Discover millions of ebooks, audiobooks, and so much more with a free trial

Only $11.99/month after trial. Cancel anytime.

Cutting Data for Turning of Steel
Cutting Data for Turning of Steel
Cutting Data for Turning of Steel
Ebook502 pages2 hours

Cutting Data for Turning of Steel

Rating: 0 out of 5 stars

()

Read preview

About this ebook

Mechanical Properties of Steel

  • Hardness
  • Carbon Steels
  • Alloy Steels
  • Stainless Steels
  • Tool Steels

Cutting Tools Materials

  • High Speed Steels
  • Cemented Carbides
  • Cermets
  • Ceramics  
  • Polycrystalline Cubic Boron Nitride (PCBN)

Machining Recommendations

  • Depth of Cut and Feed Rate
  • Cutting Speeds for Carbon Steels
  • Cutting Speeds for Alloy Steels
  • Cutting Speeds for Stainless Steels
  • Cutting Speeds for Tool Steels

Machining Power

  • Metal Removal Rate
  • Unit Power and Power Constant
  • Calculating Required Machining Power

Appendix 1: Hardness Conversion

Appendix 2: Carbon Steels

Appendix 3: Alloy Steels

Appendix 4: Stainless Steels

Appendix 5: Tool Steels

Machining is one of the most important manufacturing processes, which remove unwanted material in the form of chips from a workpiece. Material removal operations are among the most expensive; in the U.S. alone, more than $100 billion were spent on machining in 1999. These high costs put tremendous economic pressures on production managers and engineers as they struggle to find ways to increase productivity.

Machining recommendations provided in this book cover turning since it allows removing more material per unit of time and consuming more power at the roughing operations than end milling, boring or drilling. Machining recommendations relate to cutting speeds, feed rates, and depth of cuts. Such recommendations depend on the workpiece material properties and the cutting tool material. Workpiece materials described in this book are the most commonly used grades of carbon, alloy, stainless, tool, and maraging steels. Cutting tool materials are cemented carbides, cermets, and ceramics.

LanguageEnglish
Release dateOct 31, 2008
ISBN9780831191757
Cutting Data for Turning of Steel

Related to Cutting Data for Turning of Steel

Related ebooks

Industrial Design For You

View More

Related articles

Related categories

Reviews for Cutting Data for Turning of Steel

Rating: 0 out of 5 stars
0 ratings

0 ratings0 reviews

What did you think?

Tap to rate

Review must be at least 10 words

    Book preview

    Cutting Data for Turning of Steel - Edmund Isakov

    Preface

    This book provides machining recommendations for turning, which is one of the major metalcutting processes. Rough turning removes more chips and requires more machining power than milling, drilling, and boring. Machining recommendations relate to cutting speeds, feed rates, and depth of cuts. Such recommendations depend on the work-piece materials, their properties, and the type of the cutting tool materials. Workpiece materials described in this book are the most commonly used grades of carbon, alloy, stainless, and tool steels. Cutting tool materials are high-speed steels, cemented carbides, cermets, ceramics, and polycrystalline cubic boron nitride (PCBN).

    In 1966, Metcut Research Associates Inc. published its first edition of Machining Data Handbook, which became the most important source of machining data for the metalcutting industry. This Handbook had two more editions: the second edition in 1972 and the third edition in 1980. The third edition had several printings; its sixth printing includes Volumes 1 and 2, which were published in 1987 (Ref 1, Ref 2).

    For the last two decades, the metalcutting industry has undergone dynamic changes. New cutting tool materials have been developed. In the year 2000, more than 70% of all cemented carbide inserts purchased were coated by chemical vapor deposition (CVD) and physical vapor deposition (PVD). Cermets, ceramics, and polycrystalline cubic boron nitride cutting tools are becoming more popular when machining hardened steels.

    Thus, the machining recommendations of the 1980s are considered as starting cutting parameters that require certain adjustments in accordance with today’s metalcutting technology. The new book Cutting Data for Turning of Steel reflects such adjustments.

    In addition to Machining Data Handbook (Ref 1, Ref 2) and Machining, Metals Handbook (Ref 3), the new book Cutting Data for Turning of Steel will be useful to manufacturing engineers and managers, machine shop supervisors, machine tool operators, NC programmers, and cutting tool engineers and designers.

    Introduction

    Machining is one of the most important manufacturing processes, which remove unwanted material in the form of chips from a workpiece. Material removal operations are among the most expensive; in the United States alone, more than $100 billion were spent on machining in 1999. These high costs put tremendous economic pressures on production managers and engineers as they struggle to find ways to increase productivity (Ref 3, p.v).

    Machining processes are performed on a wide variety of machine tools. The primary chip formation processes are turning, milling, drilling, shaping, and abrasive machining. The majority of industrial applications of machining are in metals. In spite of their complexity, metal cutting processes are widespread in the industrial world. Metal cutting processes consist of independent or input variables and dependent variables. Independent (input) variables are:

    •Workpiece materials

    •Cutting tool materials and geometry

    •Cutting parameters

    Machine tool operators and manufacturing engineers have direct control over the independent variables and select them when setting up the machining process.

    Dependent variables are:

    •Cutting forces

    •Power consumption

    •Surface finish

    •Cutting tool conditions

    Dependent variables are determined by the machining process based on the prior selection of the independent variables. Therefore, manufacturing personnel do not have direct control over dependent variables.

    Workpiece materials (one of the independent variables) have been described by Edmund Isakov in the book published in 2000 (Ref 4). Dependent variables: Cutting forces and Power consumption were described in the book and software by Edmund Isakov published in 2004 (Ref 5) and in 2005 (Ref 6) respectively.

    The purpose of this book is to provide recommendations for selecting Machining parameters in relationship with Cutting tool materials and Workpiece materials.

    This book, which complements the author’s previous publications, will be useful for setting maximum productivity from machine tools when turning of steels.

    Chapter 1

    Mechanical Properties of Steel

    Steels represent the most widely-used category of metallic material, primary because they can be manufactured relatively inexpensively in large quantities to very precise specifications (Ref 7, p.140).

    Hardness, tensile strength, and machinability of steels are the major mechanical properties described in this chapter. These properties are taken into consideration when selecting machining parameters.

    Classification of steels by commercial name or application is the most common system consisting of the following groups: carbon steels, alloy steels, stainless steels, and tool steels (Ref 8, pp.1, 73, 243, and 431).

    1.1. Hardness

    Hardness is a measure of the resistance of a material to surface indentation or abrasion. There is no absolute scale for hardness. In order to express hardness quantitatively, each type of test has its own scale that defines hardness. Indentation hardness obtained through static methods is measured by Brinell, Rockwell, Vickers, and Knoop tests. Hardness without indentation is obtained by the dynamic (rebound) method. The typical dynamic method is the Scleroscope hardness test, also known as the Shore method.

    Traditionally, selections of machining parameters for steel have been based on Brinell or Rockwell hardness of the work materials, the cutting tool materials (high-speed steels, cemented carbides, cermets ceramics, and cubic boron nitride), and cutting tool geometry (Ref 1 and 8).

    1.1.1. Brinell hardness

    Hardness of a wide variety of materials is determined by the Brinell test, which consists of applying a constant load, usually between 500 and 3000 kgf, for a specific time (10 to 30 seconds) using a 5- or 10-mm-diameter ball.

    Brinell hardness numbers (HB) of carbon and alloy steels in the annealed, normalized, and quenched-and-tempered conditions are determined by forcing a 10-mm-diameter ball made of hardened steel or tungsten carbide into the workpiece under the 3000-kgf load. The most accurate readings are between 81 HB and 444 HB when hardened steel ball indenters are used. Tungsten carbide ball indenters are used for hardness numbers between 444 HB and 627 HB (Ref 9, pp.111–113).

    The 500-kgf load is used for testing aluminum and copper alloys. However, the same load can be used for testing unhardened steels. Keeping in mind that selection of machining conditions depends on Brinell hardness numbers obtained under the 3000-kgf load, it is necessary to convert Brinell hardness at 500-kgf load into Brinell hardness at 3000-kgf load. Such conversions can be obtained from difficult-to-use, cumbersome tables of the equivalent hardness numbers (Ref 9, pp.109–113). To make hardness conversion friendly, the author used the linear regression method and developed the necessary equations. To obtain high accuracy in conversion, the hardness numbers at 500-kgf load (HB5) and corresponding hardness numbers at 3000-kgf load (HB3) were grouped in five tables, each of which contains various numbers of data points.

    Statistical treatment of data in each table produced linear regression equations with the correlation coefficients from 0.9962 to 0.9997. These coefficients indicate a strong linear relationship between Brinell hardness at 500-kgf load (HB5) and Brinell hardness at 3000-kgf load (HB3).

    The following five equations were developed (Appendix 1, Tables 1.1–1.5).

    Table 1.1

    : HB5 range is 89–99; HB3 range is 100–112 respectively. The equation for calculating Brinell hardness at 3000-kgf load (HBC) is

    Example of calculation:

    HBC = 1.138 × 89 − 1.052 = 100.230, rounding off gives: 100 (100 HB3)

    HBC = 1.138 × 99 − 1.052 = 111.610, rounding off gives: 112 (112 HB3)

    Table 1.2

    : HB5 range is 101–120; HB3 range is 114–137 respectively. The equation for calculating Brinell hardness at 3000-kgf load (HBC) is

    Example of calculation:

    HBC = 1.224 × 101 − 9.871 = 113.753, rounding off gives: 114 (114 HB3)

    HBC = 1.224 × 120 − 9.871 = 137.009, rounding off gives: 137 (137 HB3)

    Table 1.3

    : HB5 range is 122–140; HB3 range is 139–162 respectively. The equation for calculating Brinell hardness at 3000-kgf load (HBC) is

    Example of calculation:

    HBC = 1.311 × 122 − 21.047 = 138.895, rounding off gives: 139 (139 HB3)

    HBC = 1.311 × 140 − 21.047 = 162.493, rounding off gives: 162 (162 HB3)

    Table 1.4

    : HB5 range is 142–160; HB3 range is 165–190 respectively. The equation for calculating Brinell hardness at 3000-kgf load (HBC) is

    Example of calculation:

    HBC = 1.369 × 142 − 30.005 = 164.393, rounding off gives: 164 (165 HB3)

    HBC = 1.369 × 160 − 30.005 = 189.035, rounding off gives: 189 (190 HB3)

    Table 1.5

    : HB5 range is 163–189; HB3 range is 195–228 respectively. The equation for calculating Brinell hardness at 3000-kgf load (HBC) is

    Example of calculation:

    HBC = 1.283 × 163 − 14.218 = 194.911, rounding off gives: 195 (195 HB3)

    HBC = 1.283 × 189 − 14.218 = 228.269, rounding off gives: 228 (228 HB3)

    1.1.2. Rockwell hardness

    Rockwell hardness testing is applied to most metals and alloys and consists of twenty hardness scales. Rockwell scale B is used to measure hardness (HRB) of steel at the annealed condition. The load is 100 kgf applied to a -inch- (1.588 mm-) diameter ball indenter. The accurate readings are between 41 and 100 HRB (Ref 9, pp.111–113). Rockwell scale C is used to measure hardness (HRC) of heat-treated steels harder than 100 HRB. The major load is 150 kgf applied to a 120° spheroconical diamond indenter. The accurate readings are between 20 and 69 HRC (Ref 9,p.77).

    Rockwell B hardness numbers and Rockwell C hardness numbers should be converted into the equivalent Brinell hardness numbers measured at 3000-kgf load. The same tables for conversion, mentioned earlier (Ref 9, pp.109–113), can be used. However, linear regression equations developed by the author make such conversions handy and highly accurate. The author applied the same mathematical technique, which was described earlier.

    Rockwell B hardness numbers and corresponding Brinell hardness numbers measured at 3000-kgf load were grouped in six tables, each of which contains various numbers of data points.

    Statistical treatment of data in each table produced linear regression equations with the correlation coefficients from 0.9971–0.9997. These coefficients indicate a strong linear relationship between Rockwell B hardness numbers and Brinell hardness numbers measured at 3000-kgf load.

    The following six equations were developed (Appendix 1, Tables 1.6–1.11).

    Table 1.6

    : HRB range is 62.3–69.8; HB range is 105–121 respectively. The equation for calculating Brinell hardness at 3000-kgf load (HBC) is

    Example of calculation:

    HBC = 2.136 × 62.3 − 28.345 = 104.728, rounding off gives: 105 (HB = 105)

    HBC = 2.136 × 69.8 − 28.345 = 120.748, rounding off gives: 121 (HB = 121)

    Table 1.7

    : HRB range is 71.2–79.7; HB range is 124–146 respectively. The equation for calculating Brinell hardness at 3000-kgf load (HBC) is

    Example of calculation:

    HBC = 2.579 × 71.2 − 59.950 = 123.675, rounding off gives: 124 (HB = 124)

    HBC = 2.579 × 79.7 − 59.950 = 145.596, rounding off gives: 146 (HB = 146)

    Table 1.8

    : HRB range is 80.8–85.0; HB range is 149–163 respectively. The equation for calculating Brinell hardness at 3000-kgf load (HBC) is

    Example of calculation:

    HBC = 3.303 × 80.8 − 117.887 = 148.995, rounding off gives: 149 (HB = 149)

    HBC = 3.303 × 85.0 − 117.887 = 162.868, rounding off gives: 163 (HB = 163)

    Table 1.9

    : HRB range is 86.0–89.5; HB range is 167–181 respectively. The equation for calculating Brinell hardness at 3000-kgf load (HBC) is

    Example of calculation:

    HBC = 4.031 × 86.0 − 180.006 = 166.660, rounding off gives: 167 (HB = 167)

    HBC = 4.031 × 89.5 − 180.006 = 180.768, rounding off gives: 181 (HB = 181)

    Table 1.10

    : HRB range is 90.0–95.5; HB range is 183–212 respectively. The equation for calculating Brinell hardness at 3000-kgf load (HBC) is

    Example of calculation:

    HBC = 5.289 × 90.0 − 293.668 = 182.342, rounding off gives: 182 (HB = 183)

    HBC = 5.289 × 95.5 − 293.668 = 211.432, rounding off gives: 211 (HB = 212)

    Table 1.11

    : HRB range is 96.4–100.0; HB range is 217–241 respectively. The equation for calculating Brinell hardness at 3000-kgf load (HBC) is

    Example of calculation:

    HBC = 6.736 × 96.4 − 432.340 = 217.010, rounding off gives: 217 (HB = 217)

    HBC = 6.736 × 100.0 − 432.340 = 241.260, rounding off gives: 241 (HB = 241)

    Rockwell C hardness numbers and corresponding Brinell hardness numbers measured at 3000-kgf load were grouped in eight tables, each of which contains various numbers of data points.

    Statistical treatment of data in each group produced linear regression equations with the correlation coefficients from 0.9989 to 0.99996. These coefficients indicate a strong linear relationship between Rockwell C hardness numbers and Brinell hardness numbers measured at 3000-kgf load.

    The following eight equations were developed (Appendix 1, Tables 1.12–1.19).

    Table 1.12

    : HRC range is 20.5–25.4; HB range is 229–255 respectively. The equation for calculating Brinell hardness at 3000-kgf load (HBC) is

    Example of calculation:

    HBC = 5.284 × 20.5 + 120.491 = 228.813, rounding off gives: 229 (HB = 229)

    HBC = 5.284 × 25.4 + 120.491 = 254.705, rounding off gives: 255 (HB = 255)

    Table 1.13

    : HRC range is 26.0–29.9; HB range is 258–285 respectively. The equation for calculating Brinell hardness at 3000-kgf load (HBC) is

    Example of calculation:

    HBC = 6.888 × 26.0 + 78.838 = 257.926, rounding off gives: 258 (HB = 258)

    HBC = 6.888 × 29.9 + 78.838 = 284.789, rounding off gives: 285 (HB = 285)

    Table 1.14

    : HRC range is 30.0–35.5; HB range is 286–331 respectively. The equation for calculating Brinell hardness at 3000-kgf load (HBC) is

    Example of calculation:

    HBC = 8.195 × 30.0 + 39.932 = 285.782, rounding off gives: 286 (HB = 286)

    HBC = 8.195 × 35.5 + 39.932 = 330.854, rounding off gives: 331 (HB = 331)

    Table 1.15

    : HRC range is 36.0–39.8; HB range is 336–369 respectively. The equation for calculating Brinell hardness at 3000-kgf load (HBC) is

    Example of calculation:

    HBC = 8.718 × 36.0 + 21.825 = 335.673, rounding off gives: 336 (HB = 336)

    HBC = 8.718 × 39.8 + 21.825 = 368.801, rounding off gives: 369 (HB = 369)

    Table 1.16

    : HRC range is 40.0–45.7; HB range is 371–429 respectively. The equation for calculating Brinell hardness at 3000-kgf load (HBC) is

    Example of calculation:

    HBC = 10.057 × 40.0 − 31.761 = 370.519, rounding off gives: 371 (HB = 371)

    HBC = 10.057 × 45.7 − 31.761 = 427.844, rounding off gives: 428 (HB = 429)

    Table 1.17

    : HRC range is 46.1–49.8; HB range is 433–479 respectively. The equation for calculating Brinell hardness at 3000-kgf load (HBC) is

    Example of calculation:

    HBC = 12.642 × 46.1 − 150.881 = 431.915, rounding off gives: 432 (HB = 433)

    HBC = 12.642 × 49.8 − 150.881 = 478.691, rounding off gives: 479 (HB = 479)

    Table 1.18

    : HRC range is 50.0–55.2; HB range is 481–564 respectively. The equation for calculating Brinell hardness at 3000-kgf load (HBC) is

    Example of calculation:

    HBC = 15.979 × 50.0 − 318.988 = 479.962, rounding off gives: 480 (HB = 481)

    HBC = 15.979 × 55.2 − 318.988 = 563.053, rounding off gives: 563 (HB = 564)

    Table 1.19

    : HRC range is 56.0–58.7; HB range is 578–627 respectively. The equation for calculating Brinell hardness at 3000-kgf load (HBC) is

    Example of calculation:

    HBC = 18.694 × 56.0 − 469.983 = 576.881, rounding off gives:577 (HB = 578)

    HBC = 18.694 × 58.7 − 469.983 = 627.355, rounding off gives:627 (HB = 627)

    These nineteen highly-accurate linear regression equations should be used to convert Brinell hardness numbers measured at 500-kgf load (five equations), Rockwell B-scale hardness numbers (six equations), and Rockwell C-scale hardness numbers (eight equations) into Brinell harness numbers at 3000-kgf load.

    The other mechanical properties (tensile strength–hardness relationship and machinability ratings) are described for each of the following steel categories: carbon steels, alloy steels, stainless steels, and tool steels.

    Concluding Remarks on Hardness

    1. Brinell hardness test performed at 3000-kgf load is the most commonly-used test to define the hardness characteristic of a steel.

    2. If the hardness numbers of a workpiece are available only in Brinell at 500-kgf load or in Rockwell B-scale, or in Rockwell C-scale, those hardness numbers must be converted into the equivalent Brinell hardness numbers at 3000-kgf load.

    3. The author developed nineteen linear regression equations to perform such conversions. There are five equations to convert Brinell hardness numbers measured at 500-kgf load (HB5) providing 99.1–100% accuracy; six equations to convert Rockwell hardness numbers measured on the B-scale (HRB) providing 99.5–100% accuracy; and eight equations to convert Rockwell hardness numbers measured on the C-scale (HRC) providing 99.7–100% accuracy. These nineteen equations are summarized in Table 1.1.

    Table 1.1. Conversion into Brinell hardness numbers measured at 3000-kgf load – HBC

    As can be seen from the table, the correlation coefficients range from 0.9962 to 0.99996, indicating high accuracy in converting HB5, HRB, and HRC hardness numbers into HBC hardness numbers.

    1.2. Carbon Steels

    Carbon steels are by far the most frequently used steel. In 1988 the United States produced 99.9 million tons of steel, including 86.8 million tons, or 86.9% of carbon steel (Ref 7, p.147). The feasibility of using carbon steels depends on whether or not their properties (tensile, yield, and fatigue strengths; impact resistance, need for heat treating, etc.) are suitable for the parts to be used. If the required characteristics can be obtained with carbon steel, most users select this less costly steel.

    The American Iron and Steel Institute (AISI) defines carbon steel as follows: Steel is considered to be carbon steel when no minimum content is specified or required for chromium, cobalt, molybdenum, nickel, niobium, titanium, tungsten, vanadium or zirconium, or any other element to be added to obtain a desired alloying effect; when the specified minimum for copper does not exceed 0.40 percent; or when the maximum content specified for any of the following elements does not exceed the percentages noted: manganese 1.65, silicon 0.60, copper 0.60 (Ref 7, p.147).

    Sometimes the term plain carbon steel is used instead of carbon steel. It is acceptable, but no longer considered best practice. Some of the cutting tool companies — such as KOMET of America (Ref 10), Sandvik Coromant (Ref 11), and WALTER Waukesha (Ref 12) — define carbon steel as unalloyed steel; Greenleaf Corporation defines carbon steel as non-alloy steel (Ref 13, pp. ATI 04–08). Such definitions are wrong because steel without alloying elements is nothing else but iron.

    Carbon steels are designated by an identical AISI or SAE (Society of Automotive Engineers) four-digit number. The last two digits indicate the approximate middle of the carbon range expressed in hundredth of one percent. For example, AISI 1020 grade has a carbon content of 0.18–0.23%. The first two digits of the number are also significant. The number 10 indicates nonresulfurized grades with manganese content from 0.25 to 1.00%. These grades include low-carbon, medium-carbon, and high-carbon steels. The number 11 denotes free-machining resulfurized grades. The number 12 indicates free-machining resulfurized and rephosphorized grades. The number 15 indicates nonresulfurized grades with manganese content from 0.75 to 1.65%.

    Carbon steels are also designated by UNS (Unified Numbering System). A UNS number is assigned to each chemical composition of a metallic alloy. The UNS designation consists of a letter followed by a five-digit number. Letters G or H are assigned to carbon steels. An H indicates carbon steel produced to prescribed hardenability limits. The initial four digits are the same as the four-digit number assigned by AISI or SAE and the last digit is 0 for the majority of grades. If the last digit is 4, then the steel contains 0.15 to 0.35% lead. For example, AISI 12L14 is the same as UNS G12144, AISI 1038 is UNS G10380, and AISI 1038H is UNS H10380.

    Some AISI–SAE grades are also designated by letters B and L, placed between the second and third digits, and by the letter H after the four digits. Letters B and L denote grades containing boron and lead respectively. An H indicates steel produced to prescribed hardenability limits, for

    Enjoying the preview?
    Page 1 of 1