Sunteți pe pagina 1din 12

FINAL PAPER

11th International Conference on Nuclear Engineering


Tokyo, Japan, April 20-23,2003

ICONE-11-36257

A REVIEW OF SWAGE AUTOFRETTAGE PROCESS


M. Afzaal Malik & Shahab Khushnood*
College of Electrical & Mechanical Engineering,
National University of Sciences and Technology, Rawalpindi, Pakistan.
E-Mail: mafzmlk@isb.paknet.com.pk, seeshahab@yahoo.com
Fax: 92-51-2824132
Keywords: Swage-autofrettage, thick-walled cylinder,
residual stresses, Bauschinger effect, fatigue lifetime, stress
field, crack growth.

pressure vessels subjected to very high pressures, metal


forming, jet cutting, hydrostatic extrusion chambers,
increasing strength-to-weight ratios, increasing the initial
yield pressure of a plain cylinder and increasing the fatigue
strength of components.
Apart from increasing the bearing capacity of the
vessel, the residual stress near the bore introduced by
autofrettage enhances the vessel's fatigue life. The
compressive stresses at the bore reduce the probability of
crack initiation and slow down the growth of fatigue cracks
[2-4]. In the manufacturing practice, the use of high pressure
to carry out the autofrettage process is complex, slow,
expensive and dangerous. This is because the maximum
pressures are in the range of 1000 to 2000 MPa and so at the
autofrettage pressure the cylinders are close to failure [5].
Swage-autofrettage process was developed by Davidson et
al. [5] as a way to overcome these problems and has been in
active industrial use for over 35 years.
The determination of residual stresses in
autofrettaged cylinder has been considered by many
investigators using different mathematical methods and
material models. Most of the earlier solutions were based on
the assumption that the material behaves elastically on the
release of the autofrettage process [6-12]. The effects of
non-ideal material behavior were not considered. However,
many materials, particularly the quenched and tempered,
low alloy steels generally used for high pressure vessels,
exhibit a significant Bauschinger effect [13], which is the
decrease in the yield strength of a material in compression,
resulting from the prior tensile plastic overload, thereby
reducing the residual hoop stress near the bore, than
predicted by ideal elastic-perfectly plastic solution. There
are other non-ideal behaviors of materials such as the strain
hardening and strain aging, which result in a significant
increase in yield strength.
There have been numerous papers, relating residual
stresses and associated fatigue lifetimes in autofrettaged
thick-walled cylinders. A series of effects need to be
recognized and tackled:

ABSTRACT
Autofrettage is used to introduce advantageous
residual stresses into pressure vessels and to enhance their
fatigue lifetimes. Autofrettage or self-hooping is achieved
by increasing the elastic strength of a cylinder by means of
high pressure through mechanical, hydraulic or thermal
loading. The swage or mechanical autofrettage process is
routinely used in many industries including nuclear industry,
where enhanced strength-to-weight ratio is desired. The
practical realization of the process would normally involve a
careful evaluation of the related modeling, simulation and
experimental details. There are several key issues in this
context, such as Bauschingers effect, measurement of the
level of autofrettage, fatigue life measurements, prediction
of residual stress field, temperature field for simulating
partial autofrettage, change in autofrettage level resulting
from machining, stress intensity/stress concentration and
fatigue crack growth along evacuator holes, localized
autofrettage as a design tool, crack emanating from an
erosion, and the effect of crack length-unevenness. The
current paper reviews these issues comprehensively and
attempts to provide workable design strategies for this
important industrial process.
1.

INTRODUCTION
Thick-walled cylindrical vessels such as boilers,
nuclear reactors and high-pressure containers, are used for a
variety of applications in nuclear, chemical, armament,
power, oil and food industries, where large internal
pressures have to be withstood. In the absence of residual
stresses, cracks usually form at the bore where the hoop
stress developed by the pressure is highest. To prevent such
failure and to increase the pressure-carrying capacity, a
common practice is autofrettage treatment of the cylinder
prior to use [1]. A French artillery officer (Jacob, 1907)
suggested a process for pre-stressing monobloc gun barrels,
termed autofrettage which comes from the French meaning
"self hooping". There are three types of autofrettage. These
are carried out by hydraulic pressurization, by mechanically
pushing an oversized mandrel, or by the pressure of powder
gas. The mechanical autofrettage offers economic and
processing advantages over the other methods. This paper
deals with the mechanical or swage-autofrettage. Typical
applications include; strengthening of gun barrels, design of

Determination of elastic-plastic uniaxial material behavior.


Determination of elastic-plastic uniaxial reverse loading
material behavior as a function of plastic strain during
loading.
Use of the above two points in combination with some yield
criterion to predict elastic-plastic residual stress field.
Determination of residual stress fields after material removal
from inside diameter and/or outside diameter including the
effects of any further plasticity.

Copyright 2003 by JSME

function of plastic strain during loading. It is further shown


that Sachs' experimental method, which involves removing
material from inner diameter may very significantly
overestimate autofrettage residual stresses near the bore.
Badr et al. [27] have further estimated the residual stress in
cross-bores with Baushinger effect inclusion using FEM and
strain energy density in liquid ends of positive displacement
pumps (PDPs). Perl [28] has proposed a simple, yet
improved experimental method for measuring the level of
autofrettage in thick-walled cylinders such as a gun barrel.
Perl and Alperowitz [29] have investigated the effect of
crack length unevenness on stress intensity factors. Jahed
and Dubuy [30] have proposed an axisymmetric method of
elastic-plastic analysis capable of predicting residual stress
field. It is concluded that consideration of dependency of
Baucshinger Effect Factor (BEF) on plastic strain makes
significant changes to residual hoop stress near the bore for
low-level autofrettage. However, this dependency is
insignificant for high level autofrettage. Underwood and
Parker [31] have carried out measurement and analysis of
fatigue life for over-strained tubes with evacuator holes. As
with measured life, the calculated life was significantly
affected by the amount of autofrettage of tubes. Fahmy [32]
has carried out study of the autofrettage of gun barrels,
outlining necessary tests for the autofrettage technology for
production plants. Experiments were done on specimens,
which were subjected to autofrettage by fire-pressure or
gasses. Malik [14] has outlined the salient features of
swage-autofrettage process on a thick-walled cylinder.
Generally the work reported in literature is based on steadystate autofrettage analysis. There have been very few timedependent analyses attempts [15, 17,32].
The purpose of this paper is to review the different
stages of the development of autofrettage design process and
to address various issues related to this process. Based on
the review, it is intended to provide workable design
guidelines for the process designers in this field.

Determination of stress intensity factors for cracks within


such stress fields.
Calculation of fatigue lifetimes as such cracks propagate
under fatigue loading.
Stress concentration and crack growth along evacuator holes
of autofrettaged tubes.
Methods of measuring the autofrettage level.
Effect of crack length unevenness on stress intensity factors.
Crack emanating from erosion.
Localized autofrettage as a design tool for fatigue
improvement of cross-bored cylinders.

The exact solutions such as those based on Lam model


are very restrictive. The experimental methods are not only
very expensive but also lack the exact and detailed
determination of the elastic-plastic interface. This
necessitates a detailed numerical simulation of the process.
Malik [14-15], Ahmed [16], Owen and Hinton [17], Perl &
Arone [18]. Levy et al. [19], Segall et al. [20], Perl [21,22]
to name some, and others have carried out modeling and
simulation of the process.
Since autofrettage involves partial or full plastic
deformation of the cylinder, any computation of stress under
pressure and, hence, of post pressurization residual stresses,
depends upon the assumed yield criterion. The most
commonly used criteria are those named after Tresca and
von Mises [23]. It has been found that Tresca yield criterion
is about 15% more conservative than the von Mises
criterion about the material behavior [14]. Many suggested
solutions of the problems of autofrettage assume that
Tresca's yield criterion prevails.
Following is a summary of most recent research efforts
related to swage-autofrettage. Perl [21] has developed an
analytical model for predicting the level of autofrettage
following inner, outer or combined machining of a gun
barrel based on Hill's solution for the autofrettage residual
stress field. An FEM analysis of machining process is
performed in parallel in which residual stress field is
simulated by an equivalent thermal load. Segall et al. [20]
launched an investigation into the feasibility of improving
the fatigue life of thick-walled cylinders with cross-bores by
using a localized autofrettage technique as a design tool. In
addition to prolonging the useful life of the cylinders, the
localized residual stresses were shown to be possible at
pressures below the yield threshold for the thick-walled
cylinder. Levy et al.[19]. have investigated erosion
geometry effects on the mode 1 stress intensity factor (SIF)
for a crack emanating from the erosion's deepest point in an
autofrettaged pressurized, thick-walled cylinder using FEM
method and knowledge of asymptotic behavior of short
cracks. Autofrettage based on von Mises yield criterion is
simulated by thermal loading. Parker and Underwood [24]
have described the effect of autofrettage process on fatigue
lifetime of axial residual stresses and have demonstrated that
an insignificant reduction in lifetime results from the
presence of such stresses. Badr et al. [25] have evaluated the
autofrettage effect on fatigue lives of steel with cross-bores
using a statistical and strain based method. Results of the
statistical methodology correlated with results of the reverse
plasticity criterion, showed the existence of an optimal
autofrettage presence. Parker et al. [26] have presented
Baushinger effect design procedure which encompasses
representation of elastic-plastic uniaxial loading materiel
behavior and of reverse loading material behavior as

2.

THE IDEAL MODEL


The general differential equations of deformation
of a deformable body in rectangular coordinates are given
by:
xx yx zx
+
+
+ Bx = 0
x
y
z
xy
x

yy
y

zy
z

(1)

+ By = 0

xz yz zz
+
+
+ Bz = 0
x
y
z

where

xx , yy , zz , yx = xy , zx = xz & yz = zy

are

the

stress components in x,y and z direction respectively and Bx,


By & Bz the body forces in three respective directions.
Specialization of these general equations to cylindrical
coordinates, as given by Boresi and Chong [33], and Boresi
et al [34] is shown in figure 1.
rr 1 r zr rr 0
+
+
+
+ Br = 0
r
r
r
z
r 1 z 2r
+
+
+
+ B = 0
r
r
r
z

(2)

rz 1 z zz rz
+
+
+
+ Bz = 0
r
r
z
r

Copyright 2003 by JSME

cylinder. In a cylinder, which is only partially autofrettaged,


the stress field is given by the following equations, [28].

These equations can be used to derive load stress


and load deflection relations for thick-walled cylinders.
Neglecting the body force components, the equations of
equilibrium (2) for cylindrical coordinates reduce to

rd rr = rr
dr

d
(rrr ) =
dr

or

0
3
rr =
0

(3)

and the strain compatibility condition for the thick-walled


cylinder is given by

rd = rr
dr

or

d
(r ) = rr
dr

1
r 2
1
2 ln
Q
, a r
2
2
2

b
r
b

1
1
(2 Q)
, r b

r2
b

0
3
=

0
3

(4)

(8)

2 ln r Q 1 1 , a r
2
2

b2
r
b

1
1
, r b
( 2 Q)

2
2
r
b

(9)

r = 0

where is the radius of the elastic-plastic boundary (the


autofrettage interface) and Q is given by.
Q=

a 2b2 2

1
+ 2 ln
a
b 2 a 2 b 2

The level of autofrettage is defined as the


percentage of the cylinder's wall that underwent plastic
deformation, i.e.,
a
(10)
=
x100 percent
ba

Figure 1. Stresses in thick-walled cylinder

Note that for = b( = 100 percent) , equation (7)


becomes identical to equation (8).

For the material of cylinder taken to be isotropic


and linearly elastic we have the following stress-strain
relationships.
1
[ rr ( + zz )]
E
1
= [ ( rr + zz )]
E
1
zz = [ zz ( rr + )]
E

3.

EFFECT OF MATERIAL PROPERTIES ON


STRAIN-TO-FAILURE
Material properties ranging form ductile to brittle
by heat treating the specimens have been studied [37]. It has
been shown that strain-to-failure model is a function of the
triaxiality of stress in the critical region of failure. Rice and
Tracey [38], Hancock and Mackenzie [39] and Soo Hoo et
al. [40] have shown that a strain-to-failure model is a
function of the triaxiality of stress in the critical region of
failure. This dependence is expected to be greater where the
stress triaxiality is more significant. A strain-to-failure
criterion can be extended to assess material damage and the
critical enlargement of voids in a low-cycle fatigue
applications. From the uniaxial stress-strain model, effective
or deviatoric stress-strain components are given by
1
(11)
[(1 2 )2 + (2 3)2 + (3 1)2 ]1/ 2
=

rr =

(5)

where is Poisson's ratio and E is Young's modulus of


elasticity.
For internal pressure only, the familiar Lam
equations [35] based on assumptions regarding the stressstrain relationship, yield criterion, end conditions,
compressibility, and incompressibility are given by
equations below.
Pia 2
1
b a 2

b2

r 2

P a2
1 +
= i
2
b a 2

b2

r 2

rr =

(6)

making the assumptions stated above. Since

b2
r2

1 , rr

(12)

For plane strain conditions in long uniform


cylinders, and a negligible volume change, the effective
plastic strain in a material element undergoing radial growth
can be written as
2 u
(13)
p =
ln 1 +

The analytical solution cannot be found without


is

always compressive and maximum at r=a , and is


always tensile stress, maximum at r = a.
The plane-strain stress distribution in a fully
autofrettaged cylinder using von Mises yield criterion, and
the incompressibility condition is given for the full
autofrettage case by Hill [36].
20 r
a 2 b 2 b
ln
1
ln
3 b b 2 a 2
r 2 a
20
r
a 2 b 2 b
1 + ln
1+
ln
=
b b2 a 2
3
r 2 a

r = 0

1/ 2
2
(1 2 )2 + ( 2 3 ) 2 + (3 1)2
=
3

where r is the original position and u is the radial


displacement. Effective stress is developed from equation
(11) using the insight provided by Crossland [41], i.e., plane
strain implies that the axial stress deviator is zero or axial
stress z is the mean of r and , the effective stress
becomes

rr =

(7)

3
( r )
2

(14)

Equilibrium in the radial direction can be expressed


as

where a and b are the cylinders inner and outer radii,


respectively, and 0 is the yield stress of the material of

d r r
+
=0
dR R

(15)

Copyright 2003 by JSME

where R is the current radius of the material (R=r+u).


Substitution of equation (14) into equation (15) and
integrating, gives relation between internal pressure and
effective stress
2 c+ u c
(16)
P=
a + u dR
3

(upper limit)

(20)

Marin & Weng Formula

Marin & Weng formula [44] states that for


isotropic materials which have strain hardening relation.
(21)
= K n + y
the bursting pressure can be written as
n

Pcr = u
2n

2 ln D
2n

3
D 1

D 1

iv)

The ASME Code

The ASME Boiler and Pressure vessel code Sec


viii Div. 2, calculates the bursting pressure with Pcr = u ln D .
Tables 1 and 2 give test results of sixteen
aluminum thick-walled cylinders of various D/t ratios with
material properties ranging from ductile to brittle. The test
results [37] have been compared with various theoretical
and empirical formulas which predict maximum pressures.
In particular, strain-to-failure approach [37] based on
triaxiality of stress condition at failure point (inside
diameter) of the vessel and requires complete uniaxial true
stress-strain data up to ultimate strain. Strain-to- failure
model predicts results closer to strain gage data due to
inbuilt triaxiality factor. The ASME and mean diameter
results are close but exaggerated. Marin and Weng's formula
over predicts the burst pressure.

(18)

where D is the outside to inside diameter ratio and y is the


yield stress.

Table 1. Comparison between predicted and experimental burst pressure for 6063 aluminum
D/t
Condition
1
2
3
4
u
Faupel lower Eq.
Mean
Faupel Upper
Strain- to -failure
y
Diameter Eq.
P
Eq. (19) P3
(18) P2
theory [43] 1
Pb
Pb
(20) P4
Pb
Pb

6
6
6
6

T6
T5
T4
0

8
8
8
8

T6
T5
T4
0

(22)

where n = u (strain hardening exponent) and K is strength


coefficient.

Faupel [43] assumes the cylinder to be a von Mises


material and governs the plastic flow by the distortion
energy criterion. Upper and lower bursting pressures are
defined by equations given below.
ln D

Mean Diameter Formula

iii)

Faupel Equation

(19)

D 1
Pcr = 2u

D + 1

(17)
where the constant Z is quantified by using the uniaxial
tensile test data. Failure of the thick-walled cylinder is
determined by comparing the critical strain c with the
effective plastic strain at each iteration, where m is the
mean stress.
The burst pressure provides an important basis for
the design of thick-walled pressure vessels, it is of interest to
predict this pressure using theoretical models. Following are
some of the burst pressure equations as a result of various
investigations.

ln D 2

The mean diameter formula gives ultimate pressure


explicitly in terms of ultimate stress and diameter ratio.

3
( m / )

Pcr = 2

ii)

c = Ze 2

(lower limit)

is this ultimate stress at fracture.


Equation (18) shows that a cylinder with adequate
ductility could never burst at a pressure lower than that
defined by this equation. At upper limit, the ultimate
strength of the material is involved, given by equation (19).

where

where ua and uc are the material displacements at the inside


(a) and outside (c) radius respectively.
A computer algorithm was developed by Priddy
and Roach [42] to calculate the stress-strain state and the
resulting pressure in a thick-walled cylinder. The
computation is driven by indexing an incremental
displacement at the inside radius. Effective plastic strain is
calculated using equation (13) and is then found from the
bilinear true stress-strain model. The critical value of strain
for each combination of three-dimensional stress is
determined using the stress triaxiality factor ( m / ) and the
exponential equation below

i)

Pcr = 2

5
Marin & Weng
Eq. (22) P5
Pb

6
ASME
code P6
Pb

116.4
147.4
47.7
55.5
Mean

1.01
1.05
0.97
1.16
1.05

0.72
0.70
0.3
0.51
0.5

1.1.3
1.12
0.69
0.90
0.96

1.41
1.50
1.70
1.82
1.60

1.24
1.31
1.41
1.56
1.38

1.43
1.52
1.72
1.84
1.63

116.4
147.4
487.7
555.
Mean
Mean for 6063 Aluminum

1.04
1.05
1.29
1.24
1.15
1.10

0.74
0.70
0.50
0.44
0.59
0.59

0.06
0.00
0.91
0.95
1.03
0.99

1.46
1.50
2.25
1.94
1.79
1.70

1.28
1.31
1.85
1.66
1.53
1.46

1.47
1.51
2.25
1.96
1.80
1.72

Copyright 2003 by JSME

Table 2. Comparison between predicted and experimental burst pressure for two steels (D/t = 8)
Material
Heat treat
1
2-3
4
u
(0F)
Strain-to-failure
Faupel Eq. (18) or
Mean Diameter
y
theory [43]
1045
1045
1045
1045

1200
1000
800
As
quenched

4330
4330

1000
600

P1
Pb

(19),

P2
Pb

or

P3
Pb

Eq. (20)

5
Marin & Weng

P4
Pb

Eq. (22)

6
ASME
code. P6

P5

Pb

Pb

65.9
37.4
20.7
1.0

1.07
1.18
1.39
0.94

1.12
1.20
1.42
0.96

0.93
1.03
1.22
0.83

0.83
0.94
1.14
0.82

0.94
1.04
1.23
0.83

Mean

1.15

1.18

1.00

0.93

1.01

33.3
26.1
Mean

0.93
0.88
0.91
1.03

0.95
0.90
0.93
1.06

0.81
0.77
0.79
0.90

0.74
0.72
0.73
0.83

0.82
0.78
0.80
0.83

Mean for steel

It is concluded that triaxiality of stress manifests


itself most strongly in thick-walled cylinders and the
bursting mode of failure and pressure expansion relations
are important design considerations.

relation for complete loading and reversed-loading profile


as a function of percentage plastic strain.

4.

+ 1.1619 + 0.5975 exp ( 2.175 P ) c

c
= 0.064113 + 0.34816 exp .(12.871 P )
y

[
]
[0.54959 + 0.74913 exp ( 9.6376 ) ]

BAUSCHINGER EFFECT
In terms of material modeling, a complexity that
is often ignored is the Bauschinger effect [13,26,45]
characterized by the Bauschinger effect factor[45], given
as
BEF =

where

yrev
y
y

<1

is the initial yield stress and

where

(23)
yrev is

respectively,

and
P

c are

(24)

2
c

compressive stress and strain

is the plastic strain, and

is the uniaxial

yield strength.

the reverse

yield stress.
Parker et al. [26] define BEF as the ratio of
magnitude of true yield strength in compression to yield
strength in tension. BEF is a function of prior plastic strain
and is an important factor to accurately predict the residual
stresses produced by overload in typical pressure vessel
steels. Bauschinger effect is not a constant material
property, but is strongly influenced by the magnitude of
loading overstrain. In actual engineering geometries,
different locations experience different over-strain levels,
therefore, a different BEF is required [27]. Hence defining
the appropriate BEF value as a function of over-strain
magnitude is an important consideration. Figure 2 shows
schematic representation of Bauschinger effect of a typical
gun steel, the results of the models of Jahed and Dubey
[30], Chen [45], Parker et al. [12, 46] and Chaaban et al.
[47]. Following is the outline of stress-strain profiles.
(A-B), a near-linear relationship during elastic loading
of slope E (Young's modulus);
(B-C1-C2-C3), a linear or nonlinear behavior during
loading beyond the elastic limit, B;
(C1-D1,C2-D2,C3-D3), a linear behavior during
reversal of the tensile load followed by an increasingly nonlinear behavior as the loading becomes more compressive.

Figure 2. Schematic representation of Bauschinger effect and bilinear


models [26]

Modeling the type of equation (24) has been


attempted by several researches including Parker et al.
[12], Chen [45], Chaaban et al. [47], Jahad & Dubey [30]
and Megahed & Abbas [49]. All researchers have used
bilinear representation of the unloading phase shown in
figure 2, except Megahed and Abbas [49] who have
attempted to follow a full non-linear unloading profile
which is function of P . Analysis requires a specific

Milligan et al. [13] have characterized the


Bauchinger effect in certain steels in the form of BEF for a
range of offsets (0.05, 0.10, 0.20 and 0.25%) resulting in a
series of hatched lines (figure 2) varying from unity at
plastic strain, drop rapidly with increasing plastic strain
and become flat at approximately 2% plastic strain and
remain constant thereafter. Kendall [48] has used Milligan
et al. [13] BEF data to extend the validity of the following

Copyright 2003 by JSME

analytical form which is unacceptable for a range of


materials, in particular martensitic steel [26].
5.
INFLUENCE OF BAUSCHINGER EFFECT
ON FATIGUE LIFETIME
The impact of Bauschinger effect on the fatigue
lifetime can be predicted for cyclically pressurized,
autofrettaged tubes containing pre-existing crack-like
defects. Assuming a tube of standard geometry and
pressurization, the positive cyclic range of hoop stress R at
the bore is given by [26].
5 y
(25)
+ A
R =

6.

SACHS' METHOD
An experimental technique called Sach's method
[53,54,55,56] was developed to determine the residual
stresses in an unloaded cylinder. This technique requires
that the cylinder be destroyed by machining. Tiara,
Koterazawa and Ohtani [57] have employed this technique
to determine the residual stresses in closed-ended
cylinders that had been loaded at elevated temperatures
and had deformed inelastically by creep. The method
assumes that the material behaves elastically during and
after material removal carried in a series of steps. The o.d
axial and hoop strains are measured and the differences are
then used to compute radial stress at the prospective radius
prior to material removal. Generally there are four possible
scenarios regarding the material removal from i.d. and/ or
o.d. [26].

3 3

where 5/3 term comes from Lam equations (6) and A is


the value for bore hoop stress arising from autofrettage
process. Approximate fatigue lifetime ratio (actual life
time with Bauschinger effect/ ideal lifetime) is given by
[26].

Removal of material from o.d.


Removal of material from i.d. without Bauschinger effect.
Removal of material from i.d. with Bauschinger effect.
Removal of material from o.d. followed by removal of material from
i.d.

3
Lifetime_ratio = 5 . y + A + y / 5 . y + IDEAL + y (26)

3 3

3 3 3

where IDEAL refers to the residual hoop stress obtained


from the ideal [BEF=1] solution, the term y / 3 arises

Parker et al. [26] have used a nonlinear numerical


model to simulate Sachs' experimental process, for testing
the hypothesis of elastic behavior (using 80% autofrettage
for tube of standard geometry), with Bauschinger effect
included i.e., Bauschinger affected autofrettaged tube
(BAAT) and material removal from bore or i.d. The results
shown in figure 5 indicate over-estimation in residual
hoop stress evaluation via Sachs' method. The near-bore
region shows more over-estimation and would produce
serious errors in fatigue lifetime estimates in equation
(26).

because bore pressure infiltrates and acts upon the crack


surface and the exponent 3 is typical of the exponent in the
Paris and Endogan [50] fatigue crack growth law. Figure 3
is a typical plot showing the ratios actual lifetime/lifetime
with zero autofrettage and ideal life time/lifetime with
zero autofrettage [26]. More sophisticated lifetime
predictions [46] based on stress intensity factor solutions
of extremely high accuracy (errors < 0.5 percent)
determined by the modified mapping collection technique
[51] and packaged as weight function data [52] are shown
in figure 4 which gives a similar behavior as predicted by
simpler formulation of figure 3.

Figure 5. Over-estimates in Sachs' Prediction arising from Bauschinger


effect

Parker et al. [26] further use numerical simulation


for comparison with expected Sachs' prediction and results
of reported Sachs' prediction by Davidson et al. [56] for a
54% swage-autofrettaged, over-strain are given in figure 6.

Figure 3. Approximate fatigue lifetime ratios

Figure 4. Predicted lifetimes as a function of percentage autofrettage


based upon accurate stress intensity data.

Figure 6. Overestimates in Sachs' predictions from Bauschinger effect:


comparison with experimental results.

Copyright 2003 by JSME

field to validate his procedure. Chen's residual stress field


was numerically evaluated by finite difference technique
for various percentages of autofrettage. Table 3 gives the
equivalent temperature field distributions for Chen's [11]
and Hill's [6] autofrettaged stress fields using Perl's [22]
procedure. Chen assumed strain hardening material,
whereas Hill assumed elastic- perfectly plastic material.

7.

TEMPERATURE FIELD FOR


SIMULATING PARTIAL AUTOFRETTAGE
One way to simulate residual stresses is through
an equivalent active thermal loading. Parker et al. [58]
have given analytical derivation of an axisymmatric
temperature field which is equivalent to autofrettage stress
field given by Hill [6]. Pu and Hussain [59] have
implemented this temperature field and have demonstrated
through finite element analysis of a cracked partially
autofrettaged tube that this thermal loading actually
induces the residual stresses due to autofrettage and
simulates their redistribution once the tube cracks.
Timoshenko [60] gives the resulting in-plane components
of the stress tensor for an applied axisymmetric
temperature field T(r) on a thick-walled cylinder as given
below.
rr (r ) =

E 1
.
1 r2

r2 a2

ab T(r )rdr ar T(r )rdr


b 2 a 2

(27)

(r ) =

E 1
.
1 r2

r2 + a2

ab T(r )rdr + ar T(r )r 2


2
2
b a

(28)

Table 3. [22], The equivalent temperature field distribution for


Chen's [11] and Hill's [6] autofrettage residual stress
fields.

ri
1.0
1.1
1.2
1.3
1.4
1.5
1.6
1.7
1.8
1.9
2.0

where is the coefficient of thermal expansion. Based on


Perl's [22] algorithm, the following expression is given for
the temperature field at any given radius ri.
1
(29)
[ (a ) rr (ri ) (ri )]
T (ri ) =

Tc(ri) 0F Chen's
Solution [11]
0.00
-112.67
-215.64
-310.44
-398.14
-479.49
-555.01
-625.01
-689.77
-749.4
-804.37

TH (ri), 0F Hill's
[6] solution
000
-128.40
-245.61
-353.44
-453.28
-546.22
-633.16
-714.84
-719.84
-864.67
-933.77

TC/TH
-0.877
0.878
0.878
0.878
0.878
0.877
0.874
0.871
0.867
0.861

The summary of results of radial distribution of


the induced rr for Chen [11] and Hill [6] solutions as
evaluated by finite element analysis of Perl [22] is given in
table 4. It can be seen that the equivalent temperature
fields closely reproduced original stress fields.

This algorithm is applied by Perl [22] to Chen's


[11] autofrettage stress field for an equivalent temperature

Table 4. The radial distribution of induced rr for Chen [11] and Hill [6] solution as evaluated by FEA
Material
ri/ Stress

Strain-hardening material (Chen [11]


Finite elements
Original field
1
[11] (2rr / 0 )
( / )
rr

1.0
1.1
1.2
1.3
1.4
1.5
1.6
1.7
1.8
1.9
2.0

(1rr / 2rr )

-0.0604
-0.0546
-0.0931
-0.1101
-0.1125
-0.1049
-0.0905
-0.0714
-0.0491
-0.0247
0.0219

0.0000
-0.0654
-0.1005
-0.1153
-0.1163
-0.1078
-0.0927
-0.0731
-0.0505
-0.258
0.0000

-0.83
0.93
0.95
0.97
0.97
0.98
0.98
0.97
0.96
--

-0.0693
-0.0627
-0.1069
-0.1266
-0.1296
-0.1212
-0.1049
-0.0831
-0.0573
-0.0289
-0.0258

0.0000
-0.0752
-0.1156
-0.1327
-0.1342
-0.1247
-0.1076
-0.852
-.0591
-0.0304
0.0000

-0.83
0.92
0.95
0.97
0.97
0.97
0.98
0.97
0.95
--

Milbradt [72]. The destructive approach includes SachsMesnager boring: Sachs [53] and Mesnager [73], Slitling:
Williams et al. [74] and the compliance method: Cheng
and Fannie [75].
Sachs-Mesnager method, though
reliable but has both theoretical and practical limitations
and necessitates the removal of large amount of material
which makes it time-consuming and very expensive[18].

8.

MEASUREMENT OF THE LEVEL OF


AUTOFRETTAGE
Perl and Arone [18] have reviewed most of the
experimental techniques for measuring residual stress
field. Various techniques have been developed for
measuring residual stress, which can be classified as.
1.
2.
3.

Elasto-perfectly plastic material Hill [6]


Finite elements
Original field
(1rr/2rr)
1
2
( rr / 0 )
[11] ( rr / 0 )

Non-destructive method
Semi-destructive method
Destructive method

i)

An Axisymmmetric
Numerical simulation

Stress

release

method:

Perl [28] & Perl and Arone [18, 76] have recently
attempted to measure the autofrettage level in thick-walled
cylinders considering the mentioned limitations. Perl and
Arone [18] have suggested an experimental method, based
on measuring the hoop strain, while axisymmetrically
releasing the residual stress field. An array of seven evenly
spaced inner cuts are introduced into the model. A
parametric study of the proposed experimental procedure

The non-destructive procedures include X-ray diffraction:


Hughes [61] & Clark [62], ultrasonic method: Mckannon
[63], magnetic method: Abuko and Cullity [64], neutron
diffraction: Allen et al. [65] and Stacey et al. [66], and the
indentation techniques: Opel [67] & Underwood [68]. The
semi-destructive techniques such as the hole drilling
method: Mathar [69], Beaney and Proctor [70], blind hole
method: Rendler and Vigness [71] and the ring method:

Copyright 2003 by JSME

via finite element simulation yields an empirical relation


given below, which readily enables determination of the
actual autofrettage level from the strain measurements.
1
x 100
=
1

F ( , ) =

(30)

where

( = / a ) .
2

Rearranging equation (31) yields


(32)

2 ln D = 0

where
D =1+

3 1 2 E0
.
2 1 2 0

Effect of crack length unevenness on stress intensity factors

(33)

Pu [78] was the first researcher to estimate the


SIFs for radial cracks of unequal depths emanating from
the bore of a partially autofrettaged cylinder by using
"dominant crack" approach, which models unevenness by
assuming that all cracks in the array are of equal depth
except for one which is longer. According to Perl and
Alprowitz [29], the approach is useful in terminal phase of
fatigue crack growth, which is governed by a small
number of largest cracks, and the final failure which is
usually caused by one major crack, the "dominant" one. In
the initial stages of cylinders' fatigue life, the array
consists of several uniform sub-arrays with equal depth of
cracks. Perl et al. [79], Arone and Perl [80] have suggested
a "two-crack length level model." Perl and Alperwitz [29]
have also used this 'two-crack length level model, and
have demonstrated that unevenness in SIFs depends on
three parameters, i.e., the number of cracks in the array,
the cracks' lengths, and the level of autofrettage, while
interaction range between adjacent cracks is determined
only by the relative length of the cracks and the density of
the array. They have evaluated SIFs by finite element
method using singular elements in the vicinity of the crack
tip. Modeling detail is given in Perl et al. [79]. Under
certain operating condition in a gun barrel, a longitudinal
erosion may develop at the bore of a cylinder early in its
fatigue lifetime. Becker [81] and Parker [82] have
evaluated SIFs of such eroded, cracked autofrettaged
cylinder using boundary element method, employing a
combined semi-circular erosion with a straight fronted or
elliptical crack emanating from the deepest point of the
erosion Results of Levy et al. [19] for normalized stress
intensity factors (KIP) due to internal pressure (p) only and
due to combined pressure and autofrettage for a 5-percent
eroded and a non-eroded cracked cylinder are presented in
figures 8 and 9 respectively.

The strain release curve can be approximated by a


linear function of the form

( )

=A l +B
a

(34)

Constants A and B are determined by a trial and


error linear regression of the autofrettage level estimates
based on equation (31).

l
(35)
= 2.65133 0.11305
0

where is the released strain and l is the cut depth.


ii)

An Improved Split-Ring Method

Perl [28] has suggested a new method based on


measuring the released hoop stress resulting from the
splitting of the ring specimen shown in figure 7 below.

Figure 7. Evaluation of the stress field in the "free" split-ring by


superposition: (a) the intact ring specimen, (b) the curved
beam, (c) the "free" split-ring

This method is a combination of classical splitring approach by Parker et al. [77] and the recent
axisymmetric stress release method by Perl and Arone
[18] for measuring autofrettage level in gun barrels. The
amount of hoop strain released at point A & B is given
by

( )A = 4M() (b 2 a 2 ) 1
NE

(36)

b
ln
a

(37)

2b 2

b2 a 2

( )B = 4M() (b 2 a 2 )1
NE

b
ln
a
b a

2b 2

( 2 1) 2 4 2 (ln ) 2

CRACK GROWTH IN AUTOFRETTAGED


VESSELS
The combination of manufacturing flaws or
defects at the bore, high-pressure pulses, large temperature
gradients, corrosive environment, as well as stress
concentration cause a gradual degradation, with large
arrays of radial cracks emanating from the cylinders'
internal surface, leading to failure. The fatigue crack
growth rate of these cracks is considerably reduced due to
the presence of compressive stresses at the inner portion of
the cylinder wall [19].

( 2 2 - 2 / 2 + 1) - 4 2 ln + ( 2 2 ) + 2 2 (1 2 ) ln 2 2 (1 ) ln

9.

where is the ratio of outer to inner radii and ratio of


radius of autofrettage interface to the inner radius.
Residual hoop strain at inner surface of intact cylinder
follows from equations (8) and (9) and is given below.
2 1 2 0 2

(31)
0 =
1 22 ln
3 1 2 E

(40)

M ( )
M ( = )

where M is the bending moment and N is given by


(38)
N = (b2-a2) 4a2b2 (ln b/a)2
The opening angle is given by
M ( )
( )
(39)
=
( = b)

M ( = b )

The universal correlation function for relative


released hoop strain at points A & B and the opening angle
are related to level of autofrettage prevailing in the ring
[28], is given by equation (40).

Figure 8. Normalized stress intensity factors due to internal pressure for


a 5-percent eroded and a non-eroded cracked cylinder [19]

Copyright 2003 by JSME

total = K t lame + K t Autofrettage + Pr essure

where

K t lame

K t Autofrettage

(41)

is the concentration of the Lam stress,


is the concentration of the autofrettage

stresses. Pr essure is the hoop stress, arising from internal


pressure. Parker & Underwood [24] carried out their study
on single radial vent of diameter d(d<<2R1), where 2R1 is
the outer diameter of the tube which radially pierces the
tube at an angle of 900 to the longitudinal axis. Davidson
et al. [83] and Underwood et al. [31] have considered
angled evacuators (at 300; a typical value).
The Paris law [84] relating crack growth per
loading cycle (da/dN) to stress intensity range (K ) is
given below
da
(42)
= C(K ) m

Figure 9. Normalized effective stress intensity factors due to combined


pressure and autofrettage for a 5-percent eroded and a noneroded cracked cylinder [19]

dN

where C and m are material constants, determined


experimentally. Number of cycles (N) required for the
crack to propagate from crack length ai to af is given by
integrating equation (42).

Figure 8 shows that crack emanating from the


erosion's deepest point will experience much higher SIFs
due to pressurization, resulting in higher crack growth rate
and shorter fatigue life. Figure 9 shows that SIFs for an
eroded cracked cylinder are much higher than those in
non-eroded, mainly in the vicinity of the erosion, but
difference in the two cases decrease as (a0/W) increases,
where a0 is the crack length and W is the wall thickness.
Levy et al. [19] conclude that the smaller the radius of the
arc erosion the larger its effect is on KIP. The introduction
of an erosion of a given depth results in somewhat smaller
KIA (SIF for non pressurized, autofrettaged cylinder).
Effective SIF, Kieff is highly dependent on the erosion
shape, and semicircular erosion is found to be the most
critical of the arc erosions.

af

da

N=

a i C ( K )

(43)

Using
(44)
K = 1.12 a
calculated by Bowie & Newman after Rooke & Cartwright
[84] and Paris law exponent m = 3, yields on integration.

af =
2 / a i NC(1.12 )3

(45)

Equation (45) is useful in predicting final crackdepth profiles. Paker and Underwood [24] have made
comparisons between calculated lives using equation (45),
given by equation (46) and measured fatigue lives, as
given in table 5.

Crack growth along evacuator holes and cross-bores

Crack growth in pressurized thick cylinders with


single and multiple, internal and external radial cracks,
straight fronted and semi-elliptical for partial or full
autofrettage have all been analyzed with two-dimensional
geometries prior to the introduction of cracks. Evacuator
holes are found in cannons for evacuating combustion
gases after firing and cross-bore intersections are found in
the liquid end of a positive displacement pump, figure 10.
Cross-bore intersection is the area where high elasticplastic stresses and strains develop leading to localized
cyclic plasticity, making these areas susceptible to fatigue
cracking. A simple cross-bore is obtained by drilling two
holes in a block, intersecting at right angles. Stresses
contributing to the maximum stress concentration in the
vicinity of the evacuator may arise from three different
sources, as outlined by Parker and Underwood [24], in
equation (41) for truly three- dimensional geometries prior
to the introduction of cracks.

1
1
N = 2

ai
af

3
/ C 1.12 h eff

(46)

where h is the geometrical correction factor and assuming


(h=1) the total effective stress is given by
(47)
eff = K t autofrettage + K t Lame + 2P
and is limited to

K t autofrettage yield

Table 5. Specimen configuration and test conditions[24]


Tube
No.
1A
1B
2

Tube size at evacuator


R2(mm)
R1 (mm)
53
76
53
76
53
81

Amount of
autofrettage
None
None
100 percent

Measured
life
4710
5770
9780

Calculated
life Eq(43)
4300
4300
12,200

Badr et al. [27] have carried out 3-D finite


element analysis of cross-bore geometries as found in
liquid ends of positive displacement pumps (PDPs). They
have modified Glinka [85] approach based on energy
density criterion, with Bauschinger effect inclusion. The
hoop strain z is calculated from equation below
z =

y
y

(z1 / n 1)
x
+ 0P
.( z x
)
E
E
E
2
2
( e Peak + ( BEF )S y )1 / n

(48)

where e is the equivalent stress, 0P is the plastic strain


offset and sy is initial yield stress. Badr et al. [25] have
also evaluated the autofrettage effect on fatigue lives of
steel blocks with cross-bores using a statistical and strainbased method. They have used analysis of variance
(ANOVA) and the reverse plasticity criterion. Results of
statistical methodology correlated with the results of

Figure 10. Pressurized tube with evacuator hole

Copyright 2003 by JSME

reverse plasticity, indicated the existence of an "optimal"


autofrettage pressure.
10.

LOCALIZED AUTOFRETTAGE AS A
DESIGN TOOL FOR FATIGUE
IMPROVEMENT
Segall et al. [20] have investigated the feasibility
of improving the fatigue life of thick-walled cylinders with
cross-bores. Their technique utilizes the high stress
concentration at the cross-bore to induce localized
residual stresses, using relatively low internal pressures.
Iwadate et al.[86] have derived an approximate solution
for stress intensity factor kt, assuming a quarter-circular
crack (figure 11) situated at the intersection, a single stage
cross-bore and cylinder inner diameter, given in equation
below
1/ 2

k t = 1.25P

a 2 t
a sin

tan

Q a sin
2t

F(r, a , )

Figure 12 Calculated crack-growth curves for a 4340 steel, cross-bored


cylinder showing the benefit of a localized autofrettage [20]

(49)

where the correction factor F(r, a, ) helps account for the


stress concentration and bending at the cross-bore, P is the
applied internal pressure, a is the radius of the quarter
circular flaw, is the parametric angle of evaluation along
the crack front, and t is the cylinder wall-thickness. The
correction factor F is given by equation below
2

R0
+1

R
R
a
sin
+

b
F(r, a , ) = i
+
+ 1
2
(R b + a cos ) 2

R 0 1

R i

11.

MACHINING OF AN AUTOFRETTAGED
THICK-WALLED CYLINDER
Most of today's barrels are swage-autofrettaged,
which need to be machined to final dimensions both at the
bore and outer surface. In most cases the bore is rifled.
Thus material is further removed by broaching.
Autofrettage level in the barrel changes because of the
machining processes. Prevailing level of autofrettage
determines the allowable pressure and fatigue life. Post
machining stress components for fully and partially
autofrettaged cylinder machined internally, and externally
or both are given by Perl [21].

(50)

where R0,Ri,Rb are the outer, inner, and cross-bore radii


respectively
The flaw shape factor Q is defined as
Q = 1+1.464(

1.65

12.

TIME-DEPENDENT AUTOFRETTAGE
Generally the research work reported in the
literature is based on steady-state autofrettage analysis.
Malik et al. [15] , Owen and Hinton [17], Fahmy [32] and
Vaqar [87] have attempted time-dependent analysis of the
autofrettage of gun barrels. Fahmy [32] has outlined
necessary tests for the autofrettage technology for
production plants. Malik et al.[15] have carried out
modeling and simulation of time-dependent autofrettage
process for 125 mm gun tube using von Mises yield
criteria, by employing FEM, based on computer code
TRANSDYN [17]. Validation results [15] are shown in
figure 13 and 14 respectively.

a
1
c

Figure 11. Geometry and fracture mechanics idealization of a crossbored cylinder with a quarter-circular crack

Results of Segall et al. [20], using finite element


modeling for the resulting crack growth in the critical
cross-bore region, with and without benefits of
autofrettage are shown in figure 12.
Segall et al. concluded that a significant extension
of fatigue life was possible and that the localized residual
stresses were possible at pressures below yield threshold
for the thick-walled cylinder. Thus reverse plasticity,
permanent deformations and the need of post autofrettage
machining operations that could inadvertently lessen the
beneficial results of a traditional autofrettage were
avoided.

Figure 13. Vertical displacement time history [15]

10

Copyright 2003 by JSME

4.
5.
6.
7.
8.

9.
10.

Figure 14. Vertical displacement time history [Code TRANSDYN]

13.

CONCLUSIONS
The current paper has reviewed the following issues
related to swage-autofrettage process.

11.
12.

Prediction of residual stress field and an equivalent


temperature field for simulating partial autofrettage.
Effect of material properties, ranging from ductile to
brittle, on the residual stress field
Triaxiality of stress in the critical region of failure.
Non-ideal material behavior, such as Bauschinger effect
and its impact on fatigue lifetime.
Destructive, semi-destructive and non-destructive
methods for measuring level of autofrettage.
Crack emanating from erosion, its geometry and growth
in cylinders, evacuator holes and cross-bores.
Localized autofrettage as a design tool.
Effect of internal, external and combined machining on
the level of autofrettage.
Yield criterion and time-dependent autofrettage.

13.
14.

15.

16.
17.
18.

It is recommended to carry out modeling and


simulation of the critical areas of the process by including
non-linearities such as the Baushinger effect, strainhardening, strain aging, and to carry out controlled
experimentation for the reliability of results. The survey
has indicated very few statistical and transient analysis
attempts, which could be applied to add new dimensions
to the design process. Regarding the autofrettage process
for a gun barrel there are certain potential areas in which
future research could be focused, such as the shape of neck
for clamping the barrel, the level of surface finish during
boring, required for subsequent autofrettage, the number
of steps to achieve final size with desired surface finish,
size and shape of swage required.

19.

20.
21.
22.
23.
24.

ACKNOWLEDGEMENTS
The authors are indebted to College of Electrical
& Mechanical Engineering, National University of
Sciences & Technology, Pakistan, for the completion of
this research work. We also gratefully acknowledge the
help provided by Pakistan Scientific & Technical
Information Center (PASTIC) in acquiring the related
literature.

25.

REFERENCES

28.

1.
2.
3.

26.

27.

Davidson, T.E. and Kendall, D.P, 1970, "The design of pressure vessels for
very high pressure operation", Mechanical Behavior of Materials Under
Pressure, (H.L.P. Pugh, Ed.), Elsevier Co.
Rogan,J., 1975, "Fatigue strength and mode of fracture of high pressure
tubing made from low-alloy high strength steels", High Pressure Engineering,
I. Mech. E., London, UK. pp. 287-295.
Stacey, A. and Webster, G.A., 1988, "Determination of radial stress
distribution in autofrettaged tubing," International Journal of Pressure Vessels
and Piping, Vol. 31, pp. 205-220.

29.
30.

11

Perl, M., and Arone, R., 1988, "Stress intensity factors for a radial multijacketed partially autofrettaged pressurized thick-walled cylinder", ASME
Journal of Pressure Vessel Technology, Vol. 110, pp. 147-154.
Davidson, T.E., Barton, C.S., Reiner, A.N. and Kendall, D.P., 1962, "A new
approach to the autofrettage of high strength cylinders", Experimental
Mechanics, pp. 33-40.
Hill, R., 1950, "The mathematical theory of plasticity", Oxford University
Press, London.
Bland, D. R., 1956, "Elastoplastic thick-walled tubes of work-hardening
materials subject to internal and external pressures and temperature
gradients," Journal of Mechanics and Physics of Solids, Vol. 4, pp. 209-229.
Franklin, G. J. and Morrison, J. L. M., 1960, "Autofrettage of cylinders:
prediction of pressure/external expansion curves and calculation of residual
stresses," Proceedings of the Institute of Mechanical Engineers, Vol. 174, pp.
947-974
Chen, P. C. T., 1972, "The finite element analysis of elastic-plastic thickwalled tubes," Proceedings of Army Symposium on Solid Mechanics, The
Role of Mechanics in Design-Ballistic Problems, pp. 243-253.
Chen, P. C. T., 1973, "A comparison of flow and deformation theories in a
radially stressed annular plate," Journal of Applied Mechanics, Vol. 40, pp.
283-287.
Chen, P. C. T., 1981, "Numerical prediction of residual stress in an
autofrettage tube of compressible material," Proceedings of the 1981 Army
Numerical Analysis and Computer Conference, pp. 315-362.
Parker, A. P. and Andrasic, C. P., 1981, "Safe life design of gun tubes- some
numerical methods and results," Proceedings of the 1981 Army Numerical
Analysis and Computer Conference, pp. 311-333.
Milligan, R. V., Koo, W. H., and Davidson, T. E., 1966, "The Bauschinger
effect in a high strength steel, Journal of Basic Engineering, Vol. 88, pp.
480-488.
Malik, M.A., "Swage-autofrettaged thick-walled cylinders", 10th World
Congress on The Theory of Machines and Mechanisms, International
Federation of Machines and Mechanisms (IFTOMM) Oulu, Finland , June
20-24, 1999.
Malik M.A., Khushnood, S., and Vaqar A., "Modeling and simulation of
time-dependent autofrettage for a thick-walled cylinder", Proceedings of 4th
International Conference on Mechanics and Materials in Design. Nagoya,
Japan, June 5-8, 2002.
Ahmed, A., 1998, Modeling and simulation of autofrettage process for
strengthening of gun barrels M.Sc. Research Thesis, National University of
Sciences and Technology, NUST, Pakistan.
D. R. J. Owen and E. Hinton, 1980, "Finite elements in plasticity, theory and
practice.
Perl, M., and Arone, R., 1994, "An axisymmetric stress release method for
measuring the autofrettage level in thick-walled cylinders-part 1: basic
concept and numerical simulation," Transactions of ASME, Journal of
Pressure Vessel Technology , Vol. 116, pp.384-388.
Levy, C., Perl, M., Fang, H., 1998, "Cracks emanating from an erosion in a
pressurized autofrettage thick-walled cylinder-part I: semi-circular and arc
erosions", Journal of Pressure Vessel Technology, Vol. 122, No. 4, pp.34953.
Segall, A.E., Tricou, C., Evanko, M., & J.C. Conway Jr. 1998, "Localized
autofrettage as a design tool for the fatigue improvement of cross-bored
cylinders", Journal of Pressure Vessel Technology, Vol. 122, No 4, pp.393-7.
Perl, M., 2000, "The change in overstrain level resulting from machining of
autofrettaged thick-walled cylinder", Journal of Pressure Vessel Technology,
Vol. 122, No.1, pp.9-14.
Perl, M., 1988, "Temperature field for simulating partial autofrettage of an
elastic-plastic thick-walled cylinder",Transactions of ASME, Journal of
Pressure Vessel Technology, Vol. 110, pp.100-102.
Avitzur, B., 1988, "Determination of residual stress distribution in
autofrettaged tubing: a discussion", International Journal of Pressure Vessel
and Piping, pp.147-157.
Parker, A. P., Underwood, J.H., 1996, "Stress intensity, stress concentration,
and fatigue crack growth along evacuator holes of pressurized, autofrettaged
tubes", Journal of Pressure Vessel Technology, Vol. 118, pp. 336-42.
Badr, E. A., Sorem, J.R., Jr., and Tipton, S. M., 2000, "Evaluation of the
autofrettage effect on fatigue lives of steel blocks with cross-bores using a
statistical and a strain-based method", Journal of Testing and Evaluation, Vol.
28, No. 3, pp.181-8.
Parker, A. P., Underwood, J.H., Kendall, D.P, 1999, "Bauschinger effect
design procedures for autofrettaged tubes including material removal &
Sachs method", Journal of Pressure Vessel Technology, Vol. 121, No. 4,
pp.430-7.
Badr, E. A., Sorem, J.R., Jr., and Tipton, S. M., 1999, "Residual stress
estimation in cross-bores with Bauschinger effect inclusion using FEM and
strain energy density", Journal of Pressure Vessel Technology, Vol. 121, No.
4, pp.358-63.
Perl. M., 1998, "An improved split-ring method for measuring the level of
autofrettage in thick-walled cylinders", Journal of Pressure Vessel
Technology, Vol. 120, pp. 69-73.
Perl. M., Alperowitz, D., 1997, "The effect of crack length unevenness on
stress intensity factors due to autofrettage in thick-walled cylinders", Journal
of Pressure Vessel Technology, Vol. 119 , pp. 274-8.
Jahed, H., Dubey, R.N., 1997, "An Axisymmetric Method of Elastic-Plastic
Analysis Capable of Predicting Residual Stress Field", Journal of Pressure
Vessel Technology, Vol. 119, pp. 264-73.

Copyright 2003 by JSME

31.
32.
33.
34.
35.
36.
37.

38.
39.
40.
41.
42.
43.
44.
45.
46.

47.
48.

49.
50.
51.
52.
53.
54.
55.
56.
57.
58.
59.
60.
61.
62.
63.
64.

Underwood, J.H., Parker, A.P., Corrigan, D.J., & Audino, M.J., 1996,
"Fatigue life measurements and analysis for overstrained tubes with
evacuator holes", Journal of Pressure Vessel Technology Vol. 118, pp. 424-8.
Hassan Fahmy El Shalakany, 1975, Autofrettage of thick-walled tubesbarrels by the pressure of powder gas, Research Work, Antonin Zapotocky
Military Academy.
Boresi, A.P., and Chong, K.P., 1987, "Elasticity in engineering mechanics,"
New York, Elsevier.
Boresi, A.P., Schmidt, R.J., and Sidebottom, O.M., 1993 "Advanced
mechanics of materials", 5th Edition, John, Wiley & Sons Inc.
Timoshenko, S.P. and Goodier, J.M., 1980, "Theory of elasticity" Third
Edition, Mcgraw-Hill, New York NY.
Hill, R., 1950, "The mathematical theory of plasticity Clarendon, Oxford,
U.K.
Roach, D.P., & Priddy, T.G., 1994, "Effect of material properties on the
strain-to-failure of thick-walled cylinders subjected to internal pressure",
Transactions of ASME, , Journal of Pressure Vessel Technology, Vol. 116,
pp.96-104.
Rice, J.R., and Tracey, D.M., 1969, "On the ductile enlargement of voids in
triaxial stress fields," Journal of Mechanical Phys. Solids, Vol. 17.
Hancock, J.W., and Mackenzie,A.C., 1976, "On the mechanisms of ductile
failure in high strength steels subjected to multi-axis stress states", Journal of
Mechanical Phys. Solids, Vol. 24.
Soo Hoo, M., Benzley, S.E., and Priddy, T.G., 1980, "Experimental aspects
of an investigation of microscopic ductile failure criteria". SAND 80-1917,
Sandia National Laboratories, Alburqurque NM.
Crossland, B., and Gaydon, S.A., 1982, "A review of methods of predicting
the bursting pressure of thick-walled cylinders based on material properties",
ASME, PVP. Vol. 61.
Priddy, T.G., and Roach, D.P., 1989, "Strain-to-failure of pressurized thickwalled cylinders" ASME, Pressure Vessel & Piping Conference.
Faupel,J.H.,1956, "Yield and bursting characteristics of heavy wall
cylinders", Trans. ASME.
Marin, J., and Weng, T.L., 1961, "Strength of thick-walled cylindrical vessels
under internal pressure for three steels", Welding Research Council Bulletin,
No. 67.
Chen, P.C.T., 1986, "The Bauschinger and hardening effect on residual
stresses in an autofrettaged thick-walled cylinder", ASME Journal of Pressure
Vessel Technology, Vol. 108, pp. 108-112.
Parker, A.P. and Underwood, J.M., 1998, "Influence of Bauschinger effect on
residual stress and fatigue lifetimes in autofrettaged thick-walled cylinders",
Fatigue and Fracture Mechanics: Vol. 29, ASTM STP 1321, T.L. Panontin
and S.D. Sheppard eds.
Chaaban, A., Leung, K., and Burns, D.J., 1986, "Residual stresses in
autofrettaged thick-walled high pressure vessels" ASME, PVP, Vol. 110, pp.
55-60.
Kendall, D.P., 1998, Discussion of paper "The Bauschinger effect in
autofrettaged tubes A comparison of models including the ASME code" by
Parker, A.P. and Underwood, J.H. proceedings, ASME, Pressure Vessel &
Piping conference, San Diego, CA.
Megahed, M. M., and Abbas, A.T., 1991, "Influence of reverse yielding on
residual stresses induced by autofrettage", Journal of Mechanical Sciences,
Vol. 33.2, pp. 139-150.
Paris, P.C., and Endogan, F., 1963, "A critical analysis of crack propagation
laws", ASME Journal of Basic Engineering, Vol. 85, pp. 528-534.
Andrasic, C.P., and Parker, A.P.,1984, "Dimensionless stress intensity factors
for cracked thick cylinders under polynomial crack free loadings",
Engineering Fracture Mechanics, Vol. 19, No. 1, pp. 187-193.
Andrasic, C.P., and Parker, A.P., 1982, "Spline fit weight function data for
cracked thick cylinders", Royal Military College of Science, Shrivenham,
UK, Technical Note MAT/36.
Sachs, G., 1927, "The determination residual stresses in rod and tubes"
Zcitschrift f r Metallkundi, Vol. 19, pp 352-357.
Hcindlhofer, K., 1948 "Evaluation residual stresses", McGraw Hill Book Co.,
pp 138-139.
Weiss, V., 1956, "Residual stresses in cylinders" Syracuse University
Research Institute Report No: MET 345-563 T2 Syracuse NY.
Davidson, T.E., Kendall, D.P., and Reiner, A.N., 1963, Residual stresses in
thick-walled cylinders resulting from mechanically induced overstrain"
Experimental Mechanics, pp. 253-262.
Taira, S., Koterazawa, R., and Ohatni, R., 1965, "Creep of thick-walled
cylinders under internal pressure at elevated temperature", Proc. of the 8th
Japan Congr. on Testing Materials, 53-60.
Parker, A.P. and Farrow, J.R., 1980, "On the equivalence of axisymmetric
bending, thermal and autofrettage residual stress fields", Journal of Strain
Analysis, Vol. 15, No. 1, pp. 51-52.
Pu, S.L. and Hussain, M.A., 1981,"Residual stress distribution caused by
notches and cracks in a partially autofrettaged tube", Journal of Pressure
Vessel Technology, Vol. 103, pp. 302-306.
Timoshenko, S.P., 1970, "Theory of elasticity", McGraw-Hill, 3rd edition.
Hughes, H., 1967, "X-Ray technique for residual stress measurement" Strain,
Vol. 3, pp. 26-31.
Clark, G., 1982, "Residual stress in swage-autofrettaged thick-walled
cylinders" Report MRL-R-847, Department of Defence Support, Melbourne,
commonwealth of Australia.
McKannon, E.C., 1962, "Ultrasonic measurement of stress in aluminum,"
Nondestructive Testing and Techniques, NASA, SP-5082.
Abuku, S., and Cullity, B.D., 1971, "A magnetic method for the
determination of residual stresses", Experimental Mechanics, pp. 217-223.

65.
66.
67.
68.
69.
70.
71.
72.
73.
74.
75.
76.

77.

78.
79.

80.
81.
82.

83.
84.
85.
86.
87.

12

Allen, A., Andreani, C., Hutchings, M.T., and Windson, C.G., 1981,
"Measurement of internal stress within bulk materials using neutron
diffraction", NDT International, Vol. 14. pp. 246-254.
Stacey, A., MacGillivary, H.J., Webster, G.A., Webster, P.A., and Ziebeck,
K.R.A., "Measurement of residual stresses by neutron diffraction", Journal of
Strain Analysis, Vol. 20, pp. 93-100.
Opel, G.U., 1964, "Biaxial elasto-plastic analysis of load and residual stress",
experimental Mechanics, pp.135-140.
Underwood, J.H., 1973, "Residual stress measurement using surface
displacements around an indentation," Experimental Mechanics, pp. 373-380.
Mathar, J., 1934, "Determination of initial stress of measuring the
deformation around holes." Transactions ASME, Vol. 56, pp. 249-254.
Beaney, E.M., and Proctor, E., 1974; "A critical evaluation of the center hole
technique for the measurement of residual stresses", Strain, Vol, 10, pp. 7-14.
Rendler, H.J., and Vigness, I., 1966, "Hole drilling strain gage method of
measuring residual stress", Experimental Mechanics, pp. 577-586.
Milbradt, K.P., 1951, "Ring method determination of residual stresses,
"Proceedings, SESA XI, pp. 63-74.
Mesnager, M., 1919, "Methods of internal tensile stresses in cylindrical bars,
Compts Rendus, Hebedaries des Seances de 1 Academie des sciences, Vol.
169, pp. 1391-1393.
Williams, J. G., Grey, A., and Hodgekinson, J.M., 1981. "The determination
of residual stresses in plastic pipe and their role in fracture, "Polymer
Engineering Science, Vol. 21, pp. 816-821.
Cheng, W., and Fannie, I., 1986," Measurement of residual hoop stresses in
cylinders using compliance method", ASME Journal of Engineering
Materials and Technology, Vol. 108, pp 87-92.
Perl, M., and Arone, R., 1994, "An axisymmetric stress release method for
measuring the autofrettage level in thick-walled cylinders-Part-II:
Experimental validation", Journal of Pressure Vessel Technology, Vol. 116,
pp. 389-395.
Parker,A.P., Underwood, J.H., Throop, J.F., and Andrasic, C.P., 1983, "Stress
intensity and fatigue crack growth in pressurized autofrettaged thick
cylinders", Fracture Mechanics, Fourteenth Symposium-Vol. 1: Theory and
Analysis, ASTM STP 791, J.C. Lewis and G. Sines, eds., American Society
of Testing Materials, pp. I-216, I-237.
Pu. S.L., 1985, "Stress intensity factors at radial cracks of unequal depth in
partially autofrettaged pressurized cylinders" ARLCB-TR-85018, US Army
Armament Research and Development Center, Watcrevliet NY.
Perl, M., Wu. K.H., and Arone, R., 1990, "Uniform arrays of unequal depth
cracks in thick-walled cylindrical pressure vessels. Part I Stress Intensity
Factors Evaluation" ASME Journal of Pressure Vessel Technology, Vol. 112,
pp. 340-345.
Arone, R., and Perl, M., 1989, "Influence of autofrettage on the stress
intensity factors for a thick-walled cylinder with radial cracks of unequal
length", International Journal of Fracture, Vol. 39, pp. R-29-R34.
Becker, A.A., Plant, R.C.A, and Parker, A.P., 1993, "Axial cracks in
pressurized eroded autofrettaged thick cylinders", International Journal of
Fracture, Vol. 63, pp. 113-134.
Parker, A.P., Plant R.C.A. and Becker, A.A., 1993, "Fatigue lifetimes for
pressurized eroded cracked autofrettaged thick cylinders" Fracture
Mechanics, Twenty-third Symposium, ASTM STP 1189, Ravinder Chona,
ed., ASTM, Philadelphia, PA, pp. 461-473.
Davidson, T.E., Brown, B.B., and Kendall, D.P.,1977, "Material and
processes considerations in the design of pressure vessels", 2nd International
Conference on High Pressure Technology, Brighton, I Mech. E., pp. 63-71.
Rooke, D.P., and Cartwright, D.J., 1976, "Compendium of stress intensity
factors", HMSCO, London, England.
Glinka, G., 1985, "Calculation of Inelastic notch tip stress-strain histories
under cyclic loading", Engineering Fracture Mechanics, Vol. 22, No. 5, pp.
839-845.
Iwadate, T., Chiba, K., Watanaba, J., Mima, S., Tokai, K., and Takeda, H.,
1981, "Safety analysis at a cross-bore corner of a high pressure polyethylene
reactor", ASME, PVP, Vol. 48.
Vaqar, A., 1999, "Modeling and simulation of time-dependent autofrettage
process for strengthening of gun barrels", Masters thesis, National University
of Sciences & Technology, Pakistan.

Copyright 2003 by JSME

S-ar putea să vă placă și