Anda di halaman 1dari 19

International Journal of Machine Tools & Manufacture 40 (2000) 119

A dynamic model of the rolling process. Part I:


homogeneous model
Pei-Hua Hu, Kornel F. Ehmann

Department of Mechanical Engineering, Northwestern University, Evanston, IL 60208-3111, USA


Received 5 January 1999; accepted 7 May 1999

Abstract
Mathematical models representing the static rolling process have attracted considerable attention in the
past, resulting in analytical, numerical, or graphical solutions obtained under various degrees of simplification and constraints. Most models, however, can only be used under steady conditions, and therefore are
not suitable for the study of rolling chatter. An enhanced analytical process model that can handle dynamic
variations exerted by roll vibrations in multiple directions is proposed in this paper, and its linearized form
established for easy analysis. Finally, experiments are presented that verify the accuracy of the proposed
dynamic model of the rolling process. 1999 Published by Elsevier Science Ltd. All rights reserved.
Keywords: Flat rolling; Rolling dynamics; Homogeneous deformation

1. Introduction
The most difficult part in modeling chatter in rolling is to establish a well-defined mathematical
model of the rolling process that contains all the complex relationships between different factors
and yet is still solvable. Because chatter vibrations occur as a result of dynamic interactions
between the rolling process and the mill stand structure, a model that presents the rolling force
and torque as a function of operating parameters must be established.
A number of models have been developed during the past few decades in order to better understand the rolling process. An early model developed by von Karman [1] described the pressure
distribution along the arc of contact at the roll/strip interface by a differential equation, which is
based on the assumption that plane sections remain plane during the passage of the strip through
the arc of contact. Orowan [2] discarded the assumption of constant yield stress during the rolling
* Corresponding author.
E-mail address: k-ehmann@nwu.edu (K.F. Ehmann)

0890-6955/00/$ - see front matter. 1999 Published by Elsevier Science Ltd. All rights reserved.
PII: S 0 8 9 0 - 6 9 5 5 ( 9 9 ) 0 0 0 4 9 - 8

P.-H. Hu, K.F. Ehmann / International Journal of Machine Tools & Manufacture 40 (2000) 119

Nomenclature
a..
fx
fy
Gp
h1
hc
k
M
m
p
u1
u2
up
vr
R
x1
x2
xc
xn
yp
s1
s2
ts

Transfer function matrix element


Roll horizontal force per unit width
Roll vertical force per unit width
Transfer function matrix of the rolling process
Strip thickness at entry
Roll gap spacing
Strip shear yield strength
Roll torque per unit width
Friction factor
Roll pressure
Strip velocity at entry
Strip velocity at exit
Linear model input vector
Roll peripheral velocity
Roll radius
Position of the entry plane
Position of the exit plane
Roll horizontal movement
Position of the neutral point
Linear model output vector
Strip tensile stress at entry
Strip tensile stress at exit
Strip shear stress at surface

pass and permitted the direct use of experimentally determined stressstrain curves. He further
assumed that slipping does not take place over the whole arc of contact and that plastic shear
takes place in the rolled strip, while the surface of the strip sticks to the rolls with static
friction. Also, by introducing a complicated adjusting factor into the computation, to have the
inhomogeneity taken into consideration, Orowan was able to relax the assumption of constant
horizontal velocity of the material over any vertical cross-section. To accomplish this comprehensive study, however, Orowan used a tedious and difficult graphicalnumerical method of computation in order to determine the roll pressure distribution.
Bland and Ford [3,4] and Sims [5,6] used simplified assumptions to develop solutions that, by
avoiding most of the numerical integration in Orowans theory, led to analytical expressions of
pressure distribution, roll torque, and roll force. Ford et al. [7] and Bland and Sims [8] made
considerable modifications to the simplified theory of cold rolling developed by Bland and Ford
[3,4] by taking into account the elastic deformation of contact arcs at entry and exit, and by
allowing for a certain degree of variation of the yield stress in the plastic arc of contact.
Venter and Abd-Rabbo [9] developed a comprehensive, self-contained computerized solution
for both the hot and cold rolling problems with the inhomogeneity of deformation embedded into

P.-H. Hu, K.F. Ehmann / International Journal of Machine Tools & Manufacture 40 (2000) 119

the model. Freshwater [10] simplified the equations for roll force, roll torque and roll pressure,
without sacrificing accuracy, from the works by Alexander [11] and Venter and Abd-Rabbo [9].
By eliminating the derivative of the yield stress curve, it was shown that both the homogeneous
or inhomogeneous solutions of the roll force, etc. can be calculated with equal ease even when
the material is rapidly strain hardening, as occurs in the cold rolling processes.
Tlusty et al. [12] advanced the model of rolling process dynamics by taking into account the
rate of change of the exit strip thickness, which was assumed to be the same as the rate of change
of roll gap spacing along the centerline of rolls. Yun [13] and Yun et al. [1416] further relaxed
the assumption that the exit plane coincides with the centerline of rolls, which suggested a new
conservation law of mass and an equation for determining the neutral point within the arc of
contact. Hu [17] discarded the assumption of constant strip exit velocity and the use of forward
and backward slips in velocity calculations, and therefore allowed the model to be more suitable
for chatter studies. He also took into account the roll horizontal movement allowing the effects
of complicated roll vibrations to be considered.
Most of the aforementioned studies fall into the category of static rolling analysis and are not
suitable for chatter analysis. The purpose of this paper is to address this problem by introducing
an enhanced dynamic process model based on the work of Hu [17], and therefore enable a better
study of rolling chatter. Specifically, the objective is to derive the rolling process model in the
form of a transfer function matrix that can be easily combined with a suitable structural transfer
function model of the mill to analyze the stability of the process by conventional linear system
analysis methods.

2. Rolling process model


2.1. Roll bite geometry
The basic rolling process is illustrated in Fig. 1. A strip of width w and thickness h1 enters the
gap between work rolls with velocity u1, and exits with thickness h2 and velocity u2. The radius
of the work roll is R and the roll peripheral velocity is vr. When the roll radius is much larger
than the strip thickness and the w/h ratio is greater than ten, it can be assumed that the exit width
is also w. The plastic deformation thus becomes a plane strain problem. As shown in Fig. 1, the
top work roll may vibrate in both x and y directions with velocities of xc and hc/2, respectively,
while the bottom work roll vibrates in phase along the x direction but opposite in y direction with
the top roll. This assumption is proposed in order to make the deformation process symmetric to
the center plane of the strip and thus reduce the number of degrees of freedom by a factor of 2.
Using a parabolic approximation for the roll surface [13,14], the strip thickness within the roll
bite can be established as:
(xxc)2
h(x)hc
R

(1)

where hc is the roll gap spacing measured along the centerline of the rolls, and xc the distance
between the ever-changing center of the roll and its original position (before vibrations begin).

P.-H. Hu, K.F. Ehmann / International Journal of Machine Tools & Manufacture 40 (2000) 119

Fig. 1. Roll bite geometry during chatter.

Considering the control volume of the material flow within the roll bite as shown in Fig. 2, the
flow of metal through a vertical cross-section at any arbitrary distance x from the original centerline of the rolls may be written as:
(2)
uhu1h1(x1x)hc(h1h)xc
where h is the strip thickness and u the strip velocity of that cross-section, hc is the rate of change
of roll gap spacing along the centerline of the rolls, and xc the horizontal velocity of the vibrating
rolls. The second and third terms on the right hand side of Eq. (2) represent the flow of material
across a moving boundary caused by the vertical and horizontal vibrations of the roll, respectively.
2.2. Yield criterion
The von Mises yield criterion for plane strain conditions, based on the coordinate system shown
in Fig. 3, can be written as:
1
k 2 (sxxsyy)2t2xy
4

(3)

where k is the strip shear yield strength, sxx and syy are the normal stress, and txy the shear stress.

P.-H. Hu, K.F. Ehmann / International Journal of Machine Tools & Manufacture 40 (2000) 119

Fig. 2. Conservation of material.

Fig. 3.

Coordinate system for yield criterion.

Under highly loaded conditions, which can be found frequently in the rolling process, the friction
factor model is considered more suitable than the conventional Coulombs fiction law [18]. The
shear stress at the surface of the strip can then be expressed as:
tsmk, 0m1

(4)

P.-H. Hu, K.F. Ehmann / International Journal of Machine Tools & Manufacture 40 (2000) 119

where m is the friction factor. When homogeneous deformation is considered, the shear stress
effects in Eq. (3) can be neglected and the following simplified yield criterion is obtained:
sxxsyy2k

(5)

2.3. Strip entry and exit positions


The distances measured from the centerline of the rolls to the strip entry plane can be derived
from Eq. (1) as:

x1xc R(h1hc)

(6)

To solve for the position of the strip exit plane x2, a second relationship between it and the
strip exit velocity u2 is needed. From Fig. 2, it can be established that:
x2xc
hc

sin f2
(7)
tan f2
2(u2+xc)
R
for a small angle f2 [17]. Substituting u2 from Eq. (7) into Eq. (2), and neglecting the higher
order term, the following simple expression for x2 can be obtained:
Rhchc
(8)
x2xc
2[u1h1(x1xc)hc+h1xc]

2.4. Stress distribution and neutral point position


Fig. 4 shows a slice of rolled material inside the roll bite. The static force equilibrium equation
in the x direction can be expressed as:

Fig. 4.

Stresses acting on a vertical slice.

P.-H. Hu, K.F. Ehmann / International Journal of Machine Tools & Manufacture 40 (2000) 119

dsx
dh
(psx)h 2ts0
dx
dx

(9)

where the plus/minus is determined by position x. Substituting stresses from Eqs. (4) and (5) into
Eq. (9), and assuming that sx and p are principal stresses, the horizontal stress distribution at any
location within the roll bite can be written as:

sxs1

2(xxc)
2k
m
dx
h
R

x1

(10)

where s1 is the strip horizontal stress at entry. Assuming that m is constant throughout the arc
of contact, the horizontal stress at exit can then be expressed as:

h2
s2s12k ln m
h1


tan1

x1xc

Rhc

tan1

x2xc

Rhc

2 tan1

xnxc

Rhc

(11)

The position of the neutral point within the roll bite can then be solved from Eq. (11):

xnxc Rhc tan

hc
1
h2 s1s2
1 1 x1xc
x2xc
tan1

ln
tan
2
2
2k
Rhc
Rhc
2m R h1

(12)

2.5. Roll forces and torque


From Eq. (5), the roll pressure can be rewritten as:
(13)

p2ksx
Substituting sx from Eq. (10) into the above equation yields:

p(x)2k s1

2(xxc)
2k
m
dx x2xx1
h
R

x1

(14)

It is clear that the roll forces per unit width will then be simply the integral of p and ts within
the roll bite in either the x or y direction. Since the friction force changes sign at the neutral
point, it is necessary to carry out the integration in two steps. The roll forces per unit width in
x and y directions can then be written as:


x1

fx

2k s1

x2

x1

x1

2(xxc)
2k
m
dx tan f dx mk dx
h
R
x2

(15)

P.-H. Hu, K.F. Ehmann / International Journal of Machine Tools & Manufacture 40 (2000) 119


x1


x1

2(xxc)
2k
2k s1
m
dx
h
R

fy

x2

dx mk tan f dx

(16)

x2

x1

where
tan f

xxc

R (xx )
2

Note that the minus sign in fx for the pressure term is added in order to conform with the assumed
xy coordinates. Separating the integration into two sections, x2 to xn and xn to x1, yields:


xn

fxmk(x1x22xn)

2(xxc)
2k
m
dx

h
R
xn

2(xxc)
2k
m
dx
h
R

2k s1

x2
x

xn

x1


xn

tan f dx

2k s1

x1

2(xxc)
2k
m
dx
h
R

x1

tan f dx

(18)

and

xn

fy

xn

x2

x1


xn

2(xxc)
2(xxc)
2k
2k
2ks1
m
dx
m
dx
h
R
h
R
x


x2

2(xxc)
2k
m
dx
h
R

2k s1

x1

xn

x1

x1

dx mk tan f dx mk tan f dx
xn

(19)

xn

It is rather difficult to find a closed-form solution for the force component in the x direction.
To simplify the formulation a similar approximation to the one applied in Eq. (7) will be used,
i.e. the approximation
tan fsin f

(20)

will be adopted for a small angle f. As a result, the roll force in the x direction can be explicitly
calculated as:
s1
h1
fx (h1h2)kh2ln mkh2
2
h2

R
xnxc
x1xc
x2xc
2 tan1
tan1
tan1
hc
Rhc
Rhc
Rhc

(21)

P.-H. Hu, K.F. Ehmann / International Journal of Machine Tools & Manufacture 40 (2000) 119

Under the conditions when the angle f is small, the friction component of the roll force in the y
direction becomes negligible. Therefore, the roll force per unit width in the y direction can be
expressed as:

fy(2ks1)(x2x1)4k Rhc tan1

h (x x ) 2 tan

2mk

xnxc

Rhc

x1xc

Rhc

x1xc

tan1

mkR ln

tan1

Rhc

x2xc

Rhc

tan1

x2xc

Rhc

h1
h2
h1
ln
2k(x2xc)ln
hn
hn
h2

(22)

The torque acting on the rolls caused by friction between the roll and strip can be calculated
by integrating the friction force along the arc of contact. By defining the positive z-axis to point
in the direction that goes out of the paper, by formulating the torque acting on the top roll as:

x1

M Rts ds

xn

x2
2

R mk

R2(xxc)

dx
2

xn

R2mk

R2(xxc)2

dx

(23)

and by carrying out the integration in Eq. (23), the torque per unit width on the top roll becomes:

MmkR2 2 tan1

xn+xc

R2(xnxc)2

tan1

x1+xc

R2(x1xc)2

tan1

x2+xc

R2(x2xc)2

(24)

2.6. Strip entry and exit velocities


It is desirable to eliminate the use of the forward and backward slips when calculating the strip
velocities at entry and exit since such information must be obtained through experiments. However, with the neutral point position determined through force equilibrium, it is rather easy to
formulate the strip velocities without using the forward or backward slips. Knowing that the strip
velocity equals the roll peripheral velocity vr at the neutral point, the strip velocities at entry and
exit can be obtained from Eq. (2) as:
u1

1
(xnxc)2
(vrxc)hc(vrxc)
(x1xn)hch1xc
h1
R

(25)

10

P.-H. Hu, K.F. Ehmann / International Journal of Machine Tools & Manufacture 40 (2000) 119

(x2xc)
u1h1+(x2x1)hc+ h1hc
xc
R
u 2
(x2xc)2
hc+
R
2

(26)

3. Linearized model
Although the dynamic model of the rolling process has been established in the previous section,
it does not facilitate an easy study of the interactions between the process and the structure because
of its nonlinear nature. A linearized model, therefore, is needed for rolling chatter analysis. The
linearization is achieved by applying a first-order Taylor series approximation to the equations
for the rolling process model, and eliminating the nominal value of each variable. As a result,
the process model can then be expressed in terms of the variations of system inputs and outputs.
From the construction of the dynamic rolling process model in the previous section, it is clear
that strip tensions at entry and exit s1 and s2, strip thickness at entry h1, roll gap spacing hc and
its rate of change hc, horizontal displacement of the rolls xc and its rate of change xc, and roll
peripheral velocity vr, should be defined as inputs, while the other variables, which are simply
the results of these parameters, through direct or indirect interactions as outputs. Any linearized
model output, yi, can then be expressed in terms of inputs in the following form:

dyi

yi
s1
yi
xc

ds1

ss

dxc

ss

yi
s2

yi
vr

ds2

ss

dvr

yi
h1

ss

dh1

yi
hc

ss

dhc

yi
hc

yi
dhc
dx
c
ss

dxc

ss

(27)

ss

where the subscript ss designates that the corresponding derivative is evaluated for the nominal
or steady-state conditions.
In order to take advantage of some powerful tools for linear dynamic system analysis, it is
preferable to express the linearized rolling process model in the form of a transfer function matrix.
If the model input vector is defined as:
up[ds1 ds2 dh1 dxc dhc dvr]T

(28)

and the output vector as:


yp[dfx dfy dM du1 du2]T

(29)

then the dynamic model of the rolling process in the form of a transfer function matrix can be
written as:
ypGp(s)up
where

(30)

P.-H. Hu, K.F. Ehmann / International Journal of Machine Tools & Manufacture 40 (2000) 119

afx,1 afx,2 afx,3

afx,4

afx,5+afx,6s

afy,1 afy,2 afy,3

afy,4

afy,5+afy,6s

Gp(s) aM,1 aM,2 aM,3

aM,4

aM,5+aM,6s

11

(31)

au1,1 au1,2 au1,3 au1,4+au1,5s au1,6+au1,7s au1,8


au2,1 au2,2 au2,3 au2,4+au2,5s au2,6+au2,7s au2,8

and the coefficients as are obtained through the linearization process in accordance with Eq.
(27). The particular values of the elements of the transfer function matrix, Eq. (31), are given in
Appendix A.

4. Experimental verification
Experiments were performed on a 5.6 kW Stanat rolling mill in two-high configuration, running
at speeds ranging from about 0.051 m/s to 1.525 m/s. The rolls were 101.6 mm in diameter and
152.4 mm in width, and were made of hardened alloy steel. The material used in the experiments
was 5052-O cold rolled aluminum strip with a thickness of 0.82 mm and width of 31.75 mm.
Fig. 5 shows the schematics of the experimental setup used in the rolling experiments. Dynamic
roll forces were measured using two quartz force rings placed between the top roll chocks and
the mechanical screws on both sides of the roll, along with two piezo-translators that can be used
to generate roll gap variations. A motor assembly was used to provide the dynamic component
of the strip back tension, with a third force ring mounted on it to measure the tension variation.
The strip entry and exit velocities and the roll peripheral velocity were measured individually by
a non-contact laser system. A capacitance probe was mounted to one of the chocks to measure
the roll gap variation.
4.1. Forward and backward slips
Forward and backward slips were calculated using the roll and strip velocity measurements,
and compared with the model predictions calculated using the strip velocities at entry and exit
from Eqs. (25) and (26). The forward and backward slips are defined as:
u2vr
100%
vr

(32)

vru1
100%
Sb
vr

(33)

Sf

respectively. Fig. 6 shows the results of backward slip calculation vs. reduction from both experiments and the proposed model, and Fig. 7 shows the results of forward slip calculations. It can
be seen that both the forward and backward slip predictions match the experimental results
pretty well.
The forward slip values obtained through experimenting show a more pronounced scattering

12

P.-H. Hu, K.F. Ehmann / International Journal of Machine Tools & Manufacture 40 (2000) 119

Fig. 5. Experimental setup for the rolling experiments.

pattern, as can be seen in Fig. 7, especially when the reduction is high. The forward slip is more
sensitive to variations in rolling velocity, tensions, and roll movements, while the backward slip
is rather insensitive to these factors.
4.2. Roll forces and roll gap
In order to verify the dynamic behavior of the proposed process model, external excitation in
the form of roll gap variations generated by the piezo-electric actuators situated in the roll chock
assemblies was provided for the rolling experiments, and the results were compared with simulation results. Fig. 8 shows the experimental result corresponding to a 150 Hz sinusoidal roll gap
disturbance created by the two piezo-translators inserted between the mill frame and the roll
chocks [17]. It can be seen that all the measurements prominently show the 150 Hz mode caused
by the roll gap variation. The backward tension variation appears to lead the roll gap variation

P.-H. Hu, K.F. Ehmann / International Journal of Machine Tools & Manufacture 40 (2000) 119

Fig. 6.

Backward slip vs. reduction.

Fig. 7. Forward slip vs. reduction.

13

14

P.-H. Hu, K.F. Ehmann / International Journal of Machine Tools & Manufacture 40 (2000) 119

Fig. 8. Dynamic rolling experiment with a varying roll gap.

and the roll force variation by about 90, and lags the strip exit velocity variation by about 20.
The roll force shows a peak-to-peak variation of 1000 N and the backward tension a variation
of about 50 N. The roll gap variation is estimated to be about 0.5 m, and the strip exit velocity
variation is about 0.002 m/s.
A simulation based on the linearized process model was carried out under the same conditions,
and is shown in Fig. 9. The simulation predicts a backward tension fluctuation of 36.0 N peakto-peak, a roll force variation of 1047.3 N, a roll gap variation of 1.56 m, and a strip exit
velocity variation of 0.0016 m/s. It is clear that the proposed model gives fairly accurate estimates
for the roll force, backward tension, and strip exit velocity variations. The relatively larger error
for the roll gap variation is partly the result of not being able to completely secure the capacitance
probe in the fixture.
5. Conclusion
A new dynamic model of the rolling process that takes into consideration the influence of
horizontal and vertical movements of the roll has been established, both in the complete
(nonlinear) and in linearized form. Also, a new set of independent parameters was selected to
formulate the real rolling process and to improve its suitability for chatter studies. The process
model was verified by a set of dynamic rolling experiments and has proven to be accurate.

P.-H. Hu, K.F. Ehmann / International Journal of Machine Tools & Manufacture 40 (2000) 119

15

Fig. 9. Simulation of dynamic rolling experiment with a varying roll gap.

Acknowledgements
The support of the National Science Foundation under grant #DDM 9300158 is gratefully
acknowledged. The authors also wish to thank Dr William R.D. Wilson for his contributions to
this work.

Appendix A
The transfer function matrix elements in Eq. (31) are obtained by applying first-order Taylor
approximations to all the output variables defined in Eq. (29). Individual element values are listed
below for reference:
h1
afx,1
2
afx,2

hc
2

(A1)

(A2)

16

P.-H. Hu, K.F. Ehmann / International Journal of Machine Tools & Manufacture 40 (2000) 119

s 1
afx,3
2

(A3)

afx,40

(A4)

afx,5

s 1+s 2 k h1 mk
ln
4
2 hc
2


xn

2 tan1

Rhc

tan1

x1

(A5)

Rhc

afx,60

(A6)

afy,1xnx1

(A7)

afy,2xn

(A8)

k
mR
2
afy,3 (x1xn)
h1
x1

R s 1
h1hc 2 k

(A9)

afy,40

(A10)

afy,52k

h h 2 k h

hc mkR
kx1
h1 hc
Rhc

x1

R 1
hctan

s 1

mkxn
1


x1 xn
2k
h1 hc

x1xn xn hc s 1s 2
mk k ln
h1hc hc h1
2k

(A11)

Rhc
s 1
xn
h1
afy,6 k k ln mk mk
vrhn
2
hc
hc

aM,1


2 tan1

hnR2

hnR2

x1

Rhc

(A12)

(A14)

2 R2x2n

mkR2
aM,3
2h1

Rhc

tan1

(A13)

2 R2x2n

aM,2

xn

R
h1hc

h1

R2x21

hn

R x
2

2
n

khnR2

h1 R2x2n

(A15)

P.-H. Hu, K.F. Ehmann / International Journal of Machine Tools & Manufacture 40 (2000) 119

aM,4mkR 1

mkR2
aM,5
2h1

2
1

R
h1hc

au1,2

h1

R2x21

mkR2 hc
R
2 2
aM,6

2vr hn R xn
au1,1

2R

R x R x
2

(A16)

2
n

hn

R x
2

2
n

hc s 1s 2
xn x1hn hn

ln

2
2k
R2x2n hc h1hc 2mhc h1

mkR2

(A17)

(A18)

vrxnhn
2mkRh1

(A19)

vrxnhn
2mkRh1

17

(A20)

vrhn
xn xn
au1,3 2 1
h1
mR 2x1

(A21)

au1,40

(A22)

hnh1
au1,5
h1

(A23)

vrhn
xn
xn
hc s 1s 2
2
ln
au1,6 1
h1hc
2x1 2mR h1
2k

(A24)

2x1xn
au1,7
2h1

(A25)

hn
au1,8
h1

(A26)

au2,1
au,2,2

vrxnhn
2mkRhc

vrxnhn
2mkRhc

(A27)

(A28)

18

P.-H. Hu, K.F. Ehmann / International Journal of Machine Tools & Manufacture 40 (2000) 119

vrhnxn 1
1
au2,3

h1hc mR 2x1

(A29)

au2,40

(A30)

hnhc
au2,5
hc

(A31)

au2,6

vrhnxn 1 1
hc s 1s 2
2

ln

2
2hc x1 mR h1
2k

(A32)

xn
au2,7
2hc

(A33)

hn
au2,8
hc

(A34)

References
[1] T. Von Karman, Beitrag zur Theorie des Walzvorganges, Zeitschrift fur Angewandte, Mathematik und Mechanik
5 (1925) 139141.
[2] E. Orowan, The calculation of roll pressure in hot and cold flat rolling, Proc. Inst. Mech. Eng. 150 (4) (1943)
140167.
[3] D.R. Bland, H. Ford, The calculation of roll force and torque in cold strip rolling with tensions, Proc. Inst. Mech.
Eng. 159 (1948) 144163.
[4] D.R. Bland, H. Ford, Cold rolling with strip tension Part III: an approximate treatment of the elastic compression of the strip in cold rolling, J. Iron Steel Inst. 171 (1952) 245249.
[5] R.B. Sims, Calculation of roll force and torque in cold rolling by graphical and experimental methods, J. Iron
Steel Inst. 178 (1954) 1934.
[6] R.B. Sims, Calculation of roll force and torque in hot rolling mills, Proc. Inst. Mech. Eng. 168 (1954) 191200.
[7] H. Ford, F. Ellis, D.R. Bland, Cold rolling with strip tension Part I: a new approximate method of calculation
and a comparison with other methods, J. Iron and Steel Inst. 168 (1) (1951) 5772.
[8] D.R. Bland, R.B. Sims, A note on the theory of rolling with tensions, Proc. Inst. Mech. Engrs. 167 (1953) 371372.
[9] R. Venter, A. Abd-Rabbo, Modeling of the rolling process I, Int. J. Mech. Sci. 22 (1980) 8392.
[10] I.J. Freshwater, Simplified theories of flat rolling I: the calculation of roll pressure, roll force and roll torque,
Int. J. Mech. Sci. 38 (6) (1996) 633648.
[11] J.M. Alexander, On the theory of rolling, Proc. R. Soc. London A326 (1972) 535555.
[12] J. Tlusty, S. Critchley, D. Paton, Chatter in cold rolling, Ann. CIRP 31 (1) (1982) 195199.
[13] I.S. Yun, Chatter in rolling, Ph.D. thesis, Northwestern University, Evanston, IL, 1995.
[14] I.S. Yun, W.R.D. Wilson, K.F. Ehmann, Chatter in the strip rolling process, Part I: dynamic model of rolling,
Trans. ASME: J. Manufact. Sci. Eng. 120 (2) (1998) 330336.
[15] I.S. Yun, W.R.D. Wilson, K.F. Ehmann, W.R.D. Wilson, Chatter in the strip rolling process, Part II: dynamic
rolling experiments, Trans. ASME: J. Manufact. Sci. Eng. 120 (2) (1998) 337342.

P.-H. Hu, K.F. Ehmann / International Journal of Machine Tools & Manufacture 40 (2000) 119

19

[16] I.S. Yun, W.R.D. Wilson, K.F. Ehmann, W.R.D. Wilson, Chatter in the strip rolling process, Part III: chatter
model, Trans. ASME: J. Manufact. Sci. Eng. 120 (2) (1998) 343348.
[17] P.H. Hu, Stability and Chatter in Rolling, Ph.D. thesis, Northwestern University, Evanston, IL, 1998.
[18] T. Wanheim, N. Bay, A model for friction in metal forming processes, Ann. CIRP 27 (1978) 189.

Anda mungkin juga menyukai