Sei sulla pagina 1di 18

Study on Heating Process of Ultrasonic

Welding for Thermoplastics

ZONGBO ZHANG,1 XIAODONG WANG,2 YI LUO,2,*


ZHENQIANG ZHANG1 AND LIDING WANG2
1
Key Laboratory for Precision and Non-traditional Machining
Technology of Ministry of Education, Dalian University
of Technology, Dalian 116023, China
2
Key Laboratory for Micro/Nano Technology and System of Liaoning
Province, Dalian University of Technology, Dalian 116023, China

ABSTRACT: Heating mechanisms of ultrasonic welding for thermoplastics were


studied via numerical simulation and experiment. Relationship between dynamic
modulus and static relaxation modulus of thermoplastics was theoretically analyzed.
Segment time–temperature equivalency equation for different temperature ranges
and a method for directly separating viscoelastic heat from strain energy were used in
the simulation. Temperature measurement tests with poly(methyl methacrylate)
specimens were carried out to verify the simulation results. Results of simulation and
experiment reveal that interfacial friction rather than viscoelastic heat initially start
the whole welding process. Viscoelastic heating becomes dominant when tempera-
ture reaches Tg (glass transition temperature) of the material. And viscoelastic heat
provides most required heat during welding.

KEY WORDS: ultrasonic welding, finite element method, thermoplastics, visco-


elastic.

INTRODUCTION

LTRASONIC WELDING IS a well-established process to connect thermo-


U plastic components due to its advantages: free of foreign substances,
little damage to devices, high strength, short cycle time, and localized
heating. Recently, some special usages of precision ultrasonic welding, for
instance, connecting polymer micro electro mechanical systems (MEMS)

*Author to whom correspondence should be addressed. E-mail: luoy@dlut.edu.cn


Figures 1, 5 and 13 appear in color online http://jtc.sagepub.com

Journal of THERMOPLASTIC COMPOSITE MATERIALS, Vol. 23—September 2010 647


0892-7057/10/05 0647–18 $10.00/0 DOI: 10.1177/0892705709356493
ß The Author(s), 2010. Reprints and permissions:
http://www.sagepub.co.uk/journalsPermissions.nav
648 Z. ZHANG ET AL.

devices [1,2] and nano electromechanical systems (NEMS) devices [3] had
been investigated. Improving welding accuracy becomes a hot research spot.
Since temperature distribution in the welding zone directly influences the
welding accuracy, many theoretical and experimental researches have been
done to clarify the heating mechanisms and properties of ultrasonic welding.
Initially, most researchers believed that the rising temperature during
ultrasonic welding was generated from the interfacial friction of the
contacting surfaces. However, the experimental results of Menges and
Potente [4] indicated that the main heating effect during ultrasonic welding
was from viscoelastic heating rather than interfacial friction. Until now, many
researchers confirmed that the heating mechanism could be explained by
viscoelastic dissipation of polymer under cycle loading [5,6]. Because of high
welding speed and small welding zone in ultrasonic welding the evolution of
temperature field and the melting process of the welding zone could hardly be
directly observed using existing techniques. With the improvement of
numerical calculation methods, simulation becomes an effective approach
to study this welding process. Nesterenko and Senchenkov [7] and Li et al. [8]
developed a 1D finite element model for calculation of viscoelastic heat using
Galerkin and energy variational calculus methods, respectively. Benatar and
Gutowski [9] established a 2D thermal model to simulate the temperature
field of the welding zone. And Wang et al. [10] used finite element method
(FEM) to simulate the generation of viscoelastic heat during ultrasonic
welding. In these approaches, most attention was paid on the generation of
the heat which made polymer melt. The detail heating mechanism which
made the temperature rise from room temperature to Tg of polymer has not
been thoroughly studied. Moreover, during the whole welding process, the
polymer experiences multi states such as glassy state, transitional state,
rubbery state, and viscous flow state. The mechanical properties of polymer
abruptly change as it passes through different characteristic temperature
points. And the viscoelastic property of thermoplastics is greatly influenced
by temperature. So it is reasonable to study the heating mechanisms in
different temperature ranges, respectively.
In this study, the heating mechanisms in temperature ranges both below
and above Tg of poly(methyl methacrylate) (PMMA) were studied based on
numerical simulation and temperature measuring experiment. The frequency
and temperature-dependent dynamic characteristics of polymer were
theoretically studied. Method of separating viscoelastic heat from strain
energy was presented in this article. Based on the above theoretical appro-
aches, viscoelastic heat in different temperature ranges was calculated by
FEM. Transient simulation in glassy state of PMMA was also performed and
frictional heat was calculated using thermal-structure coupling module. Temp-
erature tests with different samples were conducted during ultrasonic welding.
Ultrasonic Welding for Thermoplastics 649

Based on the results of simulation and experiment, effects of frictional heat


and viscoelastic heat in ultrasonic welding were analyzed.

THEORETICAL APPROACH

Relationship Between Dynamic Relaxation Modulus and Static Relaxation


Modulus

Thermoplastic polymers are typical viscoelastic materials. They are often


theoretically modeled as generalized Maxwell solids or generalized Voigt–
Kelvin solids, which use a combination of springs and dashpots to present
the viscoelastic properties [11]. In this article, the generalized Maxwell model
is employed, as shown in Figure 1.
The model is governed by:
8
< ¼ P
n
i
, ð1Þ
:  þ i¼1i
_ ¼  "_
i Ei i i

where  is the total stress, " is the total strain,  i is the stress on the i-th
Maxwell unit (i ¼ 1,2, . . . , n), i and Ei are the elastic modulus and viscosity
coefficient of the i-th Maxwell unit, respectively, and n is the number of
Maxwell units in the model.
Applying Laplace transform and inverse Laplace transform on Equation (1),
the static constitutive equation of the model can be obtained as:
! !
X n E
i t
X n
1 t
ðtÞ ¼ Einf þ Ei e i "0 ¼ Einf þ Ei e i "0 ¼ EðtÞ"0 , ð2Þ
i¼2 i¼1

E1 E2 Ei
Einf … …

h2 hi
h1

Figure 1. The generalized Maxwell model.


650 Z. ZHANG ET AL.

where Einf is the elastic


P modulus of the first Maxwell unit as shown in
1t
Figure 1, EðtÞ ¼ Einf þ ni¼1 Ei e i is the static relaxation modulus, i ¼ i =Ei
is the relaxation time of the i-th Maxwell unit. While under dynamic loading
the constitutive equation of viscoelastic material can be expressed as:
Z t
d"ðÞ
ðtÞ ¼ Eðt  Þ d: ð3Þ
1 d

Suppose the applied cyclic load is:

"ðtÞ ¼ "0 ei!t , ð4Þ

where "0 is the amplitude of the strain and ! is angular frequency of


the load.
And dividing relaxation modulus E(t) into two parts:

EðtÞ ¼ Einf þ Eb ðtÞ, ð5Þ

where

X
n
1 t
Eb ðtÞ ¼ Ei e i :
i¼1

Therefore, Equation (3) can be written as:


Z t
d"0 ei!
ðtÞ ¼ ½Einf þ Eb ðt  Þ d
1 d
Zt ð6Þ
¼ Einf "0 ei!t þ i!"0 Eb ðt  Þei! d:
1

Defining  ¼ t  , the above equation can be expressed as:


Z 1 Z 1
ðtÞ ¼ ½Einf þ ! sin !Eb ðÞd þ i! cos !Eb ðÞd"0 ei!t : ð7Þ
0 0

We define dynamic complex modulus as:


8 
< E ð!Þ ¼ E0 ð!Þ þ RiE00 ð!Þ
1
E0 ð!Þ ¼ EinfR þ ! 0 Eb ðtÞ sin !tdt , ð8Þ
: 00 1
E ð!Þ ¼ ! 0 cos !tEb ðtÞdt
Ultrasonic Welding for Thermoplastics 651

where E*, E0 , and E00 are complex modulus, storage modulus, and loss
modulus, respectively. So E0 and E00 can be obtained by E(t) and its Fourier
transformation [11]. Conversely, E(t) can also be obtained from E0 and E00
through inverse Fourier transform [12]:
Z
2 11 0
EðtÞ ¼ Einf þ ½E ð!Þ  Einf  sin !td! or
 0 !
Z1 ð9Þ
2 1 00
EðtÞ ¼ Einf þ ½E ð!Þ cos !td!:
 0 !

Therefore, constitutive equation of viscoelastic material is:

ðtÞ ¼ E ð!Þ"0 ei!t ¼ E ð!Þ"ðtÞ: ð10Þ

Equation (8) indicates that dynamic complex modulus is a function of load


frequency. However, most used complex modulus in the previous studies was
obtained from extrapolations of measures at lower (3–110 Hz) or higher
(500–3500 kHz) frequencies. It is hard to measure it directly at ultrasonic
frequency range [5]. Compared with measuring complex modulus, the
measurement of static relaxation modulus of viscoelastic material is easier.
As shown in Equation (8) the relationship between static relaxation modulus
and dynamic complex modulus could be a bridge from E*(!) to E(t).
The relaxation curve segments of PMMA at different temperature was
measured and the main curve was obtained according to time–temperature
equivalency equation, as shown in Figure 2. Based on the relaxation modulus,

104
40–135°C
105°C
40°C
Relaxation modulus (MPa)

100°C
103 60°C
110°C
80°C
92°C
112°C
102

115°C
Main curve
101 120°C
125°C
135°C
100
10–10 100 1010 1020
Time (s)

Figure 2. Relaxation curve segments of PMMA of various temperatures and the main curve.
652 Z. ZHANG ET AL.

the dynamic modulus at different loading frequency, called dynamic


mechanical analysis (DMA) spectrum, is calculated using Matlab 7.0
based on Equation (8), as shown in Figure 3. According to Guo and Zhao
[13], the trend of E0 and E00 coincides with the typical characteristic of
polymeric materials. Figure 4 shows the stress–strain curve at room
temperature under ultrasonic vibration using the obtained dynamic modulus.
There are evident hysteresis loops which represent energy dissipating into
viscoelastic heat.

2500
1 Hz
Storage and loss modulus (MPa)

E′ 1 kHz
2000 20 kHz
60 kHz

1500

1000

500

E ′′
0
40 60 80 100 120 140 160
Temperature (°C)

Figure 3. DMA spectrum of PMMA calculated by static modulus.

40

30

20
Stress (KPa)

10

–10 Hysteresis loops dissipating


into viscoelastic heat
–20

–30

–40
–0.02 –0.01 0 0.01 0.02
Strain (mm)

Figure 4. Stress–strain curve under ultrasonic vibration.


Ultrasonic Welding for Thermoplastics 653

Time–Temperature Equivalency Equation

The modulus of polymer is also affected by temperature. Generally, the


relaxation modulus of polymer is obtained by tension test. But the relaxation
time is often very long in room temperature. So segments of relaxation curve
are usually tested in several relatively high temperatures, as shown in Figure 1.
Thereafter, according to time–temperature equivalent law, the segments are
shifted to obtain a main relaxation curve for a long time span. Mano [14]
revealed that the Arrhenius equation can be employed to describe the
temperature-dependent properties of polymer when the temperature is below
Tg, which can be explained on the basis of Adam–Gibbs theory. And the semi-
empirical WLF (Williams Landel Ferry) equation is adopted when the
temperature exceeds Tg. Consequently, the segmented time–temperature
equivalency equation can be expressed as:
8  
< H 1 1
T  Tg
2:303R T  Tg
lg T ¼ , ð11Þ
: C1 ðTTg Þ Tg 5T  Tg þ 60 C
C2 þTTg

where lgaT is the translation amount of relaxation curve in logarithmic


coordinates of time from temperature T to Tg, H is active energy, R is molar
gas constant, C1 and C2 are constants.

Separating Vicsoelastic Heat from Strain Energy

Under a cyclic load, the strain energy of viscoelastic material in per


volume can be established by [15]:
Z t
1 1
W¼ ðtÞd"ðtÞ ¼ E0 "20 ðcos 2!t  1Þ þ E00 "20 ð2!t  sin 2!tÞ ¼ W1 þ W2 ,
0 4 4
ð12Þ

where W is strain energy, W1 ¼ 1=4E0 "20 ðcos 2!t  1Þ, W2 ¼ 1=4E00 "20 ð2!t
sin 2!tÞ. W1, named elastic energy, is a nonpositive oscillation function
between 0 and 1=2E0 "2 , which indicates that the energy from outside is stored
and then delivered fully. W2 is composed of 1=2E00 "20 !t and 1=4E00 sin 2!t,
which oscillates and rises along a line with slope coefficient of 1=2E00 "20 !. In
this process, the mechanical energy is transformed into viscoelastic heat. So
the viscoelastic heat generating rate in unit volume can be founded as:

1
Q ¼ E00 "20 ð2!  2! cos 2!tÞ: ð13Þ
4
654 Z. ZHANG ET AL.

× 10–4
12
W
W1
10
W2
8
Energy (J )

–2
0 p / 2w p /w 3p / 2w 2p /w 5p / 2w 3p /w 7p / 2w
Time (s)
Figure 5. Strain energy, elastic energy, and viscoelastic energy of PMMA during cycling load.

In FEM calculation the strain energy W of each element can be directly


obtained. In order to obtain viscoelastic energy, it is vital to accurately
separate W2 from W. Figure 5 shows the numerical calculation results of
relationship among W, W1, and W2.
As shown in Figure 5, the strain energy W is less than viscoelastic energy
W2 when t 6¼ n=! (n is a positive integer). Energy from outside is partly
stored in the form of elastic energy and the other part is dissipated into
viscoelastic heat. Especially at the time point t ¼ ð2n  1Þ=!, the elastic
energy reaches maximum in a period. If using the strain energy at this time
point is used to substitute for viscoelastic heat, the error would reach
maximum in a cycle. However, when t ¼ n=! the elastic energy is zero and
the strain energy W can substitute the viscoelastic heat W2. So it is accurate
to use strain energy as viscoelastic heat at time t ¼ n=!.

SIMULATION ANALYSIS

Simulation Strategy

The welding process of PMMA components were simulated using


ANSYS program. To improve the calculation efficiency, 2D model was
employed, as shown in Figure 6. Compared with traditional point-contact or
line-contact energy director, rectangle energy director used here has larger
contact area, which can make the temperature evolution slower and the
observation on the welding process easier.
Ultrasonic Welding for Thermoplastics 655
Ultrasonic vibration

1.0
Energy director

0.4
0.4

0.6
Constrained
Unit: mm

Figure 6. Finite element model.

Table 1 Parameters of generalized Maxwell model.

Unit Relaxation Relaxation


number (i) time  i (s) modulus Ei (MPa)
1 2  104 1.94  102
2 2  103 2.83  102
3 2  102 5.54  102
4 2  101 6.02  102
5 2 3.88  102
6 2  101 1.56  102
7 2  102 4.10  101
8 2  103 1.38  101
9 2  104 3.86
10 2  105 7.90  101
Einf ¼ 2.24

In viscoelastic heat calculation, the model was meshed by Visco88


viscoelastic elements. Viscoelastic properties of PMMA were presented
by ten units generalized Maxwell model, as Equation (3) mentioned.
The obtained relaxation data was fitted in terms of Prony series, as listed in
Table 1. These parameters were filled into the corresponding viscoelastic
material property tables using ANSYS Parametric Design Language
(APDL). Then program was developed using User Programming Features
and Fortran Language to describe the temperature-dependent properties
according to segment time–temperature equivalency equation.
The calculation schematic diagram was shown in Figure 7. After parameter
initialization, static pressure was applied on the upper surface of the model.
Then dynamic loading was applied. A loading cycle was divided into 10
substeps. As mentioned in the theoretical analysis, the strain energy was equal
656 Z. ZHANG ET AL.

Initialization of parameters

Apply static pressure

Apply substeps of cycle load


Next load cycle

Extract strain energy increase

Thermo conduation calculation

Update the temperature field

Figure 7. Strategy to simulate the viscoelastic heat.

to the viscoelastic heat at the end of each cycle. So strain energy was extracted
as viscoelastic heat at the end of the tenth substep. Then thermal conduction
process was calculated. The new temperature field was used for the
calculation of the next load step. Codes were developed using APDL to
perform the simulation process.
Thermal-structure coupling model was employed to simulate the dynamic
performance in glassy state of PMMA and to calculate the frictional heat.
Direct coupling Plane223 element was used in the simulation. And contact
element pair (Element 172 and Element 169) was employed to simulate the
contacting situation at the interface. Initial temperature of the calculation
was 258C and the interfacial friction coefficient was 0.4 which was obtained
by friction-meter (RK Tensometric, Germany).

Results of Simulation

Figures 8 and 9 show the temperature rising curves caused by viscoelastic


heat near the interface of energy director with different initial temperatures.
Figure 8 indicates that the temperature increasing rate with initial
Ultrasonic Welding for Thermoplastics 657

1.6 Initial temperature 90°C

Temperature increment (°C)


Initial temperature 25°C
1.4
1.2
1.0
0.8
0.6
0.4
0.2
0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16 0.18 0.2
Time (s)
Figure 8. Temperature curves caused by viscoelastic heat with initial temperatures of 25
and 908C.

145
140
135
130
Temperature (°C)

125
120
115
110
Tg
105
100
95
0 0.02 0.04 0.06 0.08 0.10 0.12 0.14 0.16
Time (s)

Figure 9. Temperature curve caused by viscoelastic heat with initial temperatures of 988C.

temperature of 258C is close to that of 908C, which reveals that the


increasing rate of temperature does not much affected by temperature in this
temperature range. And the increase amount is less than 28C as the welding
process lasts 0.2 s. With this rate, it is impossible to make energy director
melt in seconds as in actual welding process. It is obvious that the
viscoelastic heat effect is not dominant in this temperature range. Figure 9
shows that the temperature dramatically increases by about 408C within less
than 0.02 s when the temperature reaches Tg of the material. This trend is
consistent with experimental results in references [16, 17]. According to the
658 Z. ZHANG ET AL.

analysis above, viscoelastic heat is not evident when the temperature is


below Tg. Once the temperature reaches Tg of the material, viscoelastic heat
makes dramatic increase of temperature.
Transient analysis of welding process indicates that the relative sliding
velocity at the interface of the energy director is up to 532.463 mm/s, and the
frictional stress reaches 12.252 MPa. Moreover, simulation result shows that
the maximum sliding velocity is located at the corner of energy director’s
surface. It gradually decreases as getting away from the corner. With friction
coefficient of 0.4, the highest power density can reach 6.52 W/mm2 in the
corner, which is sufficient to make temperature of the corner rise to a
relatively high temperature.
Figure 10 shows the temperature distribution caused by interfacial friction.
Figure 11 shows the temperature curves of corner point (point 1), inner point
(point 3) on the energy director and point (point 2) in the lower polymer plate
as illustrated in Figure 10. The temperature of point 1 rises faster and higher
as a result of stress concentration and high relative sliding velocity.
Results of simulation indicate that interfacial friction effect generates the
heat which is vital to start the welding process. It makes the temperature of
the corner rise earlier than any other locations in rectangle energy director.
Once temperature of the corner reaches Tg, the intensive viscoelastic heat

NODAL SOLUTION

3
1

Unit : °C

22.319 39.26 56.201 73.141 90.082


30.789 47.73 64.671 81.612 98.552

Figure 10. Temperature distribution caused by interfacial friction.


Ultrasonic Welding for Thermoplastics 659

will be activated. By thermal conduction and viscoelastic heat generation,


the heat-affected zone (HAZ) rapidly spreads. And finally the energy
director melts to form the joint to connect the two components. Figure 12
shows the temperature distribution and the deformation of energy director,

100 Point 1
Point 2
90 Point 3
80
Temperature (°C)

70
60
50
40
30
20
10
0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1
Time (s)

Figure 11. Temperature curves caused by friction at different locations.

NODAL SOLUTION

Unit : °C

25 69.34 113.681 158.021 202.361


47.17 91.511 135.851 180.191 224.531

Figure 12. Simulation result of temperature distribution and deformation of energy director.
660 Z. ZHANG ET AL.

with the effects of frictional heating, viscoelastic heating and thermal


conduction, after 0.2 s ultrasonic welding of PMMA.

EXPERIMENTAL SETUP

Because viscoelastic heat is a kind of volumetric heat, its generation rate is


affected by stress and strain, whether there is interface will not influence its
generation rate. However, as a kind of facial heat frictional heat relies on the
existence of contacting interface. In order to study the influence of interface
on temperature distribution during the welding process, two types of
PMMA specimens with and without interface were fabricated. The
dimensions of the specimens are the same to ensure the same distributions
of strain and stress during ultrasonic welding. Figure 13 shows dimensions
of the specimens. The used PMMA were provided by Asahi Kasei
Corporation, with Tg of 1058C.
The welding process was performed on a standard ultrasonic welding
system (Branson 2000X f/aef, USA), with frequency of 30 kHz. The specimen
was clamped on the leveling anvil. Thereafter, a static pressure and an
ultrasonic range oscillation were applied to the specimen through the horn.
Energy of ultrasonic was concentrated by the energy director. The welding
process was set under ‘time’ mode. Based on experience, the welding and
holding time were 0.3 and 0.5 s. The trigger force, welding force as well as
holding force were 44, 300, and 180 N, respectively. Low temperature
rising rate is benefit for observing the evolution of temperature and
the melting process of the energy director. In order to slow down the
temperature evolution process, low ultrasonic amplitude of 18 mm was used in
the experiment.
Rapid response nickel–chromium–silicon/nickel–silicon thermocouple
(Chal-0005, Omega Company, USA) with diameter of 12.7 mm was used
to measure the temperature during welding. The diameter of thermocouple’s
head was about 15 mm, and the responding time was less than 10 ms, which
was rapid enough to capture the changing temperature during ultrasonic
welding. A multi-channel data acquisition board (NI DAQPad-6015) with
sample rate of 50 kHz per channel was employed. Amplifier (AD524) with
magnifying ratio of 1000 was used to amplify the sampled signal. Data
acquisition codes were developed using LabVIEW to realize signal sampling
and processing in a computer. During temperature testing process, the
temperature of the cold junction was measured by platinum resistance
thermometer. And temperature compensation was carried out based on the
obtained temperature profile of the cold junction. This temperature testing
system was calibrated by temperature calibrator (MKST TP 28850, Omega
Company, USA) from 20 to 2008C with a measuring error of 48C. Three
Ultrasonic Welding for Thermoplastics 661

40
20
3

3
7

2
Unit: mm
No interface
Interface
(a) (b)
Figure 13. Schematic of specimens (a) without interface and (b) with interface.

Figure 14. Embedded locations of thermocouples.

Location A
40
Location B
Temperature (°C)

Location C

30

20
0 0.5 1 1.5 2
Time (s)

Figure 15. Temperature curves of specimen without interface obtained from experiment.

thermocouples were installed in different locations of the energy director, as


shown in Figure 14.

RESULTS AND DISCUSSION

Figure 15 shows the temperature curves in three different locations of the


specimen without interface. The maximum temperature is about 408C,
662 Z. ZHANG ET AL.

which is well below Tg of the using material, after as long as 2 s ultrasonic


welding. There are burn scars on the upper surface which is in contact with
the horn of the welding machine, while no obvious deformation of energy
director is observed after welding. The temperature of point C rises faster
than that of A and B. So it is reasonable to conclude that the increase of
temperature in Figure 15 is mainly caused by thermal conduction from the
upper surface of the specimen.
Figure 16 shows the temperature curves in three different locations of the
second specimen. The temperature of the corner point (location A) rises
faster and higher than that of other points, which has been well predicted by
simulation. There is an obvious inflection point of temperature rising rate in
curve A at about 1058C which is just Tg of the used material. Before
inflection point, temperature rises by 808C in about 0.3 s. The temperature
rising rate is about 3008C/s. It is in the same order with that of 5008C/s in
the simulation result of interfacial frictional heat. As shown in the
simulation result, the temperature rising rate decreases as getting away
from the corner. In experiments, thermocouple cannot be placed at the very
corner. The difference of temperature rising rate between simulation and
experiment may be caused by this reason.
When temperature reached Tg, there is a dramatic increase of temperature
rising rate in curve A. Temperature suddenly raises by 708C within 0.025 s.
This trend is in well accordance with the simulation result of viscoelastic
heat. Temperature rising rate of point B is smaller than that of point A at
first. After about 0.3 s the rising rate increases, which is likely because of
thermal conduction from the corner of energy director. There also exists an
obvious inflection point in curve B at Tg, after which the temperature rising

220
200 Location A
180 Location B
Location C
Temperature (°C)

160
140
120 Tg
100
80
60
40
20
0 0.2 0.4 0.6 0.8 1
Time (s)
Figure 16. Temperature curves of specimen with interface obtained from experiment.
Ultrasonic Welding for Thermoplastics 663

rate dramatically increases. The temperature rising processes of points A


and B are similar except that the temperature of point A rise earlier than
that of point B, which is well predicted in the simulation results. Although
there are two distinct changes of temperature rising rate of C, yet the
temperature rising rate is low during the whole welding process. This trend is
similar with curves in Figure 15. As pointed out in the results of simulation,
the viscoelastic heat is not evident when the temperature is below Tg. So the
increase of temperature at point C may be caused by thermal conduction
from the interface.
In summary, the above results indicate that it is the combination of
interfacial friction and thermal conduction effect causes the initial temp-
erature rise near the corner of energy director’s surface when the temperature
is below Tg. Then viscoelasctic heat becomes dominant as temperature of
some locations rise to Tg. Thermal conduction effect makes temperature of
material near the interface rise to Tg, and then intensive viscoelasctic
heat above Tg is activated. The HAZ rapidly spreads as a result of
viscoelasctic heat above Tg and thermal conduction. Therefore, interfacial
friction is a key factor to get the initial heat for the welding process. And
viscoelastic effect above Tg generates the main heat to make the energy
director melt during welding.

CONCLUSION

A method for directly calculating the viscoelastic heat during ultrasonic


welding was established using FEM. Experiments using specimens both with
and without interfaces verified the simulation results. And several
conclusions are gotten from these results:
(1) The designed experiment can successfully capture details of the changing
temperature in some characteristic locations. And the results indicate that
not only the distribution of strain and stress but also the existence of
interface between the components is prerequisite to the welding process.
(2) Both interfacial frictional heat and viscoelastic heat calculated in
the simulation are in good agreement with the experimental results.
Initially, interfacial friction effect is the main reason to start the welding
process. And the majority of the heat is comes from viscoelastic heat
effect above Tg.
(3) Temperature field in rectangle-shaped energy director is nonuniform
during welding. The corner is heated up firstly follows by temperature
rising in the whole contact surface of energy director. Then thermal
conduction and viscoelastic heating effects make the interior part of
energy director melt and the welding joint forms.
664 Z. ZHANG ET AL.

ACKNOWLEDGMENT

This research was supported by the Program for New Century Excellent
Talents in University (NCET-06-0279) of Education Ministry of the People’s
Republic of China and the National Natural Science Foundations of China
(No. 50775024 and No. 50975037).

REFERENCES

1. Truckenmueller, R. and Ahrens, R. (2006). An Ultrasonic Welding Based Process for


Building Up a New Class of Inert Fluidic Micro Sensors and Actuators from Polymers,
Sensors and Actuators A, 132: 385–392.
2. Zhang, Z.B., Luo, Y., Wang, X.D., Zhang, Z.Q. and Wang, L.D. (2008). Advances in
Ultrasonic Welding of Plastics and its Usage in Polymer MEMS Bonding, Welding and
Joining, 8: 9–15.
3. Chen, C.X., Yan, L.J., Kong, E.S.W. and Zhang, Y.F. (2006). Ultrasonic Nanowelding of
Carbon Nanotubes to Metal Electrodes, Nanotechnology, 17: 2192–2197.
4. Menges, G. and Potente, H. (1996). Study on the Weldability of Thermoplastic Materials
by Ultrasound, Welding in the World, 9(1/2): 46–59.
5. Nonhof, C.J. and Luiten, G.A. (1996). Estimates for Process Conditions During the
Ultrasonic Welding of Thermoplastics, Polymer Engineering and Science, 36: 1177–1183.
6. Gutnik, V.G. Gorbach, N.V. and Dashkov, A.V. (2002). Some Characteristics of
Ultrasonic Welding of Polymers, Fibre Chemistry, 34: 426–432.
7. Nesterenko, N.P. and Senchenkov, I.K. (2004). Improvement of Ultrasound Welding of
Polymer and Thermoplastic Composite Materials, Welding Research Abroad, 50: 28–34.
8. Li, X.C., Ling, S.F. and Sun, Z. (2004). Heating Mechanism in Ultrasonic Welding of
Thermoplastics, International Journal for the Joining of Materials, 16: 37–42.
9. Benatar, A. and Gutowski, T. (1989). Ultrasonic Welding of Graphite APC-2 Composites,
Polymer Engineering and Science, 29: 1705–1721.
10. Wang, X., Yan, J., Li, R. and Yang, S. (2006). FEM Investigation of the Temperature Field
of Energy Director During Ultrasonic Welding of PEEK Composites, Journal of
Thermoplastic Composite Materials, 19: 593–607.
11. Ferry, J. D. (1980). Viscoelastic Properties of Polymer, 3rd edn, Wiley, New York.
12. Shen, T.F., Gao, M. and Zhao, B.H. (1995). Engineering Methods for the Transform
Inversion of Stress Relaxation Modulus and Complex Modulus, Acta Armamentar, 8: 40–44.
13. Guo, M.L., and Zhao, D.L. (2005). Polymer Physics, Beijing University of Aeronautics &
Astronautics Press, Beijing.
14. Mano, J.F. and Viana, J.C. (2006). Stress-Strain Experiments as a Mechanical
Spectroscopic Technique to Characterize the Glass Transition Dynamics in
Poly(Ethylene terephthalate), Polymer Testing, 25: 953–960.
15. Christensen, R.M. (1982). Theory of Viscoelasticity, An Introduction, 2nd edn, Academic
Press Inc., New York.
16. Frankel, E.J. and Wang, K.K. (1980). Energy Transfer and Bond Strength in Ultrasonic
Welding of Thermoplastics, Polymer Engineering and Science, 20: 396–401.
17. Tolunay, M.N., Dawson, P.R. and Wang, K.K. (1983). Heating and Bonding Mechanisms
in Ultrasonic Welding of Thermoplastics, Polymer Engineering and Science, 23: 726–733.

Potrebbero piacerti anche