Você está na página 1de 8

Martensite Formation, Strain Rate Sensitivity, and

Deformation Behavior of Type 304 Stainless Steel Sheet


G.L. H U A N G , D.K. M A T L O C K , and G. KRAUSS
The strain and strain rate dependence of the deformation behavior of Type 304 stainless steel
sheet was evaluated by constant temperature tensile testing in the temperature range of - 8 0 ~
to 160 ~ The strain rate sensitivity, strain hardening rate, and ductility reflected the competition of two strengthening mechanisms: strain-induced transformation of austenite to martensite
and dislocation substructure formation. At low temperatures, the strain rate sensitivity and strain
hardening rate correlated with the strain-induced transformation rate. A maximum in total ductility occurred between 0 ~ and 25 ~ and the contributions of strain rate sensitivity and strain
hardening to independent maxima with temperature of the uniform and post-uniform strains are
discussed.

I.

INTRODUCTION

T H E strain hardening rate of Type 300 series austenitic


stainless steels has long been known to be dependent on
austenite stability and sensitivity to martensite formation
during deformation. ~1,2,3~ Since austenitic stainless steels
are not hardenable by heat treatment, high strain hardening rates in steels with low austenite stability, such as
AISI Type 301 (which contains nominally 17 pct Cr and
7 pct Ni), are exploited to produce significantly higher
cold worked strengths than can be achieved with more
stable steels, such as AISI Type 304 (which contains
nominally 19 pct Cr and 10 pct Ni).
Many austenitic stainless steels are used in sheet form
and are subjected to severe forming operations, most o f
which are limited by tensile ductility. Although the
strengthening benefits of strain-induced martensite are
well known, much less is known about the effects of
deformation-induced martensite on formability of austenitic stainless steels. Formability is a complex measure
of performance which depends on many factors, including strain hardening, strain rate hardening, plastic anisotropy, friction, e t c . The following equation is often
used to quantify, in uniaxial tension, the dependence of
stress on strain (e) and strain rate (~).t4-10j
cr = K ~ m e n

[1]

where m and n are the strain rate sensitivity and strain


hardening exponent, respectively. High values of m and
n are frequently associated with good ductility and formability, t~~ The total tensile strain to failure, a commonly
used measure of ductility, is the sum of the uniform
strain and the post-uniform or necking strain. Therefore,
to completely understand the factors which control the
ductility of austenitic stainless steel sheet, the effects of
metallurgical factors on both components of the total strain
G.L. HUANG, formerly Visiting Scientist, Department of
Metallurgical Engineering, Colorado School of Mines, is Lecturer,
Department of Mechanical Engineering, Xiangtan University, Hunan,
People's Republic of China. D.K. MATLOCK, Charles F. Fogarty
Professor, Department of Metallurgical and Materials Engineering,
and G. KRAUSS, AMAX Foundation Professor and Director,
Advanced Steel Processing and Products Research Center, are with
the Colorado School of Mines, Golden, CO 80401.
Manuscript submitted April 4, 1988.
METALLURGICAL TRANSACTIONS A

must be considered. The magnitude of the uniform strain,


the strain at which the true flow stress equals the true
strain hardening rate, is controlled by a combination of
the initial yield strength and the strain hardening behavior. I41 The magnitude of the post-uniform strain is
controlled by the strain hardening rate, tTl strain rate sensitivity, m, [4'5'7'9] extent of deformation heating in the
neck, 17A1-14] and fracture processes, including void nucleation, tTl In materials with stable microstructures, the
post-uniform strain increases with m. F51 For conditions
where m decreases with an increase in temperature or a
decrease in strain rate, I131an increase in temperature within
the neck due to adiabatic heating results in a marked
decrease in both the post-uniform and total strains. To
date, limited consideration has been given to the formability of sheet materials in which the microstructure
changes during the forming operation. Of particular interest are the Type 300 series austenitic stainless steels
which transform to martensite during deformation.
The effects of stress and strain on deformation-induced
transformation of austenitic stainless steel were identified in the early 1950's. I1,2.3] The extent of transformation in Type 304 stainless steel was shown to depend on
both the deformation temperature and plastic strain. [3,6,15,16]
Figure 1, after Hecker e t a / . , ll6l shows volume percent
martensite v s true plastic strain data in the temperature
range of - 1 8 8 ~ to 50 ~ At a given strain, the extent
of transformation decreases with an increase in temperature, and a saturation level of approximately 90 pct is
observed. The behavior shown in Figure 1 for austenitic
stainless steels is referred to as a "strain-induced" transformation, and the kinetics of this process have been
modeled based on an analysis of shear band intersections. I~5,171The extent of transformation with strain in
a given alloy system depends directly on the content of
the elements which stabilize austenite. For example, an
increase in nickel from 10 to 20 pct was shown to suppress strain-induced transformations in the temperature
range of - 1 9 6 ~ to room temperature. I~8~
Strain-induced martensitic transformations significantly affect the tensile stress-strain behavior of stainless
steels. With a decrease in test temperature (and a corresponding increase in the rate of martensite formation
with strain), the stress-strain curves change from a smooth
VOLUME 20A, JULY 1989--1239

100

-80c

80

9<

-70c 25, ; / / ~ /
5(,5///-lsc

--~

__22L

oc

............

/<....---..........

60

welded to the samples) was less than 10 ~ during uniform deformation.


The majority of the tensile tests were performed at an
engineering strain rate of 1.5 x 10 2 s l ( i . e . , at a crosshead speed of 0.8 m m / s ) to obtain complete stress-strain
curves. Limited data were obtained at strain rates between 1.5 1 0 - 2 s - I and 1.5 x 10 5 s ~. Crosshead
speed change tests ( i . e . , jump tests) were used to determine strain rate sensitivity as calculated by

//7/,'

40

2a
Y
r ~

0.0

//
~

0.2

sac
I

8.4

0.8

0.8

d l n ~r In (~r2/~l)
m - - dln ~
in (e2/el)

1.0

TRUESTRAIN
F i g . 1 - - T h e e f f e c t o f strain a n d t e m p e r a t u r e o n the s t r a i n - i n d u c e d
m a r t e n s i t e v o l u m e fraction f o r m e d in T y p e 3 0 4 stainless steel. 1161D a s h e d
lines are d u e to H e c k e r et al. ,[161solid lines are original d a t a o f A n g e l , TM
a n d the d o t t e d e x t r a p o l a t i o n utilizes O l s o n ' s a n a l y s i s , uSj

parabolic behavior at temperatures greater than 25 ~ to


a sigmoidal shape at low temperatures. The sigmoidal
shaped curves correlate directly to extensive transformation during strain. [3"15'16'19'2~
The purpose of this study was to characterize as completely as possible the factors which control the total
ductility in Type 304 stainless steels. In particular, changes
in strain rate sensitivity and strain hardening behavior
with strain and temperature were evaluated and correlated with uniform and post-uniform strains in order to
determine the conditions which enhance ductility. An
important objective throughout this study was to characterize the effect of strain-induced martensite formation
on deformation behavior. All testing was performed in
isothermal baths in order to minimize heating effects on
strain-induced martensitic transformation. Without this
precaution, Bressanelli and Moskowitz tzn found that the
heating generated by high strain rate testing could completely suppress martensite formation.

II.

EXPERIMENTAL PROCEDURE

Type 304 stainless steel was received in the form of


0.88 mm thick annealed sheet with the following composition (in weight percent): 0.05C, 1.65Mn, 0.66Si,
17.9Cr, 7.7Ni, 0.31Cu, 0.20Mo, 0.10Co, 0.02S, and
0.015P. Note that the Ni and Cr contents are slightly
lower than those specified for AISI Type 304 stainless
steel. ASTM standard E-8 longitudinal tensile samples
were machined with a 50.8 mm gage length and a nominal width of 12.7 mm. The samples were slightly tapered with a 0.075 mm variation in width from the center
to the end of the reduced gage section. All tensile testing
was performed on a floor model Instron equipped with
a special assembly which was mounted to the movable
crosshead and which allowed testing of the samples
completely submerged in heating or cooling fluids. Strains
were measured with a submersible extensometer attached directly to the sample, and tests were performed
in the temperature range of - 8 0 ~ to + 1 6 0 ~
The
temperature control fluid varied with temperature range:
methanol plus liquid nitrogen for - 8 0 ~ to 0 ~ kerosene at 25 ~ and oil for 40 ~ to 160 ~ The maximum temperature rise (as measured with thermocouples
1240

VOLUME 20A, JULY 1989

[21

utilizing the 2 pct extrapolation method described by


Wagoner. [221 In Eq. [2], at the strain of the strain rate
change, the stress (o-) and strain rate (~) pairs before and
after the strain rate change are identified by the subscripts 1 and 2, respectively. Load-displacement data were
recorded on a DEC-computer based data acquisition system. The results were analyzed to obtain engineering
stress-engineering strain, true stress-true strain, and true
strain hardening rate data. Total ductilities were obtained
from direct measurements of a marked gage length after
deformation. Uniform strains were obtained from the
computer stress-strain data at a point where d o ' / d e = or,
and the post-uniform strains were calculated from the
difference between the total and uniform strains. The
volume percent of martensite was measured on selected
samples with a Frrster portable ferrite content meter.
III.

RESULTS

Figure 2 shows the effects of temperature at a constant


engineering strain rate of 0.015 s 1 on the engineering
stress-strain behavior of the 304 stainless steel. With a
decrease in temperature, there is a transition from parabolic to sigmoidal behavior. The observed variation in
stress-strain curve shape with temperature is consistent
with that observed by previous investigators, and the inflections in the low temperature curves identify the formation of strain-induced martensite. [3,15u6]
In Figure 3, the 0.2 pct offset yield strength, ultimate
tensile strength, and total strain to failure derived from

180

-80C

1513

~ - 5 0 C

~.

s~

-z

120

~o

~'~

60

135C

ku
30

304

SS

I. 5 x I O - 2 s -!
0

O.

0.1

0.2

0.3

0.4

0.5

0.6

0.7

ENGINEERING

O.

STRAIN

Fig. 2 - - E n g i n e e r i n g s t r e s s - s t r a i n tensile data f o r T y p e 3 0 4 stainless


steel tested in the t e m p e r a t u r e r a n g e o f - 8 0 ~ to 160 ~ at a n i m p o s e d engineering strain rate of 1.5 10 2 s ~.
METALLURGICAL TRANSACTIONS A

1200

'

1.0

1.5xlO-2s -I to 1.5xlO-3s -I

/~--.-..~
1000 [

304 SS

1.5x10_2s_

0.8

0.024

)c<x

Q--0.05

0.020

~
~

Q-O. IO
o-o. 15

800

600

~UTS

0.6~
c~

~ 0.016

0.4~

400

0.012

0.008
W

0.2

200

0.004

0
-90

I
-60

I
-30

I
30

I
50

I
90

I
120

I
150

0.000

TEMPERATURE C

Fig. 3 - - T h e temperature dependence of the yield strength, ov, ultimate tensile strength, ~rUTS,and total strain to failure, ex, for Type

304 stainlesssteel tested at an engineering strain rate of 1.5 x 10-2 s-<

the curves in Figure 2 are shown as a function of temperature. The ultimate strength decreases significantly
with an increase in temperature, but the 0.2 pet yield
strength is relatively insensitive to temperature. The ductility, as measured by the total elongation, goes through
a maximum between 0 ~ and 25 ~
The results of
Figure 3 are consistent with previous observations. [3,231
The tensile data summarized in Figures 2 and 3 reflect
the effects of two deformation mechanisms: strain-induced
martensite formation and plastic deformation by dislocation motion and substructure formation. With a decrease in temperature, the increase in strength and the
change in stress-strain curve shape reflects the increased
contributions of strain-induced martensite. For the data
in Figure 1, M~30 (i.e., the temperature at which a volume fraction of 50 pet martensite forms at a true strain
of 0.30) was estimated to be approximately 0 ~ r31 At
all strains for temperatures above room temperature and
at high strains for low temperatures, where the rate of
martensite formation with strain is greatly reduced,
uniform deformation is controlled by dislocation substructure formation in the deforming phases.
The strain rate sensitivity was determined as a function of strain, strain rate, and temperature, and a
summary of the results for a strain rate change between
1.5 x 10 2 s 1 to 1.5 x 1 0 - 3 S -1 is shown in Figure 4.
The data in Figure 4 were obtained by cycling at 0.05
engineering strain increments between the indicated strain
rates. These data show that the temperature dependence
of m varies significantly with the strain level at which it
is evaluated. For example, at an engineering strain of
0.05, m exhibits a m a x i m u m at approximately 25 ~ At
strains of 0.10 and 0.15, which show remarkably high
m values at temperatures below - 4 0 ~ m decreases with
an increase in temperature. For strains near the uniform
strain, m goes through a m a x i m u m between 25 ~ and
60 ~
To understand the data shown in Figure 4, the effects
of strain and temperature on the strain rate sensitivity
must be considered in conjunction with an analysis of
the effects of strain and temperature on the martensite
transformation behavior shown in Figure 1. In Figure 5,
the effects of strain on the strain rate sensitivity at ternMETALLURGICAL TRANSACTIONS A

A~// ~r-

~-~___~

0.0
-90

I
-50

I
-90

I
0

I
30

I
60

TEMPERATURE

I
90

I
120

I
150

Fig. 4 - - T h e effect of temperature on strain rate sensitivity values


evaluated at engineering strains of 0.05, 0.1, and 0.15 and near the
uniform sgain. Data were obtained with a strain rate change test cycled
between 1.5 x 10 2s-~ and 1.5 x 10 - 3 s - k

0.027

0.024
E

0.021

o. o18

1.5
/ / ~

-1 to 1.5

-1
304 SS

'.,

~0.015

W
0.012
z 0.009

7
E0. oo8
0.003
O. 000
O.

0.1

0.2

0.3

0.4

0.5

TRUE S T R A I N

Fig. 5 - - T h e strain dependence of the strain rate sensitivity evaluated


at - 8 0 ~
- 5 0 ~ 25 ~ and 135 ~ Note the m a x i m a in the low
temperature data at a strain of approximately 0. l.

peratures of - 8 0 ~ - 5 0 ~ 25 ~ and 135 ~ are shown


for a strain rate change of 1.5 x 10 2s i to 1.5 x
10 -3 s - ' . The four curves, best-fit graphical interpolation of the data, emphasize the significant differences discussed in relation to Figure 4. At - 8 0 ~ and
-50 ~
maxima in m are observed at a strain of 0.1,
while at 25 ~ and 135 ~ m decreases continuously with
increasing strain.
Similar to the stress-strain behavior, the strain rate
sensitivity reflects the effects of martensite transformation during deformation. Figure 6, obtained by differentiating the curves in Figure 1, shows the effects of strain
on the transformation rate at various temperatures. For
temperatures greater than approximately 50 ~ there is
no martensite present; thus, for all strains, the transformation rate is zero. At - 7 0 ~ the transformation rate
goes through a maximum at a strain of approximately
0.15. The maximum peak height decreases in magnitude
and shifts to higher strains with an increase in temperature, until, at room temperature, the transformation rate
is observed to increase only slightly with an increase in
strain.
VOLUME 20A, JULY 1989- 1241

/ ~ 8 0 C
304 SS
i 70 C

w3

-30C

//

z
=

'

\, ~,-15c

OC

0
0.0

I
0.2

O.

I"
0.3

~ ~-

I
0.4

22C

I
0.5

O.

TRUE STRAIN

F i g . 6 - - T h e e f f e c t o f strain a n d t e m p e r a t u r e o n the t r a n s f o r m a t i o n
r a t e , d f / d e , o f s t r a i n - i n d u c e d m a r t e n s i t e . T h e solid lines w e r e d e r i v e d
f r o m A n g e l ' s TM d a t a in Fig. 1, a n d the d a s h e d lines w e r e d e r i v e d f r o m
the c o r r e s p o n d i n g d a t a o f H e c k e r et al. t16]

A comparison of Figures 5 and 6 shows that at low


temperatures ( - 8 0 ~
the peak in the strain rate sensitivity directly mirrors the peak in transformation rate.
The peak in the transformation rate vs strain curve indicates that the martensite transformation dominates at low
temperatures at low strains. Thus, it is suggested that the
observed m values are proportional to d f / d e when the
martensitic transformation dominates. However, at high
strains at low temperatures and for all strains at 25 ~
and 135 ~ m decreases continuously with an increase
in strain.
The combined effects of strain, strain rate, and temperature on the strain rate sensitivity were evaluated, and
representative data for temperatures of 25 ~ and 135 ~
are presented in Figure 7. This figure correlates m, at
selected strains, with the average strain rate during the
strain rate change test and shows that the strain rate sensitivity increases with strain rate for temperatures where
slip deformation dominates. This observation is consistent with previous results obtained on solid solution
strengthened alloy systems, f~~ 3,24,25]

The strain dependence of the strain rate sensitivity is


interrelated with the strain dependence of the strain hardening rate and the stress-strain behavior. Figures 8 through
12 present the true flow stress, true strain hardening
rate, and strain rate sensitivity as a function of strain for
samples deformed at an engineering strain rate of 0.015 s -~
for temperatures of - 8 0 ~ - 5 0 ~ 0 ~ 25 ~ and
135 ~ respectively. In each figure, the data points for
the strain rate sensitivity data, obtained by cycling between a pair of displacement rates, are plotted and reflect measurements made with either increasing or
decreasing strain rates. Strain rate increases were made
at engineering strains of 0.05, 0.15, 0.25 . . . . . while strain
rate decreases were made at engineering strains of 0.1,
0.2, 0.3 . . . . . For temperatures of - 8 0 ~ - 5 0 ~ and
0 ~ all below Ma30, significant martensite formation
with strain occurs.
An analysis of the effects of strain on strain rate sensitivity in Figures 8 through 12 shows systematic changes
in behavior with temperature. At low temperatures, where
significant strain-induced martensitic transformation occurs, the strain rate sensitivity first increases and then
decreases with strain, resulting in well-defined maxima.
The magnitude of the peak decreases with an increase in
temperature. For temperatures greater than 0 ~ the peak
is not observed, and the strain rate sensitivity decreases
with an increase in strain, as shown for the data at 25 ~
and 135 ~ (Figures 11 and 12).
Figures 8, 9, and 10 show that at the lower temperatures, the strain hardening rate goes through a minimum
at low strains and then increases to a m a x i m u m at intermediate strains. At strains just prior to the minimum in
strain hardening, other workers have shown that in 304

5000

0.030

-80C
5000

1. S x l O - 2 s - 1

do/de

0-

~r 4 0 0 0

0.025

0.020 ~

"IJ

O. 016

, ,4

. . . . . . . .

. . . . . . . .

........

0.015

c~ 3 0 0 0

z
w

'03
kkl

O. 014
E

O. 012

z
0.010 <

(11

.....i 25C 6-0. 1

l-~ 2OO0
~'''

>_

!\

+- o. m o

m O. 008
O. 0015

1ooo-

135C Q=O. 2

\o_-

0.005

135C Q=O,3
25C Q'O. 4

~ 0. 004
m

O. 002

0.0
O. 000

10-3

0. I

0.2

0.3

0.4

0.5

O. 000
O.

TRUE STRAIN
i0 -4

i0 -3

STRAIN

iO -2

IO -j

RATE I / s

Fig. 7--The
strain rate d e p e n d e n c e o f the strain r a t e s e n s i t i v i t y at
25 ~ a n d 135 ~

1 2 4 2 - - V O L U M E 20A, JULY 1989

Fig. 8
A c o m p a r i s o n o f the true s t r e s s - t r u e strain a n d true strain
h a r d e n i n g rate d a t a o b t a i n e d at - 8 0 ~ at a n e n g i n e e r i n g strain r a t e
o f 1.5 10 -2 s ~, w i t h the strain rate sensitivity d a t a (filled circles)
o b t a i n e d b y c y c l i n g b e t w e e n 1.5 10 -2 s -I a n d 1.5 10 -3 s -1.

METALLURGICAL TRANSACTIONS A

6000,

0.030

6000

1. S x l O - 2 s - 1

0.025

5000

0.030

-50C
5000

25C
1.5xlO-2s-I

0.025
E

4000

0.020

4000

z
0.015 m

o 3000

3000

0.020

z
w

0.015 m

\
m

z
O. OLO 7

2000

1000

0,0

"-.~

0.005

0. i

0.2

0.3

0.4

0.5

TRUE

z
O. OLO 7

ooo

O. 005

1000

0.000

0.6

0.0

0.1

0.2

0,3

0.4

0.5

0.000
0.6

TRUE STRAIN

STRAIN

Fig. 9 - - A c o m p a r i s o n of the true s t r e s s - t r u e strain and true strain


hardening rate data obtained at - 5 0 ~ at an engineering strain rate
of 1.5 10 z s ~ with the strain rate sensitivity data (filled circles)
obtained by cycling b e t w e e n 1.5 10 ~ s -~ and 1.5 10 -3 s -].

Fig. 1 1 - - A c o m p a r i s o n of the true s t r e s s - t r u e strain and true strain


h a r d e n i n g rate data obtained at 25 ~ at an e n g i n e e r i n g strain rate o f
1.5 x 10 -2 s l w i t h the strain rate s e n s i t i v i t y data (filled circles) obt a i n e d b y c y c l i n g b e t w e e n 1.5 10 2 s , and 1.5 x 10 3 s - 1

stainless s t e e l , 1~71 the strain-induced martensite transformation begins. Therefore, the rapidly decreasing rate of
strain hardening at low strains is consistent with the dynamic strain softening or transformation plasticity which
accompanies deformation-induced martensite. [~5]

At strains beyond those which produce the minimum


in strain hardening rate, the transformation rate and, correspondingly, the volume fraction of martensite increase, resulting in an increase in the strength due to
composite strengthening by continual refinement of the

6000

0.030

6000

0.025

5000

0.030

OC
5000

135C

1.SxlO-2s-1

1.5xlO-2s- 1

0.025

4000

0.020

4ODD

0.020
>

0.015

o 3000

0.015

3000

z
w

'\x@

do/de

O. OLO

w 2000

w 2000

z
O. OLO 7

do/de

m
1000

0
0.0

i
0. I

i
0.2

i
0.3

i
0.4

i
0.5

0.005

1000

O. 0 0 0

0.6

TRUE STRAIN

Fig. I 0 - - A c o m p a r i s o n of the true s t r e s s - t r u e strain and true strain


hardening rate data o b t a i n e d at 0 ~ at an e n g i n e e r i n g strain rate o f
1.5 x 10 -2 s -~ w i t h the strain rate s e n s i t i v i t y data (filled circles) obtained by c y c l i n g b e t w e e n 1.5 10 2 s ~ and 1.5 x 10 3 s ~.

METALLURGICAL TRANSACTIONS A

0.0

0.005

0.1

0.2

0.3

0.4

0.5

0.000
0.6

TRUE STRAIN

Fig. 1 2 - - A c o m p a r i s o n of the true s t r e s s - t r u e strain and true strain


h a r d e n i n g rate data obtained at 135 ~ at an e n g i n e e r i n g strain rate
of 1.5 x 10 -z s ~ w i t h the strain rate s e n s i t i v i t y data (filled circles)
o b t a i n e d by c y c l i n g b e t w e e n 1.5 10 2 s ' and 1.5 10 - 3 s '.

VOLUME 20A, JULY 1989--1243

martensite and austenite mixture. As a result, measured

do-/de values significantly increase with strain. However, as the martensite volume fraction increases to near
saturation for a particular test temperature, the composite microstructure deforms, and dynamic recovery results in a decrease in strain hardening rate with strain.
As the temperature increases, the extent of transformation at a given strain is lower, and the magnitude of the
peak in the strain hardening rate decreases. For temperatures where martensite formation with strain is insignificant, the strain hardening rate continuously decreases
with an increase in strain, as shown in Figures
11 and 12.
Figures 8 through 12 show that there is a systematic
correlation between the appearance of the peaks in the
strain hardening rate and strain rate sensitivity data. Both
peaks decrease in height and shift to higher strains with
increasing temperature. Furthermore, the peak positions
do not coincide, and the strain rate sensitivity peak always occurs at lower strains than for the strain hardening
rate peak.
IV.

DISCUSSION

The results summarized above show that the strain rate


and temperature dependence of the deformation behavior
in Type 304 stainless steel sheet is complex and results
from interactions of several dynamic strengthening
mechanisms. Specifically, deformation is controlled by
the relative contributions of strain hardening due to dislocation processes and strain-induced transformation of
unstable austenite to martensite. Figure 3 shows that a
maximum in ductility developed between 0 ~ and 25 ~
This section considers the reasons for this maximum.
Figure 13 shows again the maximum in the total elongation. Shown, in addition, are the temperature dependence of the components of the total elongation, i.e., the
uniform strain, ev, and the post-uniform strain, epu. Also
included are the final strain rate sensitivity measurements, mu, evaluated at strains just below eu. Both the
uniform strain and post-uniform strain exhibit maxima.
However, the temperature ranges of the maxima do not
coincide with that for the total strain. The uniform strain

0.7

0.014

0.6

0.012

0.5

0.010 >

0.4

0.008

w
0.006

,4J, O. 3
z

wo. 2

0.004 i

O. 1
O. 0

-90

0,002

QP
I

-60

-30

30
60
TEMPERATURE C

90

120

150

O. 0 0 0

F i g . 1 3 - - A c o r r e l a t i o n o f the u n i f o r m , eu, p o s t - u n i f o r m , epu, a n d


total s t r a i n , eT, w i t h the test t e m p e r a t u r e f o r s a m p l e s d e f o r m e d in
t e n s i o n at a n e n g i n e e r i n g strain r a t e o f 1.5 x 10 ~ s '. F o r r e f e r e n c e ,
the strain rate sensitivity data w e r e e v a l u a t e d at the u n i f o r m strain.

1 2 4 4 - - V O L U M E 20A, JULY 1989

peaks at temperatures below the temperature for the


maximum in eT, while the opposite is true for the postuniform strain.
Maximum uniform strains are obtained for conditions
which maximize the strain hardening rate at high strains
while maintaining low flow stresses at low strains. [4] In
the 304 stainless steel considered here, the initial flow
stress (i.e., 0.2 pct offset yield stress shown in Figure 3)
was essentially independent of temperature. Thus, the
observed variation in eu with temperature reflects the
manner in which strength is accumulated with strain. At
low temperatures (e.g., - 8 0 ~ to - 5 0 ~
the rate of
the martensite transformation is very high at low strains,
and the strength increases rapidly due to the composite
strengthening effects of martensite in an austenitic matrix. For microstructures with significant volume fractions of martensite formed early in the deformation
process, the strain hardening rates abruptly decrease with
strain at higher strains (Figures 8 and 9). The high flow
stresses and rapidly decreasing strain hardening rates
combine to produce instability at relatively low strains.
With an increase in temperature to approximately 25 ~
the rate of the martensite transformation at low strains
decreases and transformation continues over a wider strain
range (Figure 6). As a consequence, the lower rate of
increase in flow stress due to martensite formation and
the more gradual decrease in strain hardening rate at high
strains, as shown in Figure 10, combine to produce maximum uniform elongations at temperatures around 0 ~
At deformation temperatures above 0 ~ the straininduced martensitic transformation is largely deferred to
very high strain or does not occur. As a result, the low
strain hardening rates in the absence of martensite formation at high strains produce low uniform elongations,
and ev decreases with an increase in temperature.
During necking, both the local strains and strain rates
significantly increase; thus, the post-uniform elongation
is maximized for conditions which enhance strain hardening and strain rate sensitivity at high strains. In the
304 stainless steel, these conditions produce a maximum
in eeu in the temperature range of 25 ~ to 90 ~ Figure 6
shows that strain-induced martensite formation is deferred to high strain at 25 ~ therefore, the potential for
significant strengthening due to martensitic transformation within the neck exists. Evidence that, in fact, substantial transformation within the neck occurs in the 304
stainless steel of this study is presented in Figure 14.
This figure shows the percent martensite as a function
of distance from the fracture surface for samples isothermally tested at three strain rates at 25 ~ Uniform
deformation produces martensite volume fractions in the
range of 30 to 40 pct. In the highly necked regions, i.e.,
within approximately 10 m m of the fracture surface, the
martensite volume fraction increases sharply to 60 to
70 pct. The dynamic formation of martensite within the
neck increases local strength, stabilizes flow, and results
in high post-uniform strains. Furthermore, the strain rate
sensitivity which was shown to mirror the martensite
transformation rate, Figures 5 and 6, must also be high,
further stabilizing flow and contributing to enhanced postuniform elongation J 5m
At lower temperatures, the transformation is essentially complete prior to necking. Therefore, the benefits
METALLURGICAL TRANSACTIONS A

volume expansion. Second, load in the austenitemartensite composite is redistributed due to the increase
in the volume fraction of the higher strength martensite.
Thus, the direct relationship of strain rate sensitivity to
martensite transformation rate follows from the effects
of strain rate on the dynamically formed microstructures.

100

isothermol in
80

kerosene,

1. S x l O - 3 s - 1

1.SxlO-4s-

1.5xlO-5s

-1

25C

V.
LtJ
1--

O3
Z
W

n,"
<

',

........
\

20

neck~-~--uni?orm stroin

10

20

:30

40

50

DISTANCE FROM FRACTURE, mm


Fig. 1 4 - - T h e effect of strain rate on the position dependence of the
extent of strain-induced martensite formation. Data were obtained with
a F6rster portable ferrite content meter.

of flow stabilization due to martensite formation in the


neck do not exist and post-uniform strains are low, as
shown in Figure 13. At high temperatures, the austenite
is stable at all strains. Thus, the potential for dynamic
strengthening during necking is low and, correspondingly, the post-uniform strain is low.
In summary, the relatively broad peak in total elongation is due to the summation of the uniform and
post-uniform strains. Figure 13 shows that at low temperatures, the total strain is dominated by the uniform
strain. At high temperatures, the contributions of the postuniform strain are increasingly more important.
As a final point, the data included above also show a
direct correlation of the strain dependence of the strain
rate sensitivity with the strain dependence of the martensite transformation rate. An increase in m implies that
on an increase in strain rate, the strength increases significantly. The strength increase can result from an increase in the flow stress to move a constant density of
mobile dislocations at a higher rate or from structural
changes induced by the strain rate change. In metastable
austenitic structures, the volume fraction of martensite
formed under isothermal deformation conditions increases with strain rate, as shown in Figure 14, for the
data obtained in the uniform deformation region, which
is consistent with previous studies, t231 The increase in
martensite volume fraction due to the increment of strain
rate may increase the overall strength by two mechanisms. First, a higher dislocation density may be introduced into the austenite to accommodate the martensite
METALLURGICAL TRANSACTIONS A

CONCLUSIONS

The following conclusions are based on mechanical


testing of a 300 series austenitic stainless steel in conjunction with literature studies of strain-induced martensitic transformation:
I. In the temperature range of - 8 0 ~ to 160 ~ the
deformation behavior results from the competition of
two strengthening mechanisms: strain-induced transformation of unstable austenite to martensite and dislocation interactions during deformation. At low
temperatures, martensite formation dominates, and
the strain rate sensitivity is proportional to the transformation rate. At higher temperatures, slip dominates, and the strain rate sensitivity decreases with an
increase in strain, an observation which is consistent
with most microstructurally stable metal systems.
2. The strain rate sensitivities at low temperatures go
through maxima with strain and correlate with maxima in martensite formation rates. At high temperatures, transformation at low strains is negligible and
no peaks in strain rate sensitivity develop.
3. A m a x i m u m in the total strain, composed of the sum
of the uniform and post-uniform strains, occurred between 0 ~ and 25 ~ Both components of the total
strain peaked with increasing temperature, but the
temperature of maxima differed. A m a x i m u m in the
uniform strain occurred at intermediate temperatures
(approximately 0 ~
where the peak in the martensite transformation rate was delayed to high strains. A
maximum in the post-uniform strain occurred at higher
temperatures, where the high strains associated with
neck formation were sufficient to produce significant
strength increases within the neck due to martensite
formation and strain rate hardening.
ACKNOWLEDGMENTS
The authors acknowledge the support of the Metallurgy
Program, Division of Materials Research, National Science Foundation. One of the authors (GLH) acknowledges the support of Xiangtan University, Hunan, People's
Republic of China.

REFERENCES
1. S.A. Kulin, M. Cohen, and B.L. Averbach: J. Metals, 1952,
vol. 4, pp. 661-68.
2. J.R. Patel and M. Cohen: Acta Metall., 1953, vol. 1,
pp. 531-38.
3. T. Angel: J. Iron Steel Inst. (London), 1954, vol. 177,
pp. 165-74.
4. D.K. Matlock, G. Krauss, and F. Zia-Ebrahimi: in Deformation
Processing and Structure, G. Krauss, ed., ASM, Metals Park,
OH, 1984, pp. 47-87.
5. A.K, Ghosh: Metall. Trans. A, 1977, vol. 8A, pp. 1221-32.
VOLUME 20A, JULY 1989--1245

6. A.K. Ghosh: Trans. ASMEH:J. Eng. Mater. Tech., 1977, vol. 99,
pp. 264-74.
7. K. Chung and R.H. Wagoner: Metall. Trans. A, 1988, vol. 19A,
pp. 293-300.
8. D.V. Wilson: Metall. Tech., 1980, vol. 7, pp. 282-92.
9. D.A. Woodford: Trans. ASM, 1969, vol. 62, pp. 291-93.
10. D.K. Matlock and G. Krauss: in Formabili~ and Metallurgical
Structure, A.K. Sachdev and J.D. Embury, eds., TMS-AIME,
Warrendale, PA, 1987, pp. 33-48.
11. G. Ferron: Mater. Sci. and Eng., 1981, vol. 49, pp. 241-48.
12. K.S. Raghavan and R.H. Wagoner: Metall. Trans. A, 1987,
vol. 18A, pp. 2143-50.
13. R.A. Ayres: Metall. Trans. A, 1985, vol. 16A, pp. 37-43.
14. G.L. Huang, D.K. Matlock, and G. Krauss: Colorado School of
Mines, Golden, CO, unpublished research, 1988.
15. G.B. Olson: in Deformation Processing and Structure, G. Krauss,
ed., ASM, Metals Park, OH, 1984, pp. 391-424.
16. S.S. Hecker, M.G. Stout, K.P. Staudhammer, and J.L. Smith:
Metall. Trans. A, 1982, vol. 13A, pp. 619-26.

1246--VOLUME 20A, JULY 1989

17. P.L. Mangonon, Jr. and G. Thomas: MetaU. Trans., 1970, vol. 1,
pp. 1577-86.
18. G.P. Sanderson and D.T. Llewellyn: J. Iron Steel Inst. (London),
1969, vol. 207, pp. 1129-40.
19. D.C. Ludwigson and J.A. Berger: J. Iron Steel Inst. (London),
1969, vol. 207, pp. 63-69.
20. D. Fahr: Metall. Trans., 1971, vol. 2, pp. 1883-92.
21. J.P. Bressanelli and A. Moskowitz: Trans. ASM, 1966, vol. 59,
pp. 223-39.
22. R.H. Wagoner: Metall. Trans. A, 1984, vol. 15A, pp. 1265-71.
23. I. Tamura and T. Maki: in Toward Improved Toughness and
Ductility, Climax Molybdenum Development Company (Japan),
Ltd., 1971, pp. 183-93.
24. A.K. Sachdev and R.H. Wagoner: in Novel Techniques in Metal
Deformation Testing, R.H. Wagoner, ed., TMS-AIME,
Warrendale, PA, 1983, pp. 285-98.
25. R.H. Wagoner: Scripta Metall., 1981, vol. 15, pp. 1135-37.

METALLURGICALTRANSACTIONS A

Você também pode gostar