardenability tests, variations take place in the Jominy hardenability curves.
As an example, Figure 4.25 shows the hardenability limits (also called hardenability band)
for SAE 4150H steel. The existence of the hardenability band is caused by extreme varia­
tion of chemical composition as well as by the steel’s manufacturing specifics (e.g., deoxi­
dation practice, melting specifics, etc.).
In surface hardening, the existence of the colder core could have a measurable impact on
the cooling conditions during quenching. In many surface hardening applications, the use
of short heat times, significant “workpiece diameter–to–case depth” ratio, and pronounced
178 Handbook of Induction Heating
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

Distance from quenched end


Hardenability
curve

Jominy
bar

Hardness

Quench spray

FIGURE 4.24
Sketch of Jominy end-quench test and an example of the hardenability curve.

80

70
Upper
Hardness, HRC

limit
60

50 Lower Hardenability
limit band
40 SAE 4150H

30
0 4 8 12 16 20 24 28 32
Distance from quenched end, 1/16 in.

FIGURE 4.25
Hardenability limits (hardenability band) for SAE 4150H steel. (Based on materials presented in Heat Treater’s
Guide: Practices and Procedures of Irons and Steels, ASM Int’l, Materials Park, OH, 1999.)

skin effect do not permit the temperature of the core to rise significantly during the heating
stage. Therefore, upon quenching, the cold core provides an additional cooling effect (act­
ing as a thermal cold sink), resulting in more severe quenching conditions compared to the
cooling intensity of the through-heated parts used in standard forms of both Grossman’s
hardenability test and the Jominy hardenability test. A more intensive cooling rate may
increase the steel hardenability and affect other parameters including the formation and
distribution of residual stresses, specifics of austenite transformation, and so on.

4.1.5.3 Summary of the Limitations of Standard Forms of Hardenability Tests


Unfortunately, the majority of techniques for measurement of hardenability should be
applied in induction hardening applications (particularly in surface hardening) with a
Heat Treatment by Induction 179

great deal of caution owing to the assumptions and conditions that take place during the
measurements. Some of those limitations are outlined below.
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

• According to standard hardenability tests, an entire specimen is heated to austen­


itic temperature and held at that temperature for a sufficiently long time, assuring
that a homogeneous austenite has been formed. However, in the case of induction
surface hardening when rapid heating has been used without any or with very
short holding time, the structure of the austenite and its homogenization might
differ to an appreciable extent from austenite formed on slow heating with a suf­
ficiently long holding time. This variation can cause some discrepancies compared
to standard hardenability curves owing to subtleties in the short-time austenite
transformation.
• The fact that heating during induction hardening is very intensive shifts the Ac3
critical curve toward higher temperatures (this phenomenon will be discussed in
Section 4.1.6). Quenching of specimens from temperatures that are often 100°C to
180°C higher than the temperature used in the development of hardenability tests
can also lead to certain discrepancies in the prediction of the hardenability curves
and hardness patterns.
• Immersion quenching is used during experiments in developing data for the
Grossmann’s hardenability tests. However, spray quenching is used in the major­
ity of induction hardening applications. There is a considerable difference between
spray quenching and immersion quenching [240]. The differences are both quan­
titative and qualitative, and include, but are not limited to, specifics of film forma­
tion and heat transfer through the vapor blanket during the initial stage (A) of
quenching and the kinetics of formation, growth, and removal of bubbles from the
surface of the heated component during nucleate boiling (Stage B) [17,240].
Because of the nature of spray quenching, Stages A and B are greatly sup­
pressed in time, while cooling during the convection stage (C) is noticeably more
intense, compared with the process represented by the classical cooling curves of
heat convection. Also, the thickness of the vapor blanket film during Stage A is
typically much thinner during spray quenching than when the part is submerged
in a quench tank, and depends on flow rate, impingement angle, part rotation, and
other parameters. This vapor film is unstable and could be frequently ruptured,
intensifying the cooling severity.
In addition, the transition between Stages A and B is smoother with spray
quenching than that illustrated by classical cooling curves for immersion quench­
ing. During nucleate boiling (Stage B), bubbles are smaller because they have less
time to grow. Much larger numbers of bubbles form during spray quenching and
the intensity with which they remove heat from the surface of the component is
substantially greater compared with immersion quenching [17,240].
Unfortunately, those subtleties are seldom discussed in publications devoted to
induction heat treatment, causing some confusions and misinterpretations. Section
4.2.7 provides more discussion on induction quenching.
• In its standard form, the Jominy end-quench test is primarily suitable for moderate
cooling rates and can provide noticeably misleading results for severe cooling rates.
• The presence of the cold core that often can have a measurable effect on the inten­
sity of cooling does make it possible in some cases to have “self-quenching” (also
180 Handbook of Induction Heating

known as a “mass quenching”), which allows the elimination of liquid quen­


chants. In surface hardening, the ability of the cold core to increase the severity of
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

the cooling during quenching is an important factor that might result in modifica­
tion of conventional hardenability curves.
• Experiments for developing hardenability tests are commonly conducted using
cylindrical-shaped specimens and have limited ability to be extended to square
cross sections. Unfortunately, these results cannot be easily transferred to such
geometries as rectangular, conical, and, complex geometries such as gears and the
like.

Taking into consideration the abovementioned features and in order to increase the accu­
racy of predicting a material’s hardenability in induction surface hardening, nonstandard
forms of the Grossmann’s hardenability or Jominy end-quench tests may be conducted.
These tests require the creation of certain temperature gradients before quenching and are
better suited to induction surface hardening needs. However, those tests are cumbersome,
complex, and more expensive/time-consuming and still cannot provide universal recom­
mendations. Because of these factors, nonstandard hardenability tests are not widely con­
ducted and, therefore, their results are not readily available in the literature.
As a short summary on hardenability, it should be mentioned that the data obtained
from standard Jominy or Grossmann tests should primarily be used for estimation pur­
poses only. Nevertheless, those hardenability curves provide useful information as they
“convey” important knowledge regarding achievability of certain hardness depths and
patterns.

4.1.6 Effect of Heat Intensity (Heating Rate) and Prior Structure


on Induction Heat Treatment Results in Steels
A tremendous amount of metallurgical research has been directed at the determination
of the effect of heat intensities and prior microstructures on steel’s response to induction
hardening. As mentioned earlier, a Fe–Fe3C phase transformation diagram (Figure 4.3)
is a metastable diagram and only valid for equilibrium conditions. In case of hardening
plain carbon steels, when iron is alloyed with different percentages of carbon, some prac­
titioners make an attempt to determine the critical temperatures using this diagram or,
for low-alloy steels, by using corresponding diagrams or mathematical correlations (i.e.,
formulas or equations) that indicate the effect of certain additional chemical elements on
the positioning of critical temperatures such as Equations 4.5 through 4.10. Unfortunately,
those diagrams and correlations might be misleading in the majority of induction harden­
ing applications. Transformations taking place at A1, A2, A3, and Acm critical temperatures
are diffusion-driven transformations and rapid heating permits less time for diffusion.
The intensity of heating (heating rate) in induction hardening applications often exceeds
200°C/s (Tables 4.3 and 4.4) and, in some cases, reaches 800°C/s and even higher (e.g.,
in simultaneous dual-frequency gear hardening). Therefore, such a process cannot, by
any means, be considered as equilibrium and thermal hysteresis is highly pronounced.
Actually, with induction hardening, we always deal with a certain degree of nonequilib­
rium even with seemingly slow heating.
Rapid heating considerably affects the kinetics of austenite formation, shifting it toward
higher temperatures according to the continuous heating transformation (CHT) diagrams.
As stated earlier, practically all of the carbon, which is present in the initial ferritic–­pearlitic
Heat Treatment by Induction 181

structure, for example, is contained in the pearlite. Therefore, regardless of relatively


high solubility of carbon in austenite, some minimum time is required during solid-state
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

transformation (such as austenization) for carbon to be diffused from pearlitic regions


into areas occupied by ferrite. Therefore, it is required to create conditions conducive to
the needed diffusion-driven processes in order to develop an essentially homogeneous
austenitic structure with sufficiently uniform carbon distribution before quenching that
is desirable for the great majority of heat-treating applications [1,4,17,19,53,157,190,193,237–
246,261]. Therefore, rapid heating causes a measurable increase of critical temperatures
(e.g., Ac1, and Ac3) compared to the corresponding temperatures A1 and A3 indicated by the
equilibrium Fe–Fe3C diagram. Selection of hardening temperatures based on the equilib­
rium Fe–Fe3C diagram and failure to choose the proper hardening temperatures taking
into consideration the appreciably nonequilibrium nature of induction heating result in
incomplete transformation (Figure 4.26a).
If an austenite exhibits a nonuniform distribution of carbon, then, upon quenching, a
decomposition of heterogeneous austenite begins in the lower-carbon regions shifting the
CCT curves there to the “left” compared to steels with nominal carbon content and result­
ing in greater probability of forming upper transformation products. The CCT curves for
regions having excessive amounts of carbon will be shifted in the opposite direction with
a corresponding reduction of Ms temperatures and greater probability of having a larger
amount of localized RA [238,239].
Despite the fact that austenitic transformation takes place, its markedly heterogeneous
nature might potentially lead to an unacceptably heterogeneous as-quenched microstruc­
ture. It is important to be aware that in contrast to conventional heat treatment processes
that apply relatively long heat times, as a result of rapid induction heating, a certain amount
of undissolved carbides and carbonitrides as well as various concentration gradients of not
only carbon but also alloying elements may be present in austenite. Keep in mind that
complex carbides of alloy metals often dissolve noticeably slower than Fe3C.
Metallographic evaluation helps reveal the presence of “ghost pearlite” and other upper
transformation products in the as-quenched specimens that are associated with not fully
transformed structures or with the presence of severely heterogeneous austenite before
quenching.
The degree of heterogeneity in the microstructure of the as-quenched component can be
reduced by increasing the hardening temperatures and by lengthening the holding time
at the austenite phase temperature range.

(a) (b)

FIGURE 4.26
Failure to choose proper hardening temperatures based on the nonequilibrium nature of induction heating
can produce mixed structures after quenching of heterogeneous austenite. 4% Nital etched, ×100. Two images
(a) and (b) of heterogenous as-quenched structures. Light areas represent martensitic regions and dark areas
are pearlitic regions.
182 Handbook of Induction Heating

A number of studies have been conducted to quantify the effect of heat intensities and
prior microstructures on the shift of critical temperatures and the ability to form satisfac­
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

torily homogeneous austenite. As an example, Figure 4.27 shows the effect of heating rate
on the Ac3 critical temperature of medium-carbon steel having different prior microstruc­
tures [157,237].
The inability of the equilibrium Fe–Fe3C diagram (as well as conventional mathematical
correlations, such as Equations 4.5 through 4.10, for determining critical temperatures) to
take into account heating intensity markedly limits its application for predicting the tem­
peratures required for induction hardening.
A comprehensive study regarding CHT diagrams in steels and the correlation of heat
intensity versus positioning of Ac1, Ac2, Ac3, and Acm critical temperatures, as well as the
ability to obtain homogeneous austenite was conducted by J. Orlich, A. Rose, and col­
leagues at the Max-Planck-Institut fur Eisenforschung GmbH, Dusseldorf, Germany
[241,242]. Atlases for a variety of steels were developed as a result of their study, taking
into consideration a wide range of heating rates (from 0.05°C/s to 2400°C/s). These atlases
consist of numerous diagrams that were obtained by conducting laboratory experiments.
The results presented in the CHT diagrams are more appropriate for rapid induction hard­
ening in making decisions regarding the appropriate hardening temperatures. The study
shows that differences in the Ac3 critical temperatures (rapid heating vs. equilibrium con­
ditions) could be quite high; exceeding 150°C.
An interesting observation [245] can be made from the data experimentally obtained
by J. Orlich, A. Rose, and colleagues [241,242] that are particularly applicable to induction
hardening. As an example, Figure 4.28 illustrates the effect of rapid heating on positioning
and order of critical temperatures when heating medium-carbon steel SAE 1053 (see Orlich
et al. [242], p. 114). When heat intensities exceed approximately 20°C/s (which is very typi­
cal for the great majority of induction hardening applications), instead of the normal order
of critical temperatures, Ac1, Ac2, Ac3, rapid heating can switch the order to Ac2, Ac1, Ac3. This

1100

Annealed steel
1050
Normalized steel

1000 Quenched and tempered


Temperature, °C

950

900

850

800
100 1000
Rate of heating, °C/s

FIGURE 4.27
Effect of initial microstructure and heating rate on Ac3 critical temperature for SAE 1024 steel. (From S.L.
Semiatin, D.E. Stutz, Induction Heat Treatment of Steel, ASM International, Materials Park, OH, 1986, p. 88;
W. Feuerstein, W. Smith, Trans. ASM, 46: 1270, 1954.)
Heat Treatment by Induction 183

Zeit-temperature-austenitisitiung-scheubild
(kantinui)
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

Aufneigerchwindigicit in °C/s
2400 1000 300 100 30 10 3 1 0.22 0.05
1300 Ausgange- Ausgange- Warmebehandling der
zustard gefuge Ausgangezustandes

windigicit-
1 700°C sh/ofen
gefuge

1200

1100
_homogener + Austenit
Temperature in °C

1000

_inhomogener + Austenit

AC3
900

800 Ferrite + Austenit + karbite


AC1

AC2
Ferrite + karbite
700
0.1 1 10 102 104 106 108
Zeit in s

FIGURE 4.28
Effect of rapid heating on positioning and order of critical temperatures when hardening medium-carbon steel
1053. (From Orlich et al., Copyright 1973 Verlag Stahleisen GmbH, Düsseldorf, Germany, p. 114.)

might be essential knowledge in some subcritical and intercritical heat treat applications,
turning an easy job into an almost impossible job. On the basis of the information pro­
vided by Orlich et al. [241,242], there are many steels that exhibit such behavior.
As expected, excessive increase in hardening temperatures is associated with grain
coarsening. Figure 4.29 shows the influence of the heating rate on grain size growth of
medium-carbon steel (SAE 1040) [15].
The presence of significant microstructural and chemical segregations in the prior
microstructure (also called structure of parent material, structure of “green” part, or initial
structure) could also have a noticeable effect on the degree of heterogeneity of austenite
after rapid heating as well as on the results of the heat treatment and a selection of the
process parameters.
184 Handbook of Induction Heating

28
40°C/s
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

24
Conventional furnace

Austenite grain size, µm


hardening at 840°C 50°C/s
20

75°C/s
16

12
130°C/s

4
750 800 850 900 950 1000 1050
Temperature, °C

FIGURE 4.29
Influence of the heating rate on austenitic grain size of SAE 1040 steel. (From G.F. Golovin and M.M. Zamjatin,
High-Frequency Induction Heat Treating, Mashinostroenie Press, St. Petersburg, Russia, 1990.)

As can be concluded from the discussion above, in addition to heat intensity, the prior
microstructure also makes a measurable impact on the selection of hardening temperatures,
and its effect can be summarized in the following main points [1,17–19,238–​240,248,252–256]:

• Homogeneous fine-grain Q&T initial microstructures with a hardness range com­


monly being approximately 28 to 36 HRC are the most favorable for rapid IHT,
ensuring fast transformation, which allows a reduction in the hardening temper­
atures compared to alternative initial structures. The nucleation and growth of
austenite in the rapid heating of measurably heterogeneous initial structures are
more concerning than in case of conventional austenizing, which applies longer
process times.
• Q&T structures result in a consistent response to induction hardening with mini­
mum amounts of grain growth, shape/size distortion, and required heating energy.
These prior structures are associated with a well-defined (crisp) pattern with a short
transition zone and may also produce slightly higher than expected hardness levels
(when hardening medium-carbon steels) and deeper case depths, and form greater
compressive residual surface stresses compared with other types of initial struc­
tures. As an example, Figure 4.30 shows the effect of three different initial micro­
structures on an SAE 1070 carbon steel bar in response to surface hardening [4,145].
• Normalized structures consisting of a uniformly distributed fine-grain mixture of
ferrite and pearlite also provide a rapid response to induction hardening, allowing
one to reduce the required hardening temperatures and times almost as much as
Q&T prior structures and resulting in fast and consistent transformations.
• If the steel’s prior structure has a significant amount of coarse ferrites, clusters or
thick bands of ferrites, then the structure cannot be considered “favorable.” As dis­
cussed earlier, ferrite does not contain a sufficient amount of carbon required for
martensitic transformation (practically speaking). During austenitization, large
areas (clusters or bands) of ferrite require noticeably longer times for carbon to dif­
fuse into carbon-depleted areas. These ferrite clusters and bands could be retained
Heat Treatment by Induction 185

0 mm 0.5 mm 1 mm 1.5 mm 2 mm 2.5 mm


70
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

Quenched and
60
tempered
50
Hardness, HRC 40 Normalized

30

20 Annealed
10

0
0 0.02 0.04 0.06 0.08 0.1
Distance below surface, in.
Surface

FIGURE 4.30
Effect of initial microstructure in SAE 1070 steel bars in response to surface hardening using a 450-kHz induc­
tion generator operated at a power density of approximately 2.5 kW/cm 2 (15.9 kW/in.2). (From S.L. Semiatin,
D.E.  Stutz, Induction Heat Treatment of Steel, ASM International, Materials Park, OH, 1986; T. Spencer et al.,
Induction Hardening and Tempering, ASM International, Materials Park, OH, 1964.)

in the austenite upon rapid heating [1,17–19,238–240], and after quenching, mixed
structures can be formed. These mixed as-quenched structures (Figure 4.26b) are
not typically permitted by customer specifications, because scattered soft and
hard spots and poor mechanical properties (e.g., low fatigue resistance and poor
wear resistance, just to name a few) are associated with such structures. It should
be noted that high concentration of P, Si, Mn, and N promotes banding [33].
Noticeably higher temperatures and longer heating/holding times are needed,
as an attempt to avoid the formation of these undesirable as-hardened mixed
structures. Therefore, severely banded initial microstructures of “green” parts
exhibiting severe chemical and microstructural segregation should be avoided
in short-time induction hardening (Figure 4.31a); otherwise, higher temperatures
and longer times at the austenite phase temperature range are needed, result­
ing in a lower production rate, reduced useful compressive surface residual
stresses, and potential grain coarsening, and may also negatively affect distortion

(a) (b)

FIGURE 4.31
Examples of banded initial microstructures of “green” parts. (a) Severely banded initial structure exposing
substantial chemical and microstructural segregation should be avoided in short-time induction hardening.
(b) Modest segregation and banding might be unavoidable in modern steels.
186 Handbook of Induction Heating

characteristics. It is certainly understandable that, practically speaking, modern


steels almost always consist of some degree of heterogeneity, segregation, and
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

banding (Figure 4.31b). However, avoiding severely segregated and banded prior


microstructures in rapid induction hardening is strongly suggested.
• Fully annealed and spheroidized structures (Figure 4.32a) and other initial struc­
tures that consist of coarse stable carbides also have a poor response to short-time
hardening, requiring prolonged heat cycles, noticeably higher temperatures, and
extended times at elevated temperature to complete austenitization.
Otherwise, data scatter, a combination of soft-and-hard regions, may occur. It
should be noted that regardless of achieving seemingly high hardening tempera­
tures, some stable carbides (e.g., titanium carbides TiC), nitrides, and carbonitrides
might not dissolve and remain within the austenite and respectively will be pres­
ent in the as-quenched structure of the martensite matrix.
As expected, the necessity of using higher temperatures (compared to those
that would be conventionally employed in induction hardening of friendlier struc­
tures) and longer duration at those temperatures needed to dissolve the coarse
spheroidal carbides provoke an excessive grain growth, the formation of brittle
martensite, extended HAZs, excessive oxidation, and greater amounts of RA, and
potentially worsen the shape distortion. Factors like these can also have a detri­
mental impact on toughness, impact strength, bending fatigue strength, and wear
properties [248].
However, as always in life, there are some exceptions. For example, because
of specifics of flash processing/hardening (work by Gary M. Cola, Jr. and others
[102]), it is required that initial microstructure of steel should be spheroidized with
Cr-enriched carbides. Upon rapid induction heating and intense quenching, the
obtained structure consists of the so-called Flash® bainite and carbide networking
within the martensitic structure. The hardness of such structure may exceed the
hardness of fully martensitic structures. However, this unique application and
necessity of having a spheroidized initial structure before induction hardening is
more of the exception than the rule.

Decarburized layer

(a) (b)

FIGURE 4.32
(a) Spheroidized structures and other initial structures that consist of coarse stable carbides also have poor
response to rapid induction hardening requiring higher temperatures and extended times at elevated tem­
perature to complete austenitization. (b) Decarburized layer present in the initial microstructure remained in
as-quenched sample (thin white color region indicates a free-ferrite region) (400×).
Heat Treatment by Induction 187

• Recommendations suggested in the literature including Refs. [4,237,241,242] for


a selection of induction hardening temperatures as functions of heat intensities
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

should still be used only as basic guidelines. This is particularly true for modern
plain carbon steels, which are often melted using a large percentage of scrap that
may contain microalloyed high-strength low-alloy steels. Thus, the steel may still
be considered to be nominally of the plain carbon steel composition but may also
contain trace amounts of niobium, vanadium, titanium, boron, and other elements
whose presence may noticeably affect the response of steel to heat treatment using
electromagnetic induction. For this reason, it is wise to check the complete chem­
istry on each lot of steel and to determine the proper austenitizing temperatures
experimentally.
It is important to remember about the effect of prior microstructure not only on
the positioning of critical temperatures during rapid heating but also on the ther­
mal and, in particular, electromagnetic properties of a steel of a particular grade.
This factor alone can cause the need to adjust the process recipe.
• Ref. [253] reveals the results of a study related to the response to induction hard­
ening of an SAE 10V45 steel shaft using three conditions: as-hot-rolled, 18% cold-
drawn, and 29% cold-drawn. Three nominal case depths (2, 4, and 6 mm) were
analyzed. It has been reported that induction hardening of the cold-drawn speci­
mens appeared to increase the case hardness compared to the as-hot-rolled SAE
10V45 steel, particularly for hardness case depths of 2 and 4 mm. It was postulated
that the microstructural refinement produced by cold working that was associated
with reduced ferrite grain size was responsible for this phenomenon. Some frag­
mentation and reduction in the spacing of the pearlitic cementite lamellae caused
by cold work might have also contributed to it. It was also suggested that both fac­
tors could potentially enhance the formation of austenite and carbide dissolution,
leading to a higher average carbon content of the austenite that is associated with
a higher carbon martensite and respective higher hardness. As expected, prior
cold working also resulted in a noticeably higher core hardness compared to the
as-hot-worked steel of the same grade.
• As can be expected from the above discussion, in contrast to conventional heat
treatment processes that apply relatively long heat times, a variation in prior
microstructure in combination with rapid heating nature of electromagnetic
induction hardening may produce heterogeneous austenite. Such structure may
include undissolved carbides and carbonitrides as well as various concentration
gradients of not only carbon but also alloying elements.
• Decarburization (depletion of the surface carbon) indicates the loss of surface and
near-surface carbon content from the steel as a gaseous phase after being exposed
to sufficiently high temperature (e.g., austenite phase temperature range) in an
oxidizing atmosphere (such as an air) [29,30,33,258,259]. Carbon monoxide (CO)
and carbon dioxide (CO2) gases are formed during steel surface oxidation, escap­
ing into the atmosphere and leading to a depletion of the surface carbon. This
results in the formation of a surface–subsurface carbon gradient that promotes
carbon diffusion from subsurface regions toward the surface and the creation
of a subsurface layer with noticeably reduced carbon content (decarburized sur­
face layer). The depth of the decarburized layer with depleted carbon content is a
complex function of several factors including the temperature, time at high tem­
perature, chemical composition of steel and surrounding atmosphere, moisture,
188 Handbook of Induction Heating

the counterbalancing effects of the rate of decarburization and thickness of the


decarb layer, and its removal as a result of scale formation (as it occurs in hot
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

working applications).
Regardless of the fact that decarburization is a time–temperature-dependent
(diffusion driven) process, temperature has the largest impact. Besides the time
and temperature, the extent of decarburization is also a function of steel’s alloying
compositions (e.g., Cr, Ni, and Mn restrict decarburization, but Si worsens [169]).
The presence of decarburized layers at the surfaces of “green” parts is a problem
in some heat treat practices since decarburized layers result in considerably dif­
ferent mechanical properties near the surfaces. A substantial amount of decarb is
commonly present when dealing with hot rolled bar stock, as well as as-forged,
annealed, and spheroidized structures. Because of the short process time, induc­
tion hardening itself does not typically produce a measurable amount of decarb.
Incoming decarb cannot be repaired during rapid induction hardening under nor­
mal processing conditions.
Since the carbon content determines the achievable steel’s hardness level, sur­
face decarburization causes a “soft” surface that can lead to localized severe hard­
ness and strength reduction in as-quenched steel and it even may reverse the
residual stress distribution, forming undesirable localized tensile surface stresses,
which in turn can cause premature failure of components during their service life.
Decarburized layers that are present in the initial microstructure will remain
within the as-quenched structure (Figure 4.32b), making a considerable negative
impact on other mechanical properties, including, wear resistance, fatigue, impact
and bending properties, and others.
Thus, if the prior microstructure of green parts exhibits the presence of decarb,
then the weakened near-surface material should be machined off. Metallographic
examination is the most popular way to determine the presence and extent of
decarburization.
Thus, decarburization, as well as some other conditions resulting from the steel
prior processing (e.g., seams, laps), excessive oxide scale buildup at the surface,
intergranular oxidation, severe grain coarsening, and burns should be avoided.

4.1.7 Super-Hardness Phenomenon
When surface hardening steels, a combination of using “friendly” prior microstructures,
fast heating rates, and intense quenching could result in the so-called super-hardness phe­
nomenon or super hardening [1,15,18,190,248,249]. This phenomenon refers to obtaining
greater hardness levels in the case of surface hardening compared to through hardening
or hardness levels that would normally be expected. Figure 4.33 illustrates a comparison
of typical hardness profiles along the workpiece radius (in the case of a cylindrical body) or
half-thickness (for flat parts) after induction surface hardening versus through hardening.
As can be seen in Figure 4.33, the surface hardness of case-hardened components can be
slightly higher compared with through-hardened parts. Because of this phenomenon, for
identical steel composition, the surface hardness of an induction case-hardened part could
be 2–4 HRC (1–3 HRC being more typical) higher than normally expected for a given car­
bon content for through-hardened steel [190,248,249].
The super-hardness phenomenon is not clearly understood, and its origin has not been
established or widely accepted by metallurgists worldwide. However, it has been obtained
Heat Treatment by Induction 189

Surface Core
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

Through hardening

Hardness
Surface (case)
hardening

Radius/thickness

FIGURE 4.33
Typical hardness profiles after surface hardening and through hardening.

experimentally on numerous occasions and several interpretations have been offered.


Possibly, this phenomenon can be attributed to a combination of several factors including
the magnitude of residual stresses and final microstructure (shape and density of carbides,
RA, dislocation density of the martensitic matrix, grain size, and others) [190,248,249]. The
short heating cycle in induction surface hardening results in fine austenitic grain sizes that
occur because the steel typically is at the austenitizing temperature for a short time, reduc­
ing the possibility of grain growth.
Another factor that might contribute to super hardness is the higher lattice strain from
significant residual compressive stresses at the surface of the part when its internal regions
and, in particular, the core remain at a lower temperature or at even room temperature
[248,249].
The existence of a cold core that “acts” as a thermal sink has a noticeable impact on the
severity of quenching and may also contribute to the existence of super hardness. In induc­
tion case hardening applications, the core temperature might not rise significantly owing to
a pronounced concentration of electromagnetic energy within the surface layer, high heat
intensity, and short heating time. Cooling intensity during the quenching stage depends
on the severity of surface cooling by the applied quenchant and internal cooling as a result
of thermal conduction from the hot surface toward the colder core complementing the
intensity of surface quenching. In contrast to surface hardening, hot internal areas of the
through-heated components result in thermal support counteracting the surface region that
is being quenched and reducing the quench severity of surface and subsurface areas.
Another factor that might also contribute to the presence of the super-hardness phenom­
enon is related to the formation of heterogeneous austenite associated with the short aus­
tenitization [190]. Heterogeneous austenite with a concentration gradient and a nonuniform
distribution of carbon in combination with undissolved fine carbides network (including
complex carbides of alloying elements) may produce a composite effect. Formation of high-
carbon martensite might lead to a hardness increase as well.
Prior microstructure has a marked effect on the appearance of the super-hardness phe­
nomenon. Fine-grain homogenous normalized prior structures as well as Q&T structures
have a better chance of exhibiting super hardness [249].
There is a greater probability of super hardness to occur in steels with a carbon con­
tent of 0.35% to 0.65% C having case depths of less than 3 mm and heat times of less
than 4 s. These conditions are more prone to exhibiting the super-hardness phenomenon.
As-annealed and as-spheroidized prior microstructures are not amenable to surface super
hardness [248].
190 Handbook of Induction Heating

4.1.8 Inclusions
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

Practically all commercial steels contain a certain amount of nonmetallic particles com­monly
referred to as inclusions, which are a product of steelmaking specifics [28–30,159,259,269–279].
ASTM E45-13 governs procedures for determining the nonmetallic inclusion content in
wrought steels. The origin of inclusions, their morphology, stability, type, size, distribu­
tion, and chemical composition are often used as a measure of the steel’s quality and its
purity. Inclusions may or may not have a noticeable impact on the IHT and performance of
the component. For example, a network of fine uniformly distributed inclusions acting as
nuclei sites for precipitation of carbides and nitrides might be beneficial for constraining
grain coursing. On the other hand, localized and preferentially oriented coarse platelet
inclusions can act as sites for crack initiation, negatively affecting notch sensitivity and
crack propagation rate (Figure 4.34).
On the basis of their origin, inclusions can be categorized as endogenous (also called
as indigenous) or exogenous [30,159]. The former are typically smaller in size and greater
in number, and their appearance is associated with chemical reactions during steelmak­
ing and solidification and include oxides, sulfides, nitrides, silicates, phosphides, and so
on. Thus, those inclusions are chemical compounds of metals (e.g., iron, magnesium, and
titanium, just to name a few) and nonmetallic elements (e.g., oxygen, sulfur, and nitrogen).
In contrast, exogenous inclusions are much larger in size and smaller in numbers and
may occur only occasionally as loose particles from refractories, holding vessels for molten
steel, loose dirt, or other foreign or accidental inclusion materials that were in contact with
the liquid steel. Those inclusions are not often seen in modern steelmaking.
Depending on their stability, inclusions may dissolve upon heating to the austenite phase
temperature range, but in the majority of cases, they do not dissolve and are retained upon
quenching in the as-hardened structure. As an example, Figure 4.35a shows an initial
ferritic­–pearlitic microstructure of 38MnSiVS5 forged steel showing uniformly dispersed
fine global-shaped inclusions. Figure 4.35b reveals the presence of inclusions within the
as-quenched martensitic matrix.
Metallographic analysis allows differentiating between morphologically similar inclu­
sions. Classic JK chart (developed by Jernkontorent, Sweden) and its modern modifications
help to categorize nonmetallic inclusions in steels based on their appearance.
While discussing inclusions, it would be beneficial to briefly mention the MnS inclusions
(Figure 4.34), which are the most common inclusions that heat treaters come across. This is

Elongated MnS inclusions

FIGURE 4.34
Elongated lined-up platelet MnS inclusions (also called stringers or streaks) within steel’s ferritic–pearlitic
microstructure.
Heat Treatment by Induction 191

27–28 HRC 0.100 mm


Pearlite
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

0.100 mm

Inclusions

Ferrite
57–59 HRC
(a) (b)
Prior microstructure of 38MnSiVS5 forged steel showing the Hardness case microstructure of induction surface hardened
presence of inclusions in ferritic (in the grain boundaries)— 38MnSiVS5 forged steel revealing martensitic matrix with
pearlitic (dark etched areas) structure (200×) the presence of small globular inclusions (200×)

FIGURE 4.35
Depending on their stability, inclusions may be dissolved upon heating to the austenitic temperature range, but
in many cases, they remained after quenching in the as-hardened structure. (a) Illustrates an initial ferritic–
pearlitic microstructure of 38MnSiVS5 forged steel showing uniformly dispersed fine global-shaped inclusions.
(b) Reveals the presence of inclusions within the as-quenched martensitic matrix.

so, because any commercial carbon steel used in hot working consists of a certain amount
of MnS. As discussed in Section 4.1.1, a limited amount of manganese plays an important
role in plain carbon steels, since it prevents or minimizes the formation of iron sulfides
(FeS) that could concentrate at the grain boundaries, worsening the steel brittleness and
sensitivity to intergranular cracking. Besides being very hard and brittle, FeS also has a
relatively low melting temperature.
When an adequate amount of manganese is present in the steel, instead of formation
of FeS, sulfur combines with the manganese to form MnS, creating a less crack-prone
structure.
It should also be mentioned that heating of steel to sufficiently high temperatures within
the austenitic range might cause dissolving or partial dissolving MnS inclusions and pref­
erential segregation of Mn and S at austenitic grain boundaries. This lowers the melt­
ing point in the grain boundary region compared to the nominal solidus temperature of
the steel, leading to several undesirable phenomena during and after hot working and
hardening (e.g., during cooling) and potentially promoting a temper embrittlement with
increased manganese content.
The morphology of inclusions makes a marked impact on the tendency to cracking dur­
ing steel hardening. Uniformly distributed fine inclusions of global or polyhedral shape
are preferable. In contrast, coarse platelet inclusions are potentially more susceptible to
cracking.
In the majority of induction hardening applications, the presence of coarse elongated
near-surface platelet-shaped inclusions is undesirable. This is particularly so in cases where
inclusions are lined up or have a preferential orientation. As an example, Figure  4.31b
reveals randomly located platelet MnS inclusions. In contrast, Figure 4.34 shows lined-
up elongated platelet MnS inclusions (also called stringers or streaks) within the steel’s
ferritic­–pearlitic microstructure. These inclusions may “act” as appreciable stress risers
that could potentially negatively affect certain mechanical properties of a component (e.g.,
fatigue strength) and measurable scatter in the fatigue data may occur.
A large number of coarse platelet-shaped inclusions may also have a harmful effect on
notch sensitivity, as well as on crack initiation and propagation rate owing to the presence
of stresses for a specimen being rapidly heated to high temperatures for austenitization
192 Handbook of Induction Heating

and during the quenching cycle, particularly if they are concentrated near geometrical
irregularities. In some cases, lined-up platelet shape inclusions may introduce noticeable
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

anisotropy and could potentially affect the localized eddy current flow when high fre­
quencies are used.

4.1.9 Grain Boundary Liquation (Incipient Melting)


As has been discussed in Section 4.1.6, even for plain carbon steels, because of the non­
equilibrium nature of rapid induction heating, critical temperatures are shifted to higher
temperatures compared to those indicated by the conventional Fe–Fe3C phase transforma­
tion diagram. This necessitates reaching noticeably higher temperatures in induction heat
treat operations in order to obtain appropriate metallurgical conditions. However, there
are some limits and excessively high temperatures should be avoided. There are several
metallurgical and microstructural phenomena associated with excessive temperatures.
This includes severe grain coarsening, scaling, intergranular oxidation, grain boundary
liquation, and others.
Section 4.1.4 reveals that residual elements in a limited amount can be found in any mod­
ern steel. The liquidus temperature of some residuals might be noticeably lower compared
to the expected solidus point of steel.
There is an increased use of electric arc furnaces that utilize recycled scrap, which may
be rich in Cu or other low–melting point residuals. Over the last decade, the amount of
copper residual has increased in most commercial steel grades. Copper has a relatively
low melting point (1085°C) and, as a residual element, is not readily oxidized and removed
from liquid steel in the steelmaking process using electric arc furnaces [30,187,188].
In hot working of steels, the rejection of the Cu from the iron oxide into the base metal
may create a near-surface copper-enriched zone at the FeO interface [30,33,196,268]. This
takes place because the solubility of copper in iron oxide FeO is very low, and at suf­
ficiently high temperatures, where the oxidation rate of iron is quite high, the copper is
liquid and can penetrate along grain boundaries with ease, dramatically weakening the
grain boundaries. Therefore, even in cases where the scale might be removed and a thin
surface layer is machined, there might still be areas that exhibit a localized copper surplus.
If an appreciable temperature surplus occurs at the surface during steel austenitization,
then besides the appearance of known unwanted metallurgical conditions and excessive grain
growth, a phenomenon of grain boundary liquation (also called incipient melting) may occur.
Overheating is one of the most common causes of cracking. It weakens grain structure
and greatly increases the steel’s brittleness and sensitivity to developing intergranular
cracking [1,280,281]. Grain boundary liquation is associated with liquation of low-melting
phases and impurities concentrated at grain boundaries, leading to a degradation of those
boundaries [30,33,196,268]. As an example, Figure 4.36 illustrates steel microstructures
exhibiting grain boundary liquation. It should be noted that each image has a different
magnification indicated by the scales positioned in the top left corner. A network (chains)
of liquated areas located at the grain boundaries is easily visible.
The phenomenon of grain boundary liquation can be amplified by the segregation of
manganese, sulfur, copper, and some other elements to the austenitic grain boundaries.
Both phosphorus and sulfur markedly affect the steel overheating.
As discussed in Section 4.1.8, heating of steel to sufficiently high temperatures within
the austenite phase range can be associated with partial dissolving of MnS inclusions and
might be accompanied by preferential segregation of Mn and S at the austenitic grain
boundaries. This lowers the melting point in the grain boundary region compared to the
Heat Treatment by Induction 193

~50 µm
10 µm
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

FIGURE 4.36
Examples of grain boundary liquation. Note: Each image has a different magnification indicated by scales posi­
tioned in the top left corners.

nominal solidus temperature of the steel. As reported in Ref. [168], the Fe–FeS eutectic
melts at 985°C. Reduction of both Mn and S levels is one possible solution to overheating.
However, reduction of sulfur alone might not produce the desirable result [30,33].
Ref. [282] suggests that overheating typically does not occur in steels with very low sul­
fur content (<0.002% S) that have an extremely low volume fraction of sulfides. However,
these steels are rarely used in industry because of their high cost. At the same time, steels
with 0.01%–0.02% S are more susceptible to overheating than steels with higher sulfur
content (>0.3% S).
The complexity of the influence of various elements on the propensity of a particular
steel grade to be overheated, developing a grain boundary liquation and eventual crack­
ing, can be illustrated using the following example. The study shows [283] that in SAE 1045
steel, the addition of 0.003% boron in the form of simple 12% boron alloy lowers the tem­
perature at which overheating typically occurs. In contrast, the addition of 0.0007% boron
in the form of a complex alloy containing a small amount of Al, Ti, and Zr in addition to
boron does not noticeably lower the overheating temperature of the steel.
There are several other metallurgical and microstructural phenomena associated with
steel overheating. This includes partial or complete sulfide spheroidization of the elon­
gated sulfide inclusions, grain-boundary sulfide precipitation, and severe grain coarsen­
ing. In the presence of tensile stress, the deteriorated grain boundaries can cause cracking,
resulting in a defect that is known as hot shortness (also referred to as red-shortness).
Different magnitudes of overheating cause various degrees of grain boundary liquation
and are commonly distinguished as slight, moderate, or severe. This designation is rela­
tively subjected and might be different for different corporate standards. As an example,
Figure 4.37 shows a side-by-side comparison of two SEM images of as-quenched struc­
tures. The image in Figure 4.37a shows a so-called clean structure, while the image in
Figure 4.37b illustrates moderately liquated grain boundaries. These definitions of what is
visually acceptable are always quite tricky. However, quantity measures based on area%
or maximum width or continuous length of liquated boundaries network are not included
into standards at the time of writing this book (2016).
In order to have a sound structure, it is recommended to avoid reaching very high tem­
peratures as well as any visible degree of grain boundary liquation.
As an attempt to make their machines more energy efficient with improved “miles per
gallon/liter” ratio, lightweight initiatives have been implemented by several industries
(including automotive, aerospace, off-road, and others) to meet more stringent federal
Corporate Average Fuel Economy regulations. These initiatives resulted in removing
excessive materials, leading to components with complex functionalities and containing
194 Handbook of Induction Heating
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

10 µm 10 µm

(a) (b)

FIGURE 4.37
Side-by-side comparison of two SEM images of as-quenched structures. (a) Shows a so-called clean structure,
while (b) illustrates moderately liquated grain boundaries.

FIGURE 4.38
Examples of complex geometry parts containing various holes, keyways, grooves, shoulders, flanges, diameter
changes, undercuts, splines, sharp corners, and other geometrical discontinuities.

numerous geometrical irregularities without compromising performance and safety


objectives. Certain design features could make a part become prone to localized heat sur­
plus upon induction heating. Typical examples are components and assemblies containing
longitudinal or transverse holes, keyways, grooves, shoulders, flanges, diameter changes,
undercuts, hollow areas, splines, sharp corners (Figures 2.8 and 4.38), and other geometri­
cal discontinuities and irregularities [142].
Nevertheless, such features are not unique as they are commonly found on many trans­
mission and engine components. Therefore, it might be unavoidable to have a certain
degree of grain boundary liquation when induction hardening complex geometries. In
cases like this, some corporate standards might allow a slight amount of grain boundary
liquation.
Severe grain boundary liquation must be avoided, because it is associated with extreme
brittleness and notch sensitivity, leading to failure caused by intergranular cracking
(Figure 4.39).

4.1.10 Specifics of Induction Hardening Stainless Steels and Bearing Steels


IHT of stainless steels and specialty bearing steels represents a much smaller market com­
pared to the heat treatment of plain carbon steels and low-alloy steels, yet it is an important
Heat Treatment by Induction 195
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

30 µm

FIGURE 4.39
Severe grain boundary liquation must be avoided because it is associated with extreme brittleness and notch
sensitivity leading to intergranular cracking.

sector of industry. Those steels offer several unique properties that include but are not lim­
ited to resistance to oxidation and scale formation at elevated temperatures and stain and
corrosion resistance while the material is being exposed to oxidizing atmospheres or an
aqueous, salty, or nitric acid environment [30,50,157,158,236,284–290].
It is a commonly accepted practice to classify stainless steels based on their microstruc­
ture into the following categories: ferritic, martensitic, duplex, precipitation-hardenable
(PH), and austenitic stainless steels. The first four categories exhibit ferromagnetic proper­
ties (to a different degree though); however, the last one (austenitic) is nonmagnetic.
Chromium is the major alloying element, and its minimum concentration typically
exceeds 11%. Other alloying elements include carbon, molybdenum, silicon, nickel, tung­
sten, vanadium, nitrogen, copper, and some others. Stainless steels should be considered
as at least ternary alloys (Fe–Cr–C). Alloying and residual elements can provide a com­
bined effect on the properties of a particular grade of stainless steel.
Some heat-treating applications for stainless steel products (e.g., “black,” “dull,” and
“bright” annealing of stainless steels) have been briefly reviewed in Section 2.1.4. This sec­
tion focuses on induction hardening applications of stainless steels.
Among all of the above-discussed categories, only martensitic stainless steels (MSS) and,
to a lesser degree, PH steels may be induction hardened, forming a martensitic structure in
the as-quenched condition. Typical applications for MSS include components for the aero­
space, chemical, petrochemical, and oil industries, as well as medical, pulp, and paper indus­
tries. Some examples of components would be steam, gas and jet blades, bearings, gears,
pump shafts, medical tools, knife blades, ball screws, and raceways, just to name a few.
The amount of chromium in MSS is typically within 11% to 18%. Higher chromium
content is accompanied by a higher carbon concentration. Carbon concentration in MSS
can vary to a noticeable extent. For example, carbon content of such MSS grades as ASTM
403, 410, and 416 is less than 0.15%. Some grades of MSS have medium carbon content
(e.g., grades 420 may have 0.35% C). In contrast, the carbon content of other grades may
be substantially higher. Stainless steels 440A (0.6%–0.75% C), 440B (0.75%–0.95% C), 440C,
and 440F (0.95%–1.2% C) could serve as typical examples of high carbon grades of MSS.
As expected, steels with higher carbon content can be hardened to higher hardness levels.
The PH stainless steels achieve the hardening effect as a result of precipitation of inter­
metallic compounds that are not carbides. The composition of certain martensitic PH
stainless steels (e.g., martensitic 17-4 PH steels) can produce martensite structures upon
rapid cooling [236], though achieved hardness levels are noticeably lower than commonly
used induction hardenable plain carbon steels, that is, within the 44 to 48 HRC range.
196 Handbook of Induction Heating

The induction hardening sequence of the induction hardenable stainless steel grades
comprises the same basic process stages as for plain carbon steels and low-alloy steels.
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

This includes austenitization, soaking/holding (often required within the austenite phase
region), and quenching; however, there are some subtleties.

4.1.10.1 Physical Properties
Electromagnetic and thermal properties of MSS are noticeably different compared to plain
carbon steels, measurably affecting the process recipe.
The electrical resistivity ρ of MSS is greater and relative magnetic permeability μr is
measurably lower compared to those of plain carbon steels with a similar carbon content.
For example, at room temperature, ρ is typically two- to threefold higher compared to the
respective values for corresponding plain carbon steels. Both factors (ρ and μr) produce a
greater depth of the heat generation δ for a given frequency. In addition to this, depending
on the particular grade of MSS, its Curie temperature can be 40°C–60°C lower compared to
plain carbon steels. This shortens the magnetic stage of the heating cycle.
MSS exhibit a similar behavior to carbon and low-alloy steels with respect to their criti­
cal temperatures (Ac1 and Ac3) being functions of the heat intensity. As an example, CHT
diagrams for 420A and 440C grades used in surgical tools have been published in Ref. [291].
Researchers conducted an experimental study (based on the dilatometric behavior of
alloys) while also applying computer modeling analysis. The results clearly reveal that
the critical temperatures of MSS are affected by the heat intensity, shifting them to higher
temperatures for greater heat intensities.
Though MSS have the highest thermal conductivity among other stainless steels, their
values can be 30%–45% lower than plain carbon. This means that there will be notice­
ably lesser heat transfer effect, potentially resulting in greater “surface-to-core” thermal
gradients during rapid induction heating, as well as the occurrence of a larger magnitude
of transient stresses and greater probability of crack development, suggesting the need to
apply lower heat intensities.

4.1.10.2 Hardening Specifics
The initial microstructure of MSS before induction hardening is often annealed (full, iso­
thermal, or process annealed) or martempered but never normalized.
MSS are more sensitive to a variation in chemical composition and to the repeatability
of the process recipe than plain carbon steels and low-alloy steels. This makes it markedly
more critical to ensure the ability to achieve precision of heating/quenching stages.
Somewhat longer austenizing times and higher temperatures are commonly needed
because of slow dissolution of chromium carbides during austenization. Depending on
a particular steel grade and application specifics, austenizing temperatures are typically
within the 925°C–1125°C range (higher temperatures are usually associated with grades
having a higher carbon content). Taking into consideration that longer times are required
for dissolving carbides as well as the effect of noticeably lower thermal conductivity, it is
advised that the intensity of heating should be lower compared to those intensities used
for hardening carbon steels.
MSS have much better hardenability than plain carbon steels and many low-alloy steels.
Therefore, forced air, oil, or aqueous polymer solutions can be used as quenching media.
Regardless of the fact that MSS can be air-cooled, liquid quenchants are typically used in
order to provide greater cooling intensities, allowing minimization or elimination carbide
Heat Treatment by Induction 197

precipitation during the cooling/transforming of austenite and obtaining higher surface


hardness and compressive residual stresses at the surface.
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

Table 4.9 shows typical hardening temperatures and expected as-quenched hardness levels
when heat treating selected grades of MSS (based on data published in Ref. [157]). According to
some corporate standards, in order to speed up diffusion-based processes during austenitiza­
tion, higher maximum temperatures than those shown in Table 4.9 might be applied.
Newly developed MSS occur quite regularly, offering certain application-oriented ben­
efits. For example, XD15NW (X40CrMoVN16-2) is an MSS grade that specifically targets
achieving high hardness in combination with high corrosion and fatigue resistance [288].
Table 4.10 shows the chemical composition of this complex alloy. Alloy XD15NW utilizes
the effect of a partial substitution of carbon with nitrogen. Nitrogen combined with chro­
mium and molybdenum helps improve the resistance to pitting corrosion. Molybdenum
and vanadium ensure a secondary hardening, replacing chromium in the precipitate.
Regardless of the medium carbon content of this alloy, the expected minimum hardness is
quite high being approximately 58 HRC [288].
Because of several factors, the as-quenched hardness of many MSS can be represented by a
bell-shaped curve: initially, hardness increases with a rise in austenizing temperatures, reach­
ing its maximum and then starting to decline again. As reported in Ref. [236], for some MSS
(e.g., grade 410), the use of too low or too high austenizing temperatures could manifest itself
in the presence of ferrites in the as-quenched structures. Of course, excessively high austeniz­
ing temperatures should be avoided since they produce coarse grains and a greater amount
of RA upon quenching to room temperature. At times, soaking­/holding times at austenite
phase temperatures may be needed. The as-quenched structure of many MSS and in par­
ticular grades of 440-series consists of a substantial amount of RA and often require using a
subzero cryogenic treatment followed by single or double tempering.

TABLE 4.9
Typical Hardening Temperatures and Expected As-Quenched Hardness Levels When Heat Treating
Selected Grades of Martensitic Stainless Steels
Stainless Steel Grades (SAE Classification)
403 410 414 416 420
Hardening temperatures (°C) 925–1010 925–1010 925–1050 925–1010 980–1065
Expected as-quenched 375–415 HB 375–415 HB 388–456 HB 375–415 HB 448–564 HB
hardness
422 431 440A 440B 440C
Hardening temperatures (°C) ≤1055 ≤1205 1010–1065 1010–1065 1010–1065
Expected as-quenched 45–50 HRC 401–444 HB 52–57 HRC 56–59 HRC 60–62 HRC
hardness
Source: Based on data published in Heat Treater’s Guide: Practices and Procedures of Irons and Steels, ASM Int’l,
Materials Park, OH, 1999.

TABLE 4.10
Chemical Composition of XD15NW (X40CrMoVN16-2) Martensitic Stainless Steel
C Si Mn Cr Mo V N Ni
Min 0.37 – – 15.00 1.50 0.20 0.16 –
Max 0.45 0.60 0.60 16.50 1.90 0.40 0.25 0.30
Source: XD15NW (X40CrMoVN16-2) A high hardness, corrosion and fatigue resistance martensitic grade,
Aubert & Duval, March, 2010.
198 Handbook of Induction Heating

The presence of elevated amounts of RA is associated with a stabilization phenomenon


that can be caused by an interrupted quenching of MSS. A stabilization phenomenon is not
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

exclusive to MSS and could also occur in hardening hypereutectoid steels (e.g., SAE 52100),
tool steels, and other alloys with sufficiently low Ms – Mf temperatures. This phenomenon
manifests itself in the formation of lower amounts of RA during uninterrupted quench­
ing. An interrupted quenching with a considerable quench delay at room or low-range
elevated temperatures can stabilize the austenite, exhibiting a greater amount of RA [236].
For example, as it was reported in Refs. [151,512] and other publications, if, after harden­
ing of SAE 52100 steel, it was held for different amounts of time at room temperature
before its subsequent cryogenic treatment (at −180°C), then there will be noticeably differ­
ent amounts of RA retained. Approximately 70% of the RA will be transformed if cryogenic
treatment was conducted within 5 min after hardening. Increased time delay (stabilizing
time) of cryogenic treatment for 40 min and for 50 h resulted in transformation of only 60%
and 30% of the RA, correspondingly. Other steels exhibit similar behavior.
Preheating might be beneficial as a way to reduce thermal shock during the initial stage
of heating. Component geometry and the hardness pattern specification could be deci­
sive factors affecting the process recipe (including a necessity for preheating). It is wise
to review suggestions published in Ref. [157] and the steel supplier’s information before
developing a process recipe for hardening a particular grade of MSS.
Induction hardening of specialty bearing steels has many similarities with MSS. Actually,
some MSS are used for a fabrication of bearings, raceways, ball screws, aerospace parts, and
other components. Advanced bearing steels were developed to further enhance certain
properties of MSS and high-carbon steels (such as SAE 52100). CRONIDUR-30 alloy, which
is Cr-Mo-N specialty steel (DIN 1.4108, developed by Energietechnik Essen GmbH, www​
.energietechnik-essen.de) can serve as an example of such steels. As-quenched hardness
can reach 57–58 HRC and hardness after cryogenic treatment (that is typically specified)
can be as high as 60 HRC. The permissible temperature range when surface hardening
those steels is quite tight, and in the case of CRONIDUR-30 in order to maximize achiev­
ing high hardness, austenizing temperatures are normally specified within a quite narrow
range of 1010°C–1040°C (with cryogenic treatment) or 980°C–1010°C (without cryogenic
treatment). Temperatures outside of this range result in lower hardness levels.

4.1.11 Induction Heat Treatment of Cast Irons


Up to this point, we have focused on selected aspects of heat treatment and principles
of hardening steels. The electromagnetic induction has also been successfully applied
for hardening of a variety of components (e.g., camshafts, crankshafts, sprockets, rollers,
rocker arms, crane wheels, connecting rods, etc.) fabricated using different grades of cast
irons offering numerous attractive properties and cost advantages [157,158,170,196,203–218].
As an example, Figure 4.40 shows the magnified view of a dual-spindle induction system
for hardening cast iron automotive camshafts.
Induction surface hardening of cast irons has many similarities with hardening of steels;
at the same time, there are specific features that should be taken into consideration. Some
of those features will be reviewed below.
The term cast iron does not represent one particular material but a large family of metal­
lic alloys that occupy the right side of the iron–iron carbide phase transformation diagram
(Figure 4.3) featuring the high carbon content region (2% and higher).
Generally speaking, the family of commercial cast irons can be categorized into six
groups: white, gray, malleable, ductile (also called nodular, spheroidal, or SG irons),
Heat Treatment by Induction 199
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

FIGURE 4.40
Zoomed-in view of dual-spindle induction system for hardening cast iron camshafts. (Courtesy of Inductoheat
Inc., an Inductotherm Group company.)

compacted graphite (CGI), and special high-alloy cast irons. It is important to remem­
ber that the first five groups of cast irons have a moderate amount of alloying elements,
whereas the amount of alloying in high-alloy cast irons is quite significant and usually
exceeds 3%.
There is common misconception that all cast irons are brittle. The ductility and fracture
toughness of modern ductile irons may be comparable with some steels.
Metallurgically speaking, upon the completion of solidification of cast irons, either graph­
ite particles of different morphologies (for a majority of commercial cast irons) or cementite
Fe3C (e.g., white cast iron) are formed. Graphite represents a crystalline form of carbon. It is a
more stable phase than cementite, making it more appropriate to use the Fe–C phase trans­
formation diagram when dealing with cast irons instead of the Fe–Fe3C diagram.
When hardening cast irons, it is not possible to obtain a 100% austenitic structure. However,
in contrast to high-carbon steels, instead of cementite Fe3C being present in austenite as
the second phase, graphite is present in most cast irons (an exception is white irons).
Graphite, which has relatively low density and hardness but high thermal conductivity
and good lubricity, makes a marked positive effect on several properties of cast irons.
Besides carbon, commercial cast irons consist of 0.6% to 4% Si (with 2% to 3.5% Si being
more typical), making these the two principal alloying elements. Silicon promotes a graph­
ite formtion. Therefore, because of the considerable amount of, it is more appropriate to
consider commercial cast irons not as binary alloys but at least ternary Fe–C–Si alloys.
There is a noticeable difference between the Fe–C and Fe–C–Si diagrams particularly with
respect to the maximum solubility of carbon in austenite, as well as the eutectic reaction.
Because of the presence of a considerable amount of silicon, carbon solubility in the aus­
tenite decreases, tending to dissociate iron carbides and potentially lowering the carbon
content of pearlite compared to plain carbon steels.
In order to provide particular properties for a certain type of cast iron, various alloying
elements including Mn, P, Cu, S, Ni, and others may be added [157,158,170,196,203–218].
For comparison purposes, Table 4.11 shows the typical chemical composition of the three
unalloyed cast irons: gray, malleable, and ductile irons [158]. As one can conclude, unlike
steels, different types of cast irons may have similar chemical composition but substan­
tially different morphology, microstructure, and properties. The specifics of cast iron
microstructure, the morphologies (shapes) of graphite, and the matrix structure distin­
guish different types of cast irons.
The fact that cast irons consist of a considerable amount of Si results in a shift of phase
transformation curves to higher temperatures compared to the ones shown in Figure 4.3
or in the Fe–C diagram. Thus, in contrast to the Fe–C diagram, the eutectic reactions on
the Fe–C–Si diagram occur at higher temperatures and over a range of temperatures that
increases with an increase of both the carbon and silicon content. Figure 4.41 illustrates
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4
200

TABLE 4.11
Typical Chemical Composition of Unalloyed Cast Irons
Chemical Composition (%)
Cast Iron Type TCa Mn Si Cr Ni Mo Cu P S Ce Mg
Gray 3.25–3.50 0.50–0.90 1.80–2.30 0.05–0.45 0.05–0.20 0.05–0.10 0.15–0.40 ≤0.12 ≤0.15 – –
Malleable 2.4–2.55 0.35–0.55 1.40–1.50 0.04–0.07 0.05–0.30 0.03–0.10 0.03–0.40 ≤0.03 0.05–0.07 – –
Ductile 3.60–3.80 0.15–1.00 1.80–2.80 0.03–0.07 0.05–0.20 0.01–0.10 0.15–1.00 ≤0.03 ≤0.002 0.0–0.20 0.03–0.06
Source: J.R. Davis (editor), Metals Handbook, ASM Int’l, Materials Park, OH, 1998.
a TC is a total carbon representing a sum of combined carbon (including carbon in solution and the free carbon [158]).
Handbook of Induction Heating
Heat Treatment by Induction 201

δ-Fe + γ-Fe + Liquid


Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

1500
δ-Fe
Liquid 2600°F
1400
Liquid + 2400°F
1300 Carbide
γ-Fe +
Liquid
1200 2200°F
Temperature, °C

1100 γ-Fe 2000°F

1000 cm
A
γ-Fe + Fe3C 1800°F

900
1600°F
δ-Fe + Fe
800
α-Fe + 1400°F
α-Fe A1 γ-Fe +
700 α-Fe + Fe3C Fe3C
1200°F
600
0 1 2 3 4 5
Carbon content, wt. %

FIGURE 4.41
Fe–C–Si equilibrium phase transformation diagram at 2% silicon. (From J.R. Davis (editor), Metals Handbook,
ASM Int’l, Materials Park, OH, 1998.)

this phenomenon [158]. All those features should be taken into consideration when choos­
ing the proper parameters for induction hardening.
As stated earlier, besides carbon and silicon, all commercial cast irons have, to a differ­
ent extent, a considerable amount of other purposely added alloying elements that deliver
specific properties. Also, cast irons consist of an insignificant amount of residual impuri­
ties occurring from the raw materials and specifics of the casting operation. Table 4.12 lists
some effects of the alloying elements in cast irons [218]. Table 4.13 shows the influence of Si
and three other elements commonly present in cast irons on the critical temperatures [218]
upon slow heating. A detailed description of the impacts of different alloying elements
on the heat treatment and performance characteristics of cast irons can be found in Refs.
[157,158,160,170,196,203–218].
The concept of CE discussed in Section 4.1.5.1 is also used when working with cast irons. CE
establishes the relationship between the effect of alloying elements and the amount of carbon
that would provide a similar heat treatment effect as an unalloyed cast iron. Several math­
ematical expressions have been developed over the years, allowing the calculation of CE for
cast irons [157,158,170,208–218]. This includes an expression recommended in Ref. [158]:

CE = % C + 0.3(% Si) + 0.33(% P) − 0.027(% Mn ) + 0.4(% S),



where % C is the value of the total carbon content (TC).
The eutectic composition of unalloyed cast iron is approximately 4.3% C. Therefore, if
a calculated CE is approximately 4.3% C, then the composition of that cast iron can be
roughly considered as eutectic. Cast irons with CE < 4.3% would be classified as hypoeu­
tectic. If CE > 4.3%, those cast irons would be considered as hypereutectic.
202 Handbook of Induction Heating

TABLE 4.12
Some Individual Effects of Selected Alloying Elements in Cast Irons
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

Element Effect during Solidification Effect during Eutectoid Reaction


Aluminum Strong graphitizer Promotes ferrite and graphite
formation
Antimony Little effect in amounts used Strong pearlite stabilizer
Bismuth Carbide promoter, but not carbide former Very mild pearlite stabilizer
Boron <0.15% Strong graphitizer Promotes graphite formation
>0.15% Carbide stabilizer Strong pearlite retainer
Chromium Strong carbide former. Forms very Strong pearlite former
stable complex carbides
Copper Mild graphitizer Promotes pearlite formation
Manganese Mild carbide former Strong pearlite former
Nickel Graphitizer Mild pearlite promoter
Silicon Strong graphitizer Promotes ferrite and graphite
formation
Tellurium Very strong carbide promoter, but not stabilizer Very mild pearlite stabilizer
Tin Little effect with amount normally used Strong pearlite retainer
Titanium (<0.25%) Graphitizer Promotes graphite formation
Vanadium Strong carbide former Strong pearlite former
Source: C. Walton, T. Opar, Iron Castings Handbook, Iron Castings Society, Inc., 1981.

TABLE 4.13
Influence of Silicon and Three Other Elements Commonly Presented in Cast Irons on the Critical
Temperature upon Slow Heating
Silicon Phosphorus Manganese Nickel
Range (%) 0.3–3.5 0–0.2 0–1.0 0–1.0
Upper temperature effect per % 37°C (67°F) 220°C (400°F) 37°C (67°F) 17°C (31°F)
Direction Increase Increase Decrease Decrease
Lower temperature effect per % 29°C (52°F) 220°C (400°F) 130°C (235°F) 24°C (43°F)
Direction Increase Increase Decrease Decrease
Source: C. Walton, T. Opar, Iron Castings Handbook, Iron Castings Society, Inc., 1981.

Many properties of cast irons can be related to CE. Cast irons with a low value of CE often
have better response to induction hardening. Table 4.14 shows the effect of a CE on the surface
hardness of induction-hardened gray irons [157,196]. A sizeable variation in the hardness may
be anticipated if there is an appreciable variation in the total carbon content and CE.
It is still important to remember that the value of CE is only a rough indicator of some
properties of cast irons, and its use is very limited when dealing with heat treatment of
cast irons. For example, CE does not provide any indication on an effect of the matrix
on the results of hardening. Certain iron castings might have practically identical chemi­
cal composition (which will be reflected by respected identical values of CE and TC), but
an attempt to harden those parts can vary from being “un-hardenable” to “hardenable”
depending upon its matrix (un-hardenable for a ferritic or predominately ferritic matrix
and hardenable for a fully pearlitic matrix).
The graphite appears in cast irons in different forms ranging from flakes and clumps
to spheroids. A description of these forms can be found in Refs. [157,196,218]. Gray, ductile
Heat Treatment by Induction 203

TABLE 4.14
Effect of Carbon Equivalent on Surface Hardness of Induction-Hardened Gray Irons
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

Composition (%)a Hardness HRC, Converted from Rockwell


Carbon
C Si Equivalentb As Read 30 N Microhardness
3.13 1.50 3.63 50 50 61
3.14 1.68 3.70 49 50 57
3.19 1.64 3.74 48 50 61
3.34 1.59 3.87 47 49 58
3.42 1.80 4.02 46 47 61
3.46 2.00 4.13 43 45 59
3.52 2.14 4.23 36 38 61
Source: Heat Treater’s Guide: Practices and Procedures of Irons and Steels, ASM Int’l, Materials Park, OH, 1999,
p. 822.
a Each iron also contained 0.5 to 0.9 Mn, 0.35 to 0.55 Ni, 0.08 to 0.15 Cr, and 0.15 to 0.30 Mo.

b Carbon equivalent has been calculated using the expression CE = % C + 1.3% Si.

(nodular), and, to a much lesser extent, the malleable and compacted graphite irons are
four groups of cast irons that more frequently undergo induction hardening.

4.1.11.1 Gray Cast Irons


Gray irons, being relatively inexpensive metallic materials with remarkable castability and
machinability, excellent wear resistance, and resistance to galling and seizure/spalling
(graphite flakes provide solid lubrication), are very attractive for a variety of applications.
The impressive fluidity of gray irons allows one to produce complex geometries including
a combination of thin and thick sections. The ability of gray irons to absorb energy caused
by vibration (damping capacity) is another attractive property that underlines the benefits
of gray iron castings in applications where the abovementioned properties are important.
It is quite easy to distinguish gray iron from ductile iron. Ductile cast irons consist of
graphite in shapes of spheroids or nodules. Gray iron consists of carbon as graphite par­
ticles that appear in 2-D metallographic examination in flake-like form (Figure 4.42).
Gray iron was named after the appearance of the fracture surface of its broken piece.
Graphite flakes are semi-interconnected, “acting” as stress risers; this is why most frac­
tures occur along these flakes, which appear to be gray in color.

0.02 mm

FIGURE 4.42
Gray iron consists of carbon as graphite particles that appear in 2-D metallographic examination in flake-like
form.
204 Handbook of Induction Heating

Tensile strength is typically used to distinguish different grades of gray irons, and its
code number relates to this property. In North America, the following standards are the
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

most frequently used to govern the classification of gray cast irons: ASTM A48/ASTM
A48M and SAE J431. For example, according to the ASTM A48 standard, gray iron grade
35 corresponds to minimum UTS = 35 ksi (kilopounds per square inch) or 240 MPa.
According to the SAE J431 classification, gray iron with the same properties is designated
as grade G11.
Similar to steels, there are several other national and international classifications of cast
irons, including The International Organization for Standardization (ISO 185:2005) that
also grades gray irons by minimum tensile strength in megapascals.
The properties of gray irons and their ability to be induction hardened greatly depend
on the type of the matrix structure (e.g., ferritic, ferritic–pearlitic, or pearlitic). The amount,
size, shape, and distribution of the graphite flakes also make an appreciable impact on
properties of iron castings.
Besides casting and solidification specifics, more intense cooling rates have a tendency
to form a fully pearlitic or predominately pearlitic matrix in gray irons. Slow cooling rates
promote preferential transformation of austenite into ferrite with or without some pearlite.
Cast irons with a ferritic matrix and predominantly ferritic matrix are commonly con­
sidered unsuitable for rapid induction hardening owing to the lack of ability to obtain the
typically needed hardness levels. Fully pearlitic or predominately pearlitic (e.g., a mixture
containing 90% pearlite and 10% ferrite) gray cast irons have better response to induction
hardening compared to a matrix with an increased amount of ferrite (e.g., having 80%
pearlite and 20% ferrite). Appropriate alloying elements may promote or suppress a pearl­
itic reaction.
The shape, size, and distribution of the graphite flakes also depend on the specifics of
the castings, cooling rate, and chemical composition. The length of the graphite flakes is
typically within the range of 0.06 to 1 mm.
ASTM specification (A247) establishes several types of graphite flakes designated by
capital letters “A” through “E” [219]. Table 4.15 consists of a short description of the differ­
ent types of graphite flakes. Fine graphite flakes that are uniformly distributed and ran­
domly oriented (type “A”) are the most preferable type of flakes for induction hardening
of gray irons.
Being stress-risers, graphite flakes act as crack initiation sites. In order to illustrate the
effect of those flakes on some of the properties of gray iron castings, one may imagine
that instead of being manufactured from solid metal, a workpiece was made from porous
material with a disrupted matrix consisting of a network of sharp-cornered internal micro­
cracks or notches. The larger the size of these microcracks and the greater the ratio of “the

TABLE 4.15
Types of Graphite Flakes in Gray Iron
Type “A” Type “B” Type “C” Type “D” Type “E”
Flakes are uniformly Flakes are Randomly Interdendritic Interdendritic
distributed and randomly oriented coarse segregation of segregation,
randomly oriented grouped into and moderate randomly oriented preferred
rosettes flakes (kish graphite flakes orientation
graphite)
Source: Metals Handbook, Vol.1: Properties and Selection: Irons, Steels, and High-Performance Alloys, ASM Int.,
Materials Park, OH, 1990; J.R. Davis (editor), Metals Handbook, ASM Int’l, Materials Park, OH, 1998;
ASTM A247-10 Standard Test Method for Evaluating the Microstructure of Graphite in Iron Castings.
Heat Treatment by Induction 205
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

FIGURE 4.43
Unitized machine for induction hardening of gray iron cylinder liners for commercial vehicle engines. It com­
bines two independently operated heat stations for hardening and tempering. Production rates as high as
50 liners per hour. (Courtesy of Inductoheat Inc., an Inductotherm Group company.)

volume of microcracks to the volume of the solid matrix,” the lower the tensile strength
and the poorer toughness will be.
This simplified analogy makes it obvious that the network of graphite flakes weakens
the matrix structure. The existence of these flakes with a combination of high carbon con­
tent makes the gray iron castings brittle and hard, with poor tensile strength and a limited
ability to elongate (typically, elongation <0.5%), failing in tension without any noticeable
plastic deformation. Poor ductility, toughness, and high notch sensitivity result in the ten­
dency of gray irons to fracture under bending or impact loading.
It has been reported in Ref. [158] that a combination of high carbon content, excessive
amount of silicon, high CE values, and slow cooling during manufacturing of iron castings
tends to produce larger graphite flakes. Preferentially oriented coarse flakes may bring certain
heterogeneity to the physical properties. For example, in the case of preferentially oriented
flakes, thermal conductivity along the planes of the graphite flakes is noticeably higher.
The achievable surface hardness after induction hardening of gray irons, as well as any
other cast irons, is typically reduced with an increasing amount of graphite.
Being brittle, gray irons may introduce certain challenges for induction heat treat prac­
titioners, because of the tendency toward cracking upon rapid heating as well as during
intense cooling. Preheating and the use of less severe quenching may be applied to reduce
thermal stresses and thermal shock. At the same time, there are cases when gray irons
have been successfully surface hardened using a short heat time (less than 3 s) and water
quenched.
As an example, Figure 4.43 shows a unitized machine for induction hardening of gray
iron (type A graphite) cylinder liners for commercial vehicle engines. It combines two
independently operated heat stations for hardening and tempering. High-speed, servo-
driven scanning assemblies and optimized process parameters allow very short heating
times and production rates as high as 50 liners per hour. The as-quenched hardness is
within the 47 to 49 HRC range. Nominal hardness case depth is 0.75 mm (0.03 in.). The
entire inner surface of the liner is hardened except for a 6-mm (1/4-in.) band at each end
with minimum distortion.

4.1.11.2 Ductile Cast Irons


As stated earlier, in contrast to gray irons, ductile irons have carbon particles in the form
of graphite nodules (Figure 4.44) that are isolated from each other. Formation of graphite
206 Handbook of Induction Heating
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

0.05 mm

FIGURE 4.44
Ductile irons have carbon particles in the form of graphite nodules (spheres) that are isolated from each other.

nodules is promoted by applying slower solidification rate and alloying (e.g., magnesium
and cerium alloying).
Graphite nodules serve as “crack-arresters,” providing ductile irons with important
advantages over other cast irons, including but not limited to ductility, relatively high
tensile and bending strength, moderate elongation, and much better toughness with com­
parable machinability. Figure 4.45 shows an induction surface-hardened crankshaft made
from ductile (nodular) cast iron, as well as the hardness pattern and microstructure of the
hardened case, transition zone, and “green” core.

“Green”
core

Transition
zone

Hardened
case

Surface

FIGURE 4.45
Induction surface-hardened crankshaft made from ductile (nodular) cast iron. Hardness pattern and micro­
structure of the hardened case, transition zone, and “green” core are shown. (Courtesy of Inductoheat Inc., an
Inductotherm Group company.)
Heat Treatment by Induction 207

Though there is an optimal combination of the size, number, and distribution of the
nodules for certain applications, usually graphite nodules of smaller diameters that are
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

uniformly dispersed within the matrix structure are desirable.


Ductile iron itself represents a group of materials offering versatile properties. The
group can be divided into five subgroups based on the structure of the matrix: ferritic,
pearlitic–ferritic, pearlitic, martensitic, and austempered ductile irons. Similar to gray
irons, the matrix structure of ductile irons is also controlled by the cooling intensity
during casting, as well as by additional alloying and heat treatment. A discussion of
the casting specifics and properties of different types of ductile irons can be found in
Refs. [157,158,170,196,203–218,294,295].
Induction hardening is usually applied to martensitic and pearlitic (or predominately
pearlitic) ductile irons and, to lesser extent, pearlitic–ferritic ductile irons having a con­
siderable amount of ferrite. Ductile irons consisting of a ferritic matrix structure are com­
monly considered nonhardenable by electromagnetic induction because of the inability to
obtain the hardness levels typically needed for most applications.
However, as always in life, there are some exceptions. It has been reported [292] that
upon rapid induction heating, short austenitization, and intense quenching, the fatigue
strength of ferritic ductile irons has been noticeably improved compared to an untreated
workpiece. It was suggested that several factors contributed to observed improvements.
One such factor is related to the formation of so-called ringed martensite formed around
the graphite nodules, thanks to the short-distance diffusion of the carbon from the graph­
ite nodules. The presence of ringed martensite leads to a localized hardness increase and is
associated with an increase of strength. Another possible factor may be associated with a
more favorable distribution and magnitude of compressive residual stresses.
As a result of the bending fatigue test of an untreated specimen, it was noted that the
crack initiation site was located around the graphite spheres and it propagated through the
ferritic matrix and graphite nodules [292]. In the case of the induction heat-treated speci­
men, the graphite area was somewhat smaller and the crack was initiated within the fer­
rite phase and propagated, avoiding graphite nodules surrounded by a halo of the ringed
martensite.
Nevertheless, induction hardening of ferritic ductile irons used in this case study is more
the exception than the rule, and for the great majority of induction hardening applications,
a ferritic or predominantly ferritic matrix of cast irons is highly undesirable because of the
inability to achieve the consistency and hardness levels typically needed in industry.
It should be noted at this point that it might be possible to change one form of ductile
iron to another by utilizing an appropriate ironmaking process. At the same time, once the
graphite has been formed, induction hardening cannot change its shape.
In North America, the gradation of ductile irons is commonly governed by several ASTM
standards. A classification A536-84(2014) is frequently used and it is based on the follow­
ing combined properties: “tensile strength–yield strength–elongation.” The strength units
are measured in kilopounds per square inch, and elongation is measured in percentage.
Besides the ASTM specification, the SAE J434 standard is also in use. In some countries,
ductile irons are graded based on the minimum tensile strength (in megapascals) and the
minimum elongation (in percentage); for example, SG800/2 corresponds to a strength of
800 MPa and 2% elongation.
Being inherently strong, ductile irons can handle much greater stresses during heat
treatment than gray irons. However, it is important to remember that although graphite
nodules of ductile irons serve as “crack-arresters,” their existence does not guarantee that
ductile iron castings will not crack during rapid heating or severe quenching. Caution and
208 Handbook of Induction Heating

common sense should be applied when choosing process parameters for surface harden­
ing of those materials, particularly if iron castings consist of high phosphorus content,
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

complex shapes, and geometrical discontinuities (e.g., sharp corners, holes, slots, etc.).
In the last several decades, a family of ductile cast irons was extended with a subgroup
of austempered ductile iron and bainitic ductile iron. Austempered ductile cast irons pro­
ducing bainitic matrix have gained popularity because of the attractive combination of
low cost and certain mechanical properties (including toughness and a high strength-to-
weight ratio). Its strength is almost doubled compared to ductile irons having a pearlitic
matrix structure while still exhibiting impressive toughness and elongation. Gear housing
is one of the common applications where such cast iron grades as ASTM A897 or DIN EN
1563 are used.
There are several subtleties during the induction surface hardening of these cast irons
where bainitic structures are needed instead of martensitic ones. One of such challenges
is associated with a necessity to make sure that quench intensity is appropriate and tightly
controlled. Cooling curves should “miss the nose” of the CCT diagram, “enter” the region
of the beginning of lower bainitic transformation, and “exit” that region in an area where
transformation of lower bainite has completed. It is critical that after entering the Bs region,
temperature would not excessively raise, thus not forming upper bainite, or is excessively
reduced, exiting a transformation region in the area where martensite will be formed. The
situation worsens because the bottom axis of CCT diagrams where the time is plotted has
a logarithmic scale. This means that tightly controlling the quenching intensity within
a relatively narrow temperature range would be required for a considerable amount of
time. Since most automotive components are normally of complex shapes with geometri­
cal irregularities and having regions with various thicknesses, shoulders, holes, and so on,
it might be challenging to apply spray quenching to hold a 3-D cooling intensity profile
within a narrow range for several dozens of seconds, making sure that certain critical areas
will not be cooled too low (forming martensite) and do not allow an accumulated internal
heat to “lift” cooling curves up, forming undesirable upper transformation products.

4.1.11.3 Specifics of Electromagnetic and Thermal Properties


As stated earlier, besides thermal properties, and unlike fuel-fired and infrared furnaces,
the performance of induction systems first and foremost is affected by the electromagnetic
properties of the heated materials, including electrical resistivity ρ and relative magnetic
permeability μr.
The chemical composition and volume fraction of graphite, its morphology, and the
matrix structure of cast irons affect not only mechanical properties but also electromag­
netic and thermal properties. For example, ductile irons with a ferritic matrix have a higher
thermal conductivity k compared to pearlitic or Q&T grades [217,218]. Gray irons with large
graphite flakes are known to have a higher k and a lower ρ (Figure 3.7b).
The Wiedemann–Franz law governs the relationship between the thermal conductivity
(k) and the electrical resistivity (ρ) for the majority of pure metals and metallic materials.
However, some alloys including cast irons are exceptions to this general rule, because in
addition to the chemical composition, the morphology of the graphite (e.g., nodules vs.
flakes) and the matrix constituents can alter that relationship.
The thermal conductivity of cast irons decreases with an increase in Mn and P con­
tent. It has been reported [208] that there is some impact on k if Cu content is less than
2%—it lowers thermal conductivity. Cu additions greater than 2% have no appreciable
effect on k. An increase in Si reduces k of most cast irons. For a given grade of cast iron,
Heat Treatment by Induction 209

k usually decreases with temperature. The specific heat of cast irons increases with
temperature.
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

Ref. [218] provides comprehensive data regarding the values of electrical resistivity of
different grades of cast irons and their magnetic properties and can be summarized using
the following selected points.

• Electrical resistivity ρ behaves in a complex manner, being a function of tempera­


ture, chemical composition, morphology of the graphite (e.g., nodules vs. flakes),
and microstructural specifics of matrix.
• ρ of a particular grade of cast iron gradually increases with the temperature rise.
• Cast irons with nodular or close to nodular morphology of graphite (e.g., ductile or
malleable cast irons) exhibit lower ρ compared to gray irons with graphite flakes.
• Courser graphite flakes of gray iron lead to higher ρ compared to finer flakes.
• Besides the CE and matrix constituents, the ρ of cast irons is noticeably affected
by alloying specifics. Figure 4.46a illustrates the effect of Si on ρ of pearlitic and
ferritic ductile cast irons at room temperature. Figure 4.46b shows the influence of
Al, Mn, and Ni on ρ of gray irons at room temperature [218,296].
• A pearlitic matrix results in increased ρ compared to cast irons with a ferritic
matrix. Pearlitic cast irons having fine spacing between the lamellae have greater
ρ compared to coarser pearlites. Ductile iron with a ferritic matrix has the lowest ρ.

The magnetic properties of cast irons are noticeably lower than those of plain carbon
steels and low-alloy and silicon-alloyed steels, exhibiting less residual magnetism, lower
magnetic saturation, and noticeably reduced values of relative magnetic permeability μr.
Cast irons with a ferritic matrix have higher magnetic properties compared to pearlitic cast
irons. It appears that ductile and malleable irons exhibit greater magnetic induction B and
higher μr compared to the respective magnetic properties of gray irons. More data regard­
ing the subtleties of the electromagnetic properties of cast irons including magnetization
curves and magnetic permeability curves are provided in Ref. [218].

0.8 1.4
Pearlitic matrix
Electrical resistivity, µΩ • m

Electrical resistivity, µΩ • m

2.8–3.6% ºC 1.3
0.7 Al
1.2
0.6 1.1

0.5 Ferritic matrix 1


2.9–4.1% ºC
0.9 Mn
0.4 Ni
0.8
0.3 0.7
1 1.5 2 2.5 3 3.5 4 0 1 2 3 4 5
Silicon content, % Alloy content, %
(a) (b)

FIGURE 4.46
(a) Effect of Si on ρ of pearlitic and ferritic ductile cast irons at room temperature. (Based on materials published
in www.ductile.org.) (b) The influence of aluminum, manganese, and nickel on the electrical resistivity of gray
cast irons at room temperatures. (From C. Walton, T. Opar, Iron Castings Handbook, Iron Castings Society, Inc.,
1981; I. Iitaka, K. Sekiguchi, Influence of added elements and condition of graphite upon electrical resistance of
cast iron, Reports of the Casting Research Laboratory, No. 3, Waseda University, Tokyo, Japan, 1952, pp. 23–25.)
210 Handbook of Induction Heating

4.1.11.4 Good Practices in Induction Hardening of Cast Irons and Closing Remarks


Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

a.
Ability to be hardened or obtain needed hardness levels. Among other factors, the ability
of cast irons to be induction hardened greatly depends on the amount of carbon
contained in the austenite, which, in turn, is a complex function of the matrix of
the cast iron before induction hardening, austenizing temperature, time being at
the austenite phase temperature, and chemical composition. In steels, the carbon
content is fixed by chemistry and, upon austenitization, cannot exceed this fixed
value (rare attempts to use a carbon-enriched environment are excluded from a
consideration). In contrast, in cast irons, there is a “reserve” of carbon in the pri­
mary (eutectic) graphite particles. The presence of those eutectic graphite particles
and the ability of carbon to diffuse into the matrix at temperatures of austenite
phase can potentially cause the process variability, because it may cause a local­
ized deviation/increase in an amount of carbon dissolved in the austenitic matrix
correspondingly affecting the achieved hardness levels upon quenching [316,327].
Higher temperatures of the austenite phase and longer times at those tempera­
tures are associated with a greater amount of carbon being dissolved within the
matrix. Since some of the graphite can go into solution in proximity to nodules or
flakes, it can locally increase the carbon level in the austenite, affecting martensitic
formation there and producing locally increased amount of RA. This tendency
is an important metallurgical factor representing one of the major differences
between hardening cast irons versus steels and potentially resulting in a variable
amount of carbon content of the matrix, Ms temperatures, and shift of CCT curves
and can also lead to some inconsistencies in the obtained hardness measurements
because of inadequate process control and monitoring. This is one of the reasons
why the requirements for process control and monitoring when hardening cast
irons are more stringent compared to hardening steels.
An attempt to compensate for the lack of carbon content in the matrix trying
to purposely diffuse a greater amount of carbon into austenite by applying exces­
sively high temperatures and dissolving the primary graphite particles cannot be
considered a good universal practice. This is so, because such practice is associated
with not only the necessity of reaching unduly high austenizing temperatures but
also the need to hold the cast irons at those temperatures for an extended period.
As a result, potential complications may occur. Some of those complications are
related to such undesirable metallurgical phenomena as severe grain coarsening,
presence of excessive amount of RA, incipient melting, irregular hardness pat­
terns, and delayed cracking.
While alloying elements can affect hardenability characteristics and RA as
they do in steels, an impact of alloying in cast irons is usually overwhelmed by
the effect of matrix carbon content [17,212,315–317]. Induction-hardened cast irons
usually have a well-defined case depth with a relatively short transition zone.
Thus, among other factors, the success in induction hardening of cast irons
and repeatability of obtained results is greatly affected by a potential variation
of matrix carbon content in terms of prior microstructure and hardened structure
requiring tighter control on the process recipe/protocol of induction hardening
and thermal history particularly throughout the austenitization stage.
b.
Friendly/unfriendly prior microstructures. Similar to surface hardening of steels, the
task of successful surface hardening of iron castings will be simplified by having
Heat Treatment by Induction 211

a friendly initial prior microstructure that is responsive to rapid induction heat­


ing. Taking into consideration that an austenite must contain sufficient amount of
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

carbon in order to form the required amount of martensite upon quenching, it is


preferable that cast irons have a sufficiently homogeneous Q&T or fully pearlitic
microstructure of matrix before induction hardening. Those prior microstructures
minimize needed distances for the carbon diffusion.
Ferritic cast irons or cast irons with a substantial amount of ferrite in the matrix
are not well suited for surface hardening by induction, because carbon has poor
solubility in ferrite. Therefore, the carbon could diffuse into the austenite from
eutectic graphite flakes, nodules, or initially existing regions of pearlite, bainite,
or Q&T martensite. The reaction of carbon diffusion from eutectic graphite into
the regions of a predominately ferritic matrix requires longer times. This discour­
ages one of the main advantages of induction, the short process time. Figure 4.47a
illustrates this phenomenon [204,216], showing the achievable hardness of ductile
iron with a ferritic matrix as a function of the holding time at 871°C (1600°F).
It should be mentioned at this point that a small amount of ferrite adjacent to
the graphite nodules (“bull’s eye”) usually does not affect the achievable hardness,
because the carbon might be able to quickly diffuse into those thin regions from
carbide nodules, enriching them with sufficient amount of carbon (assuming that
the temperature and time at the austenite phase are appropriate).
In some cases, regardless of the selected process recipe, a certain amount of fer­
rites might still exist within the as-quenched structure of the cast iron. These fer­
rites should not exceed a maximum amount of approximately 6% to 9%; otherwise,
a lower hardness reading and noticeable hardness scatter may occur.
Severe alloy segregation and a microstructure consisting of large graphite clus­
ters cannot be considered a friendly prior microstructure. Large graphite flakes or

600 70
Ductile iron
water quenched
500
Hardness, HRC

50
Hardness, HV

400
Four heats

300 C 3.52–3.65%
30 Si 2.22–2.35%
P 0.02–0.04%
200 Mn 0.22–0.4%
Ni 0.72–0.99%
Mg 0.045–0.065%
100 10
0 1 2 3 4 5 775 825 875 925 975
Time at temperature, min Austenitizing temperature before quenching, °C
(a) (b)

FIGURE 4.47
(a) Hardness response of ferritic ductile iron held at the austenizing temperature of 871°C (1600°F). (Based
on data from Surface Hardening of S.G. Iron, British S.G. Iron Producers’ Association Ltd., 1984; E. Rowardy
et al., Hardening characteristics of induction heated ductile iron, Trans. of the American Foundrymen’s Society,
61: 422–431, 1953.) (b) Influence of austenizing temperature on hardness of ductile iron. Specimens were
slow heated, held in air for 1 hour, and water quenched. (From ASM Handbook, Vol. 4, Heat Treating, ASM
International, Materials Park, OH, 1991.)
212 Handbook of Induction Heating

clusters having a preferred orientation of flakes being located within the hardness
depth serve as considerable stress risers (Figure 4.48). In this case, gray iron’s sen­
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

sitivity for crack development during rapid heating is increased and a redistribu­
tion of localized eddy current flow may occur (particularly when frequencies of
50 kHz and higher are used). The soft spots may occur within the as-quenched
microstructure in areas where large clusters are present.
It has been reported that the presence of eutectic carbides can promote crack­
ing in all types of cast irons suitable for surface hardening [207]. This phenom­
enon may lead to the appearance of a complex stress distribution during and after
quenching, because the carbides’ physical properties (i.e., thermal expansion, den­
sity, etc.) are quite different compared to similar properties of martensite.
Dendritic structures as well as undercooled graphite structures, which appear
as the result of rapid solidification and insufficient inoculation, should also be
avoided since, historically, they caused cracking and undesirable as-hardened
microstructures.
A friendly microstructure of the matrix of cast irons allows fast transforma­
tion at minimum required austenite phase temperatures, making it imperative
to minimize distortion compared to alternative structures and processes. For
example, a case study provided in Ref. [293] reports a reduction of distortion
when applying induction hardening compared to a similar quenching treatment
from a furnace. It has been recorded that after induction hardening, the maxi­
mum warpage in a 560-mm-long bar fabricated from gray iron was found to
be 0.03 mm compared with 0.17 to 0.25 mm for the same bars quenched from a
furnace.
Another case study provided in Ref. [299] reveals achieving an almost undetect­
able camshaft distortion of approximately 3–5 microns (based on 1.5- and 2.0-L
diesel or regular fuel engines) and, in many cases, an elimination of an entire
camshaft straightening operation after induction surface hardening. Such impres­
sive results have been obtained by applying Inductoheat’s patented nonrotational
technology (SHarP-C Technology) in combination with a friendly prior micro­
structure of ductile cast irons. Figure 4.49 shows a close-up of the unique design
of the SHarP-C inductor (a) and the as-hardened microstructure (b) that reveals a
two-phase structure (fine-grain martensite and nodular graphite) with an insig­
nificant amount of RA. SHarP-C Technology will be discussed in detail later in
this chapter.

FIGURE 4.48
Large graphite flakes or clusters having a preferred orientation of flakes being located near the surface serve as
stress raisers and should be considered to be undesirable microstructures.
Heat Treatment by Induction 213
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

(a) (b)

FIGURE 4.49
Close-up of Inductoheat’s patented SHarP-C inductor (a) and as-hardened microstructure (b) that reveals the
two-phase (fine-grain martensite and nodular graphite) structure with an insignificant amount of RA. 400×,
4% Nital etched. (Courtesy of Inductoheat Inc., an Inductotherm Group company.)

c.
Proper chemical composition. If, for some reason, cast iron does not respond to heat
treatment in an expected way, then one of the first steps in determining the root
cause for such unexpected behavior is to make sure that the cast iron has the
proper chemical composition and type of matrix as well as appropriate thermal
history (particularly throughout the austenitization phase) was used. In some
cases, what supposed to be the same type of cast iron purchased from two dif­
ferent suppliers may have appreciable variations of composition and properties.
Subtleties related to the necessity of having repeatable matrix carbon content and,
associated with it, the importance of having an accurate process recipe/protocol
were discussed above in (a).
Silicon should be closely controlled for a consistent and repeatable response
of cast irons to induction hardening because of its powerful impact on the phase
transformation, eutectic reaction, and solubility of carbon in austenite. Extreme
levels of Si are undesirable: a very low Si content promotes carbide formation,
while excessive Si levels promote more ferrite in the matrix [327].
Although carbon and silicon are two of the principal alloying elements in gray,
malleable, and ductile irons and have the most significant influence on the hard­
ened microstructure, the examination should not be limited to an evaluation of
the chemical composition of only these two elements. Commercial cast irons may
have a considerable amount of other alloying elements. As expected, inappro­
priate levels of those elements deviate results of hardening [17,212,315–317]. For
example, the phosphorus concentration should be below its prescribed maximum
level. It has been reported that an excessive amount of P (the normally specified
amount for ductile irons is ≤0.05%, whereas that for gray irons is ≤0.2%) can result
in increased brittleness of iron castings. Some of the side effects of having high P
levels include pitting and an increase in the probability of cracking, which wors­
ens with increasing hardening temperatures and phosphorus concentration.
An excessive amount of P is also associated with a greater risk of melting a phos­
phorus eutectic (melting point of approximately 930°C [327]). This could have an
undesirable impact on structure, particularly when deep case depths are obtained.
Therefore, elevated levels of phosphorus should be taken into consideration if prob­
lems arise. At the same time, it should not be assumed that elevated amounts of P
automatically causes cracking problem. On several occasions, investigations reveal
214 Handbook of Induction Heating

that cast irons containing an elevated amount of phosphorus (0.39%–0.57% P) have


been induction hardened successfully without cracking [327].
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

In addition, besides alloying elements, all commercial cast irons, to a different


extent, consist of traces of residual impurities resulting from raw materials and
the ironmaking practice. These residual impurities and alloying elements under
certain conditions may have mutual interactions markedly affecting the position­
ing of the critical temperatures, shifting the continuous cooling curves and affect­
ing the process of austenitization. As shown in Ref. [214], even seemingly small
amounts of such elements as bismuth, lead, titanium, tin, or nitrogen may have a
marked effect on the microstructure of the hardened component.
Special attention should be paid to elements, which promote graphitization. It
is also imperative to remember that some elements have a combined effect (i.e.,
carbon and silicon, sulfur and manganese, etc.); therefore, it is important to control
their combined effect. An increased Mn content exhibits a tendency to increase
the amount of RA, causing Mn segregation and reduced hardness readings [326].
On the other hand, it is necessary to have a sufficient amount of Mn to neutral­
ize S. Some investigators suggest following the Mn–S correlation for a minimum
amount of Mn in gray cast irons [327]:

Mn (%) = 1.7 × S (%) + 0.3%.

Close control must also be maintained over such properties as CE value (car­
bon equivalent) and TC value (total carbon). When dealing with ductile irons, it is
highly desirable to avoid a very high carbon content that increases the CE above
the eutectic [327].
The features of the casting process, including the rate of solidification, may also
cause some differences in respect to hardening.
Therefore, it is important that the heat treater has a clear understanding of not
only the microstructural specifics but also the chemical composition of the par­
ticular cast iron.
d.
Proper process parameters. The proper hardening parameters include but are not
limited to an appropriate temperature and time at the austenite phase, which,
besides other factors, are functions of the cast iron grade, matrix, heat intensity,
thermal history, and particularities of the IH mode (e.g., control mode: coil power
vs. coil current vs. coil voltage) and quenching specifics. The austenitizing tem­
perature should be sufficiently high (taking into consideration the nonequilibrium
nature of phase transformations during rapid heating) and time at the austenite
phase should be satisfactory long to make sure that all needed diffusion-driven
processes (required for the formation of adequately homogeneous austenite in
the matrix) will take place, and upon quenching, an as-quenched martensitic
structure will be free of the “ghost” pearlite as well as free ferrites. Insufficient
austenitization produces heterogeneous structures forming a mixture of martens­
itic and nonmartensitic products (e.g., consisting of islands of free ferrites within
the martensite). Low hardness levels, scattered hardness readings, inappropriate
mechanical properties, and a pronounced tendency toward crack development
characterize such structures.
Heat Treatment by Induction 215

Industry has accumulated several recommendations to estimate the required


austenitizing temperatures when induction hardening cast irons [29,158,205–209].
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

For a rough estimation of the minimum required austenitizing temperature of


unalloyed cast irons heated at moderate heat intensity, the following expressions
are often applied in industry:

Austenizing temperature, °C = 800 + 28(% Si) − 25(% Mn )

Austenizing temperature, °F = 1472 + 50(% Si) − 45(% Mn ).


Therefore, cast irons containing a lower silicon content can be austenitized at


lower temperatures. This is advantageous for a number of reasons, including a
reduction in the tendency for quench cracking attributed to thermal shock.
Besides the minimum austenitic temperature, there is a maximum recom­
mended temperature that should not be exceeded under any circumstances. The
temperature range of 860°C (1580°F) to 930°C (1706°F) is typical for induction sur­
face hardening of gray and ductile iron castings.
Overheating of cast irons above that maximum permissible temperature may
result in the appearance of an excessive amount of RA within the as-quenched
structure and even decarburization of the surface (if held for a sufficiently long
time), leading to hardness reduction and potentially forming an undesirable dis­
tribution of residual stresses. As an example, Figure 4.47b shows the influence of
austenitizing temperature on the hardness of water-quenched ductile iron [196].
Overheating is also accompanied by grain growth, resulting in coarse mar­
tensite and may even cause the formation of undesirable structures such as white
iron (in cases of severe overheating). As expected, high surface temperatures are
associated with large thermal gradients during the heating cycle and, especially,
during the quenching cycle, which may result in cracking.
Thus, an attempt should be made to select a process recipe that would allow
avoiding the formation of a mixture of martensite and undesirable transformation
products in the as-quenched microstructure. Figure 4.50 illustrates an example of
such undesirable mixed structures.

0.075 mm

FIGURE 4.50
An attempt should be made to avoid the formation of an undesirable mixture of martensite and upper transfor­
mation products in as-quenched microstructure.
216 Handbook of Induction Heating

As can be seen from the equilibrium phase transformation diagram, the cast
iron eutectic starts to melt at temperatures of approximately 150°C to 250°C lower
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

than the majority of carbon steels used in surface hardening. For example, in the
case of moderate heat intensity, the cast iron eutectic starts to melt at temperatures
of approximately 1130°C (2066°F). Therefore, if something goes wrong, there is a
distinct possibility for surface melting, incipient melting, or grain boundary liqua­
tion owing to the relatively low liquidus and solidus temperatures of cast irons.
The situation becomes more complicated, as certain alloying elements (such as
copper or tin, which are often added to some cast irons) have low melting tem­
peratures. Besides that, as a result of the nonequilibrium nature of rapid induc­
tion heating, all critical temperatures are shifted toward the higher temperatures.
Therefore, when hardening cast irons at sufficiently high heating rates and in order
to ensure the completion of the austenite transformation of the matrix, hardening
temperatures may approach dangerously high levels.
Unfortunately, in contrast to steels, there is very limited information in the lit­
erature with respect to the CHT diagrams suitable for heat treating of cast irons,
forcing induction heat treaters to rely solely on experiments and laboratory devel­
opments to determine the most appropriate temperatures for rapid induction
hardening of a particular cast iron grade.
All of these factors make it necessary to apply accurate process monitoring and
reliable control systems when hardening cast irons. The necessity of having pre­
cise monitoring becomes even more critical when surface hardening cast irons
with high silicon content and an excessive amount of low–melting point elements,
particularly when a scan hardening mode is used. Chapter 7 provides a discussion
on the monitoring principles and control systems.
When heating cast irons susceptible to cracking (e.g., gray irons with preferen­
tially oriented large graphite flakes), it is sometimes useful to preheat the castings
or to apply slow heating in particular during the initial stage (from room tem­
perature to approximately 550°C/1000°F). Since some cast irons and, in particular,
gray irons are by far the most susceptible to cracking upon quenching compared
to steels and ductile irons, preheating might be helpful, allowing the reduction
of thermal stresses and thermal shocks. Quenching oil at a temperature of 80°C
(176°F) to 100°C (212°F) and suitable aqueous polymer solutions may be used for
surface hardening of gray iron castings, allowing minimization of the probability
of cracking and excessive distortion.
In contrast to steels, “mass quenching” does not typically apply even in cases of
relatively small hardness case depths (e.g., less than 2 mm). On the other hand, short
quench delay (0.5 to 1.5 s) is applied quite regularly when hardening cast irons.
Since the Mf temperatures of cast irons are always below room temperature,
there will be a certain amount of RA formed in as-quenched structures (e.g.,
Figure 4.50 shows the presence of RA [light areas] in an as-hardened structure).
The amount depends on chemical composition, hardening specifics, and the pro­
cess recipe. An excessive amount of RA and its inappropriate control might nega­
tively affect the performance of the component. It has been demonstrated [297,298]
that by changing hardening temperatures, applying interrupted quenching,
developing heterogeneous austenitic structures, or having incomplete dissolution
of pearlite during austenitization, it is possible to vary the amount of RA in the
Heat Treatment by Induction 217

as-quenched structure from approximately 10% to more than 25%. Lower average
nodular graphite is associated with a tendency to lower average RA [297].
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

As expected, the greater amount of RA directly affects the as-quenched hard­


ness levels and the magnitude of residual compressive surface stresses. This could
alter critical mechanical properties (e.g., fatigue strength, toughness, and wear
resistance). Besides that, the load stresses that appear during the operation of the
component could transform the RA into untempered martensite, introducing brit­
tleness and potentially causing dimensional instability.
A research study of the complex behavior of the residual stresses as a func­
tion of the amount of RA was conducted in Ref. [297]. It was demonstrated that in
induction surface hardening a camshaft made of a pearlitic ductile cast iron, the
magnitude of the residual compressive stresses at the surface gradually increased
with the increase in RA. However, when RA exceeded 18%, the residual compres­
sive stresses started to decrease. Other studies reveal [298] that the decline in
residual stresses at the workpiece’s surface occurred earlier. Practice shows that
poorly controlled variation of RA is often responsible for the corresponding hard­
ness variations.
Different users of cast iron components develop internal corporate standards
to indicate the maximum levels of RA for a particular cast iron grade and applica­
tion specifics. For example, in some applications, having 8%–12% of RA might be
permissible.
As-quenched cast irons should be tempered as soon as possible to relieve exces­
sive residual stresses, form a tempered martensitic structure, improve toughness,
and avoid delayed cracking.
e.
Design Factors and Quality of Castings. Poor-quality castings and casting defects can
cause problems by themselves or can contribute to an abnormal combination of
factors that can cause problems. For example, the presence of such casting defects
as porosity, abnormal inclusions, sand and gas defects, blowholes, and dimen­
sional errors may cause a deviation in eddy current flow, local overheating, stress
concentration, cracking, hardness scatter data, burns, and so on.
When iron castings have a complex shape featuring a combination of “thick”
and “thin” areas, sharp corners, hollow areas, and other geometrical disconti­
nuities, the regions with substantially different masses have a tendency to heat
differently. This promotes the appearance of thermal gradients and measurable
transient stresses, which in turn might create favorable conditions for excessive
distortion and crack development. Good practices in designing hardening induc­
tors will be discussed later in this chapter.
The situation may deteriorate when transient stresses that occur during hardening
combine with an undesirable magnitude and a distribution of initial stresses retained
within the component after previous operations (i.e., casting, machining, honing, etc.).
In order to reduce the probability of cracking, stress relieving before induction hard­
ening can be very helpful. Since stress relieving is a diffusion-driven process, a certain
time–temperature correlation is needed. Stress relieving may relieve the majority of
stresses before hardening without a significant reduction of the strength of the cast
iron and its matrix modification. However, induction heating might not always be
the best choice for stress relieving of cast irons before hardening, particularly for
components with complex geometries. Instead, ovens are often more suitable for this
218 Handbook of Induction Heating

operation. One of the common practices for stress relieving of iron castings includes
heating the component and holding it at temperatures of approximately 550°C (1022°F)
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

for 1 or 2 h and then slowly cooling down to room temperature.


f.
Age-strengthening phenomenon. Some heat treaters with extensive experience in
induction hardening of gray cast irons might notice a phenomenon when seem­
ingly identical cast iron parts having the same geometry, grain size, and micro­
structure, produced at the same foundry (even from the same lot or batch) and
exhibiting practically identical composition and matrix structure, respond differ­
ently to induction surface hardening. Some of them may be processed very easily,
whereas others may have a substantial amount of cracks utilizing the identical
process recipe/protocol. One of the reasons for such a “strange” response of cast
irons might deal with the phenomenon called age strengthening. W. Nicola and
V. Richards [215] have conducted the first systematic experimental study of this
phenomenon.
It has been experimentally proven that aging at room temperature for approxi­
mately 60 days can strengthen gray cast irons up to 12%. Approximately 87% of
the gray cast irons evaluated have revealed the effect of “age strengthening.” The
Brinell hardness did not change with aging time, resulting in an increased ratio
of tensile strength to hardness. The age-strengthening phenomenon occurs in
both cupola- and induction furnace–melted cast irons. Interstitial (free) nitrogen
appears to be a controlling factor in determining whether aging will take place
or not [215]. Therefore, age-strengthened gray irons would more easily withstand
thermal gradients occurring during heating and quenching without cracking.
The age-strengthening phenomenon may explain the known practice in the
past when heat treaters would store some cast irons and, in particular, gray iron
parts for several months and then heat treat them. Therefore, it is important to
develop a process recipe and conduct a runoff using relatively new or fresh parts
and not the “aged” ones. This may help avoid unpleasant surprises in the future.

4.1.12 Specifics of Induction Hardening of Powder Metallurgy Materials


During the last decade, the use of powder metallurgy (P/M) materials in several indus­
tries has been expanded at an impressive pace; however, the biggest market for iron-based
P/M components remains to be the transportation industry, particularly the automotive
and agricultural sectors [300–306]. An ability to manufacture net-shape complex geometry
components offering competitive performance at an economical cost is an attractive fea­
ture of P/M parts (e.g., gears, timing sprockets, cams, splined hubs, shafts, shock absorbers,
etc.). Figure 4.51 shows a selection of P/M gears that regularly undergo induction surface
hardening [302].
Density and porosity in the sintered compact are major factors affecting strength and
hardenability of iron-based P/M materials; both properties have been noticeably improved
in the last decade, allowing closely approaching respected properties of fully dense
wrought steels. Carbon content is another major factor that affects achievable hardness
and strength.
Similar to steels, alloying helps enhance a desirable combination of strength, load-­bearing
capacity, ductility, and fracture resistance of P/M materials. Copper, nickel, molybdenum,
and, to a lesser extent, chromium, phosphorus, and silicon are the most commonly used
alloying elements in iron-based hardenable P/M materials.
Heat Treatment by Induction 219
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

FIGURE 4.51
Selection of P/M gears suitable for induction surface hardening. (From V. Rudnev, Intricacies of induction hard­
ening P/M parts, Professor Induction Series, Heat Treating Progress, ASM Int’l, Materials Park, OH, November/
December, 2003, pp. 23–24.)

Modern P/M materials are also noticeably more homogeneous with minimized chemi­
cal segregation. A detailed discussion on the effect of different alloying elements on the
heat treat properties, critical temperatures, and P/M hardenability can be found in Refs.
[29,196,300–306].
Still, strength, fatigue and wear resistance, and fracture toughness of P/M materials
remained somewhat lower compared to the respective properties of fully dense wrought
materials.
IHT of P/M materials has several subtleties compared to wrought steels. These features
primarily deal with the marked difference in physical properties noticeably affecting the
P/M’s response to induction heating and quenching. Sintering specifics, sintering tem­
perature, and secondary processing are other factors that affect the response of a P/M
material to induction hardening.
Challenges associated with differences in thermal properties (e.g., thermal conduction,
specific heat, etc.) of P/M materials are common for any type of heat treatment. However,
differences in electromagnetic properties are specific for processing those materials apply­
ing electromagnetic induction. Table 4.16 shows the effect of density reduction and poros­
ity increase on selected material properties and their influence on induction hardening.

TABLE 4.16
How Density Reduction (Porosity Increase) Affects Some P/M Part Properties and Induction
Hardening Parameters
Property Change Influence on Induction Process
Thermal conductivity Decrease Less soaking action from high-temperature to low-temperature
regions. Larger temperature gradients and thermal stresses
during heating. Slower cooling during quenching
Electrical resistivity Increase Larger current penetration depth
Magnetic permeability Decrease Larger penetration depth and lower coil electrical efficiency
Hardenability Decrease More severe quench is required to provide the same case depth
Structural homogeneity Worse Inconsistency of hardening; variations in surface hardness, case
depth, hardness scatter, and residual stress data. Increased
tendency for cracking during hardening
Source: V. Rudnev, Intricacies of induction hardening powder metallurgy parts, Professor Induction Series, Heat
Treating Progress, ASM Int’l, Materials Park, OH, November/December, 2003, pp. 23–24.
220 Handbook of Induction Heating

The relative magnetic permeability μr of P/M materials is significantly lower than the μr
of the corresponding steels, while their electrical resistivities ρ are greater to some extent.
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

Both factors lead to noticeably larger δ during the heating cycle, affecting not only heat
generation and temperature profiles but also the magnitude and distribution of transient
and residual stresses as well as coil electrical parameters (including coil impedance, load-
matching characteristics, etc.). These features make hardening recipes for P/M materials
noticeably different compared to hardening wrought equivalents.
When induction hardening P/M products, it is good practice to have a minimum den­
sity of at least 7.0 g/cm3 (0.25 lb/in.3). This will help obtain consistent heat treat results.
Sometimes, the density of a powder metal part is expressed as a ratio of its actual density
to the density of pure iron (7.87 g/cm3). When hardening surfaces that have cuts, shoulders,
teeth, holes, splines, slots, sharp corners, and other geometrical discontinuities and stress
risers, it is preferable to have a minimum density of 7.2 g/cm3 (0.26 lb/in.3).
One reason for low-density P/M parts to be prone to cracking is associated with a pene­
tration of the gases into the subsurface areas of the part through the interconnected pores.
Interconnected pores contribute to decreased part strength and rigidity compared with
wrought materials.
In addition, the poor thermal conductivity of porous P/M parts encourages the develop­
ment of localized hot spots and excessive thermal gradients and also requires the use of
quenchants with intensified cooling rates to obtain the required hardness and case depths,
because an increase in pore fraction and a reduction in density negatively affect the hard­
enability of P/M materials compared to their wrought equivalents.
Aqueous polymer solutions of various concentrations (with 2%–8% being the most typi­
cal) with some type of rust inhibitors and water (containing appropriate additives) are
often used in surface hardening P/M materials. Rust inhibitors help prevent internal cor­
rosion. Oil quenching is sometimes specified when hardening P/M parts, particularly for
those with stringent dimensional stability requirements and those having a pronounced
tendency to cracking. Concern about fires and environmental restrictions are obvious
drawbacks to using oils and oil-based quenchants. Note that quench oils may require
higher temperatures than polymer quenchants and water.
Highly porous P/M parts have a greater tendency to corrode and to take on an unsightly
appearance. The causes of both can be traced to residual water-based quenchant trapped
in subsurface pores [302].
It is quite common for P/M materials to absorb approximately 2% oil by weight. Therefore,
intensive ventilation must be incorporated into machine design when heat treating P/M
parts. It is also imperative to make sure that steps are taken to ensure that the reusable
quenchants remain sufficiently clean and have appropriate quenching characteristics
when running high production.
Obviously, density and porosity are not the only factors that affect the process of heat
treatment of P/M materials and probability of cracking. Other factors include the material
composition, homogeneity of the microstructure (the rate of material segregation), surface
conditions (including, surface roughness), and parameters of the heat-treating process, as
well as specifics of prior processing operations such as sintering essentials of the green
compact. In the case of sintering, factors include the process sequence, atmosphere used,
pressure, temperature, degree of sintering, and segregation. High-temperature sintering is
preferred because it improves microstructural homogeneity and ensures good diffusion.
However, decarburization of the surface before induction hardening should be avoided.
The alloying method used to produce the powder can also have a marked effect on heat
treat results. Among alloying techniques are admixing, diffusion alloying, prealloying,
Heat Treatment by Induction 221

hybrid alloying, and the MIM (metal injection molding) method. The technique used can
affect material segregation and chemical and microstructural heterogeneity, because of
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

different areas of the part undergoing abnormal phase transformations during cooling.
For example, large inclusions may form, serving as stress risers, increasing the potential
for cracking of the part and causing inconsistent hardness readings. Under certain condi­
tions, structural heterogeneity may even affect an eddy current flow.
Depending on the specific composition, some P/M parts may have a greater tendency
to crack. For example, special attention must be paid when developing induction harden­
ing recipes for P/M copper alloys. Though iron-based P/M materials with a high copper
content provide a good blend of mechanical properties at a competitive cost, they are often
more prone to cracking. Besides, there is always a legitimate concern in regard to incipi­
ent melting when using copper alloyed iron-based P/M utilizing very high hardening
temperatures. This makes it imperative to have sufficiently accurate process control and
monitoring.
For example, in order to reduce the crack sensitivity of a gear-like component with holes
made of DIN Sint D11 (ISO P2045), it was specified having gentle preheating in the fur­
nace (600°C for 40–45 min) and induction surface hardening using dunk quench in oil or
comparable high concentration polymer quenchant followed by tempering were specified.
Application called for the following composition of this iron-based P/M alloy: 0.6%–0.8% C
and 1.6%–2% Cu. Preheating and closely controlled process recipe are needed to avoid
crack developing during or after hardening. As-quenched hardness was within the 71–73
HRA range with noticeable amount of RA. After tempering, hardness was increased to
76–79 HRA accompanied by some reduction in RA (it was allowed to have 16%–18% of
RA within Q&T structure). Besides preheating and in order to avoid delayed brittleness,
it was also important to minimize the HAZ as well as peak temperature and to apply the
proprietary tempering recipe immediately after quenching.
Another case study discussed in Ref. [302] reveals that the microstructural heterogene­
ity in a copper alloyed P/M component may be so severe that, when examined under a
microscope, it almost looked like copper had been electroplated on the surface of the part.
This was the result of incomplete copper diffusion during alloying such heterogeneity can
easily result in the redirection of the localized eddy current flow if high frequencies are
applied.
Iron-based P/M components are typically more prone to inconsistencies to induction
hardening because of variations in prior microstructures, sintering conditions, composi­
tion deviations, and segregation caused by “lot-to-lot” processing differences (including
sintered density and composition) compared to wrought steels. Thus, greater inconsis­
tency compared to hardening wrought steels may be observed.
It should also be mentioned that although it is strongly recommended that P/M
induction­-hardened parts should have a density of not less than 7.0 g/cm3 (0.25 lb/in.3),
there is a number of successful applications where an O.D. hardening has been done on
P/M components with a density as low as 6.8 g/cm3.
In some cases, it might be considered beneficial for induction surface hardening if P/M
parts have a variable “surface-to-subsurface” or “O.D.-to-I.D.” density. For example, the
density at the surface to be hardened might be as high as 7.5 g/cm3 or even higher, and it
gradually decreases to a base density of 7.0–7.2 g/cm3 at the center of the part. This helps
maximize beneficial compressive residual stresses at the surface.
It is particularly critical to avoid having sharp corners/edges on P/M parts within regions
required to be induction hardened. Sufficient chamfering and radii should be applied (see
Section 4.3).
222 Handbook of Induction Heating

Shape/size distortion and warping attributed to heat treating are sometimes less for
P/M components than for their wrought steel counterparts [302]. The degree of shape
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

deformation strongly depends on part/coil geometry and hardness pattern and is usually
determined experimentally.
When determining process parameters for induction hardening of P/M parts, energies and
frequencies higher than those used for wrought equivalents are often needed. Closer process
control is also required. Preheating or pulse heating might be beneficial for obtaining the
required hardness patterns and avoiding crack initiation when hardening complex geometries.
The P/M industry continues to improve its technology. P/M parts sometimes were
tagged “low strength.” The low strength and high porosity of P/M parts held back the
widespread adoption of induction hardening. Not anymore. Improvements in P/M parts
manufacturing and a greater awareness of induction hardening specifics of P/M parts
have emerged in recent years. A number of different tools (including numerical computer
modeling) are now available to develop intelligent process recipes that will ensure success
in induction hardening of P/M components.

4.2 Induction Hardening: Subtleties of Machine


Design and Process Recipe Selection
Hardening of steels, cast irons, and P/M materials represents the most popular applica­
tion of induction heat treatment and has already been briefly discussed in Section 2.2.1.
Figure 4.52 illustrates a small array of a vast variety of geometries of induction surface-
hardened components and hardness patterns [401]. A typical induction hardening proce­
dure involves heating the entire component, or the region that needs to be hardened, to the
austenitizing temperature, holding it (if required) for a period long enough for completion
of the formation of austenite, and then rapidly cooling it below the temperature where
martensitic transformation begins (Ms temperature).
The holding stage (also referred to as the dwell or soak stage) usually takes place after
the completion of the main heating cycle and before quenching, representing a short time
delay. Coil power is switched off, or dramatically reduced, during this stage, helping to
improve radial temperature distribution and reducing the thermal shock during the initial
stage of quenching. The dwell/soak stage is typically applied in through hardening or in
surface (case) hardening when considerable hardness depth is required or when heating
alloys that exhibit high brittleness, low toughness, or low ductility (e.g., hardening gray
irons, some P/M materials, or high-carbon steels).
The three most common forms of induction hardening are surface hardening, through
hardening, and selective hardening. Depending on the application specifics, selective
hardening is sometimes considered to be a part of surface or through hardening.
Surface (case)-hardened components, having a hard outer shell, are strong and typically
have compressive residual stresses at the surface (see Figure 4.53). Among other factors,
these stresses are important for improving the fatigue properties of the workpiece, allow­
ing the delay of crack initiation and resisting the propagation of microcracks.
Depending on the heat treat specifications, required production rate, and component
geometry, induction hardening equipment can be designed as a relatively simple appa­
ratus with manual operation or can involve sophisticated machinery being completely
automated and can even apply elements of artificial intelligence.
Heat Treatment by Induction 223
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

FIGURE 4.52
Small array of a vast variety of geometries of induction surface-hardened components and hardness patterns.
(From V. Rudnev, Advanced Technologies in Induction Hardening, Invited Lecture for Caterpillar Inc., Peoria, IL,
14 January, 2015.)

Tensile Surface hardening


stresses
Through hardening
Stresses

Compressive Surface Core


stresses

Radius/thickness of workpiece

FIGURE 4.53
In contrast to majority of through-hardened parts, surface-hardened components typically have compressive
residual stresses at the surface.
224 Handbook of Induction Heating

As stated in Chapter 2, electromagnetic induction makes it possible to generate highly


controllable heat intensity quickly at specific areas of the part with high precision and
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

repeatability, ensuring the needed quality and reproducibility. In contrast to thermochem­


ical hardening systems, induction machines are often more energy efficient and envi­
ronmentally friendlier than other heat treat methods. In addition, induction equipment
requires practically no start-up and shutdown time; it uses minimum floor space and typi­
cally produces less distortion in the workpiece.
Some size/shape distortion always occurs whenever a metallic material is heated or
cooled. There are several reasons for this:

1. During the heating cycle, internal stresses formed during previous operations
(also referred to as initial stresses) are relieved to a different degree (depending on
the achieved temperature and process time), causing some change in the geom­
etry of the workpiece.
2. Thermal expansion (during heating) or contraction (during cooling) also makes
some contribution to the distortion, particularly when dealing with complex
geometries. The presence of appreciable thermal gradients during the heating and
quenching stages may also affect its magnitude.
3. Phase transformations associated with the crystal structure changes and volu­
metric variations also inevitably contributed to the shape/size distortion of heat-
treated components.

Thus, distortion is always present in hardening, but an attempt should be made to mini­
mize it. Distortion repeatability is another critical factor, since if it is repeatable, it could
possibly be compensated for during component fabrication.
Besides the metallographic features that have been discussed in Section 4.1, there are
several other factors that affect the distortion of the hardened component, including geom­
etry, hardness profile, amount of heat generation, specifics of handling workpieces during
heat treating, and some others.
The amount of heat generated within a component is one of the critical factors that have
a pronounced effect on distortion. This is particularly so when heat treating irregular-­
shaped components with a lack of symmetry such as camshafts, crankshafts, gears,
workpieces with cross-holes, keyways, shoulders, and other geometrical irregularities. In
general, the greater the amount of heated metal, the greater the expansion/contraction will
be, which worsens the shape distortion.
Usually, the distortion of the component is minimized when induction surface harden­
ing is used because heating occurs quickly with high precision and a minimum amount
of material is being heated to elevated temperatures. The presence of a colder core acts as
a shape stabilizer.
Induction surface hardening offers high-dimensional stability and repeatability, par­
ticularly on symmetrically shaped parts such as bar shafts, axle shafts, rods, pins, and the
like. Expected distortions upon hardening certain components and ways to control it will
be discussed later in the corresponding sections of this chapter.
Some of the important features of induction surface hardening are the ability to gener­
ate heat in well-defined areas and minimize the heat times that usually range from 0.5 to
20 s depending on the application (with 1.5–6 s being more typical). Heat time for hard­
ening depends on several factors, including prior microstructure, component geometry,
required hardness pattern/depth, and the available power supply.
Heat Treatment by Induction 225

Induction equipment may be easily incorporated into cellular workflow, which is quite com­
mon in most manufacturing facilities. Many IH systems are capable of producing hundreds of
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

parts per hour, and in some cases, the production rate may even exceed a thousand parts per
hour (e.g., selective hardening of fasteners). The use of robots, gentries, and conveyors can com­
pletely automate the induction heating operation. Induction equipment should follow standard
machine tool design practices in that it should be robust and repeatable for continuous use.
It has been said, and too often quoted, that the only certainties in life are death and taxes.
With these two certainties, the induction heating professionals could add another one, like
being confused by what frequency best suits a particular induction heating or heat-treating
application. Subtleties of the frequency selection induction heat treatment are discussed in
several subsequent sections of this chapter.

4.2.1 Hardness Case Depth Definitions


The first step in designing an induction surface hardening machine is to specify the required
surface hardness and hardness pattern, including an effective case depth. In some cases,
additionally, a total case and core hardness are also included in process specification. A
necessity of having a particular hardness pattern is associated with the specifics of a com­
ponent’s working condition, assuring appropriate performance characteristics and needed
mechanical properties including strength in tension, torsion, bending, fatigue, and so on.
The ability to obtain the required hardness pattern depends on the following factors:
the temperature distribution, the workpiece geometry, the microstructure of the “green”
part, its chemical composition, quenching conditions, grain size, and the hardenability of
the steel [21]. The temperature distribution is controlled by the selection of the frequency,
power density, workpiece/coil geometries, and the time of heating.
Case depth (hardness depth) can be defined in several ways: (1) specifying the thickness
of the surface layer (on outside or inside surfaces) where a certain hardness level should
be achieved, (2) requiring a particular amount of martensite needed to be formed (e.g.,
the microstructure having at least 50% martensite can be specified), or (3) specifying the
region where certain metallurgical changes can be observed. Figure 4.54 illustrates a tem­
perature profile and typical hardness distributions in surface hardening. As can be seen,

Surface Core

Transition
Temperature

zone
Ac3
Ac1

Hypoeutectoid steel
Hardness

Eutectoid steel

Radius/thickness

FIGURE 4.54
Temperature profile and typical hardness distributions in surface hardening.
226 Handbook of Induction Heating

below a certain depth, the hardness begins to decrease drastically. This occurs within the
area that is commonly referred to as the transition zone.
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

The term effective case depth is commonly used to specify the required hardness depth,
and it is defined as the distance from the surface (outside or inside surfaces) where the
specified hardness level (e.g., 50 HRC or 40 HRC or equivalent values using other hard­
ness scales) is achieved. This depth is typically measured in a cut section of the hardened
workpiece using an appropriate hardness tester and procedure to determine the hardness
at various distances from the surface.
Taking into consideration that induction hardening does not change the chemical com­
position of the steel, the hardness levels specified for determining the effective case depth
sometime is correlated with the carbon content or the CE. For example, the SAE standard
J423 suggests defining the hardness levels for an effective case depth with the steel’s nomi­
nal carbon levels [307]. Table 4.17 shows tabulated values of effective case depth hardness
versus the steel’s nominal carbon content specified in SAE J423.
Corporate standards may define different values of effective case depth hardness com­
pared to those shown in Table 4.17. In some cases, the distance below the part surface
where the hardness drops 10 HRC lower than the surface hardness might be used as a
measure of the effective case depth.
Total case depth and core hardness might be specified in addition to the effective case
depth values. The term total case depth indicates the distance from the surface where a
change in microstructural properties after surface hardening can be observed (meaning
the surface layer where the microstructural difference between the subsurface region and
the core is not visible). Thus, in contrast to the effective case depth, the extent of the transi­
tion zone is included into the total case depth.
Specification of the surface hardness, effective case depth, and total case depth can be
important for the performance of a component, since it affects not only strength but also
the magnitude and distribution of residual stresses and the distortion characteristics.
Usually, a shorter transition zone and higher hardness levels are associated with greater
surface compressive residual stresses.
In addition, as discussed in Sections 4.1.4 and 4.1.5, it is not realistic to expect that steels
of the same grade would always have an identical chemical composition. Therefore, the
steel’s carbon content and hardenability always deviate to some degree, correspondingly
affecting the achieved surface hardness and the hardness pattern. Specification of the
surface hardness, effective case depth, and total case and core hardness provides addi­
tional comfort factors that ensure the consistency of the mechanical properties of the as-­
hardened component.
Figure 4.54 illustrates a comparison in transition zone when hardening hypoeutectoid
carbon steel versus eutectoid steel.
Several opinions exist as to whether the transition zone should be short or long. In real­
ity, both types of transition zones are successfully used in industry and how large it should
be depends on the specifics of the application.

TABLE 4.17
Effective Case Depth Hardness versus Steel’s Carbon Content According to SAE Standard: SAE J423
Carbon content (%) 0.28–0.32 C 0.33–0.42 C 0.43–0.52 C ≥0.53 C
Effective case depth 35 HRC 40 HRC 45 HRC 50 HRC
hardness
Source: SAE Standard J423: Methods of Measuring Case Depth, ANSI, Feb. 1998.
Heat Treatment by Induction 227

For example, induction hardening of steels with high and low hardenability to the same
effective case depth will not be expected to provide the same strength and fatigue life.
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

In order to match similar performance characteristics, it is often necessary to provide a


greater (to some degree) effective case on the higher-hardenability steels [21,308]. In hard­
ening high-speed transmission gears, it is often advantageous to minimize the transition
zone, yet have sufficient case depth and core hardness.
A common method for checking the hardness pattern is destructive testing. Induction-
hardened components are sectioned longitudinally and transversely. Sectioning can be
carried out to an entire part or to some critical areas. SEM analysis, metallographical exami­
nation, and optical evaluation of the obtained microstructure are performed to ensure
that the desired hardening results are obtained.
The various destructive methods for measuring case depth are described in the SAE
standard J423. Nondestructive methods are also applied in industry. Issues, complications,
and good practices related to the different destructive and nondestructive techniques for
measuring hardness case depths as well as sample preparation techniques are discussed
in Refs. [21,159,307,309].

4.2.2 Induction Hardening Methods


As has been briefly reviewed in Section 2.1.1, there are four primary methods for induction
hardening:

• Scan hardening
• Continuous or progressive hardening
• Static hardening
• Single-shot hardening

These methods are related to the heating mode, inductor design, part geometry, and
processing.

4.2.2.1 Scan Hardening
In scan hardening, the inductor or workpiece or both may move linearly relative to each
other during the hardening cycle. Depending on the workflow of parts, the system can be
built as vertical, horizontal, or even at an angle, though vertical scan hardening is by far
the most popular design concept for a number of reasons, including a smaller footprint
and natural quench flow to outline a couple.
This scanning method can be applied to harden flat surfaces or irregular shapes.
However, it is most frequently used for hardening outside or inside surfaces of compo­
nents of generally cylindrical shape, such as bar shafts, axle shafts, bed-ways, bumpers,
race tracks, and so on.
As an example, Figure 4.55 shows two vertical scan hardening systems: (a) for heat treat­
ing automotive shafts, and (b) for hardening hollow bumpers. Bumpers are not rotated but
are processed in the axial direction through an inductor.
In the majority of applications, the cylinder-shape workpiece (e.g., shaft) is held between
centers or some other locating/holding tooling. The workpiece may rotate inside the induc­
tor to even out the induction hardening pattern around the circumference (Figure 4.55a)
228 Handbook of Induction Heating
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

(a) (b)

FIGURE 4.55
Vertical dual-spindle system for scan hardening shafts (a) utilizing two single-turn inductors and an induc­
tion system (b) for scan hardening of complex geometry hollow bumpers using a four-turn solenoid inductor.
(Courtesy of Inductoheat Inc., an Inductotherm Group company.)

or it may be positioned preferentially with respect to the inductor and processed without
rotation (Figure 4.55b).
Vertical systems can be set up to scan as many as four shafts at a time depending on the
size of the shafts being processed and the available power source. Loading may be manual
or automatic. Parts are loaded onto a lower center. A loading assist “Vee” block or nest may
be used to steady the part as it is being loaded. For larger parts, pneumatic cylinders may
lift the upper centers to facilitate loading.
From the load position, the part is raised up until the lowest area of the part to be heated
is in position in the heating coil. The shaft begins to rotate, the power is turned on, and the
coil may remain stationary for a period of time to provide sufficient preheating of certain
regions of the workpiece, for example, fillets. The shaft then begins to move relative to the
inductor, which is where the term scanning comes from.
In some machines, the coil remains stationary and the part moves, while in other
machines, the opposite is true [310]. To keep the high-frequency electrical leads short, it is
usually preferred to move the part through the coil.
The quench sprays down at an angle required to provide appropriate cooling without
allowing the quenchant to bubble back up into the heating area of the coil.
The quench is turned on after a short delay to rapidly cool the area austenitized dur­
ing the dwell stage. The quench delay may last a fraction of a second or a few seconds
depending on the geometry of the component, material hardenability, and hardness pat­
tern specification.
Scan hardening systems provide noticeable process flexibility with respect to the work­
piece length and, to some extent, variations in the part diameter. In addition, this scanning
mode provides the ability to vary the speed and power during the process, which controls
the amount of heat applied to different areas of the part.
The ability to detect the end of the heat-treated part allows processing a wide variety
of parts using the scanning mode without a tooling changeover. The machine chooses an
appropriate process recipe for whichever part is loaded. The use of servomotors is often
required for indexing IN and OUT of the coil at high speed to satisfy the required cycle time.
As with all induction applications, the appropriate frequency and power levels are criti­
cal for obtaining the required hardness pattern. Assuming that the power and frequency
have been chosen properly, the controlling factor for case depth in scan hardening is the
rate of scanning, which translates into the time at heat.
Heat Treatment by Induction 229

It is important to remember that quenching must follow the heating. If the scan speed
is too slow, the part may cool below the critical temperature before it enters the quench,
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

resulting in the formation of upper transformation products and low hardness.


When scanning symmetrical parts such as pins or simple shafts, often a single speed
and power level may be used. The geometry of the part being heated has a tremendous
influence on the resultant case depth (Figure 4.52). Sometimes, a slight temperature sur­
plus may not be objectionable as long as it does not result in excessive grain growth or an
unacceptable degree of grain boundary liquation (Figures 4.36, 4.37, and 4.39). The use of
preheating of fillets and changes in scan speed are commonly-used means of controlling
the pattern during scanning of irregularly shaped parts.
The more complex the workpiece is, the more complex the machine control scanning
protocol/recipe will be in order to address the complexity of the workpiece.
Scan inductors may be of different types, providing a repeatable, easily automated pro­
cess that can quickly adapt to new heat treatment tasks and be easily integrated into a
work cell.

4.2.2.1.1 Inductor Designs for Scan Hardening


Single-turn or multiturn inductors may be used in scan hardening (Figure 4.55). The
required number of turns is determined by the workpiece geometry and the ability to
properly load match (load tune) the coils to the power supply or by other specific pro­
cess requirements, such as the production rate or the hardness pattern runout, for exam­
ple. This impedance matching process is particularly important if maximum power is
required from the power supply. There should be a balance in voltage, frequency, and
current to achieve the desired power level without reaching the operational limits. The
term load matching or load tuning is used to describe this process and discussed in detail in
Section 7.4.
The longer (horizontal arrangement) or the higher (vertical arrangement) the scan coil is,
the faster the scan rate can be. This is due to the simple fact that the longer inductor leads to
a longer period when the part will be inside the coil; therefore, the scan rate can be greater.
However, limitations on the maximum length of the inductor may be associated with the
maximum permissible runout (also referred to as pattern cutoff or axial transition zone).
Single-turn inductors with narrow heating faces are used where a sharp pattern runout
is needed. An example of this would be the case where a pattern must end near a snap
ring groove. A wider heating face or more coil turns can be used when a faster scan rate
is desired and an extended longitudinal transition is permitted. The main disadvantage
to the wide heating face is that it will produce an extended/gradual pattern runout and
may not meet more stringent pattern specifications/consistency when hardening complex
geometry parts because of the undesirable appearance of electromagnetic proximity effect
and the unspecified shift of coil current density within the wide heating face.
Single-turn inductors are typically computer numerical control (CNC) machined from a
solid copper bar/block, thus making them rigid, durable, and repeatable. In other cases, a
copper tube may be used in the coil fabrication.
The quench barrel and its quench holes can be integrated into a single-turn scan induc­
tor. This type of inductor is sometimes called an MIQ (machined integral quench; see
Figure 4.56a) inductor.
Figure 4.57 shows several images of MIQ inductors. Though the great majority of scan
hardening inductors are of general cylinder shape, some scan inductors may have irregu­
lar geometry (Figure 4.57b), conforming to the shape of the workpiece.
230 Handbook of Induction Heating

Single-turn coil Shaft Single-turn coil Single-turn coil


Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

Quench
ring

Quenching

Case depth
(a) (b) (c)

FIGURE 4.56
Combined and separate quench designs. (a) MIQ-inductor machined integral quench, (b) separate inductor/
quench design, and (c) MIQ-inductor with quench ring (quench follower).

(a) (b)

FIGURE 4.57
Several examples of MIQ inductors. (a) Classical MIQ inductor of general cylinder shape; (b) cross sections
revealing water-cooling pockets and quench channels (left two images) and irregular geometry inductor
conforming the shape of the workpiece (right image). (Courtesy of Inductoheat Inc., an Inductotherm Group
company.)

The quench spray typically “impinges” the part approximately 12 mm (½ in.) to 40 mm


(1.5 in.) from the coil heating face and is angled away to prevent the quench from splash­
ing back into the inductor. This dimension can vary with different types of steel, the scan
rates, and the design specifics.
An additional quench barrel may be added to inductors designed to run parts with vary­
ing diameters. A quench barrel that follows the heating inductor (Figure 4.56b) is also used
in cases of moderate case depths and production rates, as well as for through-hardened
parts. In this case, the barrel or quench block is separated from the induction coil.
A combined design, which comprises an MIQ inductor and a quench follower, is also
applied (Figure 4.56c). This design is usually used in through hardening applications as
well as in cases of deep surface hardening or high production rates (high process speed)
when space is available for its positioning. A quench follower may be beneficial in running
parts with varying diameters, ensuring sufficient quenching of parts with geometrical
Heat Treatment by Induction 231

irregularities. This approach prevents problems with parts that are too hot to handle and
also provides sufficient quench flow to avoid the hardness reduction owing to an undesir­
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

able tempering back, which occurs if the part was not properly quenched and the residual
heat is allowed to affect the hardened area because of insufficient quench flow or time.
For scan inductors that are intended to heat fillets, an appropriate copper heating face
region must be focused into the fillet area. Coil copper profiling and the use of flux inten­
sifiers (flux concentrators) are beneficial to focus the magnetic field into the fillet. These
are critical applications that require careful design because the current will have a ten­
dency to take the shortest path and stay in the shaft area rather than flowing into the fillet.
Therefore, all efforts must be made to focus the heat generation into the fillet. Typically,
higher frequencies work better for this purpose.
Scan inductors have a better chance of achieving the desired metallurgical results, with­
out the necessity of modifying the coil in order to achieve the desired hardness pattern
thanks to the flexibility that results from being able to vary the power, scan rate, and
quench delay. Scanning is particularly advantageous for shafts where the available power
is limited.
Rotation of the part is particularly beneficial when a conventional single-turn inductor
is used without special copper profiling. In cases like this, a heat deficit may be present
in the workpiece’s region located near the area of the polarized power leads. If the part
is not rotated, there will be a tendency to develop a colder spot in the area of this split in
the inductor. Induction practitioners sometimes refer to this phenomenon as the “field-
fringing” or “fish-tail” effect, which will be discussed later in Section 4.2.8.3.
In cases where rotation of the component is not practical, alternative nonrotational solu­
tions can be used. This includes nonconcentric workpiece positioning with respect to
the inductor or special coil copper profiling; this has been successfully used in patented
SHarP-C inductors. Note that nonconcentric part positioning necessitates that the work­
piece be particularly oriented in respect to the single-turn inductor split region.
Induction scanning can also be effectively applied for hardening noncylinder geometries
(e.g., flat surfaces, raceways, teeth of large gears, etc.). Figure 4.58 shows three examples of
scan hardening non–shaft-like components.
Figure 4.58a shows surface hardening of track shoes for earth-moving machines (the
required case depths is within the 18- to 24-mm range) using a hairpin type inductor and
E-shaped laminations. “Hairpin style” (inductors) was derived from the resemblance of

(a) (b) (c)

FIGURE 4.58
Three examples of scan hardening non–shaft-like components.
232 Handbook of Induction Heating
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

FIGURE 4.59
Examples of keyway inductors.

the inductor’s loop shape to a lady’s hairpin. Such inductors are usually formed of bent or
brazed copper tubing or CNC machined. Hairpin inductors are also used for scan harden­
ing of a large bearing ring (Figure 4.58b).
Scan hardening utilizing the tooth-by-tooth concept (Figure 4.58c) is a popular technol­
ogy for heat treating large gears with wide faces used in wind turbines. Two scanning
techniques can be applied: one where the inductor is stationary and the gear is moveable,
and the other where the gear is stationary and the inductor is moveable. This technology
is discussed in Section 4.9.1.
Split-return, serpentine, clamshell, keyway, and channel-type inductors are among other
inductor styles that are occasionally used in scan hardening applications. As an example,
Figure 4.59 shows two keyway inductors that might be beneficial for scan hardening of a
localized area of cylinder components having diameter changes.

4.2.2.1.2 Case Studies of FEA Computer Modeling of Scan Hardening Applications


The subtleties of scan hardening can be better understood by analyzing the results of FEA
computer modeling, since it allows the unveiling features of a process that, in many cases,
may be difficult or even impossible to determine experimentally.
To illustrate, Figure 4.60 shows the results of FEA computer modeling of the sequential
dynamics of vertical scan hardening of a hollow shaft that comprises sharp fillet, diameter
changes, and shoulders using a two-turn MIQ inductor with an “L”-shaped flux concen­
trator ring (frequency = 9 kHz) [310,314]. Both turns are connected electrically in series,
carrying the current of the same magnitude.
To properly harden the shaft’s fillet area (in particular fillets without any radius or with
small radius), a short dwell is often incorporated at the beginning of the scan hardening
cycle where the inductor is energized but does not move. Quenching is not applied during
the dwell stage. In this case study, a 2.6-s power dwell is applied at the beginning of the
process cycle (Figure 4.60a and b).
Achieving a case depth assuring the minimum required hardness depth in the shaft
fillet area without exceeding the maximum case depth in regions next to the fillet
might present certain challenges. Those challenges can be divided into three groups
[143].
From an electromagnetic perspective, it is challenging to generate a sufficient heat source
(power density) directly into the fillet without creating excessive temperatures in neigh­
boring regions. Smaller fillets with smaller radii are even more challenging since electro­
magnetic proximity and ring effects become more powerful, unfavorably distributing the
heat sources.
Heat Treatment by Induction 233

Axis of
symmetry
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

Hollow shaft
Two-turn
MIQ-type
inductor
Magnetic flux
concentrator
Spray
quench

Dwelling stage
(fillet preheating)

(a) (b) (c)

(d) (e) (f )

(g) (h) (i)

FIGURE 4.60
Results of FEA computer modeling of the sequential dynamics of vertical scan hardening of a hollow shaft that
comprises sharp fillet, diameter changes, and shoulders using a two-turn MIQ (machined integral quench)
inductor with an “L”-shaped flux concentrator ring (frequency = 9 kHz). Panels a through i show tempera­
ture profiles during different stages of scan hardening. (From V. Rudnev, Induction Heating: Q & A, Professor
Induction Series, Heat Treating Progress, ASM Int’l, Materials Park, OH, September, 2009, pp. 29–32; V. Rudnev,
Computer modeling of induction heating: Things to be aware of, things to avoid, Industrial Heating, May, 2011.)
234 Handbook of Induction Heating

From a heat-transfer perspective, the presence of various cold areas adjacent to the fillet dur­
ing heating creates a significant irregular cooling effect (cold sink effect) in the fillet region
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

compared with neighboring areas. During heating, therefore, more heat is removed from
the fillet area as a result of thermal conduction than is removed from adjacent areas and, in
particular, the shaft. This requires an inductor designed to generate more heat directly or as
close as possible into the fillet to compensate for the intensive heat sink effect there.
From a design perspective, the shaft topology may have certain features (e.g., edges,
sharp corners, cuts, etc.) located near the fillet that are not permitted to be re-austenitized
during the fillet hardening.
In order to increase the power density into the fillet region and areas in its proximity,
the bottom turn was surrounded with an L-shaped magnetic flux concentrator, providing
a favorable condition for selectively preheating the fillet area. Closer coupling of the bot­
tom turn to the fillet compared to the shaft’s body helps focus the heating. Figure 4.60b
shows the temperature pattern at the end of the dwell stage of the fillet’s preheating. Upon
completion of the dwell stage, the shaft fillet is sufficiently preheated and scanning begins.
It is important to remember that induction hardening of steel components is a two-step
process involving heating and quenching. Quenching starts right away with the begin­
ning of scanning. Sometimes when the coil or shaft first begins to move, a higher speed is
applied to quickly move the coil into position so that the spray will strike the area heated
during the dwell. This also helps avoid overheating of the shaft region that is positioned
immediately in the vicinity to the fillet. This initial quick movement of the inductor is
particularly beneficial when a two-turn scanning inductor is used. In other cases, a short
quench delay (0.5 to 2 s) might be applied during the initial coil movement to further assist
in obtaining the adequate thermal conditions in the fillet area before quenching and to
assist in reducing the thermal shock. However, excessive quench delay may lead to a for­
mation of upper transformation products within the martensite.
The scan rate and coil power must be adjusted during scanning in order to accommo­
date changes in part shape (including shoulders, diameter changes, wall thickness varia­
tion, etc.). Computer modeling helps obtain critical information for developing the process
recipe/protocol.
When scanning outside surfaces (e.g., the shaft’s outside diameter), the spray-quench
device is positioned next to the coil. In other cases and as in this study, an MIQ inductor
is used. In either case, a quenching device consists of a quench chamber with numerous
holes (orifices) that allow the quenchant to impinge at a specific angle and distance on the
austenitized surface of the component needed to be hardened.
Figure 4.60c through i shows the results of computer modeling of heating and quench­
ing during the intermediate stages of surface hardening of a shaft.
Preheating and postheating phenomena caused by electromagnetic end effects are
clearly visible. Note that appreciable heating of the shaft begins considerably in front of
the coil leading turn (top turn), creating the preheating effect.
An electromagnetic end effect that deforms the heat source distribution outside the
induction coil is only one of the factors responsible for this preheating. A second factor is
heat flow in the axial direction owing to thermal conduction. A slower scan rate, greater
inductor-to-workpiece radial decoupling, and the use of lower frequencies make the exter­
nal preheating effect more pronounced.
The presence of the external magnetic field behind the trailing turn (bottom turn) pro­
duces the postheating of shaft areas located directly below the coil. In some cases, posthe­
ating may occur even in the regions where the quenchant impinges on the austenitized
surface of the shaft.
Heat Treatment by Induction 235

It is very important to take end effects into consideration to properly design a quench
device and determine the suitable process protocol. In cases of insufficient quenching or
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

borderline quench intensity, the impact of the external magnetic field behind the trailing
turn may be so powerful that it measurably reduces the quenching severity, potentially
creating conditions for the formation of notorious mixed structures with the formation
upper transformation products (e.g., martensitic/bainitic/pearlitic mixtures or “ghost”
networking just to name a few).
The presence of geometrical discontinuities such as shoulders near the shaft diameter
changes distorts the magnetic field generated by the inductor during scan hardening.
Figure 4.61 shows the magnified temperature distributions at intermediate heating stages
(Figure 4.60e and g).
During scanning, the magnetic field preferentially couples to the shoulder on the
shaft (larger diameter corner), generating a greater power density, which could poten­
tially result in overheating and subsequent cracking. Proximity effect can also manifest
itself as a heat source deficit occurring in the undercut region of the shaft and transi­
tion area near the shaft’s smaller diameter, potentially resulting in partially trans­
formed structures upon quenching. The complexity of the electromagnetic field (EMF)
in the area where the diameter changes requires a suitable coil design and appropriate
process recipe that address the surplus of induced power in the shoulder of the large
diameter and the power deficit in the fillet or undercut of the neighboring smaller-
diameter region.
The comet-tail effect is also clearly visible in Figures 4.60 and 4.61, manifesting itself as
heat accumulation in the shaft subsurface regions below the scan inductor [143,192,310,314].
This effect is pronounced in the areas of diameter change. Upon quenching, the tempera­
ture of the shaft surface can be cooled sufficiently below the Ms temperature. At the same
time, the heat accumulated in the shaft subsurface might be sufficient for tempering back
the as-quenched surface regions and could potentially result in the appearance of soft
spots within the case depth. Proper quenching is essential to prevent this undesirable
phenomenon.
The sequential dynamics of the temperature patterns occurring during a horizontal
scan hardening is illustrated in Figure 3.65, exposing the same basic physical phenomena
as vertical scan hardening.

(a) (b)

FIGURE 4.61
Magnified temperature distributions at intermediate heating stages (a) and (b) (Figure 4.60e and g). (From
V. Rudnev, Induction Heating: Q & A, Professor Induction Series, Heat Treating Progress, ASM Int’l, Materials
Park, OH, September, 2009, pp. 29–32; V. Rudnev, Computer modeling of induction heating: Things to be aware
of, things to avoid, Industrial Heating, May, 2011.)
236 Handbook of Induction Heating

4.2.2.1.3 Induction Scanners: Vertical Scanning versus Horizontal Scanning


As stated above, scanners are typically either vertical or horizontal. As an example, Figure
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

4.62 shows two examples of standard vertical induction scan hardening machines of the
Statiscan® family. Being a workhorse for numerous heat treaters over the last decade, the
Statiscan family of induction scanners represents a group of unitized vertical harden­
ing systems that are widely accepted as industry standards. Standard power ratings are
50–300 kW at 10–200 kHz. Equipment is coupled with a standard PLC and HMI with suf­
ficient program recipe storage to provide a very flexible platform for medium duty harden­
ing and tempering of a wide variety of parts.
The Inductoscan VSM95 (Figure 4.63) is a precision vertical-scanning induction machine
designed for handling heavier loads and can have one, two, or four spindles. This equip­
ment is ruggedly designed for high-volume scan hardening of most cylindrical parts up to
200 kg and lengths exceeding 2 m.

FIGURE 4.62
Standard vertical induction scan hardening machines the Statiscan IV. (Courtesy of Inductoheat Inc., an
Inductotherm Group company.)

FIGURE 4.63
The Inductoscan VSM95 is designed for handling heavier loads and can have one-, two-, or four-spindle,
vertical­-scanning precision induction hardening system. This equipment is ruggedly designed for high-volume
scan hardening of most cylindrical parts up to 200 kg and length exceeding 2 m. (Courtesy of Inductoheat Inc.,
an Inductotherm Group company.)
Heat Treatment by Induction 237

The Statiscan systems can be configured for multiple heat-treating operations such as
scanning, single-shot, lift/rotate, and linear transfer, just to name a few. The cooling system
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

includes a closed-loop nonferrous reservoir, heat exchanger, and recirculating system with
a centrifugal pump. Scanning speeds are user-selectable from 1 to 200 mm/s. The Statiscan
family is equipped with a user-friendly onscreen program to provide quick setup, changeover,
and diagnostic capability. To simplify setup, upper tooling centers can be adjusted without any
extra tools. A selection of standard quick-change inductor mountings is also helpful in mini­
mizing changeover time. Its description can be found at www.inductoheat.com.
Vertical scanning lends itself to manual or automatically loaded machines that require a
relatively small space in the process line. It also leads to a somewhat simpler arrangement
for quenching the part once it has been heated.
Because of possible changes in part length during the heating process, the upper center
may freely rest upon the top of the shaft with sufficient weight to ensure rotation of the
workpiece or it may be spring loaded or held by pneumatic pressure. Typically, a cup or
holder of some type is used at the top of the part or a hardened center. In some cases,
where part geometry allows, a positive mechanical “dog” is used to drive the lower end of
the part and ensure rotation. On this type system, the lower tooling is rotated and a rota­
tion sensor is located on the upper center tooling in order to ensure that the part itself is
actually rotating. Some type of rotation detection must be employed to ensure that the part
is actually rotating as it is passing through the heating coil.
On vertical systems where the part is scanned and the heating coil is stationary, the
length of the tower is directly related to the maximum length of the workpiece since it
must be able to lift the part into position at the coil and then index the full part length
until the part exits the coil. For systems utilizing a ball screw driven from the top of the
tower, this requires a tower height that is more than twice the length of the workpiece to be
heated. This is one of the reasons why vertical scanning is best for relatively short parts. A
good way to reduce the tower height on vertical systems is to use a stationary jack screw at
the center of the tower, which then requires only the length of the part and upper tooling
above the coil location.
After the part is processed, it typically would delay at the “Quench Out” position until it
is adequately cooled and then indexed to the unload position where it could be manually
or automatically unloaded.
Figure 4.64 shows a horizontal scanner that provides a maximum scan rate up to
200 mm/s (8 in./s). Workpieces of general cylindrical shape with a maximum diameter of
25 mm are moved end to end in a continuous motion through a three-turn inductor.

FIGURE 4.64
Horizontal scanner that provides a maximum scan rate up to 200 mm/s (8 in./s). (Courtesy of Inductoheat Inc.,
an Inductotherm Group company.)
238 Handbook of Induction Heating

Figure 4.65a shows a horizontal scanner to induction harden components of much larger
diameters: both ends of a trailer axle. A walking beam system was incorporated into the
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

machine for part transfer. At the heating station, the axle is lifted off the beam and the
power supply and inductor are indexed to position for scan hardening. After the comple­
tion of surface hardening of one end, the axle is then lifted off the transfer mechanism
and rotated 180° to induction harden the opposite end. Heavy-duty precision shafting and
bearings are used for stability and consistency. Figure 4.65b shows a close-up of a movable
inductor to scan harden trailer axle ends. Heating time is less than 8 s per axle end.
When lifted off the beam, an axle is indexed into the inductor and the heat can be applied
to selected areas of the workpiece or the entire length can be scanned. A walking beam
system can accommodate different sizes and irregular geometries; however, recent trends
show that it is sometimes more beneficial to use alternative mechanical devices instead of
walking beam systems, including robots for handling irregular shape components.
Induction heat treat practitioners sometimes include continuous or progressive horizon­
tal hardening systems (discussed in Section 4.2.2.2) into a family of horizontal scanners.
The difference is vague and it is a matter of terminology. Some heat treaters consider that
the difference between those systems is primarily related to the number of coils included
in the hardening machine. Systems that consist of a horizontally arranged single induc­
tor are referred to as horizontal scanners. In contrast, if a system consists of two or more
hardening inductors, it is called a continuous or progressive horizontal hardening system.
Both vertical and horizontal scanning systems are a viable means of heat treating com­
ponents. The choice of vertical versus horizontal scan hardening arrangements is usu­
ally associated with the geometry and length of heat-treated components, as well as the
workflow throughout the plant or factory in which the equipment is installed. Horizontal
hardening is often chosen when long workpieces are to be processed (e.g., 1 m or longer) or
when extremely high production rates are needed.
It may take a considerable amount of time with vertical scan hardening to heat treat the
workpiece because it must be scanned along the length up to the position where the heat­
ing process commences and then scanned back down to the load–unload position. This
takes extra time.
The horizontal scanner may exhibit an appreciable benefit from a production standpoint.
A horizontal system is typically set up as a single continuous scanning line that allows

(a) (b)

FIGURE 4.65
Horizontal scanner for hardening both ends of a trailer axle (a). A close-up of a movable inductor to scan harden
trailer axle ends (b).
Heat Treatment by Induction 239

parts to be loaded from a magazine and continuously fed along skewed rollers to the exit
of the machine. Parts may be fed end to end through the heating coil depending on the spe­
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

cific heating requirements for the end of the component and passed on to the next process.
The loading system can push parts through the inductor by a pinch drive mechanism, a
conveyor, mechanical pushers, or other means; for example, the part can be simply moved
on skewed rollers.
On a horizontal system, because of heavy-duty roller support underneath and gravity
plus any required hold down devices on top of the part, the part is maintained in the cen­
ter of the induction coil and quench ring. This may produce lower distortion than a verti­
cal system where the part geometry may change or warp to the extent that the part is not
always centered in the inductor.
However, on the other hand, it may be more difficult to maintain the exact location of
certain features of the part during heating on a horizontal system since the part is not cap­
tured and carried through the coil but is often free rolling on the skewed rollers. Because
these systems are commonly used for quite long parts of appreciable weight, it is difficult
to speed up or slow down the progress of the part along the skewed rollers as quickly as
might be done with a servo-driven carriage that captures the part.
The roller system can interfere with obtaining symmetrical cooling of the workpiece
since the location of the rollers and the rotation detection mechanism on shorter parts may
be too close to the coil or quench barrel. Also, a part hold down fixture may be required
to prevent lighter smaller-diameter parts from being occasionally moved axially by elec­
tromagnetic forces rather than the roller system. As with the vertical system, some type
of rotation detection must be employed to ensure that the part is actually rotating as it is
passing through the heating coil.
In these systems, the part to be heat treated typically has a relatively consistent cross sec­
tion. Some examples would be bars, shafts, or pins. In other cases, parts may have minor
diameter changes and shoulders. If components comprise diameter changes of appreciable
size, then it would be beneficial to have individually controlled short coils and a position
tracking system to appropriately modify the heat generation depending on the positioning
of a particular geometrical feature.
In some cases, keyways or grooves may exist on the part and the location of the particu­
lar geometrical feature or end of the shaft must be sensed and indicated in order to turn
the heat on and off at the appropriate time.
One of the main concerns with horizontal scanning is quenching. When scanning verti­
cally, the quenching is done below the inductor. This naturally allows gravity to pull the
quench fluid down. Thus, with vertical quenching, the quench fluid continues to flow on
the part long after it has passed the quench chamber. This is very beneficial while trying
to reach temperatures suitable for handling.
When quenching horizontally, the effect of gravity is somewhat different and the way
the quenchant falls from the workpiece may be altered at various sections of the part lead­
ing to the probability of nonuniform cooling along the circumference of the heat-treated
component (e.g., quenchant may be running along the top of the part but falling off on its
bottom).
It is also more critical to have a sufficient distance between the inductor exit and
the quenching device because of the higher probability for splashing the liquid quen­
chant back into the inductor, potentially leading to irregular results caused by dif­
ferent cooling rates. This may negatively affect the hardness consistency as well as
the magnitude and distribution of residual stresses. Horizontal systems with a small
angle may help.
240 Handbook of Induction Heating

Therefore, we can summarize that the main process differences between vertical or hori­
zontal scan hardening systems lie in the quenching and part handling. With some scan­
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

ners, splash shields, deflectors, and drip trays may be needed to prevent the backsplash of
the quench fluids.

4.2.2.1.4 Induction Scanners: Movable Inductor versus Movable Part


As previously stated, when a scan processing mode is chosen, either the inductor or the
part or both may be moved during the heating and quenching. When moving the inductor,
both flexible cables and hoses are used or the inductor is hard-bused to the transformer
and the transformer or heat station moves with the inductor. In some cases, the power sup­
ply itself may be moved at a moderate rate to scan a stationary workpiece.
The choice to move the inductor or to move the part is primarily based on the size,
weight, and geometry of the component compared to the size, weight, and geometry of the
inductor: in other words, it depends on which of the two is easier to move.
Weight is an important factor because the movement can occur several hundred times
each day and, in some cases of high production, even several thousand times per day.
For example, during induction surface hardening of links or bottom plates for excavators
as well as track shoes for earth-moving machines that often specify deep hardness case
depths (up to the 24 mm), it is much easier to move the inductor around the workpiece
instead of moving the workpiece, which can weigh several tons (Figure 4.58a). Another
example of moving the inductor is the case study provided above for hardening trailer
axles (Figure 4.65).
The length of the part to be heated is also an important consideration: when a component
is of moderate weight, it is obviously preferable to move the part rather than the inductor.
For example, it is much easier and more cost-effective to design a hardening system that
anticipates moving a workpiece that weighs less than 0.25 kg (<0.5 lb) rather than moving
an entire power supply, as it is shown in Figure 4.64.
In other cases, it may not be practical to move very large and elongated components. It
would consume too much floor space to move the part through a stationary inductor. In the
case of low production rates, the best choice might be to move the inductor, but the length of
the high-frequency power leads could become a problem with respect to power loss and volt­
age drop. In this case, it is preferable to move the inductor with the power supply attached.
Then, the moving cables are operating at a low frequency (50–60 Hz) with lower power loss.
In the case of high production, continuous horizontal systems may be more suitable.
The consideration of the length of the leads (e.g., cables or buses) from the power source
to the inductor is important. They should be as short as possible to conserve energy and to
allow the power source to operate properly without reaching any limits (e.g., voltage limit).
If these leads are too long, the inductance increase can be so significant that it may result in
a substantial power loss and voltage drop. The voltage drop in the leads may even exceed
the inductor voltage. Long leads could net an excessive total needed power, a measurable
reduction in energy efficiency, and potential concerns regarding the process repeatability
owing to the possibility of an inductance change of the flexible leads during their motion.
Whether moving the inductor or moving the part, the system can be designed to be
robust in order to ensure smooth and consistent operation and the production of quality
parts.

4.2.2.1.5 Induction Scanners: Single-Frequency Scanners, Dual-Frequency Scanners


The hardness requirements are often specified to ensure a certain depth, which is directly
associated with the austenitized depth. Among other factors, the depth of the austenite
Heat Treatment by Induction 241

phase is greatly affected by the frequency. In order to have a low capital cost investment,
many scanning machines utilize a single frequency. Single-frequency scanners operate in
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

a set frequency range. For example, a power supply that is said to be a 30-kHz unit may
have a range from approximately 25 to 32 kHz depending on the load-matching specifics.
When scanning a part that has an irregular cross section or large changes in diameter, the
frequency may uncontrollably change slightly. This occurs because the reactance of the
workpiece-to-coil air gap varies as the coupling gap changes. This change in frequency is
usually insignificant (but not always) as far as its influence on the process or case depth.
In the past, single-frequency scanning was the most common technique because of the
limited capability of the power supplies and fairly simple machine concepts that repre­
sented a low capital cost for the equipment.
Several different case depth requirements may be necessary on a single complex-shaped
component (Figure 4.52). If the variations in specified case depths and workpiece geom­
etry are not significantly different, a single-frequency scanner can be used very effectively.
Deeper case depths can be achieved by increasing the power or increasing the heat time
(reducing scan rate). However, caution must be used in obtaining the desired depth by
using this technique.
As discussed in Chapter 3, in some applications, it can be beneficial to use a dual-­
frequency design concept. Low frequency is used during the stage when the workpiece
surface and near-surface area retains its magnetic properties (below the Curie point). In
the next heating stage, when the workpiece becomes nonmagnetic and the current pene­
tration depth is drastically increased, it is more efficient to use a higher frequency (as often
used in through-hardening applications). This approach is not only clearly associated with
an advantage in scanning applications for through hardening but can also be beneficial in
some cases for deep case surface hardening.
Another reason for using a dual-frequency design is to accommodate a component that
has selected areas with different case depth requirements that cannot be achieved by sim­
ply adjusting the power level and scan rate. A dual-frequency scanner can be used to
harden selected areas of the part to scan with different frequencies using different power
supplies.
Section 2.1.1.1 reviews one of such applications for deep hardening (20–25 mm case
depth) of large mill rolls (up to 1.2 m in diameter) utilizing two inductors powered with
two different frequencies. The lead inductor is powered by 50 Hz and the trailing induc­
tor is powered by 250 Hz [26]. The lead inductor provides an in-depth preheating effect,
creating the heat barrier with respect to the cold core. Then, the trailing inductor utilizing
higher frequency completes the austenitization at the required hardness depth.
Another example of dual-frequency scanning is to preheat the roots of teeth in large
gears, sprockets or shafts with teeth. Preheating may also be done when a large change
in the part cross section does not allow a single pass of the inductor to heat deep enough.
In some cases, preheating can be done by scan heating selected difficult-to-heat areas of
the part, and then a soak time is applied before the final heat. Those areas will be heated
more deeply. Lower-power densities required for a preheat and higher-power densities for
a final heat may use the same frequency or a different frequency.
The material of the part being heated may also dictate a necessity of having cer­
tain preheat conditions. Some low-toughness and brittle materials (such as cast irons
or high-carbon steels that have a tendency to crack during rapid heating and spray
quenching) may require a slower heat or more time at temperature before quenching.
In cases like this, the component could be scanned to preheat it using single or mul­
tiple passes.
242 Handbook of Induction Heating

As stated above, dual-frequency scanning may be advantageous for heat treating com­
ponents with irregular shapes or for components where the selected areas require sub­
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

stantially different case depths or when hardening parts with solid and hollow regions.
Although a dual-frequency approach provides important capability to an induction sys­
tem, its utilization is limited to few specific applications. Some of the main drawbacks
to the wide spread of dual-frequency concept deal with higher cost and the necessity of
having two different power supplies instead of one single-frequency power source. This
situation has been recently changed with a development of Statitron® IFP™ Technology,
which provides a capability to instantly change power and frequency during scan harden­
ing. This technology is reviewed at the end of the next paragraph.

4.2.2.1.6 Induction Scanners: Equipment Designs for Greater Flexibility


Quality, price, and delivery have been the three key benchmarks in judging IHT equip­
ment, including induction scanners in the past. Because of a number of current develop­
ments, these three are joined by a fourth that is presently equally important—flexibility
[311–313,401,421,422].
A changing landscape in the area of automotive supply has led to a different set of
requirements for first-tier suppliers through the years. In the past, it may have been com­
mon for a contract to remain with certain suppliers for a period of many years. With the
current conditions, it may be common for a contract to change from one supplier to another on
an annual basis. When this is the case, it is necessary to win the contract, seek out and source
the required equipment to do the processing, purchase the equipment, start the equipment up,
and complete a Production Part Approval Process to be in production in a very short period. In
addition, the equipment must be high quality, very reliable, and easy to maintain. Setup and
maintenance information must be immediately on hand and easily accessible.
Finally, the concern of having a particular contract one year and possibly losing it the
next has given rise to the need for greater flexibility, the new fourth benchmark, in capital
equipment to allow easy retooling and reprogramming of the equipment if necessary to
process different parts in the future.
For capital equipment manufacturers, the consideration of these types of requirements
has led to the development of a new generation of induction heating equipment that will
meet and exceed the need for quality, price, delivery, and flexibility.
In analyzing customer purchases for the last 20 years, Inductoheat undertook a study
to categorize the most common power output levels and frequencies that had been pur­
chased. Further analysis included “what type of parts” were being processed and “what
type of equipment” had been purchased to process the particular parts. The intent was to
determine if there was a common ground where 80% of the requirements could be satis­
fied with a specific type of new equipment. Finally, an evaluation was made as to whether
historical data were consistent with currently observable trends in customer purchasing
preferences.
With consideration of the previous information, a goal was set of meeting all of the
conflicting requirements on a single equipment platform. The result was the development
of an advanced family of modular equipment system called Inductoscan (Figure 4.66)
designed to satisfy the customer’s need to have a top-quality piece of equipment delivered
promptly at a reasonable market price, and still provide the flexibility for the customer to
pick and choose the options desired to customize the equipment to their individual needs
at the present time. One of the major benefits of this type of system is that it does provide
the ability to easily retool and reuse the equipment on another project, possibly even in
Heat Treatment by Induction 243

Inductoscan® family of modern scanners


Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

Standard module Dual power supplies Extended tower

FIGURE 4.66
Three variations of the Inductoscan family of advanced modular scan hardening systems that are quickly
becoming a new industry standard. (Courtesy of Inductoheat Inc., an Inductotherm Group company.)

another plant in the future. The equipment utilizes a common utility point for all utilities,
which facilitates moving, if necessary, from one location to another.
By using a modular approach, a small number of modules can be used to provide many
different machine designs in a tightly integrated package that can be easily picked up
with a lift truck and delivered to the production site or easily moved from plant to plant
as required. Components are designed interchangeability from left to right in order to
allow the use of the same components to build a single or a dual system. This reduces the
number of spare parts that the customer must stock. A single power and plant-water utility
connection greatly simplifies equipment installation.
The power supply modules can be manufactured in a wide variety of power and fre­
quency ratings, and common components are used for single or dual versions of the
equipment. Typical standard ratings are available from 50 to 300 kW and frequencies are
available from 3 to 200 kHz. Other sizes and frequencies are also available upon request.
Part processing modules include a number of different scanning towers in vertical or hori­
zontal orientation, lift rotate mechanisms, and so on. As an example, Figure 4.66 shows
three variations of the modular system to accommodate different process specifics.
The machine control adds the ultimate in flexibility by combining a PC front end with a
PLC machine control to provide the versatility of the PC without sacrificing the reliability
of the PLC hardware for control of the system. The control offers hundreds of advanced
features possible only with a PC. Some of these include virtually unlimited program stor­
age with easy access, an onboard hyper-linked equipment manual of more than 1200 pages,
equipment drawings and parts lists, PLC logic printout, ability to view maintenance and
training videos, built-in help desk, optional integrated data acquisition system, power sup­
ply meter readings with hold settings for any point in the cycle, and so on. Programming
utilizes a simple matrix input with a touch screen, which allows simply touching the item
that is to be changed and inputting the new data through the pop-up keypad.
244 Handbook of Induction Heating

Mechanical modules or mechanisms can be easily mounted to common, predesigned


mounting points with complete access to all mechanical components by removing the
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

bolt-on cover. Three standard scan towers are available with speeds of up to more than
600  mm/s (24 in./s). X-Y adjustment is available through a standard manual base or an
optional motorized tower base.
The cover is designed with a 560-mm (22-in.) part load opening sufficient for side-by-side
robot loading of such appreciable size and weight workpieces as axle shafts. The alumi­
num door with a large viewing panel is available with an optional automatic operator,
light curtain, and safety lock. The intent with the machine and plumbing covers was to
provide a new look, similar to that of a machining center, for a traditional type of induc­
tion equipment with a control that provides state-of-the-art capabilities.
Acceptance of the Modular Induction Hardening System has been overwhelmingly
in favor of the concept, functionality, and appearance of the completed system quickly
becoming a new industry standard. With a standardized modular approach and proactive
inventory stocking, cost and delivery time decrease while quality and flexibility increase
to meet all four of the initial design benchmarks and provide a new generation of standard
heat treating equipment for the future.
As discussed in Section 4.2.2.1.5, various hardness case depths can be obtained by chang­
ing the frequency, power density, and scan rate/heat time. Thus, for a particular combination
of required case depth, initial prior microstructure, chemical composition, and production
rate, there is an optimal combination of frequency, power density, and heat time.
On the other hand, for a given frequency, it might still be possible to achieve the same
desired case depth by using different combinations of power density and scan rate. As an
example, Figure 4.67 shows sketches of the transverse cross section of a straight solid shaft
and three cases of frequency selection for surface hardening. In all three cases, it is required
to achieve the same hardness case depth (indicated as a dotted circle) using three notice­
ably different frequencies designated as “desirable” (a), “too high” (b), and “too low” (c).
The desirable frequency (Figure 4.67a) results in δ at an austenite phase temperature that
approximately exceeds 1.3- to 2-fold the required hardness case depth. Under this condi­
tion, the additional heat generation below the case depth is sufficient to compensate for the
cooling effect (owing to thermal conduction) of the colder core, allowing one to achieve the
required temperature distribution and minimizing the amount of heated metal.
If an available/selected frequency is noticeably higher than the desirable frequency, it
results in a very small δ with respect to the depth needed to be austenitized and hardened.

Required hardness depth Current penetration depth

+ + +

(a) (b) (c)

FIGURE 4.67
Illustration of three cases of frequency selection for surface hardening straight solid shaft, where it is required
to achieve the same hardness case depth (indicated as a dotted circle) using three noticeably different frequen­
cies designated as “desirable” (a), “too high” (b), and “too low” (c).
Heat Treatment by Induction 245

A very small δ might not be sufficient for austenitization of the subsurface areas (Figure
4.67b), if the remaining process parameters are kept the same as the optimal frequency.
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

Therefore, substantial additional time is needed to allow the thermal conduction to pro­
vide sufficient heat flow from the surface and near-surface areas where heat generation
takes place toward the required hardness depth to ensure its proper austenitization in
order to obtain the specified case depth upon quenching. This task could be accomplished
by increasing the heat time (or slowing the scan rate) and must be associated with a cor­
responding power density reduction. Otherwise, the surface may be severely overheated,
producing undesirable metallurgical structures as those shown in Figure 4.36, for exam­
ple. In other words, this approach relies on thermal conduction to compensate for the
drawbacks of using too high a frequency.
It should also be noted that though the lower power and slower scan rate may seem like
a very economical approach, care must be taken when translating “rules of thumb” to
actual production requirements. Though a slow scan rate may not be a problem from the
standpoint of heating the part, a primary consideration for induction hardening is getting
the heated part into the quench fast enough to ensure the transformation to martensite
without forming any of the softer constituents by delaying entering the quench too long.
Unfortunately, such a modification of the process recipe not only adds unnecessary cycle
time but also may lead to some issues related to surface overheating, potentially resulting
in excessive grain coarsening at the surface, grain boundary liquation (Figure 4.36), and
other undesirable metallurgical phenomena.
In contrast to the example discussed earlier, if the available/selected frequency is notice­
ably lower than the desirable frequency (Figure 4.67c), it might produce a much greater
depth of heat generation than necessary. This can result in greater than needed austen­
itized layers and a substantially deeper hardness case depth.
Besides, heating a greater amount of metal than necessary is associated with an unnec­
essary increase in energy utilization and potentially excessive distortion. As expected, a
critical factor that affects shape distortion is related to the amount of heat generation. This
is particularly so when heating irregularly shaped components with a lack of symmetry.
The greater the amount of heated metal, the greater the expansion produced, which wors­
ens shape distortion.
Therefore, in the case of using a lower-than-desirable frequency, it is necessary to sup­
press the thermal conduction utilizing higher than normal power densities and shorter
heat times (faster scan rates). Note: in some cases, a low frequency could produce such a
large δ compared to the maximum permissible case depth that it might not be possible to
meet the pattern specification.
Keeping in mind the aforementioned shortcomings of using other-than-optimal fre­
quencies, it should be reemphasized that, in many cases, it is possible to achieve a desirable
hardness pattern and acceptable metallurgical characteristics by suppressing or enhanc­
ing the thermal conduction and using an appropriate but a different-from-optimal combi­
nation of frequency, power density, and scan rate/heat time.
This can be easier done when hardening straight shafts without geometrical irregu­
larities and constant required case depth or its minor variations. Unfortunately, simple
geometries are the exception. In order to fulfill complex functionality requirements and
lightweight high-performance initiatives, many induction surface-hardened components
consist of discontinuities and geometrical irregularities. Typical examples are parts con­
taining longitudinal or transverse holes, keyways, grooves, shoulders, flanges, diameter
changes, fillets, undercuts, splines, threads, and so on (Figure 4.38). These features are not
unique as they are commonly found in many components and can substantially deform
246 Handbook of Induction Heating

EMF and eddy current flow, potentially compromising metallurgical results of induction
heat treatment. As expected, a better way would be using the desired frequency that is
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

optimally suited to the needed hardness case depth and addressing the geometrical spe­
cifics of a component.
The surface condition of the workpiece is another factor, which might have a pronounced
effect on heat-treating practice and workpiece life during service. Under certain condi­
tions, voids, microcracks, notches, improper chamfers, and other surface and subsurface
discontinuities and stress concentrators can initiate crack development during induction
hardening when the metal goes through the “expansion–contraction” cycle. Thermal gra­
dients and stresses can reach critical values and “open” notches and microcracks; thus, a
probability of their presence should be properly addressed during the process develop­
ment stage in particular when using other-than-optimal process parameters such as fre­
quency and power density.
Geometrical discontinuities distort the magnetic field generated by the inductor, which
is potentially affected by different electromagnetic phenomena, including the proximity
effect and end effect, as well as the edge effect, slot effect, and the ring effect, just to name
a few. These effects can cause the appearance of hot or cold spots, undesirable microstruc­
tures, excessive shape distortion, and cracking. Unfortunately, these undesirable phenom­
ena might be unavoidable when using a single frequency to heat treat complex geometry
components.
Global competitiveness demands that heat treaters must be flexible in their manufac­
turing methods and equipment purchases [17,23,143,399,401,421,422,538]. Long-term cus­
tomers may move their production at a minute’s notice. The heat treater must be able to get
new business to cover the lost business. Therefore, it would be very beneficial to not have
to replace an entire induction heating system or a significant portion of it (such as power
supply or heat station, for example) when sudden market changes occur.
Material grade, production rate, hardness pattern, case depth, workpiece size, topology,
and topography, as well as many other factors may change when market demand changes,
providing new opportunities for commercial heat treaters. All these dictate the necessity
to further maximize the flexibility and processing capability of equipment purchased in
order to meet a wide range of product requirements with minimal equipment changes.
In some cases, irregular regions of parts may require hardening to different depths.
As an example, Figure 4.52 illustrates the hardness pattern obtained with variations and
geometrical complexity when induction surface hardening different components. Thus,
it might be even more difficult to obtain the required case depths and hardness patterns
while using higher-than-desirable or lower-than-desirable frequencies.
Unfortunately, the great majority of commercially available inverters have been designed
to provide a certain frequency of output alternating current (AC) that cannot be instantly
changed during the scanning operation. Therefore, the depth of heat generation (that is
related to δ) might not be optimal for a particular portion of the heat-treated part, its geo­
metrical features, and localized min–max case depth requirements. In some cases, the
available frequency might be substantially lower (in folds) than optimal, but in other cases,
it might be significantly higher for obtaining a particular hardness case depth in a certain
area of the part during scan hardening.
Thus, conventional scan hardening technologies must rely on a compromise between
the achieved quality, production rate, and process capability. Regardless of the fact that
recipe modifications can help minimize the negative impact of using other-than-optimal
frequencies by relying on suppression or enhancement of thermal conduction, they often
cannot eliminate it. Depending on the workpiece’s geometry, this could limit the ability
Heat Treatment by Induction 247

to meet a tight pattern specification or might negatively affect the achieved metallurgical
quality of heat-treated components.
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

According to the theory of induction heating, the applied frequency is the most powerful
parameter that directly affects the depth of heat generation (see Section 3.1.2, Equations 3.6
and 3.7). Obviously, in order to address geometrical subtleties of heat-treated components
in an optimal manner, it would be much more beneficial to apply various combinations
of frequencies, power densities and scan rates at various stages of the scan hardening
operation. This would maximize production and improve quality of the heat-treated com­
ponents. Regrettably, the great majority of available inverters do not have such capability.
However, a new generation of Inductoheat’s patented transistorized inverters elimi­
nates this limitation and simplifies achieving the required hardness pattern specifications,
allowing independently and instantly change frequency and power (Figure 4.68a) dur­
ing induction scan hardening. This technology optimizes electromagnetic, thermal, and
metallurgical conditions when heat treating components with geometrical irregularities
having appreciably different masses and combinations of solid and hollow areas, diameter
changes, and various required hardness case depths.
As an example, Figure 4.68b shows the Statitron IFP inverter that provides Independent
Frequency and Power control. The ability to independently and instantly (like a CNC
machine) change both frequency and power during the heating cycle is what commercial
heat treaters have long been aspiring for, since it provides the greatest process flexibility.
Statitron IFP is an IGBT-type power supply specifically designed for induction harden­
ing applications, allowing instant and independent adjustment of both frequency (within
5–60 kHz range) and power (up to 450 kW).
For example, if a heat treater presently processes 12-mm-diameter shafts with a nominal
hardness case depth of 1.8 mm, he or she might apply a frequency in the 50- to 60-kHz
range. However, if tomorrow’s market would require processing 30-mm-diameter shafts
with a nominal 5-mm case depth, then the heat treater would possibly select a lower fre­
quency such as 5–6 kHz to ensure a more in-depth heating effect without compromising
the metallurgical quality of the product.
The unique ability to change the frequency instantly by more than 10-fold can also be
advantageous for machines that need to provide hardening and tempering operations. In

Frequency
Frequency, power

Power

Process time

(a) (b)

FIGURE 4.68
New generation of patented transistorized inverters allows instantly and independently controlled frequency
and power during induction scan hardening (a). The Statitron IFP inverter that provides Independent Frequency
and Power control (b). (Courtesy of Inductoheat Inc., an Inductotherm Group company.)
248 Handbook of Induction Heating

this case, higher frequencies can be used for hardening and lower frequencies can be used
for induction tempering and stress relieving.
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

The Statitron IFP concept effectively addresses the need of modern industry for superior
process flexibility, substantially expands heat treat equipment capabilities, and improves
the metallurgical quality of induction heat-treated components optimizing the harden­
ing patterns and processing parts with irregularities by programming power and fre­
quency changes on the fly (Figure 4.68a), which is imperative for modern scan hardening
operations.

4.2.2.2 Continuous and Progressive Hardening


This method is commonly applied when heat treating elongated workpieces, such as bars,
tubes, rods, wires, plates, beams, and others. Long parts are more readily processed in
a horizontal manner, being heated as they progressively pass through multiple induc­
tors. Inductors are positioned in-line or side by side. Each inductor may have a different
design and power/frequency setting. This type of hardening is not limited to horizontally
processed parts; vertical processing and arrangements at certain angles are possible, if
required. There are also cases when a part is statically heated to a certain temperature and
then progressively moved to another static inductor for the next heating stage. All these
processes are referred to as progressive heat treatment.
As mentioned in Section 4.2.2.1.3, induction practitioners sometimes consider continu­
ous or progressive horizontal hardening systems as horizontal scanners. The difference
is vague and it is a matter of terminology. Some heat treaters feel that it would be appro­
priate to differentiate these systems based on the number of inductors included in the
hardening machine design. Horizontal systems consisting of a single inductor are com­
monly referred to as horizontal scanners. In contrast, if a system consists of two or more
hardening inductors, then it might be referred to as a continuous or progressive horizontal
hardening system.
With the continuous hardening method, the workpiece is moved in a continuous motion
through a number of in-line inductors. Multiturn solenoid coils and, to lesser a degree,
channel-style inductors and split-return inductors are most typically used in continuous
heat treating lines. Figure 4.69 shows two continuous induction systems consisting of a
number of in-line coils. Figure 4.69a shows a side view of a horizontally arranged induc­
tion system consisting of three in-line coils. Each coil consists of three turns. Figure 4.69b
shows another example that comprises four in-line coils and a spray quench device posi­
tioned after the last inductor. Workpieces (e.g., bars, shafts, rods, pins, etc.) are processed
end to end through the inductors in a continuous motion.
Progressive multistage hardening is used when multiple workpieces are moved (via
a pusher, indexing mechanism, robot, walking beam, etc.) through a number of coils.
Therefore, the entire component or its portions are sequentially heated (in a progressive
manner) at certain predetermined heating stages inside the in-line horizontal (being more
typical) induction heater or a multiposition horizontal or vertical heater where coils are
positioned side by side.
Continuous or progressive hardening methods are typically used for through harden­
ing of elongated or moderate-length parts processing end to end and, to a lesser degree,
for surface hardening. Outside diameters for case hardening usually vary from 1/2 in.
(12 mm) to 4 in. (100 mm). In through hardening applications of solid cylinders, the diam­
eters may be as small as 3 mm.
Heat Treatment by Induction 249
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

(a)

(b)

FIGURE 4.69
Two examples of horizontal continuous induction systems consisting of a number of in-line coils. (a) Shows a
three in-line coils (each coil consists of three turns). (b) Shows a top view of four in-line coils and spray quench
device. (Courtesy of Inductoheat Inc., an Inductotherm Group company.)

Similar to induction heating before hot forming (Figures 3.54 and 3.55), in through hard­
ening applications, it is possible to recognize three heating stages:

1.
Initial or magnetic stage. Temperatures anywhere within the workpiece are below
the A2 critical temperature; thus, the steel is ferromagnetic (Figure 3.54a) and the
δ is quite small, almost regardless of applied frequencies greater than 1 kHz. Skin
effect is fairly pronounced at this stage and the heat source distribution resembles
a conventional exponential distribution. The maximum power density is located
at the surface and sharply decreases toward subsurface and the core. Heat source
generation is localized by the fine surface layer of the workpiece. This leads to a
rapid increase in temperature at the surface with a minor change in the core. This
stage is characterized by high electrical efficiency often reaching 90% or so. As can
be seen in Figure 3.54a, the temperature profile does not match the power density
distribution because of thermal conductivity, which spreads the heat from the sur­
face and near-surface area toward the core.
2.
Interim stage. During this stage, the surface and near-surface layer is heated above the
A2 critical temperature (Figure 3.54b); however, the internal region, having tempera­
tures below the Curie point, retains its ferromagnetic properties. At this point, the
power density distribution along the radius has a unique nonexponential “wave-like”
distribution, which is very different from the commonly assumed exponential distri­
bution. The cause for this behavior will be discussed in Section 4.2.2.4.
250 Handbook of Induction Heating

3.
Final heating stage. The thickness of the surface layer that exhibits nonmagnetic prop­
erties becomes greater than the δ in hot steel at a given frequency, and the “wave-
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

like” distribution disappears. The classical exponential power density distribution


will then take place. As expected, δ increased dramatically during the previous two
stages (Figure 3.14), resulting in a more in-depth heating effect. With time, the core
temperature exceeds the Curie point and the entire cross section will be nonmag­
netic (Figure 3.54c). Keep in mind that the final temperature shown in this case
study (Figure 3.54) is typical for hot forming applications and not for through hard­
ening where the maximum temperature should be lower (required hardening tem­
peratures for different alloys have been discussed above in Section 4.1).

Depending on the application specifics, the same frequency may be used for various
coils. In other cases, power levels and frequencies may be different at the different heating
stages.
When using different frequencies for the various heating stages, the coil design may need
to change as well (e.g., a number of coil turns may need to be adjusted for load matching).
Just as the eddy current penetration depth in the heated part is affected by the frequency,
the current flow in the inductor is affected as well. The wall thickness (i.e., copper tubing
wall) of the inductor might need to be adjusted to accommodate for different frequencies
to maximize the coil electrical efficiency. As the frequency is decreased, the copper wall
thickness might need to be increased. Selection of the proper copper wall thickness for
inductors is discussed later in the chapter.
As an example, Figure 4.70 shows a continuous hardening and tempering system for
magnetic steel wires with diameters from 6 mm (1/4 in.) to 12 mm (1/2 in.). The produc­
tion rate is 35 m/min. The system consists of a number of in-line solenoid multiturn coils.
Medium frequency is used for wire preheating. High frequency is used for reaching aus­
tenite phase temperatures followed by quenching. The wire is sequentially induction tem­
pered using a lower frequency after the completion of hardening.
A dual-frequency concept can be beneficial in many applications including through
heating of relatively thin products or materials that exhibit low toughness. According to
this technique, a lower frequency is used during the initial heating stage, when the steel

FIGURE 4.70
Continuous hardening and tempering system for magnetic steel wires with diameters from 6 mm (1/4 in.) to
12 mm (1/2 in.). Production rate is 35 m/min. (Courtesy of Inductoheat Inc., an Inductotherm Group company.)
Heat Treatment by Induction 251

is magnetic. In the final heating stage, when the steel becomes nonmagnetic with signifi­
cantly increased δ and becomes substantially more ductile, it is beneficial to use a higher
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

frequency. As an example, consider the induction heating of a 1/8-in.-diamater (3.2-mm-


diameter) carbon steel rod from ambient to 1100°C using both a single 10-kHz frequency
and dual 10-kHz/200-kHz frequencies (Figure 4.71) [319].
In this case, when using the single frequency of 10 kHz (Figure 4.71a), the rod’s final tem­
perature experiences very little change regardless of the coil power that is increased more
than fivefold (from 17 to 90 kW). The only noticeable difference is related to the initial slope
of the temperature–time curve, where the steel is magnetic. Upon reaching the Curie point,
there is no noticeable temperature rise. This is the result of eddy current cancellation.
In contrast, Figure 4.71b shows that a dual-frequency approach provides a remark­
able improvement in the ability to heat the rod above the Curie temperature. A power of
14 kW/10 kHz was used to heat the rod below the Curie point and a power of 19 kW/200 kHz
was used above it. The total required power is only 33 kW, compared with 90 kW using just
10 kHz, which was still unable to provide the required temperature rise.
One might think that cancellation only occurs when IH relatively thin (such as 1/8-in.-
diameter) workpieces; this is not true. The phenomenon of eddy current cancellation is
not only a function of the physical size of the workpiece but rather the ratio of diameter/­
thickness to δ. For example, similar results to the case study shown in Figure 4.71a with
severe eddy current cancellation will occur when heating 1-in.-diameter (25-mm-­diameter)
steel bars using 1 kHz or heating 4-in.-diameter (102-mm-diameter) bars utilizing line fre­
quency. The above-discussed study stresses the importance of avoiding eddy current can­
cellation in through heating applications such as through hardening.
In some not too often cases, three frequencies may be used. Lower frequency is applied
for preheating inductors, a medium frequency is used for midheat inductors, and a high
frequency is used for final heat inductors.
Sometimes, it is required that the induction system should be able to heat a variety of sizes
using a single frequency. In these cases, in order to provide efficient heating, it is necessary

Single frequency of 10 kHz and Dual frequencies (10 kHz and 200 kHz)
20 in. (0.51 m) long coil and two coils, each 10 in. (0.25 m) long
1200 1200

1000 1000 Frequency = 10 kHz


Coil power = 14 kW
Temperature, °C

Temperature, °C

800 800

600 600 Frequency = 200 kHz


Coil power = 19 kW
Frequency = 10 kHz
400 400
Coil power = 17 kW

200 Coil power = 90 kW 200

0 0
0 0.5 1 1.5 2 2.5 0 0.5 1 1.5 2 2.5
Heat time, s Heat time, s
(a) (b)

FIGURE 4.71
Illustration of dual-frequency concept when induction heating of a 1/8-in.-diameter (3.2-mm-diameter) carbon
steel rod from room temperature to 1100°C using both a single frequency of 10 kHz (a) and dual frequencies of
10 kHz/200 kHz (b). (From V. Rudnev, Systematic analysis of induction coil failures, Part 11c: Frequency selec­
tion, Heat Treating Progress, January/February, 2008, pp. 27–29.)
252 Handbook of Induction Heating

to choose a frequency that will guarantee that the O.D./δ ratio exceeds 3.6 for any workpiece
diameter or heating stage. Thus, it is important to remember that when calculating current
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

penetration depth, the values of ρ and μr of the heated material should correspond to their
values at the highest temperature that occurs during the entire heating cycle.
In applications requiring a high process speed of heated components through a multi­
turn solenoid inductor of appreciable length, there might be noticeably different kilowatt
losses in different sections of a long inductor and multiple water-cooling circuits may be
needed to assure sufficient coil copper cooling as shown in Figure 4.72.
So far, we have discussed the application of conventionally designed solenoid coils in
continuous/progressive hardening applications. However, even multiturn solenoid-type
coil geometries may have quite complex shapes accommodating the shape of induction-
hardened components. One illustration of this is shown in Figure 4.73 where two in-line
multiturn solenoid-type inductors are used for heat treating support beams.
Besides multiturn solenoid coils, channel-type multiturn inductors (also called slot or
skid inductors) are frequently used in continuous/progressive heat treating. The channel
inductor gets its name from its similarity to a long channel. This shape allows parts to be
passed through the coil in a number of ways such as a conveyor, shuttle, indexing, rotary
or carousel table, turntable, or any other indexing system.
Channel coils permit easy entry and exit of the heated components to/from the inductor.
Figure 4.74 shows images of some examples of multiturn channel inductors. The crossover
ends of channel coils are bent away to allow the part to pass through. In some cases, the
crossover ends are made high enough to ensure minimum impact on the heating of the
part at the ends of the coil and also minimizing electromagnetic forces when workpieces
enter and exit the inductor. In other cases, the opposite might be true and crossover coil
regions play an important part in providing the needed temperature distribution.

FIGURE 4.72
Multiple water-cooling circuits may be needed to assure sufficient coil copper cooling. (Courtesy of Inductoheat
Inc., an Inductotherm Group company.)

FIGURE 4.73
Two in-line multiturn solenoid inductor of a complex shape. (Courtesy of Inductoheat Inc., an Inductotherm
Group company.)
Heat Treatment by Induction 253
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

FIGURE 4.74
Images of different examples of multiturn channel inductors. (Courtesy of Inductoheat Inc., an Inductotherm
Group company.)

Channel coils are used to heat treat selected regions of parts, as well as entire compo­
nents. These inductors are often used for through hardening, annealing, and tempering
applications. However, if a specific case depth is required, rotation of the workpiece may
be needed to even case depth.
Figure 4.75 shows a “state-of-the-art” continuous fed induction system for heat treating
fasteners [320]. This system is adjustable for a wide range of fastener/bolt diameters and
lengths (12–102 mm [0.5–4.0 in.]) and is capable of production rates of up to 600 fasteners
per minute.
The unique proprietary coil design developed by Radyne Corp. maximizes electrical
efficiency and system flexibility while preventing stray heating of electrically conductive
surroundings that may potentially cause undesirable heating of structures and malfunc­
tion of electronic devices. The rotary dial tooling is designed to accept bolt fasteners from
the in-line vibratory feeder. The adjustable speed rotary table contains advanced safety
features to prevent damage and meltdown.

FIGURE 4.75
Continuous fed induction system for heat-treating fasteners [320]. (Courtesy of Radyne Corp., an Inductotherm
Group company.)
254 Handbook of Induction Heating

The quench assembly allows adjusting the quench flow for the utmost in quench con­
trol. After spray quenching, parts are stripped from the traverse assembly and dunk
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

quenched into the tank for final cooling to room temperature. After induction hardening,
the required hardness range for the first three threads is typically 52–55 HRC.
The tooling is designed with a quick change feature to ensure that all tooling can be
changed for a different part size in less than 15 min. The system is controlled through a con­
trols package and HMI for part setup and part storage of different programs. Through this
HMI, the power source coil “Z” adjustment can also be stored and adjusted for different bolt
lengths assuring superior quality fasteners. This unit includes four sizes of tooling required
for the rotary heat treat fixture and the traverse tooling: M6, M8, M10, and M12 [320].

4.2.2.3 Single-Shot Hardening
The use of scan hardening on the stepped shafts with large shoulders, multiple and siz­
able diameter changes, and other geometrical irregularities may produce severely nonuni­
form hardened patterns. In cases like this, a scan hardening inductor would be designed
around the largest diameter that would have a sufficient clearance for safe part process­
ing. However, the shaft’s diameter variations, to a significant extent, will result in a cor­
responding substantial deviation in the workpiece-to-coil coupling in different sections
of the shaft. Besides that, sharp corners have a distinct tendency to overheat owing to the
buildup of eddy currents in particular when medium and high frequencies are used. The
electromagnetic end and edge effects may also cause the shoulders to severely overheat
while the smaller-diameter area near the shoulder (including undercuts and fillets) may
have noticeable heat deficit. These factors may result in producing a hardness pattern that
might grossly exceed the required minimum and maximum case depth range, making it
unacceptable. A single-shot inductor is usually a better choice in such applications.
With the single-shot method, neither the shaft nor the coil move linearly relative to each
other; the part typically rotates instead. Single-shot inductors are made of tubing or CNC
machined from solid copper to conform to the area of the part to be heated. This type of
inductor requires the most care in fabrication because it usually operates at high power
densities and the workpiece’s positioning is critical with respect to the coil copper profiling.
According to the single-shot hardening method, the entire region to be hardened is
heated at the same time. In order to provide the required temperature distribution before
quenching, the heat is sometimes applied in several short bursts (pulse heating) with a
timed delay between them to allow for thermal conduction toward the areas that might be
difficult to heat.
When scan quenching is used, some areas of an irregular workpiece may not quench ade­
quately. These problems may also make single-shot hardening a better choice. Therefore, it
is advantageous to use single-shot inductors when the workpiece has certain geometrical
discontinuities of appreciable size including diameter changes, sharp shoulders, fillets,
and so on.
Single-shot hardening may also be the preferred choice when shorter heat times/high
production rates are desired. For example, in some applications, the time of heating for
single-shot hardening can be as short as 2 s, though 4 to 8 s is more typical.
Single-shot inductors typically require higher power levels compared to a scan harden­
ing because the entire area of the workpiece that needs to be hardened is austenitized at
once. The use of high powers and power densities combined with complex geometry can
shorten the life of the inductor. For this reason, single-shot inductors often have shorter
lives than scan inductors.
Heat Treatment by Induction 255

It is always important to keep in mind that, electrically speaking, the inductor is often
considered as the weakest link in an induction system. For this reason, most inductors
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

have a separate coil-cooling circuit and a separate part-quenching circuit. The inductor
will fail if power is increased to the point at which the water cannot adequately cool it.
Additional cooling passages may be needed with high power density single-shot induc­
tors. A high-pressure booster pump is frequently required.

4.2.2.3.1 Conventional Design of Single-Shot Inductors


Unlike scanning inductors, traditional designs of single-shot inductors produce predomi­
nantly an axial eddy current flow rather than a circumferential one. Figure 4.76 shows
a typical single-shot hardening component (top left) and a variety of conventionally
designed single-shot inductors for surface hardening shaft-like workpieces. Sometimes,
these inductors are also referred to as channel inductors.
A conventional single-shot inductor consists of two legs and two crossover segments,
also known as bridges, “horseshoes,” or half-loops. The induced eddy currents under the
legs primarily flow along the length of the part (longitudinally/axially) with the exception
of the regions of the workpiece located under the crossover segments where the flow of the
eddy current is half circumferential.
With a predominantly axial eddy current flow, the heat uniformity in the diameter
change areas of the workpiece dramatically improved and the tendency of corners and
shoulders as well as regions neighboring the diameter changes to be overheated is reduced.
Because the coil copper does not completely encircle the entire region needing to be
heated, rotation must be used to create a sufficiently uniform austenitized surface layer,
which, upon quenching, will produce a uniform case depth along the circumference of
the part. For single-shot inductors, the rotation speed usually ranges from 120 to 500 rpm.
Laminations and other types of magnetic flux concentrators complement the copper pro­
filing in order to achieve the required hardness pattern. Magnetic flux concentrators (also
called flux intensifiers, flux controllers, flux diverters, etc.) provide several considerable

FIGURE 4.76
Example of typical single-shot hardening component (top left) and a variety of conventionally designed single-
shot inductors for hardening shaft-like parts. (Courtesy of Inductoheat Inc., an Inductotherm Group company.)
256 Handbook of Induction Heating

benefits when applied in single-shot inductors, including an increase of coil electrical effi­
ciency, a noticeable reduction of coil current, and a reduction of the external magnetic field
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

[1,51,322–324].
As an example, Figure 4.77 shows a transverse cross section of a single-shot inductor and
a straight shaft. A computer-modeled EMF distribution of the case of a bare inductor (left)
compared to an inductor with a U-shaped flux concentrator (right) is shown. Note that the
scale of magnetic field intensity on both images is different. The use of U-shaped magnetic
flux concentrators typically results in a 16%–27% coil current reduction compared to using
a bare inductor while having a similar heating effect. A reduction of the external magnetic
field exposure is even more dramatic. A detailed discussion regarding the subtleties of
using magnetic flux concentrators is provided in Section 4.7.3.
Sufficient rotation is critical when using any single-shot inductor design. There should
be at least eight full rotations per heat cycle, depending on the size of the inductor and the
design specifics. Figure 4.78 shows the dynamics of the temperature profile during the
single-shot heating cycle of a 40-mm-diameter solid shaft having an inappropriate rate of
rotation. Shorter heating times require faster rotation speeds.
An appropriate inductor design with a closely controlled and monitored rotation speed
will produce a hardness pattern with minimum circumferential and longitudinal tempera­
ture deviations, which will produce sufficiently uniform hardness patterns (Figure 4.79a–d).
Failure to ensure proper rotation speed as well as the use of worn centers (lacking grabbing
force resulting in slippage) could potentially lead to severe localized overheating and even
melting (Figure 4.79e).
Though single shot hardening is commonly used to heat treat relatively short or
­moderate-length components, there are cases when the workpiece can have an appreciable
length. As an example, Figure 4.80 shows a case-hardened axle shaft (Figure 4.80a) using a
single-shot inductor with laminations selectively positioned along the length. Figure 4.80b
shows a single-shot axle shaft hardening and tempering system that is capable of process­
ing more than 175 axle shafts per hour and, at full capacity, is able to deliver approximately
1 MW of power.
Single-shot inductors can also be successfully used for hardening noncylindrical compo­
nents. This includes workpieces of general conical shapes (e.g., elliptic, parabolic hyperbolic
geometries) and some others. As an example, Figure 4.81 shows induction surface-­hardened
ball joints (ball studs) and the single-shot inductors used to harden them. Ball studs are
used in automotive, off-road, and agricultural machinery and can be different in shape and
size (Figure 4.81, compare images on the left with images on the right), requiring noticeably

(a) (b)

FIGURE 4.77
Computer-modeled EMF distribution in the case of a bare inductor (a) compared to an inductor with U-shaped
flux concentrator (b). Note: the scale of magnetic field intensity on both images is different.
Heat Treatment by Induction 257

Dynamics of single-shot induction heating of


40 mm (0.04 m) OD solid shaft using 10 kHz
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

(inappropriate rotation speed)

0.02 Legend 0.02 Legend

Initial 400 400


0.015 0.015
heating 360 360
0.01 0.01
stage 320 320

Temperature, °C
0.005 280 0.005 280
240 240
0 0
200 200
–0.005 160 –0.005 160
120 120
–0.01 –0.01
80 80
–0.015 40 –0.015 40
–0.02 0 –0.02 0
0 0.005 0.01 0.015 0.02 0 0.005 0.01 0.015 0.02

0.02 Legend 0.02 Legend

600 600
0.015 Interim 0.015
550 550
heating
0.01 500 0.01 500
stage

Temperature, °C
0.005 450 0.005 450
400 400
0 0
350 350
–0.005 300 –0.005 300
250 250
–0.01 –0.01
200 200
–0.015 150 –0.015 150
–0.02 100 –0.02 100
0 0.005 0.01 0.015 0.02 0 0.005 0.01 0.015 0.02

Legend Legend
0.02 0.02
1100 882.3 1100
0.015 0.015
1020 860.7 1020
Final
0.01 940 0.01 940
heating 888.5
Temperature, °C

0.005 stage 860 0.005 942.8


860
780 780
0 0 1007.3
700 700
–0.005 620 –0.005 1063.6 620
540 540
–0.01 –0.01
1095.7
460 460
–0.015 –0.015 1095.6
380 380
1064.6
–0.02 300 –0.02 300
0 0.005 0.01 0.015 0.02 0 0.005 0.01 0.015 0.02

FIGURE 4.78
Dynamics of the temperature profile during the single-shot heating cycle of the 40-mm-diameter solid shaft
using an inappropriate rotation.

different hardness patterns. Substantially different workpiece-to-inductor electromagnetic


coupling does not permit using multiturn solenoid inductors in these applications, yet
single-shot inductors allow the geometrical specifics of these components to be properly
addressed, producing the required hardness patterns at minimum process times.
Longitudinal leg sections of single-shot inductors and their crossover segments can be
profiled by relieving selected regions of the copper to accommodate certain geometrical
258 Handbook of Induction Heating
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

(a) (b) (c) (d) (e)

FIGURE 4.79
Inductor design with closely controlled rotation speed will produce a hardness pattern with minimum circum­
ferential temperature deviations (a–d). Failure to a proper rotation could potentially lead to severe localized
overheating and even melting (e).

FIGURE 4.80
Single-shot axle shaft hardening and tempering system that is capable of processing more than 175 axle shafts
per hour and at full capacity. It is capable of delivering approximately 1 MW of power. (Courtesy of Inductoheat
Inc., an Inductotherm Group company.)

FIGURE 4.81
Surface-hardened ball joints (ball studs) and single-shot inductors used for its hardening. (Courtesy of
Inductoheat Inc., an Inductotherm Group company.)
Heat Treatment by Induction 259

features of the workpiece (Figure 4.82). Sections of a single-shot inductor with narrower
heating surfaces facing the shaft increase the locally induced power density in the desir­
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

able regions.
Proper hardening of such components as output shafts, flanged shafts, planet carriers,
yoke shafts, sun shafts, intermediate shafts, drive shafts, and some others may require
extensive copper profiling, making a hardening inductor a very complex electromagnetic
device.
For workpieces containing fillets, it is often necessary to increase the heat intensity in
the fillet region owing to the geometrical specifics including the necessity to heat a greater
volume of metal (as discussed earlier in Section 4.2.2.1.2). Also, the larger metal mass in the
proximity of the heated fillet and behind the region to be hardened produces a substantial
thermal “cold sink” effect [319]. This draws heat from the fillet owing to thermal conduc­
tion, which must be compensated for by inducing additional heating energy in the fillet
area. The required energy surplus can be achieved by narrowing the current-carrying face
of the crossover segment of the single-shot inductor. For example, if the current-carrying
portion of the inductor heating face is reduced by 50%, there is a corresponding increase
in current density, as well as the eddy current density induced within the respective shaft
region. According to the Joule effect, doubling the induced eddy current density increases
the induced power density roughly by a factor of 4. Also, attaching a magnetic flux con­
centrator to certain areas of the hardening inductor further enhances the localized heat
intensity (Figure 4.82).
When using a single-shot inductor, it is particularly important that the workpiece is not
mislocated in the heating position and that the part-to-coil location be consistent because
it noticeably affects the heat treat pattern.
Conventionally designed single-shot inductors may exhibit high process sensitivity that
is associated with the electromagnetic proximity effect. A change in positioning of the
shaft inside the single-shot inductor attributed to bearing wear of the centers, improper
machining of the centers, incorrect part loading in the inductor, and other factors produces
variation in the hardness pattern (particularly within the fillet region). These results in a
reduced hardness depth and the formation of microstructural products associated with
incomplete phase transformation; it also alters residual stresses.
When scan hardening parts with diameter changes, the scan coil should have sufficient
gap to clear the largest diameter. When scanning the section(s) of the shaft with smaller
diameters, an inductor-to-shaft air gap might be very large, resulting in low electrical effi­
ciency and potentially exhibiting difficulties in controlling the pattern along the length

FIGURE 4.82
Longitudinal leg sections of single-shot indictors and their crossover segments can be profiled by reliev­
ing selected regions of the copper to accommodate workpiece geometrical features. Attaching a magnetic
flux concentrator to certain areas of the inductor further enhances localized heat intensity. (From V. Rudnev,
A. Goodwin, S. Fillip, W. West, J. Schwab, S. St. Pierre, Keys to long-lasting hardening inductors: Experience,
materials, and precision, Adv. Mater. Processes, October, 2015, pp. 48–52.)
260 Handbook of Induction Heating

of the part. Some difficulties may be experienced in controlling the hardness pattern in
regions (e.g., geometrical irregularities) where good control is most needed.
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

In contrast to scan hardening, a single-shot inductor can be contoured along the length
of the part. The use of flux concentrators helps drive the current into the desired areas and
allow producing a sharp well-defined hardness profile. The trade-off here is that much
more finesse is required in the design stage to produce the properly profiled single-shot
inductor at the lowest possible cost. Errors are costly since these coils are each custom
made for a given part or application and modifications can be quite costly. Thus, computer
modeling is almost always required to keep the development cost down.
In summary, there are good reasons for using either a single-shot hardening or a scan-
hardening approach. The decision must be based on many factors such as product quality,
production rate, geometry, design proficiency, limitations of available equipment, reliabil­
ity requirements, cost considerations, and so on.

4.2.2.3.2 Single-Shot Inductor Breakthrough


As discussed in Section 4.2.2.3.1, there are several ways to control heat generation at dif­
ferent regions of a heat-treated shaft utilizing single-shot inductors. This includes induc­
tor copper profiling resulting in a variation of coil-to-shaft electromagnetic coupling or
machining different widths of the current-carrying faces of the inductor. Besides that, a
magnetic flux concentrator can be attached to certain areas of the hardening inductor to
enhance localized heat intensity as shown in Figure 4.82.
When applying single-shot hardening of complex-shaped shafts, coil life is often limited
because of the necessity to dramatically “squeeze” the current in a certain area, maximiz­
ing the power density in selected regions (e.g., fillets).
The effects of intensifying the heat generation in selected areas of the shaft (i.e., exces­
sive current densities in inductor sections combined with intense heat radiation from the
workpiece surface) can cause localized coil copper overheating. This may promote water
vaporization and the formation of a steam vapor barrier, which essentially functions as
a thermal insulator inside the water-cooling pocket. Thus, the copper cooling might be
severely restricted and the copper might be overheated even when it appears that there
is sufficient water-cooling flow and regardless of the use of high-performance pumps.
To help prevent overheating, water-cooling pockets are placed as close as possible to the
current-carrying face of an inductor.
However, coil overheating can still occur and cause accelerated deterioration of the copper
surface (Figure 4.83a), which speeds up the onset of inductor copper cracking (e.g., attributed
to stress fatigue and stress corrosion) and eventual premature coil failure. As a result, coil life

(a) (b)

FIGURE 4.83
Copper cooling might be severely restricted and copper overheated even when it appears that there is sufficient
water-cooling flow and regardless of the use of high-performance pumps (a). Evidence of arc development in
the single-shot inductor region that is responsible for driving the heat into the fillet region of the shaft (b).
Heat Treatment by Induction 261

when hardening certain shafts (including output shafts, flanged shafts, planet carriers, yoke
shafts, sun shafts, intermediate shafts, drive shafts, etc.) is often shortened to 22,000–24,000
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

heat cycles (industry average). Therefore, the number of segments where localized coil cur­
rent density is increased dramatically should be kept to a minimum.
Unfortunately, as an attempt to take advantage of the electromagnetic proximity effect,
the workpiece-to-inductor distance may need to be reduced drastically. As an example,
Figure 4.83b shows evidence that arcing has taken place in the region of the single-shot
inductor that is responsible for driving the heat into the fillet region of the shaft.
A dramatically reduced current-carrying copper face resulted in the highest current
density that is required for generation of sufficient heat sources to be induced in the fillet
region. In such applications, it is particularly important to maintain a safe copper-to-part
gap in that region, minimizing the negative impact of the real-life disturbances associated
with the effect of worn bearings and the part wobbling during rotation, for example.
These seemingly unavoidable features of the great majority of conventionally designed
single-shot inductors often represent a “weak link” that limits the coil life expectancy, since
certain regions of the coil copper might carry extremely high current densities resulting in
intense heat generation, localized overheating, and arc development.
Regardless of the attempt to position water-cooling pockets as close to the current-­
carrying face of an inductor as possible and the utilization of high-performance pumps,
the coil copper might still be overheated, leading to a premature failure (Figure 4.83).
In summary, the attempt to intensify heat generation in the fillet forces designers to
minimize inductor-to-fillet air gaps. This could cause an arcing problem (Figure 4.83b).
Inductoheat has recently patented an inductor design (Figure 4.84a [321]) that dramati­
cally reduces the localized coil current density in areas prone to overheating and cracking.
The presence of a two-collar section reduces the coil current by one-half, which dramati­
cally reduces the corresponding localized heat generation in the most critical region of the
inductor copper and significantly extends coil life.
In addition, there is reduced sensitivity for a shaft positioned asymmetrically (to some
degree) within the inductor or if substantial wobbling occurs during rotation (because of
worn bearings). Reduced heating effect produced in one of the two half-collar sections
that has an increased inductor-to-shaft gap is offset by an increased induced heating
effect produced in the other half-collar section that has a reduced inductor-to-shaft gap.
Consequently, process sensitivity associated with positioning the shaft within the induc­
tor is reduced over that with a conventionally designed single-shot inductor.

(a) (b)

FIGURE 4.84
Patented inductor design (a) dramatically extends coil life in single-shot hardening of complex shaft-like auto­
motive components and manufacturer’s tool-room tag (b) showing that Inductoheat’s newly designed inductor
(which the customer named “magic coil”) is still considered in good shape after 225,000 heat cycles. (From
V. Rudnev, A. Goodwin, S. Fillip, W. West, J. Schwab, S. St. Pierre, Keys to long-lasting hardening inductors:
Experience, materials, and precision, Adv. Mater. Processes, October, 2015, pp. 48–52.)
262 Handbook of Induction Heating

In one application of the new inductor, one of the world’s largest suppliers of automotive
parts achieved a ninefold increase in single-shot coil life compared with that for conven­
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

tional inductors. This is verified by the manufacturer’s tool-room tag showing that the
inductor (which the customer named “magic coil”) was still considered in good shape after
225,000 heat cycles (Figure 4.84b). Recent study reveals that life of some of these “magic
coils” exceeded 400,000 cycles, which is a 20-fold increase compared to industry average.
Other benefits include measurable improvement in process robustness, coil reliability, and
maintainability.

4.2.2.3.3 Serpentine-Style Single-Shot Inductors


Serpentine-style inductors (Figure 4.85) may be beneficial to use in single-shot harden­
ing of thin-walled hollow stepped workpieces having diameter changes and shoulders of
appreciable size and tight permissible hardness depth deviations. These inductors generate
desirable eddy current flow, increase the “coverage” ratio (“combined copper heating face
area–to–area to be hardened” ratio) and shorten both the axial and radial transition zones.

4.2.2.3.4 Solenoid-Style Coils for Single-Shot Hardening


Single-turn solenoid hardening inductors or multiturn coils can also be used to austenize
selected areas of the component. The coil encircles the area of the component required to
be hardened and heating is applied to austenize it (Figure 4.86b). The induced eddy cur­
rents flow circumferentially.
When hardening small- and medium-sized components requiring a relatively short
heated area, part rotation may be needed in order to improve the temperature uniformity
along the circumference. Rotation helps eliminate (practically speaking) the heat deficit that

FIGURE 4.85
Serpentine-style inductors (Courtesy of Inductoheat Inc., an Inductotherm Group company.)

(a) (b) (c)

FIGURE 4.86
Multiturn solenoid coils can be used to austenize selected areas of the component. (Courtesy of Inductoheat Inc.,
an Inductotherm Group company.)
Heat Treatment by Induction 263

may occur in the workpiece’s area located near the polarized power leads (“field-­fringing”
or “fish-tail” effect) or attributed to nonsymmetrical workpiece-to-coil positioning.
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

The mechanism for moving the part in and part out of the coil can be hydraulic, pneu­
matic, or electric. A simple lift rotate configuration is one of the most popular designs for
small- and medium-sized workpieces providing a low-cost option. A cylinder lifts the
part into the working position and an electric motor begins to rotate the part. An adjust­
able hard stop is used for part positioning. A rotary turntable can be used to increase the
production rate. This type of system can also be used if the component requires hardening
different areas on the same part.
Besides the applied power, the heat intensity as well as the length of the axial transition
zone and runout can be controlled by using various turn spacing, applying profiled mag­
netic flux concentrators (Figure 4.86a), and by fabrication of coils with a different helix of
turn winding (Figure 4.86c).

4.2.2.4 Static Hardening
With static hardening, both the inductor and the workpiece are motionless during the
heating and quenching stages. Static hardening may be done when the size or geometry of
the component does not permit using alternative hardening techniques.
Numerous inductor geometries exist to statically harden a wide variety of components
accommodating their shapes and needed hardness patterns (Figure 4.87). This harden­
ing mode is commonly used when localized hardening is needed on certain areas of
the part, such as the lobes of a camshaft, steering knuckles, or working surfaces of tools,
just to name a few applications. As an example, Figure 4.87 (top middle) shows a “finger-
style” inductor for static hardening internal working surfaces of constant-velocity (CV)
joints. Another typical example of static hardening is a free-style inductor that is shown in
Figure 4.87 (top right). An assembly of smaller inductors (bottom right image) having only 8
mm outside diameter each was designed for simultaneous hardening of selected areas of
complex geometry components providing the needed wear properties.
Figure 4.87 (bottom middle) shows one-half of a patented SHarP-C inductor for station­
ary hardening of multiple lobes of camshafts [22,23,27]. This technology dramatically
reduces shape distortion, eliminating in many cases the need for a subsequent camshaft
straightening operation.

FIGURE 4.87
Numerous inductor geometries exist to statically harden a wide variety of components accommodating their
shapes or needed hardness patterns. (Courtesy of Inductoheat Inc., an Inductotherm Group company.)
264 Handbook of Induction Heating

Similar to a single-shot hardening, with static hardening, the heat is sometimes applied
in several short bursts (pulse heating) with a time delay between them to allow the heat
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

transfer toward the areas where the heat generation might be limited.
In some cases, the part to be hardened is moved into an inductor, heated, and then moved
into another position for the next operation. Figure 4.88 shows a portion of a spindle shaft
hardening walking beam line with quench and temper using solenoid-style inductors.
Spindle shafts are transferred into the working position and then the coils are moved in
and heating begins. The induced eddy currents flow circumferentially, providing uniform
heating. Both ends are austenitized simultaneously. In the majority of cases, the use of
multiturn coils eliminates the necessity to rotate the workpiece.
In static hardening, the quench may be integrated into the inductor, minimizing or elim­
inating (if desired) a quench delay and allowing quenching to be done in the same position
or station without any movement of the part. In other cases, quenching can be done at a
separate location using a specially designed quench chamber.
Figure 4.89 shows the dynamics of radial temperature distributions during surface
hardening of a carbon steel solid cylinder (24 mm diameter) using a frequency of 10 kHz
[17]. Time–temperature profiles of this case study are shown in Figure 2.6. The minimum
required hardness case depth is 3 mm and the maximum total case depth is 6 mm.
After 3 s of heating, the surface layer of the cylinder reaches the needed thermal condi­
tion for austenitization, taking into consideration the nonequilibrium nature of the phase
transformation associated with rapid heating. In this case, the average heat intensity
exceeds 300°C/s. The core temperature (R = 0 mm) is approximately 450°C at the end of the
heating cycle (where R is the radius of the heated solid cylinder and R = 12 mm represents
its surface and R = 0 represents its core).
A comparison of temperature profiles and heat source distributions at different heating
stages is shown in Figure 4.89 [17]. At the initial stage of heating, the entire cylinder is mag­
netic; thus, the δ is quite small and the skin effect is highly pronounced. As can be seen in
Figure 4.89a, after 0.5 s of heating, the surface approaches Ac1, but there is no rise of the core
temperature. The maximum of power density is located at the surface and then it rapidly
decreases (Figure 4.89b). This stage is characterized by highly efficient and intense surface
heating. The temperature profile does not match the power density distribution because the
thermal conduction spreads the heat from the surface toward the internal areas.
After 1.5 s of heating, the temperature of the surface layer (approximately 2 mm thick)
exceeds the Curie point (Figure 4.89c). Since the surface layer became nonmagnetic, there

FIGURE 4.88
Portion of Inductoheat’s spindle shaft hardening walking beam line with quench and temper using solenoid-
style inductors. Both ends of a spindle shaft are austenitized simultaneously.
Heat Treatment by Induction 265

Center Surface Center Surface


Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

1000

Power density (heat sources)


800
Temperature, ºC

600
Time = 0.5 s
400 10 kHz
Rod O.D. = 24 mm
200

0
0 0.004 0.008 0.012 0 0.004 0.008 0.012
Radius, m Radius, m
(a) (b)
Center Surface Center Surface

1000
Power density (heat sources)

800
Temperature, ºC

Time = 1.5 s
600
10 kHz
Rod O.D. = 24 mm
400

200

0
0 0.004 0.008 0.012 0 0.004 0.008 0.012
(c) Radius, m (d) Radius, m

Center Surface Center Surface

1000
Power density (heat sources)

800
Time = 3 s
Temperature, ºC

10 kHz
600 Rod O.D. = 24 mm

400

200

0
0 0.004 0.008 0.012 0 0.004 0.008 0.012
(e) Radius, m (f ) Radius, m

FIGURE 4.89
Dynamics of radial temperature distributions and power density profiles during different stages of surface
hardening of carbon steel solid cylinder (24 mm diameter) using a frequency of 10 kHz.

is a noticeable reduction of the heat intensity. A decrease of μr and an increase of ρ cause


a corresponding increase in δ compared to its values during the initial heating stage.
Though the surface layer is nonmagnetic, its subsurface retains its magnetic properties.
Note that the arrows indicate the radial positioning of the Curie point. At this point, the
power density distribution along the radius has a unique nonexponential (“wave-shaped”
266 Handbook of Induction Heating

profile), which is very different from the commonly assumed exponential distribution.
Figure 4.89d shows that the maximum of power density (heat sources) is located in the
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

subsurface area (approximately 2.2 mm below the surface). This maximum occurs where
there is a “below Curie–to–above Curie” border.
With time, the surface layer heated above the Curie point is expanded, resulting in fur­
ther reduction of the heat intensity. Figure 4.89e shows that at the end of heating (after 3 s
of heat), the entire region located at 3 mm below the surface is austenitized, assuring that,
upon sufficient quenching, the minimum hardness case depth of 3 mm will be achieved.
There are two maximums of an induced power density at this point. Figure 4.89f shows
that the first maximum of power density is located at the surface and then the power den­
sity decreases toward the core. However, once it reaches approximately 2.5–3 mm below
the surface, the power density starts to increase again, and after reaching the second maxi­
mum, at approximately 4.8–5 mm below the surface, it starts its final decline.
M. Lozinskii [7] and P. Simpson [8] independently introduced a hypothesis regarding
this unique wave-shape phenomenon of the power density (heat source) distribution along
the radius/thickness of the workpiece, which differs significantly from the commonly
assumed exponential distribution. Both scientists intuitively believed that there should be
heat stages where the power density distribution would differ from that of the tradition­
ally accepted exponential form. They provided a qualitative description of this phenom­
enon based on their intuition and understanding of the physics of induction hardening.
However, a quantitative assessment of this phenomenon could not be developed because
of the limitation in computer simulation capabilities and the lack of software that could
model the tightly coupled electromagnetic-heat transfer phenomena of induction heating.
Furthermore, it was not possible to measure power/current density distribution inside the
workpiece during heating using a laboratory setup without disturbing the eddy current
flow. The first quantitative assessment of this phenomenon with a detailed explanation of
the nonexponential heat source distribution was published in the mid-1990s [52].
A nonexponential (wave-shape) heat source distribution has a noticeable impact on the
selection of process parameters, heating protocol/recipe, final temperature distribution,
and hardness pattern in surface hardening applications. This is so, because if the fre­
quency has been chosen correctly for surface hardening, the thickness of the austenitized
layer (the nonmagnetic layer) is less than δ in austenitized steel (Figure 4.89d) and the
wave-shape heat source distribution takes place during the majority of the heating cycle
for induction surface hardening.
By comparison, in through hardening applications or when heating magnetic steels
before hot forming, the impact of this phenomenon is less significant, because the duration
of the hot stage (when the entire cross section is heated above the Curie point) occupies a
much greater portion of the heat cycle diminishing the impact of the nonexponential heat
source distribution.
Spray quenching begins after the completion of the heating stage (Figure 2.6). A dra­
matic decrease of surface and subsurface temperatures (R = 12 mm through R = 9 mm)
occurs practically instantly during spray quenching. At the same time, there is a noticeable
time delay in the cooling of internal regions (e.g., R = 6 mm) and in particular the core (R =
0 mm). Because of thermal conduction, the core temperature continues to rise during 2 s
of quenching. The temperature of internal regions located at a distance greater than 6 mm
below the surface does not exceed the Ac1 critical temperature; thus, austenitization does
not occur, meaning that the total HAZ will be slightly less than the maximum permissible
6 mm below the surface.
Heat Treatment by Induction 267

4.2.2.5 Pulse Heating Mode


Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

On several occasions, the phrase a pulse heating has been used earlier and will be used fur­
ther in this book to refer to the method of using short bursts of heat to control or maintain
desired thermal conditions. For example, it is often beneficial to use this heating mode
when induction hardening gears and gear-like components, parts made of high-carbon
steels and cast irons as well as other workpieces that comprise certain geometrical irregu­
larities or materials that exhibit low toughness and a pronounced tendency for cracking.
In some instances, these bursts of heat allow prevention of the appearance of excessive
thermal gradients, permitting the surface temperature to be maintained without overheat­
ing. The pulsing mode can also be used in combination with dual-frequency heating. In
some gear hardening applications, preheating is done utilizing a low frequency and then
the heat is pulsed for the final heat at a high frequency (see Section 4.9.1).
A typical pulse hardening cycle would include a series of heat-on and heat-off or low-
heat cycles until the desired thermal/metallurgical condition is obtained. The duration
of each pulse for hardening typically ranges between 0.5 and 8 s, depending on the size,
geometrical complexity of the component, material, and available power. Durations can
be noticeably longer, lasting dozens of seconds when heat treating large-sized workpieces.
For cast irons and cast steels exhibiting low toughness and a tendency to cracking, pulse
heating can help avoid critical thermal gradients by slowing the heat intensity at the initial
heating stage and allowing gentle conduction of the heat toward the core without over­
heating the surface of the workpiece.
Pulse heating is somewhat similar to continuous/progressive heating where the work­
piece experiences an alternation of heat-on and heat-off stages (pulsing) while processing
through a multicoil induction heating line.

4.2.3 Specialty Inductors
As has been stated earlier, an inductor or induction coil is described as an electrical current­-
carrying device located in proximity to an electrically conductive workpiece to be heated.
Thus, the location of a workpiece inside a solenoid-style induction coil (which is typi­
cally used when introducing induction heating principles) is only one of many possible
arrangements.
It might be more appropriate to introduce a heating inductor as any electromagnetic
device that provides 3-D heat source distribution that is sufficiently strong to produce a
measurable heating effect. In hardening applications, upon adequate austenitization and
quenching, the desired hardness pattern can be obtained. As stated earlier, in everyday
practice, an inductor is also called an induction coil or simply a coil. There is a nearly
endless variety of inductor/coil styles to accommodate the almost endless variety of IHT
applications. However, their geometry often does not resemble the shape of a classical coil.
Figure 3.2 shows a small array of the various designs of heating inductors, many of which
are used in induction hardening.
Over the years, induction practitioners have established what is called a family of special
or specialty inductors. Common names have been coined to describe their appearance or
function. Some of these include pancake, clamshell, hairpin, split-return, channel, and
many others. A number of inductor configurations, styles, and geometries have already
been discussed in previous sections. It is convenient to classify them into the following
main groups [17]:
268 Handbook of Induction Heating

• Solenoid-type single-turn and multiturn coils


• Serpentine and cage-style inductors
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

• Gap-by-gap gear hardening inductors


• Single-shot inductors, channel inductors, and dual-callor inductors
• Hairpin and double hairpin inductors
• Butterfly inductors and split-return inductors
• I.D. coils for hardening inside (internal) surfaces
• Pancake inductors (round, square, rectangular, and odd shape)
• Active–passive inductors
• “Z”-shaped, “D”-shaped, “C”-shaped, and “U”-shaped inductors
• Split or clamshell inductors, inductors with “doors,” and door-less inductors
• Free-style and profiled inductors, just to name a few.

An inductor may also be composed of the several design features of any of the various
styles. A particular choice of inductor design is application specific, and among other fac­
tors, it depends on the following:

• Geometry of the workpiece, the presence of geometrical irregularities and discon­


tinuities (e.g., holes, keyways, undercuts, snap rings, diameter changes, grooves,
shoulders, sharp corners, etc.), and the specifics of the required hardness pattern
• Heating mode (i.e., static heating vs. single-shot vs. scanning vs. progressive
heating)
• Specifics of process protocol, application subtleties, heating pattern, production
rate, and available space
• Material handling specifics and the part’s processing particularities (i.e., will the
part be moved into the coil, will the coil index into the part, is rotation of the part
required, or how is the part transferred after heat treatment?)
• Presence of electrically conductive bodies (e.g., centers, tooling, chucks, etc.)
located in proximity to the heating inductor
• Specifics of the material’s prior microstructure (e.g., spheroidized vs. Q&T), and so on

Single-inductor or multicoil arrangements can be used. The range of frequencies may be


from as low as line frequency to as high as 800 kHz and above. The process recipe/proto­
col may call for using a single frequency or multiple frequencies. This includes the use of
multiple frequencies sequentially or simultaneously or a combination of both.
Because of space limitations, only some of the specialty inductor types will be reviewed
here.

4.2.3.1 Pancake, Split-Return, Butterfly, and Hairpin Inductors


Pancake-style inductors are used for selective heating of flats, disks, plates, curved surfaces (e.g.,
mill rolls in the papermaking industry), and other applications. The term pancake is used to
describe the typical flat round/oval/square/rectangular shape of the inductor. This type of
inductor resembles the appearance of an electrical stovetop burner. Figure 4.90 shows a few
Heat Treatment by Induction 269
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

FIGURE 4.90
Examples of pancake inductors. The size of pancake inductors can vary from less than 20 mm diameter (bottom
left) to more than 1 m diameter (top). (Courtesy of Inductoheat Inc., an Inductotherm Group company.)

examples of typical pancake inductors. The size of a pancake inductor can vary from less than
20 mm diameter (Figure 4.90, bottom left) to more than 1 m diameter (Figure 4.90, top).
It should be noted that all pancake-style inductors exhibit eddy current cancellation in
the center (the “dead” spot), which causes a corresponding heat deficit. The workpiece
region positioned under the center of a pancake inductor is only heated via thermal con­
duction. Thus, if the heat time is short, then there will be a noticeable heat deficit in the
region positioned under the center of a pancake inductor. The temperature uniformity can
be improved by applying modified pancake designs with variable workpiece-to-turn cou­
pling having tighter coupling in the center compared to the coupling of the middle turns.
It should be noted that although a modified pancake design helps reduce the heat nonuni­
formity, it will not completely eliminate the “dead” spot in the area of the workpiece under
the center of the coil in static heating applications.
Split-return inductors also belong to the larger family of specialty inductors and are used
in a variety of applications including selective and through hardening, annealing, stress
relieving, tempering, brazing, and so on. Figure 4.91 shows several examples of split-
return inductors.
A split-return inductor has a main (center) leg that splits into two return legs, producing
a unique eddy current distribution within the workpiece. The heat generation within the
workpiece by the return legs is often undesirable and, in the majority of applications, is
commonly considered to be a waste of energy. Therefore, steps should be taken to mini­
mize these losses. In contrast, the kilowatt losses induced in the workpiece by the main leg
provide desirable heating; thus, steps should be taken to maximize these values to obtain
the maximum electrical efficiency.
Figure 4.92a shows the electrical circuit of a split-return inductor. The magnitude of cur­
rent flowing in the main leg of the inductor is double that in the return legs. Also, the
power density (heat source) induced under the main leg is four times greater than that
generated under the return legs. This ratio can be even greater if the current-carrying
face of the return legs is wider than the width of the heating face for the main leg, or if a
magnetic flux concentrator is applied to the main leg, which results in much higher power
density in a narrow band of the workpiece located under the main leg. Figure 4.92b shows
270 Handbook of Induction Heating
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

FIGURE 4.91
Split-return inductors. (Courtesy of Inductoheat Inc., an Inductotherm Group company.)

Magnetic flux Induction


concentrator coil

Tube Weld
wall
(a)
(b)

FIGURE 4.92
Electrical circuit of a split-return inductor (a) and a sketch of its transversal cross section with a U-shaped
magnetic flux concentrator (b) positioned around a main leg. (From V. Rudnev, Systematic analysis of induction
coil failures, Part 14: Split-return inductors and butterfly inductors, Professor Induction Series, Heat Treating
Progress, ASM Int’l, Materials Park, OH, March/April, 2009, pp. 17–19.)

a sketch of the transverse cross section of a split-return inductor with a U-shaped magnetic
flux concentrator positioned around the main leg.
Figure 4.93 shows the magnetic field distribution with (right) and without (left) a
U-shaped magnetic flux concentrator located around the center leg of a split-return induc­
tor [325] showing the imaginary equipotential field lines and the coil current densities.
Without a concentrator, the magnetic flux spreads around the surroundings of the induc­
tor legs (Figure 4.93a). Because current flow in the return legs is in the opposite direction
of that in the main leg, the electromagnetic proximity effect shifts all currents toward
each other. This degrades the “inductor-to-workpiece” electromagnetic coupling, which
dramatically reduces overall electrical efficiency (requiring much higher coil current
and power to provide the needed heating) and increases the coil copper kilowatt losses.
Reducing the distance between the main and return legs makes the situation even worse. The
coil electrical efficiency in such systems can drop to 20% to 30% and even lower, particularly
when heating nonmagnetic metals or heating magnetic metals above the Curie temperature.
The U-shaped magnetic concentrator positioned around the main leg forms a magnetic
path to channel the main magnetic flux in a well-defined area and separates the mag­
netic fields produced by the main and return legs (Figure 4.93b). The inductor currents
are shifted toward the workpiece surfaces, improving the heating efficiency, reducing the
required coil current, and improving the overall conditions to increase the inductor life.
This is the reason why it is strongly recommended to use flux concentrators in the majority
of split-return inductors.
Heat Treatment by Induction 271

Without flux concentrator With flux concentrator


Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

Magnetic field Magnetic field

Main leg current Main leg current

Return leg current Return leg current


(a) (b)

FIGURE 4.93
Computer-modeled magnetic field distribution (showing imaginary equipotential field lines) and coil cur­
rent densities with (b) and without (a) a U-shaped flux concentrator located around the central leg of a
split-return inductor. (From V. Rudnev, Systematic analysis of induction coil failures, Part 14: Split-return
inductors and butterfly inductors, Professor Induction Series, Heat Treating Progress, ASM Int’l, Materials
Park, OH, March/April, 2009, pp. 17–19.)

Because the edges of copper corners that face the workpiece experience the highest cur­
rent density, particularly when U-shaped concentrators are used (Figure 4.93b), the pres­
ence of sharp edges should be avoided. Instead, it is beneficial if the inductor legs have
small radii.
Butterfly inductors. Figure 4.94a shows what is commonly referred to as a butterfly induc-
tor, which is essentially a multiturn split-return inductor resembling the shape of a butterfly.
A butterfly inductor uses two pancake inductors (the “wings”), and the turns are wound so
that the electrical currents flowing in the center turns of the coil are all in the same direction
(the “body”), creating the highest power density region that provides the main heating effect.

(a) (b)

FIGURE 4.94
Examples of a butterfly inductor (a) and a hairpin inductor (b). (Courtesy of Inductoheat Inc., an Inductotherm
Group company.)
272 Handbook of Induction Heating

Similar to split-return inductors, a U-shaped flux concentrator is usually positioned


around the center turns of a butterfly inductor to focus the heat generation and increase
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

the coil efficiency in the center of the coil.


Hairpin inductors. Inductors that are formed of bent or brazed copper tubing or CNC
machined to conform to a workpiece’s geometry are sometimes referred to as hairpin induc­
tors (Figure 4.94b). This name is derived from the resemblance of the inductor’s loop shape
to a lady’s hairpin. Medium to high frequencies are commonly used with these inductors.
Hairpin inductors can be fabricated as flat inductors (Figure 4.94b) or may be bent form­
ing an “L”-shape or a “U”-shape depending on workpiece geometry and required hard­
ness patterns. The heating capability of this inductor heavily relies on the electromagnetic
proximity effect, making it particularly sensitive to “workpiece-to-inductor” gap varia­
tions. The use of a magnetic flux concentrator between forward and return legs is often
required unless the heated workpiece is positioned between inductor legs.

4.2.3.2 Specifics of Designing Inductors for Heating Interior Surfaces


When heating the interior of a workpiece (e.g., inside diameter) is required, internal or
I.D. inductors are used [1,56,328]. Induction heating of the internal surfaces of a work­
piece can be used for applications such as hardening, tempering, annealing, shrink fitting,
stress relieving, and brazing. Figure 4.95 shows a variety of different I.D. inductor styles.
Solenoid-type single-turn and multiturn coils, hairpin and double hairpin inductors, and
C-core inductors are some of the inductors that can be used for heating internal surfaces.
Single-turn and multiturn solenoids are the most popular inductors for heating surfaces of
the hollow parts with sufficient inside diameters.
The inductors are often made from copper tubing (round, square, or rectangular) that
is spiral wrapped or brazed the same way a solenoid is wrapped. In other cases, the head
of the internal inductor is CNC machined from a solid copper bar. This not only provides

FIGURE 4.95
Variety of different inductor styles for heat treating internal surfaces [328]. (Courtesy of Inductoheat Inc., an
Inductotherm Group company.)
Heat Treatment by Induction 273

a rigid, robust coil but also allows profiling of the heating face matching the specific part
geometry and minimizing the end effect and coil helix effect.
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

The effectiveness of an internal cylindrical coil depends to a much greater extent on the coil-
to-workpiece gap, compared with similar coils used for heating external surfaces or outside
diameters. The electrical efficiency of an internal coil rapidly decreases with an increase in the
coupling gap, particularly when heating nonmagnetic metallic materials. To keep the coil-to-
workpiece gap as small as possible, the return leg is usually located inside the internal coil.
The reason why solenoid-type inductors for heating I.D.’s are not as efficient as similar induc­
tors used for heating O.D.’s is related to the electromagnetic “ring” effect (see Section 3.1.5).
According to this effect, the coil current is concentrated on the I.D. of the coil that represents a low
impedance path. When heating inside diameters, this is the area farthest from the heated part.
As a result, the electromagnetic coupling between the coil and the heated workpiece is greater
than the actual air gap between the I.D. of the workpiece and the coil’s O.D. This makes for poor
coil-to-workpiece coupling and, therefore, causes a noticeable reduction of coil efficiency.
To improve the coil-to-workpiece electromagnetic coupling, use small-diameter, thin-
wall copper turns, or flattened or rectangular tubing.
Installation of a magnetic flux concentrator inside of the internal inductor is frequently
mandatory to increase the coil efficiency and reduce the coil current, particularly for
heating internal surfaces of small to moderate diameters. The flux concentrator creates
an electromagnetic slot effect that has a substantially stronger impact on the coil current
distribution than the electromagnetic “ring” effect and forces the coil current to be shifted
toward the coil outside area to be positioned closer to the surface of the heated workpiece.
This increases the magnetic field strength and heat intensity at the workpiece internal
surfaces required to be heated (Figure 4.96).
The use of magnetic flux concentrators on internal coils provides a noticeable reduction
in coil current and the required power, reduces coil water-cooling requirements, and often
simplifies load matching of the induction coil and inverter. At the same time, enlarging the
inside diameter of the heated workpiece makes these improvements less pronounced. As
expected, an increase of I.D. weakens the ring effect, which may eventually eliminate the
necessity to use concentrators when heating large-sized interior surfaces.
Sometimes, a magnetic flux concentrator may itself become the weakest link when
applied to an I.D. inductor. This is attributed to the danger of a possible magnetic satura­
tion of the flux concentrator and subsequent overheating, which could result in premature
degradation. Computer modeling can help in selecting the proper parameters, ensuring a
long-lasting design, and preventing premature failure of the flux concentrator.
Uniform winding of turns often results in greater heat generation in the middle of the
workpiece compared to the coil end areas, resulting in a distinctive “thumbnail” heat pattern.

(a) (b)

FIGURE 4.96
Results of FEA computer-modeled magnetic field of four-turn I.D. coils without (a) and with (b) a U-shaped
magnetic flux concentrator.
274 Handbook of Induction Heating

Heat source and temperature deficiencies in the end areas can be compensated for by copper
profiling and/or by spreading out the middle turns of the coil in comparison with the tighter
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

end turns, permitting an electromagnetic proximity effect to even out the pattern.
Since the center return leg of a multiturn I.D. inductor usually goes through the center
of the coil, the smallest outside diameter of an I.D. coil is typically limited to approxi­
mately 12 mm but 20 mm is more typical. Single-turn I.D. inductors can be smaller (e.g.,
Figure 2.3b). If the inside diameter of the heated part is too small for using solenoid coils,
then alternative inductor styles may be considered.
Figure 4.97 illustrates the fundamental difference in heat propagation in the workpiece when
heating O.D.’s versus I.D.’s. In O.D. heating, the heat transfers because of thermal conduction
from a larger perimeter toward a smaller perimeter. The reverse is true when heating internal
surfaces: heat transfers from a smaller perimeter area toward a larger perimeter, resulting in a
much greater “cold sink” effect. In the great majority of case hardening applications, to provide
the same case depth, higher frequencies are used for induction hardening internal surfaces
than for outside surfaces. In addition, higher power densities are usually required to compen­
sate for an inevitable loss in coil electrical efficiency and the greater cold sink effect.
The heat intensity is directly related to the frequency, which makes it easy to generate a
denser magnetic field in the workpiece surface by using a higher frequency. These process
features contribute to an I.D. surface temperature that typically is higher than that of an
O.D. surface, assuming that the heating time and the case depth are the same for both.
Depending on the application, attempting to keep the coil-to-workpiece coupling gap
as small as possible results in having air gaps in the range of 0.125 to 0.0625 in. (3.2 to
1.6 mm), and in some cases, air gaps are even as small as 1.25 mm. It is imperative to keep
in mind the expansion of the material during heating, which could measurably reduce the
“inductor-to-workpiece” gap existing at room temperature. The desire to have very small
clearances is accompanied by the danger of arc development and improper part/coil han­
dling when the inductor can accidentally touch the part. The use of an appropriate fixture
that provides a robust, reliable loading/unloading operation is essential for I.D. coils. The
use of a thin layer of ceramic coating or an appropriate electrical insulator can help prevent
arc development between the coil and the workpiece [1,56,328].
Coil cooling is another important design factor since internal inductors are commonly
less efficient compared to O.D. inductors. In many cases, high-pressure pumps are needed
with I.D. coils to provide adequate cooling water flow. In the MIQ design, coil cooling and
quench holes are integrated in the head of the internal inductor. Specifics of coil cooling
are discussed in Section 4.2.4.4.

Heating O.D. surface Heating I.D. surface

FIGURE 4.97
Difference in heat propagation in the workpiece when heating O.D.’s versus I.D.’s.
Heat Treatment by Induction 275

4.2.3.3 Proximity Induction Heating of Flat and Plane Surfaces


Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

IHT of flat and plate workpieces can be done using rectangular solenoid coils, transverse
flux coils, traveling wave inductors, or specially designed coils that apply the principle of
proximity heating (e.g., pancake or hairpin inductors). The first three techniques are dis­
cussed in the appropriate sections of Chapter 6; the use of pancake and hairpin inductors
has already been reviewed in Section 4.2.3.1.
Rectangular solenoid coils are commonly used for heating strips, slabs, and plates. These
inductors are very similar to classical solenoid coils for heating cylinders, with the excep­
tion that the coil is formed as a rectangular solenoid. With appropriate design parameters,
rectangular inductors provide through heating of the entire workpiece or selective heating
featuring high repeatability, electrical efficiency, and relatively low positioning sensitivity.
Channel coils, butterfly, split-return, or C-core inductors can be used in cases where
heating of selected areas of flat parts is required.

4.2.3.4 Inductors with Inserts


As mentioned earlier, heat treatment by induction is the preferable choice for a high-­
volume production environment. The inductor design can be optimized for a certain part
or family of similar parts. However, in some small shops, it is often required to heat many
small batches of parts, which can have different shapes and sizes.
It would not be cost-effective to have several dozen inductors to accommodate every
single part. In addition, it is often time-consuming (even with a quick-change coil design)
to install a new coil, load match it with the inverter, then run, for example, 30 or 40 parts,
and then replace it with a new coil to run another 30 or 40 parts. Significant downtime,
essential capital cost to have all those coils and their spares, and the necessity to store and
maintain them would be noticeable shortcomings of conventional induction heat treat­
ment compared to furnace batch heat treatment.
In cases like these, it is often beneficial to use inductors with changeable inserts.
Figure 4.98 shows a variety of inserts that allow one to heat treat different parts using the
same inductor and power supply.
The physics of induction heating using inserts is quite simple and is illustrated in
Figure 4.99. AC flowing within the induction coil produces a time-varying magnetic field
that, in turn, induces eddy current flow within the insert in the opposite direction. Because
of the electromagnetic proximity effect, eddy currents have a tendency to follow the path
of the source (coil) current. As discussed in Chapter 3, eddy currents have to make a loop

FIGURE 4.98
Variety of inserts allowing heat treating different parts using the same inductor and power supply.
276 Handbook of Induction Heating

Single-turn
inductor
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

Insert

Insert
Coil current
current

Coil
Workpiece current

To power source

FIGURE 4.99
Current flow within the coil insert.

(close electrical circuit). However, the slot within the insert breaks the natural eddy current
flow, forcing them to complete a loop on the internal surface of the insert.
Current flow on the inside surface of the insert creates its own magnetic field that, in
turn, induces currents within the workpiece, generating the heat as in a conventional
induction heating system.
Inserts are typically made from a copper or its alloys. It is important that the radial
thickness of the inserts should be at least six times that of the current penetration depth.
Otherwise, there will be eddy current cancellation inside the insert, resulting in low total
efficiency. Water cooling of inserts is usually required, unless the heat time is very short,
the duty cycle is low, and they can be sufficiently cooled by quenching.
Inserts provide a cost-effective solution when it is necessary to heat a variety of small
batches of workpieces having relatively small cross sections. It is easy to maintain and
store them. One noticeable advantage of using inserts deals with the fact that inserts sim­
plify load tuning of the inductor and power supply, significantly reducing the required
time to switch from the heating of one batch of parts to another.
In order to reduce inductance, the slot opening should be as small as possible and should
be chosen based on the desire to eliminate any arcing problems. Often an electrical insula­
tor is placed inside the slot of the insert.
Unfortunately, inductor inserts are not free from drawbacks. One of the main shortcom­
ings of inductors with inserts is their low electrical efficiency. The intensity of heating
and efficiency of the inductor can be improved by fabricating the face of the inserts’ O.D.,
which is larger than its I.D. face. Besides, the field fringing effect at the slot(s) area should
be kept in mind while using inserts.
In other cases, inserts are electrically attached to the coil I.D. Because of the electromag­
netic ring effect, coil current flow occurs on the inside area of the insert, improving elec­
tromagnetic coupling and accommodating a geometry of the workpiece. It is important to
make sure that there is a reliable and consistent inductor-to-insert electrical contact when
applying different inserts.

4.2.3.5 Clamshell or Split Inductors


Among other considerations, the method used to transfer parts before and after IHT
may also be a decisive factor in selecting a particular inductor style. For example, when
Heat Treatment by Induction 277

hardening irregularly shaped components (Figure 4.100), the adjoining areas can some­
times exclude the possibility of positioning the workpiece inside a cylindrical coil.
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

In other cases, the required “coil-to-part” air gap that would provide sufficient clear­
ance for loading and unloading the workpiece is so large that it dramatically reduces coil
electrical efficiency, transforming the inductor into a water heater instead of a device for
heating metals.
In addition, the necessity of having such a large coil-to-workpiece air gap may make
it impossible to obtain the required hardness patterns because of an unfavorable com­
bination of various electromagnetic phenomena and part geometry. An example is the
hardening of camshaft lobes that have a sharp “nose” and undersized “base circle” in
combination with large bearings or eccentric journals. In cases such as this, split or clam­
shell inductors (Figure 4.101a) may be used.
Clamshell inductors are so named because they are hinged on one side so that the work­
piece can be loaded in the correct heating position. For hardening applications, quenching
can be integrated into the coil or a quench device can be placed adjacent to the inductor.
Depending on the application specifics, the coil copper can be profiled to replicate the part’s
shape, which provides a minimum and consistent coil-to-part coupling (air gap). The result is a
highly efficient and uniform heating around the perimeter of an irregularly shaped component.
Locating pins are often required on the inside of a clamshell inductor or on certain
areas of the part, maintaining its location throughout the heating and quenching cycle.
If locating pins are used inside of medium- or high-frequency coils, then ceramic pins

FIGURE 4.100
When hardening irregularly shaped components, adjoining areas can sometimes exclude the possibility of
positioning the component inside a cylindrical coil.

Clam shell
Coil
inductor
current

Contact area
(a) (b)

FIGURE 4.101
Split or clamshell inductor (a) and coil electrical current flow (b) through the contact area. (From V. Rudnev,
Systematic analysis of induction coil failures, Part 9: Clamshell inductors, Professor Induction Series, Heat
Treating Progress, ASM Int’l, Materials Park, OH, Jan./Feb., 2007, pp. 17–18.)
278 Handbook of Induction Heating

are typically chosen. However, ceramics are brittle and can fail prematurely because of
improper part handling, mechanical damage, or thermal shock.
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

Short coil life and low production rates are some of the main disadvantages of clamshell
inductors. Short coil life results from the inherent need to break the electrical current path
by having high-current contacts.
When the clamshell inductor is closed, it must be clamped with sufficient pressure to
ensure that good electrical contact is made. In real life, no coil contact surfaces are per­
fectly smooth. Thus, surface roughness has the largest impact on coil electrical current
flow through the contact area (Figure 4.101b) [337]. Electrical resistance of the contact sur­
faces is usually more than 10-fold that of solid copper.
Regardless of the amount of polishing, air pockets are always present on contact sur­
faces, squeezing the coil current to flow through the randomly located solid-to-solid con­
tact points. What results is the appearance of a localized increase of current density and an
increase in electrical resistance of the contact area compared to the solid copper. Because
the same current flows through both the coil copper and clamping (hinged) contact area,
the latter region will experience much greater heat generation because of I2 R (Joule) losses.
The clamping area of the coil also contributes to a short inductor life because of the
inevitable wear after multiple clamping and the presence of contaminants, which can lead
to excessive overheating and arcing, and, ultimately, to premature coil failure. The quality
of the electrical contact and its cleanliness degrade appreciably after multiple openings
and closings. Contaminants that are always present in a real-life production environment
quickly build up on contact surfaces, increasing the electrical resistance of the contact area.
These factors cause the electrical resistance of transitional areas between contact sur­
faces to continuously change during coil operation, resulting in variation in the power
induced within the heated part and, consequently, variation in the heat pattern. With time,
heat treaters are often required to increase the contact pressure to compensate for a clam­
shell coil’s time-dependent power loss.
Silver alloy plating is commonly used in an attempt to improve the coil life and reduce the
electrical resistance of the coil clamping areas. Unfortunately, it does not appreciably raise the
inductor life to the level of conventional solenoid coils. The life of a clamshell inductor typi­
cally does not exceed 10,000 heats, and lasting for only 3000 or 4000 heats is not uncommon.
In a low-production environment, clamshell inductors may still be a valid and cost-­
effective choice; however, newly developed, innovative contactless inductor designs
including patented SHarP-C Technology are frequently better alternatives to clamshell
inductors, substantially boosting coil life expectancy and pattern repeatability particu­
larly in a high-production environment.

4.2.3.6 Profiled Inductors
Different inductor styles (e.g., Figures 4.57, 4.76, and 4.87) may require profiling to con­
form to the geometry of workpieces and to obtain the desirable hardness pattern. As an
example, Figure 4.102a shows that the electromagnetic end effect of a conventional single-
turn coil results in a thumbnail-shaped heat treat pattern. In contrast, a properly profiled
inductor (Figure 4.102b) improves the electromagnetic proximity effect and increases heat
source generation at the ends of the pattern, making it straighter. In other words, the lack
of heat generation caused by the coil end effect is compensated for by the proximity effect.
The specifics of single-turn inductor profiling are related to the applied frequency, mag­
netic field intensity, and the desired hardness pattern. When using higher frequencies (e.g.,
450 kHz), the depth of inductor profiling contour may be as small as 0.015 in. (0.4 mm),
Heat Treatment by Induction 279

Conventional coil Profiled coil


Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

Cylindrical
workpiece

(a) Heating profile (b)

FIGURE 4.102
Thumbnail heat treat pattern of a conventional single-turn coil (a). Properly profiled inductor improves electro­
magnetic proximity effect and results in flat hardness pattern (b).

whereas at lower frequencies (e.g., 3 kHz), a similar effect may be observed upon having a
depth of cut of approximately 0.125 in. (3 mm) or so.
In multiturn coils, profiling of magnetic field distribution can be achieved by changing
the turn spacing (pitch). Uniformly wound multiturn inductors have the strongest field in
the middle of the coil, producing the most intense heat there. To obtain a uniform heating
profile for static or single-shot heating, the middle turns can be spread apart or a barrel-
shaped coil can be used with the middle turns having a larger coupling gap compared to
the end turns. Figure 4.103 shows some examples of profiling of multiturn coils. Different
turn arrangements might be needed if high frequency is used in combination with larger
coil overhangs (see Section 3.1.7.1).
As mentioned earlier, because of the electromagnetic proximity effect, coil electrical effi­
ciency is maximized by keeping the workpiece-to-coil gap as small as possible. However,
if the workpiece has an irregular shape (e.g., parts with flanges, undercuts, shoulders, etc.),
there are some special considerations. Often the inductor may have to be relieved around
the edges and corners. Besides that and in order to obtain uniform heating when large
changes in masses of metal are present, the coupling gap should be changed to compen­
sate for heat sinking and flux robbing or field fringing effects. Obviously, the areas of less
mass will require less energy to heat. This is compounded by the fact that the electrical
current tends to take the shortest path (the path of smallest resistance/impedance), provid­
ing benefits in some cases and exhibiting drawbacks in others. Mathematical modeling is
an essential tool to determine the exact amount of inductor profiling needed.

Conventional coil Profiled coil


(even spacing) (non-uniform spacing)

Cylindrical
workpiece

Heating profile

FIGURE 4.103
Examples of profiling in multiturn coils.
280 Handbook of Induction Heating

4.2.4 Fabrication and Cooling of Hardening Inductors


Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

In hardening applications, inductors carry significant electrical power and often operate
in hostile working environments. This includes intense cyclic heating and cooling, being
exposed to high temperatures, and the presence of smoke, steam, quench fluid, and air­
borne particles while subjected to mechanical movement and potential sudden part con­
tact. These are only some of the causes of premature failures and the reasons why inductors
are often considered to be the weakest link in an induction hardening system [329–332,336].
Advanced inductor designs, selection of appropriate materials, precise fabrication, and suf­
ficient water cooling are paramount to ensure long life while producing high-quality parts.
Since there are innumerable inductor designs, there are a number of alloys and several
fabrication techniques for their manufacturing [330].

4.2.4.1 Material Selection


Copper and copper alloys are almost exclusively used in the fabrication of inductors because
of their reasonable cost, availability, and a unique combination of electrical, thermal, and
mechanical properties. Proper selection of the copper grade is crucial to minimize the dele­
terious effects of factors that contribute to premature coil failure including stress-corrosion
and stress-fatigue cracking, galvanic corrosion, copper erosion, pitting, overheating, and
work hardening. Cooling water pH also affects the copper’s susceptibility to cracking.
Oxygen-free high-conductivity (OFHC) copper should be specified for most hardening
inductors despite its higher cost. Besides its superior electrical and thermal properties,
OFHC copper dramatically reduces the risk of hydrogen embrittlement. The higher ductil­
ity of OFHC copper is also important, because the coil turns/legs are always subjected to
flexing and experience electromagnetic forces. The higher cost of OFHC copper is usually
offset by the improved inductor life.
The electrical conductivity of copper is an important physical property that also makes
a marked impact on the working conditions and life expectancy of the induction coil. As
discussed in Section 3.1.1, the electrical conductivity of a material σ is a measure of how
easily it conducts electric current. The reciprocal of conductivity is the electrical resistivity
ρ. Electrical resistivity varies with temperature, chemical composition, and microstructure
and depends strongly on purity. Phosphorus, tin, selenium, tellurium, and arsenic are
some of the typical impurities found in commercially pure copper. Impurities distort the
copper lattice, affecting ρ Cu to a considerable extent. For example, Figure 4.104 shows an
increase in ρ Cu of copper with admixtures of various seemingly small amounts of impuri­
ties [332–336] leading to a corresponding increase in the coil electrical resistance and Joule
losses (kilowatt losses) dissipated within the inductor.
All these factors lead to a reduction in the coil electrical efficiency, requiring higher cur­
rent and power in order to provide the same heating. Higher coil current creates favorable
conditions for the appearance of localized hot spots and increases the magnetic forces and
the growth rate of cracks in the coil copper.
Although alloying copper can noticeably improve certain industrial characteristics such
as corrosion resistance, for example, it also reduces its thermal conductivity. This decreases
the capability of water-cooling passages to reduce the localized overheating caused by the
lower rate of heat removal as a result of diminished thermal conduction [331].
Stress-corrosion cracking, stress-fatigue cracking, pitting, and some other undesirable
phenomena that have a marked negative effect on coil life are affected by residual elements
and alloying elements. For example, Figure 4.105 shows the effect of low concentrations of
Heat Treatment by Induction 281

2.4
P
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

2.3
Ti Fe
Si

Electrical resistivity, µΩ*cm


2.2

2.1
As
2

1.9
Al
1.8
Ag
1.7

1.6
0 0.02 0.04 0.06 0.08 0.1
Admixture, at.%

FIGURE 4.104
Increase in ρ Cu of copper with admixtures of various seemingly relatively small amounts of impurities. (From
G.  Joseph, K.J.A. Kundig (editors), Copper: Its Trade, Manufacture, Use, and Environmental Status, ASM Int’l,
Materials Park, OH, 1999, 452 pp.; F. Pawlek, K. Reichel. The effect of impurities on the electrical conductivity of
copper, Zeitschrift für Metallkunde, Vol. 47, 1956, p. 347.)

100,000
Arsenic
Antimony

10,000
Time to fracture, min

Silicon

1000

C12200

Phosphorus

100
0.001 0.01 0.1 1 10
Concentration, %

FIGURE 4.105
Effect of low concentrations of some elements on time-to-fracture of copper by stress-corrosion cracking under
an applied tensile stress of 70 MPa (10 ksi). (From ASM Handbook, Vol. 11: Failure Analysis and Prevention, ASM
Int’l, Materials Park, OH, 1986.)
282 Handbook of Induction Heating

some elements on time-to-fracture of copper by stress-corrosion cracking under an applied


tensile stress of 70 MPa (10 ksi) [259]. This example emphasizes the reasoning why the
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

selection of the copper grade and its purity are of great importance, making it preferable
to use OFHC copper in high–current density and high-frequency applications. In low– and
moderate–current density applications, the use of alternative copper grades (e.g., C11000,
12000) may be suitable and cost-effective.
When doing a repair work make sure that a copper grade specified by Machine Builder
OEM is used for coil fabrication.

4.2.4.2 Fabrication Techniques
Inductors for hardening applications are typically CNC machined from a solid copper bar,
making them rigid, durable, and repeatable. In other cases, copper tubing (square, rectan­
gular, round, or die-formed shaped tubes) may be used for coil fabrication.
Copper tubing is typically annealed to improve its ductility, bending properties, and
workability [330,336]. When sharp bends or complex coil shapes are required, inductor
segments made from tubing are assembled by brazing. Joints are often overlapped, creat­
ing tongue-and-groove joints. Butt-joints should not be used.
Different alloys (fillers) can be used for the brazing of copper. Good wetting and free-
flow characteristics are important. A silver-base brazing alloy that contains 35% to 45% Ag
is commonly used for brazing coil components. The alloy flows well and has lower electri­
cal resistance than the majority of other filler materials. To provide sound joints, the joint
gaps (clearances) should be held to a minimum, but must be sufficient for the silver-base
alloy to freely flow into the joint owing to capillary action and gravity force. An attempt
should be made to avoid solders with considerable amounts of bismuth and lead.
The fact that the electrical and thermal properties of pure silver are superior to those of
copper has led some inductor builders to assume that the filler metal provides electrical
contact between brazed components that is as good as solid copper, which is not the case.
Figure 4.106a shows a sketch of coil current flow in the proximity of a brazed joint [330].
Porosity (Figure 4.106b) and the presence of oxides and other elements increase the elec­
trical resistance of the brazed joint area compared with that of solid copper. As a result,
excessive heat is generated in the joint area (unless the joint is located in a portion of the
coil that does not carry electrical current). Excessive heat generation causes deterioration
of brazed joints, shortening the coil life.

Coil current Brazed joint

Porous areas
(a) of brazed joint (b)

FIGURE 4.106
Sketch of coil current flow (a) in the proximity of a brazed joint. (b) Poor-quality brazed joint of coil copper.
(From V. Rudnev, Systematic analysis of induction coil failures, Part 7: Fabrication of hardening inductors,
Professor Induction Series, Heat Treating Progress, ASM Int’l, Materials Park, OH, Sept./Oct., 2006, pp. 17–18.)
Heat Treatment by Induction 283

If heat-treated parts are not properly cleaned and contain a considerable amount of oil,
residuals, and contaminants at their surface, a burning flame may occur during heating to
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

hardening temperatures. A combination of the dirt burning and generation of aggressive/


corrosive gaseous products creates a condition for brazed joint embrittlement and for the
acceleration of degradation of the copper surface. Thus, hardening inductors will last lon­
ger when heating clean parts.
A complex geometry inductor that contains numerous brazed joints, and elbow-type 90°
joints in particular, could experience impeded water flow in the cooling coil turns, a prob­
lem more likely to occur in coils fabricated with small-diameter tubing. The use of booster
pumps to provide sufficient water pressure to cool the coil might be required. However,
this can be counterproductive as excessive water pressure adds to the electromagnetic
forces and thermal stresses experienced by the copper coil, which could further weaken
the brazed joints, leading to fatigue cracking and water leaks. Also, brazed joints and the
copper itself can weaken because of work hardening during coil service, becoming brittle
and also developing fatigue cracks.
Poor quality brazed joints are prime candidates for water leaks effecting not only the coil
life expectancy but also a quality of hardened components due to a potential soft spotting
in the areas of water leaks.
Eliminating braze joints or dramatically reducing their number, particularly in current-
carrying areas, is the key to fabricating durable, reliable, and long-lasting inductors.
Chapter 5 provides more information on brazing techniques.
Note that higher frequencies are usually associated not only with a lower magnitude of coil
current and current penetration depth but also with reduced magnetic forces. As frequency is
reduced and coil current and magnetic forces are increased, more attention must be paid to coil
support and reliability and the soundness of brazed joints, because more vibration and greater
magnetic forces are typically associated with the use of lower frequencies, particularly when
using non–solenoid-style inductors. Nonconductive materials or nonmagnetic metallic studs
held together with an insulator may also be used in providing the required support.
An epoxy coating is commonly applied for copper turns of multiturn inductors for isolation
purposes (see Figure 4.86b, for example). Ceramic guides or cast refractory liners can be added
to the system when space is available and relatively low frequency is used. If a liner is dam­
aged, it usually can be easily replaced without having to remove the inductor from the heating
system. This reduces equipment downtime in applications such as in-line hardening of rods,
pins, or tubes or through heating of bars before hot working. Of course, the space required for
positioning of refractory liners comes at the expense of having a greater copper-to-workpiece
air gap, which worsens the electromagnetic coupling and electrical efficiency (Figure 3.52a).
CNC machining or improvements in copper bending technology helps accomplish the
goal of extending inductor life. An example is given in Figure 4.107, which shows a section
of rectangular copper tubing that was bent using Inductoheat’s copper bending technol­
ogy. Note that this small rectangular copper square tubing—6.4 × 6.4 mm (¼ × ¼ in.) and
1 mm (0.04 in.) thick wall—can be easily bent into a complex shape having 90° bends and
can even be double-twisted at angles of nearly 180°. The use of this technology can have a
significantly positive effect on the coil life and uptime of IH systems.
The ability to precisely and repeatably fabricate bended or brazed inductors has always
been a legitimate concern in some applications, which may require a costly, extensive, and
time-consuming validation process after installing a new set of inductors.
At Inductoheat, the great majority of high–power density hardening inductors of small and
medium size is CNC machined from a solid copper bar regardless of their complexity. This
highly accurate and repeatable machining process produces rigid and durable inductors.
284 Handbook of Induction Heating
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

FIGURE 4.107
Bent rectangular copper tubing using proprietary Inductoheat’s bending technology. Note that this small rect­
angular copper square tubing—6.4 × 6.4 mm (¼ × ¼ in.) and 1 mm (0.04 in.) thick wall—can be easily bent into
a complex shape having 90° bends and can even be double twisted at angles of nearly 180°. (From V. Rudnev,
Systematic analysis of induction coil failures, Part 7: Fabrication of hardening inductors, Professor Induction
Series, Heat Treating Progress, ASM Int’l, Materials Park, OH, Sept./Oct., 2006, pp. 17–18.)

CAD/CAM/CNC software programs that provide the appropriate cutter-to-copper spatial


relationships are created, which produce inductors of the required shape and precision.
Figure 4.108 shows a variety of finished and semifinished CNC-machined hardening induc­
tors. In the past, most of these inductors were fabricated by brazing and bending [321]. CNC
machining is a superior method to achieve accurate, robust inductors for use in automotive,
aerospace, defense, and other industries where high process repeatability is critical.
Brazing is completely eliminated with some CNC-machined inductors (e.g., inductors
used in Inductoheat’s nonrotational SHarP-C processes for hardening crankshafts and
camshafts). Brazing is minimized in other applications being used only in non–current-
carrying regions to encapsulate water-cooling channels.
Some inductors, especially those used in selective hardening, may have very complex
geometries. A computerized 3D metrology laser scanner is used at Inductoheat to ver­
ify coil dimensional accuracy and alignment precision within approximately 25 microns
(0.001 in.) after fabrication and assembly (Figure 4.109).

FIGURE 4.108
Variety of finished and semifinished CNC-machined hardening inductors. In the past, most of these inductors
were fabricated by brazing and banding [321]. (Courtesy of Inductoheat Inc., an Inductotherm Group company.)
Heat Treatment by Induction 285
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

FIGURE 4.109
Computerized 3D metrology laser scanner is used to verify coil dimensional accuracy and alignment preci­
sion within approximately 25 microns (0.001 in.) after fabrication and assembly. (From V. Rudnev, A. Goodwin,
S. Fillip, W. West, J. Schwab, S. St. Pierre, Keys to long-lasting hardening inductors: Experience, materials, and
precision, Adv. Mater. Processes, October, 2015, pp. 48–52.)

It is important to maintain sufficient wall thickness to carry the electrical currents of


a profiled inductor. The wall thickness of an inductor’s heating face should increase as
frequency decreases. This fact is directly related to both the current penetration depth in
the copper δ Cu and the copper edge effect [329]; this holds true for machined coils as well
as those made of tubing.
It is highly desirable for the current-carrying copper wall thickness to be 1.6 times greater
than the δ Cu calculated at maximum working temperature. Increased kilowatt losses in
the copper, which are associated with reduced coil electrical efficiency and greater water-
cooling requirements, will occur if the wall is thinner than 1.6∙δ Cu.
In some cases, the copper wall thickness can be noticeably thicker than the recom­
mended value of 1.6∙δ Cu. This is because it may be mechanically impractical to use a tubing
wall thickness of, for example, 0.25 mm (0.01 in.).
Table 4.18 shows the variation of penetration depth in pure copper δ Cu versus frequency
at room temperature (20°C/68°F) [331]. For example, if a 6.35-mm-thick (0.25-in.-thick) solid
copper bus carries a current of 2000 A at a frequency of 30 kHz, in reality, the majority
of that current (approximately 1260 A) will be concentrated within a thin layer less than
0.5  mm (0.02 in.) thick. Therefore, practically speaking, the entire AC will be localized
within a surface layer equal to 1.6∙δ Cu. At this frequency, this will be less than 0.75 mm,
which is only 12% of the entire copper bus thickness. The rest of the copper primarily
serves mechanical purposes, including providing support against flexing and bending.
For quick estimation, δ Cu can be calculated with the formula:

70 2.75
δ Cu = (mm), δ Cu = (in.),
F F

TABLE 4.18
Current Penetration Depth (δ Cu) into Copper versus Frequency at Room Temperature (21°C/70°F)
Frequency (kHz)
0.06 1 3 10 30 70 200
δCu (mm) 8.8 2.2 1.2 0.7 0.4 0.26 0.15
286 Handbook of Induction Heating

where F is frequency (in hertz).


The insulating paints, epoxy, and ceramic coatings, as well as different varnishes, can be
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

used to provide certain electrical insulation and protect coil copper during its operation
from external heat exposure, extending inductor service life. Ceramic coating is used in
high-temperature applications; others are used in cases of maximum service temperatures
being less than approximately 150°C (300°F).
Electrical insulators should be properly designed and checked periodically for holes,
burns, or contamination. Otherwise, parasitic electrical current flow and micro-arcing
may develop. There is a potential for parasitic electric coil current due to real-life accumu­
lation of the dirt and scale dust caused by insufficiently sized electrical insulation between
coil legs. Initially, a parasitic coil current flow might not take place, but taking into con­
sideration a real-life working environment, with time it may occur compromising heating
quality. Depending on its magnitude, system sensors (e.g., ground fault detector) may or
may not detect micro-arcing, negatively affecting performance of the induction coil. It is
not always easy for an operator or electrician to see or hear an occurrence of micro-arcing.
Therefore, it is important to use appropriate electrical insulation with sufficient dielec­
tric properties. Regular and thorough inspection of the coil, bus network, and condition of
electrical insulation is very important. Care must be used to avoid damaging the inductors
when changing, handling, and storing them.

4.2.4.3 Surface Conditions
Copper surface conditions are another important factor in coil fabrication. The combina­
tion of the several electromagnetic phenomena discussed in Chapter 3 results in a current
concentration within a thin layer of the coil surface that faces the heated workpiece. This is
where the great majority of coil cracks initiate.
Even a seemingly small notch, indentation, tool mark, grinding damage, arcing mark,
scratch, or excessive surface roughness can become a crucial stress riser that may trigger
crack initiation.
Figure 4.110 illustrates a crack developing across the copper thickness (assuming that the
workpiece is located above the coil copper) [331]. As follows from Table 4.18, for most of the
frequencies utilized in IHT, a majority of the coil current occupies a very thin surface layer
often less than 0.5 mm. This is why even a small notch can become a crucial stress riser
that results in abnormal current flow.
Figure 4.110a shows the normal flow of coil current. However, the normal flow will be
disturbed by a fine crack, deep scratch, or tool mark on the surface of the coil. Figure 4.110b
shows how a crack blocks the normal current flow, resulting in the appearance of a local­
ized flow anomaly—the current at the surface is forced to take a deep detour around the
crack. As a result, the current density will be at a maximum at the root of the crack, where
additional heating takes place.
This abnormal current flow in combination with excessive heat generation creates a con­
dition favorable for opening the crack. Similar to a “snowball” effect, further acceleration
in crack growth will then take place, and it will be accompanied by an increased severity
of localized overheating of the crack root area, as shown in Figure 4.110c, and a reduction
of coil mechanical strength.
To complete the study of the effect of current flow on coil crack development, it is impor­
tant to keep in mind that, in reality, crack propagation has a 3-D nature and all the electro­
magnetic phenomena discussed here have complex interactions. Therefore, not only does
Heat Treatment by Induction 287

Copper surface faces Current penetration depth


heated workpiece (63% of total coil current)
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

Coil copper
thickness

(a)

Coil current Crack initiation

(b)
Severe
overheat

Coil copper
thickness
Internal area
(c)

FIGURE 4.110
Illustration of a crack development across the copper thickness (assuming that the workpiece is located above
the coil copper). Panels a through c illustrate a deviation of the coil current flow (b and c) within the copper
thickness from the normal flow (a). (From V. Rudnev, Systematic analysis of induction coil failures, Part 2: Effect
of coil current flow on crack propagation, Professor Induction Series, Heat Treating Progress, ASM Int’l, Materials
Park, OH, Sep./Oct., 2005, pp. 33–35.)

the crack deepen as it propagates, but it also widens. Figure 4.111a shows that a similar
snowball effect takes place during crack development across the width of the coil copper
that faces the heated workpiece [331].
From the very beginning, the crack blocks the normal current distribution and forces the
current to flow around the crack. This leads to a current density surplus at the crack edges.
Excessive current density produces extreme heat at the crack edges and the appearance of
hot spots that, in turn, reduce coil copper strength there and promotes further acceleration
of crack widening, as shown in Figures 4.111b through d.
This leads to the conclusion that, in addition to other factors, the direction of the coil
fracture often has a specific relationship to the coil current flow.
Abnormal service conditions that result in mechanical damage to an inductor is one of
the most typical factors that lead to premature coil failure. An example is shown in Figure
4.112a where the crossover segment of the single-shot inductor has numerous marks and
scratches resulting in severe copper surface damage. The inductor had been used for hard­
ening shafts including the spline area. Most likely, a shaft was not properly located in the
heating position. Sensors that are supposed to detect improper shaft positioning either
failed to detect it or had been disconnected for some reason. The scratches resulted from
the rubbing of the shaft spline against the induction coil during either loading/unloading
288 Handbook of Induction Heating

Current flow Crack initiation


Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

Copper
width

(a)
Current density
surplus

Copper
width

(b)
Noticable
overheat

Copper
width

(c)
Severe
overheat

Copper
width

(d)

FIGURE 4.111
Snowball effect takes place during crack development across the width of the coil copper that faces the heated
workpiece. Panels a through d illustrate a deviation of the coil current flow (b, c and d) across the coil heating
face from the normal flow (a). (From V. Rudnev, Systematic analysis of induction coil failures, Part 2: Effect of
coil current flow on crack propagation, Professor Induction Series, Heat Treating Progress, ASM Int’l, Materials
Park, OH, Sep./Oct., 2005, pp. 33–35.)

(a) (b)

FIGURE 4.112
Abnormal service conditions that result in mechanical damage to a coil is one of the most typical factors that
lead to premature coil failure. (a) Shows severe damage of coil copper face due to numerous marks and scratches
caused by improper shaft positioning. (b) Example of 3-II cracking on coil copper heating face. (From V. Rudnev,
Systematic analysis of induction coil failures, Part 2: Effect of coil current flow on crack propagation, Professor
Induction Series, Heat Treating Progress, ASM Int’l, Materials Park, OH, Sep./Oct., 2005, pp. 33–35.)
Heat Treatment by Induction 289

or shaft rotation. As would be expected, those scratches represent crack initiation sites,
culminating in coil premature failure.
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

An example shown in Figure 4.112b illustrates the 3-D nature of crack development. The
inside diameter of a single-turn coil was damaged during abnormal part loading in ser­
vice, resulting in a damaged current-carrying face, inappropriate current flow, fast crack
growth, and premature coil fracture. Note that this failure initiated at the damaged area.
The discussion above reveals that the surface of the copper (particularly its current­
carrying regions) should be smooth and clean, with a minimum amount of the geometrical
stress risers. Surface discontinuities and irregularities resulting from improper fabrication
techniques, careless assembly, storage, or shipment can have a dramatic negative impact on
coil life. Damage to the current-carrying inductor surface, be it preexisting cracks, tool marks,
notches, excessive surface roughness, or other stress risers, should be avoided [336,395].

4.2.4.4 Water Cooling of Hardening Inductors


Because of the appreciably high currents and powers applied to the coil in a majority of
hardening applications, inductors are water cooled. Water-cooling systems must be sized
properly and have sufficient chilling capability to be in the temperature range specified by
the original equipment manufacturer (OEM).
Sufficient water cooling is extremely important for equipment longevity. In some cases,
inductors have their own water-cooling circuit, but in others, they are connected to the
exit of the power supply cooling circuit. To avoid water condensation (which is harmful
for power electronic devices), the water temperature for cooling the power supply is main­
tained above the dew point, which is typically above approximately 32°C to 33°C (90°F to
92°F). If coil copper overheating occurs, it may be beneficial to use a separate water circuit
for coil cooling, allowing maintenance of the incoming water temperature to be approxi­
mately 20°C to 22°C (68°F to 72°F).
Besides a temperature range, closed-loop water-cooling recirculating systems should
satisfy certain requirements that are discussed in Section 7.12 and in Ref. [338]. Always
refer to the OEM’s specification when in doubt regarding the needed water quality because
certain process subtleties may call for specific requirements.
Typical water specifications for a closed loop cooling system are as follows:

• Deionized water is preferable.


• pH, 7–8.5.
• Hardness, 150 ppm.
• Conductivity, 400 μSiemens/cm (μmhos/cm) max.
• Chlorides, 20 ppm max.
• Sulfates, 1000 ppm max.
• Nitrates, 10 ppm max.
• Solids, 250 ppm total solids content.

Copper tubing used for coil fabrication is naturally profiled for water cooling. Water
passages are incorporated in the design if the inductor is CNC machined from a solid cop­
per block. In either case, all efforts must be made to remove all restrictions in the water
passages to allow smooth and uninterrupted water flow and to avoid excessive turbulence,
which can lead to steam pockets. When the copper tubing or the water passage becomes
too small or severely restricted, overheating will occur.
290 Handbook of Induction Heating

Industry has developed several empirical formulas to estimate the needed water-cooling
requirements and flow. One such expression is shown below:
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

PK 1 K 2
gpm = ,
K 3 T

where gpm = gallons per minute, P is total coil power (kW), K1 is the tubing coefficient (for
the great majority of high-frequency IHT applications, K1 = 0.5), K 2 = 3415 is a conversion
constant that is derived from Btu/kWh, and K 3 is a conversion constant that represents the
normalized heat capacity to allow for proper cooling of the copper. ΔT (°F) can be mea­
sured by the use of an in-line thermometer as close to the inductor as possible.
In some cases, the water passage size might not provide the above-calculated gpm and coil
life can suffer. The use of a high-pressure booster pump can increase the water flow; however,
keep in mind that it is flow (and not pressure) that is usually the most critical factor.
In most hardening applications, the workpiece surface temperature exceeds 850°C
(1562°F). The intense radiant heat from the surface of the heated component is comple­
mented by the copper heating from the Joule losses. The combined effect of both factors
can be detrimental to the coil.
In induction hardening, the coil might be positioned in proximity to the workpiece to
maximize electrical efficiency and localize the heating (refractory or liners are not typi­
cally used except in continuous/progressive hardening). The air gap between the work­
piece and the coil could be as small as 1.25 mm (0.05 in.). With the hot workpiece located
so close to the coil, the design of the cooling passage is very critical. It is obvious that the
cooling channel should be as close as possible to the inductor’s heating face, particularly if
magnetic flux concentrators are applied.
When hardening components with complex shapes, it is highly desirable that the cool­
ing passage is properly profiled, locating them sufficiently but safely close to the copper
heating face.
Consider the following example. Figure 4.113 (top left) shows an improperly drilled hole
of a water-cooling pocket for a single-shot hardening inductor [336]. An attempt was made
to position the water-cooling passage as close to the inductor’s heating face as possible.

Original coil nose Revised coil nose

FIGURE 4.113
Examples of improper water-cooling pockets resulted in premature inductor failure.
Heat Treatment by Induction 291

The drift in the positioning of the water-cooling hole weakens the mechanical strength
of the copper wall. At the same time, it also obstructs the coil current flow, squeezing
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

the electrical current into an extremely narrow copper passage. This results in a dramati­
cally increased current density and the generation of intense heat, which may promote
water vaporization and the formation of a steam vapor barrier in that region. Regardless of
what may appear to be sufficient water-cooling flow, the presence of a steam vapor barrier
will essentially act as a thermal insulator inside the water-cooling pocket and appreciably
restrict the copper cooling. The combined impact of both factors could lead to premature
coil failure.
Figure 4.113 (top middle) shows another example of improper water cooling. The longi­
tudinal legs of a single-shot inductor were profiled to accommodate the required hardness
pattern in the shaft. However, a cooling passage that was drilled straight did not accom­
modate the profile of the inductor’s heating face. This led to an unreasonably large copper
thickness (between the water-cooling passage and the current-carrying copper surface)
exceeding 8 mm (0.32 in.) and resulted in insufficient removal of the heat generated in the
current-carrying regions of the coil copper. Accelerated deterioration of the copper surface
(Figure 4.113, top right) and premature coil failure were the result.
Two sketches shown on Figure 4.113 (bottom) illustrate a comparison of originally
designed CNC machined single-turn inductor which failed because of severe overheating
of its heating face/nose (bottom left) versus revised design (bottom right) that exhibit more
than twofold improvement in coil life.
Figure 4.114 shows another example of single-shot inductors with improperly machined
water-cooling channels. Discoloration of overheated copper is clearly visible. As a solution,
the water-cooling pocket should be profiled to accommodate the features of the inductor
geometry and, in particular, its heating face, keeping the copper wall thickness sufficiently
thin for effective heat extraction, yet providing adequate mechanical support and avoiding
excessive electrical resistance.
Therefore, the geometry of the water-cooling pockets and their profiling should accom­
modate sufficient cooling of all critical regions.
According to the coil copper edge effect, the application of high frequencies leads to
a higher current density concentration at the copper turn corners. This “edge-catching”
effect of coil current density can result in “hot spots” at the copper corners [329,336]. The
combination of surplus current density at stress riser regions makes these inside corners
potential crack initiation sites, particularly if cooling is marginal. This is also the reason
why it is preferable to use rectangular water-cooling channels instead of round passages,
when using square or rectangular tubing and particularly when high frequencies are
used. Small radii on coil turn edges may also be beneficial.

FIGURE 4.114
Discoloration of overheated copper caused by improperly machined water-cooling channels of single-shot
inductors. (From V. Rudnev, Induction Heat Treating of Automotive Components, Invited lecture for Linamar Group
Companies, Guelph, Canada, July 15, 2014.)
292 Handbook of Induction Heating

When a scan inductor comprises two or more turns, the water-cooling inlet should be con­
nected to the trailing turn and the water outlet should be connected to the leading turn.
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

Case study. Figure 4.115 shows a two-turn scan inductor that has failed because of over­
heating [336]. The dark turn corresponds to the lead turn and the bright turn is the trailing
turn. Coil water-cooling requirements were incorrectly scaled down from the larger coil
to one with a smaller diameter. Because of the greater scan rate, the smaller-diameter coil
required almost the same total coil power as the larger version. Regardless of very similar
total power, this resulted in measurably lower coil volts, but noticeably greater coil current,
leading to a respective increase in copper kilowatt losses and excessive heat generation.
The water-cooling inlet was connected to the lead turn, while the water outlet was con­
nected to the trailing turn. However, the shaft’s surface temperature under the trailing
turn is substantially greater than its temperature under the leading turn. As a result, the
coil’s trailing turn was exposed to much more intensive heat (because of thermal radiation
and convection) than the leading turn. In addition, the trailing turn experiences a non­
magnetic load (because the workpiece surface is at austenizing temperature) compared to
the leading turn, which experiences a ferromagnetic load. This also affects kilowatt loss
distributions among the turns.
Therefore, it would be more appropriate in this case to connect the water-cooling inlet
to the trailing turn and the water outlet to the leading turn. Inappropriate water-cooling
scaling and incorrect connection of water-cooling circuits resulted in overheating of the
trailing turn and premature coil failure.
Foreign deposits (e.g., calcium buildup, hard water deposits, suspended solids, biological
contaminants, etc.) may build up slowly inside water-cooling passages, reducing their cool­
ing capacity and resulting in copper overheating and failure particularly at the brazed joints.
Figure 4.116 shows debrazing of the brazed joints (left) and collected hard water deposits

FIGURE 4.115
Two-turn scan inductor that has failed because of overheating. (From V. Rudnev, Induction Heat Treating of
Automotive Components, Invited lecture for Linamar Group Companies, Guelph, Canada, July 15, 2014.)

(a) (b)

FIGURE 4.116
Debrazing of the brazed joints (a) and collected hard water deposits (b) build up on the inside surfaces of the
water-cooled copper tube.
Heat Treatment by Induction 293

(right) buildup on the inside surfaces of the water-cooled copper tube. Preventive mainte­
nance, appropriate water-cooling treatments, and filtration can help avoid this problem.
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

There are some cases where water cooling of the inductor can be eliminated. For example,
low–power density coils made from Litz wire, used for low-temperature and low–power
density heating of magnetic parts, are very efficient, generating low power loss within Litz
wire; air cooling might be sufficient to cool this type of coil. In other cases, a large mass of
copper (in the coil) can provide sufficient cooling, when short heat times and low duty cycles
are applied. Thus, it might not be necessary to use water cooling, because under these condi­
tions, the copper simply does not have time to significantly heat up; the heat that is generated
is dissipated by the large mass of copper, which can be also cooled by a quenchant.
Other aspects related to water cooling are discussed in Section 7.12.

4.2.5 Frequency Choice and Power Selection


For any induction heating application, it is important to select the correct power and the
appropriate frequency. Induction equipment is rated in frequency and maximum power
output. A variety of available power supply designs are available and are discussed in
Chapter 7.
The frequency, power density, and duration(s) of heat-soak stage(s) determine the process
recipe/protocol and temperature profiles. The frequency is usually chosen first because, as
discussed earlier, all electromagnetic phenomena involved in IH are affected by frequency.
The frequency ranges for induction hardening can be as high as 4000 kHz (hardening
thin products) to as low as 60 Hz (hardening large rolls). IHT practitioners typically divide
frequencies into three categories: low frequency up to 10 kHz, medium frequency 10 to
70 kHz, and high frequency anything higher than 70 kHz. High frequency is also referred
to as radio frequency (RF), because it is above the audible range for human hearing. As the
frequency is increased, the depth of heat source generation is decreased. For example, a
system designed to operate in the 600-kHz range would have a very shallow depth of heat­
ing, whereas a 1-kHz system will produce relatively deep heating.

4.2.5.1 Through Hardening
As discussed in Sections 3.1.2 and 4.2.2.2, when through heating (e.g., through hardening)
parts, care must be taken to avoid an eddy current cancellation. There are several condi­
tions required for the component to be able to harden through its entire cross section.
These conditions include but are not limited to the following:

• It is required to heat as uniformly as possible the entire cross section of the work­
piece to sufficiently high temperatures that enable the formation of an austenitic
structure with a sufficiently homogeneous distribution of carbon and other chem­
ical elements (i.e., alloying and residual elements).
• After sufficient austenitization, it is necessary to quench the entire workpiece to
low enough temperatures developing an appropriate cooling severity to allow the
formation of fully or predominately martensitic structures (e.g., the presence of
certain carbides and nitrides within martensite as well as a certain amount of RA
may be inevitable).
• The chosen grade of steel, cast iron, or P/M material should have a sufficient hard­
enability to allow achieving the desirable hardness in the core.
294 Handbook of Induction Heating

As an example, Figure 4.18 shows the dynamics of through hardening an SAE 4340 car­
bon steel shaft (16 mm diameter) in a normalized condition using a frequency of 10 kHz.
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

The process cycle includes 8 s of heating and 0.5 s of dwell/soak. A short time delay (also
called a dwell or soak time) is often used in through hardening applications to help to
improve the “surface-to-core” temperature uniformity and reduce the thermal shock dur­
ing the initial stage of quenching.
It is beneficial in through hardening applications to use lower frequencies than in sur­
face hardening (assuming a comparable diameter/thickness of the workpiece) because of
the larger current penetration depth and a more uniform “surface-to-core” temperature
distribution.
However, if the selected frequency is too low, then eddy current cancellation may occur,
resulting in a dramatic reduction in the electrical efficiency. This is so, because eddy
currents circulating in opposite sides of the heated workpiece are oriented in opposite
directions and might start canceling each other if δ becomes comparable to the workpiece
physical size. The relative effect of applying two frequencies (F1 and F2, where F2 is greater
than F1) is shown in Figure 3.13a. As an extreme case, the heated component might become
semitransparent or even transparent to the EMF and, at a certain point, the temperature
rise may halt without reaching the needed temperatures for proper austenitization as it
was illustrated earlier in the case study shown in Figure 4.71.
When through hardening solid cylinders, the diameter-to-δ ratio should be greater
than 4. This will ensure maximization of coil electrical efficiency and helps avoid eddy
current cancellation. Keep in mind that δ should be calculated at the final heating
temperatures. If this ratio is less than 2.8, then there will be substantial eddy current
cancellation.
A ballpark estimation of the needed power for through heating applications has been
discussed in Section 3.3.1.
If the workpiece has an irregular shape (e.g., C-shaped tubes, odd-shaped parts, or slot­
ted cylinders; see Figure 4.117), the eddy current will flow on the outside and inside areas
of the part in order to provide an uninterrupted current loop and current cancellation can
occur, manifesting itself in an underheating of the entire component (bottom left) or the
appearance of localized spots exhibiting the heat deficit compared to neighboring regions
(top-right). Thus, in order to avoid eddy current cancellation, δ should be no more than
one-fourth of the thickness of the current-conducting path.

4.2.5.2 Surface Hardening
Frequencies used in surface hardening applications result in a pronounced skin effect,
which typically prevents eddy currents induced within the workpiece from canceling each
other.
In surface hardening, the desirable frequency (Figure 4.67a) results in δ at austenite
phase temperature that exceeds 1.2- to 2-fold the required hardness case depth. On the
other hand, for other-than-desirable frequency, it might still be possible to achieve the
same desired case depth by using different combinations of power densities and scan rates
(Figure 4.67b and c). Section 4.2.2.1.6 reviews these conditions.
Over the years, the industry has accumulated various simplified rules of thumb for fre­
quency selection. This includes suggestions shown in Table 4.19 that can be roughly used
for ballpark estimation of the required frequency and power density.
Heat Treatment by Induction 295

“C”-shaped and “Slotted”-shaped parts


Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

Coil Workpiece Coil Workpiece

Coil Workpiece Coil Workpiece

Workpiece is too thin Workpiece is thick enough,


compared to current preventing current
penetration depth (there cancellation. There is
is current cancellation). a distinguished current path.

FIGURE 4.117
Current cancellation in induction heating of C-shaped parts.

TABLE 4.19
Frequency and Power Density to Obtain Various Hardening Depths in Carbon Steel
Frequency (kHz) Hardening Depth (in.) Low Power Density High Power Density
450 0.015–0.045 7 12
0.045–0.090 3 8
10 0.060–0.090 8 15
0.090–0.160 5 13
3 0.090–0.120 10 17
0.160–0.200 5 14
1 0.200–0.280 5 12
0.280–0.350 5 12
Contour gear 0.015–0.045 15 25
hardeninga 450–200
Source: V. Rudnev, J. Dossett, Induction surface hardening of steels, in ASM Handbook, Volume 4A: Steel Heat
Treating Fundamentals and Processes, J. Dossett and G. Totten (editors), ASM Int., Materials Park, OH, 2013;
ASM Handbook, Vol. 4, Heat Treating, ASM International, Materials Park, OH, 1991.
a A low power density preheat at 3 or 10 kHz is recommended for contour gear hardening.
296 Handbook of Induction Heating

According to another common practice, the required frequency for surface hardening
of solid cylinders with a case depth ranging from 1.6 to 5 mm can be roughly determined
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

based on the following conditions:

2 2
 6.5   16.6 
 X  < Frequency (kHz) <  X 
CD CD

2
 9.8 
Frequency (kHz) ≅  ,
 X CD 

where XCD is the required case depth (mm).


It is important to remember that most “rules of thumb” are extremely subjective and
typically were developed for a selection of frequency and power for simple geometries and
plain carbon steels in fine-grain normalized conditions with a carbon content of approxi­
mately 0.4%–0.5%.
As shown in Figure 4.27, even for steel of the same grade (AISI 1042), the required induc­
tion hardening temperature range can vary substantially depending on the heat intensity
as well as the prior microstructure of steel:

• 1620°F to 2000°F (880°C to 1095°C), for annealed prior microstructures


• 1550°F to 1830°F (840°C to 1000°C), for normalized prior microstructures
• 1510°F to 1710°F (820°C to 930°C), for Q&T prior microstructures

It is imperative to remember that the procedure of determining the most suitable com­
bination of process parameters using “rules of thumb” could easily be misleading. Each
component to be induction hardened has its own “personality” with respect to material
specifics, prior structure, geometry, and functionality, suggesting that numerical computer
modeling is a much better option to avoid unpleasant surprises when selecting process
parameters.

4.2.5.3 Heat Duration
After selecting the appropriate frequency and power density for the needed depth of heat­
ing, based on the hardness case depth specification, the last parameter is the duration of
heat. As previously stated, components can be heated to hardening temperatures within
seconds or even a fraction of a second. Time and temperature are the two major factors
responsible for establishing the final structure and affecting shape distortion. Rapid heat­
ing tends to primarily heat the area where eddy currents have been induced, resulting in
a short transition zone. With the increase of time, the thermal conduction starts to play
a more dominant role, allowing the heat to soak from higher-temperature areas toward
lower-temperature regions. This leads to a fuzzy transition and deeper case depth.
In order to decrease the distortion of symmetrical parts, it is usually desirable to have
the heating time as short as possible. However, there are some limitations. First, the mate­
rial must reach the minimum required transformation temperature at the depth to be
hardened. If the frequency or surface power density is unreasonably high, the surface can
be overheated even though the required temperature is held at the case depth. Also, as a
Heat Treatment by Induction 297

result of the short cycle time, large temperature gradients can occur; thermal stresses can
reach their critical value and cracking can develop.
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

When heating asymmetrical parts, it may be desirable to use a longer heating time with
lower power densities and frequencies. Discontinuities such as sharp corners and edges
will heat faster with an increase of frequency and power density owing to local eddy
current concentration. Thus, a slower heat is often desired, and possibly a dwell before
quenching to allow the thermal conduction to minimize any “hot” spots.
To determine the best process recipe/protocol for each component, some experimenta­
tion with heating times as well as other process parameters may be needed. As long as
the appropriate frequency and power size are selected, the duration of heat can usually be
determined relatively quickly after several parts are sectioned and evaluated for hardness
pattern. Mathematical modeling provides valuable assistance in this respect.

4.2.6 Inductor Mounting Styles


In order to accommodate a variety of heated parts while providing high efficiency and the
required hardness patterns, inductors are frequently changed and considered as perish­
able tooling. The dilemma is that an operator should be able to quickly change inductors
and provide a reliable, low-resistive contact between the coil and power supply. The elec­
trical connection between the inductor and the transformer or bus is sometimes referred
to as the mounting foot.
There are several standard types of mounting feet that have been developed over the
years. A precision-machined keyed foot ensures a good electrical contact. It is usually fas­
tened with four stainless steel bolts.
In some cases, high-current hardening inductors have wide flat contacts and bolts
spread wide (Figure 4.118, left). This design may potentially create a poor contact between
coil foot and bus connection. If bolts loosen and contact joint opens, then bolts become an
electrical current path, which may result in excessive heat generation within the bolts. It
may also potentially make a “coil foot-to-bus work” contact less reliable. As a result, poorly
controlled transient electrical resistance and additional kilowatt losses may take place.
Control systems that rely on stabilization of the inverter output voltage, current, or power
might not properly respond to lack of heat generation within the heated workpiece caused
by additional and poorly controlled kilowatt losses in bus connections.
As an alternative, an inductor with keyed contacts and bolts close to center are preferable
(Figure 4.118, right). This design ensures good electrical contact (with reduced probability
to arcing and pitting) between coil foot and a bus, and provides electrical current flow
carried by key copper contact surfaces. Bolts in relieved area have reduced probability of
carrying significant electrical current. This type of mounting feet has become a standard

FIGURE 4.118
Two possible designs of the coil foot: (left) conventional design versus (right) keyed foot.
298 Handbook of Induction Heating

in modern induction heat treating systems providing more reliable electrical contact. If
required, some keyed feet contain orientation pin holes for proper positioning of the heat­
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

ing coil in the working position.


When fast or frequent changeover is needed owing to a wide family of parts requiring
different inductor designs, a quick-change coil adapter is often recommended. This can be
in the form of a toggle clamp, dove tail, or pneumatic. Insta-change coil adapters can also
be keyed for repeatable coil-to-part setup having the following design features:

• High force clamp springs to compensate for temperature changes.


• O-ring sealing to avoid the necessity to change hoses.
• Self-locking toggle.
• No special tools are required.
• Coils are precisely positioned using a positive stop screw adjustment.

When inductors are constructed of tubing, compression type fittings can be used.
Because the fittings are carrying the current, a special conductive sealant is used to ensure
that the connection does not leak coil coolant.
In all cases, to ensure a good electrical contact (avoid arcing), the contact areas must be
clean and free from nicks, burrs, scratches, and contaminations. Figure 4.119 shows an
example of arcing at a bus connection that could be caused by insufficient torque on bolts
holding the buswork combined with dirt accumulation. This example illustrates the impor­
tance of having appropriate and reliable electrical contacts. It is essential that fasteners are
tightened enough to do their job. Loose bolts result in increased electrical resistance in the
bolted contact area, localized overheating, arcing, and degradation, and potentially lead to
inappropriate heating of the workpiece. Unexpected and variable heat dissipation within
the coil copper itself and its electrical connections results in respective variable kilowatt
losses associated with it. This variability might not be detected by control system poten­
tially producing some variations in hardness pattern and causing a necessity to manually
adjust coil’s voltage/power levels. However, too tight is not desirable either. Quite often,
fasteners are tightened based on subjective judgment—it simply “feels” sufficiently tight. In
reality, insufficient tightening or overtightening might take place. Neither is desirable. This
can be prevented if properly calibrated torque wrenches are used for bolted joints [336].
Carbon steel bolts as well as magnetic stainless-steel fasteners should not be used for
electrical connection between the coil head and bus work. Typically, the use of ferromag­
netic fasteners in induction heating is prohibited. Ferromagnetic bolts, nuts, and wash­
ers are inductively heated to a much greater degree compared to nonferrous materials.
Repetitive heating and cooling (expansion–contraction cycles) of ferromagnetic fasteners

FIGURE 4.119
Evidence of arcing at a bus connection that could be caused by insufficient torque on bolts holding the buswork
combined with dirt accumulation. (From V. Rudnev, Induction Heat Treating of Automotive Components, Invited
lecture for Linamar Group Companies, Guelph, Canada, July 15, 2014.)
Heat Treatment by Induction 299

result in their eventual loosening. For many heat-treating applications, 10–12 mm dia.
(3/8 to 1/2 in.) brass, silicon bronze, or nonmagnetic stainless-steel bolts are used for the
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

inductor-to-bus connection. Stripped fasteners should not be used. Bolts should be exam­
ined and replaced if integrity is compromised.
All electrical connections must be properly maintained. This includes a regular and
thorough inspection of the coil and bus network (including the buses themselves, inductor
adapters, bus extensions, etc.).
Care must be used to avoid damaging the inductors when changing, handling, and stor­
ing them. Pertinent aspects of inductor storage and maintenance will be discussed later in
this chapter.

4.2.7 Quenching and Spray Quench Designs


4.2.7.1 Introduction
To this point, we have focused primarily on the heating portion of IHT. As mentioned
in Section 4.1, the hardening of steels, cast irons, and P/M components involves not only
raising and maintaining (if required) the temperature of the alloy at the austenite phase
(until the needed diffusion processes take place) but also sufficiently fast cooling below the
Ms temperature to form as-quenched martensitic structures that are typically required. It
should be also mentioned at this point that there are much less frequent cases of induc­
tion hardening when instead of forming martensitic structures it might be desirable to
form predominately bainitic or even fine pearlitic structures [155,340] (see Section 2.1.1).
Nevertheless, it is more the exception than the rule, and for the great majority of hardening
applications, the goal is to develop fully or predominately martensitic structures requiring
intense cooling.
There are two most common types of quenching techniques used with induction harden­
ing: spray quenching (Figure 4.55a) and dunk (immersion) quenching (Figure 4.120). In some
rare cases, self-quenching or slack quenching may be used when appreciably small hardness
case depths (usually less than 0.75 mm) are needed on a sufficiently massive workpiece.
Spray quenching is the most popular quench method used in the majority of induction
hardening applications. Figure 4.56 illustrates the three most typical spray quench design
concepts. In some cases, the quench system is built into the coil that is called MIQ (Figures
4.56a and 4.57), while in other cases, a quench follower (ring, barrel, or quench block) may
be used, which is separate from the induction coil (Figure 4.56b and c).
There are a variety of complex thermo-hydro-dynamic processes involved in spray
quenching, with the cooling intensity being a function of several factors including the

FIGURE 4.120
In some applications after a proper austenitization, a workpiece may be simply dunked, dropped, or lowered
into a quench medium that is static or gently agitated. (Courtesy of Inductoheat Inc., an Inductotherm Group
company.)
300 Handbook of Induction Heating

surface temperature of the workpiece, type and purity of quenchant, pressure/flow, design
of the quench assembly, number and distribution of quench holes (orifices), size of ori­
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

fices and angle of drilled holes (impingement angle), quenchant temperature, and others.
Usually, spray quench devices are directed at an austenitized surface of the workpiece
with the intent to completely cover the austenitized area with quench fluid. Quench ori­
fices are placed facing the surface of the heated component at approximately 4- to 6-mm
intervals in a staggered pattern (Figure 4.121).
The orifice diameter is related to the specifics of the quenching requirements, including
the coil and workpiece geometries, the air gap between the quench device and the work­
piece, the type of quenchant, its concentration, required flow, and others (see Table 4.20).
The quench assembly may be positioned quite closely to the inductor. In order to mini­
mize eddy current losses, it is usually made of nonmetallic materials (Figure 4.121a). Dirt
and foreign deposits (e.g., hard water deposits) can clog the orifices, which may build up
slowly in quench devices, reducing their cooling capacity and introducing cooling non­
uniformity and may necessitate extensive cleaning or replacement. Various aspects of the
maintenance of quenching devices will be discussed later in this chapter.
The primary objective of quenching is to provide the required rate of heat removal to
arrive at the desired microstructure, hardness level, and pattern, which produce certain
industrial characteristics of the materials (e.g., strength) while optimizing the residual stress
distribution [162]. It is equally important to accurately repeat both process stages (the heat­
ing and the cooling) in order to produce the same results on every part that is heat treated.
In order to facilitate this process, a number of standard tests have been developed to
characterize both the heat-treated material and the cooling capability of a particular
quench medium. Although often these curves are not directly applicable for induction
surface hardening utilizing spray quenching, they are still useful in terms of assessing the
results and determining the direction for any needed changes in the process.
At first thought, it might seem that it would be “the faster the better” to cool the part.
Unfortunately, some materials cannot be rapidly cooled without cracking or severe distor­
tion. Even in cases where a relatively forgiving material is used, the design of the part to be

FIGURE 4.121
The orifice size is related to the specifics of the quench design. They are typically placed facing the part in a
staggered pattern. (Courtesy of Inductoheat Inc., an Inductotherm Group company.)

TABLE 4.20
As the Workpiece Diameter Increases, the Orifice Size Should Also Increase to Help Provide
a More Uniform Quench
Shaft Diameter (in.) Shaft Diameter (mm) Orifice Size (in.) Orifice Size (mm)
0.25–0.50 6.5–13 0.046–0.063 1–1.5
0.50–1.50 13–38 0.063–0.094 1.5–2.5
>1.50 >38 0.115–0.156 3–4
Heat Treatment by Induction 301

hardened may include sharp corners, edges, holes, or a combination of thick and thin sec­
tions that are challenging to heat and cool uniformly. These cases require the use of spe­
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

cially designed quench devices providing the ability to tailor the localized cooling rates in
order to provide the desired hardness while eliminating cracking or excessive distortion.
If the cooling rate is insufficient to transform the needed percentage of austenite into
martensite, the material will likely transform a considerable amount of austenite into
softer constituents such as pearlite and upper bainite (Figure 4.7), which would result
in low-strength areas within the martensite, a complex and undesirable distribution of
residual stresses, and potential failure of the part in service owing to cracking or excessive
wear, just to name a few.
A typical practice for developing the cooling curve for a given alloy is to heat a sample of
the material to the austenitizing temperature and then quenching it in the chosen quench
medium under defined conditions. Thermocouples on the surface and center of the part
are used to generate a curve for the temperature versus time. Most standard cooling curves
are generated for immersion quenching, revealing three typical stages in the process of
quenching the part to the temperature of the quench medium [162,166,167,172–175,339,429].
The first stage (Stage A) is referred to as the “vapor blanket” stage when the heat from
the workpiece exceeds the amount of heat needed to produce the maximum vapor per unit
area of the workpiece. This stage is characterized by slower cooling since the vapor blanket
acts as a thermal insulator and the primary means of cooling is by thermal radiation. Some
aqueous solutions may substantially suppress this stage, dramatically reducing its impact.
The second stage of quenching is referred to as the “nucleate boiling phase” (Stage B).
This stage produces the highest rate of heat transfer and the greatest cooling rate.
The last stage of quenching (Stage C) is associated with convective heat transfer;
where the cooling is accomplished by heat convection from the workpiece to the quench
medium. The cooling intensity at this stage is much lower than the boiling phase [339].
As has been discussed in Section 4.1.5.3, classical cooling curves obtained for immersion
quenching have limited use in spray quenching. Further discussion on this subject will be
provided in Section 4.2.7.3.

4.2.7.2 Quench Media
In the quest to arrive at the most suitable cooling conditions for a given material and part
geometry, a variety of different quench media may be used, beginning with something
as simple as water and progressing through various aqueous solutions including water-
based polymer solutions, petroleum oil–based quenchants, vegetable oils, water mist, and
gaseous quenchants including forced air. Noticeable efforts in developing various quench
media as well as defining their characteristics have been done by G.E. Totten, H. Tensi,
S. MacKenzie, C. Bates, L. Canale, and others [162,170–175].
Besides being application specific, the selection of quench media is also affected by its
cost, the cost of its disposal and cleaning, and its environmental friendliness.
Water: Water alone was the original quenching medium used since the earlier stages of
steel hardening because it was readily available, simple, and inexpensive. In many cases,
water was simply disposed of after use. Cold water provides a faster cooling rate than
most of the other options and might be particularly beneficial for quenching of lower-
hardenability steels.
There are some concerns with respect to using water as a quenchant. The water quality
(soft vs. hard water) may produce a noticeably different film-boiling behavior resulting in
inconsistency in the cooling rate. The heat transfer during the vapor stage is quite erratic,
302 Handbook of Induction Heating

potentially resulting in low consistency. The unpredictable nature of the vapor blanket
stage may produce nonuniform cooling (unless uniquely designed quench systems with
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

extremely high-pressure and high–flow rate pumps are used), raising concerns related to
process repeatability and potential crack development caused by nonuniform quenching.
Besides, pure water causes the steel to rust. This is the reason why the water also contains a
certain amount of rust inhibitor to prevent the quench system and hardened components from
corroding (rusting). Care must be taken so that this does not adversely affect the cooling rate.
Aqueous salt and aqueous caustic solutions: In order to address the unpredictable behav­
ior of water quenching during the vapor blanket stage, certain types of salt (NaCl) and
caustic (NaOH) solutions have been used. These solutions disrupt and may even poten­
tially eliminate the vapor barrier phase proceeding to the boiling phase, improving stabil­
ity and consistency of cooling compared to using pure water. Brine quenching provides
the most intense quenching (assuming temperature and agitation rate are the same). As
expected, these types of solutions are more expensive, more corrosive for equipment,
and not environmentally friendly for disposal. Thus, closed-loop circulating systems are
required to contain the quenchant and to provide a special means of cleaning the work­
piece. Nevertheless, brine might be beneficial when hardening low-hardenability steels.
Oil-based quenchants: Quenching oils (petroleum-derived and vegetable oil–based quen­
chants) have been used in a variety of applications to provide a slow-to-moderate cooling
rate depending on the type of oil, its viscosity, and temperature [170,341]. Oils usually
provide more uniform quench than water or brine type of solutions, often resulting in
low distortion. However, the performance of oil quenchants and their quench severity are
dramatically affected by compositional variation and their cooling properties change with
time, demanding accurate monitoring of the oxidation, aging, and cleanliness.
Quench oils raise concerns not only with respect to their disposal but also with regard
to combustion, flammability, and fire hazards.
In induction hardening, quench oils are typically used with dunk/immersion quench­
ing only. In some cases, a flame will be seen and quickly extinguished during workpiece
immersion. Smoke is frequently produced at that time, requiring an appropriate exhaust
ventilation system. Oil-based quenchants also require a closed system to contain the oil
and special techniques to clean the parts before their transfer to the next operation.
Care must be taken with oil baths used at elevated temperatures in order to prevent con­
taminations with water in the oil, which can cause splashing or, at times, even an explo­
sion owing to the buildup of steam in an enclosed area. It has been reported in Ref. [343]
that even the presence of as little as 0.1% water in the oil may be unsafe.
In a high-production environment, there is the danger of quench oils accumulating the
heat from hot workpieces and exceeding the temperature of the flash point. Therefore,
means must be adapted for their adequate cooling.
Concerns about fires, the necessity of removing oil drag out and certain environmental
restrictions are obvious drawbacks to using oils and oil-based quenchants. The use of veg­
etable oils (including canola, soybean, corn, cottonseed, and sunflower oils) is associated
with lesser environmental restrictions compared to petroleum oil–based quenchants [170].
Still, their application in induction hardening is limited.
Quench oils are not recommended with spray quenching because of the fire hazard.
Quench oils provide lower cooling rates and can only be used with appropriate steels in
order to reduce distortion. However, because of environmental and safety concerns, they
are not commonly used in induction hardening.
Aqueous polymer solutions: Polymer quenchants have been used in industry for several
decades. Their quenching performances range from oil-based quenchants to near the
Heat Treatment by Induction 303

severity of water but with better quenching consistency and stability. Polymer concentra­
tion, temperatures, and the rate of agitation are some of the main factors in determining
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

the cooling rate. These parameters must be regularly monitored and closely controlled.
Although new polymer solutions appear quite regularly, it is appropriate at this point to
mention several basic types of polymer quenchants [162,170–175]:

• PAG—polyalkylene glycol
• PEO—polyethyl oxazoline
• PVA—polyvinyl alcohol
• ACR—sodium polyacrylate
• PVP—polyvinyl pyrrolidone

PAG and PEO are the most commonly used in induction hardening. Depending on pro­
cess specifics, polymer concentration and temperature typically vary within the following
range: 2%–25% and 25°C (80°F)–45°C (110°F), respectively.
Because of “drag out” of the quench and evaporation of the water over time, it is nec­
essary to replenish the quenchant to maintain the required concentration. This requires
monitoring the percentage, conditions of bacteria, temperature, purity, and so on.
Special care may also be needed to prevent foaming of the quenchant, which would
result in an unpredictable and greatly reduced cooling intensity. Varieties of antifoam
agents exist and can be used to prevent the formation of foam. Some polymer solu­
tions may form a jelly-like residue, which can contaminate quench system and obstruct
orifices.
Water mist, fog, forced air, and gas quenching: For high-hardenability steels (i.e., air harden­
able steels) and parts with thin sections, forced air or gas quenching may be used in order
to reduce the cooling severity. For example, in contrast to the great majority of induction
hardening applications, when hardening of high-carbon steel rails for railways, because
of the specifics of the process requirements, after hardening, it is not permitted to have
any martensite in the as-hardened structure. Depending on the selected steel grades, pre­
dominately pearlitic structures with a small amount of ferrites might be required after
induction hardening [340]. Liquid coolants of any kind are not used in those applications,
since it is imperative to “cross” the CCT-cooling curve in the higher transformation region.
Water–air mist, compressed air, or hot steam is used as quenchants in those applications.
The high probability of heterogeneity of the airflow and the specifics of the workpiece
surface conditions (e.g., the presence of oil, variations in surface roughness, etc.) raise con­
cerns with respect to excessive variations in the localized cooling rates. This is particularly
so when dealing with complex geometries.
The use of gaseous quenching might also be required to satisfy process specifications
as well as conditions related to surface appearance (e.g., bright annealing). In these appli­
cations, the use of gas quenching, which creates protective or inert atmospheres, may be
required (see Section 2.1.4).
Press or die quenching: In special cases where distortion must be minimized, a component
may be held in position by a press or die that is water or oil cooled in order to reduce the
temperature of the part while maintaining the part geometry as near to its original shape
as possible. If required, preferential cooling rates at different regions of the workpiece can
be achieved. Since parts must be handled individually, this process can be quite expensive
and time-consuming [342]. Typically, this technology is used to heat treat complex geom­
etries with stringent geometrical requirements (e.g., spiral bevel gears).
304 Handbook of Induction Heating

4.2.7.3 Quench Methods
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

Immersion or dunk quenching: In some applications after proper austenitization, a workpiece


may be simply dunked, dropped, or lowered into a quench medium that is static or gently
agitated (Figure 4.120). The parts remain in the quench medium for a sufficient time to
lower the temperature and are then removed and washed or blown off with an “air-knife,”
industrial fans, or other apparatus to retain the liquid quenchant in the tank. Immersion
quenching is the second most popular quench method used with induction hardening. It
is imperative to have an appropriately sized reservoir, agitation, temperature control, and
water replenish when using immersion quenching.
Liquid spray quenching specifics: Spray quenching is the most popular quenching method
used in induction hardening and was mentioned above on several occasions (e.g., Section
4.2.7.1 and comments for Figures 4.55 through 4.58 and 4.60). This process relies on a stream
of droplets or liquid denoting a number of technologies associated with the heat removal
via the impingement of the liquid quenchants on the austenitized workpiece surface [175].
Cooling severity for a particular liquid quenchant can be adjusted within a wide range by
changing the temperature, flow, pressure, and impingement angle [467].
In induction hardening of cylindrically shaped parts (e.g., axle shafts, spindles, rods,
etc.), spray quenching works best if the component is rotated. This ensures cooling uni­
formity. By rotating parts, the workpiece essentially experiences a constant impingement
rather than many small impingements. Uneven quenching typically has a noticeable nega­
tive effect on microstructure and could result in excessive distortion and cracking.
It should be understood that although high rotational speed can provide more uniform
circumferential heating, it may result in the deflection of the quenchant. This is particu­
larly so when hardening irregular geometry components (e.g., gears, sprockets, or splines),
causing low hardness readings. Therefore, the selection of the rotation speed should be a
reasonable compromise.
Quenchants used in spray quenching include water and aqueous polymer solutions. Because
of safety concerns, the use of oil quenchants with spray quenching is not recommended.
There is a common misunderstanding regarding the ability to apply the widely published,
classical cooling curves to induction hardening applications using spray quenching [175,240].
Classical cooling curves quantifying three stages of quenching—vapor blanket (Stage A),
nucleate boiling (Stage B), and convective cooling (Stage C)—cannot be applied directly to
spray quenching, because the great majority of those curves are obtained using the immersion
cooling technique. The differences are both quantitative and qualitative, and include, but are
not limited to, the specifics of film formation and heat transfer through the vapor blanket dur­
ing the initial quenching stage (Stage A), as well the kinetics of formation, growth, and removal
of bubbles from the surface of the heated component during the nucleate boiling (Stage B).
Because of the nature of spray quenching, Stages A and B are greatly suppressed in time,
while cooling during the convection stage (Stage C) is noticeably more intense, represent­
ing the forced heat convection mode compared with the process represented by classical
cooling curves.
Also, the thickness of the vapor blanket film during Stage A is typically much thinner dur­
ing spray quenching than when the part is submerged in a quench tank, and depends on the
flow rate, impingement angle, part rotation, and other characteristics of the quenching system.
This vapor film is unstable and could be quickly ruptured, increasing the cooling intensity.
In addition, the transition between Stages A and B is smoother with spray quenching
than shown by classical cooling curves for immersion quenching. Bubbles are smaller
because they have less time to grow during the nucleate boiling (Stage B). Much larger
Heat Treatment by Induction 305

numbers of bubbles form during spray quenching and the intensity with which they
remove the heat from the surface of the component is substantially greater compared with
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

immersion cooling [240,344]. Spray quenching intensifies cooling during the boiling (Stage
B) and convective heat transfer (Stage C) stages.
Consequently, classical cooling curves for immersion quenching are of only limited
value in induction hardening. The more intense cooling severity associated with spray
quenching provides higher hardnesses and higher compressive surface stresses, com­
pared with conventional immersion quenching.
The intensity of spray quenching (intensity of heat removal) depends mainly on the quen­
chant flow rate, the angle at which the quenchant strikes the workpiece, and the temperature,
purity, and type of quenchant, as well as the temperature of the part, geometry, the physical
properties of the steel, surface conditions (rough surface vs. polished surface), and the presence
of the “cold sink” effect. This later factor has a considerable effect on quench severity in surface
hardening and selective hardening applications and has been discussed earlier.
Both radial and longitudinal “cold sink” effects occur in induction scan hardening
applications (Figures 4.60 and 4.61). For example, in many induction surface hardening
applications, the core temperature does not rise significantly. A cold core complements
spray quenching by further increasing the cooling intensity at the surface and subsurface
regions of the part. Figure 4.122 shows the results of computer modeling of a medium-
carbon steel solid cylinder (50 mm/2 in. diameter) in a normalized condition using 16 kHz.
The required nominal case depth is 2.5 mm.
The dynamics of the induction heating and cooling during spray quenching reveal the
following phenomena. With induction hardening of carbon steels, the quenching typically
begins immediately after the required time at austenitizing temperature is reached or after a
short dwell (quenches delay). As shown in Figure 4.122, after 3.5 s of heating, the surface and
subsurface layer (where a nominal case depth of 2.5 mm is required) reaches the needed final
temperature for austenitization taking into consideration the nonequilibrium phase transfor­
mation associated with rapid heating. The core temperature does not rise significantly at the
end of heating being below 100°C. A short dwell (0.5 s) is applied before quenching, assisting
the completion of austenitization. Then, the spray quenching begins.

Heating Quenching Soaking


1000
900 Dwell 16 kHz
Surface
800
700
Temperature, ºC

2.5 mm below
600 Average
10 mm below
500
400 Core

300
200
100
0
0 2 4 6 8 10 12 14 16
Time, s

FIGURE 4.122
Results of computer modeling of medium-carbon steel solid cylinder (50 mm/2 in. diameter) in normalized
condition using 16 kHz. Required case depth is 2.5 mm.
306 Handbook of Induction Heating

In the first stage of quenching, the high temperature of the workpiece surface layer
begins to lessen. Figure 4.122 shows that after 2 s of quenching, the surface temperature is
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

reduced by more than 600°C. This results in a workpiece surface temperature of approxi­
mately 215°C. At this point, the maximum temperature will be located at a distance of
approximately 5 mm inward from the surface. Note that, at this stage, two heat transfer
phenomena take place simultaneously: the cooling of surface layers as discussed above
and the heating effect of the core as a result of thermal conduction.
After 6 s of quenching, the surface temperature will decrease to approximately 90°C. At
the same time, the core is still quite warm, being at approximately 310°C.
In some cases, heat treaters do not cool the part completely. After the workpiece is
unloaded from the induction coil, it is kept for some time on the shop floor. During this
time, the heat of the warm core soaks toward the surface. In time, the temperature dis­
tribution within the part will equalize. In this case, the remaining heat is used for some
temper back, which increases ductility and toughness.
When higher frequencies and shorter heating times are used for larger parts (greater diam­
eters or thicknesses), the cold core effect is more pronounced, which increases the overall cool­
ing severity. Note that in some induction surface hardening applications that require shallower
case depths (0.5 to 1.25 mm), self-quenching can be used. Here, the effect of thermal conduction
from the cold core may provide a cooling intensity that is sufficient enough to miss the nose of
the continuous cooling curve. This technique (also called “mass” quenching) allows a shallow
hardness case to be obtained without the use of a liquid quenchant.
As stated earlier, spray quenching works best if the workpiece is rotated during quench­
ing. When rotating irregularly shaped parts such as gears, the speed of the rotation is
slowed down for quenching; this reduction in speed is required to evenly quench the
root of the gear. In contrast, if the rotation is too fast, the quench fluid might not be able
to provide proper quenching in certain areas of the tooth (i.e., root region), resulting in an
unacceptable hardness pattern deviation.
The ability to provide identical impingement is one of considerations when spray
quenching complex geometries. The design of orifices should prevent steam pockets from
developing on the surface being quenched.
When using a single-shot inductor (e.g., Figure 4.82), it is preferable to quench from at
least two sides of the workpiece maximizing the quench coverage.
To maintain quench pressure at the point of impingement, the quench assembly should
be sufficiently close to the workpiece surface. Depending on the application, the “workpiece-­
to-quenching device” gap may be within 0.25 to 1.5 in. Too large a gap (greater than 2 in.) is
undesirable, because of the reduction of quenchant velocity and worsening of the impinge­
ment angle. This is particularly true in horizontally processed parts. On the other hand,
too short a distance (i.e., 2 mm) is not desirable as well, because it can result in a spotted
hardness pattern, particularly when static hardening is used.
In most cases of induction hardening of plain carbon and low-alloy steels (in particular
steels with low carbon content), it is important to move into the quenching position as
quickly as possible, before the part’s surface temperature cools too much (unless interme­
diate soaking is required).

4.2.7.4 Part Cleaning before Heating and after Quenching


Residues of fluids used during previous manufacturing steps (e.g., the machining) may
remain on the part surface. Before arrival of those parts at the induction hardening station,
those residues should be washed or removed from the part before the austenitizing. If left
Heat Treatment by Induction 307

on the part, the fluids may cause smoke and flames during the heating cycle and also may
contaminate the quenchant, reducing its ability to adequately cool the part. In some cases,
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

parts should also be cleaned before tempering since the heat of tempering may cause the
remaining quenchant to bake on the part, making it difficult to clean or remove during
subsequent operations.
Although some of the polymer quenches may contain rust inhibitors, tempering without
removing them can cause the parts to be more prone to rusting at a later date. This fact
and the typical odors given off during heating along with the difficulty of removing the
coating after tempering would lead to the conclusion that removal of the quenchant before
tempering might be the best course.
That being said, it should be remembered that parts that have been cleaned will be sus­
ceptible to rusting shortly after removal from the tempering furnace. Therefore, it would
be judicious to apply some type of rust preventive soon after the parts have cooled to room
temperature.

4.2.7.5 Quench Systems Design and Controls


Because of the variation in cooling rate with quench concentration, temperature, contamina­
tion, and agitation, closed systems are required to contain and maintain the quench medium
in the best possible condition, providing repeatable results every time a part is quenched.
A typical quench system would include a holding tank, a pump, heater, heat exchanger,
filter, and assorted valves to turn the quench on and off at the appropriate time. Unless
manual monitoring of the quench purity and concentration is carried out, the system may
also contain equipment to monitor the concentration of the quench media and check/alarm
for contaminants. Because of the tendency of some of the quench media to be drawn out
with the part and the water to evaporate during the cooling process, it may be necessary
to have an automatic system to dispense water and quench media to the tank to maintain
the desired concentration level.
On immersion systems, it is necessary to ensure that whatever agitation that is used in
the tank, it is uniform over the surface of the part being quenched since the cooling rate is
largely dependent on proper agitation. In some instances (e.g., quenching large gears and
sprockets), a spray quench may complement immersion quenching while the part is being
rotated (Figure 4.120).
It is also quite important on spray quenching systems to ensure that hoses feed to the
inductor from a point below the inductor in order to prevent siphoning water out of the
lines at the end of the quenching. If the lines feed from below the inductor, they remain full
and the timing of the quench arrival is the same as each part goes through the quenching.
Depending on the size of the part and inductor configuration, the required flow rate
of quench should be estimated and the pump and motor are selected based on typical
curves provided by the manufacturer. Care must be taken in the system design to include
all expected head loss through pipes and hoses to ensure that the pump and motor are
adequate to provide the required flow at the desired quenching pressure.
Because the temperature of the quench will increase as a result of cooling the hot part, it is
necessary to have some type of heat exchanger to maintain the temperature of the quenchant
within an acceptable range. The heat that is delivered to the workpiece during the heating
process must be removed from the part during the quenching cycle. In addition to this heat
load, there is the heat that is added to the quench during the process of pumping the quench
to and through the induction heating coil/quench ring and back to the tank.
308 Handbook of Induction Heating

In order to maintain the quenchant at the desired temperature, it is necessary to have both
a heater and a heat exchanger (generally water-to-water or air-to-water). It is essential that a sys­
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

tem be at the required temperature before any parts are processed. When a system “sits” for a
long period, the quenchant can cool to room temperature, which may be below the specified
temperature for processing the parts. If a system is running without processing parts, heat is
still being added to the quench media by the mechanical work of the pump. Thus, it may be
necessary to cool the quenchant even when the system is in the standby mode waiting to pro­
cess components. During the processing of parts, the heat load for the heat exchanger is much
larger and the system must be sized to handle the worst case heat load.
A well-designed system is equipped with controls to indicate when the quenchant is
at the proper temperature and concentration and to allow system operation only if the
parameters are within the correct range. Additional controls are utilized to ensure that the
pressure and flow of the quenchant is within the desired range. On some systems using
signature monitoring, all of these parameters may be recorded throughout the processing
of each part as a quality control measure and additional aid in troubleshooting problems
that may arise with a part in the production process or in the field.
Because of the need to maintain all induction hardening parameters within a relatively nar­
row band of acceptable limits, it is necessary to have a variety of components, cross-checks,
and safeguards within an induction system. In the area of quenching, many parameters must
be monitored to ensure that they remain within an acceptable range and that processing of
parts is prevented when any of the parameters are outside of the desired range.
Figure 4.123 shows a pictorial drawing of a typical quench system for IHT. Obviously,
every system may not require all of the components but a typical system might contain
components as shown.
An in-line quench filter is often used to ensure that particles of dirt, debris, metal chips,
and so on are removed from the quench before it proceeds to the induction heating coil.

Cooling Cooling
water in Quench water out
pump
Solenoid
valve Motor

A B
Heat exchanger W/W A′
B′
Quench
filter
Pressure SW Oil
Centrifugal skimmer
Pressure GA
separator
Manual
shutoff Float-type
Solenoid level control
valve
Induction
Flow heating coil
meter
Quench heater

Quench tank

FIGURE 4.123
Pictorial drawing of a typical quench system for IHT.
Heat Treatment by Induction 309

Small amounts of oxides and some scales that are inevitably generated on the surface of
hardened component (unless a protective atmosphere is used) have a tendency to accumu­
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

late over a period in a quench tank and must be regularly removed.


A centrifugal separator may be used as well to remove contaminants that may cause
degradation of the quench. Cutting fluids, machining oil residue from parts, and an inap­
propriate amount of rust preventives could cause a noticeable deviation from the typical
quenching characteristics, causing a problem. An oil skimmer may be used to help con­
tinuously remove oil residue from the quench tank. It is important to closely follow OEM
guidelines in this respect.
A pressure gauge and pressure switch are used to give a visual indication of the quench
pressure at the cooling manifold and to inhibit operation of the hardening system if the
value is insufficient to provide proper flow.
Flow meters/switches with alarm set points are used in the quench lines to the coil to
ensure that the flow rate delivered to the part is adequate to properly harden the part.
Detecting improper flow during the quenching cycle may indicate a problem with foam­
ing of the quench, lack of pressure, or a variety of other types of problems.
A high-speed solenoid valve is used to accurately turn the quench on and off at the
proper time in the hardening cycle.
During times of inactivity, the temperature of the quench may fall below the value that
is needed to ensure repeatability of the process. A quench heater is provided in the quench
tank in order to raise the temperature to the required value. The electrical control circuit
is required to monitor the temperature of the quench in the tank and turn on the heater or
turn on the cooling water to the heat exchanger if it is necessary to raise or lower the tem­
perature of the quenchant. This is done through the use of solenoid valves in the quench
and cooling water lines. Additional manual shutoff valves are often used for maintenance
on the system to prevent loss of large amounts of quench or cooling water. Mounting the
cooling water solenoid at the inlet rather than at the outlet of the heat exchanger prevents
the loss of large amounts of water in the event that the heat exchanger develops an internal
leak. One sure sign of a leak in the heat exchanger is if the quench tank level begins to rise
unexpectedly above the maximum level.
Another type of control that is typically used on IH quench systems is a float type of
level control to monitor when the quench level is too high or too low. If the level is too high,
the quench may begin to spill over onto the floor and cause a safety hazard to personnel
walking in the area. If the level is too low, it could lead to problems with inadequate cool­
ing of the quench pump and failure of seals in the pump. If the tank size is inadequate,
frequent cycling of the level control switches could cause erratic operation.
Although it may seem that it is not directly related to the quench system, fairly elaborate
mechanisms and electronics are used on induction heating systems in order to detect and
ensure rotation of the part during heating and quenching. This is due to the fact that it is
very important to ensure that the quench is properly impinging on every part of the heated
surface in the same manner. Rotation is used during the cycle to ensure uniform heating and
quenching of the part without variations in the heating or quenching simply because it is near
a certain part of the coil or quench ring that might be slightly different from another area; that
is, if a few quench holes become contaminated or plugged, the effect is less pronounced when
the part rotates during the process rather than relying on a few holes in one section of the coil.
Other types of automatic systems that have been added to induction quench systems
are electronic devices to measure the concentration of the quench media for polymer
quenches. If the system determines that the concentration is too low, then more quenchant
would be added. If the system determines that the concentration level is too high, more
310 Handbook of Induction Heating

water would be added. It goes without saying that although the quenching process may
be relatively simple in concept, certain nuances and safety requirements can lead to fairly
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

complex designs for a quenching system.


It should be mentioned here that there are a few instances when rotation is not used
during the heating or quenching stage of the hardening cycle. One example would be
Inductoheat’s patented SHarP-C Technology for hardening crankshafts, camshafts, and
other components. On these types of systems, “Flow quench” using uniquely designed
quench slots instead of spray quench holes have been successfully used. Advanced design
ensures uniform impingement of the quench media during the quenching phase.

4.2.8 Striping (Striation) Phenomena, Barber-Pole Effect,


and Snakeskin (Soft-Spotting) and Fish-Tail Phenomena
Heat treat practitioners sometimes observe unusual effects in IHT of workpieces with
seemingly smooth surfaces, such as a striping (striation) phenomenon, a barber-pole effect,
and snakeskin and fish-tail phenomena, just to name a few.

4.2.8.1 Striping Phenomena and Barber-Pole Effect


Striping phenomena have been first observed more than 60 years ago, and it may appear
on seemingly smooth surfaces of the workpiece. (Note: An appearance of stripes attributed
to geometrical irregularities of the workpiece [e.g., such as ball screws and components
with threads] will be excluded from a discussion since the cause of uneven heating is obvi­
ous and it is related to the electromagnetic proximity effect.) Since that time, there was a
number of publications devoted to its description, possible causes, and prevention [1,7,20,
54,86,87,347,348,350,468]. Generally speaking, there are three types of striping phenomena:

• Electromagnetic-caused striping (Type A)


• Electromagnetic-caused striping (Type B)
• Quench-related striping (Type C), also referred to as the barber-pole effect.

Two types of striping phenomena are attributed to an electromagnetic origin: Type A


and Type B. These phenomena may appear when a single-shot or static heat mode is used.
Although the appearance of both types is similar, the physics behind each is quite different.
The Type A striping phenomenon appears only when a multiturn coil is used, and can
be observed when heating both ferrous and nonferrous alloys, particularly those that have
relatively poor thermal conductivity.
An undesirable combination of small inductor-to-workpiece gaps, relatively high power
densities, short heat times, and loosely wound coil turns might result in an uneven stripe-
type temperature pattern (also referred to as alternating bright and dark rings or stria­
tions) along the length of the workpiece with a seemingly smooth surface, such as the case
of straight cylinder shafts.
Stripes are caused by localized power surpluses owing to electromagnetic coupling between
particular coil turns and located in the close proximity workpiece areas. The number of stripes
is directly related to the number of coil turns largely as a result of the electromagnetic proxim­
ity effect. If the heat time is sufficiently long, the thermal conductivity equalizes the thermal
gradients, making stripes disappear. The appearance of this phenomenon can easily be pre­
dicted using computer modeling. Proper coil design, selection of process parameters, and rota­
tion of the heated workpiece can eliminate this undesirable phenomenon.
Heat Treatment by Induction 311

The Type B striping phenomenon (Figure 4.124) usually occurs only during intensive
induction heating of ferrous alloys, including carbon steels where short heat times and
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

high power densities are used. Multiple stripes can be observed by the naked eye on the
surface of a heated workpiece even in the case when a single-turn inductor is used.
Because of Type B striping, the workpiece area under the coil may unexpectedly start
to heat nonuniformly. Shortly after the heating cycle begins, alternating hot bright areas
(bright stripes) and cold areas (dark stripes) become visible. When dealing with cylinder
workpieces, these bright and dark stripes encircle the cylinder, and, thus, have shapes of
rings. In contrast to Type A striping, the number of rings with Type B striping is not nec­
essarily related to the number of coil turns. For example, Figure 4.124a reveals that mul­
tiple stripes appeared when heating a steel bar utilizing a conventionally wound four-turn
cylindrical coil (frequency is 10 kHz). Figure 4.124b shows numerous stripes appearing
while heating a wide strip using a single-turn rectangular inductor.
Type B striping has been relatively seldom viewed in practical applications or in labora­
tory experiments and is considered by many as a mysterious phenomenon. In some appli­
cations, striping suddenly occurs and then disappears using practically identical process
recipe and steel grade. There were a number of attempts to explain this phenomenon.
Some researchers also tried to recreate conditions for its appearance using computer mod­
eling. However, assumptions and dramatic simplifications made in those studies did not
result in convincing explanations.
There is no single explanation for this phenomenon. M. Lozinskii attempted to explain it in
the 1960s [7]. Assume that a magnetic steel cylinder is located inside a cylindrical inductor. As
a result of the EMF produced by the induction coil, eddy currents will flow within the work­
piece. Because of the skin effect, these eddy currents appear primarily in the surface layer of
the cylinder located inside the coil, causing an increase of its surface temperature.
In reality, any material has certain microscopic and macroscopic defects, impurities,
structural heterogeneities, geometrical deviations, and different degrees of chemical seg­
regation. As a result, different surface regions of the workpiece are heated slightly dif­
ferently. Some reach the Curie temperature first and lose their magnetic properties. The
relative magnetic permeability of these areas dramatically drop to unity (μr = 1). This leads
to a significant increase in the δ. The resistances of these nonmagnetic regions drastically
decrease, creating low-resistance paths compared with neighboring surface areas that
retain their magnetic properties. As a result, the density of the induced currents in the
low-resistance regions will increase.
Eddy currents induced in areas that retain their magnetic properties (dark rings) have
a tendency to “rush” to complete their loops through the lower-resistance paths (bright

(a) (b)

FIGURE 4.124
Type B striping phenomenon occurs only during induction heating of ferrous alloys. Multiple strips appeared
when heating steel bar utilizing a conventionally wound four-turn cylinder coil (a). Numerous strips appeared
while heating a wide strip using a single-turn rectangular inductor (b).
312 Handbook of Induction Heating

rings). This current redistribution leads to a further heat source reduction in the magnetic
areas at a low temperature (dark rings) and appears as additional heat sources in the low-
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

resistance nonmagnetic areas at high temperature (bright rings). Therefore, a chain reac­
tion somewhat similar to positive feedback or “snow ball” occurs. As a result, it is possible
to observe with the naked eye a mixture of ring-shaped stripes (Figure 4.124) on the sur­
face of cylinder or rectangular shape. Hot bright stripes alternate with the relatively cold
dark stripes. Experience shows that usually the thickness of the bright and dark stripes
equals one to three current penetration depths in hot steel.
Besides the current redistribution, Type B striping potentially might also be the result
of several other electromagnetic and heat transfer phenomena, including the electromag­
netic edge effect of joined materials having different properties. This effect occurs when
conductors with different physical properties are located in a common magnetic field as
discussed in Section 3.1.7.4 (Figure 3.46).
Experience shows that striping can appear in several different ways. In some cases, very
narrow bright stripes (rings) appear at the beginning of the heating cycle. Over time, the
narrow stripes may widen and the peak temperatures may move from the center of each
ring toward the edges of each bright hot ring. During the heating, the stripes can some­
times move back and forth along the workpiece surface area, climaxing a mystery of this
phenomenon. With longer heating cycles, the striping phenomenon usually disappears.
In our opinion, the occurrence of the stripes depends on a complex function of the fre­
quency; magnetic field intensity; and thermal, electrical, and magnetic properties and
microstructure of the steel. However, as mentioned above (if it occurs), the needed condi­
tion appears to be an application of relatively high power density. If the power density
is relatively low, the temperature will equalize between the neighboring bright (high-­
temperature) and dark (low-temperature) rings because of the thermal conductivity of the
steel. It appears that steels with substantially heterogeneous structures as well as so-called
dirty steels have a greater chance of exhibiting striping. Good quality clean steels with a
homogeneous structure have a lower chance of producing this phenomenon, though there
were claims that this phenomenon appeared even when heating ultralow-carbon steels.
As stated above, there are several different types of striping phenomena. Section 6.9.2.1
discusses a striping phenomenon with the stripes/striations occurring in strip-coating
applications such as galvanizing and tin reflow that are typically longitudinally oriented.
The nature and causes of such striping phenomenon are very different compared to the
phenomena discussed above.
The Type C striping phenomenon [1,86,350] is also referred to as a barber-pole effect that
can be observed upon quenching of uniformly heated workpieces using either a single-
turn or multiturn scan inductor with or without part rotation.
Traditionally, the striping phenomenon that appeared after quenching is often called a
quench-striping effect or a barber-pole effect. The cause of its appearance might not relate
to the specifics of heating, but is primarily associated with the characteristics of quench­
ing. This includes

• Part rotation
• Specifics of spray quench flow along the workpiece surface after the spray quench
impinged (struck) its surface
• Scan speed
• Presence of quench interruptions, its nonuniformity or formation of local steam
pockets, and so on
Heat Treatment by Induction 313

At the time of writing this book (2016) and according to our knowledge, the barber-pole effect
has never been obtained by mathematical modeling. It has only been observed. Barber-pole
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

stripes are usually of spiral shapes that could occur on the surface of an as-quenched work­
piece and are typically related to nonuniform cooling, surface oxidation, and scale formation.
In some cases, the barber-pole effect might not be associated with any appreciable
microstructural variations of as-hardened areas and would not affect part performance.
Therefore, in cases when the visual appearance of the surface is not critical, the barber-pole
effect might only be aesthetically unpleasing (Figure 4.125) but has no detrimental effect
on heat-treated part performance.
However, in other instances (Figure 4.126), the barber-pole effect can alter some indus­
trial characteristics of material and the microstructure of the hardened pattern because of
improper quenching or interrupted (or partial) quench flow. Formation of upper transfor­
mation products (including lower and upper bainite and pearlite) within the martensitic
structure or the appearance of undesirable tempered martensite can be attributed to the
barber-pole effect as well, leading to a combination of hard and soft (partially hardened)
striations or rings. Experience shows that improvement in quench flow, its severity, and
uniformity, or applying a quench follower, as well as making improvements in the eccen­
tricity of a rotated part and modifying the impingement of spray strokes, helps avoid the
appearance of the barber-pole effect.
Induction heat treaters may also refer to the barber-pole effect as a phenomenon that has
no association with the appearance of any stripes, rings, or spirals. However, sometimes,
during induction heating of cylinders, instead of stripes, a shifted or squeezed tempera­
ture profile suddenly appears at the end of the hardened pattern instead of a straight heat­
ing pattern. This can take place when using a single-turn or multiturn scan inductor and
it does not appear to be related to the helix of a multiturn coil winding. This phenomenon

FIGURE 4.125
Appearance of the barber-pole effect.

Longitudinal direction

Tempered martensite
Bainite

FIGURE 4.126
Formation of upper transformation products within the as-quenched martensitic structure or the appearance
of undesirable tempered martensite can be attributed to the barber-pole effect as well. (Courtesy of Michael
O’Brien, BSc [Met], MIEAust, CPEng, Principal Consultant Metallurgist.)
314 Handbook of Induction Heating

is usually quite unstable, and when the next part is heated, this type of barber-pole effect
could disappear and may never be seen again. If this type of barber-pole effect is steady, it
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

can be eliminated by slightly changing the scan speed, coil power, or part rotation speed.

4.2.8.2 Snake-Skin Effect and Soft-Spotting Phenomena


The snake-skin phenomenon represents one of the variations of soft spotting. It appears
as alternating soft and hard spots within the as-quenched structure (looking somewhat
similar to the skin of the snake, Figure 4.127) and can often be seen by the naked eye
on the etched or even unetched workpiece surface. Soft spots represent regions where
upper transformation products were formed. This phenomenon usually takes place when
a static heating mode is used in cases of small coil-to-workpiece gaps in combination with
sparsely located spray nozzles/orifices.
To improve coil electrical efficiency, inexperienced inductor builders who lack adequate
theoretical knowledge might design MIQ inductors having a substantially small coil-to-
workpiece air gap. It is important to remember that induction hardening is a two-part
process: heating and quenching.
Lower-than-expected hardness readings (soft spots) can occur because of insufficient
quenching or trapped hot quench pockets. Properly designed MIQ inductors or use of
quenching blocks with slightly enlarged quench holes-to-workpiece gaps eliminate this
undesirable phenomenon.
Different variations of the soft-spotting phenomenon are observed when heating complex­-
shaped parts with an interrupted quench, or if quench flow is deflected or unintentionally
blocked because of geometrical complexity. The presence of steam pockets can also result
in the appearance of the soft-spotting effect.
All components of an IHT system, including tooling and fixtures, should be reviewed
when soft spotting occurs. For example, worn bearings can result in the part wobbling
during rotation. The fact that the part is rotating is sometimes deceiving, because it creates
an illusion that the workpiece rotation automatically provides the required uniformity of
both heating and quenching. However, it should be recognized that worn bearings can
lead to a situation where, regardless of part rotation, certain workpiece regions can always
be positioned closer to the induction coil, and during heating, those areas will experience
more intense heating. In addition, the same areas of the part will also be located closer to
the quenching device, often resulting in more intense quenching. In contrast, if the part
wobbling is measurable, opposite regions of the workpiece could experience appreciably
less intensive heating and a noticeably milder quench severity. Both factors potentially can
result in the appearance of surface soft-spotting and reduced hardness case depth.
When hardening gears and splines, it is important to appreciate a “quench-bouncing” phe­
nomenon where teeth of gears and splines can function as paddles bouncing off the quenching
strikes, resulting in nonuniform quenching and soft spots in flank and root regions.

FIGURE 4.127
Appearance of the snake-skin phenomenon.
Heat Treatment by Induction 315

Eliminating soft spotting requires redesigning the spray-quench system by taking into
consideration the geometrical subtleties of the particular component. Part rotation should
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

be smooth without any appreciable wobbling and excessive rotation speed.

4.2.8.3 Fish-Tail Effect (Field-Fringing Phenomenon)


Conventionally designed single-turn inductors and some multiturn coils have areas where
there is an inevitable distortion of the magnetic field, which could lead to the appearance
of regions with a heat deficit and where undesirable microstructures can be formed. One
region is related to an area where copper buses that transmit electrical current from a
power source are connected to an induction coil. These connectors sometimes have the
shape of a fish tail, and the region is often called a fish-tail (also referred to as field fring­
ing) region of the inductor. Figure 4.128 shows fish tail–shaped connection copper buses
used in the fabrication of a shaft hardening single-turn inductor (left) and multiturn strip
heating coil (right).
Figure 4.129a illustrates the location of the field-fringing effect and the instantaneous
orientation of the electrical current [350]. A magnetic field is formed by incoming and out­
going electrical currents oriented in opposite directions at the areas where the connection
buses (coil terminals) join the inductor copper.

(a) (b)

FIGURE 4.128
Fish tail–shaped connection copper buses used in the fabrication of a shaft hardening single-turn inductor (a)
and multiturn strip heating coil (b). (From V. Rudnev, Metallurgical insights for induction heat treaters, Part 7:
Barber-pole, snakeskin and fish-tale phenomena, Heat Treating Progress, ASM Int’l, Materials Park, OH, May/
June, 2009, pp. 15–18.)

Electrical current
Inductor Inductor
Compensated
Inductor “Fish-tail” Compensated
“fish-tail”
fringing “fish-tail”
effect
effect effect

Inductor Inductor
Inductor Shaft
Shaft Shaft terminals terminals
terminals

(a) (b) (c)

FIGURE 4.129
Illustration of the fish-tail (field fringing) effect and its compensation. (a) Illustration of the magnetic field fring­
ing effect; (b, c) illustration of two possible ways to compensate the field fringing effect. (From V. Rudnev,
Metallurgical insights for induction heat treaters, Part 7: Barber-pole, snakeskin and fish-tale phenomena, Heat
Treating Progress, ASM Int’l, Materials Park, OH, May/June, 2009, pp. 15–18.)
316 Handbook of Induction Heating
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

FIGURE 4.130
Inductor used in Inductoheat’s CrankPro machine implementing patented nonrotational crankshaft and cam­
shaft surface hardening technology.

This results in magnetic field distortion that leads to a heat deficit within the corre­
sponding workpiece region and is often referred to as the fish-tail effect (or flux-fringing
effect). One of the typical solutions to compensate for this fish-tail effect is to rotate the part
during heating, ensuring that all regions of the workpiece absorb the same energy during
the entire process cycle.
In recent years, some IH manufacturers have developed and patented a variety of
advanced means to effectively control magnetic flux [38,351,352], thus providing compen­
sation for the magnetic field fringing at the inductor terminals. To illustrate, Figure 4.129b
and c shows two simplified approaches that allow compensation for the fish-tail effect and
eliminate the need to rotate the heated workpiece. In both cases, improved electromag­
netic coupling (proximity effect) at the fish-tail region of the inductor compensates for the
magnetic field fringing at that location.
The fish-tail effect was one of the challenges solved during development of an innova­
tive nonrotational SHarP-C hardening technology. Figure 4.130 shows the inductor used
in Inductoheat’s CrankPro machine (Figure 2.5), implementing patented nonrotational
crankshaft and camshaft surface hardening technology [1,351,352], which is discussed in
Sections 4.9.4 and 4.9.5. The fish-tail effect was compensated for and effectively controlled
by using patented coil design innovations that can be used in other IHT applications, par­
ticularly where single-turn inductors are applied.

4.3 Holes, Keyways, Grooves, Undercuts,


and Other Geometrical Irregularities
Certain design features of components may result in eddy current flow diversion, poten­
tially resulting in appreciable heat nonuniformity. Typical examples are parts containing
longitudinal, transverse or cross-holes, keyways, grooves, shoulders, flanges, diameter
changes, undercuts, splines, sharp corners, and others (Figure 4.131) [142,281,349]. However,
such features are not unique as they are commonly found on many transmission and
engine components (Figures 4.38 and 4.52).
In such cases, induction practitioners could face certain challenges. The presence of geometri­
cal discontinuities distorts the eddy current flow and can result in the undesirable appearance
of hot and cold spots, excessive shape distortion, grain boundary liquation, and even crack­
ing. As an example, Figure 4.132 shows cracks that were initiated at transverse (radial) and
Heat Treatment by Induction 317
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

FIGURE 4.131
Examples of parts containing various geometrical irregularities.

FIGURE 4.132
Examples of cracks that were initiated at transverse (radial) and longitudinal (axial) holes.

longitudinal (axial) holes. Severe overheating is the most frequent cause of cracking and can
lead to unwanted metallurgical microstructures (Figure 4.36), which weakens grain structure
and substantially increases brittleness and sensitivity to intergranular cracking (Figure 4.39).
Therefore, it is important to carefully evaluate the eddy current flow before choosing a
process recipe and system design.

4.3.1 Longitudinal (Axial) Holes and Longitudinally Oriented Hollow Areas


The presence of longitudinal (axial) holes or hollow areas within the part can cause a redis­
tribution of eddy current flow, resulting in overheating of certain regions. Figure 4.133a

Overheated areas
Part surface

Hole B

Eddy A C
current
D

Part
Current penetration Segment of the part Current penetration
(a) (b)

FIGURE 4.133
Redistribution of eddy current flow owing to the presence of longitudinal (axial) holes within the part (a) and
sketch of relative positioning of axial holes with respect to eddy current penetration depth (b).
318 Handbook of Induction Heating

shows a transverse segment of a shaft with the normal and partially blocked current flow
within it.
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

Normal current flow is typical if the heated shaft is solid, or if a hole is located signifi­
cantly below the current penetration depth (δ); therefore, the induced eddy current does
not interact with the hole. However, if a longitudinal hole is located within the current
penetration depth, it partially blocks the normal eddy current path and squeezes current
flow into a channel between the part surface and the hole, and that may lead to a dramati­
cally increased current density and severe local overheating of the hole areas, resulting in
damage of the steel grain structure.
There are two factors that can cause the overheating:

• An increase of heat source generation owing to the redistribution of eddy current flow
• The lack of an adjacent mass of metal there, meaning that there will be reduced
heat transfer from the hot surface toward the colder core as a result of the thermal
conduction (cold sink effect)

Depending on workpiece geometry, the impact of both factors may be different. If a lon­
gitudinal hole is located within δ from the surface, then the first factor typically prevails
and is primarily responsible for the heat surplus in that area (Figure 4.133b, hole A). When
the hole is located within 1 to 2 times δ (as with holes B and C in Figure 4.133b), then both
factors have approximately the same impact on the localized heat surplus. Figure 4.135a
illustrates a “bleeding” of hardness pattern in the subsurface hole. Reduction of the heat
time or time transfer into the quench device as well as an increase in applied frequency
may help minimize a thermal effect of subsurface longitudinal holes.
If the longitudinal hole is located within 2 to 3 times δ and the heat cycle is relatively
long (8–12 s), the second factor (the lack of adjacent mass) typically makes greater contribu­
tion to the temperature surplus. When the hole is located sufficiently far away from the
current-carrying surface (hole D), the probability of overheating is very slim.

4.3.2 Transverse (Radial) Holes


Transverse holes always cause a redistribution of the eddy current flow and the matter
is to what degree. Unlike the case of longitudinal holes, eddy current redistribution attrib­
uted to transverse holes can result in both underheating and heat surplus of the hole edges
(Figure 4.134a). Because of the current concentration, overheating can occur at the hole edges
parallel to the eddy current flow. At the same time, there will be a certain lack of heat at the
hole edges perpendicular to the eddy current flow. If the hole diameter is less than one-half
of δ, then current flow distortion does not typically result in significant temperature nonuni­
formity around the hole. Lower frequencies and power densities typically reduce overheating.
An increase in the hole diameter makes overheating more pronounced. These heat nonunifor­
mities can cause grain boundary liquation and complex distribution of transient and residual
stresses, potentially initiating cracking in the vicinity of the hole edges.
It is possible to obtain a relatively uniform temperature distribution along the perimeter
of a medium-sized hole by putting metallic or nonmetallic plugs into the hole. If the plug is
made of the same alloy as the workpiece, the heat nonuniformities will be negligible and the
temperature distribution can be considered uniform, assuming it has a sufficiently tight plug-
to-metal contact (Figure 4.134b). Despite their capability in avoiding over- and underheating at
the hole edges, steel plugs present challenges because it is difficult to insert them and remove
them later after the heat treatment because they may become welded in the hole.
Heat Treatment by Induction 319
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

Part

Eddy
current

Overheat
Hole
Underheat
(a)

Carbon steel
plug

Uniform
temperature
profile
(b)

Underheat

Overheat

Copper
plug
(c)

Holes

Hot spots

(d)

FIGURE 4.134
Transverse holes always cause a redistribution of eddy current flow. (a) Transverse hole (no plug); (b) carbon
steel part and carbon steel plug; (c) carbon steel part and copper plug; (d) multi-hole part (no plugs). (From V.
Rudnev, R. Cook, J. LaMonte, Induction heat treating: Keyways and holes, Metal Heat Treating, March–April
1996, pp. 83–86.)
320 Handbook of Induction Heating
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

(a) (b)

FIGURE 4.135
(a) “Bleeding” of hardness pattern into the subsurface hole. (b) Distortion of the hardness patterns owing to the
use of copper plugs. (From V. Rudnev, R. Cook, J. LaMonte, Induction heat treating: Keyways and holes, Metal
Heat Treating, March–April 1996, pp. 83–86.)

An alternative to the use of steel plugs is the use of copper plugs. In this case, a nonuniform
current distribution still occurs and can lead to a combination of localized temperature sur­
plus and heat deficit regions. However, the location of those regions is just an opposite of their
location seen without any plugs. Figure 4.134c shows that the eddy currents will gather in the
copper plug from the neighboring carbon steel regions. This takes place because the electrical
resistivity of the copper is much less (approximately a factor of 10) than the resistivity of any
steel, and it is much easier for the eddy currents to flow through the low-resistance copper than
through the high-resistance steel. Therefore, when copper plugs are used, the hole edges that
are parallel to the eddy current flow are underheated and the hole edges that are perpendicu­
lar to the current flow may have a temperature surplus. The actual distortion of the hardness
patterns attributed to the use of copper plugs is shown in Figure 4.135b.
Besides the eddy current redistribution, copper plugs improve heat uniformity thanks to
higher thermal conductivity. Note that when using copper plugs (Figure 4.134c), the heat
surplus of the hole edges is much less pronounced compared to the case when plugs are
not used (Figure 4.134a). Thus, copper plugs may eliminate the cracking in the hole areas
during induction heating and quenching. Copper is much softer than steel and copper
plugs can be relatively easily drilled out after hardening.
In some applications, instead of metallic plugs, water-soaked wooden plugs have been used.
Because wood is not electrically conductive, the wooden plugs do not change the eddy cur­
rent flow. Therefore, there will still be regions with high and low current densities as shown
in Figure 4.134a. However, the use of water-soaked wooden plugs allows one to decrease the
magnitude of overheating of the hole edges up to a certain point owing to the heat transfer
from the high-temperature regions into the water-soaked wood by thermal conduction.
In addition, when using any plugs, there will be restricted or blocked quenchant flow
into the hole area. Therefore, the probability of cracking is reduced because of the reduc­
tion of the cooling severity at hole edges during quenching.
Inserting plugs into holes and taking them out are very delicate and time-consuming pro­
cesses. A popular attitude among heat treaters who deal with parts containing holes is that if
there is a possibility of avoiding nonuniform heating of the hole edges without using plugs
(e.g., by using a special heating recipe, coil design, or frequency choice), then the extra develop­
ment effort required is justified. Therefore, plugs should be used only as a last resort.

4.3.3 Angled Holes
Angled holes require special attention. The heating within the vicinity of angled hole areas
combines the features of nonuniform temperature distribution that are specific to both:
Heat Treatment by Induction 321

Quench flow
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

Acute Obtuse
angle angle

(a)

Acute
Obtuse angle angle
Straight
hole
(b) (c)

FIGURE 4.136
Hardness pattern deviation in the proximity of the angled hole. (a and b) Illustration of typical hardness pattern
deviation in the proximity of the angled hole and straight hole, respectively; (c) specifics of the quench flow in
case of an angled hole.

longitudinal and transverse holes. A sharper angle of the hole results in more pronounced
overheating and a greater possibility of cracking unless a profiled inductor design that
addresses those features is used.
Besides the temperature nonuniformity along the hole perimeter, there might be a notice­
able heat pattern deviation at the angled hole. This deviation results in a nonuniform case
depth in the hole area surroundings, making a marked difference of the hardness distribu­
tion in obtuse versus acute angles (Figure 4.136a) compared to a straight hole (Figure 4.136b).
Therefore, special inductor profiling might be required to properly address those geo­
metrical features and to obtain more uniform heating as it has been done with SHarP-C
inductors, for example.
It is important to recognize that besides potentially nonuniform heating, the different regions
of angled holes also exhibit various cooling intensities during quenching (Figure 4.136c). With
conventional designs, an acute angle area experiences much more severe cooling intensity,
complementing the heat surplus there. This combination makes the hole region susceptible to
cracking. Experience shows that, in many cases, advanced designs help to dramatically reduce
or even eliminate the propensity of angled holes for cracking.

4.3.4 Other Factors
It should be recognized that temperature surplus alone might not result in crack develop­
ment. There are other factors that can complement overheating, increasing crack sensitiv­
ity (see Section 4.1). Steel chemical composition is one of those factors. Steels having higher
carbon contents are more prone to cracking. Besides carbon content, there are also other
chemical elements that affect crack sensitivity and brittleness; the extent depends on the
amount and combination of elements present. An unfavorable combination of alloying ele­
ments and residual impurities could promote a tendency to cracking.
For example, sulfur and phosphorus amounts should be minimized to reduce steel brittle­
ness and crack sensitivity. Sulfur reacts with iron, producing hard, brittle iron sulfides (FeS)
that concentrate at grain boundaries. FeS also has a relatively low melting temperature, poten­
tially leading to grain boundary liquation and increased sensitivity to intergranular cracking.
FeS in carbon steels is minimized by the addition of manganese to form MnS, which are dis­
tributed within grains rather than at grain boundaries, creating a less brittle microstructure.
A high level of phosphorus can also cause excessive brittleness in both steels and cast irons.
322 Handbook of Induction Heating

Using lower-than-desirable quench temperatures and concentration, higher-than-specified


quench flow rates and pressures can initiate cracking of a hole. These and other quench
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

parameters should be verified if cracking suddenly appears.


As mentioned above, selection of appropriate process parameters and designs often
allow considerably reducing hot spots around the hole area, eliminating cracking prob­
lems. The etched section of a transmission component shown in Figure 4.52 can serve as
a good example of a successful surface hardening part comprising multiple holes.
It is still important to keep in mind that a complex stress distribution along the perim­
eter of the hole takes place during the heating and quenching cycles.
An attempt to reduce the heat surplus in some areas should not produce localized
severely underheated or improperly austenitized regions. This is illustrated in Figure
4.137. The left image shows an as-quenched microstructure at the oil hole, showing the
presence of ferrite networking in the grain boundaries (200×) near the hole. The image on
the right shows the same image at higher magnification (3000×). Insufficient austenitiza­
tion produces such structures.
It is good practice to have holes generously chamfered, having smooth surfaces and being
free of machine marks, burrs, and chattering. Excessive chattering (Figure 4.138), folds,
and large unfavorably oriented strainers or chains of coarse and preferentially oriented
inclusions might act as measurable stress risers (Figure 4.139 right three images) acting as
a nucleation site for cracking and developing preferential paths for crack propagation. The

0.100 mm

1.000 mm

0.583 mm

The case microstructure at an oil hole A high magnification of the unstransformed


showing the presence of ferrite in the grain ferrite networking of the imageon the left.
boundaries (200×). Note the slight melting of the surface (3000×).

FIGURE 4.137
As-quenched microstructure at an oil hole showing the presence of ferrite networking in the grain boundaries
near the hole. Left image, 200×; right image, 3000×. Insufficient austenitization produces such structures.

(a) (b) (c)

FIGURE 4.138
Inappropriate chamfering holes (a and b) should be avoided since it can trigger crack development (c).
Heat Treatment by Induction 323
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

FIGURE 4.139
Folds, large unfavorably oriented strainers or chains of coarse and preferentially oriented inclusions, might act
as stress risers causing cracking.

presence of nicks, deep scratches, notches, burrs, and other mechanical defects increase
the stress concentration factor and it is highly undesirable.
There are a number of other helpful practical solutions and know-how available for heat
treaters to improve heat uniformity. For example, in single-shot induction hardening, pro­
filing of inductor copper often allows dramatically reducing or eliminating hot spots in
the vicinity of holes. Several patented design concepts related to inductor profiling around
oil holes were created during the development of the nonrotational crankshaft hardening
technology [38,351,352]. Those approaches allow selectively controlling heat source distri­
bution along the oil hole perimeter by providing preferable channels for eddy current flow.
It is also beneficial to orient the part in such a way that the coil “fish tail” would be posi­
tioned near the oil hole, allowing electromagnetic fringing to help reduce the heat surplus.
Certain challenges can appear when the part consists of several closely spaced holes
(Figure 4.134d). In such cases, a sequence of cold spots and severely overheated regions
may occur.
Experience shows that in many cases, the proper choice of design parameters (applied
frequency, power density, inductor profiling, etc.) allows one to obtain the required hard­
ened pattern around holes free of cracks, even in those cases that may seem first unsuit­
able for heat treating by induction. For example, in some cases, even such complex-shaped
parts as a ball bearing cage, which consists of a number of closely located large holes, can
be surface hardened or contour hardened by induction instead of using a time- and space-
consuming carburizing processes.
Keyways and snap ring grooves at the ends of shafts are also concerns for overheat­
ing of corners and underheating in the bottom of the groove. Keyways can be considered
extreme cases of longitudinal holes (Figure 4.140). Size, shape, and orientation of keyways
with respect to eddy current flow (which depends on the coil orientation) have a substan­
tial impact on the ability to obtain the required temperature profiles within the keyway
area and to avoid undesirable hot and cold spots.
Care must be taken when hardening snap ring grooves. If only half the groove is hardened,
it typically causes an undesirable transient and residual stress distribution in the groove.
Also, grooves and undercuts require special attention because they result in redistribution
of the case pattern owing to electromagnetic end and edge effects as well as thermal edge
effects. The required hardness pattern in the undercut area depends on the application specif­
ics. In some cases, the hardness pattern has to stop at a certain distance from the groove, but
in other cases, the entire ring groove area is specified to be hardened (Figure 4.141a and b).
Metallurgical irregularities (abnormal or undesirable grain flow pattern, excessive seg­
regation, coarse inclusions, etc.) may complement geometrical stress risers of grooves,
324 Handbook of Induction Heating

Slightly
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

overheated
areas

Significantly
underheated Substantially
areas overheated
areas
(big possibility
for cracks)
Keyway Keyway

FIGURE 4.140
Size, shape, and orientation of keyways with respect to eddy current flow have a substantial impact on tempera­
ture profiles within the keyway area. (From V. Rudnev, R. Cook, J. LaMonte, Induction heat treating: Keyways
and holes, Metal Heat Treating, March–April 1996, pp. 83–86.)

Hardness pattern

(a) (b)
Snap ring groove (c)

FIGURE 4.141
Grooves and undercuts require special attention because they result in redistribution of the pattern owing
to electromagnetic end and edge effects as well as thermal edge effects. (a) Unhardened groove, (b) hardened
groove, and (c) crack at groove.

undercuts, and keyways, increasing the probability of cracking. As an example, Figure


4.141c shows that the crack follows the general direction of the flow lines (darker area on
the right-hand side indicates the runout).
In some instances, product engineers prefer to have a slightly lower hardness in the
undercut area, and instead of a fully martensitic microstructure, they sometimes prefer
to have a “martensitic–low bainite” mixed structure. This provides some ductility in the
undercut area, reduces brittleness, and decreases the probability of cracking.
When discussing heat nonuniformities owing to the presence of longitudinal and trans­
verse holes, keyways, sharp edges, and corners, it is difficult to overestimate the impor­
tance of smooth chamfers (rounding of sharp corners can be a great help in decreasing the
possibility of overheating and cracking). The repeatability of machined chamfers, their
depth, and location with respect to the inductor are of great importance as well.

4.4 Control of Distortion and Prevention of Cracking


Heat treaters are often faced with the necessity of making a reasonable compromise
between providing the required hardness and strength while obtaining a sufficiently
Heat Treatment by Induction 325

tough structure that has the desired distribution of residual stresses and their magnitude
to enhance certain mechanical properties and producing components with minimum dis­
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

tortion. The subjects of distortion control and causes and prevention of cracking are a
part of failure analysis. A vast number of publications are devoted to this very broad and
complex subject [7,30,170–181,185,186,259,280,345,346,355,375–380]. One should not expect
to find a complete discussion on this subject here. Only a short introduction to this subject
will be provided. For those who are interested in the subtleties of mathematical modeling
of phase transformations, stresses, and distortion, we suggest to review publications by Dr.
Lynn Ferguson and Dr. Zhichao Li (DANTE Solutions, Inc.), Professor H.K.D.H. Bhadeshia
(University of Cambridge), John Walters (Scientific Forming Technologies Corporation),
and some others. Several commercially available software might also be helpful in this
respect, including DANTE, DEFORM, Thermo-Calc, and others.
Shape/size distortion of heat-treated components is affected by a number of factors.
The specifics of the component’s geometry, its chemical composition and microstructure,
inductor design specifics, processing temperatures, quenching subtleties, hardness pat­
tern, magnitude and distribution of transient and residual stresses, and design of tooling/
fixtures are among the most critical factors affecting distortion.
Virtually an endless variety of components that are routinely induction hardened (Figure
4.52) call for a corresponding almost endless variety of hardening inductors (Figures 2.1,
3.2, 4.58, 4.88, 4.92, and 4.96). Each of these components and inductors has its own person­
ality that affects the outcome of heat treatment, including distortion characteristics and
probability of cracking.
Several case studies devoted to the issue of cracking have been discussed earlier in
this chapter (including Sections 4.1 and 4.3). Figure 4.142 shows the “fish bone” dia­
gram of cracking and prevention related to induction hardening [17,280], revealing the
complexity of this subject and its interrelated nature. As can be concluded from the
fish bone diagram, the potential causes of crack development can be categorized into
seven main groups:

• Material related
• Geometry of the workpiece
• Power/energy cycle (heating stage) and unspecified heating conditions
• Improper quenching conditions (including quenching severity and nonuniformity)
• Inadequate inductor design
• Improper accessories and tooling
• Other factors

Very high temperatures and stresses are prime causes of cracking and excessive shape/
size distortion. Surplus of heat generation, particularly when hardening complex geome­
tries, is inevitably associated with larger and unequal metal expansion/contraction, which
in turn causes greater distortion.
Stresses can be classified in several different ways. Depending on the distance over
which they extend, stresses can be “macroscopic” or “microscopic” [346]. Macroscopic
stresses typically appear at a distance that exceeds several grains. In contrast, microscopic
stresses take place within a grain and include stresses that appear on the atomic level.
Studies of residual stresses in steel heat treating typically focus on the distribution and
magnitude of macroscopic stresses.
326 Handbook of Induction Heating

“Fish bone” diagram of cracking and troubleshooting


Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

Improper prior Sharp edges


metal working practice and shoulders
(i.e., folds, burrs, torn metal, Part
Coil geometry
severe chatter, grain growth, geometry
microcracks)
Nonhomogeneous
structures (e.g., Coupling
severely segregated Poor (air gap)
and banded chamfer and
structures) Excessive roundings
inclusions
and foreign Location, Geometry
elements orientation and size related
Material Chemical of holes, groves, Part
related composition and undercuts, groves and positioning
residuals other geometrical
Quench irregularities
delay Fixture Tooling
integrity set-up
Flow rate
and pressure Energy
Concentration
Calibration Field
Quench
fluctuation
uniformity
Temperature Higher-order
harmonics

Contaminations Load
Quench matching Power
related Type of cycle
quenchant Copper Overheating
overheating Burns from
Quench sparks
time De-carburization
Inductor
Grain
profiling
Coil bondary
liquation Other
impedance
factors that
might indicate
Inductor Excessive tendency to
design Flux concentrator Type of crack grain coarsening potential
and scaling problem

FIGURE 4.142
“Fish bone” diagram of cracking and prevention in induction hardening revealing the complexity of this sub­
ject and its interrelated nature.

On the basis of the nature of their cause in induction hardening, several types of stresses
are encountered: thermal stresses, phase transformation stresses, and applied stresses.
Thermal stresses are caused by different magnitudes of temperature and thermal gradients,
while phase transformation stresses primarily occur because of volumetric changes accom­
panying the formation of different phases such as austenite, bainite, martensite, and others.
Applied stresses are associated with the specifics of tooling/fixture of heat-treated com­
ponents during heating and quenching. For example, shafts experience compressive and
torsion stresses during their rotation in a vertical single-shot hardened systems. If dis­
tance between support rolls is too long, an excessive bending may occur during induction
through hardening of elongated parts using progressive/continuous systems. Buckling
Heat Treatment by Induction 327

when hardening thin components (e.g., bumpers) may cause excessive distortion. If a rela­
tively thin but elongated workpiece is firmly held between two holders or nests, then it
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

will “try” to expand during the heating cycle, producing buckling.


Different techniques are used to reduce distortion. This includes applying the support/­
tooling designs that allow for part growth by utilizing a low-pressure spring on one of
tooling.
The total stress is a complex combination of a number of factors. At different stages of
heat treating, the effect of the different components on total stress will also differ.
Although stresses are 3-D—with axial, circumferential (hoop), and radial components—
discussions are often simplified by considering them as one-dimensional stresses utilizing
their magnitude.
On the basis of the timing of their appearance, stresses can be divided into three groups:
initial, transient, and residual stresses.

• Initial stresses. Their distribution and magnitude depend on the manufacturing


steps preceding IHT (casting, forging, welding, extrusion, bending, rolling, upset­
ting, drawing, etc.). For example, the maximum of tensile residual stresses of
appreciable magnitude is commonly located in the core of hot forged components.
Stress relieving is often used to ensure that repeatable results after hardening are
obtained. Some of the stress-relieving operations and schedules have been dis­
cussed in Chapter 2. Upon reaching austenite phase temperatures, initial stresses
that are “hidden” within the green part are relived to different degrees (particu­
larly in the case of through hardening applications). This factor alone might cause
a noticeable distortion. However, induction heating might be blamed for it. In
reality, the distribution and magnitude of the initial stresses formed after previ­
ous technological operation may play a major part in it. Thus, if in doubt, stress
relieving of a workpiece before IHT should be done to assess the potential for a
distortion.
• Transient stresses. In addition to initial stresses, there are transient stresses, which
occur because of volumetric expansions and contractions associated with temper­
atures, thermal gradients, and the phase transformations. Some of these stresses
may pose no problem. Others may accentuate the initial stresses and lead to exces­
sive distortion and even cracking. Transient stresses are responsible for the great
majority of cracking in induction hardening.
• Finally, there will be residual stresses that remain within the workpiece after the IHT
process is completed. Thus, residual stresses are the product of initial and tran­
sient stresses. In many cases, these stresses can be very beneficial, complementing
desirable microstructural changes in providing needed engineering characteris­
tics. For example, compressive residual surface stresses generally improve fatigue
resistance. In other cases, magnitudes and the orientation of residual stresses may
be harmful, causing excessive distortion (which may require post–heat treat oper­
ations such as grinding or straightening), increasing crack sensitivity, facilitating
crack initiation and propagation (as in case of significant tensile residual surface
stresses), and shortening the component’s service life.

The interrelationship between metallurgical and geometric features of the heat-treated


component and various process parameters is complex and multifaceted. The development
328 Handbook of Induction Heating

and control of the transient and residual stresses are the most difficult to analyze and
optimize. Figure 4.143 illustrates the representation of the simplified dynamics of macro­
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

scopic stress appearance at the surface of a carbon steel solid cylinder during induction
hardening.
Heating: As the temperature rises, surface compressive stresses form and increase in
magnitude (Figure 4.143). In the 520°C to 700°C (970°F to 1300°F) range, steels undergo
plastic volumetric expansion and surface stresses begin to decrease. Finally, when the tem­
perature exceeds approximately 850°C (1560°F), the steel’s surface plastically expands, and
the diameter of the heated area becomes greater than its initial diameter. Since the yield
point of the surface layer is considerably lower at elevated temperature, the material will
flow plastically and surface stresses will significantly decrease.
Cooling: After the quench fluid is sprayed onto the austenitized surface, the outermost
layer quickly loses its plasticity and tensile stresses appear at the surface (Figure 4.143).
Because of the drastic temperature decrease of the thin surface layer, it tries to reduce its
volume, but hot subsurface regions are still at sufficiently high temperatures preventing
that reduction. As the process of quenching continues, the greater mass of the outer layer
cools down, becoming nonplastic and trying to shrink further building tension stresses at
the surface.
The maximum of the surface tensile stress typically occurs just above the Ms (martens­
ite start) temperature. The formation of martensite (and associated with it the expansion
caused by metallurgical phase change) reduces surface tensile stresses and leads to the
development of surface compressive stresses assuming that a sufficient amount of mar­
tensite was built. Upon completion of cooling, a complex combination of compressive and
tensile stresses exists within the part (Figure 4.144). It is important to remember that the
residual stress system is self-equilibrating; that is, there is always a balance of stresses
within the workpiece owing to mechanical equilibrium. If certain regions have compres­
sive residual stresses, then somewhere else there must be offsetting tensile stresses. If the
stresses were not balanced, “movement” would then result.
Figures 4.143 and 4.144 illustrate that, in induction surface hardening, it is reasonable to
expect compressive residual stresses at the surface of as-quenched components (see also

Temperature, °F
32 392 752 1112 1472 1832
Tension
Stresses

Stage
g

0
in
nch
ue

Hea
Q

e
Compression

ting
ag
St

0 200 400 600 800 1000


Temperature, °C

FIGURE 4.143
Simplified dynamics of the appearance of macroscopic surface stress during induction hardening of a carbon
steel cylinder.
Heat Treatment by Induction 329

Heating pattern Coil


Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

O.D. of part after its


induction heating Shearing stresses
Tension Compression

Tensile stress
Quenching

maximum
Tensile stress
maximum
Tension
Radial stresses

Compression

FIGURE 4.144
Formation of residual stresses in induction hardening.

Figure 4.53). Exceptions would include very shallow case depths or incompletely trans­
formed martensitic structures (e.g., mixed structures). In some cases, too deep hardness
patterns may also result in reversing the residual surface stresses from compression to
tensile.
Surface-compressive residual stresses are considered useful and desirable in most
applications. Among other factors, these stresses are important for improving the fatigue
properties of the workpiece, delaying the crack initiation and the propagation of micro­
cracks and providing desirable protection against geometrical stress risers (e.g., micro­
scopic scratches, notches) as well as microstructural heterogeneities. Compressive residual
stresses are particularly beneficial to parts that experience bending, impact, or torsion in
service. They enhance the fatigue resistance because the applied stresses in service are
typically tensile in nature and the presence of the compressive residual surface stresses
helps minimize the magnitude of the sum of both.
Tensile residual stresses of a certain magnitude and location, on the other hand, can
be dangerous, making a negative impact on fatigue resistance and may also cause stress-
corrosion cracking in the presence of a corrosive environment. These stresses are com­
bined with the load stresses in service amplifying the total stress magnitude, which may
exceed the strength of material resulting in a crack development.
330 Handbook of Induction Heating

In surface hardening applications, the maximum of the tensile stress is usually located
just beneath the hardened case or within the transition zone, and it is often responsible
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

for subsurface crack initiation. In contrast, in through hardening applications, the surface
residual stresses are often tensile (Figure 4.53), but under certain quenching conditions,
they can be reversed into compression.
The formation of tensile residual stresses at the surface in through-hardened applications
is associated with the timing of martensite formation (at the surface vs. the core) and the
correlation between thermal contraction during quenching and a volumetric expansion
during austenite-to-martensite transformation. The fact that surface martensite experiences
a much greater quench severity compared to martensite formed in the core also makes an
appreciable impact on the distribution of the residual stresses. Ref. [177] provides a com­
prehensive computer modeling study to illustrate the distribution of the residual stresses
in a surface-hardened versus a through-hardened component and the main conditions
affecting the stress patterns.
The greater magnitude of residual stresses of the hardened component is usually associ­
ated with increased brittleness and notch sensitivity and reduced toughness. Therefore,
stress relief/tempering is commonly specified (though not always). The goals are to reduce
the maximum of the tensile stress and move it farther from regions of applied stress, while
retaining the useful surface compressive stresses.
A final grinding operation can also have a pronounced effect on the residual stress dis­
tribution. Inappropriate grinding not only can noticeably reduce the useful compressive
residual stresses at the component’s surface but also can reverse it. Therefore, the impact of
grinding should be considered not only when specifying the hardness pattern but also when
developing the required residual stress state of the as-quenched part [176,177,375,376].
In induction-hardened parts of complex shapes, the residual stress distributions are
much more complex compared to plain solid cylinders. As a result, the measurement of
residual stresses is often not an easy task, and special equipment and a great deal of time
and experience are needed. Techniques for quantifying residual stresses include the sec­
tioning, hole drilling, layer removal, bending deflection, x-ray diffraction, magnetic, and
ultrasonic methods [173,346]. At the time of the writing this book (2016), there is no reliable
method to measure/assess the magnitude of transient stresses except utilizing numerical
computer simulations.
Mathematical modeling is an essential tool that helps assess the complexity of transient
and residual stress distributions taking into consideration the intricacy of the kinetics of
displacive and reconstructive transformations based on thermodynamic and phenomeno­
logical approaches and has been discussed in many publications. Probably the most inten­
sive theoretical and computer modeling studies of stresses in induction heat treatment are
those conducted by Dr. Lynn Ferguson and Dr. Zhichao Li (DANTE Solutions, Inc.) [177].
It would be appropriate at this point to briefly review one of the case studies to illustrate
the capabilities of FEA computer modeling in discovering unique distributions of tran­
sient stresses assessing their impact on probability of cracking when induction hardening
components with angled holes.
Figure 4.145 shows the results of FEA computer simulation of the dynamics of tempera­
ture distribution before spray quenching (a) and after 0.4 s (b) and 2 s (c) of quenching an
austenitized surface of crankshaft journal containing an angled oil hole.
Figure 4.146 shows the decomposition of austenite after 0, 2, and 3 s of spray quenching
respectively. Computer modeling reveals that in contrast to the dynamics of stress forma­
tion illustrated in Figure 4.143 that shows a single peak of tensile stresses, when quenching
oil holes, there are two peaks of tensile transient stress occurring in the acute angle area.
Heat Treatment by Induction 331
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

Node
4320

Node 4344

Time = 0.37 s
Time = 3.0 s
Induction heating
(a) (b) (c)

FIGURE 4.145
Results of FEA computer simulation of the dynamics of temperature distribution before spray quenching (a)
and after 0.4 s (b) and 2 s (c) of an austenitized crankshaft journal containing an angled oil hole. (Courtesy of
DANTE Solutions, Inc.)

Time = 0.0 s Time = 2.0 s Time = 3.0 s

FIGURE 4.146
Results of FEA computer simulation of the dynamics of decomposition of austenite after 0, 2, and 3 s of spray
quenching, respectively (compare with Figure 4.145). (Courtesy of DANTE Solutions, Inc.)

The second peak of tensile stress may have a substantially greater magnitude exhibiting
an increased danger for crack development. This information helped in developing the
know-how to dramatically reduce the probability of the quench cracking in the hole area.
When discussing heat nonuniformities appearing when heating parts with geometrical dis­
continuities, it is necessary to emphasize once again that smooth chamfers on the edges and
rounding of sharp corners can be a great help in decreasing the probability of overheating and
cracking. Short quench delays and the use of indirect quenching technique may be very help­
ful to prevent cracking and was successfully applied in developing the CrankPro machine.
As stated earlier in Section 4.1.4, if steel does not respond to heat treatment in an expected
way (including sudden cracking), then one of the first steps in finding the cause for such
behavior is to check the chemical composition and assess the steel’s prior microstructures. It
is imperative to have a clear understanding regarding the exact chemical composition of steel,
the presence of residual and trace elements, and peculiarities of prior microstructure. Wide
compositional limits resulting from multiple steel suppliers may cause appreciable deviations.
Excessive distortion may also be related to nonuniform quenching. For example, nonuni­
form thermal conditions may be caused by the complexity of the workpiece geometry, failure
to quench with sufficient intensity certain regions, plugged quench orifices, improper quench
design, the eccentricity of the inductor and quench system compared to the workpiece, con­
taminated or inadequate quenchants, and so on. Excessive wobbling during a part’s rotation
may cause a phenomenon when a certain area of the heated component will always have better
electromagnetic coupling with the inductor. This can produce a localized temperature surplus.
Regardless of the part rotation, certain areas might continue to be the hottest area, because it
332 Handbook of Induction Heating

will still be the closest to the coil. This nonuniform heating may also cause uneven quenching.
Uniform quenching is very important when trying to minimize distortion and eliminate a
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

“banana” effect. Inappropriate part rotation can also be a decisive factor worsening distortion.
The fixture and tooling may also make an appreciable impact on distortion. For example,
when single-shot hardening axle shafts using horizontal machines, the support rollers
located in the center of the shaft provide needed support, keeping the shaft from bow­
ing in the center and minimizing a distortion. This roller assembly is similar to what is
referred to as a “steady rest” when turning a long part in a lathe. Because of the proxim­
ity to the inductor, these rollers are made of either ceramic or stainless steel. Properly
designed chucks and support fixtures may also noticeably help reduce distortion in verti­
cal axle shaft hardening systems.
The presence of surface decarburization (Figure 4.32b) indicates the loss of surface and
near-surface carbon and may reverse the expected residual stress distribution of the sur­
face-hardened part from surface compression to tension and this must be avoided. Surface
cracks often initiate at decarburized layers.
A combination of through-hardened and case-hardened areas as well as unbalanced
sectional mass or substantial variations in the hardness case depths of closely positioned
sections are potential causes for excessive distortion and cracking.
Measures should be taken to avoid excessive heating of the parting line regions (e.g., com­
ponents made of resulfurized forgings), because a chemical seregation, high concentration of
inclusions and residuals as well as the occasional presence of preexisting defects formed during
trimming [355]. Parting lines are particularly sensitive to rapid heating and high temperatures.
On many occasions of dealing with steels that exhibit high brittleness and a tendency
for cracking, moderate and low quenching rates or interrupted quench may help reduce
the probability of crack initiation and growth as well as an excessive size/shape distortion
of the surface-hardened components without sacrificing the ability to obtain the required
martensitic structure and hardness levels. In short, less severe quenching is typically asso­
ciated with a reduced probability of cracking.
On the other hand, some researchers (i.e., Dr. N. Kobasko [198,251] and Dr. K.
Shepelyakovskii [199]) believe that the effect of the cooling rate on crack initiation appears
as a bell-shaped curve with marked increase in probability of cracking within the 250–
400°C/s range. This postulation suggests that in cases when the cooling rate is extremely
severe, which is several times greater than conventional quenching intensity and is suf­
ficiently uniform, the probability of cracking can be reduced even when hardening high-
carbon steels.
Magnetic particle inspection (MPI or magnafluxing) is an effective nondestructive test­
ing tool that helps reveal whether surface cracks are present or not. According to this tech­
nique, the part is wetted using a solution that contains iron oxide particles and placed in a
magnetic field. The magnetized part is inspected using ultraviolet light. Particles attracted
to surface discontinuities (including shallow cracks that might be invisible to the naked
eye) become visible under the light and can be detected by an experienced operator (Figure
4.139, middle three images). Proper operator training and application of sufficient magneti­
zation and appropriate orientation of magnetic flux are imperative in ensuring a successful
inspection during MPI.
Untempered martensite is often (not always though) too brittle for commercial use. As
discussed in Section 2.1.2, tempering can provide a desirable combination of strength,
toughness, ductility, and magnitude and distribution of residual stresses.
If specified, tempering should be done as soon as possible. Very long delays between
hardening and tempering may result in the so-called delayed cracking. It is particularly
Heat Treatment by Induction 333

vital to minimize the delay time and position the tempering operation immediately after
the hardening machine or at least in the same building when dealing with low-toughness
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

materials (e.g., cast irons, high-carbon steels, etc.) and high hardness levels. It is beneficial
at this point to review a case study.
Case study. Consider this scenario. After the development stage, an induction hardening
machine is built and successfully completes a runout in June. The machine was shipped to
the customer in the middle-west of Canada. After several months of successful production
runs, parts suddenly begin to experience cracking. MPI has been done after the tempering
operation. Cracking appeared in the second half of November.
A check of the induction machine process parameters (power, frequency, quenchant
temperature, concentration, pressure and flow, part positioning inside an induction coil,
etc.) shows everything is in order. The steel’s chemical composition and prior microstruc­
ture did not reveal any abnormalities. The question is, “Why does cracking in the parts
suddenly occur?” Such an occurrence could present a challenge in determining the root
cause of the cracking and ways to eliminate it.
After a detailed evaluation, it has been found that the hardening machine and temper­
ing furnace were placed in two different buildings and after the hardening operations that
took place in one building, parts were gathered and transported to another building. Since
outside temperatures were below the freezing point (<0°C), the as-quenched parts made of
cast iron were continually quenched during their transportation to a building where the
tempering furnace was placed, transforming RA into untempered martensite.
Solution. After the tempering operation was placed in the same building as the harden­
ing machine, the problem of cracking was eliminated.
It is also important to remember that improper induction tempering may produce a
phenomenon that is typically referred to as inverse (reverse) hardening, which is associ­
ated with the appearance of “a wave” not only in the hardness distribution but also in the
reversal of residual stresses. This phenomenon will be discussed later in Section 4.6.
Short conclusion. If cracking or excessive distortion occurs, there may be a number of
interrelated factors responsible for the appearance of those undesirable phenomena. Even
a brief look at Figure 4.142 illustrates the complexity and challenges of finding the root
cause(s) of the problem and the consequential factors.
An interested reader can find numerous publications and words of wisdom shared by
industry experts. Some of the recommendations are outlined and explained in this hand­
book as well as provided in the reference list. Understanding a broad spectrum of fac­
tors associated with various failure modes is an important step in developing advanced
IHT technologies. At a minimum, an analyst must have knowledge of failure analysis,
electromagnetics, metallurgy, materials science, mechanics, heat transfer, chemistry, spec­
troscopy, and computer modeling. A certain degree of familiarity with the customer’s
required method of failure analysis is useful as well (i.e., automotive, farming industry,
government, etc.).
Technological breakthroughs in induction surface hardening appear quite regularly,
helping to dramatically reduce distortion of the components. Achieving almost undetect­
able distortion and, in many cases, elimination of an entire straightening operation is the
result of three factors:

1. An ability to form a true uniform hardness pattern


2. Minimization of peak temperature
3. Avoidance of applying any pressure/force during hardening
334 Handbook of Induction Heating

For example, when hardening gears and gear-like components, an ability to achieve a
contour-like hardness pattern also helps dramatically minimize a distortion [400].
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

The CamPro™ machine utilizing patented SHarP-C Technology can serve as another
example, revealing the ability to produce not only superior metallurgical properties of
induction-hardened lobes and bearings of automotive camshafts but also dramatically
improved straightness characteristics [22,27,299]. A testimonial of one of the users of this
advanced process published in Ref. [27] could be considered as an objective assessment
quantifying benefits of this technology based on obtained real-life records: “The SHarP-C
hardening machine helped us to reduce the camshaft’s distortion down to 3–5 microns and
we have been able to eliminate the entire straightening operation. So, our savings on elimi­
nation the straightening operation alone is about $40,000 per year. On top of that there has
been substantial improvement in the quality of the hardened camshafts, and our scrap was
reduced about 1.5%.” SHarP-C Technology is discussed in detail in Section 4.9.5.3.
Hints for troubleshooting excessive distortion

• Estimate the impact of initial stresses formed during previous manufacturing steps. In
order to establish a datum or a baseline and assess if initial stresses have measur­
able impact on excessive distortion after IHT, the geometry of “green” workpiece
should be measured and undergo tempering/stress relieving. In order to do so, a
workpiece should be placed into a tempering oven at 550°C–650°C (1050°F–1200°F)
for 1.5–2 h. As an option, parts can be normalized; however, complex geometry
components and in particular those having a combination of thick and thin sec­
tions might be cooled differently, producing certain microstructural differences in
as-normalized structures (e.g., alterations in ferritic grains sizes, a combination of
fine and coarse pearlites, just to name a few). This may alter distortion characteris­
tics. Besides that, when dealing with Q&T initial structures, their re-austenization
(in case of normalizing) might also contribute to a deviation of the results. Keep in
mind that normalized components may exhibit noticeable surface decarburization
affecting the results of IHT. Thus, the purpose of normalizing is limited to assess­
ing initial stresses only in this case. These are some of the reasons why tempering/
stress relieving is preferred. The subject of tempering is discussed in Section 4.6.
• Geometry of the heat-treated component should be within specified clearances (particu­
larly when applying RF and dealing with extremely shallow case depth).
• Review position and condition of workpiece holding devices, fixtures, nests, support rollers,
chucks, steady rests and centers, accessories, and part’s positioning during induction hard-
ening. Unspecified conditions should be avoided. This includes part’s tilting dur­
ing heating and quenching, excessive wobbling, part’s slippage during rotation,
unspecified nonsymmetrical positioning, and so on. In order to avoid warpage,
when dealing with thin elongated components, they should not fit snugly, allow­
ing expected part growth during heating. It is equally important to avoid applying
excessive force for part’s rotation. Sufficient support(s) should be provided to avoid
sagging—this is particularly critical in through hardening applications or when
the component’s core is heated to elevated temperatures, losing its capability to
“act” as a shape stabilizer. Make sure that magnetic flux concentrators (if applied)
are not shifted during heating because of the presence of magnetic forces or that
their excessive structural degradation does not take place.
• Attempts should be made to have as uniform heating as possible. Unspecified thermal
gradients (e.g., axial, circumferential, etc.) can worsen distortion characteristics
Heat Treatment by Induction 335

tremendously. An exception can be given to dealing with irregular geometries


where cold sink compensations might be needed or when it is required for achiev­
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

ing hardness gradients.


• One of the critical factors affecting a distortion is the amount of heat generation.
Excessive heat generation should be avoided since it is inevitably accompanied
by greater and unequal metal expansion/contraction, which is often associated
with greater distortion. Short HAZ complements a reduction of the mass of heated
metal and helps minimize the distortion. Also, it is typically highly desirable
to reduce a magnitude of processing temperatures that, among other factors, is
linked to thermal gradients during heating and quenching.
• Heating time has a complex impact on a distortion greatly depending on application
specifics. On one hand, prolonged heat times can reduce peak temperatures and
thermal gradients (it is beneficial). However, extended heat times may noticeably
increase the mass of metal being heated (it may aggravate distortion characteristics).
• Attempts should be made to make quenching as uniform as possible with sufficient
quench flow. Inappropriate quenching is one of the most common causes of crack­
ing and excessive distortions. Suitable quenching conditions during a runoff
should be recorded and the properties of the quenchant should be maintained
(including flow, pressure, concentration, temperature, purity, etc.). Well-thought
quench design should properly address gravity, bouncing effects in spray quench­
ing, the presence of geometrical irregularities, and hard-to-access areas. In some
not-so-frequent cases, it might be beneficial to intentionally provide a certain non­
uniform 3-D quench intensity distribution in order to control a distortion. Short
quench delay might also be helpful in some cases.
• “Locked in place” and uniformly distributed hardness pattern is also a powerful
factor in keeping distortion under control (see several case studies provided in
Section 4.9).

4.5 Accessory Equipment and Work Handling for Heat Treating


Because of the ability to heat parts quickly, induction heating readily lends itself to many
standard machine types. It requires integrating the coil design, quenching, and part hold­
ing fixtures to the chosen work handling. In automatic systems, the work handling trans­
fers the part through the heating and quenching stages as well as the preheating and
postheating operations such as loading and unloading (accessory equipment).
Often, to add automation, robots, gantries, and pick-and-place units are used. Hoppers,
magazine feeders, and rotary index tables are some other common ways to present parts
for induction heat treatment. Conveyors (Figures 5.10 and 5.13), walking beam systems
(Figures 4.80 and 4.88), and rotary tables (Figure 4.75) can be used for feeding or unloading
and processing parts through the heating and quenching operations.

4.5.1 Robots, Gantries, and Pick-and-Place Units


The use of robots and gantries is common for modern high-production equipment as
well as for handling larger irregularly shaped parts (Figure 4.63b), mainly because heavy
336 Handbook of Induction Heating

cumbersome parts are not located near conveyors or other part transfer equipment. The
time to manually load these types of parts would not be cost-efficient, or in cases of han­
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

dling heavy workpieces, manual handling may simply be unrealistic.


Robot and gantry loading can be done with virtually any size or shape component and
in some cases can be added after the induction equipment is installed. Robots and gantries
come with their own electronic control, which requires a “hand shaking”–type interface
between the machine control and the robot or gantry system to clearly define which sys­
tem has control of the part and when each system is free to move without damage to the
other systems. Because of the potential danger associated with moving machinery in prox­
imity to plant personnel, the issue of safety is of primary concern and must be addressed.
When dealing with lightweight workpieces, systems may be designed with the option
of using the induction equipment with or without the robot or gantry. This allows manual
production to continue in the event that the robot or gantry is out of service. When this is
done, positive shot pins are often used for safety to ensure that the robot or gantry can­
not move if an operator is to be running the machine within the normal robot or gantry
envelope of operation.
Pick-and-place units may be used externally for loading a machine or at times integrated
into the machine itself. A pick-and-place unit offers linear or rotary motion similar in some
respect to a robot but lacks the sophisticated control electronics and free range of motion
provided with a robot. Often, only the pick-and-place mechanics are purchased and the
control is integrated into the PLC that controls the machine operation.

4.5.2 Hoppers and Magazines


Magazines and hoppers are used when the same or a similar family of parts is to be pro­
cessed. When referring to magazines, it typically means that a given quantity of parts is
loaded into a cartridge and the parts are dropped either into the coil’s workplace for heat­
ing or onto a conveyor system for transferring to the work coil.
Hoppers are typically thought of as a tub of randomly oriented parts. From the hop­
per, the parts are sorted and oriented before they are presented to the induction system.
This can be done in a number of ways such as by vibration, magnetically, or by specially
designed mechanical fingers, just to name a few. For example, when hardening of bolts or
screws after sorting, the parts are indexed into the coil for heating and quenching.

4.5.3 Conveyors
There are many types of conveyors used for a number of induction applications, often in
combination with other feeding and loading equipment. Conveyors can be belt, chain,
or cam driven; they operate in a continuous mode or can be indexed via servomotor or
mechanical cam switches. The considerations for choosing a conveyor are usually deter­
mined by the application for which it is being used. It must stand up to harsh conditions
of hot parts and continuous quenchant being sprayed on it. Care must be taken to prevent
unspecified movement of lightweight workpieces (e.g., owing to electromagnetic forces)
while being processed through an induction system. Stainless steel mesh or chain is com­
monly used for prevention of belt corrosion. If plain carbon steel belts or conveyors are
used too close to the inductor, there is the possibility of inductively heating the belt as well
as the workpiece.
Figure 2.13a shows a conveyor to index parts through a multiturn channel inductor. The
process specifics require positioning the conveyor in proximity to the inductors. Thus,
Heat Treatment by Induction 337

means should be taken to avoid overheating the conveyor in a high-production environ­


ment. Many times, one of the critical requirements for a conveyor is to be able to handle
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

extreme temperatures since its function is to carry the hot part away after heating. When
conveyors are used at the input of a machine, consideration must be given to the possibility
of having to back the hot parts out of the machine on the entrance conveyor in the event of
any problems in the part transfer.

4.5.4 Rotary Tables
Rotary tables usually have nests for a given number of parts, and after loading the table,
it is rotated like a carousel. This type of equipment is very popular with cellular-type sys­
tems mainly because the unheated parts are loaded and heat-treated parts are unloaded
in almost the same area of the equipment. Rotary tables are also used where high volume
or high production is required, because a large number of parts can be passed through a
coil very quickly (Figure 4.75). As mentioned above, many times, several different kinds
of work-handling equipment are used in the same processes and combined to make a
completely automated system. For example, a complete system to anneal ball studs utilizes
other work-handling systems in addition to a rotary table.

4.5.5 Unscramblers and Bar Feeders


Automatic control and delivery of long thin bar stock presents its own set of challenges in
separating the bars and lining them up for continuous feeding at the correct rate through
the heating coil(s) without forming a continuous loop from end to end on the system where
ground currents could destroy rollers and bearings on the equipment.
A bar unscrambler mechanism separates the bars and places them on a table to feed on
to an index chain. The chain rapidly brings the part up to the heating position in order to
achieve the maximum production rate. Any mislocation that occurs may lead to overheat­
ing or underheating of portions of the bar.
A bar feeding mechanism may consist of a delivery table with upper and lower rollers
to pinch the part and maintain a constant linear velocity through the heating coil. Care
must be taken to prevent ground loops through the input rollers, the support rollers, and
the machine frame or drive chain, as well as to prevent mechanical damage to machined
parts.
Some rollers are equipped with a rate sensor to ensure that the part is continuously mov­
ing through the heating coil when the power is turned on.

4.5.6 Cylindrical Part Feed Mechanisms


Cylindrical parts may be delivered to a machine in a part gondola or bin, which is dumped
into a hopper feeder to sort and align the parts accommodating the handling equipment
that is typically designed to handle a variety of lengths and diameters.

4.5.7 Pipe and Tube Handling Equipment


Pipe and tube handling equipment may be required to be designed to adequately grip and
rotate the part without distorting or damaging the geometry of the part. Pipes and tubes
lack the strength of a solid core and are susceptible to denting and bending, which may
render them unusable in the final application.
338 Handbook of Induction Heating

On systems used for hardening pipe or tube, care must be taken with the handling sys­
tem to prevent the quenchant from being trapped inside the tube when the end of the tube
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

passes through the quench ring. If the quenchant is trapped inside the tube, it could cause
a lower temperature on the bottom of the tube and severe distortion in the tube (banana
effect) as some sections will be heated or cooled nonuniformly. In cases where distortion
must be held to extremely tight limits, it may be necessary to straighten the pipe or tube
after the hardening or tempering operation.

4.5.8 Basket-to-Basket Systems for Long Products


For high-speed processing of small-diameter long products such as wires or thin-wall
tubing, basket-to-basket designs are used. As an example, Figure 2.17 shows a typical lay­
out of three basket-to-basket induction annealers of ACR copper tubing [34]. Some of the
challenges of handling workpieces in such applications are associated with high process­
ing speeds and the necessity to prevent mechanical damage to the product. For example,
the diameters of processed ACR tubing range from 7 to 12.7 mm, with wall thicknesses
ranging from 0.32 to 0.52 mm at speeds of up to 600 m/min (10 m/s). Precise rotation of
both baskets is one of the key factors when using basket-to-basket workpiece handling
systems.

4.5.9 Strip Heating Mechanisms


Metal strips to be heated on induction systems may range from dozens of millimeters
(a few inches) in width to more than a meter (several feet). Smaller strips may be used for
the production of bearings or other smaller parts, while larger strips may be used for large
panels and stampings.
Often, strip products are induction heated for coating such as galvannealing or edge
hardening. On these systems, care must be taken to precisely control the position of the
strip inside the heating coil. Electromagnetic forces can cause the strip to move within the
coil if not carefully controlled in particular when transverse flux inductors are used (see
Section 6.9.2.2).
Since most strip heating systems are a part of a continuous process that involves unroll­
ing, heating, cooling, and recoiling the strip on some type of roller, it is essential that the
IH system be very robust in order to prevent expensive shutdowns of the equipment dur­
ing the processing.

4.5.10 Accessory Systems for Processing Large Steel Plates


Heating large metal plates requires a system that is capable of moving large rectangu­
lar plates of steel or other metals. Because of the large size of the parts and the required
processing speed, the power required for these types of systems may be in the megawatt
range. Care must be taken to prevent uneven heating or quenching from top to bottom
of metal plates in order to prevent severe distortion and degradation of the rollers owing
to the high operating temperature and mechanical forces involved. Flat workpieces can
be heat treated for the purpose of subcritical annealing, stress relieving, hardening, and
normalizing. A typical application for hardening large metal plates, either flat or curved
section, would be the hardening of grader blades that may be up to 12 in. wide, 6 to 8 feet
long, and 1 in. or more in thickness.
Heat Treatment by Induction 339

4.5.11 Work Holding Centers


Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

Although not a completely separate auxiliary system, centers, cups, nests, and pedestals
that are used to hold the ends of a workpiece during processing can have a major effect on
whether the required hardness pattern can be attained. Cups or centers that are made out
of magnetic steel can overheat and cause variations in the pattern runout at the end of the
workpiece. The use of stainless steel can minimize this problem and allow heating over
the end of a shaft or other type of workpiece without overheating or degrading the centers.
In order to prevent ground loops through tooling and bearings, one of the centers is usu­
ally isolated to prevent current from flowing through the center to complete a closed loop.
Even though the current is relatively small, it can do irreparable damage to bearings and
may result in costly downtime for the machine.

4.6 Tempering of Induction-Hardened Components


4.6.1 General Comments Regarding Tempering
As-quenched martensitic structures are usually associated with high hardness and
strength but low ductility and toughness. There is a belief that untempered martensite
is too brittle for commercial applications and tempering of as-quenched components is
always required. Depending on the chemical composition, hardness levels, geometry of
the component, and its function, this may or may not be true. Applications of hardening
low-carbon steels and cases where improvements in wear properties and maximization of
compressive surface residual stresses are the primary goals may serve as examples where
tempering might not be specified.
Nevertheless, the requirements of having untempered martensitic structures are more
the exception than the rule, because the as-quenched martensite is typically too brittle for
the majority of commercial uses because it promotes notch sensitivity and crack develop­
ment. Untempered martensite is also characterized by a high level of internal residual
stresses, which may be relieved during a service life resulting in shape distortion.
When making a decision as to whether tempering is required, consideration must be
given to more than just the mechanical properties [31]. If some hardened components are
left in the untempered condition, delayed cracking attributed to residual stress can occur
(see Section 4.4). Thus, it is important to minimize the time between the quench and tem­
per operations. Too long a delay or “transient time” may allow residual stresses to produce
cracking. Among other factors, the probability of delayed cracking occurring is dependent
on the hardness level, case pattern, chemical composition of the alloy (e.g., carbon content
in steels or CE in cast irons), as well as the geometry of the part.
Reheating of steel for tempering after hardening relaxes these stresses and improves
several other engineering characteristics, forming a tempered martensite microstructure.
Tempering is a form of subcritical heat treatment producing an attractive blend of micro­
structures and properties (see Section 2.1.2). Thus, tempering temperatures are always
below the lower transformation temperature Ac1. A desirable compromise between hard­
ness and strength versus toughness, ductility, and amount of residual stresses determines
the tempering process recipe/protocol. In some cases, tempering can also help improve
shape stability.
340 Handbook of Induction Heating

In induction hardening of some medium-carbon steels, the majority of high-carbon


steels, cast irons, and some alloy steels, there is a certain amount of RA present in the as-
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

quenched structure (unless cryogenic treatment is used). There is always a concern that
RA might transform to untempered martensite during the component’s service life intro­
ducing brittleness and potential shape distortion. Tempering helps decompose the RA.
Depending on the application specifics, tempering temperatures are commonly within
the range of 110°C (230°F) to 650°C (1200°F). The change in ductility may or may not be
significant with a low-temperature temper. However, an improvement of ductility is not
the only factor for tempering. Tempering can also affect the yield and ultimate strengths
as well as the fatigue life of a component. In other cases, tempering is carried out at higher
temperatures to enhance ductility and toughness to a considerable extent.
If the steel is heated to less than 110°C (230°F), then (practically speaking) there is no
change in microstructure that occurs as the result of short-time induction tempering. Low-
temperature tempering of carbon steels typically occurs at 120°C (248°F) to 250°C (482°F).
The hardness reduction in this case typically does not exceed 1 to 4 points HRC.
If high-temperature tempering is applied, the significant changes in microstructure
may lead to a measurable loss in hardness and a relaxation of the majority of the resid­
ual stresses. The hardness reduction may exceed 15–20 points HRC. Note that high-­
temperature tempering of alloyed steels may not result in a significant hardness loss.
It is imperative to mention here that for plain carbon and the majority of low-alloy
steels, an increase in tempering temperature results in a consequent reduction of the
hardness (Figure 4.147a). However, it might not be true for some alloy steels, where tem­
pering can cause an increase in the hardness of alloy steels with strong carbide formers
(chromium, molybdenum, vanadium, etc.) under certain conditions (as it is the case of
some high-speed tool steels). This phenomenon, known as secondary hardening (Figure
4.147b) [28–30], often occurs at approximately 500°C to 600°C (930°F to 1110°F), when alloy
carbides are precipitated from the solution [4]. Silicon could promote a hardness increase
at lower temperatures, approximately 315°C (600°F) [29].
Some steels may require double or triple tempering. This might be the case when there
is a concern that with the precipitation of alloy carbides, the Ms temperature may increase
and the RA will transform to martensite during cooling from the tempering temperature
[29]. Tempering of die steels and high-speed tool steels often specifies double or even triple
tempering.

%CSteel “A” > %CSteel “B” > %CSteel “C”

Steel “A” Steel “B” Alloy


steels
Steel “C”
Hardness

Hardness

Steel “D”

Steel “E”

Plain carbon steel Steel “F”

0 200 400 600 800 200 400 600 800


Tempering temperature, °C Tempering temperature, °C
(a) (b)

FIGURE 4.147
Hardness versus tempering temperature for plain carbon steels (a) versus some alloy steels with strong carbide
formers (b).
Heat Treatment by Induction 341

It is also important to remember that although a certain tempering recipe may produce
suitable properties in tension, it does not necessarily guarantee that the same improve­
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

ments will be recorded in torsion or in bending [31]. Therefore, it is imperative to evaluate a


heat-treated component under the intended working conditions to determine the benefits
of tempering and select a suitable process recipe. A review of QT properties of steels is
provided in various publications, including Refs. [2,4,28–31,33,154–156,391–393].
Basically, there are three ways to temper hardened components:

1. Self-tempering, where residual heat in the part is exploited


2. Induction tempering, where the hardened part is reheated via electromagnetic
induction
3. Furnace/oven tempering, where the part is heated by convection or thermal
radiation

As discussed in Section 4.1.3 and illustrated in Figure 4.11, martensitic transformation


occurs over a temperature range between the Ms (martensite start) and Mf (martensite
finish) temperatures. The range depends on the steel’s chemical composition and, from a
practical perspective, cannot be changed by varying the quench severity. In plain carbon
steels, the Ms and Mf temperature range is directly related to the carbon content. For plain
carbon steels with a carbon content of 0.2% to 0.5% C range, Ms temperatures are within
approximately a 300°C to 450°C range. Thus, freshly formed martensite will be imme­
diately exposed to tempering temperatures and can be potentially softened. This phe­
nomenon is referred to as auto-tempering. The degree of auto-tempering becomes more
noticeable with a reduction in quench severity and an increase of Ms temperatures and the
mass of the heated material. Alloy steels exhibit auto-tempering to a lesser degree com­
pared to plain carbon steels.
A conventional way of tempering is to run the parts through a tempering furnace/oven
(gas-fired, resistance, infrared, muffle tube, molten salt bath, or other means), which is a
well-accepted process that usually requires appreciable shop floor space, labor, and time.
Furnace tempering is a time-consuming process that often takes from 1 to 3 h. Shorter-time
higher-temperature induction tempering was developed to overcome these drawbacks.
One way or another, all microstructural changes that appear during tempering are associ­
ated with the decomposition of martensite into a structure that consists of an α-iron matrix
with a dispersion of carbide particles occurring as a result of a precipitation reaction. On
reheating of the as-quenched martensite, tempering takes place in different but overlapping
stages. Various scientific schools recognize a different number of tempering stages (typically,
from three to five stages). For example, R.W.K. Honeycomb and H.K.D.H. Bhadeshia suggest
that tempering of plain carbon steels takes place in four distinct but overlapping stages [154].
The metallurgical aspects and precipitation kinetics of tempering of different phases
(including martensite, bainite, etc.) of alloy steels is noticeably more complex compared
to plain carbon steels. Formations of transition carbides as well as the specifics associated
with complex metallic carbides are primarily responsible for that. Metallurgical kinetics
and metallography of tempering are outside the scope of this book and can be found in
numerous publications including Refs. [16,28–31,33,151–155,159,383–393].
It would be ideal if tempering would eliminate tensile residual stresses while retaining
beneficial surface compressive stresses. However, it is not possible because as discussed in
Section 4.4, there is always a balance of residual stresses (both compressive and tensile) in
the workpiece.
342 Handbook of Induction Heating

As stated earlier, surface compressive residual stresses are usually considered to be


beneficial. They provide protection against the propagation of cracks caused by micro­
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

scopic scratches and geometrical stress risers, delay fatigue cracking, and improve the
performance of parts that experience bending and torsional stresses during service (see
Section 4.4).
It is imperative to note, however, that the maximum tensile residual stress is located just
beneath the hardened case or within the hardness transition zone (Figure 4.144). This is a
zone of potential subsurface crack initiation. The maximum of applied stresses during ser­
vice is usually located at the part surface and then drops off. Therefore, one of the impor­
tant “duties” of tempering is not only the reduction of tensile stresses but also the shift
of the maximum of these stresses toward the core away from the applied tensile stresses.
Regardless of the fact that an increase in tempering temperature results in a mono­
tonic reduction of the hardness and strength in the majority of applications, a change
in impact toughness may not be monotonic with an increase of tempering temperature.
Embrittlement phenomena can occur after tempering at certain temperature ranges,
leading to a drop in impact toughness. There are several types of embrittlement that
can be associated with as-tempered structures. This includes but is not limited to the
following [30]:

• Tempered martensite embrittlement (TME) occurs when sufficiently slow tem­


pering within approximately the 200°C to 370°C range.
• Temper embrittlement (TE) occurs when sufficiently slow tempering within the
450°C to 600°C range.

As an example, Figure 4.148 shows impact toughness as a function of tempering tem­


perature of hardened low-alloy, medium-carbon steel [147].

Temperature, °C
93 204 316 427 538 649
100
Izod or V-notch Charpy at room temperature, ft-lb

80 0.40%C
V-notch Charpy

60

40

20 0.50%C
Izod

0
0 200 400 600 800 1000 1200 1400
Tempering temperature, °F

FIGURE 4.148
Impact toughness as a function of tempering temperature of hardened low-alloy medium-carbon steel. (From
M. Grossmann and E. Bain, Principles of Heat Treatment, ASM Int’l, Materials Park, OH, 1964.)
Heat Treatment by Induction 343

Besides the different temperature range, these two types of embrittlement (TME vs.
TE) are distinguished in several other ways. TE is a reversible phenomenon, while TME
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

is not. Once TME appears, there is no heat treatment to reverse the effect, other than to
re-austenitize and quench the steel, followed by tempering within an appropriate tem­
perature range. Longer tempering times have tendencies to aggravate embrittlement.
Highly alloyed steels are typically more prone to embrittlement. A detailed discussion on
these and other embrittlement phenomena can be found in a classical book authored by
Professor George Krauss [30].
Tempering can be done as either a batch or a continuous process [31]. The batch pro­
cess requires that parts be accumulated after hardening and then moved to the tempering
operation. In a continuous process, an in-line system moves from the hardening stage into
the tempering operation. Continuous processing can be very beneficial in eliminating or
dramatically reducing the probability of delayed cracking, because it minimizes the time
between hardening and tempering. The importance of minimizing the time delay was
illustrated using a case study discussed in Section 4.4.
Time and temperature are two of the most critical parameters in short-time induction
tempering. It has been shown experimentally that, under certain conditions, shorter-time
higher-temperature tempering can provide the same hardness as longer-time lower-­
temperature tempering.
There are several ways to determine the time–temperature correlation between con­
ventional longer-time lower-temperature furnace/oven tempering and shorter-time
higher-temperature induction tempering. A considerable amount of work has gone into
establishing the time–temperature relationships that result in identical hardness values
for a variety of steels. Most of these correlations show that hardness is a logarithmic func­
tion in the form of a Larsen–Miller parameter, which suggests that it is the product of the
absolute tempering temperature times the sum of a constant and the logarithm of the
tempering time

Hardness is a function of: T × [C + log 10 (t)],


where T is the temperature, t is time, and C is a constant that depends on the alloy compo­
sition. This tempering relation was first proposed by Hollomon and Jaffe [381]. It has been
established that the value of the constant C is generally between 10 and 18 for steels.
Another parametric method for correlating equivalent time–temperature conditions,
which is very similar to that of Hollomon and Jaffe, was established by Grange and
Baughman [394]. Unlike the Hollomon–Jaffe correlation, however, Grange and Baughman
revealed that the parametric constant (C) can have a constant value of C = 18 regardless of
the alloy content for a variety of steels.
Tempering data were satisfactorily simulated using the parametric equation shown below:

P = [°F + 460][18 + log (t , in hours)] × 10−3


Refs. [4,28–31,145,381,382] provide more details with respect to applying these relation­
ships. As an example, Figure 4.149 shows the time–temperature correlation for tempering
medium-carbon steels using induction tempering versus oven/furnace tempering.
Although the determination of parameters for an induction tempering process might
look straightforward, it should be understood that the abovementioned correlations
can serve only as a rough estimation technique. Some of these factors are the chemical
344 Handbook of Induction Heating

235 290 350


Furnace time Furnace time Furnace time
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

Temperature of induction tempering, °C

Temperature of induction tempering, °C

Temperature of induction tempering, °C


1h 1h 1h
225 2h 280 2h 340 2h
3h 3h 3h

215 270 330

205 260 320

195 250 310


150°C 200°C 250°C
(302°F) (392°F) (482°F)
185 240 300
0 10 20 30 40 50 60 0 10 20 30 40 50 60 0 10 20 30 40 50 60
Duraton of induction Duraton of induction Duraton of induction
tempering, s tempering, s tempering, s

FIGURE 4.149
Time–temperature correlation for longer-time furnace tempering versus shorter-time tempering by induction.

composition of the steel, the microstructural subtleties before tempering, the morphology
of martensite, grain size, the presence of residual heat, hardness profile, and thermal his­
tory (heating rate, temperature gradients, and the cooling rate after induction tempering).
The last factor (thermal history) is often the most neglected when determining tempering
parameters.
Induction tempering is a continuous process of heating and subsequent cooling (soak­
ing). Therefore, tempering conditions are affected by both heating and cooling stages.
Although the maximum tempering temperature is achieved at the heating stage, the cool­
ing stage is typically much longer (unless water cooling is used immediately after heat­
ing). Because tempering is a function of time and temperature, both factors will affect the
tempering conditions.
One of the most extensive investigations of the effects of the thermal history (heating
rate, cooling rate, pick temperature, etc.) and hardenability on tempering response was
that conducted by Semiatin, Stutz, and Byrer [382]. It has been demonstrated that the spe­
cifics of thermal history are critical in defining the deviations in the effective tempering
parameter between the center and surface of induction-tempered cylinders. Case studies
of using subroutines for predicting the results of tempering developed by Semiatin and
coworkers and comparison with experimental data are provided in Refs. [4,12,382].
The time required for induction tempering of case-hardened parts is typically two to
four times that of induction hardening. There are some rare cases of tempering thin parts
or components with shallow case depth (e.g., less than 0.5 mm) where tempering time
might be approximately equal to the heat time for austenitization.

4.6.2 Self-Tempering (“Slack Quenching”)


The principle of self-tempering after induction hardening is illustrated using the
example shown in Figure 4.122, which shows the results of numerical computer model­
ing of induction surface hardening of a medium-carbon steel solid shaft (50 mm/2 in.
Heat Treatment by Induction 345

diameter) in a normalized condition using a frequency of 16 kHz. The required nominal


case depth is 2.5 mm.
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

During the initial stage of induction heating, intensive heating of the surface and near
surface takes place. After 3.5 s of heating, the surface and the 2.5-mm-thick subsurface
layer required to be hardened have reached suitable temperatures for austenitization tak­
ing into consideration the nonequilibrium phase transformation caused by rapid heating
(see Section 4.1.6). A short dwell (0.5 s) is applied to reduce the thermal shock during the
initial stage of quenching.
The temperature at the center of the part does not increase significantly (less than 100°C)
during the heating and dwell cycles. Several reasons are responsible for that, including
pronounced skin effect, high power density, and short heating time, which do not permit
the greater amount of heat to be conducted from the surface toward the core.
During the initial quenching stage, the high temperature of the surface layer begins to
fall rapidly. After 2 s of spray quenching, the surface temperature is drastically reduced,
which is approximately 210°C (410°F). The maximum temperature reaching approximately
400°C (752°F) will be located at approximately 10–12 mm beneath the surface. Note that
the temperature at the center of the part continues rising during the first 6 s of quenching.
After 6 s of quenching, the surface temperature has decreased below 100°C (212°F); how­
ever, a considerable amount of heat is retained in the interior of the cylinder (the tempera­
ture at the core exceeds 300°C [572°F], with the average temperature being approximately
225°C [437°F]). If at this moment the supply of quenchant is cut off, the surface of the part
will begin to be heated again because of the heat that accumulated inside the workpiece.
After 5 s of soaking (heating power and quench are not applied), the surface temperature
rises to approximately 215°C (419°F) and the core temperature will be approximately 260°C
(500°F). After an additional 30 s, the difference between “surface-to-core” temperatures
will be almost undetectable. Therefore, with proper selection of quenching condition, this
retained heat can be used to temper the workpiece.
In many cases of using plain carbon and low-alloy steels for automotive applications,
the self-tempering temperatures (if applied) typically do not exceed 250°C (480°F) and are
usually in the 180°C (360°F) to 220°C (430°F) range.
Self-tempering provides several recognizable benefits [31]:

• It eliminates an additional operation owing to incorporation of self-tempering


into the hardening operation. Therefore, the capital equipment cost and total cycle
time are reduced, making it very attractive from a cost perspective.
• The time delay between hardening and tempering stages is virtually eliminated.
As discussed earlier, too long a time delay can be detrimental because of the
potential appearance of delayed cracking.
• Because self-tempering utilizes the residual heat that is retained after hardening,
there is no need to apply any additional energy for tempering, making it highly
energy efficient. A reduction in overall needed energy is associated not only with
the reheating stage but also with the fact that less energy is needed for the cooling
stage.
• There is obviously a savings in shop floor space, because there is no need for addi­
tional space to locate tempering equipment.

All these factors are very attractive and are the reasons for applying self-tempering
in some applications. However, several precautions must be taken to ensure that the
346 Handbook of Induction Heating

self-tempering process is performed correctly, and despite the considerable benefits, self-
tempering does have noticeable limitations, which restrict its broader use in industry and
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

make furnace tempering and induction tempering more popular processes. Some of those
limitations are outlined below [31]:

• Residual heat must be accurately controlled. The energy generated within the
part must be monitored closely to ensure that a constant amount of heat is pro­
duced. Modern technology allows monitoring energy levels with high precision.
However, it might be more challenging to accurately control the quench severity
with the same precision, in particular when heat treating complex geometry com­
ponents. Quench time, flow, temperature, concentration, and cleanliness should
also be monitored and held within close tolerances to ensure a consistent thermal
condition after quenching. It should be noted that the variations in quench sever­
ity are affected not only by actual conditions of the quenchant but also by the
surface condition of the workpiece, including surface roughness, presence of for­
eign residue, and so on (see Section 4.2.7). Though factors that are responsible for
potential deviations in cooling intensity as a result of the workpiece’s surface con­
ditions are always not desirable and should be minimized, they might not make
as dramatic an impact during the first two stages of quenching (vapor blanket and
nucleate boiling stages). However, they might produce a greater impact during the
third stage (convective cooling), leading to measurable variations of the residual
heat. This will inevitably negatively affect the repeatability of self-tempering. In
many cases, an infrared pyrometer is used to monitor the self-tempering tempera­
ture of the workpiece surface.
• If applicable, self-tempering can be used in static heating, single-shot heating, and,
to a lesser degree, horizontal scan hardening or continuous/progressive hard­
ening applications. It should not be used in vertical scan hardening, because of
unequal cooling conditions and variations of the accumulated residual heat in the
top and bottom regions of the vertically scan-hardened workpiece.
• It is easier to use self-tempering when dealing with simple geometries (e.g., such
as straight shafts). Geometrical irregularities may produce localized variations in
both the heating and the quenching intensities (particularly when dealing with
complex-geometry components) that might be sufficient to create too large a devi­
ation in the residual heat in self-tempering.
• Some steels and cast irons have relatively low Ms temperatures, and upon comple­
tion of the formation of the needed amount of martensite, there might be simply
an insufficient amount of retained heat accumulated within the workpiece for suf­
ficient self-tempering.
• It is more challenging to control residual heat when hardening components of
small sizes (e.g., wires, thin-walled tubing, small diameter rods, etc.). This makes
it easier to apply self-tempering in cases where there is sufficient mass. However,
large workpieces with an extremely large ratio of diameter-to-­thickness of the aus­
tenitized layer might also not be well suited for self-tempering, because the cold
core may provide such an overwhelming cold sink effect, eliminating the rise in
temperature of the hardened surface layer needed for self-tempering.
• Self-tempering should be avoided when profiled hardening is used (e.g., con­
tour hardening of gears and gear-like components). The variations in the
Heat Treatment by Induction 347

neighboring masses may substantially produce a nonuniform tempering effect.


The amount of heat stored as well as the heat sink of neighboring regions must
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

be the same or very similar; otherwise, the temperatures obtained during self-
tempering will be substantially different, resulting in an unacceptable temper­
ing structure.

The challenges discussed above prevent the wide use of self-tempering in industry, with
furnace/oven tempering and induction tempering becoming more popular choices.

4.6.3 Induction Tempering and Its Features


Induction tempering or furnace/oven tempering can be applied to those parts that cannot
be self-tempered. Space limitation does not permit a detailed discussion on the specifics
of furnace/oven tempering. An interested reader can review the following publications on
this subject: Refs. [16,28–30,33,151–155,353,354,383]. This section focuses on the specifics
of tempering using electromagnetic induction. Table 4.21 provides a brief comparison of
induction tempering versus oven/furnace tempering.
Some process features of induction tempering have already been outlined in Sections
2.1.2 and 4.6.1. There are several major factors to consider when applying induction
tempering:

1. Quantification of the purpose of tempering (e.g., what hardness range is needed


after induction tempering or what ductility is expected)
2. Alloy type and as-quenched hardness pattern as well as the expected distribution
of residual stresses after hardening
3. Geometry of the component (including shape and proximity factors), presence of
stress risers, nonuniform heat sinks, and specifics of tooling and fixtures
4. Anticipated heating mode, production rate and flow, presence of residual heat, and
its distribution
5. Residual magnetism requirements after induction tempering
6. Maximum permissible time delay between hardening and tempering operations
7. Frequency/power selection, and so on

Since tempering temperatures are always below the Curie point, the carbon steel is
always ferromagnetic. The primary mechanism of the heat generation in induction tem­
pering is associated with the heat generated by eddy current according to the Joule effect
(I2 R). The second mechanism of heat generation that occurs in any ferromagnetic material
is related to the hysteresis energy losses when energy is dissipated during the reversal of
magnetic domains (it does not exist at hardening temperatures). As discussed in Section
3.4.1, magnetic hysteresis losses are associated with magnetization–demagnetization
AC cycles that occur when ferromagnetic materials are exposed to an alternating mag­
netic field. Heat generation takes place as the loss of energy attributed to internal friction
between molecules [48]. Magnetic hysteresis heat generation is proportional to the applied
frequency and the area of the hysteresis loop, which is a complex function of chemical
composition, grain size, temperature, magnetic field intensity, and frequency.
The first mechanism of heat generation, eddy current losses (I2 R), has a greater impact
on overall heat generation during induction tempering compared to hysteresis losses,
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4
348

TABLE 4.21
Comparison of Induction Tempering versus Oven Tempering
Induction Tempering Oven Tempering
Temperature According to the Hollomon–Jaffe law and the Grange–Baughman According to the Hollomon–Jaffe law and the Grange–Baughman
correlation, it requires higher temperatures. correlation, it requires lower temperatures.
Time According to the Hollomon–Jaffe law and the Grange-Baughman Noticeably longer process times: typically 1.5–2 h.
correlation, it requires using shorter times.
Ergonomics Environmentally friendly, minimum of thermal exposure to Greater thermal exposure (“heat” wave) being absorbed by
operator, no “green” gas generation. environment and operator.
Required shop space Less shop floor space. Greater shop floor space and height are required.
Temperature uniformity Less uniform, since it primarily heats areas that are More uniform heating. Whole workpiece is always heated.
electromagnetically coupled with the induction coil.
Control Better controllability: Less controllability:
• Tempering temperature can be easily changed • Takes some time if there is a necessity to change tempering
• Allows using power pulsing (“heat–soak” cycles) temperature because of the considerable thermal inertia of
to enhance the heat flow. the furnace.
• Various regions of a hardened workpiece can be selectively
induction tempered.
Maintenance Less maintenance is required. More maintenance is required.
Readiness to be used Ready to use (immediate readiness of induction tempering Typically takes at least a couple of hours to heat up refractory of
equipment). No time required for start-up, resulting in oven from ambient to approximately 1050°F (e.g., after holidays,
minimum downtime. or after an interruption in a workflow—when working a single
shift). This also increases the amount of downtime.
Energy efficiency Very efficient. Less efficient.
1. Electrical efficiency when heating magnetic bodies (below 1. Lower efficiency owing to the need to hold oven’s
Curie temperature) is above 90%. temperature, since there is constant heat leaking through the
2. It does not waste energy for heating workpiece areas that are refractory and furnace doors.
not required to be heated. 2. Oven always heats the whole body of the workpiece regardless
3. Since no time is required for start-up, there is no need to of whether it is required or not.
waste energy and run an empty furnace just to be ready 3. Energy is wasted for nonproductive start-ups (also see
when parts will be available. “Readiness to be used”).
Handbook of Induction Heating
Heat Treatment by Induction 349

particularly when approaching the Curie temperature (A2 critical temperature). The impact
of hysteresis heating in tempering applications can be measurable and, among other fac­
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

tors, differentiates induction heating used for tempering versus hardening.


A progressive/continuous heating mode is typically used in tempering. In other
cases, the selection of the heating mode for induction tempering might be correlated
with the particular mode of hardening (e.g., single-shot, scanning, or static mode, see
Section 4.2.2).

4.6.3.1 Coil Design and Process Parameters


It is possible to harden and temper parts using the same coil and the same power supply.
In some cases, it is the best concept and has an obviously low capital cost advantage and
less tooling to store. In other cases, it might not best suit the customer’s requirements.
There are several reasons why, in the great majority of applications, the inductor designed
for hardening is not used for tempering.

• The power densities used during hardening are much higher than those used for
tempering. In tempering, it is necessary to heat the surface at a much lower rate to
obtain a “gentle” temperature gradient from the heated surface to the case depth.
Too high a power density could cause the surface temperature to exceed the opti­
mum tempering temperature, which would result in an unacceptably soft surface.
Depending on the type of power supply, load-matching restrictions might occur
when trying to run it at very low power levels (e.g., 2%–3%) using an inverter
designed for a hardening application.
• In induction hardening, to obtain the required hardness pattern for a workpiece
of complex shape, it is necessary to redistribute the EMF so as to introduce more
energy in certain areas. However, an optimal field distribution for hardening
might not be desirable for tempering. It is highly recommended that the temper
inductor should not heat only selected hardened area but a much larger region or
in some cases, even the entire workpiece.
• It is often preferable to use a lower frequency for tempering compared to hardening
because the tempering temperatures are always below the Curie point. Since the
workpiece retains its magnetic properties, the skin effect is pronounced. In addition,
the μr of steel during induction tempering is significantly higher than the magnetic
permeability of steel even during the initial stage of induction hardening. This takes
place because the power densities used in induction hardening are much higher
than the power densities used during tempering, resulting in much lower mag­
netic field intensities used for tempering (Figure 3.8a). Therefore, the μr during the
tempering stage is much higher than the μr during hardening for the same grade
of steel using the same frequency. This results in a noticeably smaller δ within the
workpiece for tempering as compared to the depth during any stage of harden­
ing. For example, during the magnetic stage of typical hardening applications (at
temperatures of 400°C to 500°C, the μr is typically within a range of 6 to 18). During
induction tempering, the μr at the same temperatures for the same grade of steel
heated using the same frequency is within a range of μr = 50 to 150 depending on
the application and can even be higher. Of course, at the final stage of hardening
(above the Curie point μr = 1), the differences in δ are even greater. Table 4.22 shows
typical δ in medium-carbon steel at room temperature (21°C) and high tempering
350 Handbook of Induction Heating

TABLE 4.22
Typical Current Penetration Depths in Carbon Steel for Room Temperature (20°C) and High
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

Tempering Temperature (620°C) and Frequency Compared to an Austenite Phase Temperature


900°C for Hardening
Current Penetration Depth (mm)
Frequency
Temperature (°C) 60 Hz 500 Hz 3 kHz 10 kHz 30 kHz
20 4.7 1.6 0.7 0.4 0.2
620 15.5 5.4 2.2 1.2 0.7
900 70 24.3 9.7 5.5 3.2
Source: V. Rudnev, R. Cook, D. Loveless, M. Black, Induction heat treatment: Basic principles, computation, coil
construction, and design consideration, Chapter 11A of the Steel Heat Treatment Handbook, G. Totten and
M. Howes (editors), Marcel Dekker, New York, 1997.

temperature (621°C) using five different frequencies (60 Hz, 500 Hz, 3 kHz, 10 kHz,
and 30 kHz) compared to δ at an austenite phase temperature (900°C). As can be
seen in Table 4.22, the “skin” effect will be very pronounced in the great majority
of tempering applications where δ represents a small fraction of the workpiece size
and the radius-to-δ ratio could be quite large. The use of higher frequencies further
increases this ratio, negatively affecting radial temperature uniformity. Therefore,
since both the temperature range and magnetic field intensity used in tempering
result in a much higher μr, compared to the respective values during hardening, it is
advantageous to apply substantially lower frequencies for induction tempering. The
typical inductor that is designed for hardening at a high frequency might perform
poorly in low-frequency tempering and vice versa.
• Unlike coils used for induction hardening, temper coils almost never require
the use of magnetic flux concentrators, because there is no essential gain in coil
efficiency in low-temperature tempering and there could be a decrease in effi­
ciency. Besides, flux concentrators might provide the needed heat distribution at
austenizing temperatures but promote the appearance of localized “hot spots” in
low-­temperature tempering.
• The time required for induction tempering is typically two to four times that of
hardening. Thus, the production and power supply utilization might suffer if the
same inductor were used for hardening and tempering.
• Heating for tempering with a specially designed coil and dedicated power sup­
ply is a more costly solution from a capital investment standpoint, but at the same
time, it has several noticeable advantages. The power density distribution can be
optimized specifically for the tempering operation. A separate, loosely coupled,
encircling or channel-type, single- or multiturn coils can be used effectively for
this purpose. Tempering is a diffusion-driven process; therefore, the time needed
to complete it might be noticeably longer than that for hardening. It may take sec­
onds to induction harden a part but tens of seconds or even minutes to induction
temper it. For high production: one high-power hardening machine can operate in
conjunction with two or three low-power tempering machines.

There is a common misconception that low-temperature tempering removes all internal


residual stresses. It does not. However, it relaxes some of those stresses. In surface-­hardened
Heat Treatment by Induction 351

steel or cast iron, the tensile stress maximum is typically located just beneath the hardened
case (Figure 4.144). Tempering not only reduces this maximum but also shifts the location
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

of the potentially dangerous, maximum tensile residual stresses farther away from the
surface and the maximum of applied load.
To induction temper complex-shaped parts, the choice of frequency, power density,
and coil geometry is dictated by the need to apply enough energy to certain areas of
the component. For example, in tempering of gears or large splines, the challenge is to
induce a sufficient amount of energy into the root of the tooth without overheating its
tip. The root is the critical area that is where the maximum concentrations of applied
stresses are typically located. As a result, fatigue cracks occur primarily in the root area.
This is why the root area is a primary target of tempering. However, several factors
complicate this task.

• One of them deals with poor electromagnetic coupling between an encircling-


type coil and the tooth root compared with its tip. This makes it more challenging
to induce the needed energy in the root area.
• The second factor deals with the existence of a significantly larger cold sink that
is located under the pitch circle and in particular under the tooth root compared
to its tip.
• The third factor derives from the fact that the tempering temperatures are always
below the Curie point. Therefore, the steel is magnetic and the skin effect is always
pronounced (Figure 3.8a), resulting in a power surplus in the tip of the tooth com­
pared to that in the root. In addition, the use of high frequency for induction tem­
pering will result in more heat sources being generated in the tip compared to the
root. Also, there will be a tendency to overheat edges, sharp corners, and so on. In
order to overcome these difficulties, low frequency, loose coil coupling, and low
power density are more suitable for tempering.

Over the years, in order to overcome the above mentioned challenges, IH manufacturers
have developed a variety of advanced induction tempering machines and designs.
Examples of induction tempering of general cylindrically shaped parts with I.D. harden­
ing and O.D. hardening are shown in Figure 4.150. The part shown in Figure 4.150a has
been hardened on a portion of its inside surface. This is the type of application in which
induction tempering is most effective. The tempering coil is positioned around the part, so
that the heat can be gently transferred from the outside surface toward the hardened (dark
color) layer on the inside. Proper energy/temperature control enables tempering without
overheating the hardened region of cross-groove disks (Figure 4.151).

(a) (b)

FIGURE 4.150
Examples of induction tempering of complex-shape parts with I.D. hardening and O.D. hardening.
352 Handbook of Induction Heating
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

FIGURE 4.151
Proper energy/temperature control enables tempering without overheating the hardened region of cross-
groove disks. (Courtesy of Inductoheat Inc., an Inductotherm Group company.)

A tempering coil of similar design can be applied to a hollow part that has been hard­
ened on its outside surface. In this case, however, it is preferable to position the coil inside
the workpiece if possible.
Unfortunately, in many cases, there is not sufficient space available inside the hollow
workpiece to position a tempering inductor, making it necessary to apply coils positioned
outside (externally) of the workpiece. This introduces some subtleties. As an example, the
spindle in Figure 4.150b requires a special tempering coil that will allow a predetermined
amount of energy to be induced into each hardened section, so that all sections are heated
to the same required tempering temperature. Because of the wide variation of the neigh­
boring mass near the case-hardened regions, an inductor designed to heat only the hard­
ened case (as was used for hardening) is not recommended. Because of the different heat
sink effects, the tempering inductor should also induce heat into areas that have not been
hardened, such as the flange. These areas will serve as a heat buffer. Note that sufficient
heat also must be transferred into the transition zone beneath the hardness case, where the
tensile residual stresses are at a maximum.
Because the temper inductor is loosely coupled to the workpiece, the heating of hard-to-
reach areas is facilitated without having excessive localized hot spots. Through heating of a
part without overheating its surface may also be achieved by adopting a gentle heat–soak cycle.
Figure 2.13a shows a channel inductor used to temper a variety of parts having complex
shapes such as CV joints, which are hardened both inside the bell and on the outside of the
shaft (Figure 2.7). In cases of induction tempering large-size components, they can be rotated
during the heat cycle, ensuring uniform heating while using channel-style inductors.
In other cases, complex designs may be used for induction tempering machines uti­
lizing different styles of inductors, which provide a 90° shift in current flow within the
workpiece, eliminating or dramatically reducing the effect of “dead” spots or “proximity”
heating surpluses.
A cooling cycle usually follows the completion of tempering. The cooling station can
be incorporated in the tempering machine, or it may be positioned separately on an exit
conveyor. The use of a separate cooling station reduces the complexity of the tempering
machine. The appropriate process parameters can be found by hardness measurements
and by microscopic evaluation of the as-tempered microstructure as well as conducting
residual stress measurements.
Heat Treatment by Induction 353

Because a ferromagnetic body naturally attracts a magnetic field, the coil-to-workpiece


proximity effect is typically not as pronounced as it is in the case of heating nonmagnetic
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

materials, dictating the development of certain process strategies when tempering parts
that consist of various diameters. This phenomenon is illustrated in Figure 4.152.
Multiple pins were progressively processed end to end through a multiturn solenoid
coil. Since tempering temperatures were noticeably below the Curie point, the electro­
magnetic proximity effect was not as pronounced, exposing minor differences in the
electromagnetic efficiency of both smaller-diameter and larger-diameter sections and gen­
erating an insignificantly different amount of eddy current in both sections. However, the
smaller-diameter section has a significantly smaller mass of metal. This results in a sub­
stantially lower temperature of the larger-diameter section. The difference in discoloration
of sections of pins with small and large diameters reveals the difference in the achieved
temperatures (note that there is no surface discoloration on the larger O.D. section), and
substantially higher temperatures were achieved on the smaller-diameter regions. A com­
parison of a discoloration of the pin on the left versus the pin on the right (Figure 4.152)
reveals that the greater difference between neighboring diameters produces greater tem­
perature surplus. This phenomenon can be reversed when heating nonferrous materials
using a certain frequency selection, making it possible to have a higher mean temperature
of the larger section or vice versa.
Aside from the factors discussed above, the appearance of electromagnetic end effect is quite
different in induction tempering compared to hardening. As stated in Section 3.1.7.1, end effect
manifests itself as a distortion of the EMF at both ends of the induction coil (Figure 3.35): the
extreme end of the cylinder (the so-called hot end) and the opposite end (the cold end).
The electromagnetic end effect at the extreme end (hot end) of the cylinder (Figure 3.35,
region “a”) manifests itself as heat deficit or heat surplus of the extreme end area, and it is
defined primarily by the following variables:

• Skin effect, R/δ


• Coil overhang, σ
• Ri/R and wall thickness-to-δ ratios
• Power density
• Presence of flux concentrator
• Space factor of coil turns Kspace (density of windings of coil turns)

where R is the radius of the heated cylinder, Ri is the inside coil radius, and δ is the current
penetration depth. The effects of frequency F and the electromagnetic physical properties
of steel (ρ and μr) are included in the skin effect ratio R/δ.
In static tempering using a conventional solenoid coil design, a measurable heat deficit at
the extreme end of the cylinder commonly occurs (specifics of electromagnetic end effect
of magnetic materials vs. nonmagnetic materials have been discussed in Section 3.1.7.1).

FIGURE 4.152
Difference in discoloration of sections of two progressively processed stepped pins with small and large diam­
eters reveals the difference in the achieved temperatures.
354 Handbook of Induction Heating

Experience shows that increasing the heating time in an attempt to promote thermal
conduction and to raise the temperature of the butt end of the cylinder does not noticeably
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

improve the temperature uniformity but reduces the production rate. The use of a higher
frequency can compensate for the heat deficit at the extreme end, but it also makes the
system more sensitive to variations in part position with respect to the induction coil and
worsens the radial temperature nonuniformity, which is critical in tempering applications.
Figure 4.153 shows several design concepts in the case of using solenoid coils that allow
improving the heat deficit at the extreme end of a ferromagnetic workpiece. An increased
coil overhang helps reduce the heat deficit, but there is a limit. First, upon increasing the coil
overhang, an underheating of the end will improve (Figure 4.153a). However, after reach­
ing a certain value of coil overhang, the temperature of the end might no longer improve.
The end might remain underheated regardless of the further increase in coil overhang,
and only a reduction of coil electrical efficiency and power factor will be recorded with no
change in temperature distribution.
One way to obtain a more uniform longitudinal heating pattern is to vary the radial gaps
along the coil length by reducing the “coil-to-workpiece” gap near the coil ends (Figure
4.153b) or using profiled inductors with various turn pitches, with tighter turn windings
near the coil ends and looser windings in the central region (Figure 4.153d). As an exam­
ple, Figure 4.154 illustrates achieving sufficiently uniform temperature distribution thanks
to a combination of appropriate process parameters (e.g., medium frequency, adequate coil
overhang and variable turn pitch) while tempering a carbon steel tube.
Another option is to use multilayer windings near the coil end (Figure 4.153e) and a
single-layer winding in the central region. This approach can be effectively applied only
when using sufficiently low frequencies (e.g., 50 to 200 Hz). Other design approaches are

Induction coil Axis of symmetry

(a) (d)

Pipe
Second layer

(b) (e)

Flux concentrator

(c) (f )

FIGURE 4.153
Several design concepts allow improving the heat deficit at the extreme end of a ferromagnetic workpiece.
(a) Conventional multi-turn coil, where a selection of a coil overhand allows to control heating of the workpiece
end area. (b) Application of coil turns with various inside diameters. (c) Application of various heating faces of
copper turns. (d) Application of tighter turn winding pitch in the coil end region. (e) The use of localized multi-
layer design. (f) Application of magnetic flux concentrator at coil end region.
Heat Treatment by Induction 355

817.6
900
743.7
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

669.7 800
700

Temperature
595.8
1 521.8 600
447.9 500
373.9 400
300.0
300
226.0
200
152.1
0 1 2 3 4 5 6 7 8
78.1
°F Distance from end (in.)

FIGURE 4.154
Uniform temperature distribution for high-temperature tempering application attributed to an appropriate
combination of frequency, power density, coil overhang, and variable turn pitch. (Courtesy of Inductoheat Inc.,
an Inductotherm Group company.)

also shown in Figure 4.153. These designs can help improve the end heating of ferromag­
netic materials; however, it will require dedicated coils for various workpiece sizes, driv­
ing up capital costs and reducing both flexibility and robustness.
Manufacturers of induction equipment have developed a number of unique tempering
designs optimizing heat uniformity for certain styles of components. One such apparatus
is referred to as Fluxmanager® Technology [357]. It represents a patented design of an
induction heater, providing an effective way of controlling the end effect and providing
the needed temperature uniformity for tempering/stress relieving of thick-walled carbon
steel pipes used in the manufacture of high-quality connections for oil country tubular
goods (see Section 2.1.2). This unique technology is discussed in detail in Section 4.6.3.2.
When tempering workpieces of general circular shape including disks or rings of apprecia­
ble size (e.g., wheel hubs) using oval solenoid-style coils or channel inductors, there might be
uneven heating along the circumference of the workpiece. An excessive localized heat genera­
tion could take place as a result of the electromagnetic proximity effect. A noticeable heat sur­
plus can occur in areas of the workpiece closest to the coil current-carrying copper at a greater
rate than in the areas located farther away and particularly at 90o from the closest regions
(Figure 4.155a). This phenomenon may result in a noticeable temperature nonuniformity,

Circular workpiece Channel inductor Friction guide rail Uniform heating pattern

Nonuniform heating pattern Workpiece pusher element


(a) (b)

FIGURE 4.155
Patented design concept helps to substantially improve the heat uniformity and the results of tempering of
workpieces of general circular shape. A comparison of circumferential temperature distribution when work­
piece of general circular shape proceeds through a conventionally designed coil arrangement (a) versus novel
patented technology (b). (From V. Rudnev, D. Loveless, Electric induction heat treatment of workpieces having
circular components, US Patent #9,060,390, June 16, 2015.)
356 Handbook of Induction Heating

causing undesirable tempering conditions along the circumference. Tempering nonunifor­


mity is aggravated with an increase in the diameter of the workpiece.
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

Rotation of the workpiece during induction heating has been recognized as a solution to
uneven heating while components are being processed through an oval coil or a channel
inductor. However, a conventionally designed apparatus for accomplishing the rotation
utilizing numerous motors or a specially designed chain driven-assembly carrying work­
pieces through an inductor dramatically increases the complexity and total equipment
cost as well as complicates the system’s maintainability.
A novel approach eliminates the necessity of using motors for part rotation and helps
to substantially improve the heat uniformity and the results of tempering associated with
it [402]. The sketch shown in Figure 4.155b illustrates the basic concept of this patented
technology. According to this approach, a circular workpiece is pushed through the induc­
tion coil arrangement (e.g., channel or oval inductors) by a pusher element positioned off
center relative to the circular workpiece. Thus, the pusher element introduces a force that
moves the workpiece linearly forward between the two guide rails and rotationally by
kinetic friction of the circular workpiece with the friction guide rail. Therefore, the circular
workpiece will be simultaneously moved forward and rotated during its heating without
utilizing any additional rotating devices.
In an alternative embodiment of the invention, the guide rails may be skewed relative to
each other from horizontal, increasing the kinetic friction force, which, in turn, enhances
rotation of the workpieces as it advances through the induction coil from the workpiece
coil entry and exit positions. The apparatus can optionally include a means for adjusting
the skewing angle to suit particular workpiece geometry specifics [402].

4.6.3.2 Flux Manager Technology for Stress Relief of Oil Country Tubular Goods
One of the specific induction tempering/stress-relieving applications is related to the man­
ufacturing of high-quality connections for oil country tubular goods [357,358]. After the
well is drilled, steel pipes (casing) are lowered into the hole and secured in place. This
provides a structural component of the wellbore. Casing is available in a range of sizes
and material grades and is usually manufactured from carbon steel that is heat treated to
obtain particular mechanical properties (e.g., varying strengths, API standard J55 has a
minimum yield of 55,000 psi; P-110 has a minimum yield of 110,000 psi) or a combination of
properties. For special applications, casing can be fabricated from stainless steel, titanium,
and other metals and metallic alloys exhibiting the needed properties [358].
An area above the reservoir is packed off’ inside the casing and connected to the surface
with a smaller-diameter pipe (tubing). The production string that provides the path to
bring the reservoir fluids to the surface is connected to components allowing control and
flow of oil from the well. Any mechanical failure of the pipe could result in the loss of the
drill string down the wellbore. Estimates of the cost to operate a drilling operation over
an offshore deep-water site could reach as much as $500,000 per day. Lost drilling time as
well as retrieval and environmental costs make connection integrity even more important
because a mechanical failure of the wellbore casing could cause leakage into groundwater
aquifers or other environmental catastrophes.
Taking into consideration the necessity of drilling deeper wells in combination with a harsh
drilling and operation environment, the providers of casing and tubing demand higher-
quality tubular products including providing heavier wall pipes and reliable connections.
Companies within the oil industry have developed special thread designs that are sup­
plied to oil and gas exploration companies. These are referred to as premium connections
Heat Treatment by Induction 357

[357]. These high-performance threads assure superior hydraulic sealing, improved tensile
capacity, and ease of makeup.
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

Premium connections are typically long tapered threads with separate sealing surfaces.
The female end is referred to as the box and the male end is called the pin. The manu­
facture of the connection begins with a straight length of pipe approximately 9–10 m in
length. The ends of the pipe are inserted into an upsetting press and mechanically formed
to the rough taper of the thread.
Upsetting creates changes in the grain structure of the metal in the area of the upset.
Following the upsetting operation and in order to improve mechanical properties and
prevent failure of the connection, the formed ends of the pipe undergo stress relieving.
The stress relief operation (also called high-temperature tempering or subcritical annealing)
is an important step in the manufacture of a quality connection. Improper heat treatment
could result in several undesirable phenomena from total joint failure to a type of bimetal­
lic corrosion known as “ring-worm corrosion” that occurs in improperly stress-relieved or
normalized pipes. This corrosion takes the form of a ring around the pipe usually located
a few inches up from the pipe upset.
Stress relieving is typically done before machining of the thread. In order to achieve the
best stress relief, the upset end must be uniformly heated along the entire swage length as
well as through the wall thickness of the pipe. Superior axial, radial, and circumferential
temperature uniformity is imperative for manufacturing quality tubular goods.
A modern connection manufacturer can have as many as 250 different pipe diameter and
wall thickness combinations to thread. Pipe size diameters can vary from 2 3/4 in. (70 mm)
to 18 5/8 in. (473 mm) with wall thicknesses ranging from 0.250 in. (6.4 mm) to 1.250 in. (32
mm) or even greater. In addition, the swage length can vary from 5 in. (127 mm) to nearly
18 in. (457 mm) depending on the pipe diameter, wall thickness, and application specifics.
The ability to provide uniform heating, high quality, and cost-effectiveness with induc­
tion heating machinery has been the traditional key benchmark deliverables in the past.
Today, in many industries, these three are joined by two additional requirements that are
equally important: flexibility and robustness. The flexibility of machinery reflects its abil­
ity to process a wide variety of parts without compromising the quality of the product.
Robustness reflects sensitivity of industrial machinery to withstand real-life disturbances/​
tolerances, ensuring heating quality not only under ordinary working conditions but also
under extraordinary or unexpected conditions. For example, taking into consideration the
physical size (a combination of diameters, wall thicknesses, and lengths), it is not unusual
to have nonsymmetrical (to some degree) positioning of the statically heated pipe inside
the induction coil. Therefore, the ability to reduce downtime after changeover of induction
coils while processing different products is important.
Induction heating for stress relieving and tempering of pipe ends is generally accom­
plished by placing the end of the pipe into an appropriately designed multiturn induction
coil where it statically heats for a specified amount of time reaching the target temperatures.
Customers typically specify production rate, final temperature, required heated length
(RHL) at the pipe end, and allowable maximum temperature deviations: radial (±ΔTR),
circumferential (±ΔTC), and longitudinal (±ΔTL). Some applications call for a sharp longi­
tudinal transition zone while others require certain axial temperature runouts; therefore,
the thermal gradient within the “hot-to-cold” longitudinal transition zone might be also
specified.
In stress relieving/tempering of oil country tubular steel goods and depending on
the particular steel grades and application specifics, the typical temperature uniformity
requirements are ΔT = ±18°C at target temperature levels of 500°C to 650°C. However,
358 Handbook of Induction Heating

when heating a variety of pipe sizes, the actual achieved temperature uniformity in most
conventional induction systems may noticeably exceed the needed uniformity.
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

4.6.3.2.1 Achieving Radial Heat Uniformity


As can be seen from Table 4.22, when heating thick-wall pipes, the “skin” effect can be
very pronounced, being a fraction of the pipe’s wall thickness, particularly during the ini­
tial and intermediate stages of heating for high-temperature tempering and the pipe wall
thickness–to–δ ratio could be quite large. The use of higher frequencies further increases
this ratio, negatively affecting radial temperature uniformity. For example, the average
pipe wall thickness–to–δ ratio can easily exceed 10-fold when heating pipes with wall
thicknesses greater than 16 mm (5/8 in.), utilizing a frequency of 1 kHz. With 10 kHz, this
ratio can exceed 30-fold.
For a given power, the temperature distribution across the wall thickness of a carbon
steel pipe is primarily a function of the applied frequency and the heat time. Frequencies
of 500 Hz and higher result in a greatly pronounced surface heating effect, generating the
majority of the induced power within a fraction of the pipe’s wall. This produces notice­
able radial thermal gradients when heating heavy-wall pipes requiring pulse heating (e.g.,
using multiple heat on–heat off cycles) as an attempt to avoid surface overheating while
providing sufficient heating of the pipe’s internal regions and its inside diameter. The
necessity of using multiple heat-soak pulses of extended lengths increases the complexity
and reduces the production rate; still, there is a danger of overtempering the outside diam­
eter of heavy-wall pipes while trying to provide sufficient heating of the inside diameter
while utilizing frequencies of 500 Hz and greater. An attempt to increase the production
rate by shortening the heat times suppresses thermal conduction (less time for O.D.-to-I.D.
temperature equalization), worsening the heat nonuniformity and leading to significantly
greater radial thermal gradients, particularly when heating heavy-wall pipes.
In contrast, the use of line frequency (50–60 Hz) is associated with a significantly deeper
heating effect as can be seen from Table 4.22. Deeper heating naturally helps provide more
intense subsurface heat flow and, therefore, more uniform radial temperature distribution
across the wall thickness. Therefore, as expected, the O.D.-to-I.D. thermal gradient will be
substantially reduced when applying line frequency. Table 4.23 shows a comparison of the
wall thickness–to–δ ratio for typical pipe wall thicknesses at different temperatures and
various frequencies based on the data of Table 4.22.

TABLE 4.23
Wall Thickness–to–δ Ratios for Typical Pipe Wall Thicknesses and Temperatures Using Different
Frequencies
Frequency
Pipe wall, mm (in.) Temperature, °C (°F) 60 Hz 500 Hz 3 kHz
12.7 (1/2) 21 (70) 2.7 7.8 19
621 (1150) 0.8 2.4 5.8
16 (5/8) 21 (70) 3.4 9.8 23.9
621 (1150) 1.03 3 7.3
19 (3/4) 21 (70) 4 11.7 28.4
621 (1150) 1.23 3.5 8.6
25.4 (1) 21 (70) 5.4 15.6 37.9
621 (1150) 1.64 4.7 11.5
Heat Treatment by Induction 359

For example, in the case of heating 19-mm (¾-in.) wall pipes using 500 Hz, the wall-to-δ
ratio changes from 11.7 (initial stage of heating) to 3.5 (at 621°C). This means that practi­
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

cally all of the induced power will be generated in the surface layer that occupies only
 1 
8.5%  × 100% = 8.55%  of the pipe wall thickness during the initial stage of heating
 11.7 
 1 
and 28.6%  × 100% = 28.6%  at 621°C. Note that the condition worsens since there is a
 3.5 
nonuniform (an exponential) distribution of the heat sources below the Curie point even
within the δ (Figures 3.54, top left, and 4.90, top right).
The use of higher frequencies will make these ratios even worse. In contrast, the corre­
1   1 
sponding values at 60 Hz are 25%  × 100% = 25%  and 81.3%  × 100% = 81.3%  ,
4   1.23 
which means that the heating is significantly deeper.
As the thickness of the pipe increases (e.g., when heating pipes with thicker walls), the
O.D.-to-I.D. temperature nonuniformity (ΔTR) will rapidly worsen if frequencies of 500 Hz
and higher are used regardless of having relatively long heat-soak time as an attempt to
facilitate more gentle heating.
As expected, with 50–60 Hz, the O.D.–to–I.D temperature uniformity will be substan­
tially improved because of the more in-depth nature of the heat generation.

4.6.3.2.2 Achieving Longitudinal (Axial) Heat Uniformity


As discussed in Sections 3.1.7.1 and 4.6.3.1, a uniform power density (heat source) distribu­
tion along the extreme (hot) end of the pipe does not correspond to a uniform temperature
distribution because of the additional heat losses (owing to thermal radiation and heat
convection) at the pipe’s hot end compared to its central region. Proper selection of coil
design and process parameters can help compensate for the additional surface heat losses
at the end of the pipe by generating a desirable power surplus properly utilizing the elec­
tromagnetic end effect. This allows one to obtain a satisfactory uniform temperature dis­
tribution within the RHL of the pipe.
Another important feature that affects the required coil length is the fact that in zone
“b” (Figure 3.35), which is sometimes defined as the HAZ or axial transition zone, there is
a considerable longitudinal temperature gradient that results in a corresponding heat flow
owing to thermal conduction from the pipe’s high-temperature region toward its colder
area manifesting itself as a “cold sink” phenomenon.
Note that the extreme end of the pipe (Figure 3.35, zone “a”) offers no axial thermal path
for the heat conduction in contrast to its cold end (zone “b”) that provides a ready thermal
sink and an intense heat conduction path. An extended HAZ is directly associated with
the additional energy.

4.6.3.2.3 Conventional Induction Heating Technology Utilizing Line Frequency (50–60 Hz)


The use of line frequency was the earliest technique for the heating of carbon steel pipe
ends. The larger δ associated with the application of 50–60 Hz has provided the best radial
temperature uniformity across the pipe’s wall thickness (ΔTR). Unfortunately, the use of
line frequency in combination with conventionally designed coils has been associated
with an often inadequate axial (longitudinal) temperature distribution (ΔTL) attributed to
poor controllability of the electromagnetic end effects resulting in a heat deficit at the pipe
extreme end.
360 Handbook of Induction Heating

“Bell-shaped” temperature profiles (Figure 4.156a) are quite typical when using conven­
tionally designed inductors powered by line frequency for end heating of magnetic pipes
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

below the Curie temperature. Note that temperature control nodes #1 and #2 correspond
to the temperatures near the extreme butt end of the pipe (hot end). Control nodes #3 and
#4 represent temperatures in the middle of the RHL and near the “cold” end, respectively.
The highest temperature was observed in the middle of the RHL with its gradual reduc­
tion toward the butt end of the pipe (“hot” end) as well as toward its “cold” end. Therefore,
the longitudinal temperature distribution resembles the shape of a bell. The thermal image
reveals that the surface temperature in the middle of the RHL was approximately 980°F
(527°C), but the temperature at the pipe’s butt end was only approximately 810°F (432°C).
Heat deficit at the pipe’s butt end has been associated with the limited controllability
of conventionally designed coils to compensate for the electromagnetic end effect that is
associated with the inability to balance the deficit of the induced power at the pipe end
area regardless of substantial coil overhangs. In this particular case, the coil overhang σ at
the extreme end exceeded 4 in. (102 mm) but still fell short of generating sufficient heat at
the pipe’s butt end. Note that the necessity of having very large coil overhangs is always
accompanied by such undesirable phenomena as reduced electrical efficiency and poor
power factor, as well as increased vibration and audible noise.
Practice shows that an attempt to increase the heating time with the hope that the ther­
mal conduction can produce a sufficient longitudinal heat flow and equalize the tempera­
ture in the axial direction does not noticeably improve the heat uniformity.

1000 “Bell”-shaped hot profile


970
940
910
4 3 2 1
880
850 5
820
790
760
730
700
(a) ˚F
1000
970
940
910
880 4 3 2 1
5
850
820
790
6
760
730
700 Local “hot” spot
(b) ˚F

FIGURE 4.156
“Bell-shaped” temperature profiles (a) and the appearance of localized “hot” (b) and “cold” spots are quite typi­
cal when using conventionally designed coils. (From P. Ross, V. Rudnev, R. Gallik, G. Elliott, Innovative induc­
tion heating of oil country tubular goods, Industrial Heating, May, 2008, pp. 67–72.)
Heat Treatment by Induction 361

Figure 4.154 shows several coil designs that allow improvement in pipe end heating that
have been discussed earlier. All these approaches necessitate having a dedicated coil for
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

a particular pipe geometry, which would significantly increase the number of required
coil sets for heating pipes of different diameters, wall thicknesses, and heated lengths.
Such an approach would not only appreciably increase the equipment capital cost but also
negatively affect the overall system flexibility and robustness. Equipment flexibility is
extremely important because most customers prefer utilizing a minimum number of coil
sets for heating a variety of different pipe sizes.

4.6.3.2.4 Call for Higher Frequencies (500 kHz to 10 kHz)


The use of higher than 50- to 60-Hz frequencies represents another attempt to improve
longitudinal temperature uniformity at the pipe end area and product quality. In the past,
there have been a number of successful induction pipe heating systems, utilizing frequen­
cies in the range of 500 Hz to 10 kHz supplied to industry. Recent trends to increase pipe
wall thicknesses in combination with tighter requirements for heat uniformity, however,
have outlined several drawbacks of using frequencies of this range versus line frequency
when heating thick-wall magnetic steel pipes to temperatures suitable for stress relieving
and tempering. These include the following:

1. When heating thick-wall pipes (e.g., 15-mm wall), as can be seen in Table 4.22,
even with a frequency of 500 Hz, the “skin” effect is very pronounced during the
majority of the heat cycle and the wall thickness–to–δ ratio is quite large. Higher
frequencies further increase this ratio. For example, with frequencies 3 kHz and
higher, the wall thickness–to–δ ratio can exceed 20, meaning that practically all of
the induced power will be concentrated within less than 5% of the pipe wall thick­
ness. Such a pronounced surface heating effect is directly related to the danger
of overheating the outside surface while trying to obtain a uniform radial tem­
perature distribution and to provide a sufficient heat flow toward the pipe’s inside
diameter. Radial temperature nonuniformity could result in undesirable heteroge­
neous stress-relieving/tempering properties.
2. Higher frequencies are noticeably more sensitive to asymmetric pipe positioning
inside the induction coil, meaning that even slight variation in pipe positioning,
coil overhangs, or “coil-to-pipe” radial gaps could lead to measurable temperature
variations at the pipe end area because of the electromagnetic proximity effect.
This worsens process repeatability when using frequencies of 500 Hz and higher,
negatively affecting controllability and robustness. Difficulty in controlling the
proximity effect is typically associated with the appearance of localized “hot” and
“cold” spots (Figure 4.156b), which occur when the pipe is not positioned perfectly
symmetrically inside the induction coil. At the same time, asymmetric pipe posi­
tioning (to some degree) is quite common when various diameter pipes are heated
in the same inductor or because of abnormal pipe loading.
3. Higher frequencies are “less forgiving” for variations of “coil-to-pipe” electromag­
netic coupling (e.g., radial air gaps), which is also the concern when heating pipes
with different tapered connections (“pin”- and “box”-type tapered ends).

FEA computer modeling provides important insights regarding the sensitivity of induc­
tion pipe heating systems using medium and high frequencies [403]. Figure 4.157 shows a
362 Handbook of Induction Heating

120 mm
Inductor
25 mm
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

Pipe

#2 #4
#1
#3

FIGURE 4.157
Sketch of a simulated induction system. Pipe O.D. = 178 mm (7 in.), wall = 9.25 mm (3/8 in.). Required heated
length is 120 mm (4.75 in.). Total pipe length is 9 m. Frequency is 3 kHz. Required heating time is 30 s.

sketch of a simulated induction system. Pipe O.D. = 178 mm (7 in.), wall = 9.25 mm (3/8 in.).
RHL is 120 mm (4.75 in.). Total pipe length is 9 m. Frequency is 3 kHz. Heating time is 30 s.
First, it was assumed that there was perfectly accurate pipe positioning inside the induc­
tor. Longitudinal temperature variations along outside and inside diameters at the end of
heating were ΔTL = ±22°C and approximately ±25°C, respectively. However, the overall
temperature nonuniformity within RHL was ΔT = ±58°C owing to the presence of an
appreciable radial temperature gradient within the pipe’s wall. Though a 9.25-mm wall
might be considered to be relatively thin, 3 kHz is sufficient to produce an unacceptable
temperature distribution after 30 s of heating, and considerably longer times are needed
to assist radial heat flow.
It is important to recognize a wave-like temperature distribution at the pipe end area.
Heat surplus occurs at the pipe’s butt end. There is also a region with a heat deficit (that is
approximately 15–30 mm long) located immediately behind the surplus region.
The effect of coil overhang σ variation within ±6 mm (±¼ in.) was investigated [403].
Because the coil current in all cases was kept the same, the temperatures of the O.D. and
I.D. in the middle of the heated region (0.1 m from the pipe’s butt end) were about the
same.
Regardless of the fact that ±6 mm (±¼ in.) coil overhang variation might seem to some
people to be relatively insignificant (particularly taking into consideration the pipe physi­
cal size), it is sufficient to produce a significant temperature variation that primarily occurs
in the pipe extreme end region. A coil overhang increase compared to the base value pro­
duces noticeable end overheating while its reduction results in an appreciable heat deficit
there. Table 4.24 summarizes the temperature variations of the control nodes. Similar pro­
cess sensitivity takes place when using 1 kHz.

TABLE 4.24
Temperatures of Control after 30 s of Heating versus Coil Overhang Variations [403]
Variation of Coil Temperature of Control Nodes after 30 s of Heating
Overhang
Compared to Its
Base Value Node #1 Node #2 Node #3 Node #4
0 683°C (1261°F) 651°C (1204°F) 580°C (1076°F) 664°C (1227°F)
+6 mm (¼ in.) 730°C (1346°F) 673°C (1243°F) 600°C (1112°F) 657°C (1215°F)
−6 mm (¼ in.) 635°C (1175°F) 627°C (1161°F) 557°C (1035°F) 672°C (1242°F)
Note: Pipe O.D. = 178 mm (7 in.), wall = 9.25 mm (3/8 in.). Required heated length is 120 mm (4.75 in.). Frequency
is 3 kHz.
Heat Treatment by Induction 363

As can be seen in Table 4.24, the use of medium and high frequencies is associated with
extreme process sensitivity and poor overall robustness of the induction system. Even rela­
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

tively small deviations in pipe positioning inside an induction coil (including longitudinal
and radial positioning) can produce substantial temperature gradients: radial (±ΔTR), cir­
cumferential (±ΔTC), and longitudinal (±ΔTL), as well as localized “hot” and “cold” spots
that can dramatically exceed permissible thermal deviations.

4.6.3.2.5 The Use of a Faraday Ring (“Robber” Ring Effect)


While using medium- and high-frequency heating systems, some induction manufactur­
ers have made an attempt to balance nonuniform end effects by utilizing a Faraday ring.
According to this approach, a water-cooled ring that is typically fabricated from copper
(though some other metals and alloys can also be used) is located in proximity to the end
of the heated pipe. A copper ring “robs” power concentration at the pipe butt end and is
often called a “robber” ring.
The physics of this “flux robbing” phenomenon can be explained as follows: An alternat­
ing (changing) coil current (AC) produces in its surroundings an AC magnetic field. This
field will have the same frequency as the coil current. The AC magnetic field induces eddy
currents in the pipe and in other electrically conductive objects that are located near the
coil including the copper ring. Induced currents have the same frequency as the coil cur­
rent. However, their direction is the opposite of the coil current. The copper ring current
produces its own magnetic field, which has an opposite direction to the main magnetic
field of the coil. By locating a robber ring in proximity to the pipe end, a partial cancella­
tion or “robbing” effect of the main field can be achieved. Therefore, by selecting the pro­
cess parameters in such a way that it would normally lead to a heat surplus at the pipe end
and by properly adjusting a flux robber ring, it would be possible to achieve the needed
temperature there.
Unfortunately, the application of a “robber” ring is associated with several drawbacks;
some of them are related to the necessity of using frequencies of 500 Hz and higher [403].

1. The effect of “robber” ring remains very sensitive to pipe positioning (radial, lon­
gitudinal, and circumferential) inside the induction coil. It is critically important
to maintain uniform and consistent gaps (radial and axial) between the copper
ring and the pipe end. Such sensitivity negatively affects the system’s robustness
and can easily produce heat nonuniformity when the loading mechanism does not
provide an accuracy of pipe positioning better than ±3 mm (±1/8 in.).
2. The application of a robber ring is associated with a time-consuming skilled setup,
training, and special experience in the proper adjustment of positioning when
changing pipe sizes, and there is still the possibility for human error. Besides, it
is simply one more water-cooled and movable component that can fail. Extended
setup time negatively affects equipment uptime.
3. Additional power losses induced in a robber ring add no value to the end product.
This additional energy cost is compounded because not only is it not transmitted
to the work, but it must also be dissipated as waste thermal load on the cooling
system, which is then eventually exhausted into the atmosphere. These are unre­
coverable costs that translate into wasted dollars when considering that a machine
will run thousands of cycles per year.
4. One of the positive effects of using “robber” rings is related to a reduction of the
heat nonuniformity that is caused by a coil helix. Unfortunately, the application of
364 Handbook of Induction Heating

“robber rings” can only reduce but does not eliminate the effect of a helix leading
to localized temperature nonuniformity.
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

5. A number of “robber” rings having different diameters is typically required to


cover the range of heated pipe sizes.
6. The use of frequencies of 500 Hz and higher is inevitably related to an increased
radial (O.D.-to-I.D.) temperature nonuniformity when heating thick-wall pipes
that is produced by the pronounced surface heating effect. This means that practi­
cally all induced power will be generated in the surface layer that occupies a frac­
tion of the pipe wall thickness. Therefore, the ability to achieve the required heat
uniformity within the pipe’s wall thickness relies greatly on thermal conductivity.
The necessity of stress relieving heavy-wall pipes and the desire to increase pro­
duction rates by shortening the heat times suppress thermal conduction, worsen
radial heat nonuniformity, and lead to measurable thermal gradients across the
wall thickness. All of these factors raise a concern with respect to ending up with
metallurgically undesirable heterogeneous structures and hardness distribution,
compromising the quality of the stress-relieved/tempered products.
7. Eddy currents induced in robber rings inevitably produce magnetic forces lead­
ing to vibration and audible noise, which greatly affects the overall noise level
and could lead to a measurable emitting of resonant sound waves. It is widely
accepted that the audible noise generated by frequencies of approximately 1 kHz
range is considered to be one of the most unpleasant noises for the human ear.
Therefore, it is highly desirable to eliminate a number of flexible components
by replacing them with rigid and robust designs. With time, excessive vibration
can also produce leaks in water-cooled robber rings. High maintenance costs
and wearing out of the robber rings, flexible cables, and moving mechanisms are
always a concern.
8. In recent years, several international organizations have raised concerns related to
external EMF exposure, developing awareness regarding nonionizing radiation
and evaluation of the health risks associated with external EMF exposure. Some
of these organizations are the Institute of Electrical and Electronic Engineers
(IEEE), the International Radio Protection Association (IRPA), the World Health
Organization (WHO), the Occupational Safety & Health Administration (OSHA),
and others. Studies have been conducted to evaluate the direct and indirect effects
of EMF exposure on health, passive and active medical implants, hypersensitiv­
ity, and so on. A number of international standards, guidelines, and regulations
have been put in force [406]. The use of bare coils is unavoidably associated with
an increase in external magnetic field exposure. Bare coils with bigger diameters
and larger coil-to-pipe radial gaps are prone to have a greater magnitude of exter­
nal EMF exposure. Therefore, measures should be taken to reduce EMF exposure,
ensuring that its magnitude does not exceed the maximum recommended levels.

4.6.3.2.6 Fluxmanager Technology
Fluxmanager is a patented technology (Figure 4.158a) that was developed [357,358] to
improve the shortcomings of conventional tempering/stress relief processes that rely on
using frequencies of 500 Hz and higher. At the same time, this technology dramatically
improves the drawbacks related to the heat deficit at the pipe end regions associated with
Heat Treatment by Induction 365
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

1000.00
970.00
940.00 7
910.00
880.00 5 4 3 2 1
850.00 6
820.00
790.00
760.00
730.00
700.00
°F
(b)

(a)
Thermal image of heated end of the steel pipe, 194 mm (7-5/8 in.) diameter; 19 mm (¾ in.) wall utilizing Fluxmanager
technology. Required heated length is 382 mm (15 in.). Heat time – 93 s. Coil power is about 100 kW/60 Hz.

FIGURE 4.158
Fluxmanager is a patented technology that was developed [357,358] to improve the shortcomings of con­
ventional tempering/stress relief processes (a) and the thermal image of the heated end of the steel pipe,
194 mm (7 5/8 in.) diameter; 19 mm (¾ in.) wall utilizing this technology (b). (Courtesy of Inductoheat Inc., an
Inductotherm Group company.)

using conventional line frequency (50–60 Hz) induction heaters. These advantages have
been achieved by addressing the following seven factors:

• Improved equipment flexibility, repeatability, and robustness with reduced sensi­


tivity to real-life process disturbances, tolerances, and imperfections.
• Induction heater components have a rigid and reliable design.
• Elimination of the use of “robber” rings of any kind.
• Ensuring superior radial, circumferential, and longitudinal temperature unifor­
mity, resulting in improved quality of stress relieving/tempering of magnetic
steel pipes. There is a significant reduction of the probability of excessive tempera­
ture surplus on the outside diameter while heating thick-wall pipes.
• Elimination of temperature nonuniformity caused by a coil helix.
• Dramatic reduction of the external magnetic field around the induction coil. FEA
computer modeling and field tests reveal that EMF exposures are kept well below
maximum levels recommended by the international professional organizations
including IEEE, WHO, IRPA, OSHA, and others.
• Total cost reduction.

As an example, Figure 4.158b shows the thermal image of the heated end of the steel pipe
(194 mm [7 5/8 in.] diameter; 19 mm [¾ in.] wall) utilizing Fluxmanager technology. RHL
is 382 mm (15 in.). Heat time is 93 s. Average coil power was approximately 100 kW/60 Hz.
ΔT= ±18°C (±32°F). Control node #1 corresponds to the temperature near the butt end of
366 Handbook of Induction Heating

the pipe (“hot” end). Control nodes #2 through #4 correspond to temperatures along the
RHL. Nodes #5 and #6 represent the extreme end of the RHL located near the “cold” end.
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

Fluxmanager Technology consists of several distinctive design features, including the


following:

• The application of line frequency (50–60 Hz) provides a significantly more in-depth
heating effect (as discussed earlier), allowing dramatically improved temperature
uniformity within the pipe wall thickness, particularly for heating heavy-wall
pipes, and short cycle times. Line frequency inductors are much less sensitive and
more robust compared to medium- and high-frequency heaters.
• In order to eliminate the effect of the coil helix, a Fluxmanager induction coil con­
sists of a number of two-layer disc-pair copper windings, connected electrically
in series forming several coil sections. Such a design has a threefold advantage.
First of all, it allows for keeping the coil turn windings straight, which eliminates
the negative effect of the coil helix (Figures 3.42 and 3.43, left) on temperature
uniformity. Second, depending on application specifics, it allows distributing coil
sections in an optimum way, improving the longitudinal temperature uniformity
of the heated pipe. Third, one design concept allows connecting the Fluxmanager
directly to a step-down 240-V outlet without the necessity of having a solid-state
inverter. A simple, reliable, and cost-effective SCR control system is used to control
power. A set of capacitors adjusts to the lack of reactive power, improving the coil
power factor. As an optional design, a system includes a low-frequency inverter
(up to 150 Hz) to improve electrical efficiency when heating smaller-diameter
pipes (less than 90 mm diameter).
• Unlike other induction heating systems, Fluxmanager utilizes a proprietary
designed flux concentrator that boosts power density at the coil end regions and
can be adjusted to accommodate different pipe diameters, walls, or the “pin” and
“box” geometry, resulting in superior process flexibility and temperature unifor­
mity. A stack of laminated steel is used to fabricate a magnetic flux concentrator,
helping to make an entire structure very rigid and robust. It also protects coil
copper windings from accidental impact and excessive wear during abnormal
pipe loading–unloading operations. Several flux concentrator designs were devel­
oped. A flux concentrator also positively affects the EMF distribution in the “hot-
to-cold” transition. The transition zone, the area at the back of the heated length
above the swage, typically exhibits a steep temperature gradient; evidence that the
energy consumed is focused at the work area and is primarily applied where it is
required. If it is desirable, the length of this zone can be adjusted.

Computer simulation is an ideal tool to explain the principle of the Fluxmanager


Technology. Because of the system’s symmetry, only the top half of the pipe and induc­
tor is shown in Figure 4.159, while explaining the physics of the process and demon­
strating the capabilities of the Fluxmanager Technology by analyzing the magnetic field
distribution.
As it is known from college physics, the degree of electromagnetic coupling between the
heated workpiece and the magnetic field of the coil is determined by the number of imagi­
nary magnetic flux lines that enter the workpiece (e.g., pipe). More flux lines intercepted by
the pipe results in higher efficiency and greater induced heat sources.
Heat Treatment by Induction 367
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

Pipe
Pipe

10¼ in. (260 mm) O.D 6.7 in. (170 mm) O.D
Axis of symmetry ¾ in. (19 mm) wall Axis of symmetry ¾ in. (19 mm) wall

FIGURE 4.159
Results of computer modeling demonstrating the capabilities of the Fluxmanager Technology. (From V. Rudnev,
Flux Manager Technology, unpublished presentation for the Tenaris Group, 2013.)

Flux leakage refers to the magnetic flux that does not thread through the heated mate­
rial. The degree of flux leakage depends, in part, on the electrical gap between the current-
carrying coil surface and the pipe. This gap is of greater concern as frequency increases,
because flux lines tend to stay closer to the surface of coil conductors as frequency is
increased. That is one of the reasons why line frequency is more forgiving to an air gap
variation compared to medium and (especially) high frequencies. Thus, the same geomet­
rical variations (e.g., caused by tolerances or pipe positioning inside the coil) will produce
less temperature variations when lower frequencies are used.
The use of flux concentrators helps reduce the external flux leakage assisting to keep the
coil field intact, reduce external magnetic field exposure, and improve effectiveness (see
Section 4.7).
The specifics of the Fluxmanager Technology calls for the use of laminations. Laminations
are punched out of grain-oriented magnetic alloys (Ni–Fe and Si–Fe alloys). Stacks of lami­
nations are particularly effective for low-frequency applications. Lamination stacks are
rigid, having relatively high impact strength and can withstand higher temperatures than
other materials. They can also be used to support and protect the induction coil copper
while being electrically insulated from it. Another advantage of laminations is related to
their high relative magnetic permeability (in strong magnetic fields) and a saturation flux
density (1.4 to 1.9 T) higher than any other flux concentrator material. This means that
laminations are better able to retain their magnetic properties in the induction fields.
Without a concentrator, the magnetic flux would spread around the coil and link with
the electrically conductive surroundings (e.g., auxiliary equipment, metal supports, tools,
and fixtures). The flux concentrator forms a magnetic path to channel the coil’s main mag­
netic flux in a well-defined area outside the coil or near the coil assembly and facilitates
the concentration of magnetic flux lines in desired regions by providing a low-reluctance
path with minimum energy dissipation. This low-reluctance path reduces the stray flux
and tends to gather flux lines, thereby concentrating the EMF.
If required, Fluxmanager can allow modification of a magnetic field distribution and
minimization of flux leakage while effectively heating different-sized pipes (e.g., vari­
ous diameters/wall thicknesses). As an example, Figure 4.159 shows the ability to control
the coil’s EMF when heating 260-mm-diameter (left) and 170-mm-diameter (right) pipes.
Regardless of such large variations of coil-to-pipe air gap, the ability of a flux concentrator
to form a low reluctance magnetic path helps divert the magnetic field in a desirable way,
significantly improve the electromagnetic “coil-to-pipe” coupling, dramatically reduce the
flux leakage, and improve pipe heating when heating pipes of various diameters.
368 Handbook of Induction Heating

The use of line frequency (50/60 Hz) provides substantially more in-depth heating com­
pared to using medium frequencies, because it does not rely as much upon the thermal
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

conduction as heating using medium and high frequencies. This explains why the tem­
perature uniformity of line frequency induction heaters is affected to a much smaller
degree by a reduction of the heat time. When applying Fluxmanager Technology for heat­
ing smaller-sized pipes (less than 90 mm diameter) having thin wall thicknesses, elevated
frequencies of up to 150 Hz using specially designed inverters are more beneficial.
As an example, Figure 4.160 shows a specially designed 150-kW inverter that utilizes a stan­
dard 3 phase input (380, 480, or 575 V at 50/60 Hz) and provides a variable 60- to 150-Hz output.
Equipment includes infrared temperature control, pneumatic coil shuttle, data logging
capability for traceability and archival recipe management, and standard safety features.
Power supply dimensions (W × D × H): 915 mm (36 in.) × 1930 mm (76 in.) × 2030 mm
(80  in.). Standard coil assembly utilizing only six coil sizes provides quality heating of
pipes with outside diameters from less than 2.5 in. to more than 20 in. [404,405]:

• Coil “A”: 61–143 mm (2.375–5.625 in.)


• Coil “B”: 102–193 mm (4.000–7.625 in.)
• Coil “C”: 193–269 mm (7.625–10.625 in.)
• Coil “D”: 269–371 mm (10.625–14.625 in.)
• Coil “E”: 371–448 mm (14.625–17.625 in.)
• Coil “F”: 448–524 mm (17.625–20.625 in.)

Depending on pipe geometry, this patented system provides superior process flexibil­
ity. For heating larger pipes with heavier walls, lower frequencies are more appropriate.

FIGURE 4.160
Inductoheat’s 150-kW inverter provides a variable 60- to 150-Hz output. (Courtesy of Inductoheat Inc., an
Inductotherm Group company.)
Heat Treatment by Induction 369

However, higher frequencies (range, 120–150 Hz) would be more suitable for smaller-
diameter/thinner wall pipes. Thanks to this unique technology, a control board can gen­
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

erate a control signal to change output frequency during the heat cycle from 60 to 150 Hz
almost instantly. In reality, in order to change frequency, there will be a delay of a fraction
of a second, with 1 s being the maximum. During this time, heating power is turned OFF,
the control signal changes settings, and then power is turned back ON with a different
frequency. Practically speaking, a time delay of less than 1 s does not affect the process,
resulting in a short (approximately 1 s) soak time.
An increase in frequency of almost three times results in a corresponding reduction
in δ (as a square root of 3). If line frequency is used for heating smaller pipe diameters to
a tempering/stress-relieving temperature range, then there will be partial eddy current
cancellations occurring within the heated pipe. Besides that, an electromagnetic end effect
in this case produces heat deficit and the underheated end of the pipe. In contrast, the use
of 150 Hz drastically improves this situation, avoiding eddy current cancellation and com­
pensating for the heat deficit at a pipe end region.
In case of using a conventional multiturn solenoid coil, with a frequency of 120–​150 Hz,
there will still be measurable heat deficit at the pipe end region caused by the electromag­
netic end effect of the ferromagnetic body. However, the patented Fluxmanager Technology
significantly modifies an occurrence of the end effect thanks to the formation of a preferen­
tial magnetic path that allows compensating for the heat deficit at the pipe’s end.
Even when heating large pipes, it might still be beneficial in some cases to use a fre­
quency of 120–150 Hz at the very end (during the last few seconds) of the heat cycle to
better accommodate an electromagnetic end effect (if required) and raise pipe butt end
temperature. In this case, a frequency of 60–90 Hz is used for a majority of cycles, provid­
ing an in-depth heating effect of large thick-wall pipes resulting in improved O.D.-to-I.D.
heat uniformity during the majority of the heating cycle and, thus, producing better stress-
relieving quality compared to frequencies in the 1- to 10-kHz range.
Enhanced system flexibility is the greatest and most obvious advantage of this technol­
ogy. If required, a system can be programmed to use multiple frequencies within the 60- to
150-Hz range applying any sequence. Reduced process sensitivity and improved tempera­
ture uniformity compared to using 1-kHz frequencies and higher are other advantages of
this remarkable technology.
Figure 4.161 shows an example of FEA electromagnetic (a) and thermal (b) computer mod­
eling of a Fluxmanager system for induction heating and for stress relieving of a carbon

Carbon steel pipe

Variable pitch
Axis of symmetry coil

(a) (b)

FIGURE 4.161
Results of FEA electromagnetic (a) and thermal (b) computer modeling of a Fluxmanager system for stress
relieving/tempering of a carbon steel pipe [370 mm (14.57 in.); O.D., 16 mm (5/8 in.)] utilizing a frequency of
60 Hz. (From V. Rudnev, Flux Manager Technology, unpublished presentation for the Tenaris Group, 2013.)
370 Handbook of Induction Heating

steel pipe (370 mm [14.57 in.] O.D., 16 mm [5/8 in.] wall) using Fluxmanager Technology
and a frequency of 60 Hz. According to the customer specification, it required heating a
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

285-mm (11.2-in.) length of the pipe end area from room temperature to 605°C (1121°F).
Maximum heat time should not exceed 40 s per pipe end.
The induction coil comprises five groups of turns (8 + 4 + 4 + 4 + 12). Lamination shunts
placed outside the induction coil enhance heat efficiency and repeatability, dramatically reduce
external magnetic field, and focus heating in the area that is required to be heated. Expected
temperature uniformity immediately after 40 s of the heating cycle was ±16°C (±30°F).
The Fluxmanager inductor is easy to set up. Several design options are available, but all
of them are oriented on heating pipes and tubes of moderate and large size. Because of
space limitations, it is difficult to discuss all of the features of this patented technology.
Interested readers are welcome to contact Inductoheat Inc. (www.inducto​heat.com).

4.6.3.3 C-Core Inductors
Figure 4.162 shows several sketches of design options of one of the oldest inductor designs
for induction tempering. It is commonly called a C-core inductor. The induction heater
consists of a typical transformer-type design with one or several C-shaped magnetic cores.
Laminated low-carbon steel thin sheets are used for core fabrication. One or several multi­
turn coils are wound around the C-core to create the common magnetic flux. The intensity
of heating of the C-core inductors depends on the strength of that magnetic flux.
Because of known limitations, these types of heaters are not as popular and not as
widely used as the tempering inductors shown in Figures 2.13, 4.151, and 4.154. There are
still quite a few applications, including shrink fitting and tempering/stress relieving of
steel/cast iron parts where these types of inductors can be used quite effectively.
There are a number of approaches to applying this technique. The first deals with the
heating of solid workpieces. With this approach, a part is located within the open space of
the C-shaped core (e.g., see Figure 4.162a). Therefore, the heated workpiece becomes part
of the magnetic circuit. The major portion of the magnetic flux generated by the multiturn
coils (located below and above the workpiece) and diverted by the magnetic core will close
the magnetic loop through the magnetic workpiece. As the workpiece is ferromagnetic
and is not laminated, the induced eddy currents will heat it quite effectively as a result of
the Joule effect and hysteresis losses.
Among other factors, the effectiveness of C-core inductors depends on the gaps between
the core and workpiece. Smaller gaps result in higher efficiency. To ensure minimum gaps,
special clamping devices may be applied. Several C-shaped magnetic cores may also be
used to improve the heat uniformity.
The second approach to using C-core inductors is applicable to the hollow parts that are
placed around a certain portion of the core (see Figure 4.162b through d). In this case, the
heated workpiece represents the secondary winding of a transformer.
Both approaches apply low frequency (single phase or multiphase). The most commonly
used frequencies are in the range of 50 to 300 Hz. Higher frequencies lead to a reduction of
the electrical efficiency of the induction system owing to the higher losses in the laminated
C-core and also lead to localized heat surpluses in the heated workpiece.
Figure 4.163 shows a plot of the EMF of a C-core inductor for induction heating two
different workpieces: a solid carbon steel cylinder (Figure 4.163b through d) and a ring-
shaped workpiece (Figure 4.163e and f) using line frequency.
Figure 4.163b shows a plot of the imaginary field lines when a carbon steel cylinder is
below the Curie temperature. Figure 4.163c is a magnified version of the EMF distribution
Heat Treatment by Induction 371

“Lift-drop”
Induction action
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

coil Induction
Laminated core coil Laminated core

Ring

Heated
solid
part

Induction
Induction coil
coil Dropped-down core
(a) (b)

Induction “Open-close” Induction


“Open-close”
coil action coil
action

Hinged core Hinged core

Induction Ring
coil

Heated
ring
Laminated core Laminated core

(c) (d)

FIGURE 4.162
Sketches of design options of a C-core inductor (a) C-core design for heating of solid workpiece. (b) C-core
design for heating hollow parts applying lift-drop moveable I-section. (c) C-core design for heating hollow parts
applying a hinged core design. (d) Double C-core design (also referred to as E-core) for heating hollow parts
applying a hinged core design.

within the magnetic billet shown in Figure 4.163b. Figure 4.163d shows the EMF distribu­
tion when the workpiece is in a nonmagnetic state.
Because the temperature range of all tempering applications is below the Curie point, the
carbon steel/cast iron workpiece is always magnetic, and as shown in Figure 4.163b and c,
it acts as a continuation of the C-core. Actually, the C-shaped core becomes an O-shaped
core. It is much easier for magnetic flux to complete its loop through a magnetic body of the
heated workpiece than through the air. This is why the majority of the magnetic flux lines
pass through the workpiece. There is a negligible amount of leakage flux.
Pneumatic and hydraulic clamps can be used in order to ensure a minimum air gap between
the billet’s butt ends and the C-core faces which in some cases consist of thin thermal insulator.
372 Handbook of Induction Heating
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

(a) Sketch of C-core inductor (b) Below Curie point (c) Below Curie point (zoomed) (d) Above Curie point

800.0
725.0
652.0
578.0
504.0
430.0
356.0
282.0
208.0
134.0
50.0
Field lines °F
(e) Below Curie point (f) Above Curie point (g) Below Curie point

FIGURE 4.163
Plots of the EMF distribution of a C-core inductor for induction heating a solid carbon steel cylinder (b through
d) and a ring-shaped workpiece (e and f) using line frequency and thermal image of temperature distribution
of ring bevel gear (g).

Laminations of the C-core should be electromagnetically thin enough. For low-frequency


applications, individual lamination is typically 0.2 to 0.5 mm thick. Because laminations
are electromagnetically thin, there will be a current cancellation of eddy currents induced
in each lamination and, therefore, the Joule losses will be small enough and will not gener­
ate a large amount of heat (Section 4.7.3 provides more discussion on laminations as well
as other magnetic flux concentrators).
A large portion of the magnetic flux conducted by the laminations will be conducted by
the magnetic workpiece. However, the workpiece is electromagnetically thick and will not
experience current cancellation. Therefore, because of Joule and hysteresis losses, there
will be a significant amount of heat generated there.
Electromagnetically speaking, there is no difference between the conventional induc­
tion heating of billets using a solenoid coil or the C-core inductor shown in Figure 4.163a.
In both cases, it is possible to observe similar electromagnetic effects, including skin and
end effects (Figure 4.163b and c).
The use of line or low frequency results in larger eddy current penetration depth into the
heated material and, therefore, produces a smaller surface-to-core temperature difference
compared to using medium or high frequencies.
Note that, occasionally, there are some not-so-frequent attempts to heat billets to tem­
peratures that exceed the Curie point. This includes heating of relatively small billets from
room temperature to temperatures of warm working applications.
Figure 4.163d shows the EMF when heating a nonmagnetic solid cylinder in a C-core
inductor. As one can see, a nonmagnetic cylinder cannot conduct a magnetic field in the
same way as a ferromagnetic body. There are fewer magnetic lines crossing the part,
Heat Treatment by Induction 373

meaning a noticeable increase of magnetic flux leakage and a significant reduction in elec­
trical efficiency. The taller the nonmagnetic billet and the larger the air gap between the
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

billet’s butt ends and the C-core heating faces, the lower is the efficiency of the C-core
inductor.
As in the case of IH with a conventional solenoid coil, the electromagnetic end effect
when using a C-core inductor varies depending on whether the workpiece is in a magnetic
or nonmagnetic state.
A C-core inductor can be particularly effective for heating ferromagnetic ring-shaped
workpieces. Practically all the features discussed above with respect to the IH of magnetic/
nonmagnetic solid cylinders hold true when heating ring-shaped parts in a C-core inductor.
The distribution of the magnetic field when heating a ring-shaped workpiece in a mag­
netic state and in a nonmagnetic state is shown in Figure 4.163e and f, respectively. One can
conclude from a comparison of both figures that, when heating a ring-shaped workpiece,
the density of the eddy currents induced on the outside diameter is similar to the density
of eddy currents induced on the inside diameter. However, this phenomenon might be
misleading. People observing the heating process might assume that there is a uniform
current density distribution within the ring/tube wall and therefore temperature unifor­
mity through the wall has been achieved as well. As can be seen in Figure 4.163e, there
is still a skin effect within the heated wall thickness. The eddy current density decreases
from the ring surface toward the internal regions.
C-core tempering inductors have several beneficial features, including the following:

• The ability to use line and low frequencies leads to system simplicity and low-cost
power supplies.
• Radial and circumferential temperature gradients are typically quite small (par­
ticularly when heating a ring-shaped workpiece), producing fairly uniform tem­
peratures even when heating workpieces with geometrical irregularities (e.g., ring
gears, Figure 4.163g). In some applications of heating parts with sizable diameters
instead of C-core, using an E-core inductor or double or triple E-core inductors
might be beneficial. The number of poles might be further increased, improving
heat uniformity. As an example, Figure 4.164 shows an induction line frequency
system utilizing an E-core (left) and multiple cores (right) that can be used for
tempering/stress relieving, as well as paint curing, shrink fitting, disassembling,
and other applications (also see Section 5.4).

(a) (b)

FIGURE 4.164
Induction line frequency system utilizing an E-core (a) and multiple cores (b). (Courtesy of Inductoheat Inc., an
Inductotherm Group company.)
374 Handbook of Induction Heating

• The heated part does not need to be rotated to obtain circumferential heat
uniformity.
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

• The process provides the required temperature uniformity for medium- and
large-sized (200 mm O.D. and even larger) ring/tube-shaped workpieces, includ­
ing ring gears and bearings. Height is typically not a problem. There are cases of
successfully heated tubular workpieces that are as long as 1 m.
• Design simplicity.
• High electrical efficiency (typically 85% to 92%) and inductor power factor (COS
ϕ 0.75 to 0.85).

However, the wide utilization of C-core inductors is limited by several shortcomings:

• Complications in handling certain types of parts for in-line processing.


• Ineffective for heating asymmetrical complex-shaped workpieces as well as solid
nonmagnetic materials.
• Relatively low production rate.
• The presence of significantly sized air gaps breaking the magnetic path can result
in reduced coil efficiency and worsened heat uniformity as well as local overheat­
ing of lamination faces.
• Equipment might be quite noisy (85–90 db) at higher power levels.
• Does not always suit well for tempering selective areas and for certain geometries.
• For solid parts with different sizes/shapes, it is necessary to have adjustable lami­
nated ends (inserts).
• The use of medium and high frequencies may produce localized hot spots.
• If an inverter is not used, there could be challenges associated with power factor
correction and phase balancing (when a single-phase inductor is used).
• Parts can be magnetized and the value of retained magnetism could exceed the
permissible maximum level requiring post-degaussing. In some cases, power con­
trol of C-core inductors may incorporate an option of degaussing, but it compli­
cates the system.

4.6.4 Final Remarks
There is a group of applications where induction tempering has been successfully
applied in conjunction with self-tempering, combining the benefits of both processes.
For example, a combination of self-tempering and multipulse induction tempering
is successfully used in nonrotational crankshaft hardening (SHarP-C Technology)
[22,351,352]. In this case, the journals of a crankshaft are stationary heat treated. For
most automotive crankshafts, it takes approximately 3 to 4 s to austenitize a journal
surface layer for hardening using frequencies in the range of 10 to 30 kHz (depend­
ing on the specifics of the automotive crankshaft and the required case depth). After
completion of austenitization, quenching is applied for only 4 to 5 s, followed by 3
to 5 s of the first soaking that accomplishes the first stage of self-tempering. Then,
low-power induction tempering is applied for approximately 3 to 5 s, followed by the
second soaking and the second induction tempering. The process may be repeated
to achieve desirable tempering conditions, providing a multipulse tempering effect
Heat Treatment by Induction 375

combined with self-tempering and allowing optimization of the tempered structure.


SHarP-C Technology is reviewed in Section 4.9.4.2.
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

The decision on whether to temper the workpiece and what tempering method to be
used (self-tempering, oven tempering, or induction tempering, or combinations of these)
should be carefully weighed. Studies show that sometimes, regardless of having a similar
hardness pattern, certain properties of tempered parts subjected to higher-temperature/
shorter-time induction tempering may be different from those of parts subjected to lower-
temperature/longer-time oven/furnace tempering.
For example, in a comparison of the properties of carbon steel parts processed with
both types of tempering, it has been shown that regardless of identical hardness (48 HRC),
the tempered carbon steel part (0.8% C) subjected to the higher-temperature/shorter-time
tempering (30 s at 465°C) showed three times higher brittle strength compared to a part
tempered for 3000 s at 350°C. The improvement in brittle strength considerably extends the
impact toughness. Similar results have been obtained for some alloyed steels, including
chromium steels.
One hurdle to the acceptance of induction tempering is that some metallurgists are
not comfortable with shorter-time/higher-temperature induction tempering. They feel
that an oven that heats the entire part and holds it at a temperature for hours is more
suitable.
There are numerous studies devoted to comparison of various workpieces tempered
under different conditions: as-quenched versus oven tempered versus induction tem­
pered, including materials published in Refs. [31,408].
For example, K. Madler and J. Grosch conducted a comprehensive study using speci­
mens made of three plain carbon steels and three low-alloy steels having various carbon
contents (from 0.4 to 0.97% C) and induction-hardened case depths [408]. Some specimens
were examined under the as-quenched (untempered) condition, but others were oven tem­
pered (180°C for 2 h) or induction tempered (peak temperature was 250°C with a heating
rate of 46°C/s). Steels with a higher carbon content were selected to evaluate the impact of
high-carbon martensite and RA.
Tests were performed to determine the bending strength of the various materials. The
study revealed that the bending strength for the tempered samples versus the untempered
samples was substantially increased for medium-carbon steels regardless of the temper­
ing method. In contrast, steels with a higher carbon content show a lower crack resis­
tance under the as-quenched condition with less pronounced benefits of either type of
tempering.
The toughness of medium-carbon steels was increased appreciably by both temper­
ing methods. The steels with higher carbon contents, particularly 100Cr6 steel, showed a
considerably lower increase in toughness by tempering, but generally, conventional and
induction tempering have a very similar effect [408].
Other materials related to the QT properties of steels are provided in various publica­
tions, including Refs. [2,4,15,16,28–31,33,149–158,392–394,408–416].
Although both methods are proven technologies, they can provide the tempered parts
with different properties. However, it does not mean that one tempering method is bet­
ter than another. The proof of any production process is how well the part performs in
service. The higher-temperature/shorter-time induction tempered part, like any other
machine component, should be thoroughly tested and evaluated.
Fatigue and failure test data for induction- and furnace-tempered parts should be com­
pared. It is important to remember that the surface tempering temperature alone is not a
valid indication of proper tempering.
376 Handbook of Induction Heating
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

FIGURE 4.165
Two examples of compact in-line induction tempering machines. (Courtesy of Inductoheat Inc., an Inductotherm
Group company.)

If tempering has been done correctly, there will only be an allowable reduction in hard­
ness, which will be more than offset by the benefits obtained: internal stress relief, improved
ductility or toughness, shifting of the dangerous maximum of tensile stresses farther away
from the applied stresses, and improved machinability, just to name a few benefits. Other
benefits of in-line induction tempering may include part-to-part processing, minimization
of the probability of delayed cracking, savings in investment cost, environmental friendli­
ness, low floor space requirements, maintenance and labor savings, high energy efficiency,
and precise control and monitoring.
Figure 4.165 shows two examples of compact in-line induction tempering machines that
occupy a fraction of the floor space of tempering furnaces assuming a similar production
rate. The sophisticated design allows increased reliability, simplified operating conditions,
and decreased maintenance requirements. Alternative temper inductor designs can also
be applied with these systems. A variety of parts can be processed on these machines hav­
ing multiple hardened areas with both inside and outside case-hardened surfaces.

4.7 Magnetic Flux Control Techniques: Concentrators,


Intensifiers, Shunts, and Shields
Before embarking on a discussion of controlling magnetic flux, we need to review the defi­
nition of magnetic flux and establish how we observe the result of flux control techniques.
Flux is a term used to describe the rate of transfer of fluid, particles, or energy across a
given surface interface. Magnetic flux is a measure of the total number of imaginary mag­
netic field lines (Figure 3.23a) crossing a chosen surface.
Although magnetic flux lines cannot be seen, they can be represented in mathematical
calculations and graphs. Because an isolated magnetic source or monopole has never been
observed to exist in nature, the net flux or divergence of the magnetic flux density vector
(B) is always equal to zero (Equation 3.40). This simply means that the imaginary lines of a
magnetic field represent closed loops, and for a given volume, the number of lines entering
the volume will equal the number of lines leaving.
Analyzing the distribution of the imaginary field lines, we can move on to the question of
interest, which is “Can the position of these lines be altered in any way?” The answer to this
Heat Treatment by Induction 377

question is yes. As has been illustrated above utilizing various case studies, the position of
the field lines and the strength of the field in a given area can be modified in several ways. The
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

density of the lines can be modified by changing the cross-sectional area through which the
flux is flowing. The location of the lines can be changed by the use of special materials that
provide a preferable path or restrict the path of the magnetic flux. A third way that the flux
lines can be changed is by the presence of other current-carrying conductors in the vicin­
ity of the existing field according to several electromagnetic phenomena including but not
limited to the proximity effect, as well as faraday ring phenomenon.

4.7.1 Electromagnetic Shields
In the area of IH, magnetic fields are generated for the purpose of heating electrically
conductive materials. Often, the AC field generates the heat not only within the desired
workpiece but also in electrically conductive areas adjacent to the coil. Further concerns
can arise as the EMF couples into an enclosure, tooling, or fixtures producing undesirable
heating and potentially causing disturbances in control, metering, power circuit opera­
tion, or power wiring. For these reasons, it is often necessary to use magnetic shields to
reduce the effect of the EMF and to protect other structures, components, controls, equip­
ment or people in the immediate area.
The effectiveness of a magnetic shield is determined by comparing the field strength of
the incident wave and the reflected or attenuated wave. The common formula used is

SE(db) = 20 × log(H1/H 2)

where H1 and H2 are the field strengths of the incident and attenuated waves, respectively
[407]. Higher numbers represent better shielding effect. A 40-db attenuation would repre­
sent a field strength of the reflected or attenuated wave (H2) of 100 times less than the field
strength of the incident wave (H1). A specific area may be shielded from an EMF by either
reflecting the wave or absorbing the energy. Common materials used for the purpose of
reflecting the wave would be copper or aluminum. The main materials used for absorp­
tion would be carbon steel or high-permeability materials.
At higher frequencies, shielding by reflection is very effective. Absorption may be used
but is most often unnecessary. When reflection is the primary mechanism, a relatively thin
shield can produce quite good results. Often, a thin copper or aluminum foil or thin sheet
may be used above 100 kHz for a 60-db attenuation and one reference depth for frequen­
cies as low as 10 kHz for a 40-db attenuation. At frequencies below 10 kHz, absorption is
the most effective choice. Materials used for absorption should be highly permeable and
the minimum thickness should be approximately four times the reference depth of the
frequency for a 35-db attenuation.

Absorption (db) = 8.68* S.D.,


where S.D. is shield thickness/reference depth.


The effectiveness of a given shield is evaluated by the equation

SE (db) = R (db) + A (db) + B (db).


This equation includes the effects of reflection and absorption as well as a factor that is
used for reflection on very thin shields at high frequencies (Figure 4.166).
378 Handbook of Induction Heating
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

FIGURE 4.166
Application of thin shields at high frequencies.

For shields operating by the principle of reflection, the location of the shield may have
a large impact on its effectiveness. For shields operating by the absorption principle, the
losses are independent of the distance between the shield and the source. Shields utilized
in the presence of very strong EMFs can be a major source of power dissipation in the sys­
tem and may require water cooling. Shields typically reduce the inductance of the circuit.
This may affect load-matching characteristics.
Figure 4.167 shows the typical attenuation versus shield thickness for a copper shield
and illustrates the rule of thumb of using a shield that is equal in thickness to four times
the material δ at the given frequency. Figure 4.168 gives an indication of how the shielding

100

80
Decibels

60

40
10 kHz
20

0
0 1 2 3 4 5 6
Skin depth of shield

FIGURE 4.167
Typical attenuation versus shield thickness for a copper shield.

40
Attenuation, dB

20

0
0 2 4 6 8 10
Frequency, kHz

FIGURE 4.168
Illustration of how the shielding effectiveness falls off with a given thickness of shield as the frequency is
lowered.
Heat Treatment by Induction 379

TABLE 4.25
Shield Effectiveness at 10 kHz Frequency
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

Case Description (in.) Flux Density (G)


1 No shielding 40.0
2 0.125 Copper shield 1.16
3 0.125 Aluminum shield 1.18
4 0.125 Stainless shield 4.0
5 0.125 Mild steel shield 4.7
6 0.125 Mag. Mat’l. (perm. = 60) 20.0
7 0.66 Stainless shield 0.89
8 0.4 Concentrator @ coil 27.0
9 0.4 Concentrator and Cu shield 5.5
Note: Calculations at 3 in. from coil surface using relatively short shield.

effectiveness falls off with a given thickness of shield as the frequency is lowered. The
effectiveness of the shield illustrated in Figure 4.168 could be doubled simply by increas­
ing the length to include the full area of coverage of the EMF. In an attempt to shield an
inductor, this may require a shield that extends at least one and one-half times the inductor
outside diameter beyond the center line of the inductor.
Table 4.25 shows a brief comparison for shielding the effectiveness of different materi­
als at a 10-kHz frequency, assuming that the two conductors in the examples above form
an induction coil to be shielded and the shield represents a closed-loop system (electri­
cally speaking).

4.7.2 Magnetic Shunts
A magnetic shunt generally consists of a large stack of thin steel laminations placed along
the axis of an inductor and parallel to it. The shunt provides a low-reluctance path for
the magnetic flux. Shunts can considerably reduce the external magnetic field and pre­
vent heating of surrounding metallic structures. Shunts can also be a source of significant
power dissipation in the system and, depending on application, might or might not need
to be cooled.

4.7.3 Magnetic Flux Concentrators


Magnetic flux concentrators (also called flux intensifiers, flux controllers, diverters,
shunts, or magnetic cores) have become an acknowledged standard in induction design
[2,18,20,44,51,322–324,417–420].
High-permeability soft-magnetic materials having low electrical conductivity are rou­
tinely used in a manner similar to that of magnetic flux cores in power transformers or
motors. The soft-magnetic nature of flux concentrators means that they are magnetic only
when an external magnetic field is applied. Upon exposure to an AC magnetic field, these
materials can change their magnetization rapidly without much friction. Narrow mag­
netic hysteresis loops of small area are typical for these materials.
In the absence of magnetic field, magnetic domains are randomly arranged. Such
arrangement corresponds to a minimum energy configuration when the magnetic effects
of the domains cancel each other, resulting in a negligible overall magnetization. The
magnetic domains can be easily rearranged by applying an external magnetic field. The
380 Handbook of Induction Heating

direction of domain rearrangement corresponds to the direction of the applied EMF, mak­
ing them behave as temporary magnets.
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

Different applications may call for various materials used as magnetic flux concentrators
including stacks of silicon-steel laminations, pure ferrites, and various proprietary multi­
phase composites. Successful development of powdered metal composites based on Fe, Ni,
Co, and other elements have dramatically increased the effectiveness and popularity of
flux concentrators in the last few decades.
New flux concentrator materials and different grades appear quite regularly. For exam­
ple, at one time, it was considered that laminations could only be applied to frequencies
below 10 kHz or so. Recently developed nanotechnology has led to the development of
novel materials that substantially expand this range. Laminations made of the nano­
crystalline alloy NANO comprise a proprietary composition of 82% Fe, with a balance
of silicon, nickel, boron, copper, carbon, and molybdenum, and have a frequency range
exceeding 50 kHz [420].
The selection of a particular material for fabricating a flux concentrator depends on a
number of factors, including the following:

• Applied frequency, power density, and duty cycle


• Operating temperature and ability to be cooled
• Geometries of workpiece and inductor
• Machinability, formability, structural homogeneity, and integrity
• An ability to withstand an aggressive working environment resisting chemical
attack by quenchants and corrosion
• Brittleness, density, and ability to withstand impact force
• Ease of installation and removal, available space for installation, and so on

4.7.3.1 Physics of the Magnetic Flux Concentration


The number of imaginary magnetic flux lines that enter the workpiece determines the
degree of electromagnetic coupling between the workpiece and the magnetic field of the
inductor. This magnetic flux density is proportional to the coil current. Flux lines tend to
stay closer to the coil conductors as frequency is increased. Flux leakage refers to the mag­
netic flux that does not thread through the heated material.
In addition to some other factors, the degree of flux leakage depends on the coupling
gap between the current-carrying surface of the inductor and the workpiece. Flux leakage
can be reduced by tightening the workpiece-to-inductor gap or by the use of flux concen­
trators, which are purposely placed on or near the coil assembly to provide a path of low
reluctance with minimum energy dissipation to facilitate the concentration of flux lines
in the desired region and thus decrease the stray flux. Flux concentrators can play a very
important role in enhancing the electrical efficiency, improving the coil power factor and
may also noticeably reduce inductor current requirements.
Let us examine what happens when a magnetic flux concentrator is applied. Without
a concentrator, the magnetic flux would spread around the coil or current-carrying con­
ductor and link with the electrically conductive surroundings (e.g., auxiliary equipment,
metallic fixtures, support, tooling, etc.).
To illustrate the effect of a flux concentrator, it is beneficial to review a case study. Figure
3.16 shows the magnetic field (a) and current distribution (b) in an isolated rectangular
Heat Treatment by Induction 381

conductor with clearly visible skin effect. The current redistribution within this con­
ductor after locating an electrically conductive body (i.e., workpiece) in its proximity
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

can be observed in Figure 3.22 (right). Because of the electromagnetic proximity effect,
a significant part of the conductor’s current will flow near the surface of the conductor
that faces the workpiece. The remainder of the current will be concentrated in the sides
of the conductor.
When an external magnetic flux concentrator is placed around the conductor, practically
all of the conductor’s current will be concentrated on the surface facing the workpiece
(Figure 3.23, right). The magnetic concentrator squeezes the current to the “open surface”
of the conductor according to the electromagnetic slot effect. Current concentration at the
inductor surface facing the workpiece results in improved electromagnetic coupling and
therefore enhances the electrical efficiency. The current density distribution in the con­
ductor depends on the frequency, geometry, and material properties of the conductor, the
workpiece, and the concentrator.

4.7.3.2 Design and Application Features


When a concentrator is used, an excessive localized heat generation in the coil copper may
take place. Note that Figure 4.83a exhibits severe copper overheating in the region where
a magnetic flux concentrator is positioned. Therefore, consideration must be given to the
coil current density, wall thickness, copper selection, water cooling, and other subtleties of
the inductor design that were discussed in Sections 4.2.2 through 4.2.4. Original inductor
design should be evaluated if stress cracking attributed to the copper work hardening has
occurred over time.
It is also important to be aware that impedance of the bare coil can be much differ­
ent compared to an identical inductor with a flux concentrator. Therefore, after the flux
concentrator has been installed, it is necessary to check that the inductor properly load
matches the power supply.
Utilization of a magnetic flux concentrator may lead to an increase in power density
not only in the induction coil but also in certain workpiece areas. This is the reason why
the electromagnetic slot effect plays a particularly important role in the proper design
of inductors for selective induction hardening, including channel, hairpin, and pancake
inductors, just to name a few.
Special care should be taken when applying flux concentrators to a multiturn coil,
because the voltage across the turns can be significant, and a short current path and arcing
can develop through the concentrator. Thus, the appropriate electrical insulation should
be applied.
One of the major concerns when heating a complex geometry workpieces to the austen­
ite phase temperatures is that of undesirable heating of adjacent areas of the workpiece
that have previously been hardened This is often referred to as temper back or the anneal­
ing effect of adjacent regions. This is a concern in the induction hardening of camshafts
and other critical components. The complexity of this problem arises from the fact that,
because of EMF propagation, the eddy currents are induced not only in the workpiece that
is located under the inductor but in adjacent areas as well.
Figure 4.169a shows a sketch of an induction camshaft hardening and the EMF field dis­
tribution around a bare single-turn coil. Without a concentrator, the magnetic flux spreads
around the inductor and links with electrically conductive surroundings, which include
neighboring areas of the part (e.g., adjacent journals and cam lobes) as well as electrically
conductive components of the machine, work hold tooling, and fixture. As a result, the
382 Handbook of Induction Heating
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

(a) (b) (c)

FIGURE 4.169
Sketch of an induction camshaft hardening system and the EMF field distribution around it using a bare single-
turn coil (a) versus a coil with a U-shaped flux concentrator (b) versus a combination of U-shaped flux concen­
trator and Faraday rings (c).

heat will be generated. This heat can cause undesirable metallurgical changes. For exam­
ple, edge areas of camshaft lobes that were induction hardened in previous process stages
can be softened because of the “bleeding-over” of the heat during the subsequent heating
stage when the next lobe is induction hardened.
At different stages of the process cycle, the heating intensity of the adjacent areas can
vary. At the initial stage of the heating cycle, the entire workpiece is magnetic, resulting
in high electrical efficiency and intense heating of any surface areas of the workpiece sur­
rounded by the coil compared to any other regions.
After a short time, the surface of the heated lobe reaches the Curie temperature, the μr
drops to 1, the surface layer becomes nonmagnetic, and its heating intensity drastically
decreases. Although the surface of the heated lobe has lost its magnetic properties, the
adjacent areas retain theirs. Consequently, during this heating stage, a greater portion of
the EMF will link with the adjacent areas, including the corner of the adjacent lobe, poten­
tially resulting in its undesirable temper back or even re-austenization.
After placing a U-shaped magnetic flux concentrator around the single-turn coil
(Figure 4.169b), a much smaller portion of the EMF of the inductor will link with adjacent
lobes, reducing their heating. The external field will be dramatically reduced while the
internal field has increased.
Depending on a camshaft’s geometry and coil design specifics, it is reasonable to expect
a 4- to 12-fold reduction in the power density induced in adjacent lobes compared to using
a bare coil. This substantially reduces the probability of undesirable softening. Therefore,
flux concentrators allow decoupling of the induction coil to the adjacent electrically con­
ductive regions while noticeably improving effectiveness of heating areas being enclosed
by an inductor.
Further reduction of the flux leakage can be achieved by using a combina­
tion of  U-shaped magnetic flux concentrators and Faraday rings (“robber” rings)
(Figure 4.169c).
In some applications, several induction coils are involved. Because of the relatively
small distances between coils, strong magnetic field interactions can occur, potentially
adversely affecting operation where neighboring coils are powered by different power
supplies. Besides that, the generation of a sufficiently intense stray field can cause mal­
functions of robots, grippers, sensors, pyrometers, and other electronic devices. Thus,
the use of flux concentrators as electromagnetic shields can help prevent those undesir­
able phenomena.
Heat Treatment by Induction 383
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

(a) (b) (c)

FIGURE 4.170
Side-by-side comparison of power density distribution when using a magnetic flux concentrator with a profiled
single-turn inductor in selective heat-treating of fasteners. (a) Bare coil vs. coil with U-shaped flux concentrator;
(b) both sides have U-shaped magnetic flux concentrators; and (c) coil with L-shaped concentrator vs. coil with
U-shaped flux concentrator. (From V. Rudnev, Workshop on induction heating and heat treating technologies
for Alcoa’s Fasteners Group, Tucson, AZ, Sept. 19–20, 2012.)

The necessity to comply with international standards and regulations related to control­
ling external exposure to EMF in the workplace (including IEEE, OSHA, WHO, IRPA, and
others) can also be assisted by using flux concentrators acting as shields.
Properly applied flux concentrators can play an important role in enhancing induction
heating. Figure 4.170 shows a side-by-side comparison of power density distribution when
using a magnetic flux concentrator with a profiled single-turn inductor in selective heat
treating of fasteners [419]. The current-carrying face of the coil copper was fabricated to be
noticeably narrower. This helps increase the density of coil current and the corresponding
intensity of the generated heat. Figure 4.170a shows a comparison of power density distri­
bution generated by a bare coil (left) compared to a coil with a U-shaped flux concentrator
(right). Figure 4.170b shows the difference in power density distribution and its magnitude
when only an L-shaped concentrator is applied (left) versus using a U-shaped encircling
concentrator (right). The application of the U-shaped concentrator (right) provides a sig­
nificant benefit compared to the L-shaped concentrator by increasing the heat intensity in
the required region, reducing the axial HAZ, and decreasing undesirable heating of the
neighboring areas (e.g., reducing heating of the head of the fastener). Figure 4.170c shows
the U-shaped concentrator on both sides.
This computer modeling case study emphasizes the importance of having an uninter­
rupted magnetic path when applying flux concentrators, its effect on the generation of heat
sources, and shielding effect. Care should be taken at the corners of flux concentrators
because of their tendency to overheat because of increased flux density.

4.7.3.3 Selection of the Flux Concentrator Materials


Some factors affecting the choice of concentrator material have been outlined above.
Usually, the higher the value of μr, ρ, k, Curie point, and saturation flux density, the better
the situation. Other important factors rely on lower values for better performance, includ­
ing magnetic hysteresis heat, and eddy current losses, and low anisotropy.
Magnetic materials with high ρ have reduced eddy current losses, thereby decreasing
the tendency to overheat. High thermal conductivity flux concentrators usually have a
longer life because of reduced probability of local overheating. Local overheating can be
caused by several factors including thermal radiation from the heated workpiece, exces­
sive localized flux density, or intense heat flow due to thermal conduction from higher-
temperature regions of the inductor.
384 Handbook of Induction Heating

One of the most important magnetic properties of concentrator materials is hysteresis


loss. This quality is derived from the magnetization curve [48]. A typical magnetization
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

curve, representing the magnetization process, includes the following:

1. A cycle of magnetization in one direction


2. A reversal of the applied magnetic field, which results in demagnetization of pre­
viously magnetized material and its magnetization in the opposite direction
3. Another reversal process resulting in magnetization in the original direction

Hysteresis loss is characterized by the conversion of electromagnetic energy into ther­


mal energy owing to rearrangement of the magnetic domains during the magnetic hyster­
esis cycle. Magnetic hysteresis loss should be as small as possible because it contributes to
a temperature rise of the flux concentrator, which can cause a loss of its magnetism, struc­
tural degradation, and a deviation of the heat intensity and pattern. Magnetic hysteresis
loss is proportional to the area of the hysteresis loop and the frequency.
Materials used for flux concentrators should have a coercive force as small as possible.
A perfect (idealistically speaking) flux concentrator providing maximum efficiency would
have no magnetization remaining after the external magnetic field has decreased to zero
and would have infinite values of μr, ρ, and k. A wide opening in the magnetization curve
in combination with a high frequency corresponds to a high value of hysteresis loss. The
flux concentrator properties can be determined from the manufacturer’s data sheet or
measured with appropriate test equipment.
The types of materials most commonly used in induction heat treatment for flux concen­
trators are as follows [2,18,20,44,51,322–324,417–420]:

1. Laminations
2. Pure ferrites and ferrite-based and iron-based composites utilizing compressed
magnetic powder particles
3. Complex multiphase magnetic composites
4. Soft formable materials

Grain-oriented and nonoriented magnetic alloys used in laminations are nickel–iron


alloys and cold- and hot-rolled silicon–iron alloys. Packets of conventional laminated steel
stampings are effectively used from line frequency to 30 kHz. However, there are cases
where novel laminations applying nanotechnology have been successfully used at higher
frequencies (i.e., 50 kHz and higher).
Laminates are insulated with mineral and organic coatings and must be electrically iso­
lated from each other and used at the proper frequency. The thickness of the individual
laminates should be held to a minimum (commonly being 0.06 to 0.8 mm thick) to keep
low kilowatt losses. Thin laminations are used for higher frequencies. Laminates with a
thickness greater than 0.3 mm are typically used for frequencies below 500 Hz. Compared
to the majority of alternative flux concentrator materials, laminations have the highest μr
and saturation flux density in typical IH applications, which is considered an important
advantage.
When laminations are applied, some challenges may occur. Laminations may be labor
intensive to install. They are heterogeneous in nature (thermally and electromagneti­
cally) and particularly sensitive to aggressive environments such as certain quenchants,
which lead to rust and degradation problems (Figure 4.171, top right). Degradation of the
Heat Treatment by Induction 385
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

FIGURE 4.171
Typical failures of different magnetic flux concentrators.

magnetic properties of laminations is caused by an increase of coercive force and corro­


sion. If the laminations are not firmly clamped, they could be bent, and start vibrating,
resulting in mechanical damage, noise, and subsequent coil failure (Figure 4.171, top left).
Care should be taken with the corners of laminations and their ends, because of their ten­
dency to overheat as a result of electromagnetic end effects.
It may be beneficial to use soft, nonelectrically conductive spacers between lamination
keepers and lamination packs to allow expansion to some degree during the heat cycle.
One of the main advantages of using laminations is that they are relatively inexpensive
and can withstand high temperatures better than alternative materials. Lamination pack­
ets can also be used to support the inductor while remaining insulated from it. Another
advantage of laminations deals with the fact that laminations have not only a high μr but
also an exceptionally high saturation flux density (1.4–1.9 T). No other types of commonly
used flux concentrators, including magneto-dielectric materials and ferrites, have such
a high saturation flux density. This means that laminations better retain their magnetic
properties in the strong magnetic fields.
Laminations made of the nanocrystalline alloy NANO comprise a proprietary composi­
tion of 82% Fe, with a balance of silicon, nickel, boron, copper, carbon, and molybdenum,
and have the following impressive properties [420]:

• Curie temperature is 570°C (1060°F).


• Density is 7300 kg/m3 (230 lb/ft3).
• Thickness of individual laminations is 0.02 mm.
• Stacking factor is 0.75.
• Magnetic flux density 13 kilogauss (kG).
• Coercive force is 0.04 Oersted.
• Namlite coating is applied to NANO material to reduce eddy current losses.
• Frequency range exceeds 50 kHz.

Iron-based composites have been widely applied since the 1980s. They can be machined
by conventional methods, come in different sizes, and are available in μr = 15–40 range.
386 Handbook of Induction Heating

Some materials are rated for higher frequencies and others are rated for lower frequen­
cies. Typically, they do not significantly degrade over time or rust and can be easily
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

removed and replaced for coil repairs. There are a number of these materials on the
market today (2016).
Other materials are pure ferrites or ferrite-based composites. Ferrites are dense ceramic
structures made by mixing iron oxide (FeO) with oxides or carbonates of one or more
metals such as nickel, zinc, or magnesium [336]. They are pressed and then fired in a
kiln at high temperature and machined to suit the coil geometry. In relatively weak EMF,
ferrites have very high magnetic permeability (μr = 2000+). Ferrites are quite brittle, and
this is one of their main drawbacks. Other disadvantages of ferrites are low saturation
flux density (typically 3000 to 4000 G), low Curie point (approximately 220°C or 450°F),
poor machinability, and inability to withstand thermal shocks or impact force. Because of
their high ρ, ferrite-based concentrators are particularly attractive with high frequencies
(50 kHz and higher).
Ferrites and composites that utilize compressed magnetic particles have noticeably more
homogeneous physical properties than laminations, though some degree of heterogeneity
is still present in concentrators made of composite materials depending on the direction of
applied pressure during their fabrication.
Some concentrator materials are provided in a soft formable state that can be molded
and baked in place (thanks to a thermally sensitive catalyst) to a desired shape for develop­
ment purposes and later machined.
In some applications, magnetic flux concentrators can be made from a single material.
Others may be constructed of several materials. For example, in a split-return inductor,
laminations can be located in the middle area of the coil and iron- or ferrite-based compos­
ites can be placed at the coil ends. Such designs are cost-effective and electrically efficient
because they take into account the 3-D field distortion owing to the electromagnetic end
effect that would result in additional losses within laminations positioned at the inductor
ends (see Section 4.2.3.1).
Table 4.26 compares the basic properties and characteristics of laminations and iron- or
ferrite-based powder composites [323]. The table can be used as a general guide for deter­
mining the application range of use of concentrators.

4.7.3.4 Advantages and Drawbacks of Using Magnetic Flux Concentrators


As stated earlier, the use of magnetic flux concentrators may be associated with substantial
benefits:

• Energy savings attributed to a reduction of the required power and coil current
as well as increasing the electrical efficiency and power factor of the inductor. For
example, the performance of coils used to heat the inside or bore of a hollow work­
piece can be dramatically improved by placing the flux concentrator inside the coil
(Figures 3.27 and 4.96). The ability to apply certain inductor styles may solely rely
on the use of flux concentrators (Figures 4.93 and 4.94).
• The heat cycle time can be shortened measurably.
• Making it possible to localize the heat generation, minimizing energy consump­
tion and distortion, and improving hardness patterns. For example, flux concen­
trators applied at coil ends increase heat intensity there and make the hardness
pattern more uniform, transforming it from less desirable thumbnail patterns to
straighter ones.
Heat Treatment by Induction 387

TABLE 4.26
Comparison of Selected Characteristics of Laminations versus Iron- or Ferrite-Based Powder
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

Composites
Iron-Based or Ferrite-Based
Property or Characteristic Laminations Powder Materials
Homogeneous properties Fair (pronounced heterogeneous High (still heterogeneous)
properties)
Magnetic properties (magnetic High Fair (particularly in strong fields)
permeability, magnetic
saturation, etc.)
Curie temperature High Fair
Electrical resistivity High (if installed properly in High
respect to the field orientation
and frequency)
Thermal conductivity Good to fair (depending on Fair
and an ability to be cooled orientation of stack to magnetic
filed)
Maximum long-term use Good (depends on thermal insulation Fair (less than 250°C [482°F] because
temperature between sheets of laminations, of low working temperature
typically above 500°C (932°F) of bonders) [21]
Machinability Fair Good
Resistance to corrosion Fair Good
Typical working frequency range 50 Hz to 50 kHz 50 Hz to 800 kHz+
Localized overheating Fair Fair
Ability to withstand high- Good Fair
temperature radiation and
heat convection
Impact strength Good to fair (depending on Fair to poor
orientation of stack to impact)
Ability to be applied to Fair Good
complex-shaped coils
Source: V. Rudnev, Systematic analysis of induction coil failures, Part 5: Effect of flux concentrators on coil life,
Professor Induction Series, Heat Treating Progress, ASM Int’l, Materials Park, OH, March/April, 2006,
pp. 21–26.

• Acting as a shield in helping to minimize the probability of the electronic devices


working in proximity to induction coils malfunctioning (e.g., robots, gentries,
work-hold tooling or transporting mechanisms, etc.). Assisting to avoid/mini­
mize an undesirable heating of adjacent regions (shielding effect, Figure 4.169).
Eliminating or dramatically reducing exposure of personnel to external EMFs.

Addition of a flux concentrator to an existing inductor involves an additional expense.


A common saying among induction heating professionals is that if a good product can
be produced without a flux concentrator, there is no reason to add to the inductor cost;
besides, it might simply be an additional part that can go wrong. All concentrators degrade
in service. As soon as they are installed, their ability to concentrate magnetic fields begins
to slowly decline as a result of degradation of binders used to hold magnetic powder par­
ticles together and rusting, just to name a few factors.
Another major concern is the reliability of the installation. Flux concentrators fabri­
cated from iron- and ferrite-based powder composites are usually soldered, screwed, or
388 Handbook of Induction Heating

sometimes even simply glued to the inductor copper using so-called thermally conductive
adhesives. However, that label could be misleading. These adhesives actually have much
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

poorer thermal conductivities than copper or even steel. Users should not automatically
assume that the adhesive thermal conductivity is sufficiently high to provide the required
cooling of a flux concentrator. If used, the thickness of the adhesive layer should be mini­
mized to make the best of its capability to provide cooling from the water-cooled copper
turn toward the surface of the flux concentrator [336].
There is another mispostulation that can lead to cooling-related failures. Some users
mistakenly think that if the surface of a powder-type flux concentrator has good con­
tact with the water-cooled inductor, then the entire flux concentrator will be suffi­
ciently cooled. Iron- and ferrite-based powder flux concentrators are fabricated using
special binders. Their thermal conductivities are typically even lower than those of
so-called thermally conductive adhesives, which may result in localized overheating
almost regardless of how well the concentrator is in contact with the water-cooled
copper.
Flux concentrators are typically positioned in areas of high magnetic flux density, where
electromagnetic forces can be substantial [323]. Those forces can reach a considerable mag­
nitude, and over time, they can cause the concentrator to loosen and unexpectedly shift or
self-relocate to an improper working position.
Another potential cause of concentrator loosening is unstable thermal conditions.
During the heat treating cycle, the concentrator can be rapidly heated to elevated tempera­
tures, followed by intense cooling during quenching to room temperature. These repeti­
tive “heating–cooling” cycles are accompanied by corresponding “expansion–contraction”
cycles, which can cause it to loosen or be unintentionally moved to an improper working
position, altering the hardness pattern.
Practitioners should be aware that ferrites as well as ferrite-based and iron-based com­
posites usually have quite low maximum working temperatures. For example, according
to data obtained from one of the manufacturers of flux concentrators, the maximum tem­
perature of some concentrators is only 250°C for long-term use and 300°C for short-term
use. If the temperature of those composites exceeds its maximum permissible level, its
binder (different types of epoxies) starts to deteriorate and disintegrate uncontrollably
(Figure 4.171, bottom), negatively affecting the overall robustness, repeatability, and reli­
ability of the induction system.
An unexpected change in the hardening pattern can cause serious damage. In the auto­
motive industry, for example, this situation can result in the recall of many thousands of
vehicles to replace defective parts. To prevent such a situation, flux concentrators should
be examined regularly on a scheduled basis and repaired if necessary (the subject of main­
tenance of flux concentrators is discussed in Section 4.8.6). In some cases, special monitors
can be installed to indicate changes in concentrator performance; however, they may add
substantially to total system cost.
There is a common misconception that the use of flux concentrators automatically leads
to an increase in efficiency [419]. Flux concentrators improve the efficiency of the process
partly by reducing the coupling distance between the inductor and the workpiece, as
well as by reducing the stray losses (by reducing the reluctance of the air path). However,
because the flux concentrator is an electrically conductive body and conducts high-density
magnetic flux, there is some power loss generated as heat within it because of the Joule
effect and magnetic hysteresis heat. This phenomenon could cause a reduction of electrical
efficiency and the need to design a special water-cooling system to remove the heat from
the concentrator.
Heat Treatment by Induction 389

Therefore, in some applications, flux concentrators will improve efficiency; in others, no


improvement will be observed, or efficiency may even drop. For example, there is usu­
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

ally no improvement in coil efficiency when concentrators are used in applications such
as tempering. The same can be said for the electromagnetically long solenoid coils used
to heat ferrous or nonferrous billets, bars, and rods before forging, rolling, upsetting, and
extrusion or in some induction melting furnaces.
However, measurable improvements in electrical efficiency can be achieved when flux
concentrators are applied to certain types of induction coils. Examples include hairpin and
split-return inductors, electromagnetically short single-turn or multiturn coils, and induc­
tors used to heat internal surfaces, just to name a few.
An improvement in heating efficiency can be attributed to the concentrator’s abil­
ity to localize the EMF in a specific area. Typically, the major portion of the EMF will
not propagate behind the concentrator and the heated area will be localized. As a
result, the heated mass of metal will be reduced; therefore, less energy will be required
to accomplish the needed job and thus less energy will be wasted on nonproductive
heating.
Manufacturers of induction equipment have found the use of flux control technology
increasingly important in improving the quality of IHT components. Before beginning a
project, detailed mathematical modeling or laboratory tests should be conducted to deter­
mine the cost-effectiveness of using flux concentrators in a particular application.

4.8 Heat Treating Equipment Maintenance


Components of any heat treat system are subject to change, exhibiting certain vari­
ability over a given period. Preventive maintenance (PM) is a relatively simple yet very
effective way to assure specified working conditions of system components, allowing
one to avoid a costly recall of heat-treated parts and to also extend the life of induc­
tion equipment [1,162,167,175,336,423,424,428,432,433]. PM maximizes “uptime” and
minimizes the number of scrap parts. Regardless of its obvious effectiveness, PM is
often neglected, leading to premature failure of induction heating machine compo­
nents including problems associated with coils, fixtures, tooling, cooling/quenching
systems, power supply, and so on. Recognizing that each heat-treating machine has
unique features and its own PM procedures, some general guidelines for PM will be
discussed here. Although induction coils are the workhorses of the induction harden­
ing operation, they are often considered to be the weakest link in the system. The PM
program for induction coils is frequently delayed, incomplete, or simply ignored. In
order to increase equipment life expectancy, a number of PM actions must be periodi­
cally conducted.
Equipment maintenance for heat treating can be broken down into six general areas,
including power supply, heat station, water recirculating system, quench recirculating
system, heating coils, and the heat-treating machinery. The first step in planning PM is to
thoroughly read the equipment manual provided by the OEM.
The three senses vision, smell, and hearing are your best assistants in the search for
any signs of abnormal operation or deterioration and in preventing problems before they
occur. Before doing any maintenance work, please remember: SAFETY FIRST. Workplace
wellness, injury prevention, and safety programs should be given the highest priority.
390 Handbook of Induction Heating

4.8.1 Power Supply
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

The subject of power supplies is discussed in detail in Chapter 7 of this handbook.


Dust, dirt, moisture, improper torque, presence of foreign particles, and localized
“hot” spots are the most frequent causes of failures of electrical/electronic devices in
power supplies. The manufacturer’s equipment manual will usually provide the spe­
cifics for safely maintaining their equipment, but a general overview would include
the following items.

4.8.1.1 Visual Inspection
A good first step with all power turned off and equipment safely discharged is to closely
inspect electrical control connections, control boards, power connections, and electrical
components for any signs of overheating, arcing or mechanical deformation, swelling,
breakage, condensation, and so on.
All electrical insulators (e.g., capacitor insulating bushings and insulators between
current-carrying buses, etc.) should be wiped completely clean of any dirt or foreign par­
ticles to leave the surface in a “like new” condition.
If there is evidence of excessive dust or dirt accumulation inside the cabinets, then the
door gaskets may need to be replaced.
Another area to carefully inspect is the condition of the copper tube on the positive side
of DC bus cooling connections. If hoses are too short and electrolysis is in progress, por­
tions of the positive bus cooling tube will be eaten away and deposited on the negative
cooling tube. This often results when someone has decided to save some space or money
by making the hoses as short as possible.
On high-voltage vacuum tube oscillator systems, a “sacrificial” anode is often used in
order to prevent erosion of the actual fitting that the hose is connected to. The anode will
degrade over time and will need to be replaced at the end of its life as directed by the OEM
maintenance manual.
If it is necessary to flush out or clean the inverter or heat station to eliminate any buildup
in water passages, it is essential to follow the procedure outlined in the OEM maintenance
manual. Some types of fluids used in cleaning water passages are highly conductive and
should not be used on any buswork when high voltage is present.

4.8.1.2 Mechanical Test
The next step is to check the tightness of all control and power connections, capacitor hard­
ware, transformer tapping hardware, and so on.

4.8.1.3 Ohm-Meter Tests
Ohm-meter tests can be made on all power semiconductors per the manufacturer’s recom­
mendation in the equipment manual. A test should be done to ensure that no components
that should be electrically isolated are connected to ground and that no components are
shorted when they should indicate a resistance or capacitance value.
Another very helpful test at this point, since the power supply is very closely tied to the
heat station, would be to use a signal generator or load frequency analyzer to determine
the resonant frequency of the power and load circuit.
Heat Treatment by Induction 391

4.8.1.4 Power Measurements (Meter or Oscilloscope)


Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

• Control Circuit—DC voltage supplies to control components should be tested to


ensure the proper value of DC voltage and that the ripple of the DC voltage is at
an acceptable level. (Very large ripple voltage on a DC supply can signal the failure
of a filtering capacitor.)
• Power Circuit—Carefully follow the manufacturer’s recommendation in the equip­
ment manual to safely measure the voltage and current in the power circuit with
an oscilloscope or an appropriate True-RMS reading voltmeter.

4.8.1.5 Meters and Limit Lamps


Observe the power supply meters and limit lights during operation of the system and
compare the readings to the data recorded on the data sheet. (Note: It cannot be stressed
enough that complete data for all settings of the power supply, heat station, and machine
must be carefully documented in order to quickly ascertain any differences that occur
when a problem arises. Often, a simple comparison to the standard data sheet can allow a
quick diagnosis of what otherwise could be a costly and time-consuming problem.)

4.8.1.6 Audible and Visual Observation


During the operation of the power supply, it is very important to listen to and care­
fully observe the operation of the power supply. If any unexplained “ticking” sound or
erratic operation begins to occur, it can be a sign of electrical arcing or the failure of a
component.
Another helpful instrument would be an infrared camera. The use of infrared cameras
is an essential and widely accepted method for preventive inspection of electrical devices
and circuits, useful for determining abnormal conditions, poor connections, and various
hot spots on enclosures, cabinets, bus networks, cables, and components. As an example,
Figure 4.172a shows the thermal image of an area inside the power supply cabinet display­
ing the temperature of various hot spots on enclosures, bus bars, wiring, or other electrical
components.

26.8
25.5
24.1
22.8
21.5
20.2
18.8
17.5
16.2
14.8
13.5
°C
(a) (b) (c)

FIGURE 4.172
Thermal images inside a power supply cabinet displaying the alarming temperature of various hot spots.
(a) Inside power supply, (b) overheated loose bolt, and (c) discolored and corroded loose bolt.
392 Handbook of Induction Heating

If any measurements are to be made with enclosure doors open, the manufacturer
should be contacted and only qualified personnel with proper personal protective equip­
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

ment may perform the service closely following recommended procedures and applying
safety principles. Remember: SAFETY FIRST.

4.8.2 Heat Station, Bus Connections, and Inductor Foot


4.8.2.1 Visual
With all power turned off and equipment safely discharged, closely inspect electrical con­
trol connections, control boards, power connections, and electrical components for any
signs of arcing, mechanical deformation, discoloration, swelling, breakage, and so on.
Figure 4.173 shows two examples of arcing at or near a bus connection that could be
caused by insufficient torque on bolts holding the buswork or a breakdown of insulation
between two high-voltage bus connections. Any failures of this type must be thoroughly
cleaned and repaired to leave no trace of the carbon tracks generated by the arcing prob­
lem. If the area is not completely cleaned and repaired, it will simply become an area of
repeated failure. All electrical insulators (e.g., capacitor insulating bushings and insulators
between current-carrying buses, etc.) should be wiped completely clean of any dirt or for­
eign particles to leave the surface in a “like new” condition.
Figure 4.119 shows an example of arcing in an inductor foot caused by loose bolts com­
bined with dirt accumulation. Loose bolts resulted in an increased electrical resistance in
the bolted contact area, localized overheating, and arcing.
Figure 4.172b shows the thermal image of an alarmingly high temperature because of an
insufficiently tightened electrical connection (loose bolt) between a cable and bus network.
Discoloration of the overheated bolt is shown in Figure 4.172c.
These examples clearly illustrate the importance of having appropriate and reliable electrical
contacts. It is essential for fasteners to be tight enough to do their job—but not too tight. Quite
often, fasteners are tightened based on subjective judgment: It simply “feels” sufficiently tight.
In reality, insufficient tightening or overtightening takes place. Neither is desirable. This can be
prevented if properly calibrated torque wrenches are used for bolted joints.
Another vital step in preventing failures related to inappropriate electrical connections is
proper and regular maintenance, including thorough inspection of the coil and bus network
(including the buses themselves, inductor adapters, and bus extensions). While performing the
inspection, all power should be turned off and equipment should be safely discharged.
It is also imperative to consider the EMF in the immediate vicinity of the inductor. The
use of low-resistivity nonferrous materials or, where feasible, electrically nonconductive

FIGURE 4.173
Examples of arcing at or near a bus connection.
Heat Treatment by Induction 393

materials in the area close to the inductor is recommended. Fasteners and washers used
to provide electrical connections should be of the appropriate size and type (according to
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

OEM specifications) and must be nonferrous (e.g., silicon bronze or nonmagnetic stainless
steels). In some rare cases of low-stress applications, where nonmagnetic metallic bolts,
nuts, and washers do not work (e.g., because of electromagnetic overheating or excessive
corrosion), ceramic, polymer, and other plastic fasteners can be used. As a rule of thumb,
any ferrous materials should be located at least one coil diameter away from the coil. More
accurate recommendations can be obtained by numerical computer modeling.
Electrical insulators should be checked periodically for holes, wear, or contamination.
Otherwise, arcing may develop. Depending on its magnitude, a ground fault detector sig­
nal may or may not shut down the machine. It is not always easy for an operator or elec­
trician to see or hear an occurrence of arcing. Sometimes, after the appearance of a series
of occasional ground faults, the decision may be made by personnel to reduce ground
fault sensitivity. This may temporarily fix the problem of shutting down the machine.
However, this action may trigger more severe problems in the future and destroy equip­
ment. Therefore, it should not be done. It is more appropriate to immediately contact the
OEM for inspection and an appropriate fix for the problem.
Appearance of the ground fault signal may also indicate unspecified coil-to-workpiece
positioning and the presence of a physical contact between the workpiece and inductor
(i.e., the workpiece or flux concentrator may be shifted into an inappropriate position).
Therefore, if the ground fault signal appears, it is important to review all factors that may
cause its appearance.
Coil terminals and quick-disconnect devices should be checked for flatness, traces of
burns, and arc marks. It is imperative to keep contact faces clean and free of dust, oil, and
debris. The presence of contaminants and uneven surfaces results in poor electrical con­
tacts that are associated with localized overheating.

4.8.2.2 Mechanical Test
The next step is to check the tightness of all control and power connections, capacitor hard­
ware, transformer tapping hardware, and so on.

4.8.2.3 Ohm-Meter/Capacitor/Meter/Load Frequency Analyzer


Ohm-meter tests can be made on all power components per the manufacturer’s recom­
mendation in the equipment manual. A test should be done to ensure that no components
that should be isolated are connected to ground or that no components are shorted when
they should indicate a resistance or capacitance value.
At this point, similar to the above discussion, a helpful test would be to use a signal
generator or load frequency analyzer to determine the resonant frequency of the power
and load circuit.

4.8.2.4 Audible and Visual Observation


During the operation of the power supply, it is very important to listen to and carefully
observe the operation of the heat station. If any unexplained “ticking” sound or erratic
operation begins to occur, it can be a sign of electrical arcing or the failure of a component.
As previously discussed, another helpful instrument at this time would be an infrared
camera to display the temperature of various hot spots on heat station enclosures, bus bars,
394 Handbook of Induction Heating

or components. If any measurements are to be made with enclosure doors open, the manu­
facturer should be contacted to advise the technician with respect to safety equipment and
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

recommended procedures.

4.8.3 Water Recirculating System


4.8.3.1 Visual
A good visual observation would include checking for water leaking on the floor or in other
areas that should not be wet and checking the level of the water in the recirculating system
as well as the appearance of the water leaks. It should be clean with no fungi or algae form­
ing in the tank. Hoses and hose connections should be checked regularly, making sure that
there are no leaks and that hoses are not kinked or clogged but are free to provide continuous
water flow at the flow rate recommended by the manufacturer. With time, the hoses inevita­
bly become stiff and breakable, requiring replacement. Before replacing the hoses, the OEM
manual should be reviewed and recommended types of hoses should be used. Make sure that
hose bundles do not have crimps or other irregularities that may restrict water flow.
Any filters should be checked to make sure that they are clean and free of debris. Any
covers for the tank should always be in place to prevent contamination except for times
when maintenance is being done on the system.

4.8.3.2 Meter Measurement
The water level should be checked regularly and, if required, water of appropriate qual­
ity according to the instructions of the OEM (including, pH, conductivity, etc.) should be
added. The quality of city water and well water is typically inappropriate for cooling an
induction coil and electrical/electronic devices.
The conductivity of the water should be measured with a conductivity meter to ensure
that it is within the manufacturer’s specifications for the power supply and heat station
components. On systems with a DC link in the power system, the water conductivity can
be a critical parameter to prevent damage and equipment failure owing to electrolysis in
water lines. A typical starting level might be 10 μSiemens/cm (μmhos/cm) with water in
the tank changed when it reaches 100 μSiemens/cm. On systems without a DC link, a level
of 400 μSiemens/cm may be more typical.
In some cases where extremely cold temperatures are possible, it may be necessary to
add anti-freeze protection to the water system. It should be noted that this is not standard
automotive antifreeze. It should be propylene glycol or a very low conductivity ethylene gly­
col at the appropriate concentration or anti-freeze specified by OEM to provide protection
of the equipment from freezing water in pipes, hoses, or water-cooled components. On
closed systems, it is also very important to make sure that there is no trapped air in the
water system; the system must be purged of air if necessary.
Note that if proper freeze protection is not provided, the freezing of water in thin-wall
copper tubes is often sufficient to rupture the tube, resulting in leaks that are very difficult
to locate and repair. In this case, prevention is the best cure.

4.8.3.3 Pressure Flow and Temperature Measurement


With the system operating, the value of pressure, flow, and temperature should be mea­
sured and compared to the standard data sheet for the equipment. Safety switches for
Heat Treatment by Induction 395

flow and pressure should be tested to ensure proper operation. If a high-pressure pump
is used for the inductor cooling, separate tests should be done for pressure and flow after
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

the high-pressure pump.

4.8.3.4 Audible and Visual Observation


The system should be observed during normal operation to evaluate any excessive vibra­
tion or noise coming from water pumps or piping. If the water tank begins to overflow dur­
ing normal operation, it may be attributed to a failure in a water-to-water heat exchanger
or a fill valve that may be stuck in the open position.
All hosing should be color coded for supply and return lines and fitted with the proper
connectors that cannot be confused with or inserted where a quench line should be con­
nected. Preferably, the so-called foolproof connections should be used to avoid misunder­
standing and misinterpretations.
Again, a key factor in the observation stage is to have accurate data for normal operation
of the equipment that will flag anything that is operating differently from normal.

4.8.4 Quench Recirculating System


4.8.4.1 Visual
A good visual observation would include checking for quench leaking on the floor or in other
areas that should not be wet and checking the level of the quench in the recirculating system
as well as the appearance and smell of the quench. It should be clean with no fungi or algae
forming in the tank, no hydraulic oil film collecting at the surface, and excessive foaming [428].
Any filters and oil skimmers should be checked to make sure that they are clean and
free of debris. Excessive amount of scale flakes, dirt, foreign particles, and other contami­
nants found in the tank may indicate the failure of a bag filter. Scale flakes from heat-
treated parts can clog quench holes of spray quenching devices (Figure 4.174). They are
very abrasive and can cause damage to inductors, spray quench devices, pumps, or piping
when pumped around the system at high velocity. Electrically conductive particles can be
deposited by quenchants in the edge regions of buses and between coil turns, resulting in
arc development.
The presence of the transmission oil and lubricants in the quench system can decrease
the effectiveness of the quench cooling, lead to excessive foaming and inconsistent quench­
ing, and result in irregular hardness readings. Any covers for the tank should always be
in place to prevent contamination except for times when maintenance is being done on the
system. An unusual smell may indicate an uncontrolled growth of biological contamina­
tion in the quench tank.

FIGURE 4.174
Scale flakes from heat-treated parts can clog quench holes of spray quenching devices.
396 Handbook of Induction Heating

4.8.4.2 Meter Measurement
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

The concentration of the quenchant should be measured with a refractometer to ensure


that it is within the manufacturer’s specifications for the cooling rate required.

4.8.4.3 Pressure Flow and Temperature Measurement


With the system operating, the value of pressure, flow, and temperature should be mea­
sured and compared to the standard data sheet for the equipment. Safety switches for flow
and pressure should be tested to ensure proper operation. Often, these data are recorded
on a PLC or HMI for an accurate history of normal operating performance.

4.8.4.4 Audible and Visual Observation


The system should be observed during normal operation to evaluate any excessive vibra­
tion or noise coming from quench pumps or piping. If the quench tank begins to overflow
during normal operation, it may be attributed to a fill valve being stuck in the “open” posi­
tion. All hosing should be color coded and fitted with appropriate connectors.
Often, the first indication of a problem with the quench system may be inadequate hard­
ness or pattern on the part as a result of one of the above parameters drifting out of range.
A systematic visual observation of the operating parameters with comparison to normal
values could head off many of these types of problems before they ever reach the stage of
stopping the production line to correct an “unknown” problem.
Again, a key factor in the observation stage is to have accurate data for normal operation
of the equipment that will flag anything that is operating differently from normal. Other
aspects of quench maintenance have been already discussed in Section 4.2.7.

4.8.5 Heat Treating Inductor Maintenance and Storage


Heat treating inductor maintenance can be summarized in five categories: maintain­
ing the tooling and part holding devices, keeping the inductor clean, providing a visual
inspection for monitoring a deterioration process, maintaining good electrical contacts,
and maintaining spares [1].

4.8.5.1 Consistent Workpiece Holding


Because heat-treating coils that are made from copper are an electrical as well as a mechan­
ical device and are in proximity to hot parts and quenchants, they are subjected to deterio­
ration and wear. The tooling, which holds the workpiece during heating, should be robust,
corrosion resistant, and temperature resistant, and provide a consistent inductor-to-part
location. Care should be used when manually loading parts, to avoid damaging the induc­
tor. Operational specifics of several inductor styles were discussed in previous sections.

4.8.5.2 Keeping the Inductor Clean


Most hardening applications produce some amount of scale from the steel part as it is heated
to temperatures in the scale-forming range and quenched. The scale (small metallic flakes)
combined with the quench medium can stick or cling to the inductor. If this is allowed to
build up on or within the coil, it can cause arcing or premature failure of the inductor. Some
Heat Treatment by Induction 397

parts may have oil residuals that can produce intense smoke or even flame during heating.
This typically accelerates copper surface degradation (i.e., erosion). Wiping or washing off
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

the coil on a regular basis is recommended. The application and its cleanliness will deter­
mine the frequency of cleaning. A mild soap can be used to clean the inductor. If a brush is
used, it should have plastic bristles; a metallic wire brush can leave metal bristles and other
stress risers in the coil copper, which could reduce the service life of the coil.

4.8.5.3 Visual Inspection
Routine visual inspections should be done to look for deterioration or arcing of the coil.
High-powered inductors can develop copper cracks owing to stress corrosion or stress
fatigue in the heating face; this is more likely to happen when the inductor has quench
holes drilled through the heating area of the coil. Microcracks in the coil will obstruct the
AC flow and cause an accelerated loss of efficiency, because the current will have to travel
around the crack (see Section 4.2.4.3 and Ref. [331]). This will be more noticeable at higher
frequencies, because it is like having a larger equivalent air gap between the workpiece
and the inductor and increases electrical resistance of coil copper. If the crack in the induc­
tor becomes large enough, it will eventually cause a water cooling leak and the inductor
will be unable to process parts.

4.8.5.4 Maintain Electrical Contacts


There are two basic ways to install an inductor: either several bolts are used or, for frequent
changeover, a quick-change type can be used.
For many heat-treating applications, 10- to 12-mm (⅜- to ½-in.) brass, silicon bronze (pref­
erable), or nonmagnetic stainless steel bolts are used for the inductor-to-bus connection. It
is commonly specified that the stainless steel bolts should be tightened to approximately
40 to 45 lbf/ft. Overtightening of the bolts can pull the threads from the soft copper, but
undertightening will not produce a good electrical contact and lead to overheating, arc­
ing, and premature failure (Figure 4.119). Scheduled inspection of the fasteners to check
for tightness is important. Higher power and lower frequencies require a more frequent
inspection for tightness, as a result of more vibration when the heat is applied.
If a tapped hole is to be used at the connection point, stainless steel threaded inserts
should be utilized to maintain the threaded hole. Heavy-duty washers are recommended
to ensure that adequate force is applied in the right areas and to distribute the force over a
larger area. Figure 4.172b shows the results of not using the proper washers.
In some rare cases, low-stress applications, coil fixturing for example, where metal bolts,
nuts, and washers do not work (e.g., because of electromagnetic overheating or excessive
corrosion), ceramic, polymer, and other plastic fasteners can be used. There is no measur­
able heating effect of ceramic fixtures by induction. Besides, it can withstand quite high
temperatures (easily exceeding 1000°C) and, being exposed to many corrosive environ­
ments, is practically immune to corrosion. Unfortunately, the use of ceramic fasteners is
associated with a danger of its brittle fracture. Ceramics have extremely low resistance to
torque or impact. Because of its extreme brittleness, it is tricky to install.
The use of fasteners made from polymer and plastic materials (nylon, PTEE, Delrin,
PEEK, etc.) is restricted by low torque and, in particular, temperature restrictions. For
example, the maximum applicable temperatures for the above-discussed materials are as
follows: nylon, 65°C–70°C; Derlin, 80°C–85°C [430]; PEEK, 245°C–250°C [431].
398 Handbook of Induction Heating

It is important to be aware that some of those materials might be hygroscopic absorbing


of moisture (e.g., from liquid quenchants) leading to appreciable swelling. Some of them
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

have poor chemical resistance that causes their degradation owing to contact with liquid
quenchants. Therefore, carefully review the manufacturer’s data sheet before using those
fasteners.
When using a quick-change arrangement, it is important to maintain and inspect the
clamping features such as springs and stainless steel clamp bars.
In all cases the electrical connection surfaces must be clean and free from various elec­
trically conductive deposits (i.e., scale, graphite particles, etc.), nicks, burrs, scratches, and
arc marks. Sandpaper should not be used to clean electrical contacts. Inductor and bus
mounting feet are carefully machined for an accurate interface and hand sanding can lead
to high and low areas that can inevitably redistribute contact area and current density
redistribution, producing hot spots and subsequent failures.
Sometimes, a section of copper connection surface is heavily oxidized and covered with
what appears to be a “greenish” powder. This copper oxide has a much higher resistance
than clean copper, and if it is found on a coil-to-bus or a bus-to-bus connection, it can also
lead to localized overheating and failure of the coil or bus at that point.
Besides keeping the surfaces clean, providing good electrical contacts, coating with
thin silver plating is often done in an effort to keep the resistivity low at the connection
joint. In addition, the plating prevents the copper from corroding, which could cause
poor contact.
An often overlooked area of the coil is the electrical insulator or any thermal insulator
(refractory type of liners) used. Electrical insulators should be periodically removed from
a coil in order to check for any signs of arcing at “feed through” holes in the coil terminals
or buses. If a refractory type of liner is used, it should be cleaned out often and checked
for cracks to ensure that no hot scale comes in contact with coil turns that could lead to a
short circuit.

4.8.5.5 Maintaining Spares
As with any critical component in production, it is important to have and maintain spares
to minimize equipment downtime. It is recommended to have at least two of each induc­
tor on hand. In the event that one inductor is being repaired, there will still be one on the
equipment and one spare in storage. It is important to keep inductors stored (when not
in operation) in a way that they stay clean and undamaged. If adequate protection is pro­
vided, the shipping containers can often be used as a storage device. The container also
provides a label to describe the contents; the label includes vendor identification numbers
and the customer’s part and tooling number. These are used not only for inductor storage
but also to ship the inductor back and forth for any maintenance repairs [423,424]. Because
of the cost and value of inductors, it sometimes helps to remember that they should be
handled as carefully as one would handle a thousand dollar bill (i.e., it would not simply
be left lying around on the floor!).

4.8.6 Flux Concentrators
Sometimes, magnetic flux concentrators (ferrites, iron-based and ferrite-based powder
bonded materials, and laminations) are added to the induction coil. The most typical fail­
ure mode of the flux concentrators is their degradation as a result of overheating or chemi­
cal attacks.
Heat Treatment by Induction 399

There are three causes of concentrator overheating [336]:


Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

1. Heat generated within a concentrator (owing to Joule losses and magnetic hyster­
esis), particularly when high coil current densities and frequencies are used. The
binders that hold particles together in magnetic powder concentrators are often
called dielectric binders. As a result, some mistakenly assume that a magnetic flux
concentrator is a dielectric body (electrically nonconductive); thus, no Joule heat is
generated. In reality, they are far from being dielectric; their electrical resistivities
are not infinite by any stretch of imagination but have some appreciable values.
Therefore, because they are electrically conductive objects, there will be some heat
generated within them under an applied AC field.
2. Excessive heat transfer from the surface of the heated workpiece (owing to thermal
radiation and heat convection). In the majority of induction hardening applications,
the coil-to-workpiece air gap is typically in the 1.6- to 9-mm range (greater gaps
are associated with lower frequencies). Because hardening temperatures are usually
within the 845°C to 980°C (1550°F to 1800°F) range (though in some cases tempera­
tures can be even higher), there can be appreciable heating of a flux concentrator
that is located in proximity to the hot surface of the workpiece. Different types of
epoxy-based binders are often used for the fabrication of composite concentrators.
Any epoxy is a relatively poor thermal conductor with limited ability to provide
sufficient cooling of hot spots on a concentrator surface. When short heat times are
applied in combination with relatively low duty cycles, the heat flow within a con­
centrator might be sufficient to minimize the negative impact of the localized heat
surplus. However, a combination of increased duty cycles and longer heat times (e.g.,
scanning elongated workpieces utilizing relatively slow scan rates) might lead to
excessive surface overheating of flux concentrators, degradation of the epoxy-based
binder, and its eventual premature failure because of the material’s disintegration.
3. Heat transfer from the inductor copper to which the flux concentrator is attached
owing to thermal conduction. Regardless of the fact that most coils are water
cooled, there are cases where there are physical restrictions to locating water-
cooling pockets sufficiently close to the current-carrying face of the coil copper.
Although it is not desirable, because of certain design restrictions, the distance
between water-cooling pockets and coil copper current-carrying surface may
exceed 5–6 mm. Profiled inductors for heating fillets of shafts can serve as an
example. Temperatures of the copper surface of a working inductor can reach
temperatures that exceed maximum allowable temperatures of flux concentrators.
Thus, it is critical that coil copper working temperatures are held below maximum
working temperatures of the flux concentrator, to avoid heating of a concentrator,
potentially causing its premature failure.

Although laminations can withstand much higher working temperatures than pow­
der-based composite concentrators, they can still be overheated, particularly in higher-
frequency applications. Laminations are also sensitive to aggressive environments, such
as quenchants. Rust and degradation can result (Figure 4.171, top right).
Another major concern with any flux concentrator is the reliability of its installation.
This subject was discussed earlier in Section 4.7.3.4.
Flux concentrators are usually positioned in areas of high magnetic flux density, where
localized magnetic saturation may occur and the magnitude of electromagnetic forces can
400 Handbook of Induction Heating

be substantial. Over time, these forces can cause the concentrator to loosen and unexpect­
edly shift out of position, negatively affecting the heat pattern.
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

Taking these into consideration, it is very important to regularly review the condition
and positioning of magnetic flux concentrators, making sure that no discoloration, exces­
sive degradation, or unexpected repositioning has taken place.

4.8.7 Heat-Treating Machinery Maintenance


Although the hope of every manufacturer is to provide equipment that will last indefi­
nitely, machines do lose their effectiveness, utility, or economy over time. An evaluation of
machine failures over time has indicated that there are three main reasons for the loss of
usefulness of a machine. Studies show that seventy percent of machines lose their useful­
ness because of surface degradation. Fifteen percent of the machines lose their usefulness
due to obsolescence, and 15% lose their usefulness as a result of some type of accident that
renders them unusable [425–427]. The 70% that lose their usefulness as a result of surface
degradation can be further broken down into two categories. The first category is corro­
sion (20%) and the second is mechanical wear (50%). Often, a well-intentioned repair crew
can cause irreparable damage to a piece of equipment due to a misdiagnosis or improper
repair technique.

4.8.7.1 Visual
For the reasons stated above, a good visual inspection of the equipment can reveal and
result in the correction of many situations that could lead to premature failure of machin­
ery. A good visual inspection would include looking for leaks, contamination of lubri­
cants, corrosion on sliding components, mechanical wear attributed to interferences or
insufficient lubrication, and misalignment of components during machine setup or attrib­
uted to a machine jamming condition. Moving components of the equipment should be
sufficiently and properly lubricated according to OEM instructions.
A misalignment of the workpiece during the heating cycle may be something that is
a minor consideration for the part handling machine but could lead to the failure of an
expensive heating coil or tooling. Recognition of these types of potential problems dur­
ing a cursory visual inspection can be very economical in terms of preventing the loss
of time and money that may have been incurred by an unexpected breakdown of the
machine.

4.8.7.2 Audible and Visual Observation


Audible and visual observation of the machine during production operation should
include looking for vibration, erratic linear or rotary movement of cylinders and motors,
index tables, noise in pumps, binding of bearings, and so on. A hydraulic pump that is
experiencing cavitation will sound as though it is pumping marbles. This can lead quickly
to overheating and destruction of the pump due to excessive vibration and failure of the
seals.
The positioning accuracy of the workpiece within the inductor during the heat treat
cycle is a primary concern in order to ensure proper heating and quenching of the part.
Positioning accuracy is also important to prevent damage to the heating coil and tooling
from accidental contact and abuse.
Heat Treatment by Induction 401

Any wobbling of the workpiece during heating can result in certain areas of the part
being closer to the coil during heating and to the quench block during quenching. This can
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

result in localized overheating or inappropriate localized quench severity. Worn or dam­


aged components (such as bearings, centers, etc.) should be replaced.
Any mechanical stops or “dogs” that are instrumental in positively driving the part to
ensure rotation during heating should be regularly inspected and repaired if necessary.
Figure 4.79e illustrates what can happen when the center is rotating but the workpiece is
not. In this case, the contact friction between the part and the center was insufficient to
ensure the needed rotation of the workpiece.
Gages and warning lights should be observed to ensure that hydraulic systems are oper­
ating within the required flow and pressure limits.
An infrared camera could be used to measure the temperature of the various compo­
nents of the system to locate areas of concern. It is best to record and save data for normal
operation when the equipment is initially set up in order to compare and identify any
changes that have occurred in the operating parameters or temperature readings. Process
recipes should be backed up regularly and reliably.

4.8.7.3 Safety Circuits
IHT machines will have a variety of safety circuits to prevent improper or erratic opera­
tion. One of the most critical is called the inductor ground circuit. This circuit is most often
connected to the heating coil and gives immediate indication of the part touching the heat­
ing coil. This should never occur in the normal operation of the system. If it does occur, the
movement of the system is inhibited until the workpiece is physically moved away from
the inductor via an operator override circuit in the machine control.
The inductor ground circuit should not be tested by running the workpiece into the coil,
but by using a small jumper wire to simulate a part touching the coil. This prevents any
damage to the coil or tooling during the testing phase.
From time to time, the sensitivity of a ground circuit may need to be adjusted due to the
conductivity of the quench medium but care should be taken to ensure that this circuit is
never disabled during machine motion.
Needless to say, it is never advisable to jumper out a safety switch in order to see if it is
working properly. The better course is to test the switch outside of the circuit by measuring
and verifying its operational parameters. Always remember: SAFETY FIRST.

4.9 Review of Selected IHT Applications


4.9.1 Gear, Pinion, and Sprocket Hardening
The traditional gear heat treatment is gas carburizing in batch or continuous furnaces for
an extended time and then oil quenching. The gears are typically low-temperature fur­
nace tempered for a minimum of 1 h at temperature. After carburization, the carbon con­
tent, which is typically within the 0.7%–1.0% C range, is distributed nonuniformly within
the case depth, having a higher concentration at the surface and then gradually decreas­
ing from the surface toward the subsurface area. Post–heat treat processing may include
grinding and shot-peening to improve fatigue strength.
402 Handbook of Induction Heating

Over the years, gear manufacturers have increased their knowledge of the production
of quality gears and gear-like components. This knowledge has led to many improve­
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

ments, including lower noise, lighter weight, and lower cost, as well as increased load-
carrying capacity to handle higher speeds and torque with a minimum amount of
generated heat.
In contrast to carburizing and nitriding, induction hardening does not require heating
the whole gear or pinion. With induction, the heating can be primarily localized to the
areas where metallurgical changes are desired (e.g., the flank, root, and gear tip can be
selectively hardened) and the heating effect on adjacent areas is minimum [1,434–438].
Selective hardening of specific areas of gear teeth producing a fine-grain martensitic layer
helps not only to optimize gear performance but also to minimize distortion.
Improvement in wear resistance, contact fatigue strength, endurance, and impact
strength helps eliminate premature gearbox failure. Another goal of induction gear hard­
ening is to produce considerable compressive residual stresses at the surface and in the
subsurface region. Compressive stresses help inhibit crack development and resist tensile
bending fatigue. Depending on the application, tooth hardness typically ranges from 48
to 62 HRC.
Not all gears and pinions are well suited for induction hardening. External spur and
helical gears, worm gears, internal gears, racks, and sprockets are among the parts that are
typically induction hardened (Figure 4.175).
As mentioned earlier, it is not always possible to obtain a fully martensitic case depth.
Depending on the steel chemical composition, the presence of a certain amount of RA
within the case depth is unavoidable (unless cryogenic treatment is used). Up to a certain
point, some amount of RA does not noticeably reduce the surface hardness. However, it
might bring some ductility and provide better absorption of impact energy, which could
be imperative for heavily loaded gears. In addition, because of its unstable nature, RA may,
with time, transform into martensite. It introduces additional compressive residual stresses
and increases the surface hardness. From this perspective, a small amount of RA may not
only be harmless but might even be considered beneficial in some cases. However, in the

FIGURE 4.175
Spur and helical gears, worm gears, internal gears, racks, and sprockets are among the parts that are typically
induction hardened. (Courtesy of Inductoheat Inc., an Inductotherm Group company.)
Heat Treatment by Induction 403

great majority of applications, an excessive amount of RA is undesirable and can even be


detrimental because it may noticeably reduce the surface hardness, weaken bending fatigue
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

properties, decrease wear resistance and the magnitude of useful compressive residual
stresses within the case depth, promote shape distortion, and can result in the appearance
of a crucial amount of brittle untempered martensite during the gear’s service life.
The kind of steel, cast iron, or P/M material; their prior microstructure; and the required
gear performance characteristics (including load condition and operating environment)
dictate the needed surface hardness, core hardness, hardness profile, and needed residual
stress distribution.

4.9.1.1 Material Selection and Required Gear Conditions before Heat Treatment


Several guidelines, recommendations, and standards help select an appropriate material
for gears. This includes the following:

• ANSI/AGMA 2004-C08 standard: “Gear Materials, Heat Treatment and Processing


Manual”
• AGMA 923-B05 standard: “Metallurgical Specifications for Steel Gearing”
• ASTM A536-84 standard: “Specification for Ductile Iron Castings”
• ASTM A48/A48M-03 standard: “Specification for Gray Iron Castings”
• ANSI/AGMA 6008-A98 standard: “Specifications for Powder Metallurgy Gears”
• Heat Treater’s Guide: Practices and Procedures for Irons and Steels, ASM Int’l.

Since induction hardening does not change the chemical composition of steel, the steel
grade must have sufficient carbon and alloy content and be capable of achieving a certain
surface and core hardness. Low-alloy and medium-carbon steels with 0.4% to 0.55% C (e.g.,
SAE 1040, 15B41, 4140, 4340, 1045, 4150, 1552, and 5150) are commonly used in gear harden­
ing [19,28–30,400]. H-steels (1050H and 4340H) are preferred because they produce more
consistent results. In some cases, high-carbon steels (1065, 1080, and 52100) are used, as
well as MSS, cast irons, P/M components, and proprietary microalloy steels.
It is particularly important to have “favorable” prior microstructure when hardening
gears. Repeatability and the stability of the hardness pattern are grossly affected by the con­
sistency of the microstructure before hardening and by the steel’s chemical composition.
A homogeneous fine-grained Q&T initial microstructure with a hardness range of 30
to 36 HRC leads to a fast and consistent steel response to induction hardening. Section 4.1
provides a discussion on the response of various prior structures to rapid induction
hardening.
In contrast to a Q&T prior microstructure, steels with large carbides (i.e., spheroidized
microstructures) have poor response to induction hardening and also result in the need
for prolonged heating, requiring higher temperatures for austenitization. A combination
of high temperatures with longer heat time leads to grain growth, the formation of coarse
martensite, data scatter, an extended transition zone, and excessive gear distortion. Coarse
martensite has a negative effect on tooth toughness and impact strength.
These are the reasons why steels under normalized and Q&T conditions are commonly
used in the majority of induction gear hardening.
As opposed to other heat-treating techniques, hardening by induction is noticeably
affected by variations in the steel’s chemical composition. Therefore, a favorable initial
metal condition also includes tight control of the specified chemical composition of steels
404 Handbook of Induction Heating

and, in the case of cast irons, control of the chemical composition, matrix, as well as the
shape, size, and the distribution of graphite particles. Wide compositional limits cause sur­
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

face hardness and case depth variation. However, tight control of the composition reduces
possible variation of the results of hardening that could be caused, among other factors,
by using materials from multiple sources. It is highly desirable to use homogeneous (both
chemically and structurally) prior structures without excessive segregation and banding.
The surface condition of the gear is another factor that can have a distinct effect on gear
performance and its heat-treating practice. Voids, microcracks, notches, and other surface
and subsurface discontinuities as well as stress concentrators can initiate crack develop­
ment during rapid induction hardening when the metal goes through the expansion–­
contraction cycle; thermal gradients and stresses can reach critical values and “open”
notches and microcracks.
Medium and high frequencies have a tendency to overheat sharp corners and edges;
therefore, gear teeth should be generously chamfered (if possible) for optimum results.
Because gears provide transmission of motion and force, they belong to a group of the
most geometrically accurate power transmission components. A gear’s geometrical accu­
racy and ability to provide a required fit to its mate greatly affect gear performance char­
acteristics. Typical required gear tolerances are measured in microns; therefore, the ability
to control such undesirable phenomena as gear warpage, ovality, conicality, out-of-flatness,
tooth crowning, bending, growth, shrinkage, and the like plays a dominant role in provid­
ing quality gears. This is why hardness pattern consistency, minimum shape/size distor­
tion, and its repeatability are among the most critical parameters that should be satisfied.
Gears are often manufactured with holes to reduce weight. In induction hardening of
gears with internal lightening holes, including hubless spur gears and sprockets, cracks
can develop below the case depth in the inner hole areas. This crack development results
from an unfavorable stress distribution during or after quenching. Proper material selec­
tion, improved quenching technique, and modification in gear design or the required
hardness pattern can prevent crack development in the lightening hole areas.

4.9.1.2 Overview of Hardness Patterns


The first step in designing a gear hardening machine is to specify the required surface
hardness and hardness profile. Insufficient hardness as well as an interrupted (broken)
hardness profile at the tooth contact areas will shorten gear life owing to poor load-
carrying capacity, premature wear, tooth bending fatigue, rolling contact fatigue, pitting,
and spalling, and can even result in some plastic deformation of the teeth.
A through-hardened gear tooth with a hardness reading exceeding 60 HRC may be
sometimes too brittle, exhibiting lack of toughness and potentially causing a premature
brittle fracture. Hardened case depth should be adequate (not too large and not too small)
to provide the required gear tooth properties.
There is a common misconception that a uniform contour profile is always the best pat­
tern for all gear hardening applications. It is not. In some cases, a certain hardness gradi­
ent can provide a gear with superior performance, although, in many applications, a true
contour hardening pattern is highly desirable.
Among other factors, operating load condition (whether there are occasional, intermit­
tent, or continuous loads) has a pronounced effect on the tooth geometry and hardness
profile. Loads lasting up to 30 min/day are considered occasional loads. Loads lasting
several minutes per hour are considered to be intermittent-type loads. Continuous loads
last from 10 to 24 h [439,440].
Heat Treatment by Induction 405

Let us briefly evaluate a variety of hardening patterns achievable with induction hard­
ening (Figure 4.176) and their effect on gear load-carrying capacity and life [441,442].
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

Pattern a is a flank hardening pattern that has been used since the late 1940s for hard­
ening large sprockets and gears (with teeth modulus of eight and larger). The hardened
pattern occupies the tooth flank area and ends before the tooth fillet. This pattern might
provide the required wear resistance, but the typical failure mode in this case is bending
fatigue. In the hardened–nonhardened transition region of tooth fillet or root land, the
residual stresses change from compression in the hardened area to tensile in the nonhard­
ened area. The maximum tensile residual stresses are located just below the end of the
hardened pattern. A combination of applied tensile stresses with tensile residual stresses
creates a favorable condition for early crack development in the root/fillet area, particu­
larly for moderate and heavily loaded gears. Therefore, a mechanical hardening (i.e., roll
or ball hardening) of the roots is required, developing the needed compressive residual
stresses that will resist bending fatigue. However, it dramatically increases the cost, and it
is normally preferable to use alternative hardness profiles such as Pattern g or h.
Pattern b is a flank and tooth hardening pattern. This pattern has a shortcoming simi­
lar to the previous one, featuring poor load-carrying capacity, yet might be used in cases
where wear resistance is of primary concern. Patterns e, f, and g provide better results
when a combination of wear, tear, and bending fatigue resistance is required.
Pattern c is a tooth tip hardening pattern. In this case, the gear has minimum shape distor­
tion. However, the application of gears with this pattern is extremely limited because the two
most important tooth areas (flank and root) are not hardened. As a matter of fact, because of
the unfavorable residual stress distribution, the bending fatigue strength of a gear with this
pattern, as well as Patterns a and b, can even be 25% lower compared to the gear strength
before hardening (“green” gear) [443]. In most cases, Patterns f and g would be better choices.
Pattern d is a root hardening pattern. The maximum bending stresses are located in the
tooth fillet area; therefore, this pattern provides good fillet/root strengthening. Since the
root is reinforced, the maximum tensile residual stresses are shifted far away from the
surface to a depth where tensile residual stresses will not complement the tensile applied

(a) (b) (c) (d)

(e) (f ) (g) (h)

FIGURE 4.176
Variety of hardening patterns achievable with induction hardening. (a) Flank hardening; (b) Flank and tooth
hardening; (c) Hardening of tooth tip; (d) Root hardening; (e) Hardening of entire tooth and root; (f) Profile
hardening. Non-uniform pattern; (g) Profile hardening. Uniform pattern; (h) Flank and root hardening. (From
V. Rudnev, D. Loveless, R. Cook, M. Black, Handbook of Induction Heating, Marcel Dekker, 2003.)
406 Handbook of Induction Heating

stresses during service, reducing the probability of bending fatigue fracture. However,
application of this pattern is quite limited. Since the tooth flank is not hardened, this pat­
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

tern provides poor wear resistance that may result in removal or displacement of particles
from the gear surface and tooth deformation. Theoretically, it is possible to imagine the
necessity of using this pattern as well as the previous one; however, it is more practical to
use another pattern, such as Pattern h.
Pattern e is one of the most typical patterns when hardening gears with small teeth as
well as sprockets and splines. Because the tooth is through hardened, it might exhibit
some brittleness unless fine-grain and ultrafine-grain martensite is formed. There is a
danger of having a brittle fracture in gears with through-hardened teeth, particularly
those subjected to shock/impact loads. The core of the tooth that exhibits sufficient tough­
ness, ductility, and strength should be able to withstand impact loads and prevent plastic
deformation of the gear teeth. Low-temperature tempering helps provide a combination of
those properties. Low-temperature tempering lowers the final hardness down to 54 to 58
HRC (depending on application specifics).
Patterns f and g are popular patterns for medium-size gears. According to Pattern f,
a case depth in the tooth root area is typically 30% to 40% of the depth in the tooth tip.
Slightly larger hardness depth at the tooth pitch line compared to the root may be benefi­
cial in some cases as a preventive action against spalling and pitting. It is very important
to harden the entire gear perimeter, including the flank and root area. An uninterrupted
hardened pattern of all contact areas of the tooth indicates good wear properties of the
gear and it ensures the existence of an uninterrupted distribution of desirable compressive
stresses at the tooth surface. Because gear teeth are not hardened through, a relatively duc­
tile tooth core (30 to 44 HRC) and a hard surface (56 to 62 HRC) provide a good combina­
tion of such important gear properties as exceptional wear resistance, toughness, bending
strength, and the needed gear durability (assuming an appropriate gear design).
Pattern h is the most popular choice for induction hardening of large gears and pinions
(i.e., gears with 300 mm and even larger O.D.) with coarse teeth (modules greater than
eight). This pattern provides an exceptional combination of fatigue and wear strength as
well as resistance to shock loading and scuffing (severe adhesion wear where metallic
particles transfer from one tooth to another), which is very important for heavily loaded
gears and pinions experiencing loads of appreciable magnitude. It is recommended that
for these applications, surface hardness should not be too high, typically in the range of
55 to 59 HRC. If surface hardness exceeds 61 to 62 HRC, the gear might be too brittle and
could experience some tooth bending failures.

4.9.1.3 Coil Design and Heating Mode


The variety of required hardness profiles calls for different coil designs and process pro­
tocols. Depending on the size of the gear, the required hardness pattern, and tooth geom­
etry, gears are induction hardened by encircling the whole gear (external or internal) with
an induction coil (the so-called spin hardening, Figure 4.177) or, for larger gears, hard­
ening them tooth by tooth with either gap-by-gap or tip-by-tip heat-treating techniques
(Figures 4.58c) [1,2,4,7–17,196,400,434–460].
Spin hardening provides higher production rates but requires a considerably greater
amount of power and capital equipment investment because of the necessity of heating
the entire perimeter of the gear at once (particularly when hardening large gears). In con­
trast, the power demand of tooth-by-tooth inductors is substantially lower; however, the
production rate is also noticeably reduced.
Heat Treatment by Induction 407
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

FIGURE 4.177
Examples of spin hardening of gears. (Courtesy of Inductoheat Inc., an Inductotherm Group company.)

4.9.1.3.1 Tooth-by-Tooth Hardening of Gears


As the name implies, each tooth is heated individually. The tooth-by-tooth method com­
prises two noticeably different techniques: tip-by-tip or gap-by-gap hardening.
The tip-by-tip method can apply a static heating mode or scanning mode (though static
is more typical). A gap-by-gap technique exclusively applies a scan hardening mode. Both
techniques (tip-by-tip and gap-by-gap hardening) are not suitable for small and fine-pitch
gears (modules smaller than six).
In tip-by-tip hardening, an inductor encircles the body of a single tooth. Inductor geom­
etry depends on the shape of the teeth and the required hardness pattern. In the case of
static hardening of large sprocket teeth, a solenoid coil or a split-return inductor can be
used. In other cases, a hairpin inductor can be applied to scan harden the tooth working
surfaces. The use of this method produces hardening patterns a, b, and c (Figure 4.176).
One of the typical concerns associated with tip-by-tip hardening is the problem of unde­
sirable heating and softening (tempering back) of the areas adjacent to the hardened area,
because of the external magnetic field of the inductor. This effect can be minimized to
some extent using flux concentrators as shields. However, in cases when the allocation of
concentrators can be difficult because of space limitations, the undesirable heating of the
adjacent teeth can be reduced by applying thin copper shields (copper caps).
Presently, tip-by-tip hardening is very rarely used because the hardening patterns
usually do not provide the needed fatigue and impact strength. This is the reason why
the term tooth-by-tooth hardening is often exclusively associated with the gap-by-gap
hardening method and below we will use both tooth-by-tooth and gap-by-gap hardening
interchangeably.
The tooth-by-tooth hardening concept can be applied to external and internal gears and
pinions, and it requires the inductor to be symmetrically located between two flanks of
adjacent teeth (Figure 4.178). It can be designed to heat only the root and flange of the tooth,
leaving the tip and tooth core tough and ductile.
This is one of the oldest induction hardening techniques; however, recent innovations
continue to improve the quality of gears that are heat treated using this method. Induction-
hardened gears can be fairly large, with outside diameters easily exceeding 3 m and can
weigh several tons. Gears used in wind turbines are typical examples where tooth-by-
tooth induction hardening is effectively used [435]. There is a limitation to applying this
method for hardening internal gears. Typically, it is required that the internal diameter of
an internal gear exceed 200 mm (8 in.) and, in some cases, 250 mm (10 in.) or more.
Since most wind turbines are constructed on remote sites, the size and weight of tur­
bines in combination with the expenses associated with their repair demand superior
strength and higher quality of wind energy generator components. Therefore, the qual­
ity of case-hardened large gears directly affects the longevity of wind turbines and their
408 Handbook of Induction Heating
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

(a) (b) (c)

FIGURE 4.178
Inductors for tooth-by-tooth (also referred to as gap-by-gap) surface hardening of large gears. (a) Double-
butterfly style inductor; (b) single-butterfly style inductor; (c) close-up view of a single-butterfly style inductor
and spray quench. (Courtesy of Inductoheat Inc., an Inductotherm Group company.)

competitiveness. This emphasizes the importance of taking steps to ensure the quality and
repeatability of hardness patterns.
As stated above, the gap-by-gap technique applies the scanning mode, requiring a high
level of skill, knowledge, and experience. Scanning rates can reach 10 mm/s and even
higher (though 6 to 8 mm/s is more typical).
Gap-by-gap hardening is a time-consuming process with low production rates. Power
requirements for these techniques are usually quite low. This can be considered a signifi­
cant advantage, because if spin hardening is used, a large gear would require an enormous
amount of power, which could diminish the cost-effectiveness of the heat treating or make
it not even feasible.
There is a variety of tooth-by-tooth inductor designs to accommodate the vast variety of
gear types, tooth profiles, and sizes. Some of the most popular inductor designs are shown
in Figure 4.178. Originally, the tooth-by-tooth inductor was developed in the 1950s by the
British firm Delapena.
As one can see in Figure 4.179, the path of the induced eddy current has a butterfly-
shaped loop. The maximum current density is located in the root area (the center part of
the butterfly). This is the reason why this inductor style is also called a butterfly inductor.
In order to further increase the power density induced in the root, a magnetic flux
concentrator is applied. A stack of laminations is typically used as flux concentrators.
Laminations are oriented across the gap. In some cases, it is advantageous to use a so-
called double-butterfly inductor (Figure 4.178a). Figure 4.180 shows the hardness pattern
profile that is typical for gap-by-gap hardening.

FIGURE 4.179
Path of the induced eddy current resembles a butterfly-shaped loop. (Courtesy of Inductoheat Inc., an
Inductotherm Group company.)
Heat Treatment by Induction 409
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

FIGURE 4.180
Typical hardness pattern for gap-by-gap induction hardening. (Courtesy of Inductoheat Inc., an Inductotherm
Group company.)

Although the eddy current path has a butterfly shape, when applied with a scanning
mode, the temperature is distributed within the regular area of the gear roots and
flanks quite uniformly. At the same time, since the eddy current makes a return path
through the flank and, particularly through the tooth tip, proper care should be taken
to prevent overheating the tooth tip. Overheating of the tip can substantially weaken
the tooth.
Applied frequencies are usually in the range of 3 to 30 kHz. At the same time, there are
cases when a frequency of 70 kHz or higher has been used. For example, the NATCO sub­
merged technique [459] applied an RF of 450 kHz.
Pattern uniformity is quite sensitive to coil positioning. Asymmetrical positioning of
an inductor results in a nonuniform hardness pattern. For example, an increase in the air
gap between the coil copper and the flank surface on one side will result in a reduction of
hardness and shallower case depth there, altering the mechanical properties.
Because of relatively small inductor-to-tooth air gaps (with 0.8 to 2 mm being typical) and
harsh working conditions, these coils require intensive maintenance and have a relatively
short life compared to inductors that encircle the gear.
Too small an air gap can result in local overheating or even melting of the gear surface.
Some arcing can occur between the inductor and the gear surface if the air gap is too small
[336,398].
Precise inductor fabrication techniques, inductor rigidity, and superior alignment
techniques are essential. Special locators or electronic tracking systems are often used
to ensure proper inductor positioning in the tooth space. Thermal expansion of metal
during heating should also be taken into consideration when determining the proper
inductor-to-tooth gap. After loading and initial coil positioning, the process runs auto­
matically based on the application process protocol.
When developing tooth-by-tooth gear hardening, particular attention should be paid
to electromagnetic end/edge effects and the ability to provide the required pattern in the
gear end areas, as well as along the tooth perimeter. To obtain the required temperature
uniformity, it is necessary to use a complex control algorithm: Power versus Scan Rate versus
Inductor Position.
A short dwell at the initial and final stages of inductor travel is often used. Thanks to
preheating attributed to thermal conduction, the dwell at the end of coil travel is usually
shorter compared to the dwell at the beginning of travel or is not applied at all.
With the scanning mode for hardening gears with wide teeth, two techniques can be
used: a design concept where the inductor is stationary and the gear is moveable, and a
concept that assumes the gear is stationary and the inductor is moveable.
There can be substantial shape/size distortion when applying tooth-by-tooth hardening.
Shape distortion is particularly noticeable in the last heating position where the last tooth
can be pushed out 0.1 to 0.8 mm. Hardening every second tooth can minimize distortion.
410 Handbook of Induction Heating

Obviously, this will require two revolutions to harden the entire gear. Therefore, final
grinding may be required. There is a linear relationship between the volume of required
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

metal removal and the grinding time. Thus, excessive distortion leads to a prolonged
grinding operation and increases the cost. Repeatable distortion may be compensated for
during gear fabrication.
Even though there might be appreciable distortion when hardening large gears and pin­
ions (e.g., mill, marine and large transportation gears, etc.), its magnitude is not typically as
significant when compared to carburized gears. Carburizing requires soaking of gears for
many hours (in some cases, up to 30 h or longer) at temperatures of 850°C (1562°F) to 950°C
(1742°F). At these temperatures, the large masses of metal expand to a much greater extent
compared to the case when only the surface layer is heated by electromagnetic induction.
The expansion of a large mass during prolonged heating during carburizing and the con­
traction during cooling/quenching “move” the metal to a much greater degree, resulting
in larger gear distortion.
Besides that, large gears being held at temperatures of 850°C (1562°F) to 950°C (1742°F)
for many hours have little rigidity; therefore, they can sag and have a tendency to follow
the movement of their supporting structures during soaking and handling. With induc­
tion hardening, areas unaffected by heat as well as areas with temperatures correspond­
ing to the elastic deformation range serve as shape stabilizers and lead to not only lower
but more predictable distortion.
One typical concern when applying tooth-by-tooth hardening is the challenge related to
undesirable heating and softening of the areas adjacent to the hardened region (tempering
back). The main cause of undesirable softening/temper back is associated with the ther­
mal conductivity phenomenon. Heat is transferred by thermal conduction from a high-
temperature region of the workpiece toward a lower-temperature region and is a function
of the temperature difference, distance, and the value of thermal conductivity. Most met­
als are relatively good thermal conductors. During hardening, the surface temperature
exceeds the critical temperature Ac3. Therefore, when heating one side of the tooth, there
is a danger that the opposite side of the tooth will be heated by thermal conduction to an
inappropriately high temperature, resulting in undesirable softening of previously hard­
ened areas (Figure 4.181, first two images).
Whether a hardened side of the tooth will be unacceptably softened depends on several
factors, including the applied frequency, gear module, tooth shape, heat time, case depth,
and so on. In the case of shallow case depths and thick teeth, the root of the tooth, its fillet,
and the dedendum area are typically not unduly softened because of thermal conduction.
The massive area below the tooth root serves as a heat sink, which helps protect the hard­
ened side of the tooth from tempering back.

Coil
Previously
hardened
Cooling
block

FIGURE 4.181
Undesirable softening of tooth tip area of previously hardened areas (first two images). Additional cooling pro­
tects already hardened areas while austenizing unhardened areas of the gear (last two images).
Heat Treatment by Induction 411
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

(a) (b)

FIGURE 4.182
Two tooth-by-tooth induction machines built by Inductoheat Inc. (a) for hardening large gears and by
Inductoheat-Europe (b) for heat treating large bearing rings for a wind energy turbine with teeth located
on inside diameter of the ring. (Courtesy of Inductoheat Inc., an Inductotherm Group company.)

Conversely, the tooth tip region can be considered as an area of concern because there
is a relatively small mass of metal. In addition, heat has a short distance to travel from
one (heating) side to the other (already hardened) side of the tooth (Figure 4.181, first two
images).
To overcome the problem of tempering back, special cooling spray blocks are applied.
Additional cooling protects already hardened areas while heating unhardened areas of
the tooth (Figure 4.181, last two images). Even though external cooling is applied, there still
may be some unavoidable softening depending on the tooth shape and process param­
eters. This tempering back is typically insignificant, well controlled, and acceptable.
Tooth-by-tooth hardening can be applied for gears submerged in a temperature-
controlled quench tank. This technique was applied in the original Delapena induction
hardening process. In this case, quenching is practically instantaneous. However, notice­
ably higher power is needed to compensate for the cooling effect of the quenchant during
heating. The fact that a gear is submerged in quenchant also helps prevent the tempering
back problem. In addition, the quenchant serves as a coolant to the inductor. Therefore, in
submerged hardening, an inductor might not have to be water cooled. On the other hand,
there may be some challenges to set up this system because of the additional consideration
of the quenchant requirements (e.g., quenchant obstruction issues).
Figure 4.182 shows two tooth-by-tooth induction gear hardening machines built by
Inductoheat Inc. (a) for hardening large gears and by Inductoheat-Europe (b) for heat treat­
ing large bearing rings for a wind energy turbine with teeth located on the inside diameter
of the ring. A bearing ring O.D. can be as large as 140 in. (3556 mm), and the maximum
weight exceeds 11,000 lb (5000 kg). The required case depth is 2.5 to 3.5 mm. The z-axis
scan height (tooth width) is 13.75 in. (350 mm).

4.9.1.3.2 Gear Spin Hardening (Encircling Inductors)


Spin hardening of gears utilizes a single or multiturn inductor that encircles the gear
(Figure 4.177). It is typically used for small- and medium-sized gears. Unfortunately, spin
hardening sometimes cannot be easily used for certain gears with complex geometries
412 Handbook of Induction Heating

because of the difficulty in obtaining a contour-like (uniform) austenized surface layer


before quenching. Besides, in case of appreciable size gears, it might also require an exces­
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

sively large amount of power owing to the necessity of short heat times suppressing ther­
mal conduction in order to obtain the desired hardness pattern uniformity. Still, gear spin
hardening is the most popular technology for hardening a variety of gears of small and
moderate size (Figure 4.175). It has been reported [457,460] that greater surface compressive
stresses within the tooth root were achieved with spin hardening than with tooth-by-tooth
hardening.
Gears are rotated during heating to ensure even distribution of energy compensating for
the field fringing (fish-tail) effect (see Section 4.2.8.3).
When applying encircling coils, there are five parameters that play a dominant role in
obtaining the required hardness pattern: frequency, power density, heat time, quenching
conditions, and coil geometry. Certain combinations of these parameters can result in dif­
ferent hardness patterns. Figure 2.12b shows a diversity of hardening patterns that were
obtained on the same carbon steel shaft with teeth using different combinations of heat
time, frequency, and power.
Figure 4.183 illustrates the eddy current flow in two extreme cases: application of high
frequency (a) and low frequency (b). When high frequency is applied, δ is relatively small
and the induced eddy currents generally follow the contour of the gear teeth (Figure
4.183a). This leads to a heat source surplus in the tip of the tooth compared to that in
the root. Besides that, the tip of the tooth has a substantially smaller mass of metal to be
heated, compared with the dedendum and root, where a much larger thermal heat sink
is located. These two major factors result in greater heat intensity at the tip, with a corre­
sponding temperature rise and deeper case depth upon quenching.
Application of low frequency (Figure 4.183b) is associated with a dramatic increase in δ,
potentially leading to eddy current cancellation at the tooth tip (top land) and possibly at
the addendum area. This makes it much easier for induced current to take a shorter path,
following the root circle of the gear instead of following along the tooth profile, leading to
more intensive heat generation in the tooth root area compared with its tip.
FEA computer modeling illustrates the effect of frequency on temperature distribution
and hardness pattern while applying different frequencies. Figures 4.184 through 4.187

High frequency Low frequency

Gear

Coil

Eddy
current

Gear
Coil
(a) (b)

FIGURE 4.183
Illustration of an eddy current flow in two extreme cases: application of high frequency (a) and low frequency (b).
(From V. Rudnev et al., Gear heat treating by induction, Gear Technology, March, 2000.)
Heat Treatment by Induction 413

show the dynamics of temperature distribution during heating and quenching of a fine-
pitch gear using various frequencies: RF (300 kHz), moderate frequency (30 kHz), and
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

lower frequency (10 kHz) [461]. Advantage was taken with respect to periodicity and sym­
metry of the gear teeth, allowing the modeling of only ½ of the tooth (right side).
As expected, when an RF of 300 kHz is applied, the eddy current induced in the tooth
follows its contour (Figure 4.184). Because the highest concentration of current density will
be in the tip of the tooth, there will be a heat source surplus in the tip compared to the root.
Also, taking into account that the tip of the tooth has a smaller mass of metal to be heated
compared with the root, the tip will experience the most intensive temperature rise during
the heating cycle. In addition, from a thermal perspective, the amount of metal beneath
the root circle represents a much larger cold sink compared with the addendum area and
particularly in the tooth tip region. Another factor that contributes to more intensive heat­
ing of the tip is a better electromagnetic proximity effect between the inductor and tooth
tip versus its root. Higher frequency has the tendency to make the proximity effect more

Heating—300 kHz

Time = 0.3 s Legend Time = 0.6 s Legend Time = 1.2 s Legend


Y Y Y
0.027 1100 0.027 1100 0.027
1100
1 990 1 990 990
0.026 0.026 0.026 1
880 880 880
770 770 770
0.025 0.025 0.025
660 660 660
0.024 550 0.024 550 0.024 550
2 440 2 440 2 440
0.023 0.023 0.023
330 330 330
220 220 220
0.022 0.022 0.022
110 110 110
0.021 0 0.021 0 0.021 0
0 0.002 X 0 0.002 X 0 0.002 X
N1 = 768.2 N2 = 407.4 N1 = 804.5 N2 = 616.3 N1 = 889.0 N2 = 738.8

Time = 1.0 s Legend Time = 2.0 s Legend Legend


Y Y
0.027 1100 0.027 1100 65
1 990 1 990 60
0.026 0.026
880 880 55
0.025 770 0.025 770
50
660 660
46
0.024 550 0.024 550
41
2 440 2 440
0.023 0.023 36
330 330
220 220 31
0.022 0.022
110 110 26

0.021 0 0.021 0 21
0 0.002 X 0 0.002 X
16
N1 = 30.6 N2 = 86.0 N1 = 24.4 N2 = 49.3

Quenching Hardness

FIGURE 4.184
Temperature distribution during heating and quenching of a fine-pitch gear using RF (300 kHz); heat time is
1.2 s. (From V. Rudnev, Spin hardening of gears revisited, Professor Induction Series, Heat Treating Progress, ASM
Int’l, Materials Park, OH, March/April, 2004, pp. 17–20.)
414 Handbook of Induction Heating

pronounced. These factors provide rapid austenitization of the tooth tip (Figure 4.184, top
images). Temperatures of two critical nodes (N1, tip; N2, root) are in degrees Celsius and
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

X–Y dimensions are in meters.


Spray quenching results in a martensitic layer at the tooth tip only (Figure 4.184, bottom
right). Note that the dedendum and the root area have not been hardened, because insuf­
ficient heat was generated for austenitization.
The next simulation was conducted using a moderate frequency, which is 10-fold lower
(using 30 kHz, Figure 4.185). At the end of the heat cycle, the whole tooth was heated
quite uniformly, achieving the austenite phase temperature (Figure 4.185, bottom right)
and resulting in a through-hardened tooth upon quenching.
The study was continued by further reduction of frequency, using 10 kHz, represent­
ing a 30-fold reduction compared to the case study shown on Figure 4.184. With 10 kHz,
the eddy current flow and temperature distribution in the gear tooth are quite different
(Figure 4.186). Keep in mind that the heat time in all three cases discussed so far was kept
the same (1.2 s).

Heating—30 kHz

Time = 0.3 s Legend Time = 0.6 s Legend Time = 1.2 s Legend


Y Y Y
0.027 1250 0.027 1250 0.027 1250
1 1127 1127 1 1127
0.026 1 0.026
1004 0.026 1004 1004
881 881 881
0.025 0.025 0.025
758 758 758
0.024 635 0.024 635 0.024 635

2 512 2 512 2 512


0.023 0.023 0.023
369 369 369
266 266 266
0.022 0.022 0.022
143 143 143
0.021 20 0.021 20 0.021 20
0 0.002 X 0 0.002 X 0 0.002 X
N1 = 739.4 N2 = 724.6 N1 = 782.6 N2 = 932.2 N1 = 905.5 N2 = 1011.6

Time = 0.2 s Legend Time = 1.0 s Legend Legend


Y Y
0.027 1250 0.027 1250 65
1 1127 1 1127 60
0.026 0.026
1004 1004 55
0.025 881 0.025 881 50
758 758
46
0.024 635 0.024 635
41
2 512 2 512
0.023 0.023 36
369 369
266 266 31
0.022 0.022
143 143 26
0.021 20 0.021 20 21
0 0.002 X 0 0.002 X
16
N1 = 236.8 N2 = 467.8 N1 = 48.9 N2 = 160.1

Quenching Hardness

FIGURE 4.185
Temperature distribution during heating and quenching of a fine-pitch gear using moderate frequency
(30 kHz); heat time is 1.2 s. (From V. Rudnev, Spin hardening of gears revisited, Professor Induction Series,
Heat Treating Progress, ASM Int’l, Materials Park, OH, March/April, 2004, pp. 17–20.)
Heat Treatment by Induction 415

Heating—10 kHz
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

Time = 0.3 s Legend Time = 0.6 s Legend Time = 1.2 s Legend


Y Y Y
0.027 1250 0.027 1250 0.027 1250
1 1127 1 1127 1 1127
0.026 0.026 0.026
1004 1004 1004
0.025 881 881 881
0.025 0.025
758 758 758
0.024 635 0.024 635 0.024 635
2 512 2 512 2 512
0.023 369 0.023 0.023
369 369
266 266 266
0.022 0.022 0.022
143 143 143
0.021 20 0.021 20 0.021 20
0 0.002 X 0 0.002 X 0 0.002 X
N1 = 536.2 N2 = 515.9 N1 = 657.7 N2 = 790.5 N1 = 733.4 N2 = 919.3

Time = 0.5 s Time = 1.0 s Legend


Y Legend Y Legend
0.027 0.027 65
1250 1250
60
1 1127 1 1127
0.026 0.026 55
1004 1004
881 881 50
0.025 0.025
758 758 46
0.024 635 0.024 635 41
512 512 36
2 2
0.023 0.023 31
369 369
266 266 26
0.022 0.022
143 143 21
0.021 20 0.021 20 16
0 0.002 X 0 0.002 X
N1 = 84.3 N2 = 262.7 N1 = 45.5 N2 = 144.1

Quenching Hardness

FIGURE 4.186
Temperature distribution during heating and quenching of a fine-pitch gear using lower frequency (10 kHz);
heat time is 1.2 s. (From V. Rudnev, Spin hardening of gears revisited, Professor Induction Series, Heat Treating
Progress, ASM Int’l, Materials Park, OH, March/April, 2004, pp. 17–20.)

A frequency reduction from 300 to 10 kHz noticeably increases the eddy current penetra­
tion depth in the hot steel—from 1 to 5.4 mm. In a fine-toothed gear, such an increase in cur­
rent penetration depth results in current cancellation in the tooth tip and pitch circle area
heated above A2 critical temperature. This makes it easier for eddy current to take a short
path, following the root circle of the gear instead of following the tooth contour. The result
is more intensive heating of the tooth root area compared with its tip and the development
of martensite upon quenching. In contrast to using 300 kHz, with 10 kHz upon reaching the
Curie point, the temperature rise in the tip and entire addendum region was stopped; thus,
proper austenization does not occur and hardening did not take place. Notice the effect of a
slight increase in heat time from 1.2 to 1.5 s when using 10 kHz (compare Figure 4.186 with
Figure 4.187), leading to hardening not only the root but also the dedendum region.
As a rule, when it is necessary to harden only the tooth tips, a higher frequency and
high to moderate power densities are applied (Figure 4.183a). When hardening the tooth
root, a lower frequency in combination with short time and high power density is used
416 Handbook of Induction Heating

Heating—10 kHz
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

Time = 0.3 s Legend Time = 1.2 s Legend Time = 1.5 s Legend


Y Y Y
0.027 1250 0.027 1250 0.027 1250
1 1127 1 1127
0.026 1 1127
1004 0.026 0.026
1004 1004
881 881 881
0.025 0.025 0.025
758 758 758
0.024 635 0.024 635 0.024 635
2 512 2 512 2 512
0.023 0.023 0.023
369 369 369
266 266 266
0.022 0.022 0.022
143 143 143
0.021 20 0.021 20 0.021 20
0 0.002 X 0 0.002 X 0 0.002 X
N1 = 536.2 N2 = 515.9 N1 = 733.4 N2 = 919.3 N1 = 754.2 N2 = 942.2

Time = 0.2 s Time = 1.0 s


Y Legend Y Legend Legend
0.027 1250 0.027 1250
65
1127 1127
0.026 1 0.026 1 60
1004 1004
55
0.025 881 0.025 881
50
758 758
0.024 635 0.024 635 46

512 512 41
2 2
0.023 0.023 36
369 369
266 266 31
0.022 0.022
143 143 26
0.021 20 0.021 20 21
0 0.002 X 0 0.002 X
16
N1 = 211.3 N2 = 458.1 N1 = 49.5 N2 = 165.3

Quenching Hardness

FIGURE 4.187
Temperature distribution during heating and quenching of a fine-pitch gear using lower frequency (10 kHz);
heat time is 1.5 s. (From V. Rudnev, Spin hardening of gears revisited, Professor Induction Series, Heat Treating
Progress, ASM Int’l, Materials Park, OH, March/April, 2004, pp. 17–20.)

(Figures 4.177a and 4.183b). Low power density and extended heat time produces a deep
pattern with a wide transition zone.
It is imperative to keep in mind that depending on gear geometry, besides frequency, the
variation of applied power density can shift the heat intensity within the gear tooth. For
example, the application of sufficiently high frequency in combination with a relatively low
power density could result in tip hardening or hardening only the tip and addendum area.
However, if the power density is substantially increased (keeping the same frequency), it
could result in more intense heating of the root area instead of the tooth tip. Higher power
density reduces μr by saturating the steel and leading to a considerable increase of δ, thus
modifying the eddy current flow within the gear tooth.
Figure 4.188 shows an example of applying excessive power density and the use of lower-
than-optimal frequency when heating a gear in an encircling coil. The teeth had some
undercuts and were approximately 8 mm tall. Regardless of the fact that the tips of the
teeth were located much closer to the current-carrying face of the coil copper compared to
Heat Treatment by Induction 417
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

FIGURE 4.188
Example of applying an excessive power density and the use of very low frequency when hardening gears using
encircling coil.

the roots, the great majority of eddy currents have completed their loop following the root
circle, resulting in severe overheating there.
As expected, the terms high frequency and low frequency have relative meanings. For
example, depending on the tooth geometry, a frequency of 3 kHz may act as a high fre­
quency for coarse teeth and 300 kHz may act as a low frequency for splines, threads, fine
teeth, or skinny teeth, for example.
In addition to the factors discussed above, the hardness pattern and its repeatability
depend strongly on the relative positioning of the gear with respect to the induction coil
and the ability to maintain gear concentricity.
There are several ways to accomplish quenching for gear spin hardening. One technique
is to submerge the gear in a quench tank upon completion of austenization (Figure 4.120).
This technique is applicable for moderate and relatively large-sized gears. Small- and
medium-sized gears are usually quenched in place, using an integrated quench. The third
technique requires the use of a separate concentric quench block (quench ring) located
below the inductor or in its proximity.
Quite often, in order to optimize gear performance and reduce distortion, it is highly
desirable to obtain a contour-like hardening pattern (Figure 4.176f and g). This also often
maximizes the beneficial compressive stresses within the case depth. In some instances,
obtaining a true contour hardened pattern can be a challenging task because of the above-
discussed differences in current density distribution and heat transfer conditions within
the gear tooth.
The five most popular induction-heating concepts that employ encircling-type coils are
as follows [1]:

1. Conventional single-frequency concept


2. Pulsing single-frequency concept
3. Pulsing dual-frequency concept
4. IFP technology
5. Single-coil, dual-frequency concept (also referred to as simultaneous dual-frequency
concept)

Conventional single-frequency concept (CSFC). CSFC [1] is used for hardening gears with
medium and small teeth and splines. In these applications, it is often acceptable to
through harden teeth (Figure 4.176, Patterns b and e), and CSFC can be a cost-effective
approach to do so. Quite frequently, CSFC can also be successfully used for medium-
sized gears without through hardening the entire tooth, but having deeper case depth
at the tip and shallower case depth at the root with a pattern somewhat similar to that
shown in Figure 4.176f.
418 Handbook of Induction Heating

Although CSFC is the most cost-effective approach for small- and medium-sized
gears, there are cases where this concept has been successfully used for large gears
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

as well. For example, Figure 4.120b shows an induction machine for hardening large
sprockets. A multiturn encircling coil is used for hardening sprockets with a major
diameter of 27.6 in. (701 mm), a root diameter of 24.3 in. (617 mm), and a thickness of
3.125 in. (79 mm). In this particular case, it was in the customer’s best interest to harden
and temper in the same coil using the same power supply. In other cases, it might not
be the best solution.
Figure 4.189a shows another example of an induction gear-hardening machine that
applies this concept. The gear-like part being heat treated in this application is an auto­
motive transmission component with helical fine teeth on the inside diameter (I.D.) and
coarse teeth on the outside diameter (O.D.) for a parking brake. Both the inside and outside
diameters require hardening (Figure 4.189b). Because of the difference in size of teeth, the
hardening of the fine teeth requires a noticeably higher frequency than the outside diam­
eter. Therefore, a frequency of 200 kHz was chosen for I.D. heating and 10 kHz was chosen
for the O.D. Regardless of using an RF of 200 kHz, the fine teeth are through hardened,
but the coarse teeth reveal some contouring of the hardness pattern regardless of using a
much lower frequency of 10 kHz (Figure 4.189c).
At the end of the austenization, the quenchant is applied in place; that is, no reposition­
ing is required. This practically instantaneous quenching provides a consistent metallur­
gical response. During quenching, there is minimal rotation to ensure that the quenchant
penetrates all critical areas of the teeth. Quenching reduces the gear temperature to the
level suitable for handling. A cam-operated robot is used to convey the gears. Parts are
monitored at each station and accepted or rejected based on all the major factors that
affect the hardening process. This includes energy input, quench flow rate, temperature,
quench pressure, heat time, and others. An advanced control/monitoring system verifies
all machine settings to provide confidence in the quality of processing for each individual
gear. Precise control of the hardening operations and appropriate coil design minimize the
distortion, forming the desirable residual stresses. The hardened gear is then inspected
and moved to the next operation.
As stated earlier, quite often, in order to optimize gear performance and reduce distor­
tion, it is desirable to make a hardness distribution as close to contour hardening as pos­
sible. Unfortunately, CSFC does not typically allow doing so.

(a) (b) (c)

FIGURE 4.189
Automotive transmission component with helical fine teeth on the inside diameter (I.D.) and coarse teeth on the
outside diameter (O.D.) for a parking brake being induction hardened (a). Both the inside and outside diameters
require hardening (b and c). (Courtesy of Inductoheat Inc., an Inductotherm Group company.)
Heat Treatment by Induction 419

Pulsing single-frequency concept (PSFC). In order to overcome the drawbacks of CSFC and
to improve the hardness distribution, PSFC was developed. In many cases, PSFC allows
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

the user to noticeably improve the shortcomings of CSFC and obtain a hardening pattern
that is more contoured than patterns obtained using the CFSC concept. Pulsing assists
in providing the desirable heat flow toward the root without overheating of the tooth tip.
A typical PSFC protocol consists of a moderate- or low-power preheat, soaking stage,
short high-power final heat, and quenching. This approach is also known as dual pulse
hardening. Preheating improves the heated depth at the root, enabling an enhancement of
the metallurgical result. Preheat times typically range from several seconds to a minute,
depending on the size and gear geometry.
After preheating, there may be a soak time ranging from 2 to 10 s to achieve more suit­
able thermal conditions. Final heat times can range from less than 1 s to a few seconds.
Depending on the application, the preheat can consist of several pulses. Obviously, pre­
heating reduces the amount of energy required in the final heat.
Pulsing dual-frequency concept (PDFC). The idea of using two different frequencies to
closely approach, producing the desired contour hardness pattern, has been around since
the late 1950s. This concept was primarily developed to obtain a contour hardening profile
for helical and straight spur gears. Obviously, since that time, the process has been refined
and several innovations have been developed.
According to PDFC, the gear is preheated using lower frequency to a temperature deter­
mined by the process features. This temperature is usually 350°C to 100°C below the
critical temperature Ac1. Preheat temperature depends on the gear specifics and available
power/frequency source. It should be mentioned that the higher the preheat temperature,
the lower the power required for the final heat. However, excessively high preheat tem­
peratures can result in increased distortion.
As with PSFC, PDFC can be accomplished using a single-shot mode or scanning mode.
The scanning mode is typically applied for shafts with teeth.
Preheating is usually accomplished by using a medium frequency (3 to 10 kHz).
Depending on the type of gear, its size, and material, higher frequency (30 to 450 kHz) in
combination with high power density is applied during the final heat, where the selected
frequency allows the current to penetrate only to the desired depth, helping to obtain a
contour-like hardness pattern.
Most commonly, a two-coil arrangement was used. One coil provides preheating, and
another coil provides final heating. Both coils work simultaneously if the scanning mode
is applied. In the case of a single-shot heating mode, a two-step index-type approach is
used.
In some cases, dual-frequency machines produce components with lower distortion hav­
ing a more favorable distribution of residual stresses compared to using single-frequency
techniques. However, depending on the tooth geometry and presence of undercuts, the
time delay between low-frequency preheating and high-frequency final heating could
have a detrimental negative effect for obtaining truly contour hardened patterns. Attempts
have been made to minimize the gear indexing time, keeping it within the 0.25–0.6 s range
[400]. However, further reduction is limited by inertia and the mechanical capability of
the transferring/indexing mechanisms. Even such a seemingly short time delay between
preheating and final heating may have an appreciably negative impact that discourages
the achievement of optimal heat treat properties.
Independent frequency and power control concept (Statitron IFP Technology). One of the
impressive technologies is referred to as Statitron IFP and has been reviewed in Section
4.2.2.1.6 (Figure 4.68). IFP Technology allows independent and instantaneous changing of
420 Handbook of Induction Heating

the power and frequency (within the 5- to 60-kHz range) (Figure 4.68, left) during induction
hardening. Therefore, this technology allows eliminating some shortcoming associated
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

with PSFC. The patented Statitron IFP inverter’s (Figure 4.68, right) ability to indepen­
dently and instantly (like a CNC machine) change both frequency and power during the
heating cycle allows heat treaters to use a lower frequency for preheating the root areas,
while a higher frequency helps ensure sufficient heating of tooth flanks and tips when
hardening moderate-sized gears. This technology eliminates the need for using expen­
sive high-speed transferring/indexing mechanisms and noticeably improves the hardness
pattern.
Single-coil, dual-frequency concept (also referred to as simultaneous dual-frequency concept).
Limitations of the PDFC concept have initiated the development of an alternative technol­
ogy called simultaneous dual-frequency gear hardening as a way to further improve the
quality of induction-hardened gears [400,436].
The core of simultaneous dual-frequency gear-hardening technology is associated with
the development of solid-state inverters capable of producing two substantially different
frequencies simultaneously. Figure 4.190 shows some examples of the waveforms gen­
erated by a simultaneous dual-frequency inverter. Waveforms representing coil voltage
(top) and coil current (bottom) comprise two appreciably different frequencies that can
be applied at the same time to a single inductor [436]. The lower-frequency output of the
power supply helps austenitize the roots of the teeth and the high frequency helps austen­
itize the flanks and tips. Thus, two appreciably different frequencies can be applied to a
hardening inductor, making it much easier to obtain a true contour hardening of the gear
teeth (some of true contour hardening examples are shown in Figure 4.175).

Heating coil voltage

1.530 m 1.550 m 1.570 m 1.590 m 1.610 m 1.630 m


Time (ms)

Heating coil current

1.530 m 1.550 m 1.570 m 1.590 m 1.610 m 1.630 m


Time (ms)

FIGURE 4.190
Examples of the waveforms generated by a simultaneous dual-frequency inverter. Waveforms representing coil
voltage (top) and coil current (bottom) comprise two appreciably different frequencies that can be applied at the
same time to a single inductor. (From V. Rudnev, Single-coil dual-frequency induction hardening of gears, Heat
Treating Progress, ASM Int’l, October, 2009, pp. 9–11.)
Heat Treatment by Induction 421

480 V

Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

Disconnect Cs Cs
High-frequency module: Ls Induction
600 kW, 150–400 kHz
current-fed inverter coil
Cs Cs

Medium-frequency module:
L10
600 kW, 10 kHz C10
voltage-fed inverter
(a) (b)

FIGURE 4.191
Inductoheat’s single-coil dual-frequency induction gear hardening system utilizing a total output power
of 1.4 MW (a) and a sketch of one possible circuitry of this technology (b). (Courtesy of Inductoheat Inc., an
Inductotherm Group company.)

FIGURE 4.192
Effect of different combinations of time–power–frequency of hardening patterns of spur gears. (Courtesy of
Inductoheat Inc., an Inductotherm Group company.)

Experience reveals that it is not always advantageous to have two different frequencies
working simultaneously. Many times, depending on the gear geometry, it is preferable
to apply lower frequency at the beginning of the heating cycle, and after achieving the
desirable root heating, the higher frequency can complement the initially applied lower
frequency, completing the job by working together.
Figure 4.191a shows Inductoheat’s single-coil, dual-frequency induction gear harden­
ing system utilizing a total output power of 1.4 MW. As expected, for different applica­
tions, the output power might not need to be so high. Figure 4.191b shows a sketch of
one possible circuitry of this technology. In this scheme, two single-frequency modules
generate substantially different frequencies. Specially designed filters prevent undesir­
able interactions between inverters. Those modules can work not just simultaneously
but in any sequence desirable to optimize the desired properties of the gear hardening
process [436].
The relatively high cost is the main shortcoming of power supplies that produce two
simultaneous frequencies. Nevertheless, single-coil simultaneous dual-frequency technol­
ogy has a number of obvious advantages over conventional single-frequency hardening.
Higher cost can be justified using this technology when it is appropriate. As an example,
Figure 4.192 illustrates the effect of different combinations of time, power, and frequency
of hardening patterns on the spur gears.

4.9.2 Hardening of Steering Racks


Steering racks (Figure 4.193a) are an essential part of automotive power steering sys­
tems. Size and geometry specifics of the toothed steering racks as well as their heat treat
422 Handbook of Induction Heating
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

(a)
(b)

FIGURE 4.193
Steering racks are an essential part of automotive power steering systems (a) being heat treated using medium
frequency (b). (Courtesy of Inductoheat Inc., an Inductotherm Group company.)

specification depend on the design of the particular steering system. Minimization of


shape distortion after hardening without compromising the metallurgical characteristics
of the components is particularly important, because any elongated products have a ten­
dency to grow, bow, and contract during the hardening cycle, particularly in cases when
only the toothed area of the rack is specified to be heat treated.
The necessity to minimize the weight of vehicles has initiated the use of smaller-
diameter solid/hollow steering racks in combination with improved hardness patterns
to provide the needed strength.
There are two most frequently used processes for hardening of toothed steering racks.
The first approach is scan hardening using horizontal or vertical (more typical) arrange­
ments with single or multiple spindles/axes. The second approach is static hardening.
Machines can be built with single or multiple stations for high production rates. In other
cases, steering racks consist of toothed and ball-screw sections needed to be surface
hardened.
Specially designed scan inductors (including but not limited to conventional single-
turn and multiturn coils, as well as so-called “D”-shaped or “Z”-shaped inductors, etc.)
are applied for scan hardening of steering racks. Certain inductor design configura­
tions may necessitate part rotation while others may process without rotation, requir­
ing a particular orientation of the inductor with respect to the toothed area of the rack
and its back region. Various sensing devices may be used as an attempt to measure
and minimize distortion during scan hardening, though they noticeably complicate
the system.
With single-shot processing, a modified split-return, butterfly, or modified hairpin
inductor can be used. In other applications, highly effective electrical resistance heaters
can apply medium-frequency or high-frequency systems (Figure 4.193b). A number tech­
nique has been developed over the years in industry to minimize the distortion of toothed
steering racks.

4.9.3 Induction Hardening Raceways for the Wind Energy Industry


Similar to gear hardening, there are two basic approaches to heating medium- and large-
sized raceways, namely, a single-shot (static) hardening process and a scanning process.
In single-shot or static induction heating, the ring is surrounded by a single-turn or
multiturn solenoid coil having a profiled copper heating face. For example, to harden
the surface of the inside diameter of a bearing race, an induction coil is positioned inside
Heat Treatment by Induction 423

the ring. However, if it is required to have a hardness layer on the outside diameter of the
ring, then the induction coil is placed around its outside diameter, encircling the raceway
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

being heated. The ring may be rotated (spun) during the heat treatment process to ensure
an even distribution of energy along its perimeter over the entire heating cycle or can be
heated using special means to control the field fringing effect (fish-tail effect) as discussed
earlier in Section 4.2.8.3.
Upon completion of the heating stage, the ring is lowered into a separate concentric
spray quench device positioned below the inductor and spray quenched in place, or it is
submerged in a quench tank and quenching takes place inside the tank while the quen­
chant is usually agitated.
In induction scan hardening, a short inductor moves (scans) along a circular path con­
centric with the ring (along the ring’s periphery). A spray quench block (quench jet) follows
the inductor or is incorporated into the inductor design. This method requires signifi­
cantly less power than the single-shot or static hardening since only a small portion of the
ring is consequently heated. As an example, Figure 4.194 shows an Inductoscan base com­
pact machine for bearing ring scan hardening. The transformer is mounted on a spring-
loaded x–y slide. Rings are loaded with a crane into the load position, and the component
is moved into the hardening station where it is moved against the coil (which is equipped
with guides). Thus, a precise coupling distance is ensured for the entire hardening cycle.
The part then rotates for scan hardening via a circular axis. The heated portion of the ring
is spray quenched during scan hardening. The scan rate is 0.5 in./s (12 mm/s). The out­
side diameter of the bearing ring is up to 55 in. (1400 mm), and the typical case depth is
approximately 0.100 to 0.120 in. (2.5 to 3 mm).
Scan hardening of raceways using a single inductor is simple and the most economical
approach, requiring minimum capital cost and using the simplest control and machine
design. However, a narrow straight or angled soft band (transition zone) is inevitably cre­
ated with this technique because of the tempered back effect of the region adjoining the
final ring section to be heated. Special techniques have been developed to minimize the
length of the transition zone.

FIGURE 4.194
Inductoscan base compact machine for bearing-ring scan hardening. (Courtesy of Inductoheat-Europe, an
Inductotherm Group company.)
424 Handbook of Induction Heating

In some cases, an angled band is preferable instead of a straight transition band. To


prevent the occurrence of the tempered back soft zone without the necessity of using
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

an unduly large, expensive power source, an alternative process has been developed.
In this method (Figure 4.195a [14]), instead of using a single inductor, a pair of heating
inductors is used (Nos. 3 and 4). At the start of the heating process, identical power is
supplied to both inductors while each inductor travels in an arc of a little less than 180°
in opposite directions at a uniform speed. Trailing each inductor is a corresponding
quench jet incorporated into each inductor, or separate quenching blocks (Nos. 5 and
6), until the inductors meet again at the end of the heating process in the final heat­
ing position, where they are de-energized and, simultaneously, an auxiliary quench
is automatically applied to quench the final heated portion of the ring. The adjacent
inductors complement each other’s magnetic field in the final heating position, elimi­
nating any soft zones caused by inevitable temper back when a single scan inductor
is used.
The description of the process concept that allows eliminating soft spots using two scan­
ning inductors (Figure 4.195a) was adapted from the Russian literature [14] published in
the late 1970s. In fact, this was not the first description of scan hardening that eliminates
soft spots using two independently moving heaters. Other descriptions can be traced back
to the early 1940s [462–465]. For example, Figure 4.195b was adapted from a patent issued
in the early 1960s [462].
In its continuing tradition to further improve on existing technologies, Inductoheat
has developed novel patented technology that further perfects the known general
process of eliminating soft spots when hardening large ring-shaped parts using two
independently scanning inductors [466]. The technology uses numerous techniques
specifically developed to meet the need for induction case hardening of windmill com­
ponents. Some innovations are related to the unique ability of Inductoheat’s inverters
to independently control power and frequency during the scanning operation, which
helps optimize the thermal conditions at the initial and final scanning stages. Other
features are related to process subtleties and recipes, as well as effective handling of
large-sized parts.

38
3 36
4 42
44
5 46 44 48

47 68
110 120
104
6 70 108 86 84
46
48 88
98
72
76 74
66 94
106 80
78 82 100 96
102 92
90
18
(a) (b)

FIGURE 4.195
Seamless hardening (without a transition band) of circular components (a) (adapted from A.D. Demicheiv,
Surface Induction Hardening, St. Petersburg, Russia, 1990) and image adapted from a patent issued in the early
1960s (b) (D.A. Hay, Heat treating device, U.S. Patent #3,036,824, May 29, 1962).
Heat Treatment by Induction 425

4.9.4 Induction Hardening of Crankshafts


Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

4.9.4.1 Introduction
Induction heat treatment plays an important role in the manufacturing of quality crank­
shafts. Crankshafts are widely used in internal combustion engines, pumps, compres­
sors, and the like. They belong to a group of critical auto components, typically weighing
between 13 kg (30 lb) and 40 kg (85 lb) depending on the engine (Figure 4.196). However,
the weight of some crankshafts can exceed 1 ton (e.g., crankshafts for the ship industry or
stationary engines of power generators). The materials most frequently used for crank­
shafts are steel forgings, nodular iron castings, microalloy forgings, and austempered duc­
tile iron castings.
Structurally, crankshafts are composed of a series of crankpins (also called pin journals
or pins) and main journals (mains) interconnected by counterweights (webs or side walls)
and oil seal or flange [22].
Taking into consideration the specifics of the working conditions of a crankshaft, typical
causes for failures are fatigue (attributed to cycling, bending, and torsional work load­
ing and the presence of stress concentrations at the fillets), impact cracking (attributed
to the occasional impact loads), bearing seizure, and excessive wear of bearing surfaces.
Overheating of journal surfaces as a result of improper cooling caused by insufficient or
contaminated lubricants and metal-to-metal contact is another typical cause for a crank­
shaft failure.
Some of the most important requirements for automotive crankshafts are as follows:

• High strength
• Good wear and fatigue resistance
• Short length and lightweight
• Low cost
• Geometrical accuracy for vibration free and quiet performance

Important process-related factors are as follows:

• Repeatability—the same hardness pattern should be achieved over many cycles.


• Process robustness—takes into consideration real-life deviations (including geo­
metrical tolerances), making it a reliable and reproducible process.
• Minimization of energy consumption.

FIGURE 4.196
Crankshafts are widely used in internal combustion engines, pumps, compressors, and the like and belong to
the group of the critical auto components.
426 Handbook of Induction Heating

• Producing of low-roughness surfaces.


• Maximization of beneficial residual compressive surface stresses.
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

Many of these requirements have multiple and often interrelated impacts on vehicle
performance. For example, vibration does not only affect psychoacoustic characteristics,
harshness, and noise experienced by the occupants of the cabin, but vibration also affects
fatigue life through an acceleration of crack development.
These requirements continue to become more stringent with rising horsepower ratings,
the need for better fuel efficiency, the desire to reduce weight, and tightening environmen­
tal regulations.
Depending on design specifics of the crankshaft, induction hardening of journals is
used for either [22]

• Band hardening with fillet deep rolling


• Band-and-fillet hardening (also simply called fillet hardening)

Both techniques are presently used in industry depending on the specifics of the crank­
shaft application. Only the bearing surfaces of journals are induction case hardened with
the band hardening. Typically, there is an undercut in the fillet that is roll hardened in
order to improve fatigue properties and form compressive residual stresses there. Roll
hardening is applied after induction hardening.
As the term indicates, with band-and-fillet hardening, not only bearing surface but also
the fillet regions are induction hardened. Crankshaft distortion is typically substantially
greater when the band-and-fillet hardening is used, meaning greater forces are needed
during straightening operation.
The main goal of induction hardening is to produce metallurgically sound and mechani­
cally strong products with minimum distortion. A properly designed and heat-treated
crankshaft lasts well beyond the life of the vehicle.

4.9.4.2 Technologies for Crankshaft Heat Treating


Irregular geometry of crankshafts and the presence of counterweights (Figure 4.196)
exclude the possibility of using conventional encircling coils. Because the diameters of the
crank’s journals (mains and pins) are much smaller compared to the external dimensions
of the counterweights, the conventional encircling-type coils could not freely pass from
one heat-treated feature to another, dictating a necessity of developing special inductor
designs. Over the years, several induction technologies have been used for surface harden­
ing of crankshafts, including

1. Induction hardening using clamshell (split) inductors


2. U-shaped (half-shell) inductors requiring crankshaft rotation
3. Nonrotational (stationary)—SHarP-C Technology.

The Inductotherm Group (including Inductoheat Inc.) built machines utilizing all these
processes. Depending on application specifics, certain technology might be preferable. For
example, Figure 4.197 shows a low-cost, low-production system for band-and-fillet harden­
ing of crankshafts built by Inductotherm-India.
Heat Treatment by Induction 427
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

FIGURE 4.197
Low-cost low-production system for band-and-fillet hardening of crankshafts. (Courtesy of Inductotherm-
India, an Inductotherm Group company.)

As discussed in Section 4.2.3.5, clamshell inductors are so named because they are typi­
cally hinged on one side (Figure 4.100) so that the crankshaft journal can be located in
the appropriate heating position. These inductor styles may be beneficial to use in low-
production systems (e.g., hardening crankshafts for race cars). Figure 4.198 shows Radyne’s
clamshell inductor that is being effectively used for hardening a small crankshaft. Some of
the advantages of clamshell inductors are related to simplicity of operation and lower cost.
High-production systems, such as those used in automotive industry, reveal some limi­
tations in using clamshell inductors. Short inductor life and challenges with maintainabil­
ity in a high-production environment are some of the main drawbacks associated with the
use of clamshell inductors. The contact area is the weakest link and the main reason for
their short life, which does not typically exceed 10,000 cycles, and obtaining only 3000 to
4000 cycles is not uncommon.
From the 1960s to the year 2000, the majority of existing induction crankshaft harden­
ing machines for automotive industry used nonencircling U-shaped (half-shell) inductors
requiring a crankshaft rotation. This process can be used for both band hardening and fillet
hardening [22].
In a high-production environment, multiple pins and mains can be hardened simultane­
ously. Crankshafts have a lack of symmetry in particular around pin journals (the axes of the
pins are offset radially from the main axis). Therefore, during rotation, the pins orbit the main
axis. According to this process, each crank journal is heated by bringing the U-shaped induc­
tor close to the pin or main bearing surface. The inductor physically rides on the journal utiliz­
ing carbide guides (also called locators or spacers). The crankshaft rotates about its main axis
during the entire heat treat cycle. The rotational speed typically varies between 20 and 32 rpm.
Consequently, the U-shaped inductor and other quite massive components of the hardening
machine travel with the orbital motion of the connecting-rod journals.

FIGURE 4.198
Clamshell inductor that is being effectively used for hardening a small crankshaft. (Courtesy of Radyne Corp.,
an Inductotherm Group company.)
428 Handbook of Induction Heating

Precise control of circular rotation requires that the machine have complex hydrauli­
cally counterbalanced heat stations. The weight of moving components often exceeds
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

1 ton (Figure 4.199). The circular orbital motion of such a massive system (including a set
of water-cooled inductors, buswork, cables, transformers, etc.) must be maintained quite
precisely with respect to the rotated crankshaft using a special control tracking system.
The induction system must provide a complex time-dependent power modulation for
each heated journal during its rotation depending on geometry of counterweights, the
presence of oil holes, and so on.
On average, each U-shaped inductor contains 6 to 10 carbide guides (Figure 4.199, middle).
Each carbide guide physically “rides” on the hot pin/main surface (920°C and even higher).
This technology demands critically small journal-to-inductor air gaps (0.25 to 0.4 mm). As
expected, it is quite challenging to monitor the wear of each carbide guide that could lead
to an inevitable negative impact on the process and quality of heat-treated cranks.
Though carbides provide good wear resistance, they are still subject to wear. Excess
wear can cause the U-shaped inductor to accidentally contact/affect the rotating journal,
resulting in premature coil failure, water leaks, and an appearance of “soft” spots on the
surface of the heat-treated journal. This negatively affects process robustness and quality
of hardening.
Different wear rates of various carbide guides can also result in unexpected tilting of the
U-shaped inductors with respect to the journal surface, leading to air gap (coupling) varia­
tions and potentially producing pattern drift owing to localized excessive heating or heat
deficit. Carbide guides are also associated with a time-consuming, skilled setup training
and special experience in the proper adjustment of the locators, which allows for human
error. Besides, each carbide locator is simply one more part that can go wrong. Thus, it
would be beneficial to eliminate the necessity of using carbide guides.
As expected, U-shaped inductors inherently produce a nonsymmetric heating pattern
at any given time because heat is applied to only a less-than-half portion (35%–40%
being typical) of the crankshaft pin/main journal. The rest of the journal undergoes an
air cooling/soaking cycle.
The inevitable nonsymmetrical heating nature of the rotational process demands
extended time of the heating cycle in an attempt to minimize the circumferential tempera­
ture nonuniformity that could also negatively affect distortion characteristics.

FIGURE 4.199
Crankshaft hardening system utilizing multiple U-shaped (half-shell) inductors requiring crankshaft rotation.
Heat Treatment by Induction 429

Total indicated runout (TIR) is another important characteristic of the crankshaft hard­
ening. TIR directly affects grinding/straightening requirements. There are several factors
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

that affect crankshaft distortion, including its geometry, physical properties, prior micro­
structure, heat uniformity, hardness profile, and so on. One of the most important factors
is the amount of heat generated within the crankshaft body during surface austenization.
The greater the amount of heated metal, the larger the amount of metal expansion, which,
in turn, causes more distortion.
There are several reasons why excessive shape and size distortion of a crankshaft
matters:

• Excessive distortion is associated with a necessity to use extreme force during a


straightening operation, which is an elastic–plastic operation, causing a hetero­
geneous plastic deformation of the heat-treated component that may lead to the
following effects:
• Potential for crack development when trying to straighten an already hard­
ened crankshaft.
• Noticeable reduction of beneficial compressive residual surface stresses with a
possibility of their 3-D redistribution, particularly in the fillet area.
• Excessive grinding may be needed.
• Complex processes take place during surface grinding; however, monitoring
of grinding operations is not an easy task, potentially producing unspecified
results [176,375,376].
• Extreme heat generation and inappropriate grinding conditions/protocol can
alter the crankshaft’s performance and the wear resistance of working surfaces
by developing undesirable microstructures and burns, potentially resulting in
spotted hardness readings.
• Aggressive/abusive grinding can also cause cracking.
• The amount of grinding stock that is removed from the hardened case directly
affects the life of the cutting tool and process robustness.
• The above factors could negatively affect the quality of crankshafts, their perfor­
mance characteristics, and overall process cost-effectiveness; thus, it is imperative
to minimize crankshaft distortion after heat treating and is one of the reasons
why, in certain applications, the band hardening technique is preferable compared
to the band-and-fillet hardening.

Some induction manufacturers and suppliers of the rotational technology attempt to carry
out simultaneous hardening and straightening operations. According to this approach, dur­
ing the heating cycle, a certain dynamic contact pressure of the inductor guides on the heated
and rotating crankshaft is applied. This dramatically increases the complexity, worsening
the coil life expectancy (since, besides its heating functionality, the inductor intends to per­
form a straightening operation, deforming a steel crankshaft), and makes cumbersome mon­
itoring capability of the carbide guides, their wear, and overall process reliability. Besides,
localized excessive heat generation is applied as an attempt to further reduce a distortion.
The high cost of equipment maintenance, hardened pattern repeatability, and substan­
tial machine downtimes are other concerns expressed by users of rotational technology.
Although the life of the U-shaped coils is substantially longer (within 50,000–70,000 cycles)
430 Handbook of Induction Heating

compared to the clamshell coils, it is still considered by many as a noticeable drawback of


equipment that is expected to work in a high-production environment. It is highly desir­
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

able to improve inductor life expectancy, that is, to reach at least 100,000 cycles but prefer­
ably more than 120,000 cycles. There are several factors that lead to short coil life while
using U-shaped inductors:

• The combination of a necessity to have small air gaps (usually less than 0.5 mm),
relatively prolonged heat time (8 to 18 s), crankshaft surface temperature often
exceeding 920°C, and a moist working environment provided a favorable condi­
tion for stress-corrosion cracking of the inductor’s copper.
• The presence of numerous inductor joints (Figure 4.199, two images on the right)
in combination with the presence of electromagnetic forces promotes stress-
fatigue failure of copper and its surface degradation, particularly in the heat-
exposed joint areas of the inductor.
• Small inductor-to-journal gap and challenges to reliably monitor carbide guides
(locators) may produce failure attributed to uncontrollable wear of the carbide
guides, causing an inductor to have accidental physical contact with a rotating
crank, or, in other approaches, to an error of the noncontact inductor position
tracking system.

High maintenance cost and the inductors’ short life caused the users of crankshaft hard­
ening machines with rotational technology to spend an average up to $2 per crankshaft
in maintenance costs (based on data obtained from one of V-6 crankshaft manufacturer’s
practice).
In order to overcome the abovementioned drawbacks and some other customer concerns
associated with the rotational crankshaft hardening process, a nonrotational technology
was developed.
Nonrotational crankshaft hardening (SHarP-C Technology). The patented stationary harden­
ing process for crankshafts (SHarP-C Technology) has been considered by many experts to
be a revolutionary induction technology.
After its first appearance in earlier 2000 [351,352], this technology was further “fine-
tuned” [22,23,359], becoming a proven advanced process that eliminates the need to rotate
the crankshaft and utilize U-shaped coils while at the same time removing the drawbacks
of the interrupted high-current contacts associated with clamp-type inductors. Figure 2.5
shows the two-station CrankPro™ machine, which implements the SHarP-C Technology.
This patented technology was specifically developed for band hardening of automotive
crankshafts.
The SHarP-C induction system consists of a number of top (passive) coils and cor­
responding set of bottom (active) coils. Each coil has two semicircular areas where the
respected crankshaft’s journal is located (Figure 4.200). The bottom coils (being active and
connected to a power supply) are stationary, while the top (passive) coils represent elec­
trically closed-loop systems. While being unpowered, top inductors can be opened and
closed during crankshaft loading and unloading. Each inductor has profiled areas where
the crankshaft’s journals to be heat treated can be located, while the top inductors being
unpowered are in an “open” position.
After loading the crankshaft into the heating position, the top coils pivot into a “closed”
position and the power is applied from the power supply to the bottom set of (active) coils.
The AC flowing in the bottom/active coils will instantly generate the eddy currents that
Heat Treatment by Induction 431

Top (passive) coil


Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

Laminations

To power supply

Bottom (active) coil

(a) (b)

FIGURE 4.200
Circuitry of a SHarP-C induction system (a) and the pallet for simultaneous hardening of five journals
(b). (Courtesy of Inductoheat Inc., an Inductotherm Group company.)

start flowing into the top set of inductors (passive coils) thanks to a set of lamination packs
that serve as magnetic flux couplers similar in effect to a transformer core.
AC flowing in passive coil will be oriented in the opposite direction from that of the
source current flowing in the active coil. Therefore, the crankshaft journals will experience
heating in the SHarP-C inductor as it would be in an encircling induction system, making
the heating efficient and symmetric to produce a consistent hardness profile that includes
the so-called fish-tail area (or split region of the coil) utilizing patented design features to
compensate for a field fringing effect (see Section 4.2.8.3).
If the design parameters have been chosen correctly, the difference between the source
current flowing in the bottom coil (active) and the current induced in the top coil (passive)
will be negligible (less than 3%) and can be further compensated for by copper profil­
ing. It is typically sufficient to provide a “passive inductor-to-journal gap” of 0.25–0.5 mm
smaller compared to the “active inductor-to-journal gap.”
Specially designed quench slots are used to accomplish the process of quenching as well
as coil copper cooling (Figure 4.200b).
The CrankPro machine, which uses SHarP-C Technology, is available as two-station
(Figure 2.5) or three-station (Figure 4.201) depending on a production rate and crankshaft
topology. This induction hardening and tempering system provides several principal ben­
efits such as dramatically reduced distortion, simple operation, better metallurgical qual­
ity, superior reliability, and equipment maintainability, as well as substantial life cycle cost
reduction.
The customer’s materials, manufacturing, and quality engineers recognized the advan­
tages of the nonrotational hardening and tempering system immediately. Besides the fac­
tors discussed above, other benefits include the following:

1. U-shaped inductors are often fabricated using one of two techniques: copper is
bent or brazed, having multiple bents/joints. In both cases, precision and repeat­
ability of fabrication of complex geometry brazed or bent inductors are always a
legitimate concern. This leads to an extensive process validation after installing
a new set of inductors. SHarP-C inductors are more robust, rigid, and repeatable,
being CNC machined from solid copper block. All brazed or bent components
432 Handbook of Induction Heating
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

FIGURE 4.201
Three-station CrankPro machine, which realizes a SHarP-C Technology. (Courtesy of Inductoheat Inc., an
Inductotherm Group company.)

in manufacturing SHarP-C inductors are eliminated. This, in turn, dramatically


reduces the possibility of inductor distortion during its fabrication and eliminates
the associated hardness pattern drift. Compared with the complexity of U-shaped
inductors, there are far fewer components involved in the nonrotational induc­
tor design. Figure 4.202 shows side-by-side comparison of inductors used in rota­
tional (a) versus nonrotational SHarP-C (b) processes.
2. Instead of 35%–40% of coil coverage when using U-shaped inductors, SHarP-C
provides 100% coverage (Figure 4.202, compare sketches shown on the top left
vs. those on the top right), allowing minimum heat time and eliminating heat–
cool cycles that are inevitable in the crankshaft rotation process. With SHarP-C
Technology, the heat time is typically in the range of 2.5 to 4 s (compared to 8 to
18 s for the rotational process). Because of such a short heat time, only a small
mass of metal will be heated. This results in minimization of metal expansion and,
obviously, in minimization of size and shape distortions (with SHarP-C, process
distortion is typically less than 45 μm). TIR is therefore dramatically reduced.

“Node #1” “Node #2”

(a) (b)

FIGURE 4.202
Side-by-side comparison of inductors used in rotational (a) versus nonrotational SHarP-C (b) processes.
Heat Treatment by Induction 433

3. Short heat time improves the metallurgical properties of the hardened zone,
including a reduction of grain growth. The hardened zone is clearly defined and
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

crisp without the fuzzy transition zone (Figure 4.203) that is present when longer
heat times are used.
4. The inductor design incorporates a number of innovative electromagnetic fea­
tures, which dramatically reduce sensitivity of SHarP-C technology to differences
in adjacent masses. No wearing of locators/guides is involved.
5. Heat patterns are locked in place and very repeatable because neither the crank­
shaft nor the coils are moving during heat treating. The same pattern is achieved
over many cycles.
6. With rotational hardening, the inductor power can only be modulated in time and
then over an entire arc length of the U-shaped inductor. It is not possible to adjust
the heat source generation across the width of the journal. Also, the quenching
intensity cannot be modified in the oil hole area. In contrast, the spatial copper
profiling capability of SHarP-C inductors addresses the presence of irregularly
shaped areas adjacent to the journal, allowing not only circumferential compensa­
tion for irregular counterweight masses but axial as well. This permits the induc­
tor design to appropriately address the specifics of obtuse or acute oil holes, by
localized copper profiling of the inductor (Figure 4.204). It is possible to preferen­
tially change not only the heat intensity but quench severity as well, minimizing
a probability of excessive quench severities in acute regions of angled oil holes
(Figure 4.136).
7. SHarP-C Technology provides essential energy savings for the user.
a. Higher electrical efficiency is one reason for energy savings.
b. Shorter heat time and smaller mass of metal being heated result in a notice­
ably reduced heat generation within the crankshaft, which is another factor for

FIGURE 4.203
Typical SHarP-C hardness patterns. (Courtesy of Inductoheat Inc., an Inductotherm Group company.)

FIGURE 4.204
Spatial copper profiling capability of SHarP-C inductors addresses the presence of irregularly shaped
areas adjacent to the journal, allowing circumferential and axial compensation for irregular counterweight
masses.
434 Handbook of Induction Heating

energy savings (because electrical energy is used more effectively, dramati­


cally reducing the heating of internal areas that do not require phase transfor­
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

mation changes).
c. After completion of crankshaft hardening, for safety and handling purposes,
it is necessary to cool the crankshaft down. The cooling requirements are
reduced because of less energy generated within the crankshaft. This is why
the nonrotational short heat time approach will require spending much less
energy for the cooling/quenching cycle.
d. The combined effect of the factors above results in a measurable reduction
of energy consumption during both heating and quenching/cooling cycles.
Combined savings on energy consumption alone may exceed 15%–20%,
compared to rotational processes depending on case depth and crankshaft
geometry.

The SHarP-C inductors have several variations. The design described above is referred
to as “passive–active” style. Another patented approach is called an “active–active” design.
According to Inductoheat’s “active–active” approach, both the top and bottom coils are
connected to two output transformers of an inverter. Both output transformers are phase
locked, providing currents flowing in opposite directions in the respective coils. This
approach does not require lamination packs for electromagnetic coupling and allows for
individual control of the magnitude of currents in the top and bottom inductor sections.
Tempering specifics used with SHarP-C crankshaft hardening. A combination of self-
tempering and multipulse induction tempering is effectively used in crankshaft hardening
using the nonrotational stationary hardening process (SHarP-C Technology) [22]. For most
automotive crankshafts, it takes approximately 3 to 4 s to austenitize a journal surface layer
using frequencies in the range of 10 to 30 kHz (depending on the specifics of the process
requirements). After completion of austenitization, quenching is applied for only 4 to 5 s,
followed by 3 to 5 s of the first air soaking that provides self-tempering. Then, low-power
induction tempering is applied for approximately 3 to 5 s, followed by the second soaking
and the second induction tempering. It is usually sufficient to apply 2 or 3 pulses for tem­
pering. However, the process may be repeated to achieve desirable tempering conditions.
A combination of induction tempering (in which heat flows from the surface toward a
subsurface region) and self-tempering (in which heat flows from the internal region toward
the surface) delivers significant advantages compared to either one being used alone [22].
One such advantage is the substantially more uniform thermal conditions and better con­
trollability. Self-tempering, if applied alone, relies only on heat accumulated in internal
regions of the crank journal after a hardening cycle. Therefore, when only self-tempering
is used, the residual heat distribution might be quite nonuniform because various metal
masses of counterweights adjacent to hardened journals produce thermal conduction
effects of different magnitudes.
In contrast, when self-tempering and induction tempering are used simultaneously, the
processes complement each other and produce a more uniform heat distribution. Reduced
thermal gradients and favorable residual-stress profiles are created as a result of such tem­
pering protocol, while minimizing process cycle time. Keep in mind that, in some cases,
single-pulse induction tempering can result in an undesirable (reversed) residual-stress
distribution if applied improperly. A combination of multipulse induction tempering with
self-tempering provides gentle tempering conditions, optimizing tempered martensite
structures.
Heat Treatment by Induction 435

4.9.4.3 Summary Regarding SHarP-C Technology


Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

A CrankPro machine that utilizes patented SHarP-C crankshaft hardening/​tempering


technology eliminates the need to rotate the crankshaft and any movement of the inductor
during heating and quenching cycles. This stationary heat-treating process provides sev­
eral practical and technological benefits that include but are not limited to the following:

1. Superior quality of heat-treated crankshafts (producing metallurgically sound and


mechanically strong crankshafts).
2. Precise, localized heating removes the need for “heat-soak” cycling, which is
inevitably associated with rotational systems. This allows to dramatically reduce
crankshaft distortion (typically less than 45 microns).
3. High production rate (90 cranks per hour—45 s per crankshaft).
4. Significant reduction of the heat time (three- to fourfold reduction compared to rota-
tional process).
5. No wearing of the locators/guides is involved. The nonrotational process uses
inductors, which do not require contact guides or complex and expensive noncon­
tact coil positioning tracking systems of any kind.
6. Improved repeatability (the same hardness pattern is achieved over many cycles).
7. Robust and reliable process (coil life expectancy is at least doubled compared to rota-
tional process).
8. Energy savings (ranging from $150K to $260K per year).
9. The necessity of having the crankshaft journal surface at high temperature for
prolonged times, as required by rotational technology, is sometimes associated
with the undesirable metallurgical phenomenon of grain boundary liquation. This
phenomenon substantially increases brittleness and sensitivity to intergranular
cracking, particularly around the oil holes. The current-carrying copper face of the
inductor has relief, providing needed localized decoupling of the EMF around the
oil hole region.
10. Oil holes do not experience direct strike of the quenchant, reducing a probability
of cracking there.
11. Modular common base allows switching pallets for different crankshaft topology
and configuration (production ready “Quick change” pallet assembly).
12. This novel crankshaft hardening/tempering process is designed with ergo­
nomics in mind. There is easy access to all parts of the machine. CrankPro
machines are easy to operate and maintain with significant reduction of indus­
trial noise and a major improvement in coil life. Studies show that availability
of the CrankPro machine is significantly higher compared to machines that
require crankshaft rotation. The life cycle cost of conventional crankshaft heat
treatment equipment has been shown to be 3.32 times higher compared to
advanced SHarP-C technology (simple operation, improved uptime, and consider-
ably reduced maintenance cost).

Nonrotational SHarP-C Technology is an increasingly popular choice for band hard­


ening crankshaft journal bearing surfaces when medium- and high-production rates
are required. At the time of writing this book (2016) and thanks to advancements in IFP
436 Handbook of Induction Heating

Technology, the horizons of applying nonrotational SHarP-C Technology expand to fillet


hardening applications.
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

Whenever it is required to induction harden split pins, the rotational process using a
nonencircling U-shaped inductor continues to be the most widely used process at the time
of writing this book (2016).

4.9.5 Induction Hardening of Camshafts


Camshafts are used in engines controlling the timing and the speed of the opening and
closing of the intake and exhaust valves. A camshaft consists of several sets of bearings
and cam lobes that are irregular and eccentrically shaped components. As an example,
Figure 4.205a shows typical camshafts used in automobiles and truck engines. The num­
ber of cam lobes, their size, profile, positioning, and orientation are dependent on the
camshaft type and engine specifics (Figure 4.205b). Figure 4.206 illustrates common terms
associated with the functionality of a typical camshaft lobe.

(a) (b)

FIGURE 4.205
Examples of camshafts used in automobiles and truck engines (a) and variety of cam lobes (b).

Base
circle

Opening Closing
ramp ramp

Opening Closing
flank flank
Nose

Rotation

FIGURE 4.206
Illustration of common terms associated with functionality of typical camshaft lobe. (From G. Doyon, V.
Rudnev, J. Maher, R. Minnick and G. Desmier New technology straightens out camshaft distortion, Industrial
Heating, December, 2014, pp. 46–52.)
Heat Treatment by Induction 437

Since the cam lobe profile is the working surface having a contact with the follower at
high revolutions per minute, it requires sufficient strength and wear resistance. Therefore,
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

it is imperative to obtain sufficient surface hardness. Since maximum cyclic shear stress is
located beneath the surface, it is equally important to have a sufficient hardness case depth
to prevent a subsurface crack initiation and plastic deformation. The Hertzian analysis can
be used in fatigue calculations.
Depending on application specifics and economic reasons, various materials can be
used for the manufacturing of camshafts. This includes nodular cast irons, gray cast irons,
plain carbon steels, and low-alloy or microalloy steels. Cast irons are attractive because of
their production cost-effectiveness, while forged steels maximize tolerated compressive
stresses, increasing strength.
In the past, the hardening of camshafts was accomplished with flame and then submerg­
ing in an oil bath. There are a number of well-known drawbacks associated with flame
hardening, including excessive distortion and metallurgically undesirable structures that
often cause cracking.
Induction hardening utilizing medium frequencies (usually from 3 to 40 kHz) has elimi­
nated many of the problems of flame hardening. Induction hardening is used for cam­
shafts as large as ship and train cams to as small as cams for lawnmower engines. The
size of the camshaft usually influences the selection of the processing mode. For material
handling purposes, large camshafts for ship or locomotive engines are typically processed
in a horizontal manner. Automotive and truck camshafts can be processed vertically or
horizontally.
A variety of inductor designs and heating modes have been used for induction hard­
ening of camshafts. Depending on application specifics, the following heating modes
are used [22]:

• Scan hardening
• Single-shot hardening
• Static (nonrotational) hardening

4.9.5.1 Scan Hardening
This camshaft hardening mode is typically chosen for relatively low production rates and
wide and large-diameter lobes or when available power is limited. This mode results in a
lower initial capital cost of the machine. Lobes can be rotated during the heat treating or be
motionless. When lobes are scan hardened without rotation, the inductor copper is shaped
(Figure 4.207c) to accommodate lobe geometry and compensate (to some degrees) for the
lack of heat generation in the lobe base circle (also referred to as a heel). An appropriate
lobe orientation in respect to an inductor is required. This process is particularly beneficial
for hardening moderate- and large-sized lobes.
The heating face of scan inductor is relatively narrow (Figure 4.207a), minimizing axial
HAZ. Therefore, only a portion of the single lobe is heated at a particular moment of time,
requiring a minimum amount of power as opposed to a single-shot mode that requires
heating the whole feature at once. Another advantage of scanning is the ability to vary the
power and scan speed.
The main disadvantage of the scanning mode is longer cycle time compared to heating
the whole feature at the same time. Some difficulties may occur when scan hardening nar­
row lobes as well as lobes that are positioned close to each other.
438 Handbook of Induction Heating
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

(a) (b) (c)

FIGURE 4.207
Conventional camshaft hardening inductors. (a) Scan inductor applying narrow heating face. (b) Multiturn
coil for hardening simultaneously multiple lobes. (c) Profiled inductor. (Courtesy of Inductoheat Inc., an
Inductotherm Group company.)

4.9.5.2 Single-Shot Hardening
In contrast to scan hardening, single-shot heating of single lobe or multiple lobes with
camshaft rotation is commonly used when surface hardening small- and medium-sized
camshafts with similar lobes. Multiple lobes can be heat treated simultaneously using
multiturn coils to increase the production rate (Figure 4.207b). For example, if a four-turn
coil provides heating of a camshaft that consists of eight lobes, it will require only two
heating positions to do a job of hardening. Though application of multiturn coil drastically
increases the production rate, it also requires more power, because several lobes are being
heated at one time.
The possibility of using multiple turns is related to a topology of the camshaft. The
current-carrying face of different turns is frequently profiled to properly address the field
interaction between neighboring turns. Flux concentrators are often installed on the top
and bottom of the inductor.
In order to harden some truck camshafts, both a multiturn inductor to heat the intake,
injector, and exhaust lobes, and a single-turn inductor to heat the bearings are used.
Single-shot hardening is not as flexible as scan hardening because inductor geometry
and copper profiling are usually specific to a particular camshaft, and only relatively
small deviations of cam lobe topology (including size and their spacing) are tolerable. This
necessitates having a number of inductors in order to process camshafts with considerably
different geometries and practically excludes the use of off-the-shelf inductors.
Single-shot hardening (either a single lobe or multilobe) applying camshaft rotation usu­
ally produces deeper hardness depth in the nose of the cam lobe compared to a base circle/
heel (Figure 4.208), because the nose has better coupling with the coil. Therefore, because
of the electromagnetic proximity effect, a greater amount of heat is generated within the
closer coupled areas (such as the nose of the cam lobe).
To illustrate this, Figure 4.209 shows results of FEA computer modeling of EMF dis­
tribution in the case of single-lobe hardening using a bare coil (Figure 4.209a) and
a coil with a “U”-shaped magnetic flux concentrator (Figure 4.209b) [468]. As can be
seen, because of the proximity effect, more imaginary lines of the EMF are closing
their loops in air without reaching the surface of the heel of the cam lobe compared to
Heat Treatment by Induction 439

Right 3.5
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

1 2 3 4 5 6 7 8
2.0 2.5

1.8 Lobe 1l 2.4


Left

1 2 3 4 5 6 7 8

2.0

FIGURE 4.208
Hardness patterns applying a conventional camshaft hardening processes. (Courtesy of Inductoheat Inc., an
Inductotherm Group company.)

Single-turn Flux
coil concentrator

“A” “B” “C” “A” “B” “C”

(a) (b)

FIGURE 4.209
Results of FEA computer modeling of field distribution in case of a single-lobe hardening using a bare coil (a)
and a coil with a “U”-shaped flux concentrator (b). (From V. Rudnev, Aspects of Induction Heat Treating, Invited
Lecture for Peoria ASM Chapter Meeting, Peoria, IL, January 13, 2015.)

its nose producing more intense and deeper heating at the nose compared to the base
circle (the heel).
In order to obtain a proper austenization of the poorly coupled heel region, using con­
ventional camshaft hardening processes, it is usually necessary to extend the heating time
or reach appreciably high maximum temperatures at the nose. This leads to a considerably
nonuniform hardness pattern along the circumference of the lobes and to the possibility
of crack development. There is another challenge associated with this process. When heat
treating camshafts with closely positioned lobes and journals, there is a danger of causing
undesirable softening of adjacent areas that have previously been hardened [299]. As can
be seen from Figure 4.209, an EMF spreads around an induction coil and also links with
electrically conductive surroundings, including adjacent cam lobes and journals. Heat
generated can cause undesirable metallurgical changes in corner and edge areas of the
camshaft lobes that were hardened in a previous process stage.
440 Handbook of Induction Heating

4.9.5.3 Static (Nonrotational) Hardening


Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

With static (nonrotational) hardening, both the inductor and camshaft are motionless.
Several inductor designs have been used over the years to statically harden cam lobes.
Many of those designs were similar to the designs used for hardening crankshafts (dis­
cussed in Section 4.9.4) and includes the following:

• Conventional single-turn coil with profiled heating face to accommodate the


geometry of the lobe. In an attempt to reduce the heat generation in the nose of the
cam lobe and taking advantage of the flux frigging effect, the nose is positioned in
the area where buses that transmit electrical current from a power source are con­
nected to the induction coil (Figure 4.207c). Flux concentrators can also be applied
in selective locations of the inductor to intensify heat generation in poorly coupled
areas of the base circle. Before the beginning of the heat cycle, the coil can also be
moved closer to a base circle of the lobe and de-couple its nose as an additional
attempt to balance the circumferential heat generation. Regardless of those efforts,
this design is still associated with poor controllability of the hardness pattern, low
heating efficiency, and large camshaft distortion, and its utilization is therefore
very infrequent.
• Clamshell or split inductors are used in camshaft hardening but not too often. The
pros and cons of these inductors have been discussed in Section 4.2.3.5.
• Nonrotational SHarP-C Technology developed for crankshafts (see Section 4.9.4)​
has also been effectively used for true contour hardening of cam lobes
[22,23,27,299,359]. The inductor consists of a top (passive) inductor and a bottom
(active) inductor (Figure 4.200). After the loading of the camshaft into the heating
position (Figure 4.210), the top coil pivots into a “closed” position and the power
is applied from the power supply to the bottom (active) coil. The current starts to
flow in the bottom coil, and thanks to a lamination pack that serves as a magnetic
flux coupler, both coils (top and bottom) are tightly electromagnetically coupled.
Therefore, cam lobes “see” the nonrotational inductor as an encircling and highly
electrically efficient coil having uniform or profiled air gaps (whatever is desired)
to obtain true contour-hardened profiles. Figure 4.211 shows several examples of

(a) (b) (c)

FIGURE 4.210
CamPro machine utilizing patented SHarP-C Technology allows one to dramatically improve metallurgical
characteristics and straightness of camshafts. (a and b) Close-up views of SHarP-C inductor in open and close
positions, respectively; (c) CamPro machine utilizing patented SHarP-C Technology. (Courtesy of Inductoheat
Inc., an Inductotherm Group company.)
Heat Treatment by Induction 441
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

FIGURE 4.211
Examples of true contour hardening patterns of camshaft lobes. (Courtesy of Inductoheat Inc., an Inductotherm
Group company.)

true contour hardening patterns. Figure 4.210a shows that eight lobes were heat
treated at once. This advanced process lends itself to simultaneously hardening
several lobes. Therefore, high production rates can be achieved.

The compound benefits of patented SHarP-C Technology for heat treating camshafts can
be summarized as follows:

• Achieving almost undetectable camshaft distortion (approximately 3–5 microns;


based on 1.5- and 2.0-L diesel or regular fuel engines) and, in many cases, an elim­
ination of an entire straightening operation [27,299]. This is the combined result of
three factors: (1) the ability to form a true uniform hardness pattern, (2) reduction
of localized temperature peaks during austenization, and (3) avoidance of apply­
ing any pressure/forces during camshaft hardening.
• Experience of using SHarP-C camshaft hardening technology reveals producing
not only superior straightness but also better metallurgical properties forming
fine-grain martensitic structures and minimizing a probability of crack develop­
ment and grain boundary liquation owing to an improvement in temperature uni­
formity along the cam lobe surface.
• Elimination of undesirable softening (temper back) of previously hardened neigh­
boring lobes.
• The energy consumption during both heating and cooling is reduced. Depending
on the specifics of the camshaft’s geometry and heat treat specifications, com­
bined savings on energy consumption may exceed 12%–18% compared to other
processes depending on material, case depth, camshaft shape, size, topology, and
topography.

The CamPro machine (Figure 4.210c) utilizing patented SHarP-C Technology can serve
as an example of dramatically improved straightness characteristics [22,27,299]. A testimo­
nial of one of the users of this advanced process published in Ref. [27] could be considered
an objective assessment quantifying benefits of this technology based on obtained real-life
records: “The SHarP-C hardening machine helped us to reduce the camshaft’s distortion
down to 3–5 microns and we have been able to eliminate the entire straightening opera­
tion. So, our savings on elimination the straightening operation alone is about $40,000 per
year. On top of that there has been substantial improvement in the quality of the hardened
camshafts, and our scrap was reduced about 1.5%.”
442 Handbook of Induction Heating

4.10 Solutions for Solving Typical IHT Problems: Questions and Answers


Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

4.10.1 Case Study #1. The Power–Scan Rate–Hardness Relationship


Question: “When purchasing an induction scan hardening machine, it is not unusual to
choose a system with an extra margin of available power (10% to 20%) to accommodate a
potential future growth in production. It seemed to make good sense: the increase in scan­
ning speed required to increase production rate could be simply compensated for by a cor­
responding increase in applied coil power. However, once the scanner started producing
parts at a higher rate using a higher scan speed, it was suddenly observed that hardness
specs aren’t being met and hardness readings are inconsistent (scattered soft and hard
spots). Why is that?”
Answer: A common problem that can occur when trying to increase scan speed with
a corresponding increase in applied power involves inconsistent hardness readings that
may be attributed to incomplete austenitization. It is important to remember that raising
the heating rate affects the kinetics of austenite formation by increasing all critical tem­
peratures (see Figures 4.27 and 4.28) [469].
Depending on the initial microstructure of the part—annealed, normalized, or quenched
and tempered—it might not be enough to just increase the power. Steels with large stable
carbides (annealed and spheroidized microstructures) naturally have a poor response to
rapid induction hardening and thus require prolonged heating and higher temperatures
for austenitization. Microstructures having coarse ferrites or large bands of ferrites would
also require a certain minimum amount of time to allow carbon to diffuse into the carbon-
lean ferrite.
Solution: If metallographic examination reveals “ghost pearlite” or “an excessive
amount of free ferrite,” then an appropriate increase in coil heating face must accom­
pany the increase in applied power to properly compensate for the faster scan speed.
Bear in mind, however, that an increase in coil heating face also results in a longer
axial transition zone.
Another potential cause of lower-than-expected hardness readings could be attributed
to insufficient quenching that can occur when scan speed is increased and a sufficient
increase in power is applied to the coil. Adjustment of only the power level or coil length
(heating face) might not be sufficient. To ascertain the cause of the problem, remember that
hardening is a two-part process: heating and quenching. If scanning speed is increased,
both heating time and quenching time will be shortened. As already mentioned, a reduced
heating time can often be compensated for by an appropriate increase in applied power
and by adjusting coil length. However, compensating for quench time reduction requires a
different approach. The objective is to ensure a martensitic case, avoiding the formation of
upper transformation products, such as upper bainite and pearlite, which can result when
quench severity is less than that required for martensitic transformation or when quench­
ing is not long enough. This occurs when the cooling curve “enters” the upper transfor­
mation region of the CCT diagram. Insufficient cooling might also lead to an undesirable
“tempering back,” which results in reduced hardness of the as-quenched part.
If lower hardness is observed, the first step is to conduct a metallographic analysis of
the as-quenched part. If the microstructure contains tempered martensite or upper trans­
formation products, then a very short quench is probably the cause. To ensure proper
hardening, increase the quenchant flow rate and add or enlarge the quench follower
(Figure 4.56b).
Heat Treatment by Induction 443

4.10.2 Case Study #2. Cryogenic Coil Cooling


Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

Question: “Since it is beneficial to fabricate coils from low-resistive metals, why not use
cryogenic cooling of the induction coil instead of water cooling? Cryogenic cooling will
further dramatically decrease copper resistance, reduce kilowatt losses, and therefore
improve coil efficiency and life.”
Answer [470]: Power loss in induction heating systems includes losses in power supply
and capacitors, transmission losses, unproductive power generation in the coil surround­
ings (e.g., enclosures, tooling, fixtures, water-cooled liners, and rollers), thermal losses
from the workpiece surface, and actual losses in coil copper windings. When heating non­
magnetic, low–electrical resistivity metallic materials (e.g., copper), the power loss in coil
copper can be greater than all other losses combined. Therefore, reducing kilowatt losses
in coil copper represents the main approach to improving total efficiency. An obvious way
to decrease coil losses is to reduce coil resistance, Rcoil, which can be expressed in a simpli­
fied manner as Rcoil = ρcoil (L/A), where ρcoil is the electrical resistivity, L is the length of the
current-carrying conductor, and A is the area through which the current flows. Generally
speaking, ρcoil can be reduced by using electrically conductive materials that exhibit low
electrical resistivities. Unfortunately, there are not too many options in selecting materials
for inductor fabrication, since copper is among the metals that have the lowest electrical
resistivity (see Section 4.2.4.1).
Theoretically, the phenomenon of superconductivity can be used to measurably reduce
Rcoil and, therefore, significantly reduce unproductive kilowatt losses in coil copper and
improve heating efficiency, particularly when heating low-resistivity metals. Copper’s elec­
trical resistivity falls as temperature decreases. It has been shown that by taking advan­
tage of the superconductivity phenomenon when heating aluminum billets, coil electrical
efficiency can be increased from approximately 45% to as high as 88%. Unfortunately, this
requires extremely low temperature, which necessitates the use of liquid nitrogen for cool­
ing coil turns instead of water. Such an approach would result in a significant increase in
the cost of the induction heater. At the same time, although coil electrical efficiency would
be noticeably improved by cooling with liquid nitrogen, total system efficiency could still
suffer because of low efficiency of the cryogenic cooling system itself, resulting in a further
drop in overall efficiency. Besides that, the use of such low temperatures will inevitably be
associated with increased brittleness of coil copper turns and potential crack development
due to electromagnetic forces experienced by the heating inductor.

4.10.3 Case Study #3. Chain Hardening


Question: “We have a chain-hardening application. Is it possible to have different hard­
nesses along the link perimeter while utilizing a multiturn induction coil?”
Answer [470]: Sometimes, it is preferable to have a certain hardness distribution along
the perimeter of chain links. Some customers choose to have slightly higher hardness in
link areas that experience wear, while having slightly more ductility in areas that experi­
ence tension. Certain combinations of process parameters (power, frequency, and heating
time) might make it possible to induction harden chain links to produce a specific hard­
ness gradient along the link’s perimeter depending on the link size and the ability to con­
trol eddy current cancellation. Of course, if required, it is also possible to obtain the same
hardness along the perimeter. Power supplies that provide independent frequency and
power regulation (such as Statitron IFP) are very helpful in achieving the needed thermal
conditions.
444 Handbook of Induction Heating

4.10.4 Case Study #4. Hardness at Diameter Change


in Scan Hardening of Stepped Shafts
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

Question: “For quite a while after scan hardening, we experience low hardness readings
(52–55 HRC) in areas where the diameter of our 1045 steel hollow shaft changes. We have
consulted with induction heating specialists, but still have not eliminated the problem.
What causes soft spots at these locations and how we can fix this problem?”
Answer [310]: On the basis of the hardening properties of SAE 1045 steel, it is reason­
able to expect achieving a maximum as-quenched surface hardness of approximately
60–62 HRC [157]. Under normal conditions, a surface hardness in the range of 52–55 HRC
is lower than the expected as-quenched hardness. Generally, there are three groups of fac­
tors that could cause the appearance of soft spots:

• Metallurgical/microstructural factors (e.g., flaws in the steel’s prior structure, severe


chemical and phase segregation, presence of decarburization after previous oper­
ation, excessive amount of RA, and some others)
• Inappropriate heating conditions (e.g., insufficient austenitization or formation of
heterogeneous austenite, localized overheating of a larger-diameter shoulder, or
insufficient localized austenization of the smaller-diameter area near the shoulder)
• Quench-related factors, including inappropriate impingement angle, insufficient
quench flow, and temper back effect

Taking into consideration that hardness soft spots are associated with the diameter
change area, most likely they are caused by one of the reasons related to the second or
third group of factors.
If lower-than-expected hardness readings appeared at the large-diameter shoulder near
the diameter change area, then most likely this is caused by the heat surplus there. This
could produce unwanted RA, an excessive heat accumulation that may result in temper
back softening.
If lower-than-expected hardness readings appeared at the small-diameter area (e.g., fil­
let region or undercut) of the diameter change area, then most likely this is caused by
insufficient austenization or improper quench impingement.
Therefore, if you have a handheld pyrometer, compare temperatures of larger and
smaller diameters of the shoulder area to determine which of those two heat-related fac­
tors might contribute to this problem. Adjustments of scan hardening recipe, using nar­
rower heating face scan coils, can help fix this problem. Also, instead of scan hardening,
consider using a single-shot inductor, which noticeably improves heating of diameter
change regions.
Shaft orientation during scan hardening can also have an impact on quench inten­
sity and uniformity, potentially causing lower-than-expected hardness readings within
the diameter change region. For a stepped shaft processed in the undesirable direc­
tion, certain areas of the shaft shoulder may be hard to access and not be sufficiently
quenched out—spray quench strokes could miss the smaller diameter located near the
shoulder, resulting in insufficient quenching and the appearance of soft spots. Therefore,
if spray quench impingement angle (the angle at which the quenchant strikes the work­
piece) is too small, then its increase can help better quench out of the smaller-diameter
area near the diameter change. This is one of the reasons why quench assemblies fre­
quently require redesign when processing shafts with geometrical irregularities that
are processed in different directions. Keep in mind that an increase in an impingement
Heat Treatment by Induction 445

angle should not be too drastic, which may lead to splash back of quenchant into the
heating coil and cause a spotted hardness pattern. In order to sufficiently quench out
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

hard-to-access regions, some practitioners dramatically increase quench pressure or


flow. Excessive quench can also lead to overflooding and splashing in the scan inductor,
resulting in a nonuniform hardness pattern.

4.10.5 Case Study #5. Quenchant pH and Corrosion


Question: “After running powder metallurgy parts with a heavier production schedule
for about 2 to 3 months, we started seeing rust stains mostly located in the single-shot
induction-hardened area. When we changed the quenchant, the corrosion disappeared.
After 1 to 2 months, we again saw corrosion. Changing the quenchant cleared the prob­
lem again. We keep our polymer quenchant between 4% and 6% solution, and it is moni­
tored daily. We heard that adding small amounts of bleach helps to stretch out the time
between flushes. What is the recommended practice? Does induction heating promote
corrosion?”
Answer [310]: Biological degradation of the quenchant can result in a reduction of pH
value, causing the solution to become acidic, which can promote excessive corrosion as a
result of acid attack. Among other factors, the rate of corrosion also depends on a type of
quenchant. For plain water and most water-based polymer-type quenchants, the severity
of corrosion rate dramatically increases when the pH is less than 6. The intensity of the rate
of corrosion increases when pH decreases, somewhat of a “snowball” effect. This is why it
is preferable to have pH values greater than 7. However, too high of a pH level (e.g., pH >12)
is also not desirable, because it can potentially cause phenomena called caustic cracking
and caustic gouging. Typically, pH values between 8 and 9 are considered to be the most
desirable unless a membrane separation is used, in which case there is a specific pH range
for the solution to be maintained to ensure the stability of the membrane [471]. You should
consult with your quench suppliers and review the ASTM Standard Guide for Evaluation
of Aqueous Polymer Quenchants.
Electromagnetic induction does not enhance any substantial corrosion. However, it is
well established that corrosion intensity increases with increasing temperature (regardless
of whether the temperature increase is after induction heating or any other heating means,
assuming that the rest of the conditions remain the same). Therefore, if certain regions of
the heat-treated component have different temperatures, then areas at higher tempera­
tures will experience greater corrosion compared with areas at lower temperatures. The
reason why rust stains were mostly local to the induction-hardened area is most likely
related to the presence of residual heat at those locations.
Under no circumstances should bleach or swimming pool chemicals be added to quen­
chants in an attempt to stretch out the time between flushes. This will cause further deg­
radation of the polymer and possibly lead to the formation of chloroethers, which are
considered by many as carcinogenic [471]. A study conducted by K.P. Kirkbride and H.J.
Kobus [472] shows that at a certain concentration, there is even danger of an exothermic
reaction leading to an explosion and skin burns.

4.10.6 Case Study #6. Copper Coil Wall Thickness vs. Coil Life
Question: “I use induction heating to anneal selected areas of my nonferrous metal part.
However, coil life is quite short. Our coil supplier sent a quote suggesting that we rebuild
the coil using copper tubing with a thicker wall, claiming that this would strengthen the
446 Handbook of Induction Heating

coil and improve its life. Will increasing wall thickness make the coil last longer? Presently,
we use rectangular copper tubing 3/8 × 1/4 in., 0.048 in. wall (9.5 × 6.35 mm, 1.22 mm wall).
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

The frequency is 30 kHz.”


Answer [470]: The idea that thicker-wall copper automatically leads to longer life is
deceiving. Increasing tubing wall thickness can improve or shorten coil life, depending on
the specifics of the application. There is a combination of factors that interact:

1. Increasing copper wall thickness will increase tubing strength. This is a step in
the right direction.
2. At a given frequency, if the copper wall thickness does not exceed the coil cur­
rent penetration depth, then an increase in wall thickness will reduce coil copper
resistance and, therefore, reduce copper losses, improve electrical efficiency, and
most likely increase coil life. At 30 kHz, the copper penetration depth is less than
0.4 mm at a water-cooled temperature. Taking into consideration that the pres­
ently used copper tube wall thickness is 0.048 in. (1.22 mm), that ratio will be
1.22 mm/0.4 mm, or 3.05 times. Therefore, from an electromagnetic perspective,
your copper wall is already sufficiently thick, and in this case, a thicker wall will
not appreciably reduce coil copper kilowatt losses.
3. In contrast to Factor 1 discussed above, increasing copper wall thickness reduces
the size of the water-cooling channel. Originally, the copper tubing was 3/8 ×
1/4 in. (9.5 × 6.35 mm), with a 0.048-in. (1.22-mm) wall. This means that the water
channel area was 0.278 × 0.154 in. (7.06 × 3.91 mm). Let us say that you increase
your copper wall thickness to 1/16 in. (1.59 mm), and then the new water chan­
nel area will be only 0.249 × 0.125 in. (6.32 × 3.17 mm). Depending on coil current,
this might be too small a water channel to sufficiently cool coil copper that carries
several thousand amperes at 30 kHz. Therefore, from this perspective, an increase
in copper wall thickness will be a step in the wrong direction.
4. Another phenomenon that takes place when you increase copper wall thickness
deals with the fact that the current-carrying copper layer is the area that experi­
ences the highest temperature (maximum current density and heat radiation from
the workpiece). Even though copper has good thermal conductivity, increasing
the copper wall thickness reduces the effect of water cooling on the coil. This
would also be a step in the wrong direction. In this specific case, coil life can
be noticeably improved if you use larger copper tubing. By substituting 3/8 ×
5/16 in. (9.5 × 7.9 mm) tubing, then copper current density and copper heating
will decrease and coil life will improve. Actually, in this case, it is also possible
to increase copper wall thickness without reducing the water-cooling channel
cross section. By the way, the electrical efficiency of a longer coil is also greater
than that of a short coil. However, it is important to remember that if a heat-
treating process has been established, an increase in copper cross section could
potentially cause a change in the heating profile. Therefore, hardness pattern
verification is required.

4.10.7 Case Study #7. Residual Magnetism


Question: “RFQ of our potential customer for normalizing, stress relieving and anneal­
ing of carbon steel bars demands guaranteeing control of the residual magnetism to less
than 5 G level. How much is it? And how can I control it? Another project required the
Heat Treatment by Induction 447

following: ‘Do not introduce any additional magnetism into the product after induction
heating.’ In the past, we supplied several copper annealing systems and nobody asked us
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

to guarantee not exceeding a certain maximum permissible value of residual magnetism.”


Answer [310]: In your copper annealing lines, you did not have to worry about residual
magnetism since copper is a diamagnetic (nonmagnetic) material.
In normalizing or full annealing applications dealing with carbon steels, you typically
do not have any problems with residual magnetism either, because target temperatures
are way above the Curie point. However, temperatures for stress relieving, tempering,
and subcritical annealing are always below the Curie point. This makes the situation quite
different.
The customer requirement “do not introduce any additional magnetism into the work­
piece as it was heated” might be very tough to fulfill. Actually, some residual magnetism
after induction full-body preheating at below Curie temperature applications is inevitable.
The magnitude of residual magnetism depends on many factors. This includes but is not
limited to the following:

1.
Magnetic properties of ferromagnetic materials (e.g., soft magnetic vs. hard magnetic,
steel grade, its prior treatment, and structure). For example, low-carbon steels (e.g.,
SAE 1006) have a typically greater amount of residual magnetism compared to
high-carbon steels (e.g., SAE 1080).
2.
Type of inverter, presence of harmonics, and the inductor style. Usually, progressive
and continuous heating applications utilizing loosely coupled solenoid coil(s)
and systems with a parallel connected load produce less residual magnetism
because a parallel circuit “provides” a degaussing effect to some degree. In con­
trast, static heaters and series output voltage-fed inverters with tightly coupled
non-solenoid inductors (e.g., channel coils) usually produce a greater amount of
residual magnetism.
3.
Distribution of coil power and frequency along the heating line. Low-density soak coil(s)
positioned at the end of the line typically allow reducing a residual magnetism.
Coils with tighter coupling and lower frequencies produce a greater magnitude of
residual magnetism compared to loosely coupled higher-frequency coils.
4.
Temperature range. Lower processing temperatures are associated with higher
magnitudes of residual magnetism.

The required maximum magnitude of residual magnetism of less than 5 G might be


quite challenging to achieve in some applications. A value of 5 G is quite small. To illus­
trate, consider the following: 5 G is equal to 0.0005 T. Steel SAE 1045 saturates at approx­
imately 1.6 T. Therefore, this means that the maximum allowable residual magnetism
must be 3200 times lower than the magnetic saturation of steel. As a matter of fact, the
magnetism of Earth is approximately 0.6 G or so. This is the reason why one of the criti­
cal demands when building plants for building lines to process long products (e.g., tubes,
pipes, bars, rods, etc.) is associated with a critical magnetization requirement in that it
is important to orient lines along Earth latitudes and not along longitudes. Actually, in
some cases, it is prohibited to process long products along longitudes because the Earth
magnetism will inevitably magnetize the long products.
It is not easy to measure the residual magnetism in a production environment, particu­
larly in progressive/continuous IH lines. Sometimes, professionals apply a simple test by
bringing small-sized plain carbon steel particles/pieces (1 mm × 1 mm × 1 mm, 5 mm ×
448 Handbook of Induction Heating

5 mm × 1 mm, 10 mm × 10 mm × 1 mm, etc.) or even small screwdrivers or wrenches close


to the workpiece. Estimation can be made regarding the magnitude of residual magnetism
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

based on the size/weight of those test pieces that would stick to the magnetized workpiece.
Applying industrial degaussing devices positioned at the end of the heating line can
help reduce the magnitude of residual magnetism. However, most available degaussing
devices do static degaussing. In your case, you need to use a degaussing system for a
continuous line. A small, lower-power soaking inductor (particularly a graded/profiled
coil with turn spacing gradually increasing toward the exit end) positioned at the end of
the line will “act” as a degaussing system to some degree. However, it is hard to quantify
demagnetization and predict whether a magnitude of residual magnetism will be less
than a certain level. Thus, I suggest you to talk to your local manufacturer of degaussing
systems and see what level of residual magnetism they would guarantee to you. Then,
review the above-discussed factors with your customers.

4.10.8 Case Study #8. Heating with Variable Axial Gaps between Long Rods
Question: “We are induction heat treating long steel rods of about 0.75–1 in. (19–25 mm)
diameter using two in-line multiturn coils (for hardening) followed by four in-line coils
(for tempering). Sometimes, rods are processed end to end, but in other cases, there may
be variable gaps between rods being within 25–100 mm (1–4 in). We notice hardness varia­
tions in the end zones of rods, as well as grain coarsening and an excessive amount of
retained austenite present there. Any clues?”
Answer: The phenomenon that you are experiencing is typically referred to as electro­
magnetic transient end effect. This subject will be discussed in Chapter 6. In your applica­
tion, a complexity of this phenomenon is twofold.
When heating carbon steel rods for hardening, an axial gap of an appreciable size
between ends of neighboring rods causes higher austenizing temperature of the end
zones. This factor could potentially produce grain coarsening, excessive amount of RA,
and hardness variation.
Besides that, the presence of an axial gap between rods in tempering lines may produce
lower or higher temperatures at the end zones of rods compared to their regular regions
and also resulting in respected variations in tempering hardness drop.
When rods travel end to end through an induction heating line, the nose-to-tail tem­
perature uniformity is typically not a problem (excluding the first and the last bars of the
group). However, the existence of appreciable size axial gaps between rods could result
in noticeable axial temperature nonuniformity particularly at rod end region because of a
distortion of magnetic field there. The task of obtaining nose-to-tail temperature unifor­
mity is typically a more difficult one compared to minimizing the surface-to-core temper­
ature gradient. An explanation of this phenomenon and recommendations to minimize
temperature variations are provided in Section 6.2.3.

4.10.9 Case Study #9. Hollow Shafts with Holes and Splines


Question: “We are processing solid and hollow shafts made of hardenable stainless steels
and low-alloy steels using MIQ scan inductors. Shafts have different holes and splines
and various diameters. The frequency is 120 kHz. Regardless of an absence of flaws in the
steel’s conditions before hardening, we sometimes observe soft spots. What are the factors
that might be causing those soft spots?”
Heat Treatment by Induction 449

Answer [310]: Several factors might be responsible for the appearance of lower-than-
expected hardness readings (soft spots), assuming an absence of flaws in the steel’s prior
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

structure:

1. Check part prints and expected tolerances. Excessive tolerances can have a signifi­
cant impact on workpiece-to-coil coupling, resulting in temperature deviations in
particular when RF frequencies are used. This can also affect the amount of RA in
particular when dealing with MSS or alloy steels. The amount of RA affects hard­
ness readings.
2. If scattered soft spots are localized to fillets, undercuts, holes, and areas of transi­
tion from larger to smaller diameters, and if metallographic examination reveals
low-hardness readings associated with the appearance of “ghost” pearlite, or
an excessive amount of free ferrite, then, most likely, the problem is related to
incomplete austenitization. The microstructure formed upon quenching incom­
pletely transformed austenite typically consists of a combination of lower and
upper transformation products, causing the appearance of soft spots. Solution:
Appropriate increases in coil power, a slight reduction of scan speed near the low-
hardness region, and using the appropriate frequency can improve the kinetics of
austenite formation there.
3. If the appearance of lower-than-expected hardness occurs only in a shoulder area,
soft spots can be caused by the following factors: (a) Reduced hardness could be
a result of undesirable tempering back owing to the comet-tail effect (see Section
4.2.2.1.2) if metallographic examination reveals the presence of tempered martens­
ite within the as-quenched structure. (b) If metallographic examination reveals
the presence of upper transformation products, quench severity is less than
required for martensitic transformation and the cooling curve entered the upper
transformation region of the CCT diagram. Solution: It is important to continue
quenching the shaft after coil power is turned off regardless of whether the shaft
surface appears to be sufficiently cooled. This ensures removal of the heat retained
in the shaft subsurface and sufficient compensation for the comet-tail effect. In
both cases, improvement in quench severity (i.e., flow rate increase) and extending
quench time by using a quench follower will fix the problem.
4. It should be understood that although high rotational speed can provide more
uniform circumferential heating, it may also result in the deflection of the quen­
chant and a reduction in cooling severity, particularly if geometrical irregularities
are present. If the rotation is too fast, the quench fluid might not be able to provide
proper quenching in certain areas, resulting in soft hardness readings.
5. Because of the complexity of the EMF in a diameter change area, there is always
a reasonable compromise between a surplus of induced power in the shoulder
of the large diameter and its deficit in the fillet or undercut of the neighboring
smaller-diameter area. Solution: Process controllability can be improved by apply­
ing a coil having a narrower face or by lowering the frequency or using a single-
shot or a serpentine-style inductor. Shaft orientation during scan hardening can
influence the spray quenching uniformity and the appearance of scatter hardness
within the diameter change region. Review comments for Case Study #4 above.
6. The presence of radial holes, grooves, and keyways can also cause lower-than-
expected hardness regions because of a localized heat deficit. This subject was
discussed in Section 4.3.
450 Handbook of Induction Heating

4.10.10 Case Study #10. Graphs of Hardened Case Depth vs. Power
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

Question: “I am a Corporate Metallurgist and preparing a short course on ‘Introduction to


Induction Hardening’ for engineering personnel and came across a number of graphs and dia­
grams (some of them are migrating from book to book; therefore, I assumed that they represent
reliable pieces of knowledge). All makes sense in it except for a few things, and this is my first
question: (1) curves for hardened case depth versus surface power densities have a negative
slope. This means that, for a particular frequency, if surface power goes down, the hardened
case depth goes deeper. It makes sense to me only from the point of view of skin effect—with
lowering the power input, the skin gets deeper and thus will affect hardened case depth in the
same way. But I think time is an independent variable in this case and has no relevance.
On the other hand (and this is my second question), (2) when I talk to people who set
the machine recipe (hands-on people on the shop floor), they all say that this is not correct
as they see the reverse—when they increase power without touching time, the case depth
gets deeper (which I instinctively agree with because I experienced this myself many
times). What is your take on that?”
Answers: The first statement is not quite correct. Above the Curie temperature (at aus­
tenizing temperatures), the skin effect, practically speaking, is independent of the power
density. The eddy current penetration depth (δ) is the quantitative measure of the skin
effect and it is a function of only three variables: electrical resistivity, relative magnetic
permeability, and frequency. Above the Curie point, μr = 1 and both electrical resistivity
and frequency are independent of power density. The use of high power densities in com­
bination with the short heat time can only reduce an effect of thermal conduction (cold
sink effect of the cold core) that in turn could have a thermal impact and not an electro­
magnetic one.
Below the Curie point, the situation is quite different since power density does have
a measurable impact on electromagnetics owing to the impact of μ r. With lowering
power density, the field intensity of the coil is reduced, which, in turn, increases μ r (see
Figure 3.8a) that decreases δ in turn.
Your second statement is correct. If time is kept as a constant, then an increase in power
density results in a corresponding increase in applied energy (assuming the same fre­
quency is used) and, therefore, deeper case depths. However, in this case, an increase in
the case depth will be associated with an increase in the surface temperature. In some
cases, it is ok, but in other cases, severe surface heat surplus could lead to grain coarsening,
excessive distortion, grain boundary liquation (incipient melting), and cracking.
In some charts, an increase in power density results in reduction of the hardness case
depth. It might look strange and contradict suggestions indicated by alternative charts
that are contained in other publications. However, it should be noted that those charts
were developed under a condition of having the same surface temperature while applying
different power densities. Results might be just an opposite if it will be allowed to have
noticeably higher surface temperatures with power density increase.
Graphs that you are referring to indicate the following:

a. If you need to increase the case depth, then it is advantageous to use a lower fre­
quency (deeper heating effect) or a lower power density (but heat time must be
longer in this case). Both factors will help minimize the surface heat surplus and
obtain the required deeper case depth.
b. If the frequency is fixed, then the combination of increased heating time and lower
power density should be used to obtain deeper hardness case depths.
Heat Treatment by Induction 451

Over the years, the industry has accumulated numerous charts to help have a rough
estimation of basic process parameters. Some of those charts, diagrams, and rules of
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

thumb were published in Refs. [4–16,196] and were widely used in the 1970s up to
the 1980s. Though appearing very convenient, those ballpark estimation techniques
should be used with great caution because they were obtained for a certain steel grade
and under particular conditions (including prior microstructure, quenching, process
speed, classical geometries, etc.), usually after running a few experiments on straight
solid cylinders utilizing solenoid-type coils. Therefore, results may be different under
different conditions.
It is imperative to be aware that in many applications (especially in the cases of heat
treating irregular-shaped workpieces), erroneous and inadequate results can be obtained
when such simplified estimation techniques are used. Rather than using estimation tech­
niques with many restrictions and disputable accuracy, it is advisable to use available and
highly effective numerical computer modeling.

4.10.11 Case Study #11. Induction Tempering and Reverse


Residual Stress Distribution
Question: “We are doing single-shot hardening and tempering of carbon steel shafts.
Recently, a customer has requested to slightly reduce the surface hardness to be on the
middle of the specification. Our thoughts were that the easiest and risk-free way to drop
1.5–2 HRC points on a surface-hardened shaft is to increase power for tempering, which
would correspondingly increase tempering temperature and make steel softer. Suddenly,
we were surprised to notice that approximately 5% of all our hardened shafts had small
surface cracks. What causes this cracking and how could we avoid this problem?”
Answer: According to a conventional theory of tempering [16,28–30,33,151–155,383], it
might be expected that a stress-relieving process takes place during induction tempering
as a result of decomposition of fresh martensite by partial loss of tetragonality in mar­
tensite, precipitation of transition carbides (i.e., ε-iron [epsilon] carbides Fe2.4C and η-iron
[eta] carbides Fe2C), replacement of transition carbides by cementite (Fe3C), and formation
of tempered martensitic structures. This leads to a reduction in the magnitude of both
compressive and tensile stresses and the creation of more ductile, less brittle, and stress-
relieved structures. It is exactly what happens during conventional tempering of plain
carbon steels using furnaces/ovens.
As expected, under normal surface hardening conditions, the useful compressive resid­
ual stresses formed within as-quenched hardness case depths. However, compressive
residual stresses are balanced by tensile stresses. The maximum of tensile residual stresses
that is typically located in the hardened–unhardened transition region (Figure 4.144) is a
potential danger zone, and this is where subsurface cracks may initiate. Therefore, one of
the important “duties” of tempering is not only to reduce subsurface tensile stresses but
also to shift the maximum of these stresses toward the core away from the surface and
applied tensile stresses while not reversing compressive surface stresses within the case
depth.
Sufficiently mild induction tempering utilizing low power densities and adequately long
cycle times yields results that are very similar to conventional oven/furnace tempering.
However, under certain conditions, the heat intensity used in induction tempering (e.g., by
applying excessive power densities) may reverse residual stress distribution, and instead of
expected compressive residual stresses within the case depth, there may be a localized fine
surface layer where tensile residual stress of a considerable magnitude may occur.
452 Handbook of Induction Heating

Those undesirable surface tensile stresses can initiate crack formation owing to the ever-
present macroscopic and microscopic heterogeneities, structural features, and various
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

types of stress raisers (including but not limited to chemical and microstructural segrega­
tions, micronotches, burrs, scratches, chattering, etc.). Figure 4.212 illustrates a variation
of residual stresses along the radius of the induction-hardened solid cylinder applying
different induction tempering recipes.
In some cases, this phenomenon of reverse stress distribution is accompanied by a local­
ized reduction of hardness and is known in the theory of heat treating as the “reverse
hardness” phenomenon. However, in other cases, reverse stress distribution might not be
accompanied by a “reverse hardness” and a localized hardness reduction at the surface is
not observed.
Most likely, your temper time is too short but your power density increase was drastic
enough to reverse residual surface stresses during induction tempering. Therefore, instead
of simply increasing coil power for tempering, it would be better to increase time of tem­
pering, which will also result in an increase in tempering temperature and reduce surface
hardness. Of course, cycle time will be slightly increased. Multipulse induction tempering
might also be beneficial in this case (see Section 4.6.3). As an alternative, instead of modify­
ing temper cycle, you can also try to reduce quench severity by increasing the temperature
of the quenchant or its concentration.
Keep in mind that residual stress distributions are not nearly as simple when deal­
ing with complex geometries as they are in a plain cylinder. Geometrical discontinui­
ties including holes, fillets, sharp edges, shoulders, slots, keyways, undercuts, and other
discontinuities and stress risers can alter the induction tempering conditions, leading to
localized hardness and stress variations.
It is imperative to understand that, in some cases, tempering represents a combination
of tempering using additional power of the heat cycle(s) and self-tempering that utilizes
residual heat accumulated during the heating cycle for austenizing (see Section 4.6.2). The
use of both self-tempering and induction power tempering utilizing low power densities

Case depth Unaffected region

Improper induction
Proper induction
tempering cycle
tempering cycle
Tensile
Residual stresses

Radius
Compression

As-quenched

Transition
area

Surface Core

FIGURE 4.212
Variation of residual stresses along the radius of the induction-hardened solid cylinder applying different
induction tempering recipes.
Heat Treatment by Induction 453

allows a more uniform heat distribution for tempering to be produced. Thus, an additional
option to fix your cracking problem is to reduce quench time, allowing greater accumula­
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

tion of residual heat and applying self-tempering.

4.10.12 Case Study #12. Soft Spots after Grinding


Question: “We experience getting soft spots after grinding of cam lobes of our induction-
hardened camshafts and not on bearings. Any suggestions?”
Answer: Complex processes take place during surface grinding. Insufficient cooling
and excessive heat generation attributed to inappropriate grinding conditions/protocols
can produce thermal damage such as rehardening burns, temper burns, or a deviation/
reversal of residual surface stresses. There are a number of parameters of the grinding
process that can go wrong and need to be optimized. This includes excessive cutting force/
speed, dirty or dull wheel, and others, which can negatively affect the wear resistance of
the camshaft’s working surfaces by altering the microstructure, causing low hardness
readings (soft readings), developing grinding cracks, and decreasing or even reversing
beneficial compressive residual stresses within the case depth. There are a number of
excellent research work conducted by different researchers to study grinding conditions
on induction-hardened components. This includes the work by Professor Janez Grum
[176,376]. Therefore, we suggest reviewing those publications and also contacting a manu­
facturer of your grinding machine.
Since soft spots occur only on cam lobes, circumferential as-quenched hardness dis­
tribution might be worth checking. This can help ensure that differences in coil-to-lobe
coupling in the nose area versus the base circle region (assuming that you are using
a conventional system) did not contribute to the hardness variation, especially if a
conventional camshaft hardening process is used for both induction hardening and
tempering.
It would be highly beneficial to minimize the amount of grinding stock and produce
camshafts that are as straight as possible. Both factors are associated with the ability to
produce a uniform hardness pattern along the circumference of the cam lobes and bear­
ings and to minimize the total amount of metal being heated to elevated temperatures.
Therefore, consider using the CamPro machine utilizing patented SHarP-C Technology
instead of conventional processes. This technology eliminates many drawbacks of con­
ventional camshaft IHT, dramatically improves straightness characteristics, and has been
discussed in Section 4.9.5.

4.10.13 Case Study #13. Safety Warning for People


with Pacemakers, Metallic Implants, etc.
Question: “We have an induction process and need to include physical barriers and/
or a safety warning for people with medical devices such as metallic implants, cardiac
pacemakers, and so on to stay away X number of feet/meters, providing visible guarding.
What is the required distance for people with medical devices to stay away from induc­
tion heating coils? Induction heating is a complicated process, but it can be modeled and
controlled if you know what you are doing….. Thanks for any help you can provide on
stand-off distances.”
Answer: Many studies have been conducted to evaluate the direct and indirect effects of
EMF exposure on health, passive and active medical implants, hypersensitivity, and so on.
454 Handbook of Induction Heating

Several national and international organizations including the Institute of Electrical


and Electronic Engineers (IEEE), the International Radio Protection Association (IRPA),
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

the World Health Organization (WHO), the Occupational Safety & Health Administration
(OSHA), and others have developed awareness programs regarding nonionizing radiation
and evaluation of the health risks associated with external field exposure when using any
electromagnetic devices (such as cell phones, household electrical appliances, computer
monitors, transmitting antennas, microwave ovens, and others). A large body of scientific
research exists, and a number of international standards, guidelines, and regulations have
been put into effect. For example, in the United States, these include but are not limited to
the following two standards:

• IEEE: C95.6 Safety levels with respect to human exposure to EMFs, 0 to 3 kHz
• IEEE: C95.1 Safety levels with respect to human exposure to EMFs, 3 kHz to 300 GHz

All manufacturers of electromagnetic devices (including induction heating manufac­


turers) must comply with international/national standards and regulations related to
controlling external exposure to EMF in the workplace.
There is no universal guideline for so-called safe stand-off distances for employees with
medical devices or people without them, because such safe distances are application-
specific and greatly depend on coil design details, applied power and frequency, the pres­
ence of other electromagnetic devices in coil surroundings, and other factors.
A magnitude of external magnetic field can be either computer modeled or measured.
For example, at Inductoheat, we use both techniques. In some cases, induction coil is not
the main source of EMF exposure; other electromagnetic sources (e.g., transformers, bus
bars, electromagnetic lifting devices, etc.) might contribute an even greater degree to its
magnitude of EMF.
There are a number of devices available on the market to measure an EMF exposure at
the workplace. They range in applied frequency, design, and so on. On the basis of our
experience, three-axis measuring devices provide acceptable accuracy if properly used
and calibrated. For example, for applications where it is suitable, we have been using
Holaday 3-axis VLF Magnetic Field Meter HI-3637 for a number of years (Figure 4.213).
This device is capable of providing isotropic measurements of magnetic field distribution.
The instrument is designed to conform to the requirements issued in the IEEE protocols
and guidelines for suitable frequencies.

FIGURE 4.213
Holaday 3-axis VLF Magnetic Field Meter HI-3637. This device is often used for isotropic measurements of
magnetic field distribution around an induction system.
Heat Treatment by Induction 455

Here, there are general guidelines for monitoring and supervision of EMF exposure.
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

Running conditions of the induction system: The machine should be running with the
maximum running parameters for output kilowatts or the maximum allowable
running current/voltage to the coil.
Measurement specifics: The machine envelope should be drawn and measurements
should be taken at the operator station (or where other people may be present)
and at a number of points around the perimeter of the machine at distances
from the machine of 0.25, 0.5, 1, and 1.5 m (whatever is the most appropriate)
and different heights from the floor. The data should be reviewed and a report
should be issued to indicate whether the measured values comply with safety
requirements. All meter readings should be taken and data should be properly
documented, including setup parameters for capacitor and transformer set­
tings. If necessary, measurements should be made to calculate the coil running
current for the specified meter readings.

Over the years, manufacturers of induction machinery have developed a number of


ways to minimize external magnetic field exposure in case if it exceeds maximum permis­
sible levels. This includes but is not limited to a number of patented designs and propri­
etary techniques including passive and active magnetic shields, magnetic shunts, Faraday
rings, just to name a few.
We also suggest to review an excellent article “Control of Professional Magnetic
Field Exposure—International Standards and Regulations” written by Loris Koeing,
SFinduction, France on this subject (see Ref. [406]).

4.10.14 Case Study #14. Re-Hardening of Already Induction Hardened Parts


Question: “We cast camshafts with FCD 700 grade ductile iron (JIS), which is a predomi­
nantly pearlitic grade. Our process requires us to induction harden and temper the cam
lobes. This brings me to the general question… If we process them incorrectly and end up
with a case depth that is too shallow… and we have oven tempered them. Can we correct
the case depth by just re-running them thru the induction process again and re-tempering
them to get the targeted case depth? What would be concerns with this thinking?”
Answer: On numerous occasions, we have been contacted with similar questions related
to rehardening of improperly (insufficiently) hardened parts made of steels, cast irons, or
P/M components. In many cases, there were no obvious problems with this practice (par­
ticularly if parts were underheated, producing very shallow case depth) and rehardened
parts lasted as long as normally hardened parts, passing tests with flying rates. There are
many applications where induction heating is successfully used on reheating carburized
parts (but these are different applications having its own specifics).
Formally speaking, induction rehardening of already induction-hardened parts is not
considered to be a good metallurgical practice. Here are some of the potential concerns for
rehardening of already induction-hardened parts:

1. After rehardening, there is a danger of crack development. It is particularly so,


if the part has been hardened to a considerable case depth (e.g., if effective case
depths are greater than 3 mm [1/8 in.]). Of course, if the required case depth is
5  mm, but for whatever reason parts were improperly induction hardened to
1/64  in. (0.4 mm) or 1/32 in. (0.8 mm) of effective case depth, then, most likely,
456 Handbook of Induction Heating

rehardening to the required effective case depth of 5 mm might not exhibit any
detrimental effects.
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

2. Grain coarsening might appear upon rehardening. There might be several reasons
for that. One such reason is related to a fact that the electromagnetic properties
of martensitic (as-quenched or Q&T) initial microstructures are different from
ferritic–​pearlitic structures. If process settings (process recipe) are kept identi­
cal in both cases (hardening of a “green” part vs. rehardening), then components
with martensitic prior structure will most likely be heated to noticeably higher
temperatures compared to parts with ferrite–pearlite initial structures. Therefore,
most likely, it will be necessary to properly modify the process recipe/protocol for
reheating.
3. Residual stresses of rehardened parts might be different to some degree (depend­
ing on geometry, hardness pattern, and other factors) compared to components
that were induction hardened only once. It is not necessarily true that the residual
surface compressive stresses of the rehardened part will be lower compared to
stresses of the part that was hardened once. Actually, it might be the other way
around, but the matter of the fact is that the residual stress distribution in both
cases might be noticeably different. Besides that, the shape/size distortion charac­
teristics might worsen after rehardening.
4. In order to salvage expensive components, some practitioners rehardened parts
more than twice, which is associated with a danger of crack development. Besides,
after multiple rehardening, there might be some issues with carbides and nitrides
networking. This is a very complex issue and it is related to the material’s specifics
and heating recipes and, therefore, should be carefully analyzed. Normally, such
practices of multiple rehardening should be avoided.

These are at least four main factors that must be considered when discussing an issue
of making a decision whether to reharden parts or not. Again, we would like to reiter­
ate that, in some cases, rehardening did not produce measurable negative effects on part
performance. Of course, it does not look as fair deal to not be able to reharden parts that
upon rehardening look practically identical to those that have been properly hardened
and it is natural to try to salvage some parts instead of scraping them. However, a decision
should be carefully assessed on a case-to-case basis, and still “formally speaking,” induc­
tion rehardening of previously improperly hardened parts is not considered to be a good
metallurgical practice.

4.10.15 Case Study #15. Etched vs. Unetched Samples in Crack Detection


Question: “We are routinely Magnafluxing our surface-hardened parts made of cast iron
to make sure that there is no cracking. If cracks occur, we tried to look using a micro­
scope to see the details. We noticed that it is relatively easy to find surface cracks on our
Magnaflux machine, but have difficulties seeing them using an optical microscope. We use
our standard sample preparation practice: sectioning, polishing, etching (4% Nital), and so
on. We also heard that Picral etchant might better reveals cracks. Any suggestions?”
Answer: Etching is very beneficial to evaluate microstructural subtleties. However, it is
easier to reveal crack on an unetched sample. Figure 4.214 compares images (×100) of Nital-
etched (left) versus unetched (right) samples made of nodular cast iron. Picric acid etchant
(Picral) helps to better reveal certain transformation products. However, it exhibits greater
Heat Treatment by Induction 457
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

Depth
Length: 0.744 mm

(a) (b)

FIGURE 4.214
Comparison of images of etched (a) versus unetched (b) samples of a part made of nodular cast iron (×100). It is
easier to observe the crack on unetched samples.

safety and hazard issues (including explosive hazard in dry condition, low flash point,
and autoignition temperature). Thus, your personnel must have proper training in using
this etchant. Again, in our opinion, it is easier to observe the crack on unetched samples.

4.10.16 Case Study #16. Induction Heating of Ultra-Thin Conductive Coatings


Question: “I do have a general question. Is it possible to inductively heat very thin nonfer­
rous coating that is applied to flat, electrically nonconductive base material? The size of
the base material is 20 mm × 200 mm. Coating thickness is about 15 μm, and it is electric
conductive, having an electrical resistivity of approximately 8–10 times greater than stain­
less steel. The target temperature is about 350°C–380°C.”
Answer: Since electrical resistivity of coating (that is applied on a nonelectrically con­
ductive base material) is quite high (8–10 greater than ρ of stainless steel) and its thickness
is only approximately 15 μm, there is no chance to heat it using traditional solenoid induc­
tors because of severe eddy current cancellation. Actually, a coating will be simply trans­
parent to EMF. The only chance you might have is to use transverse flux (TFX) inductors.
Taking into consideration that the sample is flat, you can probably try to use a scanner to
conduct a test applying a hairpin inductor and positioning ferrites on opposite sides of the
workpiece. Usually, TFX inductors (which are discussed in Section 6.9.2.2) apply relatively
low frequencies, but in your case and because of such a thin thickness of coating, you must
use frequencies of RF range. As you know, the major drawback of TFX heaters is related to
severe temperature inconsistency at edge areas, which is associated with poor repeatabil­
ity and high sensitivity to a film-to-inductor positioning. Therefore, you can probably do a
simple R&D project to see if you will be able to heat a sample by positioning two hairpin
inductors on opposite sides of the sample and applying scan heating or, as an option, using
a single hairpin inductor and ferrites positioned on its opposite side.

4.10.17 Case Study #17. Distortion of Shafts in Induction Hardening


Question: “We are experiencing an issue that has had me confused for about a month
now, and I’m wondering if you can provide some insight as to what might be causing it.
We do scanning induction hardening of axle shafts, and we recently took on a new pro­
vider of steel. With this new provider, we are experiencing problems after induction case
458 Handbook of Induction Heating

hardening. When the process is finished, we are typically seeing 0.350–0.500 in. (9–12 mm)
TIR on our parts. They are warping very bad. If we switch back to the old steel provider
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

and run another set with the same program, the TIR is somewhere between 0.030 and 0.075
in. (between 0.75 and 1.9 mm), which is normal. The new provider assures us that the steel
meets our required specifications, and our plant metallurgist has confirmed that. We have
even tried different suppliers, and theirs all came back with good TIR. Parts are centered
in the coils, and we have tried this on several different machines to verify. The same issue
with each one. We are at the bottom of our case depth requirements, so there is not a lot of
energy we can take out of the part. Is there anything you suggest we take a look at, either
in the steel, or in our process?”
Answer: It appears that your description indicates that the problem of excessive dis­
tortion is related to steel and not to a process recipe. It is so, because you can make the
problem of excessive distortion “to appear” or “to disappear” depending on whether
you are using steel from a particular steel supplier or from others. This fact is a clear
proof that the problem of excessive distortion is somehow related to steel and not a
process of hardening. Unfortunately, there are too many factors related to steel specifics
that might affect the shaft’s distortion: steelmaking subtleties, steel cleanliness, chemical
composition deviations, residuals, grain size, microstructural and chemical heterogene­
ity, initial stresses and their distribution before induction scan hardening, just to name
a few factors. Deviations in prior microstructure of steel can also affect electrothermal
physical properties of steel (including electrical resistivity, thermal conductivity, and
others), which will also affect heating conditions (including but not limited to achieved
temperatures) while using the same process recipe to harden shafts made from different
steel suppliers.
You stated that your plant metallurgist has confirmed that steel composition is accord­
ing to a specification. However, something must be different to cause an increase of a
shaft’s warpage by almost 10-fold, resulting in 0.350–0.500 in. TIR instead of 0.030–0.075 in.
(0.75–1.9 mm) with the original steel supplier. It would be interesting to know your plant
metallurgist’s judgment regarding the fact that the problem “appears” or “disappears”
depending on whether steel is supplied by that particular provider compared to several
alternative suppliers. I suggest you to send samples of both steel samples to independent
metallurgical laboratory for evaluation of the chemical compositions of both steels (old
and new), including assessing all residual elements, segregation characteristics, banding,
grain size, and so on to try to have a clear understanding of what are you dealing with
when hardening shafts made of that particular steel. This will be an added expense, but
there is no doubt that having such a bad distortion is also a huge expense, and the sooner
the cause of an excessive distortion is eliminated, the better it will be.
In addition, it might be also beneficial to put two unhardened shafts (made of “old” steel
and “new” steel) into high-temperature tempering furnace/oven for a few hours at tem­
peratures of at least 650°C–670°C and then assess the difference in the shape distortion
after high-temperature tempering of both unhardened samples. This will give you an indi­
cation regarding the magnitude of internal stresses before induction hardening. Internal
stresses are relieved during induction scan hardening. Therefore, different magnitudes
of internal stresses before hardening could cause differences in distortion characteristics.
Also, please check geometrical tolerances in both cases when shafts are made of “good”
steel and “bad” steel.
Again, there could be a number of different steel-related factors affecting a distortion
and these are some ideas for you and your plant metallurgist to evaluate in trying to find
the root cause problem of excessive distortion of your shafts.
Heat Treatment by Induction 459

4.10.18 Case Study #18. Uniform Heating of Steel Plates


Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

Question: “We need to heat a flat component (36 mm thick × 88 mm wide × 360 mm long)
made of SAE 1035 steel from room temperature to about 960°C ± 10°C. We have an old
300-kW/6-kHz inverter from a previous scan hardening job and would like to use it to
statically heat this plate. Production rate is not so critical for us and heat time can be as
long as 3–4 min if needed. Can you please comment on this?”
Answer: Induction heating of rectangular workpieces such as plates has several unique
features compared to heating cylinders. One of them is related to electromagnetic trans­
verse edge effect (Figure 3.34). Among other factors, a selection of optimum frequency is a
reasonable compromise between providing needed temperature uniformity: “end to end”
(controlled by an electromagnetic end effect), “across width” (controlled by an electromag­
netic edge effect), and “surface to core” (controlled by a skin effect). These effects have been
discussed in corresponding sections of Chapter 3. In your attempt to heat this plate using
the old scan hardening system, you will most likely have difficulties in satisfying your
process requirements. There are several reasons for this.

1. Scan hardening applications typically apply single-turn, two-turn, or three-turn


coils. This means having relatively low coil voltage. In order to statically heat your
plate, you will be required to use a multiturn coil, resulting in noticeably different
coil impedance and requiring noticeably higher voltage. Therefore, you may face
some difficulties with load-matching coil and an inverter.
2. Even if you can solve a load-matching problem, the frequency of 6 kHz will not be
able to provide the needed ±10°C heat uniformity. Figure 4.215a shows the results
of FEA computer modeling of temperature distribution in a top right quarter of
statically heated steel plate using 6 kHz. A comparison of temperatures of criti­
cal areas reveals that after 240 s of heating, the coldest temperature will be in the
plate core (Node #3) being 921.4°C. The hottest area is represented by Node #4
and its temperature is 980.9°C. Therefore, the temperature uniformity achieved
is approximately ±30°C. It might be reasonable to assume that a lower frequency,

Node #1 Node #2

Plate geometry:
Node #3 Node #4 0.036 m (thick) × 0.088 m (wide) × 0.24 m (long)

6 kHz; 240 s 3 kHz; 240 s


0.018 0.018
0.016 1 2 0.016 1 2
Thickness, m
Thickness, m

0.014 0.014
0.012 0.012
0.01 0.01
0.008 0.008
0.006 0.006
0.004 0.004
0.002 3 4 0.002 3 4
0 0.02 0.04 0 0.02 0.04
Width, m Width, m

N1 = 950.6; N2 = 953.9; N3 = 921.4; N4 = 980.9 N1 = 985.1; N2 = 948.3; N3 = 958.0; N4 = 982.8


Temperature, ºC Temperature, ºC
(a) (b)

FIGURE 4.215
Results of FEA computer modeling of temperature distribution in a top right quarter of statically heated steel
plate using 6 kHz (a) versus 3 kHz (b).
460 Handbook of Induction Heating

producing more in-depth heating effect, would improve “surface-to-core” heat


uniformity. Figure 4.215a shows the use of 3 kHz. 2-D corner (Node #2) has the
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

coldest temperature of 948.3°C. Surface temperature in the middle of the plate


width (Node #1) is the hottest area, being approximately 985.1°C. Uniformity of
heating is approximately ±20°C. Underheating of a corner occurs because the most
eddy current close their loops earlier without reaching a corner. Keep in mind that
the heat deficit in the corner will further worsen during plate transportation to a
subsequent operation or if heating time is increased.

Therefore, we suggest reviewing the results of FEA computer modeling with your cus­
tomer and finding out if the required temperature uniformity of ±10°C can be relaxed.
Otherwise, a dual-frequency concept must be applied. An explanation of factors respon­
sible for temperature nonuniformity when heating rectangular bodies and the subtleties
of using a dual-frequency design are provided in Section 6.8.

4.10.19 Case Study #19. Circumferential Hardness


Variation Using Single-Shot Inductors
Question: “For a number of years, we scan hardened relatively straight shafts using a
30-kHz machine. Our new component has about 4.8 mm radial step (sharp shoulder) that
was melted when we tried to scan harden. We heard that a single-shot inductor could
reduce overheating of shoulder and still provide sufficient hardening of the fillet. We asked
a local coil builder to make a single-shot coil for us. Corner overheating is reduced, but to
our surprise, now we have a 4–6 HRC surface hardness variation around the circumfer­
ence of the as-quenched component when applying single-shot hardening. We have not
seen such hardness variation around the shaft surface after scan hardening. What might
cause it and how should we fix it?”
Answer: Appropriate inductor/quench design and suitable process recipe should not
produce such a wide hardness variation along the circumference of the shaft. A number of
possible causes exist, with two of the most common discussed here.

1.
Heat-related factor. With scan coils, induced eddy current flow is circumferential
producing very uniform temperature distribution at any instant of the heat time.
Only minor heat variation occurs in a fish-tail area or flux fringing region of coil
terminals (see Section 4.2.8.3). Part rotation makes this impact negligible (prac­
tically speaking), normally providing very uniform temperature distribution
(assuming an absence of slots, keyways, holes, and similar geometrical discon­
tinuities). Unlike scanning inductors, majority of single-shot inductors predomi­
nantly produce an axial eddy current flow (Figures 4.76 and 4.77) rather than a
circumferential one. Therefore, if the heating time is relatively short, the coil heat­
ing face is narrow, and rotation speed is insufficient, then there might be a notice­
able nonuniform circumferential temperature pattern at a time when quenching is
applied. Solution: An increase in rotation speed during a single-shot heating might
solve this problem. The use of serpentine-type inductor is another option.
2.
Quench-related factor. MIQ coils are often used with scan hardening, normally pro­
viding sufficiently uniform circumferential quenching. Quench rings/barrels/­
followers used in scan hardening also produce reasonably uniform quenching.
With a single-shot inductor, there might be obstructions for providing 360° access
Heat Treatment by Induction 461

for spray quenching. In some cases, quenching is done from one side, covering
only 90°–120° of the austenized surface. This could result in considerable circum­
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

ferential hardness deviation. Solution: Redesign the quenching device, making an


attempt to “cover” as much surface as possible. In addition, try to change the rota­
tion speed during quenching (e.g., a 25% increase and then a 25% decrease) and
compare results. Both insufficient rotation and excessive rotation could result in
circumferential quench nonuniformity (for different reasons though).

4.10.20 Case Study #20. Non-Martensitic Structures in Induction Hardened Parts


Statement: In the past, some heat treaters have been confused by observing the presence
of nonmartensitic structures in induction-hardened parts where, based on conventional
metallurgical practice and widely used heat treating diagrams and charts, a fully martens­
itic structure (practically speaking) is expected to be obtained [247]. An example follows:

• Fact 1: After induction heating and quenching of a steel component, there was no
evidence of ferrite/pearlite networking. This led the heat treater to conclude that
austenitization of the steel had been completed.
• Fact 2: Metallographic examination revealed no evidence of excessive amount
of RA, severe grain coarsening, scale formation, or decarburization, which sup­
ported the idea that the final temperature had not been too high. This was verified
by measuring the surface temperature of the component at the end of the heat
cycle.
• Fact 3: Based on data provided by the steel’s manufacturer, the Mf temperature was
above the temperature of the quenchant. Quench time, flow, and concentration
were sufficient to cool the austenized component below the Mf temperature and
miss the “nose” of the CCT curve.

Analysis: These conditions explain why the heat treater expected a fully martensitic
structure (practically speaking) after quenching. However, metallography revealed a
small amount of nonmartensitic transformation products (bainite) in an overwhelmingly
lath martensite structure. An improper assumption had been made: The absence of ferrite/
pearlite networking in martensite does not mean that homogeneous austenite had been
formed. One characteristic of heterogeneous austenite is a nonuniform distribution of
carbon.
After rapid heating, ferrite/pearlite networking may not exist, but there might still be
an inhomogeneous distribution of carbon in the austenite. Thus, there could be some
localized carbon-enriched areas and regions where the carbon concentration is noticeably
reduced. Since both the Ms and Mf temperatures depend on carbon content, the high- and
low-carbon areas of a component having heterogeneous austenite will have different CCT
curves upon quenching. As a result, local subareas have different critical cooling rates, and
other, nonmartensitic transformation products may form in the predominantly martens­
itic structure. The existence of appreciable size regions of martensite having measurably
different carbon concentration may also cause spotted hardness readings. Besides that, the
heterogeneous nature of austenite formed as a result of rapid induction heating might also
be associated with various concentration gradients of chemical elements. Keep in mind
that those complex carbides (e.g., involving substitutional chemical elements) often require
noticeably longer times and higher temperatures to be dissolved than Fe3C.
462 Handbook of Induction Heating

It should also be noted that concentration gradients influence not only the maximum
achievable surface hardness and case depth but also the transition zone.
Downloaded By: 10.3.98.93 At: 07:56 17 Sep 2018; For: 9781315117485, chapter4, 10.1201/9781315117485-4

Solution: A longer time at austenitizing temperatures and higher temperatures might be


required to reduce the heterogeneity of the austenite and fix this problem. It should also be
noted that the presence of a network of fine undissolved carbides and carbonitrides might
still be present in predominantly martensitic structure.
The questions and answers provided above represent only a very small fraction of hundreds of
questions received by the authors since the publication of the first edition of Handbook of Induction
Heating. Space does not permit reviewing all questions. Many case studies have been discussed
earlier in this chapter and answers have been provided. If more questions arise, please contact the
authors directly.

Potrebbero piacerti anche