Metallurgical Modelling of Welding
Metallurgical Modelling of Welding
Metallurgical Modelling of Welding
Modelling
of Welding
SECOND EDITION
0YSTEIN GRONG
Norwegian University of
Science and Technology,
Department of Metallurgy,
N-7034 Trondheim, Norway
Editor: H. K. D. H. Bhadeshia
The University of Cambridge
Department of Materials Science
and Metallurgy
T H E INSTITUTE OF MATERIALS
Book 677
First published in 1997 by
The Institute of Materials
1 Carlton House Terrace
London SWlY 5DB
Originally typeset by
PicA Publishing Services
Additional typesetting and corrections by
Fakenham Photosetting Ltd
Printed and bound in the UK at
The University Press, Cambridge
TO TORHILD, TORBJ0RN AND HAVARD
without your support, this book would never have been finished.
Preface to the second edition
Besides correcting some minor linguistic and print errors, I have in the second edition in-
cluded a collection of different exercise problems which have been used in the training of stu-
dents at NTNU. They illustrate how the models described in the previous chapters can be used
to solve practical problems of more interdisciplinary nature. Each of them contains a 'prob-
lem description' and some background information on materials and welding conditions. The
exercises are designed to illuminate the microstructural connections throughout the weld
thermal cycle and show how the properties achieved depend on the operating conditions ap-
plied. Solutions to the problems are also presented. These are not complete or exhaustive, but
are just meant as an aid to the reader to develop the ideas further.
0ystein Grong
Preface to the first edition
The purpose of this textbook is to present a broad overview on the fundamentals of welding
metallurgy to graduate students, investigators and engineers who already have a good back-
ground in physical metallurgy and materials science. However, in contrast to previous text-
books covering the same field, the present book takes a more direct theoretical approach to
welding metallurgy based on a synthesis of knowledge from diverse disciplines. The motiva-
tion for this work has largely been provided by the need for improved physical models for
process optimalisation and microstructure control in the light of the recent advances that have
taken place within the field of materials processing and alloy design.
The present textbook describes a novel approach to the modelling of dynamic processes in
welding metallurgy, not previously dealt with. In particular, attempts have been made to ra-
tionalise chemical, structural and mechanical changes in weldments in terms of models based
on well established concepts from ladle refining, casting, rolling and heat treatment of steels
and aluminium alloys. The judicious construction of the constitutive equations makes full use
of both dimensionless parameters and calibration techniques to eliminate poorly known ki-
netic constants. Many of the models presented are thus generic in the sense that they can be
generalised to a wide range of materials and processing. To help the reader understand and
apply the subjects and models treated, numerous example problems, exercise problems and
case studies have been worked out and integrated in the text. These are meant to illustrate the
basic physical principles that underline the experimental observations and to provide a way of
developing the ideas further.
Over the years, I have benefited from interaction and collaboration with numerous people
within the scientific community. In particular, I would like to acknowledge the contribution
from my father Professor Tor Grong who is partly responsible for my professional upbringing
and development as a metallurgist through his positive influence on and interest in my re-
search work. Secondly, I am very grateful to the late Professor Nils Christensen who first
introduced me to the fascinating field of welding metallurgy and later taught me the basic
principles of scientific work and reasoning. I will also take this opportunity to thank all my
friends and colleagues at the Norwegian Institute of Technology (Norway), The Colorado School
of Mines (USA), the University of Cambridge (England), and the Universitat der Bundeswehr
Hamburg (Germany) whom I have worked with over the past decade. Of this group of people,
I would particularly like to mention two names, i.e. our department secretary Mrs. Reidun
0stbye who has helped me to convert my original manuscript into a readable text and Mr.
Roald Skjaerv0 who is responsible for all line-drawings in this textbook. Their contributions
are gratefully acknowledged.
0ystein Grong
Contents
1.1 Introduction
The symbols and units used throughout this chapter are defined in Appendix 1.1.
Since heat losses from free surfaces by radiation and convection are usually negligible in
welding, the temperature distribution can generally be obtained from the fundamental differ-
ential equations for heat conduction in solids. For uniaxial heat conduction, the governing
equation can be written as:1
(i-D
where T is the temperature, t is the time, x is the heat flow direction, and a is the thermal
diffusivity. The thermal diffusivity is related to the thermal conductivity X and the volume
heat capacity pc through the following equation:
(1-2)
d-3)
and
(1-4)
The above equations must clearly be satisfied by all solutions of heat conduction problems,
but for a given set of initial and boundary conditions there will be one and only one solution.
A pre-condition for obtaining simple analytical solutions to the differential heat flow equa-
tions is that the thermal properties of the base material are constant and independent of tem-
perature. For most metals and alloys this is a rather unrealistic assumption, since both X, a,
and pc may vary significantly with temperature as illustrated in Fig. 1.1. In addition, the
thermal properties are also dependent upon the chemical composition and the thermal history
of the base material (see Fig. 1.2), which further complicates the situation.
By neglecting such effects in the heat flow models, we impose several limitations on the
application of the analytical solutions. Nevertheless, experience has shown that these prob-
lems to some extent can be overcome by the choice of reasonable average values for X, a and
pc within a specific temperature range. Table 1.1 contains a summary of relevant thermal
properties for different metals and alloys, based on a critical review of literature data. It should
be noted that the thermal data in Table 1.1 do not include a correction for heat consumed in
melting of the parent materials. Although the latent heat of melting is temporarily removed
during fusion welding, experience has shown this effect can be accounted for by calibrating
the equations against a known isotherm (e.g. the fusion boundary). In practice, such correc-
tions are done by adjusting the arc efficiency factor Tq until a good correlation is achieved
between theory and experiments.
Carbon steel
Hx-H0 = PC(T-T0 ),J/mm3
Temperature, 0C
Fig. 1.1. Enthalpy increment H7-H0 referred to an initial temperature T0 = 200C. Data from Refs.
2-4.
Table 1.1 Physical properties for some metals and alloys. Data from Refs 2 - 6 .
X9 W/mm 0C
Temperature, 0C
(b)
High alloy steel
X, W/mm 0C
Temperature, 0C
Fig. 1.2. Factors affecting the thermal conductivity X of steels; (a) Temperature level and chemical
composition, (b) Heat treatment procedure. Data from Refs. 2-4.
The concept of instantaneous heat sources is widely used in the theory of heat conduction.1 It
is seen from Fig. 1.3 that these solutions are based on the assumption that the heat is released
instantaneously at time t - 0 in an infinite medium of initial temperature T0, either across a
plane (uniaxial conduction), along a line (biaxial conduction), or in a point (triaxial conduc-
tion). The material outside the heat source is assumed to extend to x = + °° for a plane source
in a long rod, to r = °° for a line source in a wide plate, or to R = °° for a point source in a heavy
slab. The initial and boundary conditions can be summarised as follows:
T-T0 = oo for t = O and x = O (alternatively r = O or R = O)
T-J 0 = O for t = O and x * O (alternatively r > O or 7? > O)
7-T 0 = O for O < t < oo when x = ± oo (alternatively r = oo or R = oo).
It is easy to verify that the following solutions satisfy both the basic differential heat flow
equations (1-1), (1-3) and (1-4) and the initial and boundary conditions listed above:
where Q is the net heat input (energy) released at time t = O, and A is the cross section of the
rod.
(1-6)
where d is the plate thickness.
Equations (1-5), (1-6) and (1-7) provide the required basis for a comprehensive theoretical
treatment of heat flow phenomena in welding. These solutions can either be applied directly
or be used in an integral or differential form. In the next sections a few examples will be given
to illustrate the direct application of the instantaneous heat source concept to problems related
to welding.
(a) T
Fig. 1.3. Schematic representation of instantaneous heat source models; (a) Plane source in a long rod.
(b) T
X y
(C)
Fig. 1.3.Schematic representation of instantaneous heat source models (continued); (b) Line source in a
wide plate, (c) Point source in a heavy slab.
1.5 Local Fusion in Arc Strikes
The series of fused metal spots formed on arc ignition make a good case for application of
equation (1-7).
Model
The model considers a point source on a heavy slab as illustrated in Fig. 1.4. The heat is
assumed to be released instantaneously at time t = 0 on the surface of the slab. This causes a
temperature rise in the material which is exactly twice as large as that calculated from equation
(1-7):
(1-8)
(1-9)
— Dimensionless time:
d-10)
(1-11)
where qo is the net arc power (equal to Qlt(), and (Hc-Ho) is the heat content per unit volume at
the reference temperature.
(1-12)
(1-13)
Heat source
Isotherms
Fig. 1.4. Instantaneous point source model for assessment of temperatures in arc strikes.
T1
^i
Equation (1-13) has been solved numerically for different values ofCT1and T1. The results
are presented graphically in Fig. 1.5. Due to the inherent assumption of instantaneous release
of heat in a point, it is not possible to use equation (1-13) down to very small values OfCT1 and
T1. However, at some distance from the heat source and after a time not much shorter than the
real (assumed) time of heating, the calculated temperature-time pattern will be reasonably
correct. Note that the heavy broken line in Fig. 1.5 represents the locus of the peak tempera-
tures. This locus is obtained by setting 3In(OAi1VdT1 = 0:
from which
(1-14)
where Qp is the peak temperature, and e is the natural logarithm base number.
Example (1.1)
Consider a small weld crater formed in an arc strike on a thick plate of low alloy steel. Calcu-
late the cooling time from 800 to 5000C (Af875), and the total width of the fully transformed
region adjacent to the fusion boundary. The operational conditions are as follows:
where r| is the arc efficiency factor. Relevant thermal data for low alloy steel are given in
Table 1.1.
Solution
In the present case it is convenient to use the melting point of the steel as a reference tempera-
ture (i.e. 0 = 0m = 1 when Tc = TJ. The corresponding values OfZi1 and 9 (at 800 and 5000C,
respectively) are:
and
Alternatively, the same information could have been read from Fig. 1.5. Although it is
difficult to check the accuracy of these predictions, the calculated values for Ats/5 and ARlm are
considered reasonably correct. Thus, because the cooling rate is very large, in arc strikes a
hard martensitic microstructure would be expected to form within the transformed parts of the
HAZ, in agreement with general experience.
Equation (1-6) can be used for an assessment of the temperature-time pattern in spot welding
of plates.
Model
The model considers a line source which penetrates two overlapping plates of similar thermal
properties, as illustrated in Fig. 1.7. The heat is assumed to be released instantaneously at time
Heat source
Heat source d
Fig. 1.7. Idealised heat flow model for spot welding of plates.
t = 0. If transfer of heat into the electrodes is neglected, the temperature distribution is given
by equation (1-6).
This equation can be written in a dimensionless form by introducing the following group of
parameters:
— Dimensionless time:
(1-15)
(1-16)
(1-17)
(1-18)
e/n2
6/n2
T
2
T2
Figure 1.8 shows a graphical representation of equation (1-18) for a limited range of a 2 and
T2. A closer inspection of the graph reveals that the temperature-time pattern in spot welding
is similar to that observed during arc ignition (see Fig. 1.5). The locus of the peak tempera-
tures in Fig. 1.8 is obtained by setting d\n{^ln7}ldx2 - 0.
which gives
and
(1-19)
Example (1.2)
Consider spot welding of 2 mm plates of low alloy steel under the following operational con-
ditions:
Calculate the cooling time from 800 to 5000C (Af8/5) in the centre of the weld, and the cooling
rate (CR.) at the onset of the austenite to ferrite transformation. Assume in these calculations
that the total voltage drop between the electrodes is 1.6 V. The M^-temperature of the steel is
taken equal to 475°C.
Solution
If we use the melting point of the steel as a reference temperature, the parameters n2 and 6 (at
800 and 5000C, respectively) become:
and
and
Since the cooling curves in Fig. 1.8 are virtually parallel at temperatures below 8000C (i.e.
for QZn2 < 0.15), the computed values of Ar8/5 and CR. are also valid for positions outside the
weld centre-line. In the present example the centre-line solutions can be applied down to
(°"2m)2 ~ 2. According to equation (1-19), this corresponds to a lower peak temperature of:
Thermit welding is a process that uses heat from exothermic chemical reactions to produce
coalescence between metals and alloys. The thermit mixture consists of two components, i.e.
a metal oxide and a strong reducing agent. The excess heat of formation of the reaction prod-
uct provides the energy source required to form the weld.
Model
In thermit welding the time interval between the ignition of the powder mixture and the com-
pletion of the reduction process will be short because of the high reaction rates involved.
Assume that a groove of width 2L1 is filled instantaneously at time t = 0 by liquid metal of an
initial temperature Tt (see Fig. 1.10). The metal temperature outside the fusion zone is T0. If
heat losses to the surroundings are neglected, the problem can be treated as uniaxial conduc-
tion where the heat source (extending from -L 1 to +L1) is represented by a series of elementary
sources, each with a heat content of:
(1-20)
At time t this source produces a small rise of temperature at position JC, given by equation (1 -5):
(1-21)
(1-22)
Isl'srau*'*'=]
Fusion zone
Fig. 1.9. Calculated peak temperature contours in spot welding of steel plates (numerical solution). Op-
erational conditions: / = 23kA, 64 cycles. Data from Bently et al1
Fusion zone
Fig. 1.10. Idealised heat flow model for thermit welding of rails.
where erf(u) is the Gaussian error function. The error function is defined in Appendix 1.3*.
Because of the complex nature of equation (1-22), it is convenient to present the different
solutions in a dimensionless form by introducing the following groups of parameters:
*The error function is available in tables. However, in numerical calculations it is more convenient to use the
Fortran subroutine given in Appendix 1.3.
Dimensionless temperature:
(1-23)
Dimensionless time:
(1-24)
Dimensionless jc-coordinate:
(1-25)
(1-26)
Equation (1-26) has been solved numerically for different values of Q and T3. The results
are presented graphically in Fig. 1.11. As would be expected, the fusion zone itself (Q < 1)
cools in a monotonic manner, while the temperature in positions outside the fusion boundary
(Q > 1) will pass through a maximum before cooling. The locus of the HAZ peak temperatures
in Fig. 1.11 is defined by 3673T3 = 0. Referring to Appendix 1.3, we may write:
which gives
(1-27)
The peak temperature distribution is obtained by solving equation (1-27) for different com-
binations of Qm and T3m and inserting the roots into equation (1-26).
Example (1.3)
Consider thermit welding of steel rails (i.e. reduction of Fe2O3 with Al powder) under the
following operational conditions:
Calculate the cooling time from 800 to 5000C in the centre of the weld, and the total width
of the fully transformed region adjacent to the fusion boundary. The Ac3-temperature of the
steel is taken equal to 8900C.
Definition of parameters:
91
T
3
Solution
For positions along the weld centre-line (Q. = 0) equation (1-26) reduces to:
and
The locus of the 8900C isotherm in temperature-time space can be read from Fig. 1.11.
Taking the ordinate equal to 0.40, we get:
By inserting this value into equation (1-27), we obtain the corresponding coordinate of the
isotherm:
Unfortunately, measurements are not available to check the accuracy of these predictions.
Systematic errors would be expected, however, because of the assumption of instantaneous
release of heat immediately after powder ignition and the neglect of heat losses to the sur-
roundings. Nevertheless, the present example is a good illustration of the versatility of the
concept of instantaneous heat sources, since these solutions can easily be added in space as
shown here or in time for continuous heat sources (to be discussed below).
Friction welding is a solid state joining process that produces a weld under the compressive
force contact of one rotating and one stationary workpiece. The heat is generated at the weld
interface because of the continuous rubbing of the contact surfaces, which, in turn, causes a
temperature rise and subsequent softening of the material. Eventually, the material at the
interface starts to flow plastically and forms an up-set collar. When a certain amount of up-
setting has occurred, the rotation is stopped and the compressive force is maintained or slightly
increased to consolidate the weld.
(1-28)
(a)
Continuous heat source
(b) q
t
Fig. 1.12. Idealised heatflowmodel for friction welding of rods; (a) Sketch of model, (b) Subdivision of
time into a series of infinitesimal elements dt'.
At time / this heat will cause a small rise of temperature in the material, in correspondance
with equation (1-5):
(1-29)
If we substitute t"=t-1'into equation (1 -29), the total temperature rise at time t is obtained
by integrating from t"= t (t'= 0) to /"= 0 (t'= t):
(1-30)
In order to evaluate this integral, we will make use of the following mathematical transfor-
mation:
where
and
The latter integral can be expressed in terms of the complementary error function* erfc{u)
by substituting:
(1-31)
If the temperature of the contact section at the end of the heating period is taken equal to Th,
equation (1-31) can be rewritten as:
(1-32)
where t'h denotes the duration of the heating period (t < t'h). Measured contact section tem-
peratures for different metal/alloy combinations are given in Table 1.2.
Equation (1-32) may be presented in a dimensionless form by the use of the following
groups of parameters:
Dimensionless temperature:
(1-33)
Dimensionless time:
(1-34)
Based on direct readings of the voltage drop between the two work-pieces.
— Dimensionless .^-coordinate:
(1-35)
(1-36)
Equation (1-36) describes the temperature in different positions from the weld contact sec-
tion during the heating period. However, when the rotation stops, the weld will be subjected to
free cooling, since there is no generation of heat at the interface. As shown in Fig. 1.13(a) this
can be accounted for by introducing an imaginary heat source of power +qo at time t = t'h
which acts simultaneously with an imaginary heat sink of negative power -q o. It follows from
the principles of superposition (see Fig. 1.13b) that the temperature during the cooling period
is given by:9
(1-37)
where 6"(x4) and 6"(T 4 - 1) are the temperatures calculated for the heat source and the heat
sink, respectively, using equation (1-36).
Equations (1-36) and (1-37) have been solved numerically for different values of Q'and T4.
The results are presented graphically in Fig. 1.14. Considering the contact section (Q'= 0), the
temperature increases monotonically with time during the heating period, in correspondance
with the relationship:
(1-38)
q
(a)
e"
(b)
Heating
period
$ffl9
Fig. 1.13. Method for calculation of transient temperatures during friction welding; (a) Sketch of
imaginary heat source/heat sink model, (b) Principles of superposition.
(1-39)
Outside the contact section (Q / > 0), the temperature rise will be smaller and the cooling
rate lower than that calculated from equations (1-38) and (1-39).
Heating Cooling
e"
Example (1.4)
Consider friction welding of 026mm aluminium rods (Al-Cu-2Mg) under the following con-
ditions:
Calculate the peak temperature distribution across the joint. Assume in these calculations
that the thermal diffusivity of the Al-Cu-2Mg alloy is 70mm2 s"1.
Solution
Readings from Fig. 1.14 give:
Next Page
In this particular case, it is possible to check the accuracy of the calculations against in situ
thermocouple measurements carried out on friction welded components made under similar
conditions. A comparison with the data in Fig. 1.15 shows that the model is quite successful in
predicting the HAZ peak temperature distribution. In contrast, the weld heating and cooling
cycles cannot be reproduced with the same degree of precision. This has to do with the fact
that the present analytical solution omits a consideration of the plastic straining occurring
during friction welding, which displaces the coordinates and alters the heat balance for the
system.
In most fusion welding processes the heat source does not remain stationary. In the following
we shall assume that the source moves at a constant speed along a straight line, and that the net
power supply from the source is constant. Experience shows that such conditions lead to a
fused zone of constant width. This is easily verified by moving a tungsten arc across a sheet of
steel or aluminium, or by moving a soldering iron across a piece of lead or tin. Moreover,
zones of temperatures below the melting point also remain at constant width, as indicated by
the pattern of temper colours developed on welding ground or polished sheet.
It follows from the definition of pseudo-steady state that the temperature will not vary with
time when observed from a point located in the heat source. Under such conditions the tem-
perature field around the source can be described as a temperature 'mountain' moving in the
direction of welding (e.g. see Fig. 15 in Ref. 11). For points along the weld centre-line, the
temperature at different positions away from the heat source (which for a constant welding
speed becomes a time axis) may be presented in a two-dimensional plot as indicated in Fig.
1.16. Specifically, this figure shows a schematic representation of the temperature in steel
welding from the base plate ahead of the arc to well into the solidified weld metal trailing the
arc. If we consider a fixed point on the weld centre-line, the temperature will increase very
rapidly during the initial period, reaching a maximum of about 2000-22000C for positions
immediately beneath the root of the arc.11 When the arc has passed, the temperature will start
to fall, and eventually (after long times) approach that of the base plate. In contrast, an ob-
server moving along with the heat source will always see the same temperature landscape,
since this will not change with time according to the presuppositions.
It will be shown below that the assumption of pseudo-steady state largely simplifies the
mathematical treatment of heat flow during fusion welding, although it imposes certain re-
strictions on the options of the models.
In this particular case, it is possible to check the accuracy of the calculations against in situ
thermocouple measurements carried out on friction welded components made under similar
conditions. A comparison with the data in Fig. 1.15 shows that the model is quite successful in
predicting the HAZ peak temperature distribution. In contrast, the weld heating and cooling
cycles cannot be reproduced with the same degree of precision. This has to do with the fact
that the present analytical solution omits a consideration of the plastic straining occurring
during friction welding, which displaces the coordinates and alters the heat balance for the
system.
In most fusion welding processes the heat source does not remain stationary. In the following
we shall assume that the source moves at a constant speed along a straight line, and that the net
power supply from the source is constant. Experience shows that such conditions lead to a
fused zone of constant width. This is easily verified by moving a tungsten arc across a sheet of
steel or aluminium, or by moving a soldering iron across a piece of lead or tin. Moreover,
zones of temperatures below the melting point also remain at constant width, as indicated by
the pattern of temper colours developed on welding ground or polished sheet.
It follows from the definition of pseudo-steady state that the temperature will not vary with
time when observed from a point located in the heat source. Under such conditions the tem-
perature field around the source can be described as a temperature 'mountain' moving in the
direction of welding (e.g. see Fig. 15 in Ref. 11). For points along the weld centre-line, the
temperature at different positions away from the heat source (which for a constant welding
speed becomes a time axis) may be presented in a two-dimensional plot as indicated in Fig.
1.16. Specifically, this figure shows a schematic representation of the temperature in steel
welding from the base plate ahead of the arc to well into the solidified weld metal trailing the
arc. If we consider a fixed point on the weld centre-line, the temperature will increase very
rapidly during the initial period, reaching a maximum of about 2000-22000C for positions
immediately beneath the root of the arc.11 When the arc has passed, the temperature will start
to fall, and eventually (after long times) approach that of the base plate. In contrast, an ob-
server moving along with the heat source will always see the same temperature landscape,
since this will not change with time according to the presuppositions.
It will be shown below that the assumption of pseudo-steady state largely simplifies the
mathematical treatment of heat flow during fusion welding, although it imposes certain re-
strictions on the options of the models.
The main purpose of this section is to review the classical models for the pseudo-steady
state temperature distribution around moving heat sources. The analytical solutions to the
differential heat flow equations under conditions applicable to arc welding were first presented
(a)
Time, s
(b)
Peak temperature, 0C
Observed relationship
Predicted relationship
Fig. 1.15. Comparison between measured and predicted temperatures in friction welding of Al-Cu-2Mg
alloys; (a) Temperature-time pattern, (b) Peak temperature distribution. Data from Tensi et al.10
by Rosenthal,1314 but the theory has later been extended and refined by a number of other
inve stigators .9'n*15"20
where qo is the net power received by the weldment (e.g. measured by calorimetry), / is the
welding current (amperage), and U is the arc voltage.
For submerged arc (SA) welding the efficiency factor (r\) has been reported in the range
from 90 to 98%, for SMA and GMA welding from 65 to 85%, and for GTA welding from 22
to 75%, depending on polarity and materials.11
A summary of ranges is given in Table 1.3.
(1-41)
where is the time available for conduction of heat over the distance
to point P.
For a convenient presentation of the pseudo-steady state solution, the position P should be
referred to that of the moving heat source. This is achieved by changing the coordinate system
from O to O'(see Fig. 1.17):
and
Weld pool
Root of arc
Solidified
weld metal
X=Vt
Relative position along weld centre-line
Fig. 1.16. Schematic diagram showing weld centre-line temperature in different positions from the heat
source during steel welding at pseudo-steady state.
Table 1.3 Recommended arc efficiency factors for different welding processes. Data from Refs 11,21.
where
and
into equation (1-42), and integrating between the limits u = (R2l4af)m and u = o°.This gives
(after some manipulation):
(1-43)
It is well-known that:
(1-44)
If u is sufficiently small (i.e. when welding has been performed over a sufficient period), we
obtain the pseudo-steady state temperature distribution:
(1-45)
This equation is often referred to as the Rosenthal thick plate solution,1314 in honour of
D. Rosenthal who first derived the relation by solving the differential heat flow equation di-
rectly for the appropriate boundary conditions.
attained after a transient heating period. The duration of this heating period is determined by
the integral in equation (1-44).
Taking the ratio between the real and the pseudo-steady state temperature equal to K1, we
have:
(1-46)
(1-47)
Dimensionless time:
(1-48)
(1-49)
where and
Equation (1-49) has been solved numerically for a limited range of cr3 and x. The results are
presented graphically in Fig. 1.18. A closer inspection of Fig. 1.18 reveals that the duration of
the transient heating period depends on the dimensionless radius vector a 3 . In practice this
means that the Rosenthal equation is not valid during the initial period of welding unless the
distance from the heat source to the observation point is very small. It should be noted, how-
ever, that a dimensionless distance o~3 may be 'short' for one combination of welding speed
and thermal diffusivity, while the same position may represent a 'long' distance for another
combination of V and a. Similarly, the dimensionless time T may be 'short' or 'long' at a
chosen number of seconds, depending on the ratio v/2a.
Example (1.5)
Consider stringer bead deposition on a thick plate of aluminium at a constant welding speed of
5 mm s"1. Calculate the duration of the heating period when the distance from the heat source
to the point of observation is 17 mm.
Solution
Taking a = 85 mm2 s"1, the dimensionless radius vector becomes:
Ki = (T-T0)/(T-T0)p.s.
T = v2t/2a
Fig. 1.18. Ratio between real and pseudo-steady state temperature in thick plate welding for
different combinations ofCT3and T.
It is seen from Fig. 1.18 that the pseudo-steady state temperature distribution is approached
when T ~ 3, which gives:
The above calculations show that the Rosenthal equation is not valid if the ratio between R
and L2 exceeds a certain critical value (typically 0.15 to 0.30 for aluminium welds and 0.4 to
0.6 for steel welds). This important point is often overlooked when discussing the relevance of
the thick plate solution in arc welding.
(1-50)
Dimensionless jc-coordinate:
(1-51)
Dimensionless ^-coordinate:
(1-52)
Dimensionless z-coordinate:
(1-53)
where 6 and a 3 are the dimensionless temperature and radius vector, respectively (previously
defined in equations (1-9) and (1-47)).
Equation (1-54) has been solved numerically for chosen values of a 3 and £. A graphical
presentation of the different solutions is shown in Fig. 1.19. These maps provide a good
(a)
e/n3
%
Fig, /./P.Dimensionless temperature maps for point sources on heavy slabs; (a) Vertical sections paral-
lel to the ^-axis.
overall indication of the thermal conditions during thick plate welding, but are not suitable for
precise readings. Consequently, for quantitative analyses, the following set of equations can
be used:11
Fig. /.iP.Dimensionless temperature maps for point sources on heavy slabs (continued): (b) Isothermal
contours in the £-\}/-plane for different ranges of 0/n3.
(1-55)
Equation (1-55) can be used for calculations of isothermal zone widths V|/w and cross sec-
tional areas A1. From Fig. 1.20 we have:
(1-56)
and
(1-57)
A graphical presentation of equations (1-55), (1-56), and (1-57) is shown in Fig. 1.21.
(1-58)
where £' and £"are the distance from the heat source to the front and the rear of the enclosure,
respectively.
x,S
Heat source
y.v
z, C
1
V9P
Fig. 1.21. Dimensionless distance a3m, half width \|/m and cross sectional area A1 vs n3 /Qp.
The coordinates £' and £" are found by setting a 3 = ± ^ in equation (1-54). This gives:
(1-59)
and
(1-60)
(1-61)
(1-62)
(1-63)
The dimensionless volume T is related to the real volume of the enclosure V(mm3) through
the following equation:
(1-64)
d-65)
Equation (1-65) provides a basis for calculating the cooling time within a specific tempera-
ture interval (e.g. from O1 to B2):
(1-66)
For welding of low alloy steel the cooling time from 800 to 5000C is widely accepted as an
adequate index for the thermal conditions under which the austenite to ferrite transformation
takes place. From equation (1-66), we have:
(1-67)
(1-69)
(1-70)
(1-71)
Example (1.6)
Consider stringer bead deposition (GMAW) on a thick plate of low alloy steel under the fol-
lowing conditions:
Sketch the contours of the fusion boundary and the Ac3-isotherm (9100C) in the £-\|/ (x-y)
plane at pseudo-steady state.
Solution
As shown in Fig. 1.22(a) it is sufficient to calculate the coordinates in four different (character-
istic) positions to draw a sketch of the isothermal contours. If we neglect the latent heat of
melting, the d/n3 ratio at the melting point becomes, according to equations (1-9) and (1-50):
End-points
The end-points are readily obtained from equations (1-59) and (1-60):
and
Maximum width
The maximum width of an isothermal enclosure can generally be calculated from equations
(1-55) and (1-56), or read from Fig. 1.21. Taking O^ /n3 = QIn3 = 0.088, we obtain:
(a) V
Pos. (3)
Pos. (2)
Pos. (4) Pos.(1)
%
(b) ¥
\
x(mm)
y(mm)
Fig. 1.22. Calculation of isothermal contours from Rosenthal thick plate solution; (a) Calculation proce-
dure, (b) Sketch of fusion boundary and Acj-isotherm in position £ = 0 (Example 1.6).
and
Intersection point with y/ (y)-axis
In this case £ = £ = O, and cr3 = \|/. Hence, equation (1-54) reduces to:
which gives
Similarly, the contour of the Ac3-isotherm can be determined by inserting BZn3= 0.052 into
the same set of equations. Figure 1.22(b) shows a graphical representation of the computed
isothermal contours.
Example (1.7)
Consider GTA welding on a thick plate of low alloy steel under the following conditions:
Calculate the weld pool volume, the weld bead cross section, the width of the fully trans-
formed HAZ, the cooling time from 800 to 5000C, and the cooling rate at the onset of the
austenite to ferrite transformation (e.g. at 6500C). The arc efficiency factor is taken equal to
0.5.
Solution
If we neglect the latent heat of melting, the GM3 ratio at the melting point becomes:
We can now use equations (1-63) and (1-64) to calculate the weld pool volume. This gives:
and
Weld bead cross section
The weld bead cross section can be read from Fig. 1.21. Taking n3 /6 p = 1/0.445 = 2.196, we
obtain:
and
which gives
This value is also valid for positions outside the weld centre-line, since the cooling curves
at such low temperatures are reasonably parallel down to Tp ~ 9100C (see Fig. 1.19(a)).
(1-72)
According to the assumptions this amount of heat will remain in a slice of thickness dx due
to the lack of a temperature gradient in the welding direction. Since symmetry demands that
the isotherms in the y-z plane are semi-circles*, the situation becomes identical to the tempera-
ture field around a linear instantaneous heat source in a thin plate, provided that the space
above the slab is replaced by solid material and the strength of the source is doubled (see Fig.
1.23(b)). The solution is then given by equation (1-6) if we replace QId by Iq0 Iv:
d-73)
(1-74)
A graphical presentation of equation (1-74) gives a family of curves which resembles the
thermal cycles shown in Fig. 1.19(a). Although the cooling conditions close to the weld cen-
tre-line are similar to those calculated from the Rosenthal equation, the predicted width/depth
of the isotherms will always be greater than that inferred from the general thick plate solution
as illustrated in Fig. 1.24 due to the assumption of 2-D heat flow. The parameter o~4m in Fig.
1.24 is obtained by differentiating equation (1-74) with respect to time:
*The isotherms must meet the plate surface at right angles in the absence of a temperature gradient across the bound-
ary.
(a)
(b)
Symmetry
plane
Fig. 1.23. Fast moving high power source on a semi-infinite slab; (a) Sketch of model, (b) Analogy be-
tween a fast moving high power source and an instantaneous line source.
4m m
Asymptote
0 /n
P 3
Fig. 1.24. Theoretical width of isotherms under 2-D and 3-D heat flow conditions, respectively at pseudo-
steady state (thick plate welding).
which gives
(1-75)
It is interesting to note that the dimensionless width a4m will approach \j/ m when the dp /n3
ratio becomes sufficiently small (i.e. less than about 0.1). Under such conditions the isotherms
will be strongly elongated in the welding direction (see Fig. 1.19(b)), which forces the heat to
flow primarily in directions normal to the x-axis.
A general graphical representation of the weld thermal programme can be obtained by
combining equations (1-74) and (1-75):
(1-76)
Equation (1-76) has been plotted in Fig. 1.25. This graph provides a basis for calculating
the retention time within specific temperature intervals under various welding conditions.
e/ep
W ° J
Fig. 1.25. Temperature-time pattern in thick plate welding at high arc power and high welding speed.
Example (1.8)
Consider SA welding on a thick plate of low alloy steel under the following conditions:
Calculate the retention time within the austenite regime (T > 9100C) for points located lmm
outside the fusion boundary.
Solution
If we use the melting point of the steel as a reference temperature, the parameter n3 becomes:
A comparison with Fig. 1.24 shows that the assumption of 2-D heat flow is justified when
dp < 1. Hence, the width of the fusion zone (8p = 1) can be calculated from equation (1-75):
which gives
The peak temperature lmm outside the fusion boundary is thus:
The total time spent in the thermal cycle from 0 = 0.59 (T= 9100C) to G^ = 0.78 and down
again to 0 = 0.59 can now be read from Fig. 1.25. Taking the ordinate 0/0^ equal to 0.76, we
obtain:
which gives
and
(1-77)
where is the time available for conduction of heat over the distance
to point P.
If we refer the position P to that of the heat source at time t, we shall expect a solution
independent of time at pseudo-steady state. This is achieved by changing the coordinate sys-
tem from O to O' (see Fig. 1.26):
Fig. 1.26. Moving line source in a thin sheet.
K0(U)1K1(U)
U
Fig. 1.27. Graphical representation of the Bessel functions Ko(u) and Kx{u).
Hence, we may write:
d-78)
where
from which
d-79)
It is well-known that:
where K0(G5) is the modified Bessel function of the second kind and zero order. A graphical
representation of A^(W) is shown in Fig. 1.27.
Hence, the general thin plate solution can be written as:
(1-80)
When w is sufficiently large (i.e. when welding has been performed over a sufficient pe-
riod), we obtain the pseudo-steady state temperature distribution:
(1-81)
Equation (1-81) is often referred to as the Rosenthal thin plate solution, in memory of its
originator D. Rosenthal.1314 It follows that this model is applicable to all types of welding
processes (including electron beam, plasma arc and laser welding), provided that a full through-
thickness penetration is achieved in one pass.
(1-82)
Example (1.9)
Consider butt welding of a thin aluminium plate at a constant travel speed of 5mm s"1. Calcu-
late the duration of the transient heating period when the distance from the heat source to the
point of observation is 17mm.
Solution
Taking a = 85 mm2 s"1, the dimensionless radius vector becomes:
It follows from Fig. 1.28 that the pseudo-steady state temperature distribution is approached
when T ~ 5, which gives:
and
This minimum bead length is nearly twice as large as that calculated for 3-D heat flow for
the same combination of welding speed, thermal diffusivity and radius vector (see Example
K2=(T-To)/(T-To)p.s.
t = v2t/2a
Fig. 1.28. Ratio between real and pseudo-steady state temperature in thin plate welding for different
combinations of a 5 and T.
(1.5)). Consequently, the duration of the transient heating period is significantly longer in thin
plate welding than in thick plate welding due the pertinent differences in the heat flow condi-
tions.
(1-83)
Plots of this equation are shown in Fig. 1.29. It follows that the pseudo-steady state tem-
perature distribution in thin plate welding depends on the parameter 68//?3. In practice, this
means that the shape of the isotherms is not influenced by the welding speed, since both 8 and
n3 are proportional to v.
(1-84)
\
\
%
\
V
Fig. 1.29. Dimensionless temperature maps for line sources in thin plates for different ranges of 98/n3.
Heat source
This gives:
(1-85)
and
(1-86)
Equation (1-86) can be used for calculations of isothermal zone widths (v|/m) and cross
sectional areas (A2) in thin plate welding. Referring to Fig. 1.30, we have:
d-87)
and
(1-88)
Figure 1.31 shows a graphical presentation of equations (1-86), (1-87), and (1-88).
(1-89)
and
(1-90)
^5m.Vm-A 2 /8
n3/6p5
Fig. l.Ji.Dimensionless distance a5m, half width \|/m and cross sectional area A2/S vs n3/Qph.
s\-?.*t
n 3 /98
Fig. 1.32. Dimensionless distance from heat source to front (£') and rear (£") of isothermal enclosure vs
«3/68 (thin plate welding).
A graphical presentation of equations (1-89) and (1-90) is shown in Fig. 1.32. Included is
also a plot of the total length of the enclosure £t vs the parameter n3 /08.
(1-91)
Equation (1-91) provides a basis for calculating the cooling time within a specific tempera-
ture interval (e.g. from O1 to 02):
(1-92)
The dimensionless cooling time from 800 to 5000C is thus given by:
(1-93)
(1-94)
Taking as average values X = 0.025 W mm"1 0C"1, pc = 0.005 J mm"3 0C"1, and T0 = 200C
for welding of low alloy steels, we have:
d-95)
(1-96)
(1-98)
Example (LlO)
Consider GTA butt welding of a 2mm thick sheet of cold-rolled aluminium (Al-Mg alloy)
under the following conditions:
Sketch the contours of the fusion boundary and the Ar-isotherm in the £-\|/ (x-y) plane at
pseudo-steady state. The recrystallisation temperature Ar of the base material is taken equal to
2750 C. Calculate also the cross sectional area of the fully recrystallised HAZ and the cooling
rate at 2750C for points located within this region.
Solution
Referring to Fig. 1.22(a) (Example (1.6)) it is sufficient to calculate the coordinates in four
different (characteristic) positions to sketch the contour of the fusion boundary. If we neglect
the latent heat of melting, the n3/68 ratio at the melting point becomes:
End-points
The end-points can be read from Fig. 1.32:
and
Maximum widths
The maximum width of an isothermal enclosure can generally be calculated from equations
(1-86) and (1-87) or read from Fig. 1.31. When n3/QpS = 0.84, we obtain:
and
Intersection point with y/(y)-axis
In this case £ = 0 and cr5 = \j/. Hence, equation (1-83) reduces to:
which gives
Similarly, the contour of the Ar-isotherm can be determined by inserting n3/db = 2.08 into
the same set of equations. Figure 1.33 shows a graphical representation of the calculated
isothermal contours.
§
x(mm)
y(mm)
Fig. 1.33. Calculated contours of fusion boundary and Ar-isotherm in GTA butt welding of a 2mm thick
aluminium plate (Example 1.10).
Cross sectional area of fully recrystallised HAZ
In general, cross sectional areas can be read from Fig. 1.31. Taking the n3/OpS ratio equal to
0.84 (Qp= 1) and 2.08 (8 p = 0.48), respectively, we have:
which gives
(1-99)
According to the assumptions this amount of heat will remain in a rod of constant cross
sectional area due to the lack of a temperature gradient in the welding direction. Under such
conditions the mode of heat flow becomes essentially one-dimensional, and the temperature
distribution is given by equation (1-5):
(1-100)
Equation (1-100) represents the simplified solution for a fast moving high power source* in
a thin sheet, and is valid within a limited range of the more general Rosenthal equation for two-
dimensional heat flow (equation (1-81)).
By substituting the appropriate dimensionless parameters into equation (1-100), we obtain:
* Since the shape of a given isotherm in the x-y plane is determined by the qjd ratio, the minimum welding speed
which is required to maintain 1-D heat flow increases with decreasing qjVd ratios. Hence, the term 'fast moving high
power source' is also appropriate in the case of the thin plate welding.
Fig. 1.34. Fast moving high power source in a thin plate.
(1-101)
The locus of the peak temperatures is readily evaluated from equation (1-101) by setting
3ln(e8M3)/3T = 0:
which gives
and
(1-102)
It is evident from the plot in Fig. 1.35 that the predicted width of the isotherms is always
greater than that inferred from the general thin plate solution (equation (1-83)) due to the
assumption of one-dimensional heat flow. However, such deviations become negligible at
very small Qpb/n3 ratios because of a small temperature gradient in the welding x direction
compared to the transverse y direction of the plate.
A general graphical representation of the weld thermal programme (similar to that shown in
Fig. 1.25 for a fast moving high power source on a heavy slab) can be obtained by combining
equations (1-101) and (1-102):
¥m(1-D)/¥m(a-D)
Asymptote
0 p 8/n 3
Fig. 1.35. Theoretical width of isotherms under 1-D and 2-D heat flow conditions, respectively at pseudo-
steady state (thin plate welding).
e/ep
2t/(vm)2
Fig. 1.36. Temperature-time pattern in thin plate welding at high arc power and high welding speed.
Next Page
(1-103)
Example (LU)
Consider butt welding of a 2mm thin plate of austenitic stainless steel with covered electrodes
(SMAW) under the following conditions:
Calculate the retention time within the critical temperature range for chromium carbide
precipitation (i.e. from 650 to 8500C) for points located at the 8500C isotherm.
Solution
If we use the melting point of the steel as a reference temperature, the parameter n3/5 becomes:
A comparison with Fig. 1.35 shows that the assumption of 1-D heat flow is justified when
Qp< 1. Hence, the total time spent in the thermal cycle from 6 = 0.43 (T = 6500C) to 0p = 0.56
(Tp = 8500C) and down again to 0 = 0.43 can be read from Fig. 1.36. Taking the ordinate 0/0p
equal to 0.76, we obtain:
which gives
and
(1-103)
Example (LU)
Consider butt welding of a 2mm thin plate of austenitic stainless steel with covered electrodes
(SMAW) under the following conditions:
Calculate the retention time within the critical temperature range for chromium carbide
precipitation (i.e. from 650 to 8500C) for points located at the 8500C isotherm.
Solution
If we use the melting point of the steel as a reference temperature, the parameter n3/5 becomes:
A comparison with Fig. 1.35 shows that the assumption of 1-D heat flow is justified when
Qp< 1. Hence, the total time spent in the thermal cycle from 6 = 0.43 (T = 6500C) to 0p = 0.56
(Tp = 8500C) and down again to 0 = 0.43 can be read from Fig. 1.36. Taking the ordinate 0/0p
equal to 0.76, we obtain:
which gives
and
(1-104)
where
Note that equation (1-104) is simply the general Rosenthal thick plate solution (equation
(1-45)) summed for each source.
Fig. 1.37. Real and imaginary point sources on a medium thick plate.
*The number of imaginary heat sources necessary to achieve the required accuracy depends on the chosen values of
R0 and vd/2a.
By substituting the dimensionless parameters defined above into equation (1-104), we ob-
tain:
(1-105)
where
It follows from equation (1-104) that the thermal conditions will be similar to those in a
thick plate close to the centre of the weld. Moreover, Rosenthal1314has shown on the basis of
a Fourier series expansion that equation (1-104) converges to the general thin plate solution
(equation (1-81) for points located sufficiently far away from the source. However, at interme-
diate distances from the heat source, the pseudo-steady state temperature distribution will de-
viate significantly from that observed in thick plate or thin plate welding because of variable
temperature gradients in the through-thickness direction of the plate. Within this 'transition
region', the thermal programme is only defined by the medium thick plate solution (equation
(1-104)).
Y
m
(b)
Thin plate solution
(2-D heat flow) 1-D heat flow
V"3
Y
m
Fig. 1.38. Peak temperature distribution in transverse direction (\|/ = \\fm) of plate; (a) Upper plate surface
(^ = 0), (b) Lower plate surface (J = 8).
1. Close to the heat source, the thermal programme will be similar to that in a thick plate
(Fig. 1.38(a)), which means that the temperature distribution is determined by equation (1-
54). For large values of the dimensionless plate thickness, the mode of heat flow may
become essentially two-dimensional. This corresponds to the limiting case of a fast mov-
ing high power source in a thick plate (equation (1-74)). Under such conditions the slope of
the Qp/n3-\ym curves in Fig. 1.38(a) attains a constant value of-2.
2. With increasing distance from the heat source, a transition from three-dimensional to
two-dimensional heat flow may occur, depending on the dimensionless plate thickness and
the operational conditions applied. Considering the upper surface of the plate (Fig. 1.38(a)),
the extension of the transition region is seen to decrease with increasing values of 8 as the
conditions for thick plate welding are approached. The opposite trend is observed for the
bottom plate surface (Fig. 1.38(b)), since a small dimensionless plate thickness generally
results in a more rapid equalisation of the temperature gradients in the t,(z) direction. When
the curves in Fig. 1.38(b) become parallel with the jc-axis, the temperature at the bottom of
the plate reaches its maximum value. Note that within the transition region, reliable predic-
tions of the pseudo-steady temperature distribution can only be made from the medium
thick plate solution (equation (1-105)).
3. For points located sufficiently far away from the heat source, the temperature gradi-
ents in the through-thickness direction of the plate become negligible. This implies that the
temperature distribution at the upper and lower surface of the plate is similar, and can be
computed from the thin plate solution (equation (1-83)). When the conditions for one-
dimensional heat flow are approached (equation (1-101)), the slope of the dp/n3-\\fm curves
in Fig. 1.38(a) and (b) attains a constant value o f - 1 .
An inspection of Fig. 1.40 reveals a complex temperature-time pattern. In this case it is not
possible to determine the exact field boundaries between the respective solutions, since the
Thin plate solution
(2-D heat flow) (1-D heat flow)
e/n3
e/n3
T= ^
X
Fig. 1.39. Cooling programme for points located on the weld centre-line (\\f = £ =0).
e/n 3
AXf
Fig. 1.40. Total time spent in a thermal cycle from 9 through 9p to 9 for a chosen reference temperature of
9 = 0.59p.
mode of heat flow may vary within a single thermal cycle. Hence, the extension of the differ-
ent fields is not indicated in the graph. The results in Fig. 1.40 provide a systematic basis for
calculating the retention time within specific temperature intervals under various welding con-
ditions.
Isothermal contours
Because of the number of variables involved, it is not possible to present a two-dimensional
plot of the isotherms without first specifying the dimensionless plate thickness. Examples of
calculated isotherms in different planes are shown in Figs. 1.41 and 1.42 for 8 equal to 0.5 and
5, respectively. It is evident that an increase in the dimensionless plate thickness from 0.5 to 5
has a dramatic effect on the shape and position of the isothermal contours. However, in order
to explain these observations in an adequate manner, it is necessary to condense the results into
a two-dimensional diagram. As shown in Fig. 1.43, this can be done by plotting the calculated
field boundaries in Fig. 1.38(a) at maximum width of the isotherms vs the parameters Qp/n3
and vdlla.
It is seen from Fig. 1.43 that a large plate thickness generally will favour three-dimensional
heat flow. With decreasing values of Qp/n3, the conditions for a fast moving high power source
are approached before the transition from the thick plate to the thin plate solution occurs. In
such cases the isotherms at the bottom of the plate will be strongly elongated in the welding £
direction and shifted to positions far behind the heat source. The opposite trend is observed at
small values of vdlla, since a rapid equalisation of the temperature gradients in the through-
thickness direction of the plate will result in elliptical isotherms at both plate surfaces, located
in an approximately equal distance from the heat source. In either case the temperature at
which the cross-sectional isotherms approach a semi-circle or become parallel with the XXz)-
axis can be obtained from Fig. 1.43 by reading-off the intercept between the line for the
dimensionless plate thickness and the respective field boundaries.
(b) The thermal conductivity, density, and specific heat are constant and
independent of temperature.
(c) The plate is completely insulated from its surroundings, i.e. there are no
heat losses by convection or radiation from the boundaries.
(e) The electrode is not consumed during welding, and all heat is concen-
trated in a zero-volume point or line.
(a) ¥
(b)
%
(C)
(b)
%
(C)
\-
Thin plate
solution
Thick plate
solution
(2-D heat flow)
5 = vd/2a
Fig. 1.43. Heat flow mechanism map showing calculated field boundaries in transverse direction (i|/
= i|/m) of plate vs Qp/n3 and 8 = vdlla.
(f) Pseudo-steady state, i.e. the temperature does not vary with time when
observed from a point located in the heat source.
In general, the justification of these assumptions relies on a good correlation between theory
and experiments. However, since the analytical solutions ignore the important role of arc
energy distribution and directed metal currents in the weld pool, predictions of the weld ther-
mal programme should be restricted to positions well outside the fusion zone where such
effects are of less importance (to be discussed below).
Example (1.12)
Consider stringer bead welding (GMAW) on a 12mm thick plate of aluminium (> 99% Al)
under the following conditions:
Based on Fig. 1.43, sketch the peak temperature contours in the transverse section of the
weld at pseudo-steady state.
Solution
If we neglect the latent heat of melting, the parameter n3 at the chosen reference temperature
(Tc = Tm) becomes:
Similarly, when v = 3mm s l and a = 85mm2 s"1 we obtain the following value for the di-
mensionless plate thickness:
0.50 —• 1.0 340 —• 660 Medium thick Heat flow in x and y directions,
plate solution partial heat flow in z direction
0.17 -> 0.50 130 -> 340 Thin plate solution Heat flow in x and y directions,
(2-D heat flow) negligible heat flow in z direction
< 0.17 < 130 Thin plate solution Heat flow in y direction, negligible
(1 -D heat flow) heat flow in x and z directions
A sketch of the HAZ peak temperature contours in the transverse section of the weld is
shown in Fig. 1.44.
Fusion
zone
Fig. 1.44. Schematic diagram showing specific peak temperature contours within the HAZ of an alu-
minium weld at pseudo-steady state (Example 1.12).
(a) WELD A1
y(mm)
z(mm)
(b) WELD A2
y(mm)
z(mm)
Fig. 1.45. Computed peak temperature contours in aluminium welding at pseudo-steady state (Case study
1.1); (a) Weld Al, (b) Weld A2. Black regions indicate fusion zone.
Aluminium welding
In general, the maximum width of the isotherms at the upper and lower surface of the plate can
be obtained from Fig. 1.38(a) and (b), although these maps are not suitable for precise read-
ings. A comparison with the computed peak temperature contours in Fig. 1.45(a) and (b)
reveals a strong influence of the welding speed on the shape and position of the cross-sectional
isotherms at a constant gross heat input of 1.5 kJ mm"1. It is evident that the extension of the
fusion zone and the neighbouring isotherms becomes considerably larger when the welding
speed is increased from 2.5 to 5 mm s"1. This effect can be attributed to an associated shift
from elliptical to more elongated isotherms at the plate surfaces (e.g. see Fig. 1.43), which
reduces heat conduction in the welding direction. It follows from Fig. 1.43 that the upper left
corner of the map represents the typical operating region for aluminium welding.
Because of the pertinent differences in the heat flow conditions, the temperature-time pat-
tern will also vary significantly between the respective series as indicated by the maps in Figs.
1.39 and 1.40. Hence, in the case of aluminium welding the usual procedure of reporting arc
power and travel speed in terms of an equivalent heat input per unit length of the bead is highly
questionable, since this parameter does not define the weld thermal programme. In general,
the correct course would be to specify both qo, v and d, and compare the weld thermal history
on the basis of the dimensionless parameters n3 and 8.
Steel welding
If welding is performed on a steel plate of similar thickness, the operating region will be
shifted to the lower right corner of Fig. 1.43. Under such conditions, the isotherms adjacent to
the fusion boundary will be strongly elongated in the x-direction even at very low welding
speeds (see Fig. 1.42). This implies that the thermal programme approaches a state where the
temperature distribution is uniquely defined by the net heat input r\E, corresponding to the
(a) W E L D S1
y(mm)
z(mm)
(b) WELD S2
y(mm)
z(mm)
Fig. 7.46.Computed peak temperature contours in steel welding at pseudo-steady state (Case study 1.1);
(a) Weld Sl, (b) Weld S2. Black regions indicate fusion zone.
qo V d E n3 8
Material Series (W) (mras"1) (mm) (kJmrrr 1 )
Ordinate:
Abscissa:
Relevant literature data for the cooling time between 800 and 5000C are given in Fig. 1.49.
A closer inspection of the figure reveals a reasonable agreement between observed and pre-
dicted values in all cases. For welding of thick plates, the ordinate attains a constant value of
1. Similarly, in thin sheet welding, the slope of the curve becomes equal to 1.
In aluminium welding, the thermal conditions will be much more complex because of the
resulting higher base plate thermal diffusivity (see Fig. 1.39). Hence, it is not possible to
describe the weld cooling programme in terms of equations (1-74) and (1-101), which apply to
fast moving high power sources. The plot in Fig. 1.50 confirms, however, that the medium
thick plate solution is also capable of predicting the cooling time within specific temperature
intervals (e.g. from 300 to 1000C) in aluminium weldments.
These groups of variables can be obtained from equations (1-74) and (1-101).
Measured
Predicted
Time (sec)
Fig. 1.47. Comparison between measured and predicted weld thermal cycles in aluminium welding for
fixed values of Tp. Data from Myhr and Grong.20
A tr(s), observed
T
P
A tr(s), predicted
Fig. /.4#. Comparison between measured and predicted retention times in aluminium welding for fixed
values of Tp. Data from Myhr and Grong.20
SMAW
SMAW
SAW
SAW
THICK PLATE
1
SOLUTION
n 3 [-i--4-
8/5
AT
THIN PLATE
SOLUTION
-2Hr + Ir 1
5 6SOO 6SOO
Fig. 1.49. Comparison between observed and predicted cooling times from 800 to 5000C in steel welding
(solid lines represent theoretical calculations). Data from Myhr and Grong.20
T(0C)
t(sec)
At (sec), predicted
Fig. 1.50. Comparison between observed and predicted cooling times from 300 to 1000C in aluminium
welding. Data from Myhr and Grong.20
Peak temperatures and isothermal contours
Figure 1.51 shows a comparison between measured and predicted HAZ peak temperatures in
aluminium welding. It is evident that the relative positions of the HAZ isotherms can be
calculated with a reasonable degree of accuracy from the medium thick plate solution, pro-
vided that the equation is precalibrated against a known isotherm (i.e. the fusion boundary).
Additional information on the HAZ peak temperature distribution in aluminium welding
can be obtained from the data of Koe and Lee21 reproduced in Fig. 1.52. These numerical
calculations* showed a good correlation with experimental measurements. A comparison with
the medium thick plate solution in Fig. 1.52 reveals a fair agreement between numerically and
analytically computed isothermal contours. It is interesting to note that even though the plate
thickness is small (i.e. 3.2mm), the mode of heat flow becomes essentially three-dimensional
close to the fusion boundary. This important point is often overlooked when discussing the
relevance of the analytical heat flow models in thin sheet welding.
Tp, 0C , predicted
Fig. 1.51. Comparison between observed and predicted HAZ peak temperatures in aluminium welding.
Data from Myhr and Grong.20
Z (mm)
Fig. 1.52. Comparison between numerically and analytically computed peak temperature contours in
GTA welding of a 3.2mm thin aluminium sheet. (Broken lines: numerical model; solid lines: analytical
model.) Data for welding parameters and material properties are given in Ref. 21.
1. Considering heat flow and temperature distribution in fusion welding, there exists no
defined plate thickness which can be regarded as 'thick' or 'thin'. Accordingly, in a real
welding situation, the mode of heat flow will vary continuously with increasing distance
from the heat source.
2. Close to the centre of the weld, the thermal conditions will be similar to those in a
heavy slab. This means that the temperature distribution is approximately described by the
Rosenthal thick plate solution.
3. At intermediate distances from the heat source, the temperature distribution will de-
viate significantly from that observed in thick plate welding because of variable tempera-
ture gradients in the through-thickness direction of the plate. Within this transition region,
reliable predictions of the pseudo-steady state temperature distribution can only be made
from the medium thick plate solution.
4. For points located sufficiently far away from the heat source, the temperature gradi-
ents in the through-thickness direction of the plate become negligible. Under such condi-
tions, the weld thermal programme is approximately defined by the Rosenthal thin plate
solution.
5. In general, a full description of the weld thermal history requires that both the arc
power qo, the travel speed V, and the plate thickness d are explicitly specified. Hence, the
usual procedure of reporting welding variables in terms of an equivalent heat input per unit
length of the bead, £(kJ mm"1), is highly questionable, since this parameter does not define
the weld thermal programme. An exception is welding of thick steel plates, where the
temperature distribution approaches that of a fast moving high power source because of a
low thermal diffusivity of the base metal.
6. A comparison between theory and experiments shows that the medium thick plate
solution predicts adequately both the peak temperature distribution and the temperature-
time pattern within the HAZ of stringer bead weldments for a wide range of operational
conditions (including aluminium and steel welding). A requirement is, however, that the
equation is calibrated against a known isotherm (e.g. the fusion boundary) due to the sim-
plifying assumptions inherent in the model.
1.10.5 Distributed heat sources
In some cases it is also necessary to consider the important influence of filler metal additions,
arc energy distribution, and convectional heat flow in the weld pool on the resulting bead
morphology to obtain a good agreement between theory and experiments. Of particular inter-
est in this respect is the formation of the so-called weld crater/weld finger, frequently observed
in GMA and SA stringer bead weldments (see Fig. 1.53). Although the nature of these phe-
nomena is very complex, they can readily be accounted for by applying an empirical correc-
tion for the effective heat distribution in the weld pool.
(1-106)
where
y
Crater
Finger
HAZ"
Fusion line
Fig. 1.53. Schematic diagram showing the weld crater/weld finger formation during stringer bead weld-
ing.
(a) y
(b)
Z
Fig. 1.54. General heat flow model for welding on medium thick plates; (a) Physical representation of the
heat distribution by elementary point sources, (b) Method for calculating the temperature field around an
elementary point source displaced along the y-axis.
(C)
Fig. 1.54. General heat flow model for welding on medium thick plates(continued): (c) Method for cal-
culating the temperature field around an elementary (submerged) point source displaced along the z-axis.
Similarly, for a submerged point source located along the z-axis (Fig. 1.54(c))> we obtain:
(1-107)
where
and
Note that equation (1-107) correctly reduces to equation (1-106) when A approaches zero.
The total temperature rise in point P is then obtained by superposition of the temperature
fields from the different elementary heat sources, i.e.:
(1-108)
where
In practice, we can subdivide the heat distributions into a relatively small number of el-
ementary point sources, and usually a total number of 8 to 10 sources is sufficient to obtain
good results (i.e. smooth curves). However, the relative strength of each heat source and their
distribution along the y- and z-axes must be determined individually by trial and error by
comparing the calculated shape of the fusion boundary with the real (measured) one.
Figure 1.55 shows the results from such calculations, carried out for a single pass (bead-in-
groove) GMA steel weld. It is evident that the important effect of the weld crater/weld finger
formation on the HAZ peak temperature distribution is adequately accounted for by the present
model. A weakness of the model is, of course, that the shape and location of the fusion bound-
ary must be determined experimentally before a prediction can be made.
(1-109)
where and
y (mm)
Fig. 1.55. Calculated peak temperature contours in the transverse section of a GMA steel weldment (Op-
erational conditions: / = 450A, U - 30V, v = 2.6mm s"1, d = 50mm).
Fig. 1.56. Distributed heat source of net power density qJ2L on a semi-infinite body (2-D heat flow).
(1-110)
where erf(u) is the Gaussian error function (previously defined in Appendix 1.3).
(1-111)
Dimensionless time:
(1-112)
Dimensionless y-coordinate:
(1-113)
Dimensionless z-coordinate:
(1-114)
(1-115)
Example (1.13)
Consider GMA welding with a weaving technique on a thick plate of low alloy steel under the
following conditions:
Sketch the contours of the fusion boundary and the Ac r isotherm (71O0C) in the y-z plane.
Compare the shape of these isotherms with that obtained from the point heat source model.
Solution
If we neglect the latent heat of melting, the operating parameter at the chosen reference tem-
perature (Tc = Tm) becomes:
Fusion boundary
Here we have:
Ac j-temperature
In this case the peak temperature should be referred to 7200C, i.e.:
ep/nw
V -
Fig. /.57. Calculated peak temperature distribution in the through-thickness direction of the plate at dif-
ferent positions along the weld surface.
ep/nw
y m /L
Fig. 1.58. Calculated peak temperature distribution in the transverse direction of the plate at position y
(Z) = 0.
y m /z m
nw
from which
Similarly, equation (1-75) provides a basis for calculating the width of the isotherms in the
limiting case where all heat is concentrated in a zero-volume point. By rearranging this equa-
tion, we obtain:
which gives
Figure 1.60 shows a graphical presentation of the calculated peak temperature contours.
Implications of model
It is evident from Fig. 1.60 that the predicted shape of the isotherms, as evaluated from equa-
tion (1-110), departs quite strongly from the semi-circular contours required by a point heat
source. Moreover, a closer inspection of the figure shows that inclusion of the heat distribution
also gives rise to systematic variations in the weld thermal programme along a specific
isotherm, as evidenced by the steeper temperature gradient in the v-direction compared with
the z-direction of the plate. This point is more clearly illustrated in Fig. 1.61, which compares
the HAZ temperature-time programme for the two extreme cases of z = 0 and y = 0, re-
spectively. It is obvious from Fig. 1.61 that the retention time within the austenite regime is
considerably longer in the latter case, although the cooling time from 800 to 5000C, Af8/5, is
reasonably similar. These results clearly underline the important difference between a point
heat source and a distributed heat source as far as the weld thermal programme is concerned.
Model limitations
In the present model, we have used the simplified solution for a fast moving high power
source (equation 1-73)) as a starting point for predicting the temperature-time pattern. Since
the equations derived later are obtained by integrating equation (1-73), they will, of course,
apply only under conditions for which this solution is valid.
Moreover, a salient assumption in the model is that the heat distribution during weaving
can be represented by a linear heat source orientated perpendicular to the welding direction.
Although this is a rather crude approximation, experience shows that the assumed heat dis-
z, mm
Fig. /.60. Predicted shape of fusion boundary and Acpisotherm during GMA welding of steel with an
oscillating electrode (Example 1.13). Solid lines: Distributed heat source; Broken lines: Point heat source.
Temperature, 0C
Time, s
Fig. L61. Calculated HAZ thermal cycles in positions y = 0 and z = 0 (Example 1.13).
tribution is not critical unless the rate of weaving is kept close to the travel speed. However,
for most practical applications weaving at such low rates would be undesirable owing to an
unfavourable bead morphology.
Fusion line
z, mm
Stainless steel
y, mm
Weld S4
z, mm
Stainless steel!
y, mm
Weld S5
z, mm
Fig. 1.62. Comparison between observed and predicted Ac3 and Ac1 contours during strip electrode weld-
ing (Case study 1.2). Data from Grong and Christensen.19
Table 1.5 Operational conditions used in strip electrode welding experiments (Case study 1.2).
I U v 2L nw
1 12
WeIdNo. (A) (V) (mms" ) (mm) "V=O- T] = 0.23
Calibration procedure
In general, a comparison between theory and experiments requires that the arc efficiency fac-
tor can be established with a reasonable degree of accuracy. Unfortunately, the arc efficiency
factor for GTA welding has not yet been firmly settled, where values from 0.25 up to 0.75 have
been reported in the literature (see Table 1.3). Additional problems result from the fact that
only a certain fraction of the total amount of heat transferred from the arc to the base plate is
sufficiently intense to cause melting. This has led to the introduction of the melting efficiency
factor T]m, which normally is found to be 30-70% lower than the total arc efficiency of the
process, depending on the latent heat of melting, the applied amperage, voltage, shielding gas
composition, or electrode vertex angle.23 Consequently, since these parameters cannot readily
be obtained from the literature, the following reasonable values for r\m and r\ have been as-
sumed to calculate nw in Table 1.6, based on a pre-evaluation of the experimental data: j \ m =
0.12 (fusion zone), TI = 0.23 (HAZ).
It should be noted that the above values also include a correction for three-dimensional heat
flow, since the assumption of a fast moving high power source during low heat input GTA
welding is not valid. Hence, both the arc efficiency factor and the melting efficiency factor
used in the present case study are seen to be lower than those commonly employed in the
literature.
z, mm
Weld B1
Fusion line
y, mm
Weld B2 z, mm
Fig. 1.63. Comparison between observed and predicted fusion line, Ac3 and Ac1 contours during GTA
welding under full weaving conditions (Case study 1.3). Data from Grong and Christensen.19
experiments is largely improved by inserting 2Leq = 7.5mm into equation (1-110), when com-
parison is made on the basis of the point source model. In contrast, at a heat input of 2.5 kJ
mm"1 (Fig. 1.64(b)), the measured shape of the HAZ isotherms is seen to approach that of a
semi-circle, and hence the deviation between the present model and the simplified solution for
a fast moving high power point source is less apparent.
Intermediate weaving
At intermediate amplitudes of weaving (2L = 5 and 7.5mm, respectively), convectional heat
flow in the weld pool will also tend to increase the bead width to depth ratio beyond the
theoretical value predicted from the present model, as shown in Fig. 1.65. The plot in Fig. 1.65
Weld B3
y, mm
Fusion line
z, mm
Weld B4 y, mm
z, mm
Fig. 1.64. Comparison between observed and predicted fusion line, Ac3 and Ac1 contours during GTA
welding with a stationary arc (Case study 1.3). Solid lines: Distributed heat source, Broken lines: Point
heat source. Data from Grong and Christensen.19
includes all data obtained in the GTA welding experiments with an oscillating arc, as reported
by Grong and Christensen.19 These results suggest that the applied amplitude of weaving must
be quite large before such effects become negligible. Consequently, adaptation of the model to
the weld series considered above would require an empirical calibration of the weaving ampli-
tude similar to that performed in Fig. 1.64 for stringer bead weldments to ensure satisfactory
agreement between theory and experiments.
Theoretical
curve
n
w
Fig. 1.65.Comparison between observed and predicted weld width to depth ratios during GTA welding
with an oscillating arc (Case study 1.3). Data from Grong and Christensen.19
T,e
t,t
Fig. 1.66. Idealised heat flow model for prediction of transient temperatures during interrupted welding.
Model (after Rykalin9)
The situation existing after arc extinction may be described as shown in Fig. 1.66. From time
t = t* there is no net heat supply to the weldment. This condition is satisfied if the real source
q0 is considered maintained by adding an imaginary source +qo and sink -qo of the same
strength at t*. The temperature at some later time t** in a given position R0 (measured from the
origin 0") is then equal to the difference of temperatures due to the positive heat sources qo
and +qo and the negative heat sink -qo. Each of these temperature contributions will be a
product of a pseudo-steady state temperature Tps, and a correction factor K1 or K2 (given by
equations (1-49) and (1-82), respectively). Hence, for 3-D heat flow, we have:
(1-116)
where
and
(1-117)
where
(ro is the position of the weld with respect to the imaginary heat source at time f* in the x —y
plane).
Example (1.14)
Consider repair welding of a heavy steel casting with covered electrodes under the following
conditions:
Suppose that a 50mm long bead is deposited on the top of the casting. Calculate the tem-
perature in the centre of the weld 5 s after arc extinction.
Solution
The pseudo-steady state temperature for points located on the weld centre-line (i|/ = £= 0) can
be obtained from equation (1-65). When t** - t* = 5 s, we get:
Referring to Fig. 1.67, the position of the weld with respect to the imaginary heat source at
time f* is 10mm, which gives:
Fig. 1.67. Sketch of weld bead in Example 1.14.
(a)
3-D heat flow
(b)
3-D heat flow
(C)
Fig. /.6#. Recommended correction factor/for some joint configurations; (a) Single V-groove, (b) Dou-
ble V-groove, (c) T-joint.
Moreover, the dimensionless times T** and T** - x* are:
and
At these coordinates, the correction factor K1 is seen to be 1 and 0.62, respectively (Fig.
1.18). The temperature in the centre of the weld 5 s after arc extinction is thus:
which is equivalent to
(1-118)
Recommended values of the correction factor/for some joint configurations are given in
Fig. 1.68.
Example (Ll5)
Consider deposition of a root pass steel weld in a double-V-groove with covered electrodes
(SMAW) under the following conditions:
Calculate the cooling time from 800 to 5000C (Ar875), and the cooling rate (CR.) at 6500C
in the centre of the weld when the groove angle is 60°.
Solution
The cooling time, Ar875, and the cooling rate, CR., can be obtained from equation (1-68) and
(1-71), respectively:
The above calculations show that the thermal conditions existing in root pass welding may
deviate significantly from those prevailing during ordinary stringer bead deposition due to
differences in the effective heat diffusion area. These results are in agreement with general
experience (see Fig. 1.69).
(1-119)
(1-120)
Fig. 1.69. Comparison between observed and predicted cooling times from 800 to 5000C in root pass
welding of steel plates (groove preparation as in Fig. 1.68(b)). Data from Akselsen and Sagmo.34
Recommended values of k'/p for some arc welding processes are given in Table 1.7. In
practice, the deposition coefficient k'/p will also vary with current density and electrode stickout
due to resistance heating of the electrode. Consequently, the numbers contained in Table 1.7
are estimated averages, and should therefore be used with care.
Example (1.16)
Consider stringer bead deposition (S AW) on a thick plate of low alloy steel under the follow-
ing conditions:
Table 1.7 Average rates of volume deposition in arc welding. Data from Christensen.32
SMAW 0.3-0.5
GMAW, steel 0.6-0.7
GMAW, aluminium -0.9
SAW, steel -0.7
Calculate the mixing ratio Bf(B + D) at pseudo-steady state.
Solution
The amount of fused parent material can be obtained from equation (1-75). If we include an
empirical correction for the latent heat of melting, the dimensionless radius vector a4m be-
comes:
This gives:
Similarly, the amount of deposited metal can be calculated from equation (1-120). Taking
ifc'/p equal to 0.7mm3 A"1 s"1 for SAW (Table 1.7), we get:
Example (L 17)
Consider stringer bead deposition with covered electrodes (SMAW) on a thick plate of low
alloy steel under the following conditions:
Solution
In this particular case the conditions for a fast moving high power source are not met. Thus, in
order to eliminate the risk of systematic errors, the amount of fused parent metal should be
calculated from the general Rosenthal thick plate solution (equation (1-45)) or read from Fig.
1.21. When Tc = Tm (i.e. 8^ = 1), we obtain:
Moreover, the amount of deposited metal can be calculated from equation (1-120). Taking
k'/p equal to 0.4mm3 A"1 s~l for SMAW (Table 1.7), we get:
and
The above calculations indicate a small difference in the mixing ratio between SA and
SMA welding, but the data are not conclusive. In practice, a value of BI(B + D) between 1/3
and 1/2 is frequently observed for SMAW, while the mixing ratio for SAW is typically 2/3 or
higher. The observed discrepancy between theory and experiments arises probably from diffi-
culties in estimating the amount of fused parent metal from the point heat source model.
A summary of Jackson's data is shown in Table 1.8. It is seen that the constant C in equa-
tion (1-121) has a value close to 0.024 for SAW and SMAW with E6015 type electrodes, and
about 0.050 for GMAW with CO2 -shielding gas. Penetration measurements of GMA/Ar + O2,
GMA/Ar, and GMA/He welds, on the other hand, show a strong dependence of polarity, and
shielding gas composition, to an extent which makes the equation useless for a general predic-
tion. Such data have therefore not been included in Table 1.8.
Example (1.18)
Based on the Jackson equation (equation (1-121)), calculate the bead penetration for the two
specific welds considered in Examples (1.16) and (1.17). Use these results to evaluate the
applicability of the point heat source model under the prevailing circumstances.
Solution
From equation (1-121) and Table 1.8, we have:
Table 1.8 Recommended bead penetration coefficients for some arc welding processes. Data from
Jackson.33
and
The corresponding values predicted from the point heat source model are:
and
Provided that the Jackson equation gives the correct numbers, it is obvious from the above
calculations that the point heat source model is not suitable for reliable predictions of the bead
penetration during arc welding. This observation is not surprising.
(1-122)
T
Temperature profile
att = 0
Weld
Preheated
zone
z
Fig. 1.70. Sketch of preheating model.
where T* is the local preheating temperature, and L* is the half width of the preheated zone.
Equation (1-122) can be written in a general form by introducing the following groups of par-
ameters:
— Dimensionless temperature:
(1-123)
Time constant:
(1-124)
Dimensionless time:
(1-125)
(1-126)
(1-127)
It is evident from the graphical representation of equation (1-127) in Fig. 1.71 that the
predicted weld cooling programme falls within the limits calculated for Q"-^ 0 (no preheating)
and Q /7 -> oo (global preheating). The controlling parameter is seen to be the dimensionless
half width of the preheated zone Q", which depends both on the actual width L*, the base plate
thermal properties a, X, and the net heat input qo Iv.
Example (1.19)
Consider stringer bead deposition with covered electrodes (SMAW) on a thick plate of low
alloy steel under the following conditions:
Calculate the cooling time from 800 to 50O0C (A%5), and the cooling time £1Oo measured
from the moment of arc passage to the temperature in the centre of the weld reaches 10O0C.
Solution
First we calculate the time constant to from equation (1-124):
^6
from which
This cooling time is only slightly longer than that calculated from equation (1-68) for T0 =
200C (6.9s), showing that moderate preheating up to 1000C is not an effective method of
controlling Ar875.
The above value should be compared with that evaluated from the numerical data of Yurioka
et al.,35 replotted in Fig. 1.72 (see p.104). It follows from Fig. 1.72 that the weld cooling
programme in practice is also a function of the plate thickness d, an effect which cannot read-
ily be accounted for in a simple analytical treatment of the heat diffusion process. For the
specific case considered above the parameter ^100 varies typically from 500 to 900s, depending
on the chosen value of d. This cooling time is significantly shorter than that calculated from
equation (1-127), indicating that the analytical model is only suitable for qualitative predic-
tions.
References
1. H.S. Carslaw and J.C. Jaeger: Conduction of Heat in Solids; 1959, Oxford, Oxford University
Press.
2. British Iron and Steels Research Association: Physical Constants of some Commercial Steels
at Selected Temperatures; 1953, London, Butterworths.
3. R. Hultgren, R.L. Orr, RD. Anderson and K.K. Kelly: Selected Values of Thermodynamic
Properties of Metals and Alloys; 1963, New York, J. Wiley & Sons.
4. E. Griffiths (ed.): J. Iron and Steel Inst., 1946,154, 83-121.
5. J.E. Hatch (ed.): Aluminium — Properties and Physical Metallurgy; 1984, Metals Park (Ohio),
American Society for Metals.
6. Metals Handbook, 9th edn., Vol. 2, 1979, Metals Park (Ohio), American Society for Metals.
Previous Page
from which
This cooling time is only slightly longer than that calculated from equation (1-68) for T0 =
200C (6.9s), showing that moderate preheating up to 1000C is not an effective method of
controlling Ar875.
The above value should be compared with that evaluated from the numerical data of Yurioka
et al.,35 replotted in Fig. 1.72 (see p.104). It follows from Fig. 1.72 that the weld cooling
programme in practice is also a function of the plate thickness d, an effect which cannot read-
ily be accounted for in a simple analytical treatment of the heat diffusion process. For the
specific case considered above the parameter ^100 varies typically from 500 to 900s, depending
on the chosen value of d. This cooling time is significantly shorter than that calculated from
equation (1-127), indicating that the analytical model is only suitable for qualitative predic-
tions.
References
1. H.S. Carslaw and J.C. Jaeger: Conduction of Heat in Solids; 1959, Oxford, Oxford University
Press.
2. British Iron and Steels Research Association: Physical Constants of some Commercial Steels
at Selected Temperatures; 1953, London, Butterworths.
3. R. Hultgren, R.L. Orr, RD. Anderson and K.K. Kelly: Selected Values of Thermodynamic
Properties of Metals and Alloys; 1963, New York, J. Wiley & Sons.
4. E. Griffiths (ed.): J. Iron and Steel Inst., 1946,154, 83-121.
5. J.E. Hatch (ed.): Aluminium — Properties and Physical Metallurgy; 1984, Metals Park (Ohio),
American Society for Metals.
6. Metals Handbook, 9th edn., Vol. 2, 1979, Metals Park (Ohio), American Society for Metals.
Heat input:
E=1.7 kJ/mm
Fig. 1.72. Cooling time to 1000C, tm, in steel welding for different combinations of T0*, L*, d and E.
Data from Yurioka et a/.35
7. K.P. Bentley, J.A. Greenwood, R McKnowlson and R.G. Bakes: Brit. Weld. J., 1963,10, 613-
619.
8. N.N. Rykalin, A.I. Pugin and V.A. Vasil'eva: Weld. Prod., 1959, 6, 42-52.
9. N.N. Rykalin: Berechnung der Warmevorgdnge beim Schweissen; 1953, Berlin, VEB Verlag
Technik.
10. H.M. Tensi, W. Welz and M. Schwalm: Aluminium, 1981, 58, 515-518.
11. N. Christensen, V. de L. Davis and K. Gjermundsen: Brit. Weld. J., 1965,12, 54-75.
12. Welding Handbook, 8th edn., Vol. 2, 1991, Miami (Florida), American Welding Society.
13. D. Rosenthal: Weld. / , 1 9 4 1 , 20, 220s-234s.
14. D. Rosenthal: Trans. ASME, 1946, 68, 849-866.
15. CM. Adams: Weld. J., 1958, 37, 210s-215s.
16. RS. Myers, O.A. Uyehara and G.L. Borman: Weld. Res. Bull., 1967, 123, 1-46.
17. T.W Eagar and N.S. Tsai: Weld. J., 1983, 62, 346s-355s.
18. M.F. Ashby and K.E. Easterling: Ada Metall, 1984, 32, 1935-1948.
19. 0. Grong and N. Christensen: Mater. ScL Tech., 1986, 2, 967-973.
20. O.R. Myhr and 0 . Grong: Acta Metall. Mater., 1990, 38, 449-460.
21. S. Kou and Y. Le: Metall. Trans., 1983,14A, 2245-2253.
22. O.R. Myhr and 0. Grong: Unpublished work, 1990, University of Trondheim, The Norwegian
Institute of Technology.
23. R.W. Niles and CE. Jackson: Weld. J., 1975, 54, 25s-32s.
24. G.M. Oreper, T.W. Eagar and J. Szekely: Weld J., 1983, 62, 307s-312s.
25. Y.H. Wang and S. Kou: Proc. Int. Conf. on Trends in Welding Research, Gatlinburg, TN, May,
1986, pp. 65-69, Publ. ASM International.
26. S.A. David and J.M Vitek: Int. Mater. Rev., 1989, 34, 213-245.
27. K.C. Mills and BJ. Keene: Int. Mater. Rev., 1990, 35, 185-216.
28. R.L. UIe, Y. Joshi and E.B. Sedy: Metall. Trans., 1990, 21B, 1033-1047.
29. T. Zacharia, S.A. David, J.M. Vitek and H.G. Kraus: Metall. Trans., 1991, 22B, 243-257.
30. A. Matsunawa: Proc. 3rd Int. Conf. on Trends in Welding Research, Gatlinburg, TN, 1992,
pp.3-16, Publ. ASM International.
32. N. Christensen: Welding Metallurgy Compendium, 1985, University of Trondheim, The Nor-
wegian Institute of Technology.
33. CE. Jackson: Weld. J., 1960, 39, 226s-230s.
34. O.M. Akselsen and G. Sagmo: Technical Report STF34 A89147, 1989, Trondheim (Norway),
Sintef-Division of Metallurgy.
35. N. Yurioka, M. Okumura, S. Ohshita and S. Saito: HW Doc. XII-E-10-81, 1981.
Appendix 1.1
Nomenclature
Appendix 1.2
Refined Heat Flow Model for Spot Welding
The refined model is based on the assumption that all heat is released instantaneously at time
t = 0 in a point located at the interface between the two overlapping plates, which implies that
equation (1-7) is valid. However, in order to maintain the net flux of heat through both plate
surfaces equal to zero, it is necessary to account for mirror reflections of the source with
respect to the planes z = dt/2 and z = - dt/2. This can be done on the basis of the method of
images, as illustrated in Fig. A 1.1. By including all contributions from the imaginary sources
"Q-2 >Q-i ,Q\,Qi >••• located symmetrically at distances ± idt below and above the centre-axis
of the joint, the temperature distribution is obtained in the form of a convergent series:
(Al-I)
where
y
Z
Fig. Al.l.Refined heat flow model for spot welding of plates.
Appendix 1.3
The Gaussian Error Function
The eiTor function erf(u) and the complementary error function erfc(u) are special cases of the
incomplete gamma function. Their definitions are:
and
The functions have the following limiting values and symmetries:
and
The following Fortran subroutine can be used for calculations of the error functions with a
fractional error less than 1.2 X 10~7:
FUNCTION ERFC(U)
Z=ABS(U)
T=l./(l.+0.5*Z)
ERFC=T*EXP(-Z*Z-1.26551223+T*(1.00002368+T*(.37409196+
* T%09678418+T*(-.18628806+T%27886807+T*(-l.13520398+
* T*(1.48851587+T*(-.82215223+T*.17O87277)))))))))
IF (U.LT.O.) ERFC=2.-ERFC
RETURN
END
Appendix 1.4
Gaussian Heat Distribution
Following the treatment of Rykalin,9 the situation may be described as shown in Fig. Al.2.
Here we consider a distributed heat source of net power density (in W mm"1):
(Al-2)
The total power of the source qo is obtained by integration of equation (Al-2). Substituting
and integrating from u = -©© to u = +<*>, gives:
from which
(Al-3)
It follows from equation (1-73) that an infinitesimal source dqy> located between j ' a n d
will cause a small rise of temperature dTy> in point P at time t, as:
(Al-4)
where
and
Integration of equation (Al-4) between the limits y'= -°o and >>'= +00 gives:
(Al-5)
q(y)
P
2-D heat flow
z
Fig, Al.2. Distributed heat source of net power density q(y) on a semi-infinite body.
where is a time constant)and n = Aat.
from which
(Al-6)
If we replace the Gaussian heat distribution by a linear source of the same strength, which
extends from -L0 to +L0 on either side of the weld centre-line in the y-direction (see Fig. A1.3),
we may write:
(Al-7)
In practice, the parameter L0 has the same physical significance as the weaving amplitude
L in equation (1-110). Consequently, these solutions are equivalent in the sense that they
predict a similar temperature-time pattern.
q(y)
Fig. Al.3. Physical representation of a Gaussian heat distribution by a linear source of width 2LO.
2
Chemical Reactions in Arc Welding
2.1 Introduction
The weld metal composition is controlled by chemical reactions occurring in the weld pool at
elevated temperatures, and is therefore influenced by the choice of welding consumables (i.e.
combination of filler metal, flux, and/or shielding gas), the base metal chemistry, as well as the
operational conditions applied. In contrast to ladle refining of metals and alloys where the
reactions occur under approximately isothermal conditions, a characteristic feature of the arc
welding process is that the chemical interactions between the liquid metal and its surroundings
(arc atmosphere, slag) take place within seconds in a small volume where the metal tempera-
ture gradients are of the order of 1000°C mm"1 with corresponding cooling rates up to
1000°C s"1. The complex thermal cycle experienced by the liquid metal during transfer from
the electrode tip to the weld pool in GMA welding of steel is shown schematically in Fig. 2.1.
As a result of this strong non-isothermal behaviour, it is very difficult to elucidate the
reaction sequences during all stages of the process. Consequently, a complete understanding
of the major controlling factors is still missing, which implies that fundamentally based pre-
dictions of the final weld metal chemical composition are limited. Additional problems result
from the lack of adequate thermodynamic data for the complex slag-metal reaction systems
involved. However, within these restrictions, the development of weld metal compositions
can be treated with the basic principles of thermodynamics and kinetic theory considered in
the following sections.
The symbols and units used throughout this chapter are defined in Appendix 2.1.
In ladle refining of metals and alloys, the reaction kinetics are usually controlled by mass
transfer between the liquid metal and its surroundings (slag or ambient atmosphere). Exam-
ples of such kinetically controlled processes are separation of non-metallic inclusions from a
deoxidised steel melt or removal of hydrogen from liquid aluminium. In welding, the reaction
pattern is more difficult to assess because of the characteristic non-isothermal behaviour of the
process (see Fig. 2.1). Nevertheless, experience shows that it is possible to analyse mass
transfer in welding analogous to that in ladle refining by considering a simple two-stage reac-
tion model, which assumes:1
(i) A high temperature stage, where the reactions approach a state of local
pseudo-equilibrium.
(ii) A cooling stage, where the concentrations established during the initial stage
tend to readjust by rejection of dissolved elements from the liquid.
Gas nozzle
Contact tube
Shielding gas
Filler wire
Arc plasma
temperature~10000°C Electrode tip droplet
(1600-20000C)
Cold part of Falling droplet (24000C)
weld pool (< 19000C)
Hot part of weld pool
(1900-22000C)
Fig. 2.1. Schematic diagram showing the main process stages in GMA welding. Characteristic average
temperature ranges at each stage are indicated by values in parenthesis.
As indicated in Fig. 2.2 the high temperature stage comprises both gas/metal and slag/metal
interactions occurring at the electrode tip, in the arc plasma, or in the hot part of the weld pool,
and is characterised by extensive absorption of elements into the liquid metal. During the
subsequent stage of cooling following the passage of the arc, a supersaturation rapidly in-
creases because of the decrease in the element solubility with decreasing temperatures. The
system will respond to this supersaturation by rejection of dissolved elements from the liquid,
either through a gas/metal reaction (desorption) or by precipitation of new phases. In the latter
case the extent of mass transfer is determined by the separation rate of the reaction products in
the weld pool. It should be noted that the boundary between the two stages is not sharp, which
means that phase separation may proceed simultaneously with absorption in the hot part of the
weld pool.
In the following sections, the chemistry of arc welding will be discussed in the light of this
two-stage reaction model.
As shown in Table 2.1, gases such as hydrogen, nitrogen, oxygen, and carbon dioxide will be
widely dissociated in the arc column because of the high temperatures involved (the arc plasma
temperature is typically of the order of 10 0000C or higher). From a thermodynamic stand-
point, dissociation can be treated as gaseous chemical reactions, where the concentrations of
the reactants are equal to their respective partial pressures. Hence, for dissociation of diatomic
gases, we may write:
(2-1)
Grey j
zonei
Absorption Rejection of
of elements dissolved elements
Time
Fig. 2.2. Idealised two-stage reaction model for arc welding (schematic).
Table 2.1 Temperature for 90% dissociation of some gases in the arc column. Data from Lancaster.2
CO 2 3800
H2 4575
O2 5100
N2 8300
Next, consider a shielding gas which consists of two components, i.e. one inert component
(argon or helium) and one active component X2. When the fraction dissociated is close to
unity, the partial pressure of species X in the gas phase px is equal to:
(2-2)
where H1 and nx are the total number of moles of components / (inert gas) and X, respectively
in the shielding gas, andptot is the total pressure (in atm).
It follows from equation (2-1) that two moles of X form from each mole of X2 that dissoci-
ates. Hence, equation (2-2) can be rewritten as:
(2-3)
where nXl is the total number of moles of component X2 which originally was present in the
shielding gas.
If nXl and H1 are proportional to the volume concentrations of the respective gas compo-
nents in the shielding gas, equation (2-3) becomes:
(2-4)
Taking vol% / = (100 - vol% X2) andp,ot = 1 atm, we obtain the following expression for
Px-
(2-5)
(2-6)
and
(2-7)
It is evident from the graphical representations of equations (2-5) and (2-7) in Fig. 2.3 that
the partial pressure of the dissociated component X increases monotonically with increasing
concentrations of X2 and YX2 in the shielding gas. The observed non-linear variation of px
arises from the associated change in the total number of moles of constituent species in the gas
phase due to the dissociation reaction. Moreover, it is interesting to note that the partial pres-
sure px is also dependent on the nature of the active gas component in the arc column (i.e. the
stoichiometry of the reaction). This means that the oxidation capacity of for instance CO2 is
only half that of O2 when comparison is made on the basis of equal concentrations in the
shielding gas (to be discussed later).
Px
Vol%)^ f VoRGYX2
(i) Transport of reactants from the bulk phase to the gas/metal interface.
(iii) Transport of dissolved elements from the interface to the bulk of the metal.
(2-8)
where Dx is the diffusion coefficient of the transferring species X (in mm2 s~*).
Although the validity of equation (2-8) may be questioned, the thin film model provides a
simple physical picture of the resistance to mass transfer during gas absorption.
Partial pressure
Distance
(2-9)
where A is the contact area (in mm 2 ), R is the universal gas constant (in mm3 atm K"1 mol"1),
T is the absolute temperature (in K), px is the partial pressure of the dissociated species X in the
bulk phase (in atm), and px is the equilibrium partial pressure of the same species at the gas/
metal interface (in atm).
Based on equation (2-9) it is possible to calculate the transient concentration of element X
in the hot part of the weld pool. Let m denote the total mass of liquid weld metal entering/
leaving the reaction zone per unit time (in g s"1). If Mx represents the atomic weight of the
element (in g mol"1), we obtain the following relation w h e n / ? x » p°x:
(2-10)
It follows from equation (2-10) that the transient concentration of element X in the hot part
of the weld pool is proportional to the partial pressure of the dissociated component X in the
plasma gas. Since this partial pressure is related to the initial content of the molecular species
X2 or YX2 in the shielding gas through equations (2-5) and (2-7), we may write:
Arc column
Bulk gas
phase
Stagnant gaseous
boundary layer
Gas/metal interface
Metal phase
Hot part of
weld pool
Fig. 2.5. Idealised kinetic model for gas absorption in arc welding (schematic).
(2-11)
and
(2-12)
where C1 and C2 are kinetic constants which are characteristic of the reaction systems under
consideration.
Although the above analysis presupposes that the element absorption is controlled by a
transport mechanism in the gas phase, the transient concentration of the active component X in
the hot part of the weld pool can alternatively be calculated from chemical thermodynamics by
considering the following reaction:
By introducing the equilibrium constant K{ for the reaction and setting the activity coeffi-
cient to unity, we get:
(2-14)
This equation should be compared with equation (2-10) which predicts a linear relationship
between wt% X and px. If the above analysis is correct, one would expect that the partial
pressure px at the gas/metal interface is directly proportional to the partial pressure of the
dissociated component in the bulk phase. Unfortunately, the proportionality constant is diffi-
cult to establish in practice.
During the subsequent stage of cooling following the passage of the arc, the concentrations
established at elevated temperatures will tend to readjust by rejection of dissolved elements
from the liquid. When it comes to gases such as hydrogen and nitrogen, this occurs through a
desorption mechanism, where the driving force for the reaction is provided by the decrease in
the element solubility with decreasing metal temperatures.
(2-15)
where k'd is the mass transfer coefficient (in mm s 1X and p°x is the equilibrium partial pres-
sure of component X2 at the gas/metal interface (in atm).
Bulk gas
phase
Stagnant gaseous
boundary layer
Gas/metal interface
Metal phase
Cold part of
weld pool
Fig. 2.6. Idealised kinetic model for gas desorption in arc welding (schematic).
The partial pressure pX2 can be calculated from chemical thermodynamics by considering
the following reaction:
2X(dissolved) = X2 (gas) (2-16)
from which
(2-17)
where K2 is the equilibrium constant, and [wt% X] is the concentration of element X in the
liquid metal (in weight percent). Note that the activity coefficient has been set to unity in the
derivation of equation (2-17).
The equilibrium constant K2 may be expressed in terms of the solubility of element X in the
liquid metal at 1 atm total pressure Sx. Hence, equation (2-17) transforms to:
(2-18)
(2-19)
Data for the solubility of hydrogen and nitrogen in some metals up to about 22000C are
given in Figs. 2.7 and 2.8, respectively. It is evident that the element solubility decreases
steadily with decreasing metal temperatures down to the melting point. This implies that the
desorption reaction is thermodynamically favoured by the thermal conditions existing in the
cold part of the weld pool.
(2-20)
Equation (2-20) is known as the Sievert's law. This relation provides a basis for calculating
the final weld metal composition in cases where the resistance to mass transfer is sufficiently
small to maintain full chemical equilibrium between the liquid metal and the ambient (bulk)
gas phase.
2.7 Overall Kinetic Model for Mass Transfer during Cooling in the Weld Pool
Because of the complexity of the rate phenomena involved, it would be a formidable task to
derive a complete kinetic model for mass transfer in arc welding from first principles. How-
(a) (b)
Aluminium
ml H2/100 g fused metal
Solid Cu
Temperature, 0C Temperature, 0C
(C) (d)
ml H2/100g fused metal
ml H2/100g fused metal
Iron Nickel
Temperature, 0C Temperature, 0C
Fig. 2.7. Solubility of hydrogen in some metals; (a) Aluminium, (b) Copper, (c) Iron, (d) Nickel. Data
compiled by Christensen.4
ever, for the idealised system considered in Fig. 2.9, it is possible to develop a simple math-
ematical relation which provides quantitative information about the extent of element transfer
occurring during cooling in the weld pool. Let [%X]eq denote the equilibrium concentration of
element X in the melt. If we assume that the net flux of element X passing through the phase
boundary A per unit time is proportional to the difference ([%X] - [%X]eqX the following
balance is obtained:3
(2-21)
where V is the volume of the melt (in mm 3 ), kd is the overall mass transfer coefficient (in
mm s"1), and A is the contact area between the two phases (in mm 2 ).
Temperature, 0C
Iron
log (wt% N)
104AT1 K
Fig. 2.8. Solubility of nitrogen in iron. Data from Turkdogan.5
Phase I i
Concentration
Fig. 2.9. Idealised kinetic model for mass transfer in arc welding (schematic).
By rearranging equation (2-21) and integrating between the limife [%X]( (att = O) and [%X]
(at an arbitrary time t\ we get:
(2-22)
(2-23)
It follows that the final concentration of element X in the weld metal depends both on the
cooling conditions and on the intrinsic resistance to mass transfer, combined in the ratio t/to.
When [%X]eq is sufficiently small, equation (2-23) predicts a direct proportionality between
[%X] and [%X\t (i.e. the initial concentration of element X in the weld pool). This will be the
case during deoxidation of steel weld metals where separation of oxide inclusions from the
weld pool is the rate controlling step. Moreover, when t/t0 » 1 (small resistance to mass
transfer), equation (2-23) reduces to:
(2-24)
Under such conditions the final weld metal composition can be calculated from simple
chemical thermodynamics.
Because of this flexibility, equation (2-23) is applicable to a wide range of metallurgical
problems at the same time as it provides a simple physical picture of the resistance to mass
transfer during cooling in the weld pool.
(X-X^)Z(X1-Xeq)
t,s
Some of the well-known harmful effects of hydrogen discussed in Chapters 3 and 7 (i.e. weld
porosity and HAZ cold cracking) are closely related to the local concentration of hydrogen
established in the weld pool at elevated temperatures due to chemical interactions between
the liquid metal and its surroundings.
(i) Loosely bound moisture in the coating of shielded metal arc (SMA) electrodes and in
the flux used in submerged arc (SA) or flux-cored arc (FCA) welding. Occasionally, moisture
may also be introduced through the shielding gas in gas metal arc (GMA) and gas tungsten arc
(GTA) welding.
(ii) Firmly bound water in the electrode coating or the welding flux. This can be in the
form of hydrated oxides (e.g. rust on the surface of electrode wires and iron powder), hydro-
carbons (in cellulose), or crystal water (bound in clay, astbestos, binder etc.).
(iii) Oil, dirt and grease, either on the surface of the work piece itself, or trapped in the
surface layers of welding wires and electrode cored wires.
It is evident from Fig. 2.11 that the weld metal hydrogen content may vary strongly from
one process to another. The lowest hydrogen levels are usually obtained with the use of low-
moisture basic electrodes or GMA welding with solid wires. Submerged arc welding and flux-
cored arc welding, on the other hand, may give high or low concentrations of hydrogen in the
weld metal, depending on the flux quality and the operational conditions applied (note that the
former process is not included in Fig. 2.11). The highest hydrogen levels are normally associ-
ated with cellulosic, acid, and rutile type electrodes. This is due to the presence of large
amounts of asbestos, clay and other hydrogen-containing compounds in the electrode coating.
Table 2.2 (shown on page 132) gives a summary of measured arc atmosphere compositions
in GMA and SMA welding. Included are also typical ranges for the weld metal hydrogen
content.
Fig. 2.11. Ranking of different welding processes in terms of hydrogen level (schematic). The diagram is
based on the ideas of Coe.6
(i) The Japanese method (JIS Z 313-1975), which has been adopted with important ad-
justments from the former ASTM designation A316-48T. This method involves collection of
released hydrogen from a single pass weld above glycerine for 48h at 45 0 C. The total volume
of hydrogen is reported in ml per 10Og deposit. Only 5 s of delay are allowed from extinction
of the arc to quenching.
(ii) The French method (N.F.A. 81-305-1975) where two beads are deposited onto core
wires placed in a copper mould. Hydrogen released from this bead is collected above mercury,
and the volume is reported in ml per 10Og fused metal (including the fused core wire metal).
(iii) The International Institute of Welding (HW) method (ISO 3690-1977), where a single
bead is deposited on previously degassed and weighed mild steel blocks clamped in a quick-
release copper fixture. The weldment is quenched and refrigerated according to a rigorously
specified time schedule. Hydrogen released from the specimens is collected above mercury
for 72 h at 25°C, and the results are reported in ml per 10Og deposit, or in g per ton fused metal.
To avoid confusion, it is recommended to use the symbol HDM for the content reported in terms
of deposited metal (ml per 10Og deposit), and HFM for the content referred to fused metal (ml
per 100 g or g per ton fused metal). The relationship between HDM and HFM is shown in Fig.
2.12.
As would be expected, these three methods do not give identical results when applied to a
given electrode. Approximate correlations have been established between the HW criteria
HDM and HFM and the numbers obtained by the Japanese and the French methods (designated
HJIS and HFR, respectively). For covered electrodes tested at various hydrogen levels, we
have:7
Fig. 2.12. The relation between HDM and HFM (0.9 is the conversion factor from ml per 10Og to g per ton).
(2-25)
(2-26)
The conversion factor from HFR to HFM applies to a ratio of deposited to fused metal,
DI(B + D), equal to 0.6, which is a reasonable average for basic electrodes.
The use of HFM in preference of HDM is normally recommended, because it is a more ra-
tional criterion of concentration. Moreover, HDM values would be grossly unfair, if applied to
high penetration processes like submerged arc welding. In GTA welds made without filler
wire HDM cannot be used at all, since there is no deposit.
It should be noted that the present HW procedure gives the amount of 'diffusible hydrogen'.
For certain purposes the total hydrogen content may be wanted. It is obtained by adding the
content of 'residual hydrogen' determined on the same samples by vacuum or carrier gas
extraction at 6500C. A very small additional amount may be observed on vacuum fusion of the
sample, tentatively labelled 'fixed hydrogen'. There is no clear line of demarcation between
these categories of hydrogen. As will be discussed later, the extent of hydrogen trapping
depends both on the weld metal constitution and the thermal history of the metal. In single-
bead basic electrode deposits the diffusible fraction is usually well above 90%.
(i) An inner zone of very high temperatures which is characterised by absorption of atomic
hydrogen from the surrounding arc atmosphere.
Electrode
Hydrogen
trapped in
weld metal
Weld pool
Fig. 2.13. Idealised reaction model for hydrogen pick-up in arc welding.
(ii) An outer zone of lower temperatures where the resistance to hydrogen desorption is
sufficiently small to maintain full chemical equilibrium between the liquid weld metal and the
ambient (bulk) gas phase.
Under such conditions, the final weld metal hydrogen content should be proportional to the
square root of the initial partial pressure of diatomic hydrogen in the shielding gas, in agree-
ment with Sievert's law (equation (2-20)).
An interesting effect of oxygen on the weld metal hydrogen content has been reported by
Matsuda et al.9 Their data are reproduced in Fig. 2.15. It is evident that the hydrogen level is
significantly higher in the presence of oxygen. This is probably due to the formation of a thin
(protective) layer of slag on the top of the bead, which kinetically suppresses the desorption of
hydrogen during cooling.
Table 2.2 Measured arc atmosphere compositions in steel welding. Also included are typical ranges for
the weld metal hydrogen content. Data compiled by Christensen.4
*The arc atmosphere composition can vary within wide limits, depending on the operational conditions applied.
Example (2.1)
Consider GTA welding (Ar-shielding) on a thick plate of low-alloy steel under the following
conditions:
The shielding gas contains 0.1 vol% moisture (H2O) and is supplied at a rate of 15Nl mhr 1 .
Calculate the 'potential' hydrogen level, assuming that all hydrogen introduced through the
shielding gas is absorbed in the weld metal.
Solution
First we calculate the total mass of hydrogen per mm:
The resulting bead cross section and total mass of weld metal per mm can be estimated
from the Rosenthal equation by considering the dimensionless operating parameter at the melting
point (equation (1-50)):
It is evident from the above calculations that the 'potential' hydrogen level is at least one
order of magnitude higher than the expected weld metal hydrogen content (1 to 3 ppm). This
shows that the hydrogen pick-up in GTA welding is not determined by the total amount of
hydrogen which is introduced through the shielding gas, but is mainly controlled by the result-
ing partial pressure of hydrogen in the ambient (bulk) gas phase.
(2-27)
The parameter pw can be estimated on the basis of combustion measurements of the elec-
trode coating, assuming that no carbon is picked up or lost from the system in excess of the
amount calculated from an analysis of the base plate and the electrode wire. For a recorded
content of mw g H2O and mc g CO2 per 100 g of electrode coating, we obtain:
(2-28)
(2-29)
where K3 is the equilibrium constant for the H 2 O-H reaction, and [%O] is the weld metal
oxygen content. In practice, the correction term ^ K3/(K3+[%0]) does not depart signifi-
cantly from unity, which means that Sw ~ SH.
During welding with basic covered electrodes considerable amounts of CO2 may form as a
result of decomposition of calcium carbonate, according to the reaction:
(2-30)
Modern basic electrodes contain between 20 to 40 weight percent CaCO3, which is equiva-
lent with a CO2 content of 9 to 18 percent. Taking as an average mc equal to 15 g CO2 per 100 g
electrode coating, we obtain:
(2-31)
In Fig. 2.16 the validity of equation (2-31) has been checked against relevant literature data
(compiled by Chew10). A closer inspection of the data reveals that the weld metal hydrogen
content falls within the range calculated for chemical equilibrium at 1520 to 2000°C, taking Sw
equal to the solubility of hydrogen in pure iron at the indicated temperatures (i.e. 27 and 40 ml
H2 per 100 g fused metal, respectively). Although the observed scatter in the effective reaction
temperature is admittedly large, equation (2-31) points out a very interesting effect, namely
that the hydrogen content of SMA steel weld metals is controlled by the combined partial
pressure of H2 and H2O in the ambient gas phase. For this reason it is frequently recom-
mended that calcium carbonate is added to the electrode coating, which on decomposition
produces considerable amounts of shielding gas in the form of CO2. Hydrogen shielding can
also be achieved by additions of volatile alkali-fluorides, which on heating will evaporate and
dilute the atmosphere with respect to hydrogen.
Solution
First we calculate the CO2 content per 100 g of electrode coating. Taking the atomic weight of
CaCO3 and CO2 equal to 100.1 and 40.0, respectively, we obtain:
The combined partial pressure pw can now be estimated from equation (2-28). Before
drying we have:
From Fig. 2.7(c) it is evident that the solubility of hydrogen in liquid iron at 18000C is
about 37 ml H2 per 100 g fused metal. This corresponds to a modified solubility Sw (in ppm)
of:
Substituting this value into the expression for Sievert's law gives:
(before drying)
(after drying)
It follows from the above calculations that a low weld metal hydrogen level requires the use
of 'dry' basic electrodes. In practice, this can be achieved by protecting the electrodes against
moisture pick-up during storage (see Fig. 2.17). However, in certain cases it is necessary to
differentiate between strongly bound and loosely adsorbed moisture in the coating of basic
electrodes. This point is more clearly illustrated in Fig. 2.18, which shows the HDM content of
hydrogen in basic electrode deposits at various levels of coating moisture. It is seen that water
remaining from an insufficient baking treatment is more dangerous than moisture picked up by
exposure of a properly dried coating. This has to do with the fact that loosely adsorbed mois-
Water content in electrode coating, wt%
Fig. 2.17. Moisture content in basic electrode coating as a function of exposure time and relative humid-
ity (R.H.) in ambient gas phase. Data from Evans.11
medium
low
very low
The flux contains 0.04 wt% H2O and is consumed at a rate of 0.6 g per g weld deposit.
Estimate both 'potential' and 'equilibrium' hydrogen levels when the total oxidation loss of
carbon in the weld pool is 0.03 wt%.
Solution
First we calculate the total amount of fused parent metal and weld deposit formed on welding.
From equations (1-75) and (1-120), we have:
and
When the dilution ratio DI(B + D) is known, it is possible to calculate the total flux con-
sumption per gram fused weld metal:
If we assume that all CO produced by reactions between dissolved carbon and oxygen is
infiltrated in the arc column, the following balance is obtained:
Since the effective reaction temperature of hydrogen absorption in SA welding is not known,
the maximum solubility of hydrogen at 1 atm total pressure is taken equal to 33 ppm, similar to
that in the previous example. By inserting this value in the expression for Sievert's law, we
obtain:
In practice, the 'potential' hydrogen level represents an upper limit for the hydrogen con-
centration which cannot be exceeded. Thus, the contradictory results obtained in the present
example clearly illustrate the difficulties involved in estimating the effective partial pressure
of hydrogen in SA welding.
(2-32)
As seen from equation (2-32), the first traces of hydrogen added to the atmosphere are
completely absorbed in the metal. At increasing partial pressures the fraction of hydrogen
picked up in the metal will gradually decrease, finally attaining a threshold of (SH/2) in the
case of pure H2. This shows that the concept of 'potential' hydrogen content frequently used to
characterise filler materials (see Fig. 2.11) is a dangerous one, since the rates of absorption are
so different in the high and low ranges of the hydrogen potential.
HAZ
Distance from fusion line (mm)
Condition Weld Metal Oto 5 5 to 10 10 to 15
Copper: 16500C
Aluminium: 19000C
Nickel: 19000C
At present, it is not known whether these reaction temperatures also apply to conventional
GTA or GMA welding of the same materials or are mainly restricted to the operational condi-
tions employed in the arc melting experiments.
It is generally recognised that interstitial nitrogen embrittles steel (e.g. see discussion in Chap-
ter 7). In steel weld metals the associated loss of toughness due to free nitrogen has been
attributed to solid solution hardening and dislocation locking effects. In addition, excessive
nitrogen pick-up can cause porosity in steel weldments because of gas evolution during solidi-
fication.
2.9.7 Sources of nitrogen
Since the total nitrogen level in most welding consumables and shielding gases is quite low,
the main source of nitrogen contamination is air infiltrated in the arc column. For this reason,
the weld metal nitrogen content is very sensitive to variations in the operational conditions
(e.g. arc length, electrode stick-out, shielding gas flow rate etc.). The overall reaction of nitro-
gen absorption is similar to that of hydrogen:
(dissolved) (2-33)
(2-34)
where SN is the maximum solubility of nitrogen at 1 atm total pressure,/^ is the activity coef-
ficient, and pNl is the resulting partial pressure of diatomic nitrogen in the gas phase.
The solubility of nitrogen in liquid iron is approximately given by:
(2-35)
Similar features are seen from Fig. 2.20(b) for welding of stainless steel. Again, the devia-
tion becomes more pronounced as the oxidation potential of the gas mixture is increased in the
sequence H 2 -Ar-CO 2 -O 2 . Moreover, a comparison with Fig. 2.20(a) reveals that the displace-
ment of the nitrogen concentrations in the presence of chromium is larger than expected from
the calculated reduction of the nitrogen activity coefficient.
The trends observed in Fig. 2.20 have been confirmed by O'Brien and Jordan17 who studied
nitrogen pick-up during CO2-shielded welding of low-alloy steel. As can be seen from Fig.
2.21 (a) their curves are similar to those of Kobayashi et al.16 for short circuiting metal transfer,
while a mixed spray/globular transfer gives a sharp rise of nitrogen absorption up to pNi = 0.3
followed by a constant or slightly decreasing concentration (Fig. 2.21(b)). Both patterns are
clearly not in accordance with predictions based on Sievert's law (equation (2-34)).
An interpretation of the observed trends should be made with a view to absorption of hy-
drogen, where the concept of pseudo-equilibrium has proved useful for a semiquantitative
prediction. In both cases the molecular species H2 and N 2 are known to dissociate in the arc
column (see Table 2.1), and would therefore dissolve in the metal to an extent far beyond the
solubility controlled by pH or PN . The excess of dissolved hydrogen is probably released as
gas at weld pool temperatures. This will also be the case with nitrogen in the absence of oxy-
gen, as shown previously in Fig. 2.20(a) and (b). However, under oxidising conditions the
desorption of gaseous nitrogen becomes suppressed by the presence of oxygen at the gas/metal
interface and hence, nitrogen is retained at a level which by far exceeds the solubility limit at
1 atm total pressure of N2. This has been confirmed experimentally by Uda and Ohno18 in their
classic work on surface active elements (i.e. oxygen, sulphur and selenium) in liquid steel. A
similar phenomenon was quoted in Section 2.8.4.1 from the work of Matsuda et al9 even in
the case of hydrogen, where increased entrapment of hydrogen was observed in the presence
of oxygen (see Fig. 2.15).
It appears thus that excessive absorption of nitrogen (and in some cases also hydrogen)
should be interpreted as a state of incomplete release of solute, as described previously in
Sections 2.6 and 2.7. As a consequence, Sievert's law cannot be used for an estimate of nitro-
gen pick-up in steel welding, unless the weld metal oxygen content is extremely low.
(b)
Fig. 2.20. Nitrogen pick-up in GMA welding at different concentrations of N2 in the shielding gas;
(a) Low-alloy steel, (b) Stainless steel. Data from Kobayashi et al.16
(a)
Low-alloy steel
Nitrogen content, wt%
Experiment
Low-alloy steel
Nitrogen content, wt%
Example (2.4)
Consider SA (single pass) welding on a thick plate of low-alloy steel under the following
conditions:
Based on the 'rule of mixtures', calculate the weld metal nitrogen content. Assume in these
calculations that the nitrogen content of the base plate and the filler wire is 0.005 and 0.012
wt%, respectively.
A B C D
Welder No.
Fig. 2.22. Natural fluctuations in nitrogen pick-up during SMA welding due to variations in the arc
length. Data from Morigaki et al.19
Next Page
loss
gain
Fig. 2.23. Nitrogen pick-up in SA welding at different levels of nitrogen in the electrode wire. Data from
Bhadeshia et a/.20
Solution
First we calculate the total amount of fused parent metal and weld deposit formed on welding.
From equations (1-75) and (1-120), we have:
and
The 'rule of mixtures' gives us the nominal weld metal nitrogen content, which is defined
as:
The above calculations show that the nitrogen content of single pass SA steel welds is close
to that of the base plate because of the high dilution involved. This is in agreement with gen-
eral experience.
Previous Page
Partial oxidation almost invariably accompanies the welding of steel. It is well established that
considerable interaction takes place between the liquid weld metal and its surroundings (arc
atmosphere, slag) when welding is performed in the presence of oxygen. For slag-protected
processes, the flux is the main source of oxygen because of its content of easily-reduced ox-
ides, such as iron oxide, manganese oxide, silica, or rutile. In gas metal arc (GMA) welding,
oxygen is often deliberately introduced through the shielding gas to improve the arc stability
and bead morphology, but at the expense of an increased oxygen content in the weld metal and
intensified losses of alloying elements. The oxidation reactions proceed very rapidly under the
prevailing conditions owing to the high metal temperatures and the large interfacial contact
area available for interactions.
A general survey of oxygen contents in fusion welds is shown in Table 2.5.
Table 2.5 Summary of measured weld metal oxygen contents. Data compiled by Christensen.4
Proper sampling techniques are needed for basic studies of this kind, which allow sampling
of the falling droplets during their flight through the arc column. An additional requirement
is that the speed of quenching is sufficiently high to freeze-in the metal composition estab-
lished at elevated temperatures. If not, spontaneous reactions and subsequent losses of dis-
solved elements due to CO gas formation and manganese silicate slag precipitation may take
place on cooling.
To overcome these problems, a special 'melt spinning' technique has been developed by
Grong and Christensen,1 utilising the same principles as those employed in production of amor-
phous alloy ribbons. By using a water-cooled, fast rotating copper wheel (spinner) as a cath-
ode, rapid crystallisation of droplets released from the electrode tip can be obtained in the
absence of a weld pool*. The reactions which normally occur in the pool during regular multi-
run deposition may then be assessed by comparison of analytical data for chilled metal (i.e.
falling droplets) and normal multi-layer weldment, respectively.
Vol% O 2 in Ar
(b)
Electrode tip
Chilled metal (failing droplet)
Multi-layer weld metal
Electrode wire
wt%c
V o l % C O 2 in Ar
Fig. 2.24. Measured carbon contents in electrode tip droplets, chilled metal and multi-layer weld deposit
vs the oxygen potential of the shielding gas; (a) Ar-C>2 gas mixtures, (b) Ar-CC>2 gas mixtures. Data from
Grong and Christensen.1
the former case. Normally, homogeneous nucleation of CO gas within the liquid metal is con-
sidered impossible, which means that the CO nucleation in practice must take place heteroge-
neously. However, the most probable site for CO evolution during droplet formation at the
electrode tip will be the gas/metal interface itself, which allows carbon to be oxidised simulta-
neously with silicon and manganese. It is reasonable to assume that most of the observed
carbon oxidation is located to the hot layers facing the arc, where the reaction is thermody-
namically favoured. At other surface positions Si and Mn are expected to prevent carbon from
reacting due to a rather low metal temperature, stated to be only slightly above the melting
point at the time of detachment.21
It is evident from the data in Fig. 2.24(a) and (b) that the carbon losses increase with in-
creasing O2 or CO2 contents in the shielding gas up to a certain critical level. Hence, supply of
oxygen to the tip droplet surface is the rate controlling step for oxidation of carbon at low
oxygen potentials. This conclusion is also consistent with calculations made by Corderoy et
al.,22 who found that transport of atomic oxygen through a stagnant gaseous boundary layer
close to the metal surface controls the oxidation rate of alloying elements at this stage of the
process.
The carbon oxidation will gradually decline with increasing oxygen concentrations in the
shielding gas, probably as a result of build-up of carbon monoxide in the surrounding atmos-
phere. When the critical CO gas pressure is reached, the carbon reaction is blocked, silicon
(and manganese) now exercising control of the oxygen level, as indicated in Fig. 2.25. For Ar-
O 2 gas mixtures this critical pco pressure is attained at about 10 to 15 vol% O 2 in the shielding
gas, corresponding roughly to 0.05 wt% C oxidised in chilled metal. When welding is per-
formed in Ar-CO 2 mixtures the reaction is blocked at a much earlier stage of carbon oxidation
(equal to about 0.02 wt% C lost in chilled metal), since dissociation of CO2 in this case will
produce an additional amount of CO to concentrate in the surrounding gas phase.
atm
Silicon control
Carbon control
Pco>
Temperature, 0C
Fig. 2.25. Break even equilibrium partial pressure of CO vs temperature for silicon control of oxygen
level at 0.8 wt% Si and silica saturation. Data from Elliott et al. 23
The mechanism indicated above is supported by the data presented in Fig. 2.26, which
show that a CO2-rich atmosphere even may act carburising if the initial carbon content of the
electrode wire is sufficiently low. Moreover, it is interesting to note that the carbon-oxygen
reaction is also influenced by the rate of droplet detachment. Since the highest carbon oxida-
tion losses are normally associated with a coarse globular droplet transfer mode, this suggests
that the reaction time is more important than the effective contact area available for interaction
which depends on the droplet size.
(Gain)
(Loss)
Vol% O 2 in Ar
(b)
Chilled metal (falling droplet)
Multi-layer weld metal
Electrode wire
Wt% Si
Vol% CO 2 in Ar
Fig. 2.27. Measured silicon contents in chilled and multi-layer weld metals vs the oxygen potential of the
shielding gas; (a) Ar-02 gas mixtures, (b) Ar-CC>2 gas mixtures. Data from Grong and Christensen.1
product. Evaporation losses can in this case be excluded due to a very low vapour pressure of
silicon at the prevailing temperatures. It is therefore reasonable to assume that SiO(g) forms as
a result of chemical reactions occurring at the electrode tip. The observed decrease in chilled
metal Si content with increasing O2 and CO2 contents in the shielding gas is probably caused
by the presence of CO at the gas/metal interface, which facilitates SiO formation according to
the reaction:
(slag) (gas) (gas) (gas) (2-36)
Figure 2.28 shows a plot of the equilibrium partial pressure of SiO vs temperature at silica
saturation for three different CO levels. Note that stoichiometric amounts of CO2 have been
assumed to form in order to calculate psi0. It is evident from the thermodynamical data pre-
sented in Fig. 2.28 that the formation of SiO is strongly dependent on the CO partial pressure
at the slag/metal interface. Thus, from a thermodynamic standpoint the silicon loss at the elec-
trode tip should be most pronounced during CO2-shielded welding due to the resulting higher
CO pressure. The data shown in Fig. 2.27(a) and (b) support this assumption. It is seen that the
silicon loss is increased by a factor of 3 to 5 in presence of CO2 when comparison is made on
the basis of equal oxygen potential of the shielding gas (i.e. equal loss of deoxidants in the
weld pool). Similar observations have also been made by Heile and Hill27 from determination
of silicon in collected GMA welding fumes. The recorded chilled metal Si loss in Fig. 2.27(a)
and (b) is in good agreement with the reported fume formation rates of silicon. Since all CO
consumed in reaction (2.36) is expected to be regenerated immediately by decomposition of
CO2 at the metal surface, the high CO partial pressure required for SiO formation at the elec-
trode tip is maintained even in the case of extensive silicon losses. Consequently, the net reac-
tion can be written as:
PSiO,10-3atm
Temperature, 0C
Fig. 2.28. Equilibrium partial pressure of SiO vs temperature at different CO levels. Data from Refs. 25
and 26.
(slag) (dissolved) (gas) (2-37)
When welding is performed in Ar-O 2 gas mixtures, pSiO andpco may be taken proportional
to the recorded loss of silicon and carbon in chilled metal (see data in Fig. 2.29). It is evident
from this plot that the silicon loss is directly proportional to the corresponding loss of carbon
up to a certain critical level. Thus, during the initial period of carbon oxidation at the electrode
tip the SiO formation is probably controlled by the resulting partial pressure of CO at the slag/
metal interface, according to reaction (2-36). When the carbon reaction is blocked, the chilled
metal silicon loss becomes independent of the CO partial pressure (see reaction (2-37)), since
all CO consumed in the SiO formation will immediately be recirculated within the system.
However, at the break even point for silicon control of the oxygen level at the electrode tip, the
CO pressure in the surrounding gas phase will be the same for both Ar-O 2 and Ar-CO 2 gas
mixtures. Hence, the recorded loss of silicon in chilled metal at 20 vol% CO2 in Ar is seen to be
similar to that in Ar + 10 vol% O2, as indicated by the heavy broken line in Fig. 2.29.
At the temperatures where liquid steel is normally deoxidised, silicon and manganese have
a strong affinity to oxygen. Their ability to form stable oxides decreases rapidly with increas-
ing temperature, and above approximately 180O0C silicon and manganese do no longer act as
efficient deoxidation agents. Precipitation of manganese silicate slags is therefore favoured by
the lower metal temperatures prevailing at the electrode tip and in the cold part of the weld
pool. At higher temperatures, these oxides become unstable. Consequently, as a result of the
metal superheating occurring during droplet transfer through the arc column, the macroscopic
slag phase formed earlier at the electrode tip surface (as reported by Corderoy et ah12) will
redissolve in the metal. This gives rise to a relatively high chilled metal oxygen content (to be
discussed below).
Ar+10vol%O2
Based on the data presented in Fig. 2.27(b), calculate the fume formation rate (FFR) of
silicon (in mg per min) due to SiO formation at the electrode tip. The wire feed rate is 125mm s"1.
Solution
The total loss of silicon due to SiO formation may be taken equal to the observed difference
between the filler wire and the chilled metal silicon contents. For welding in pure CO2, we get:
The corresponding fume formation rate of silicon (in mg min"1) can readily be calculated
when the wire feed rate (WFR) is known. Taking the density of the steel equal to 7.85mg mm"3,
we obtain:
A comparison with the measured FFR of silicon in Table 2.6 (at / = 13OA) shows that the
calculated value is reasonable correct. Moreover, these data support our previous conclusion
that the SiO formation is favoured by a high CO2 content in the shielding gas due to the
dissociation reaction. In fact, more detailed studies of the reaction kinetics have confirmed that
the rate of SiO formation is proportional to the resulting partial pressure of CO at the gas/metal
interface,28 in agreement with equation (2-36).
250 29 1 1 22
Argon 300 31 1 0 12
350 35 5 4 51
150 28 15 8 134
Ar+ 2 % O 2 200 28 12 10 75
300 29 5 4 35
400 34 18 16 86
100 28 29 33 273
Ar+ 5% O 2 200 28 16 22 129
300 28 10 23 76
100 23 11 20 75
Ar + 25% CO2 150 27 13 35 105
300 35 29 62 191
130 27 13 59 86
150 30 14 63 120
Pure CO2 200 30 19 73 126
250 30 31 112 216
300 30 36 125 214
in chilled metal, corresponding to a total loss of 1.7 wt% (Fe + Mn) or 500mg (Fe + Mn) per
min. The reported fume formation rate of Fe and Mn under nearly similar conditions is only
one half of that stated above, which indicates that the calculated flux of metal vapour may be
somewhat overestimated. However, if these two values are considered to represent borderline
cases, the volume of metal vapour is in the range from 300 to 500 times that of the droplets
and will therefore be more than sufficient to protect the metal from arc atmosphere oxidation
at this stage of the process. This conclusion is consistent with statements made by Distin et al?1
who claim that iron vapour acts as an effective oxygen getter already at about 19000C.
Very large amounts of manganese are also lost in the weld pool stage, as shown by the
difference between the measured Mn contents in chilled and multi-layer weld metals. This
situation appears to be quite similar to that of silicon. For Ar-O 2 mixtures the manganese loss
increases steadily with increasing oxygen contents in the shielding gas, and reaches a value of
about 1.08 wt% Mn (or 0.31 wt% O) removed from the weld pool at 30 vol% O2 in Ar. In the
case OfAr-CO2 shielded welding the Mn oxidation loss starts to drop off at a CO2 content of
about 20 vol%, finally attaining an upper limit of about 0.50 wt% Mn corresponding to 0.15
wt% O removed from the weld pool.
A more detailed discussion of oxidation reactions in GMA welding is given in Section
2.10.1.5.
Example (2.6)
Consider GMA welding low-alloy of steel under conditions similar to those in Example 2.5.
(a)
Chilled metal
Wt% Mn
(falling droplet)
Vol% O 2 in Ar
(b)
Wt% Mn
Vol%CO 2 inAr
Fig. 2.30. Measured manganese contents in chilled and multi-layer weld metals vs the oxygen potential
of the shielding gas; (a) Ar-02 gas mixtures, (b) Ar-CO2 gas mixtures. Data from Grong and Christensen.1
Based on the data presented in Fig. 2.30(b), calculate the fume formation rate (FFR) of manga-
nese due to evaporation losses occurring during droplet transfer through the arc column. Esti-
mate also the effective mass transfer coefficient for manganese evaporation under the prevail-
ing circumstances by utilising the vapour pressure data in Fig. 2.31. The surface temperature
of the falling droplets is assumed constant and equal to 26000C.
LogpMn,atm
PMr/PFe
Temperature, 0C
Fig. 2.31. Equilibrium manganese vapour pressure and corresponding vapour pressure ratio pMn to PFe vs
temperature at 1.27 wt% Mn in iron. Data from Kubaschewski and Alcock. 30
Solution
The total loss of manganese due to evaporation may be taken equal to the observed difference
between the filler wire and the chilled metal manganese contents. For welding in pure CO2, we
have:
The corresponding fume formation rate of manganese (in mg min"1) can readily be calcu-
lated when the wire feed rate (WFR) is known. Taking the density of the steel equal to 7.85mg
mm~3, we obtain:
A comparison with the data in Table 2.6 reveals that the reported fume formation rate of
manganese (at / = 13OA) is much lower than computed in the present example. This discrep-
ancy can probably be attributed to differences in the filler wire manganese content. Assuming
that the evaporation loss of manganese is controlled by a transport mechanism in the gas phase,
we can estimate the effective mass transfer coefficient from equation (2-15) by inserting a
reasonable average value for the manganese vapour pressure at the gas/metal interface (pMn in
the bulk phase is taken equal to zero). Reading from Fig. 2.31 (at T= 26000C) gives:
If the diameter of the falling droplets is taken equal to the diameter of the filler wire, it is
possible to calculate the total loss of manganese associated with one droplet (in mol):
Since the average flight time of large, globular droplets through the arc column is of the
order of one second, the corresponding flux of manganese vapour per unit time is close to 9.97
X 10~8 mol s"1. Thus, by rearranging equation (2-15), we obtain the following value for the
effective mass transfer coefficient:
Although the above value is rather uncertain, the calculated mass transfer coefficient is of
the expected order of magnitude. This supports our previous conclusion that the evaporation
kinetics are controlled by a transport mechanism in the gas phase.
(2-38)
Taking the Si to Mn mass ratio in precipitated slag equal to 0.66,22 the corresponding loss of
silicon is equal to:
(2-39)
Based on equations (2-38) and (2-39) it is possible to calculate the oxygen absorption at the
electrode tip which is associated with the MnO and the SiO2 slag formation. If corrections also
are made for the amount of oxygen simultaneously removed as iron oxide*, we obtain:
(2-40)
*An approximate correction can be made from an analysis of the Fe to Mn mass ratio in precipitated slag, which under
the prevailing circumstances is close to 0.16.22
(a)
Chilled metal (falling droplet)
Multi-layer weld metal
Wt% O Electrode wire
Vol%O2inAr
(b)
Vol% CO2 in Ar
Fig. 2.32. Measured oxygen contents in chilled and multi-layer weld metals vs the oxygen potential of
the shielding gas; (a) Ar-C>2 gas mixtures, (b) Ar-CO2 gas mixtures. Data from Grong and Christensen.1
(2-41)
It is seen from the graphical representation of equation (2-41) in Fig. 2.33(a) and (b) that
most of the oxygen pick-up takes place in the weld pool. However, at present it is not clear
whether the calculated values represent a real transient concentration in the hot part of the
weld pool or a number in concentration units representing precipitation of manganese silicate
slags. According to Fischer and Schumacher33 the solubility of oxygen in liquid iron is 0.94
wt% at 19000C and 1.48 wt% at 20000C This temperature range is probably relevant with re-
spect to the hot part of the weld pool, although the surface metal temperature will be even
higher.34 However, in steel weld metals the oxygen concentrations should be well below this
solubility limit in the presence of silicon and manganese. Taking the average weld pool silicon
content equal to 0.50 wt%, the corresponding oxygen content in equilibrium with a silica-
saturated slag is roughly 0.10 and 0.20 wt% at 1900 and 20000C, respectively.23 On the other
hand, under the prevailing circumstances there will be no solid/liquid interface available for
nucleation of oxide particles, and hence homogeneous nucleation is the only possibility.
According to Sigworth and Elliott35 this requires a relatively high degree of supersaturation,
which means that the oxygen concentration most likely will exceed the Si-Mn-O equilibrium
value before slag precipitation occurs.
The data in Fig. 2.33(a) and (b) indicate that the weld pool oxygen absorption is controlled
by a complex transport mechanism in the gas phase. However, since mass transfer in gas-jet/
liquid systems is not fully understood,3 we shall only consider the limiting case where the
resistance to mass transfer is confined to a stagnant gaseous boundary layer adjacent to the
metal surface. Under such conditions equation (2-10) predicts that the transient oxygen con-
centration is determined by the partial pressure of atomic oxygen in the plasma gas, as shown
Vol% O2 in Ar
(b)
Calculated total oxygen
absorption
Analytical weld metal
oxygen content
Wt% O
Vol% CO2 in Ar
Fig. 2.33. Calculated total oxygen absorption in GMA welding at different oxygen potentials of the
shielding gas; (a) Ar-O 2 gas mixtures, (b) Ar-CO 2 gas mixtures. The analytical weld metal oxygen
content is indicated by the broken lines in the graphs. Data from Grong and Christensen.1
by the plots in Fig. 2.34. If we as a borderline case assume that all oxygen present in the plasma
gas is immediately absorbed in the liquid metal, the slope of the curve in Fig. 2.34 for Ar-O 2
gas mixtures (equal to about 2.67 wt% dissolved oxygen at one atmosphere total pressure of
atomic oxygen) is representative of the total amount of gas passing through the arc column.
Taking the mass of liquid metal leaving/entering the reaction zone equal to 3Og min"1 under
the prevailing circumstances,1 the following gas flow rate is obtained (in Nl min"1).
This value corresponds to about 6 per cent of the total shielding gas flow rate.
In contrast to the situation described in Fig. 2.34 for Ar-O 2 gas mixtures, a large deviation
from the expected relationship is observed for welding in CO2-rich atmospheres. At present,
the reason for this shift in the reaction kinetics is not known. However, it is evident from the
data presented in Fig. 2.34 that the effective mass transfer coefficient decreases by a factor of
about 2.5 when the CO2 content in the shielding gas increases from 10 to 100 vol%. This
implies that the oxidation capacity of pure CO2 is comparable with that of Ar + 13 vol% O2,
although the partial pressure of atomic oxygen in the plasma gas is equivalent with an oxygen
content of about 33 vol%.
Example (2.7)
Consider GMA welding of low-alloy steel in Ar + 20 vol% CO2 and pure CO2, respectively.
Based on the results in Figs. 2.27(b) and 2.30(b), calculate the total weight of top bead slag (in
gram per 100 gram weld deposit) which forms as a result of deoxidation reactions. Assume in
these calculations that the iron to manganese mass ratio in precipitated slag is equal to 0.1.
[%O]t0t,wt%
Pure CO2
P0, atm
Fig. 2.34. Calculated total oxygen absorption in GMA welding at different partial pressures of atomic
oxygen in the plasma gas. Data from Grong and Christensen.1
Solution
The amount of silicon and manganese lost as a result of deoxidation reactions is equal to the
observed difference in the chilled and multi-layer weld metal Si and Mn contents. Assuming
that these elements are removed from the weld pool as SiO2 and MnO, respectively, the fol-
lowing balance is obtained:
Thus, for welding in Ar + 20 vol% CO2, the total weight of slag amounts to:
A comparison with the experimental data in Fig. 2.35 shows that there is a fair agreement
between the amount of slag recorded by weighing and that calculated from a simple mass
balance of silicon, manganese and iron.
Vol% CO2 in Ar
Fig. 2.35. Comparison between measured and calculated weight of top bead slag in CO2-shielded weld-
ing. Data from Grong and Christensen.1
2.10.1.6 Classification of shielding gases
The data in Fig. 2.34 provide a basis for evaluating the oxidation capacity of various shielding
gases. For welding in Ar + O2 and Ar+CO2 gas mixtures up to 10 vol% CO2 in argon, the total
oxygen absorption is approximately given by the following equation:
(2-42)
Similarly, for welding with CO2-rich shielding gases (i.e. between 10 and 100 vol% CO2),
we obtain:
(2-43)
Equal oxidation capacity means that the total weld metal oxygen absorption is the same for
both shielding gas mixtures. Hence, we may write:
(2-44)
when the CO2 content in the shielding gas is less than 10 vol%, and
(2-45)
when welding is performed with CO2-rich shielding gases (more than 10 vol% CO2 in Ar).
Based on equations (2-44) and (2-45) it is possible to compare the oxidation capacity of
various shielding gases (see Table 2.7). Included in Table 2.7 is also a slightly modified ver-
sion of the International Institute of Welding (HW) classification system,36 which is based on
an evaluation of retained (analytical) oxygen in the weld deposit. It is evident from these data
that both systems are applicable and mutual consistent, although the former one utilises a more
rational criterion for the shielding gas oxidation capacity.
Example (2.8)
Consider GMA welding of low-alloy steel under the following conditions:
Table 2.7 Proposed shielding gas classification scheme for GMA welding of low-alloy steel according to
equations (2-44) and (2-45). Included is also a modified version of the corresponding IIW's classification
system.36
The analytical weld metal oxygen content (in wt%) is given by the values in brackets.
The shielding gas is pure CO2 and is supplied at a constant rate of 15Nl min *. Based on the
composition data in Table 2.8 calculate the resulting CO content in the welding exhaust gas.
Solution
First we calculate the nominal weld metal chemical composition by neglecting oxidation loss
of alloying elements due to chemical reactions:
The dilution ratios BI(B + D) and DI(B + D) can be estimated from the classic heat flow
theory presented in Chapter 1. From equations (1-75) and (1-120), we have:
Table 2.8 Chemical composition of filler wire, base plate and weld metal used in Example 2.8.
C O Si Mn
Element (wt%) (wt%) (wt%) (wt%)
This gives:
The extent of gas/metal interaction can then be evaluated from the observed concentration
displacements:
Calculated values for the concentration displacements of carbon, oxygen, silicon and man-
ganese utilising the composition data in Table 2.8 are given below.
Element C O Si Mn
\-%X\nom. 0
- 1 2 0
- 0 0 8 0
' 6 3 L 3 9
The total CO evolution (in mol min l) can now be computed from an overall oxygen bal-
ance for the reaction system. In these calculations we shall assume that Si and Mn lost as
SiO(g) and Mn(vap.) immediately react with CO2 to form SiO2 and MnO, respectively*. Tak-
ing the density of steel equal to 7.85 X 10~3 g mm"3, the total mass of weld metal produced per
unit time amounts to:
Oxidation of carbon:
Oxidation of silicon:
:
CO 2 consumed in oxidation of iron vapour is disregarded.
Oxidation of manganese:
Total CO evolution (sum):(13.4 + 53.5 + 28.3 + 9.6) X 10~3 mol CO min - 1 = 104.8 X 10~3 mol
CO min' 1
Based on this information it is possible to calculate the resulting CO content in the welding
exhaust gas:
A comparison with the data in Table 2.2 shows that a CO content of about 15 vol% is
reasonably close to that determined by analysis.
Amperage
When the welding current is raised, the time available for interaction decreases due to the more
rapid detachment of the electrode tip droplets. At the same time the interfacial contact area
increases as the average droplet size becomes smaller. From measurements of fume formation
rates in GMA welding,27 it has been shown that these two counteracting effects will almost
cancel, i.e. the total amount of emitted dust (in mg per g deposit) is found to be constant and
nearly independent of the applied amperage. On the other hand, the total fume formation rate
is probably not a reliable index for the burn-off of Si and Mn, since the evolution of iron
vapour during droplet transfer will tend to conceal the corresponding loss of alloying ele-
ments.
The effect of amperage (or more correct the droplet detachment frequency) on the burn-off
of carbon, silicon, and manganese in CO2-shielded welding has been investigated by Smith et
al.38 They found that the recovery of alloying elements in the weld deposit increased with
increasing welding current (i.e. droplet detachment frequency). In view of the previous discus-
sion, it is reasonable to assume that the higher weld metal carbon and silicon contents reported
by Smith et al3S are a result of a reduced CO and SiO gas evolution at the electrode tip due to
the shorter time available for chemical interaction. In the case of manganese reduced evapora-
tion losses because of a more rapid transfer of the droplets through the arc column offers a
reasonable explanation to the increased element recovery. This shows that the weld metal
chemistry is sensitive to variations in the welding current.
Arc voltage
Since the arc voltage neither affects the melting rate nor the droplet size to any great extent,39
variations in the arc voltage should only have a minor effect on the weld metal chemistry. This
conclusion is apparently in conflict with observations made by Lindborg,40 who found that the
oxidation reactions in GMA welding were strongly voltage dependent and at the same time
independent of the welding current, the droplet detachment frequency, and the mode of metal
transfer (spray or short-circuiting). Consequently, further investigations are required to ex-
plain these discrepancies.
Welding speed
It can be inferred from the data of Grong and Christensen1 that the analytical weld metal
carbon and oxygen contents are virtually independent of the welding speed v within the nor-
mal range of GMAW (i.e. from 0.4 to 6 mm s"1). However, the intensified losses of silicon and
manganese observed at low welding speeds indicate that more oxygen is absorbed in the weld
pool under such conditions. This point is more clearly illustrated in Fig. 2.36 which shows a
plot of [%O]tot vs v for a series of multi-pass GMA welds deposited under the shield of Ar +
10 vol% O2. It is evident that the total oxygen absorption increases nearly by a factor of two
when the welding speed decreases from 6 to 0.4mm s"1. This shows that the welding speed has
a marked effect on the transient oxygen pick-up in the hot part of the weld pool during GMA
welding, since it controls the time available for element absorption.
where MxOy denotes any oxide component in the slag phase (e.g. SiO2, MnO or FeO).
The basicity index (B.I.), originally adopted from steel ladle refining practice, is most fre-
quently employed for assessment of oxygen pick-up in SA welding, since it gives an approxi-
mate measure of the flux oxidation capacity. A number of different expressions exists in the
literature, but for the purpose of convenience the basicity index defined by Eagar47 has been
adopted here:
basic oxides"
non- basic oxides
(2-47)
Acid fluxes
(2-48)
and
(2-49)
If we assume that rejection of Si and Mn in the weld pool occurs as a result of MnSiO3
microslag precipitation and subsequent phase separation, [%Mn]rej^ is bound to [%Si]rej. through
the following stoichiometric relationship:
(2-50)
Taking the ratio between absorbed Mn and Si in the weld metal equal to k, a combination of
equations (2-48), (2-49) and (2-50) gives:
(2-51)
The value of k is difficult to evaluate in practice, but in view of the reported mass transfer
coefficients for Mn and Si a reasonable estimate would be about 0.5 in the case of manganese
silicate fluxes.46 Under such conditions the total oxygen absorption, [%O]abs is given by:
(2-52)
where [A%O] is the observed concentration displacement of oxygen in the weld metal, and
[%O]rej. is the amount of oxygen rejected from the weld pool as a result of deoxidation reac-
tions.
Based on equation (2-52) it is possible to estimate the total oxygen absorption during SA
welding of C-Mn steels from an analysis of measured concentration displacements of oxygen,
silicon, and manganese in the weld metal. The results of such calculations are shown graphi-
cally in Fig. 2.38, using data from Indacochea et a/.44 It is evident from this plot that the total
oxygen absorption during SA welding is much larger than that inferred from an analysis of
retained oxygen in the weld deposit. The situation is thus quite similar to that observed experi-
mentally in GMA welding (see Fig. 2.33).
It should be noted that the calculated values for [%O]abs. in Fig. 2.38 may be encumbered
by systematic errors due to the number of simplifying assumptions inherent in equation (2-52).
However, this does not affect our main conclusion regarding the significance of the oxygen
absorption, since more refined calculations give a pattern similar to that observed above (see
Fig. 2.39).
Retained oxygen
Absorbed
oxygen
Retained
oxygen
past.47"51 Most of these investigators have interpreted their results as a high-temperature equi-
librium between the slag and the weld metal, but a verified quantitative understanding of the
transfer of elements during welding is lacking. This situation arises mainly from the lack of
adequate thermodynamic data for the complex slag/metal systems involved.
0.1 wt%Cat10bar.
Temperature,°C
Fig. 2.40. The break even equilibrium temperature for silicon control of oxygen level at 0.1 wt% C and
0.3 wt% Si. Data from Elliott et al.23
Table 2.9 Contents of ferrosilicon and iron powder in the electrode coating of experimental
consumables used in the hyperbaric welding experiments.
Electrode R
A
B
C
Carbon content, wt%
Fig. 2.41. Carbon absorption in hyperbaric SMA welding. Data from Grong et al.51
is at least qualitatively correct would expect a correlation between the weld metal carbon
content and the concentrations of oxygen, silicon and manganese, both under atmospheric and
hyperbaric welding conditions. The main effect of pressure on weld metal chemistry is thus to
suppress the carbon-oxygen reaction in the weld pool at the expense of intensified oxidation
losses of silicon and manganese, as indicated by the thermodynamic data in Fig. 2.40.
This interpretation is further supported by the results from the oxygen determination con-
tained in Fig. 2.42. Although there is considerable scatter in the data in this figure, it is evident
that the recorded enhancement of the weld metal carbon content at high ferrosilicon levels in
the electrode coating is accompanied by a corresponding reduction in the oxygen concentra-
tions.
Assuming the activity ratio (tf Mn0 ) / (aSio2 )in precipitated slag to be constant and inde-
pendent of pressure, equation (2-55) may be rewritten as:
(2-56)
In Fig. 2.45 the weld metal manganese content has been plotted versus the square root of
the silicon content by inserting data from Figs. 2.43 and 2.44. As it appears from Fig. 2.45, the
Oxygen content, ppm
Electrode R
A
B
C
Fig. 2.42. Oxygen absorption in hyperbaric SMA welding. Data from Grong et al.51
Electrode R
A
B
C
Silicon content, wt%
Electrode R
A
B
C
Manganese content, wt%
Fig. 2.44. Manganese oxidation in hyperbaric SMA welding. Data from Grong et al.51
experimental data cluster around a straight line passing through the origin, which confirms
that the silicon and manganese concentrations are balanced by a reaction according to equation
(2-55).
Electrode R
A
B
C
Manganese content, wt%
1/2
[Silicon content, wt%]
Fig. 2.45. Correlation between weld metal manganese and silicon contents. Data from Grong et a/.51
Here K5 is the equilibrium constant for reaction (2-53) (equal to about 2.0 X 10"3 at 16000C
and 2.6 X 10~3 at 20000C),23 Nco is the mole fraction of CO in the reaction product (equal to
the partial pressure of CO at 1 bar), andptotis the total ambient pressure.
During the initial stage of cooling in the weld pool, the oxygen content in an assumed
equilibrium with carbon would be expected to be higher than the analytical values. This situa-
tion applies in particular to welds made under hyperbaric conditions, where significant quan-
tities of oxygen clearly are removed from the weld pool in the form of oxide inclusions after
the completion of the carbon oxidation.
The concentration of dissolved oxygen at the break even temperature for silicon control of
the oxygen level can be estimated from the measured concentration displacements of oxygen,
silicon and manganese in the weld deposit with increasing pressures, relative to 1 bar (desig-
nated A[%(9], A[%Si] and A[%Mrc], respectively). If the total amount of oxygen which reacts
with silicon and manganese at 1 bar, as a first approximation, is taken equal to the analytical
weld metal oxygen content, the following balance is obtained:
(2-58)
Next Page
Here [%0]eq. is the oxygen concentration in an assumed equilibrium with carbon at a given
pressure, and [%O]anaL is the analytical weld metal oxygen content. For this correlation, minor
vaporisation losses of manganese as well as possible reactions between oxygen and liquid iron
have been neglected.
In Fig. 2.46 the product m = [%C] [%O] is plotted vs the total ambient pressure. Calcula-
tions of m have been done both on the basis of [%O]anaL and [%O]eq. It can be seen from Fig.
2.46 that the former set of data (i.e. open symbols in the graph) cannot be represented by a
straight line passing through the origin, which should apply to a true equilibrium reaction.
However, when proper corrections are made for the amount of oxygen removed from the weld
pool after the completion of the carbon oxidation, such a correlation may be obtained as shown
by the solid line in Fig. 2.46. No clear effect of the electrode deoxidation capacity (i.e. ferrosilicon
content) on the product m = [%C] [%O] can be observed within the precision of measure-
ments. This result is to be expected if the weld metal carbon content is controlled by a local
equilibrium with oxygen established at elevated temperatures in the weld pool. Also, inspec-
tion of the slope of the curve (i.e. heavy solid line in Fig. 2.46) indicates that the product
K5Nco is about 1.14 X 10~3 under the prevailing circumstances. If a reasonable average value
for the equilibrium constant K5 of 2.3 X 10~3 is assumed within the specific temperature range
of the reaction, we get:
Nco~0.5 (2-59)
The above calculations suggest that the controlling partial pressure of CO in the reaction
product is significantly lower than the ambient pressure under hyperbaric welding conditions.
This probably arises from an extensive infiltration of helium in the nucleating bubbles at the
slag/metal interface which, thermodynamically, will enhance the deoxidation capacity of car-
bon according to Le Chatelier's Principle. The conditions existing in hyperbaric SMA welding
thus appear to be similar to OBM/Q-BOP steelmaking, where simultaneous injection of oxy-
gen and inert gas from the bottom of the convenor during the decarburisation stage results in a
steel carbon content which is typically below the value calculated for equilibrium between
oxygen and carbon at 1 atm partial pressure of CO.52
During cooling, the metal concentrations established at high temperatures due to dissolution of
oxygen tend to readjust by precipitation of new phases. Accordingly, a supersaturation with
respect to the various deoxidation reactants initially increases and thus provides the driving
force for nucleation of oxides. Subsequently, the deoxidation reactions will proceed rapidly
through growth of nuclei above a critical size. Equilibrium conditions will finally establish the
limits for the degree of deoxidation that can be achieved.
In spite of the fact that large amounts of oxygen are removed from the weld pool during the
deoxidation stage, the analytical weld metal oxygen content exceeds by far the value predicted
from chemical thermodynamics, assuming that equilibrium conditions are maintained down to
the solidification temperature (see data in Table 2.5). This situation cannot be ascribed to a
large deviation from chemical equilibrium, but is mainly a result of an incomplete phase sepa-
ration. Consequently, due consideration must be given to the kinetics. The three basic consecu-
tive steps in steel deoxidation are shown in Fig. 2.47.
Previous Page
Here [%0]eq. is the oxygen concentration in an assumed equilibrium with carbon at a given
pressure, and [%O]anaL is the analytical weld metal oxygen content. For this correlation, minor
vaporisation losses of manganese as well as possible reactions between oxygen and liquid iron
have been neglected.
In Fig. 2.46 the product m = [%C] [%O] is plotted vs the total ambient pressure. Calcula-
tions of m have been done both on the basis of [%O]anaL and [%O]eq. It can be seen from Fig.
2.46 that the former set of data (i.e. open symbols in the graph) cannot be represented by a
straight line passing through the origin, which should apply to a true equilibrium reaction.
However, when proper corrections are made for the amount of oxygen removed from the weld
pool after the completion of the carbon oxidation, such a correlation may be obtained as shown
by the solid line in Fig. 2.46. No clear effect of the electrode deoxidation capacity (i.e. ferrosilicon
content) on the product m = [%C] [%O] can be observed within the precision of measure-
ments. This result is to be expected if the weld metal carbon content is controlled by a local
equilibrium with oxygen established at elevated temperatures in the weld pool. Also, inspec-
tion of the slope of the curve (i.e. heavy solid line in Fig. 2.46) indicates that the product
K5Nco is about 1.14 X 10~3 under the prevailing circumstances. If a reasonable average value
for the equilibrium constant K5 of 2.3 X 10~3 is assumed within the specific temperature range
of the reaction, we get:
Nco~0.5 (2-59)
The above calculations suggest that the controlling partial pressure of CO in the reaction
product is significantly lower than the ambient pressure under hyperbaric welding conditions.
This probably arises from an extensive infiltration of helium in the nucleating bubbles at the
slag/metal interface which, thermodynamically, will enhance the deoxidation capacity of car-
bon according to Le Chatelier's Principle. The conditions existing in hyperbaric SMA welding
thus appear to be similar to OBM/Q-BOP steelmaking, where simultaneous injection of oxy-
gen and inert gas from the bottom of the convenor during the decarburisation stage results in a
steel carbon content which is typically below the value calculated for equilibrium between
oxygen and carbon at 1 atm partial pressure of CO.52
During cooling, the metal concentrations established at high temperatures due to dissolution of
oxygen tend to readjust by precipitation of new phases. Accordingly, a supersaturation with
respect to the various deoxidation reactants initially increases and thus provides the driving
force for nucleation of oxides. Subsequently, the deoxidation reactions will proceed rapidly
through growth of nuclei above a critical size. Equilibrium conditions will finally establish the
limits for the degree of deoxidation that can be achieved.
In spite of the fact that large amounts of oxygen are removed from the weld pool during the
deoxidation stage, the analytical weld metal oxygen content exceeds by far the value predicted
from chemical thermodynamics, assuming that equilibrium conditions are maintained down to
the solidification temperature (see data in Table 2.5). This situation cannot be ascribed to a
large deviation from chemical equilibrium, but is mainly a result of an incomplete phase sepa-
ration. Consequently, due consideration must be given to the kinetics. The three basic consecu-
tive steps in steel deoxidation are shown in Fig. 2.47.
Electrode R
A
B
C
[%C][%O]x102
Fig. 2.46. The product [%C][%0] in hyperbaric SMA welding. Solid symbols: calculations based on
[%O]eq.. Open symbols: calculations based on [%O]anai.- Data from Grong et al. 51
Diffusion of reactants in the
Heterogeneous nucleation
Homogeneous nucleation
Surface tension
Separation Stage
Average particle size
Particle coalescence
Buoyancy
Nucleation Stage
Stirring
Growth Stage
Reaction time
Fig. 2.47. The three major consecutive steps in steel deoxidation (schematic).
Although rate phenomena in ladle refining of liquid steel are extensively investigated and
reported in the literature,53"55 only recently attempts have been made to include such effects in
an analysis of deoxidation reactions in arc welding. 1 ' 56 ' 57
(2-60)
where AH° is the standard enthalpy of reaction, G is the oxide-steel interfacial energy (as-
sumed to be constant and independent of temperature), and Vm is the molar volume of the
nucleus.
The derivation of equation (2-60) is shown in Appendix 2.2.
Example (2.9)
Assume that precipitation of manganese silicates in the weld pool occurs according to the
following reaction:
where
AG°(J) = 858620 + 3457.
Based on equation (2-60), calculate the critical temperature interval of subcooling for ho-
mogeneous nucleation OfMnSiO3. Typical physical data for liquid steel and manganese sili-
cate slags are given in Table 2.10.
£
Time
Fig. 2.48. Idealised model for homogeneous nucleation of oxide inclusions in steel weld metals (sche-
matic).
Table 2.10 Physical data for liquid steel and manganese silicate slags at 16000C. Data from Refs. 3 and
53.
Solution
First we estimate the molar volume of the nucleus:
By inserting the appropriate values for Vm, AH°, and a in equation (2-60), we obtain:
from which
If we assume that the supersaturation is released at T2 = 16000C (1873K), the initial tem-
perature of the liquid T1 becomes:
The critical temperature interval of subcooling is thus:
Similar calculations can also be carried out for other types of oxide inclusions, e.g. FeO(I),
SiO2(s), and Al2O3(S). The results of such computations are presented graphically in Fig. 2.49,
using data from Refs 55 and 58. It is evident from these plots that the critical temperature
interval of subcooling depends on the interfacial energy, G. Although data for oxide-steel
interfacial energies are scarce, the following average values are frequently used in the litera-
ture, 5558 i.e. a(FeO-Fe) = 0.3 J m"2; G(SiO2-Fe) = 0.9 J m~2; and G(Al2O3-Fe) - 1.5 J mr2. If
these values are accepted, the results in Fig. 2.49 indicate that the critical temperature interval
of subcooling for homogeneous nucleation of FeO (1), SiO2(s), and Al2O3(S) is of the order of
200 to 30O0C. Considering the fact that the liquid weld metal spans a temperature range of
about 2200 to 15000C,34 it is not surprising to find that nucleation of oxide inclusions occurs
readily in the weld pool during cooling. It should be noted that the quoted data for G are
representative of ladle-refined steel deoxidised at 16000C. At higher metal temperatures, in the
presence of large amounts of dissolved oxygen, the oxide-steel interfacial energies would be
expected to be significantly lower.59 Hence, it is reasonable to assume that the actual tempera-
ture interval of subcooling required for spontaneous oxide precipitation in a weld pool is well
below 2000C.
Critical temperature interval, AT (0C)
Interfacial energy,o(J/m2)
Fig. 2.49. Critical temperature interval of subcooling for homogeneous nucleation of oxide inclusions in
steel weld metals at 16000C as a function of the interfacial energy a.
From deoxidation and ladle refining of liquid steel it is well established that the flotation
rate of the oxides generally depends on their growth rates, since large inclusions separate
much more rapidly than small ones, in agreement with Stokes law.55 Growth of the oxides can
proceed either through diffusion of reactants in the melt to the oxide nuclei or by collision and
coalescence of ascending inclusions, and is therefore influenced by factors such as the number
density of the nuclei, interfacial tensions, and the extent of melt stirring.53"55 The last factor is
of particular importance in welding, because the stirring action will increase the possibilities
for collision and coalescence of inclusions and, depending on the direction of flow, can give
rise to circulation of inclusion-laden metal to the surface. As a result, the separation of small
oxide particles, i.e. microslag, is strongly favoured by the turbulent conditions existing in the
hot part of the weld pool immediately beneath the root of the arc.
Upward welding
Downward welding
Oxygen content, wt%
Horizontal
(2-62)
where g is the gravity constant, dv is the particle diameter, Ap is the difference in densities
between the liquid steel and the inclusions, and |x is the steel viscosity.
Taking the Stokes parameter, gAp/18|ji, equal to 0.6 jjinr1 s"1 for manganese silicate slags in
steel (Table 2.10), equation (2-62) becomes:
(2-63)
Generally, the majority of non-metallic inclusions in steel weld metals are of a diameter
below 2 jxm.56'57 According to equation (2-63), such particles have a relative velocity less than
2.4 |xm s"1. This implies that the buoyancy effect alone is far too insignificant to promote
flotation of the inclusions out of the weld pool before solidification, when it is recognised that
the average fluid flow velocity in the weld pool is four to five orders of magnitude higher (to
be discussed later).
Stokes law is based on the assumption that the inclusions are completely wetted by the
liquid steel, i.e. there is no slip at the oxide/metal interface.55 Normally, interfacial tension
effects promote slip at the particle/metal interface, which, in turn, enhances the flotation rate
of the ascending oxides (often referred to as the Plockinger effect).60 The concept of a wetting
angle has been used in this context. But, the slip phenomenon is probably perceived better in
terms of a secondary flow in the interfacial region between the liquid metal and the solid
precipitate, which is produced by gradients in interfacial tension. In the rare case of no wetting
(0 = 180°), it has been shown that the average terminal velocity of the inclusions is approxi-
mately 50% higher than that given by equation (2-62).55 For silica slags, the wetting angle (in
the presence of air) is close to 115°,54 indicating that the correction in the particle terminal
velocity due to the slip at the oxide/metal interface in less than about 30%. Consequently,
interfacial tension effects between the slag and the steel do not significantly affect the flotation
rate of the particles, and, therefore, can be ignored.
(2-64)
where Cd is the drag coefficient, p is the fluid density, u is the bulk velocity of the fluid
relative to the particle, and dv is the particle diameter. The drag coefficient is, in turn, a func-
tion of the particle Reynolds number, A^,:3
(2-65)
where |x is the fluid viscosity. Typical physical-property data for liquid steel are given in Table
2.10. By inserting the values for p and JJL from Table 2.10, then using the relative velocity
calculated in equation (2-63) (2.4 |xm s"1) as one extreme and the fluid flow velocity (0.4 m s"1) as
the other (i.e. assuming a stationary particle), one can demonstrate that the particle Reynolds
numbers obtained never exceed 1.3 for most weld metal inclusions. The limit calculated above
for the Reynolds number is within the so-called creeping flow region where Stokes law is
indeed valid. Under such conditions, the ratio between the particle drag force Fd and the corre-
sponding gravity force Fg is given by the expression:3
(2-66)
Note that equation (2-66) is the basis for obtaining Stokes law (equation (2-62)).
For steady motion of the particle, Fd and Fg are, of course, equal. However, the significance
of Fd can also be interpreted in a transient sense, by considering the limiting case where the
particle is stationary and its instantaneous motion is governed by the forces acting on it. In this
case, the ratio of Fd (max) acting on the particle relative to the net gravity force (given by
Electrode
Pulsating cavity
Base plate
Fig. 2.51. Typical flow pattern in SA weld pools. For clarity, the arc, slag and flux have been omitted.
The sketch is based on the ideas of Lancaster.32
equation (2-66)) can determine the dominant force responsible for the particle's trajectory. In
Fig. 2.52 the above ratio has been calculated and plotted against the particle diameter, dv, for
values of u' equal to 0.025 and 0.4 m s"1, which represent the typical velocities reported for
SA steel weld pools. 63 ' 64 It can be seen from the figure that the drag force is always several
orders of magnitude greater than the gravity force for particles within the typical size range of
weld metal inclusions (i.e. less than 2 |xm in diameter). This result is to be expected, since the
relative velocity, based on Stokes law and calculated in equation (2-63), is negligible com-
pared with the liquid velocities in the weld pool (2.4 \xm s"1 vs 0.02 to 0.4 m s"1). This calcu-
lation thus supports our previous conclusion that the separation of the oxide inclusions is con-
trolled solely by the fluid flow behaviour in the weld pool. The fact that the phase separation
proceeds under strongly turbulent conditions is also evidenced by the large number of iron
droplets being mechanically dispersed in the top bead slag of GMA steel welds, as shown by
the optical micrographs in Fig. 2.53.
In the case of GMA welding, the non-metallic inclusions that are brought to the upper
surface behind the arc coalesce rapidly to form large slag clusters that float on the top of the
bead. Generally, re-entrapment of the slag does not occur owing to the decrease in the total
surface free energy of the system, which is caused by the emergence of the inclusions from the
weld metal.54 Consequently, the slag will remain floating on the top of the bead even when
welding is performed in the overhead position, as shown previously in Fig. 2.50.
Fig. 2.53. Optical micrographs showing characteristic "comet's tails" of trapped iron droplets (light
areas) in collected top bead slags of GMA steel welds; (a) Ar + 5vol%O2, (b) Ar + 5vol%CO2.
Electrode
(a)
Base plate
(b) S e c t i o n A-A
Cold part of weld pool: Hot part of weld pool: Final weld metal,
Deoxidation/phase oxygen concentration
Temperature
Deoxidation/incomplete separation
phase separation
Oxygen content
Fig. 2.54. Schematic diagrams showing the sequence of reactions occurring during weld metal deoxidation;
(a) Longitudinal section of weld pool, (b) Cross section of weld pool along A-A. The diagrams are based
on the ideas of Grong and Christensen.1
(i) The hot part of the weld pool, characterised by simultaneous oxidation and deoxidation
of the metal, where the separation of microslag takes place continuously as a result of highly
turbulent flow conditions.
(ii) The cold part of the weld pool, where precipitated slag will largely remain in the metal
as finely dispersed particles as a result of inadequate melt stirring.
Under such conditions equation (2-23) predicts a direct correlation between absorbed and
retained (analytical) oxygen in the weld metal, i.e.:
(2-67)
in agreement with experimental observations (see plots in Figs. 2.33, 2.38, and 2.39). Typi-
cally, the proportionality constant C4 varies between 0.1 to 0.2, which corresponds to a range
in the t/to ratio from 2.3 to 1.6. This shows that the boundary between 'hot' and 'cold' parts of
the weld pool is not well defined, but depends on the welding system under consideration as
well as on the operational conditions applied.
To allow for the decrease in the silica activity with increasing manganese-to-silicon ratios,
it is essential to establish a correlation that links the activity of silica to the concentrations of
the deoxidation elements in the weld metal. A semi-empirical corr61ation of this kind has been
presented by Walsh and Ramachandran,65 derived from a re-analysis of activity data for silica
in the Fe-Mn-Si-O system previously published by Hilty and Crafts.66 Within the tempera-
ture range from 1550 to 16500C, they showed that the silica activity in the deoxidation product
can be approximately expressed as:
(2-70)
where K1 represents the manganese-to-silicon ratio at which the activity of silica becomes
unity for a given temperature. A check of this equation against more recent data for silica
activities in MnO-SiO2 slags reported by Turkdogan67 supports the findings of Walsh and
Ramachandran65 that the activity of silica is approximately given by equation (2-70) for a wide
range in the steel manganese-to-silicon ratio (i.e. from 0.1 to 50). By combining equations (2-
69) and (2-70), it is possible to obtain an expression for the equilibrium oxygen content, solely
in terms of the silicon and manganese concentrations:
(2-71)
(2-72)
when pure SiO2 is used as the standard state for the silica activity.
For a rough estimate of the temperature dependence of equation (2-70), the results of
Turkdogan55 can be used. It should be noted that Walsh and Ramachandran65 did calculate the
temperature dependence of K1 within the range from 1550 to 16500C. However, because equa-
tion (2-70) is empirical, the function cannot be extrapolated beyond these temperature limits.
The data quoted in Ref. 55 are derived directly from the Si-Mn reaction (equation (2-55)) and
activity data for MnO at silica saturation. On introduction of the equilibrium constant for
equation (2-55), we obtain:
(2-73)
By using data from Ref.55, the initial [%Mn]2/[%Si] ratio for precipitation of silica satu-
rated slags (equal to K9(aMnO)2) at 1500,1550,1600 and 16500C has been recorded and replotted
against temperature, as shown in Fig. 2.55. The figure shows that the critical [%Mn]2/[%Si]
ratio for precipitation of silica saturated slags is temperature dependent and decreases from
about 5 at 15000C to below 1.5 at 16500C. On the basis of a crude extrapolation of the data to
higher metal temperatures, it can, however, be argued that the ratio would approach a constant
value of approximately 0.75 at temperatures beyond about 17500C (indicated by the broken
horizontal line in Fig. 2.55). The above observation reflects the fact that the Si-Mn-O reaction
equilibrium (equation (2-55)) is not very sensitive to a change in temperature (i.e. the enthalpy
for the reaction is small). Over the composition range normally applicable to Si-Mn deoxidation
of steel weld metals the observed threshold for the critical [%Mn]2/[%Si] ratio for precipita-
tion of silica-saturated slags in Fig. 2.55 corresponds to a manganese-to-silicon ratio closely
equal to unity. Consequently, at temperatures higher than about 17500C, K1 can, as a first
approximation, be taken constant and independent of temperature (i.e. K1- Y). The tempera-
ture dependence of equation (2-71) is thus simply (1/^ 6 ) 05 or:
(2-74)
Temperature, 0C
Fig. 2.55. The critical [%Mn]2/ [%Si] ratio for precipitation of silica-saturated slags as a function of
temperature. Data from Turkdogan.55
Inclusions commonly found in steel weldments will either be exogenous or indigenous, de-
pendent on their origin. The first type arises from entrapment of welding slags and surface
scale, while indigenous inclusions are formed within the system as a result of deoxidation
reactions (oxides) or precipitation reactions (nitrides, sulphides). The latter group is almost
always seen to be heterogeneous in nature both with respect to chemistry (multiphase parti-
GMA Welding
Oxygen content, wt% SMA Welding
([%SI] [%Mn]f°"25
Fig. 2.56. Examples of pseudo-equilibrium in GMA and SMA welding of C-Mn steels. The solid lines in
the graph represent thermodynamical calculations at indicated temperatures. Data from Refs. 56 and 68.
cles), shape (angular or spherical particles), and crystallographic properties as a result of the
complex alloying systems involved69 (see Fig. 2.57). An exception may be C-Mn steel welds,
where the oxide inclusions will be predominately glassy, spherical, manganese silicates.1
A survey of important weld metal inclusion characteristics is given in Table 2.11.
*Vy: volume fraction; dv: arithmetic mean (3-D) particle diameter (jxm); Nv: number of particles per unit volume
(mm"3); Sv\ total particle surface area per unit volume (mm2 per mm3).
Fig. 2.57. Digital STEM brightfield image and Si, Al, S, Mn and Ti X-ray images of a multiphase weld
metal inclusion. After Kluken and Grong.57
Solution
First we calculate the total weight of retained MnOSiO2 per 100 g weld metal:
Similarly, we can calculate the weight and volume fraction of MnS in the weld metal:
and
In practice, the stoichiometric conversion factors for oxygen and sulphur are virtually con-
stant for a wide spectrum of inclusions70 and hence, they can be regarded as independent of
composition. Taking the solubility of sulphur in solid steel equal to 0.003 wt%, the following
relationship is obtained for steel weld metals:57'70
(2-75)
The validity of equation (2-75) has been confirmed experimentally by comparison with
microscopic assessment methods.5771
In steel weld metals the majority of the inclusions will be in the submicroscopic range
owing to the limited time available for growth of the oxides. From the histogram in Fig. 2.58 it
is seen that particles with diameters between 0.3 to 0.8 |xm contribute to nearly 50 percent of
the total inclusion volume fraction. This trend is not significantly changed by additions of
strong deoxidisers, such as aluminium and titanium, or by a moderate increase/decrease in the
heat input.57
Fig. 2.58. Percental contribution of different size classes to the total volume fraction of non-metallic
inclusions in a low-alloy steel weld metal. Data from Kluken and Grong.57
(b)
High Al (0.053 wt%)
High Ti (0.053 wt%)
Frequency (%)
Fig. 2.59. Three-dimensional (3-D) size distribution of non-metallic inclusions in two different low-
alloy steel weld metals; (a) Low weld metal aluminium and titanium levels, (b) High weld metal alu-
minium and titanium levels. Data from Kluken and Grong.57
(2-76)
Here do is the initial particle diameter, a is the oxide-steel interfacial energy, Dm is the
element diffusivity, Cm is the element bulk concentration, V'm is the molar volume of the oxide
per mole of the diffusate, and t is the retention time.
Frequency (%)
For welding of thick plates, the time available for growth of particles in the 'cold' part of
the weld pool can be estimated from the Rosenthal equation, i.e. equation (1-45) in Section
1.10.2 (Chapter 1). If the characteristic length of the cooling zone is taken equal to the weld
ripple lag (defined in Fig. 2.63), the retention time t is approximately given by the following
relationship:
Temperature, C°
Time, seconds
Fig. 2.62. Measured temperature level in the trailing edge of the weld pool during GMA welding. Data
from Kluken and Grong.57
Retention time
Weld ripple lag
Welding speed
Max. width
Heat source
Fusion boundary
Fig. 2.63. Definition of weld ripple lag xm, and retention time t.
(2-77)
where J\ is the arc efficiency (equal to about 0.95 for SAW and 0.80 for GMAW /SMAW), and
E is the gross heat input (kJ mm"1). Note that equation (2-77) assumes constant values for the
steel thermal diffusivity and volume heat capacity (5 mm2 s"1 and 0.0063 J mm"3 0C"1, respec-
tively), and no preheating (T0* = 200C).
In Fig. 2.64 the Ostwald ripening theory has been tested against relevant literature data,
which may be considered representative of the 3-D particle size distribution. Although there is
some scatter in the data, the observed inclusion growth rates fall within the range calculated
for oxygen diffusion-controlled coarsening of SiO2 and Al2O3 at 15500C, using the Wagner
equation. In these calculations, a reasonable average value for the bulk diffusivity of oxygen
has been assumed (i.e. 10~2 mm2 s"1).55 If the effective inclusion growth rate constant for low-
alloy steel weld metals is taken equal to the slope of the curve in Fig. 2.64, the following
relationship is achieved:
GMAW
SAW
SMAW
dv,jLim
•1/3
i i sJ/3
Fig. 2.64. Relation between arithmetic mean (3-D) inclusion diameter dv and retention time t for different
arc welding processes. Data compiled by Kluken and Grong.57
1
'Hot' part of CoId1 part of
weld pool weld pool
Incomplete phase;
Phase separation
Deoxidation
Deoxidation
separation
Growth of
inclusions
Oxidation
Gas/metal- Homogeneous
slag/metal nucleation Ostwald ripening
reactions
Time
Fig. 2.65. Proposed deoxidation model for steel weld metals (schematic). The diagram is based on the
ideas of Kluken and Grong.57
(2-78)
By substituting equation (2-77) into equation (2-78), we obtain a direct correlation between
the arithmetic mean 3-D inclusion diameter dv and the net heat input T\E:
(2-79)
Equation (2-79) predicts that dv is a simple cube root function of E, in agreement with the
experimental data in Fig. 2.61.
It should be noted that the measured shape of the particle distributions (see Figs. 2.59 and
2.60) deviates somewhat from that required by the Wagner equation, which assumes a quasi-
stationary distribution curve, and that the maximum stable particle diameter is about 1.5 times
the mean diameter of the system.73 Although the origin of this discrepancy remains to be
resolved, this suggests that particle clustering is also a significant process in steel weld metals
as it is in other metallurgical systems. In fact, such effects would be expected to be most
pronounced at high aluminium levels because of a large interfacial energy between Al2O3 and
liquid steel, in agreement with experimental observations.57'69
Example (2.11)
Consider SA welding with a basic flux on a thick plate of low-alloy steel under the following
conditions:
Previous experience has shown that this steel/flux combination gives a weld metal oxygen
and sulphur content of 0.035 and 0.008 weight percent, respectively. Based on the stereometric
relationships given below, calculate the total number of particles per unit volume Nv, the total
number of particles per unit area Na, the total particle surface area per unit volume Sv, and the
mean particle centre to centre volume spacing Xv in the weld deposit:74'75
(2-80)
(2-81)
(2-82)
(2-83)
Solution
First we calculate the total volume fraction of oxide and sulphide inclusions from equation
(2-75):
Vv = 10-2 [5.0(0.035) + 5.4(0.008 -0.003)] = 2.0 X 10~3
The arithmetic mean (3-D) inclusion diameter can then be evaluated from equation (2-79):
This gives:
28 wt%
20 wt%
Frequency (%)
number of counts recorded for copper in discrete particles is not significantly higher than the
corresponding matrix value, which shows that the copper content of the inclusions is low.
Since these measurements were carried out on mechanically polished specimens and not on
carbon extraction replicas, as done in the EDX analysis, the indications are that the higher
inclusion copper level observed in the latter case mainly results from surface copper contami-
nation inherent from the extraction process. In contrast, the WDX analysis of sulphur revealed
evidence of sulphur enrichment in most of the particles. Considering the fact that these parti-
cles also contained significant amounts of manganese, it is reasonable to assume that most of
the sulphur is present in the form of MnS (possibly with some copper in solid solution).76
(2-84)
The activity coefficients for MnO and FeO in the ternary system SiO 2 -MnO-FeO can be
computed from the equations presented by Sommerville et al.:S2
(2-85)
and
(2-86)
(2-87)
and
(2-88)
TiN (secondary reaction product)
Fig. 2.67. Schematic diagram showing the presence of primary and secondary phases in weld metal
inclusions.
Under such conditions the activity ratio aMnO/aFeO in the slag is given by:
(2-89)
The activity ratio aMnO/aFeO can also be estimated from the equilibrium constant for the Fe-
Mn reaction at 15500C,23 i.e.:
(2-90)
The corresponding mole fractions of MnO and FeO in the slag phase are then obtained by
combining equations (2-89) and (2-90):
(2-91)
and
(2-92)
Equations (2-91) and (2-92) provide a basis for calculating the chemical composition of the
inclusions under different deoxidation conditions. A requirement is, however, that the weld
metal Si to Mn ratio is sufficiently high to promote precipitation of silica-saturated slags at
15500C.
and
This gives the following chemical composition of the inclusions (in wt%):
The above calculations should be compared with the composition data presented in Fig.
2.68. It is evident from this plot that the agreement between predictions and experiments is
reasonably good both at high and low weld metal manganese levels. This justifies the
simplifications made in deriving equations (2-91) and (2-92).
Aluminium
The average aluminium content of the inclusions, [%Al]inch can be estimated from the meas-
ured difference between total and acid soluble aluminium in the weld metal, (k%Al)we[d. This
difference is, in turn, equal to the total mass of aluminium in the inclusions:
(2-93)
where mind and Vv (cal) are the total mass and volume fraction of non-metallic inclusions in
the weld deposit, respectively.
*The numerical constants in the constitutive equations given below could alternatively be expressed in terms of
atomic weights etc. to bring out more clearly their physical significance (e.g. see the treatment of Bhadeshia and
Svensson83).
Ar - O2 gas mixtures
Ar - CO2 gas mixtures
Mole fraction
Fig. 2.68. Comparison between measured and predicted microslag composition in GMA welding of
C-Mn steels. Solid lines represent theoretical calculations based on equations (2-91) and (2-92). Data
from Grong and Christensen.1
Mn
Si
Ti
Al
(2-94)
However, since data for acid soluble aluminium (and titanium), in practice, may contain
large inherent errors, the following restriction applies:
(2-95)
Titanium
Similarly, the average titanium content of the inclusions, [%Ti]incl, can be estimated from the
measured difference between total and acid soluble titanium, (A%Ti)weld. However, since TiN
dissolves readily in strong acid, it is necessary to include an empirical correction for the amount
of titanium nitride which simultaneously forms at the surface of the inclusions during solidifi-
cation. This can be done on the basis of published data for the solubility product of TiN in
liquid steel.84 If we assume that the nitrogen content of the inclusions is proportional to the
calculated difference between total and dissolved nitrogen at the melting point (15200C), the
following relationship is obtained:
(2-96)
where [%N]anai is the analytical weld metal nitrogen content, and [%Ti]soL is acid soluble
titanium. Note that the correction term for TiN, in practice, neither can be negative nor exceed
[%Ti]soh
In this case, the maximum amount of titanium which can be bound as Ti2O3 is determined
by the overall oxygen balance:
(2-97)
Sulphur
If the solubility of sulphur in solid steel is taken equal to 0.003 wt%,70 the average sulphur
content of the inclusions, [%S]ind is given by:
(2-98)
MnO Al2O3
wt% AI2O3
Fig. 2.70. Measured inclusion compositions in low-alloy steel weld metals. Data from Saggese et alP
(2-99)
Considering manganese, proper adjustments should also be made for the amount of MnS
formed at the surface of the inclusions during solidification. Hence, the average manganese
content of the inclusions, [%Mn]inci, is given by the sum of the oxygen and the sulphur contri-
butions:
(2-100)
Titanium
Manganese
Measured composition, wt%
Silicon
Implications of model
It can be inferred from equations (2-94) and (2-95) that the chemical composition of the in-
clusion oxide core is directly related to the aluminium to oxygen ratio in the weld metal.
Referring to Fig. 2.72, the fraction of MnOAl2O3 and 7-Al2O3 in the inclusions is seen to in-
crease steadily with increasing (A%Al)wdJ[%O]anal ratios until the stoichiometric compo-
sition for precipitation of aluminium oxide is reached at 1.13. At higher ratios, the deoxida-
tion product will be pure Al2O3, since aluminium is present in an over-stoichiometric amount
with respect to oxygen. When titanium is added to the weld metal, titanium oxide (in the form
of Ti2O3) may also enter the reaction product. At the same time both TiN and a-MnS form
epitaxially on the surface of the inclusions during solidification. Consequently, in Al-Ti-Si-Mn
deoxidised steel weld metals the total number of constituent phases within the inclusions may
approach six. The kinetics of inclusion formation are further discussed in Ref. 87.
SiO2
MnO
Fig. 2.72. Coexisting phases in inclusions at different weld metal aluminium-to-oxygen ratios. Shaded
region indicates the approximate composition range for the oxide phase. The diagram is constructed on
the basis of the model of Kluken and Grong57 and relevant literature data.
Example (2.13)
Consider SA welding of low-alloy steel with two different basic fluxes. Data for the weld
metal chemical compositions are given in Table 2.12. Based on Fig. 2.72, estimate the total
number of constituent phases in the inclusions in each case.
Solution
It follows from Fig. 2.72 that the chemical composition of the deoxidation product is deter-
mined by the aluminium to oxygen ratio in the weld metal. For weld A, we have:
Since this ratio is higher than the stoichiometric factor of 1.13, all oxygen is probably tied up
as aluminium oxide. In addition, weld A contains small amounts of titanium and sulphur,
which may give rise to precipitation of TiN and MnS at the surface of the inclusions during
solidification. Hence, the three major constituent phases in the weld metal inclusions are
7-Al2O3, TiN, and ot-MnS.
Table 2.12 Chemical composition of SA steel weld metals considered in Example 2.13.
*Data for acid soluble Al and Ti are given by the values in brackets.
In the case of weld B the situation is much more complex due to a higher content of oxygen
and titanium. From Table 2.12, we have:
and
These calculations show that Al and Ti are not present in sufficient amounts to tie-up all oxy-
gen. Under such conditions Fig. 2.72 predicts that the total number of constituent phases in the
inclusions is six, i.e.: SiO2, MnOAl2O3, 7-Al2O3, Ti2O3, a-MnS and TiN.
References
Appendix 2.2
Derivation of equation (2-60)
(A2-1)
where J0 is a constant (with the unit nuclei per m3 and s) and AG* is the energy barrier for
nucleation.
By rearranging equation (A2-1) and inserting reasonable values for J and J0 for the specific
case of homogeneous nucleation of oxide inclusions in liquid steel,55 we obtain:
(A2-2)
(A2-3)
where NA is the Avogadro constant, o is the interfacial energy between the nucleus and the
liquid (in J nr 2 ) and AGV is the driving force for the precipitation reaction (in J nr 3 ).
The parameter AGV can be expressed as:
(A2-4)
where AH° and AS° are the standard entalpy and entropy of the precipitation reaction, respec-
tively and Vm is the molar volume of the nucleus (in m3 mol"1).
It is evident from Fig. 2.48 that AGV = 0 when T=Tu which gives \S° = A//7 T\. Hence,
equation (A2-4) may be rewritten as:
(A2-5)
By combining equations (A2-2), (A2-3) and (A2-5), we obtain the following relationship
between T\ and T^-
(A2- 6)
3
Solidification Behaviour of Fusion Welds
3.1 Introduction
Inherent to the welding process is the formation of a pool of molten metal directly below the
heat source. The shape of this molten pool is influenced by the flow of both heat and metal,
with melting occurring ahead of the heat source and solidification behind it. The heat input
determines the volume of molten metal and, hence, dilution and weld metal composition, as
well as the thermal conditions under which solidification takes place. Also important to solidi-
fication is the crystal growth rate, which is geometrically related to weld travel speed and weld
pool shape. Hence, weld pool shape, weld metal composition, cooling rate, and growth rate
are all factors interrelated to heat input which will affect the solidification microstructure.
Some important points regarding interpretation of weld metal microstructure in terms of these
four factors will be discussed below.
Since the properties and integrity of the weld metal depend on the solidification microstruc-
ture, a verified quantitative understanding of the weld pool solidification behaviour is essen-
tial. At present, our knowledge of the chemical and physical reactions occurring during solidi-
fication of fusion welds is limited. This situation arises mainly from a complex sequence of
reactions caused by the interplay between a number of variables which cannot readily be ac-
counted for in a mathematical simulation of the process. Nevertheless, the present treatment
will show that it is possible to rationalise the development of the weld metal solidification
microstructure with models based on well established concepts from casting and homogenising
treatment of metals and alloys.
The symbols and units used throughout this chapter are defined in Appendix 3.1.
During ingot casting, three different structural zones can generally be observed, as shown
schematically in Fig. 3.1. The chill zone is produced by heterogeneous nucleation in the re-
gion adjacent to the mould wall as a result of the pertinent thermal undercooling. These grains
rapidly become dendritic, and dendrites having their <100> direction (preferred easy growth
direction for cubic crystals) parallel to the maximum temperature gradient in the melt will
soon outgrow those grains that do not have this favourable orientation. Competitive growth
occurring during the initial stage of the solidification process leads to an alignment of the
crystals in the heat flow direction and eventually to the formation of a columnar zone. 12 Fi-
nally, an equiaxed zone may develop in the centre of the casting, mainly as a result of growth
of detached dendrite arms within the remaining, slightly undercooled liquid.
A similar situation also exists in welding, as indicated in Fig. 3.2 However, in this case the
chill zone is absent, since the partly melted base metal grains at the fusion boundary act as seed
crystals for the growing columnar grains.3 In addition, the growth direction of the columnar
Shrinkage pipe
Chill zone
Columnar zone
Equiaxed zone
Mould
Fig. 3.1. Transverse section of an ingot showing the chill zone, the columnar zone and the equiaxed zone
(schematic).
grains will change continuously from the fusion line towards the centre of the weld due to a
corresponding shift in the direction of the maximum temperature gradient in the weld pool.
This change in orientation may result in a curvature of the columnar grains (Fig. 3.2(a)). Alter-
natively, new grains can nucleate and grow in a columnar manner, producing a so-called 'stray'
structure as shown schematically in Fig. 3.2(b). Finally, if the conditions for nucleation of new
grains are favourable, an equiaxed zone will form near the weld centreline similar to that
observed in ingots or castings (see Fig. 3.2(b)).
Although the process of weld pool solidification is frequently compared with that of an
ingot in 'miniature', a number of basic differences, already mentioned, exist which strongly
influence the microstructure and properties of the weld metal. Of particular importance is also
the disparity in cooling rate between a fusion weld and an ingot (see Fig. 3.3). For conven-
tional processes such as shielded metal arc (SMA), gas metal arc (GMA), submerged arc (SA)
or gas tungsten arc (GTA) welding the cooling rate may vary from 10 to 103 0 C s"1, while for
modern high energy beam processes such as electron beam (EB) and laser welding the cooling
rate is typically of the order of 103 to 106 0 C s"1.4 Consequently, to appreciate fully the impli-
cations of these differences in general solidification behaviour between a weld pool and an
ingot, it is necessary to consider in detail the sequence of events taking place in the solidifying
weld metal beginning with the initiation of crystal growth at the fusion boundary.
It is well established that initial solidification during welding takes place epitaxially, where the
partly melted base metal grains at the fusion boundary act as seed crystals for the columnar
grains. This process is illustrated schematically in Fig. 3.4.
(a)
Welding direction
(b)
Columnar zone
Columnar zone •
Fig. 3.2. Examples of structural zones in fusion welds (schematic); (a) Curved columnar grains,
(b) Stray grain structure.
Cooling rate, °C/s
Process
Fig. 3.3. Disparity in cooling conditions between casting, welding and rapid solidification.
boundary
Fusion
Fig. 3.4. Schematic illustration showing epitaxial growth of columnar grains from partly melted base
metal grains at fusion boundary.
Liquid (L)
Substrate (S)
(3-2)
where VE is the volume of the solid embryo, AGV is the free energy change associated with the
embryo formation, AEL and AES are the areas of the embryo-liquid and embryo-substrate inter-
faces, respectively, and/(P) is the so-called shape factor, defined as:
(3-3)
The critical radius of the stable nucleus, r / , is found by differentiating equation (3-2) with
respect to rs and equating to zero:
(3-4)
By substituting equation (3-4) into equation (3-2), we obtain the following expression for
the energy barrier to heterogeneous nucleation (AG^ r ):
(3-5)
where AHm is the latent heat of melting, Tm is the melting point, and AJT is the undercooling.
It is easy to verify that the first term in equation (3-5) is equal to the energy barrier to
homogeneous nucleation, AG^om. Hence, we may write:
(3-6)
Equation (3-6) shows that AG^ is a simple function of the wetting angle O). Since the
chemical composition and the crystal structure of the two solid phases are usually very similar,
we have:6
Under such conditions equation (3-1) predicts that the wetting angle 3 ~ 0 (cos(3 ~ 1),
which implies that there is a negligible energy barrier to solidification of the weld metal (№}*het
~ 0), i.e. no undercooling of the melt is needed, and solidification occurs uniformly over the
whole grain of the base metal. This is in sharp contrast to conventional casting of metals and
alloys where some undercooling of the melt is always required to overcome the inherent en-
ergy barrier to solidification (see Fig. 3.6).
AG
r
s
Casting
Welding Homogeneous
nucleation
Fig. 3.6. The free energy change associated with heterogeneous nucleation during casting and weld
metal solidification, respectively (schematic). The corresponding free energy change associated with
homogeneous nucleation is indicated by the broken curve in the graph.
Fusion
line
HAZ Weld
Weld metal
metal
Fig. 3.7. Correlation between HAZ prior austenite grain size at the fusion boundary and the correspond-
ing weld metal prior austenite grain size. Data from Grong et al?
2. pass
1. pass
HAZ
Base metal
Fig. 3.8. Optical micrograph showing renucleation of columnar grains during multipass GMA welding
of a P-titanium alloy.
3.4 Weld Pool Shape and Columnar Grain Structures
Growth of the columnar grains always proceeds closely to the direction of the maximum ther-
mal gradient in the weld pool, i.e. normal to the fusion boundary. This implies that the colum-
nar grain morphology depends on the weld pool geometry.
(3-7)
where qo is the net arc power, v is the welding speed, a is the thermal diffusivity of the base
plate, and Hm-Ho is the heat content per unit volume at the melting point.
As shown in Fig. 3.9(a), a tear-shaped weld pool is favoured by a high n3 value, which is
characteristic of fast moving high power sources. In contrast, at a low arc power and a low
welding speed the shape of the weld pool becomes more elliptical because of a shift in the
mode of heat flow (see Fig. 3.9(b)). Note, however, that the thermal properties of the base
metal is also of importance in this respect, since the n3 parameter is a function of both a and
Hm-Ho. Consequently, a tear-shaped weld pool is usually observed in weldments of a low
thermal diffusivity (e.g. austenitic stainless steel), whereas an elliptical or spherical weld pool
is more likely to form during aluminium welding owing to the resulting higher thermal diffu-
sivity of the base metal.
In addition to the factors mentioned above, the geometry of the weld pool is also affected by
convectional heat transfer due to the presence of buoyancy, electromagnetic or suface tension
gradient forces. Recently, attempts have been made to include such effects in heat flow mod-
els for welding.8"11 Referring to Fig. 3.10(a) the buoyancy force will promote the formation of
a shallow, wide weld pool because of transport of 'hot' metal to the surface and 'cold' metal to
the bottom of the pool. In the presence of the electromagnetic force the flow pattern is re-
versed, since the latter force will tend to push the liquid metal in the central part of the pool
downward to the root of the weld. This makes the weld pool deeper and more narrow, as
shown in Fig. 3.10(b).
Moreover, it is generally accepted that surface tension gradients can promote circulation of
liquid metal within the weld pool from the region of low surface tension to the region of higher
surface tension.9 In the absence of surface active elements such as oxygen and sulphur, the
surface tension decreases with increasing temperature as illustrated in Fig. 3.10(c), which forces
the metal to flow outwards towards the fusion boundary. This results in the formation of a
relatively wide and shallow weld pool. However, if oxygen or sulphur is present in sufficient
quantities a positive temperature coefficient of the surface tension may develop, which facili-
tates an inward fluid flow pattern and an increased weld penetration (see Fig. 3.10(d)). The
important influence of surface active elements on the resulting bead morphology is well docu-
HAZ isotherms
Fusion boundary
V
Weld
pool
(a)
HAZ isotherms
Fusion boundary
V
Weld
pool
(b)
Fig. 3.9. Theoretical shape of fusion boundary and neighbouring isotherms under different operational
conditions; (a) High n3 values, (b) Low ^-values.
merited for ordinary GTA austenitic stainless steel welds. 1 2 1 3 The indications are that such
effects become even more important under hyperbaric welding conditions. 14
Arc
Weld pool
(a)
Electrode
Arc
Weld pool
(b)
Fig. 3.10. Schematic diagrams illustrating the major fluid flow mechanisms operating in a weld pool;
(a) Buoyancy force (b) Electromagnetic force.
columnar grains as shown in Fig. 3.11(b), since the direction of the maximum temperature
gradient in the melt does not change significantly during the solidification process. The lat-
ter condition is known to promote formation of centre-line cracking because of mechanical
entrapment of inclusions and enrichment of eutectic liquid at the trailing edge of the weld
pool.
Arc
Weld pool
(C)
Surface tension
Temperature
Electrode
Arc
Weld pool
(d)
Fig. 3.10. Schematic diagrams illustrating the major fluid flow mechanisms operating in a weld pool
(continued); (c) Surface tension gradient force (negative gradient); (d) Surface tension gradient force
(positive gradient).
Heat source v
(a)
Heat source v
(b)
Fig. 3.11. Schematic comparison of columnar grain structures obtained under different welding condi-
tions; (a) Elliptical weld pool (low n3 values), (b) Tear-shaped weld pool (high n3 values). Open arrows
indicate the direction of the maximum temperature gradient in the weld pool.
Fusion boundary
Heat source
Crystal
growth direction and the welding direction equal to a, the steady state growth rate, R N , be-
comes:
(3-8)
(b)
Stainless steel (1 mm plate thickness)
Nominal growth rate (RN), mm/s
Equiaxed zone
Fig. 3.13. Measured crystal growth rates in thin sheet electron beam welding; (a) Niobium, (b) Stainless
steel. Data from Senda et al.16
Example (3.1)
Consider electron beam (EB) welding of a lmm thin sheet of austenitic stainless steel under
the following conditions:
Estimate on the basis of the Rosenthal thin plate solution (equation 1-83) the steady state
growth rate of the columnar grains trailing the weld pool.
Solution
The contour of the fusion boundary can be calculated from the Rosenthal thin plate solution
according to the procedure shown in Example (1.10). If we include a correction for the latent
heat of melting, the QbZn3 ratio at the melting point becomes:
Substitution of the above value into equation (1-83) gives the fusion boundary contour
shown in Fig. 3.14. It is evident from Fig. 3.14 that the weld pool is very elongated under the
prevailing circumstances due to a constrained heat flow in the ^-direction. This implies that the
angle a will not change significantly during the solidification process. Taking a as an average,
equal to about 70°, the steady-state crystal growth rate R N becomes:
This value is in reasonable agreement with the measured crystal growth rates in Fig. 3.13(b).
Fusion boundary
+yjmm)
Fig. 3.14. Predicted shape of fusion boundary during electron beam welding of austenitic stainless steel
(Example (3.1)).
(a)
(b)
Fig. 3.25.Examples of faceted cubic crystals; (a) Crystal delimited by {100} planes, (b) Crystal delim-
ited by {111} planes.
Columnar grain
Liquidus temperature
(3-9)
which gives:
(3-10)
Equation (3-10) shows that the local growth rate increases with increasing misalignment of
the crystal with respect to the direction of the maximum temperature gradient in the weld pool.
Since such crystals cannot advance without a corresponding increase in the undercooling ahead
of the solid/liquid interface (see Fig. 3.17), they will soon be outgrowed by other grains which
have a more favourable orientation. Fusion welds of the fee and bcc type will therefore de-
velop a sharp <100> solidification texture in the columnar grain region, similar to that docu-
mented for ingots and castings. The weld metal columnar grains may nevertheless be sepa-
rated by 'high-angle' boundaries, as shown in Fig. 3.1&, due to a possible rotation of the grains
in the plane perpendicular to their <100> length axes.
Example (3.2)
Consider electron beam welding of a 2mm thick single crystal disk of Fe-15Ni-15Cr under
the following conditions:
The orientation of the disk with respect to the beam travel direction is shown in Fig. 3.19.
Calculate on the basis of the minimum velocity (undercooling) criterion the growth rate of
the dendrites trailing the weld pool under steady state welding conditions (assume 2-D
heat flow). Make also schematic drawings of the solidification microstructure in different
sections of the weld. Relevant thermal properties for the Fe-15Ni-15Cr single crystal are
given below:
Solution
Since the base metal is a single crystal, separate columnar grains will not develop. Neverthe-
less, under 2-D heat flow conditions growth of the dendrites can occur both in the [100] and
the [010] (alternatively the [010]) direction. Referring to Fig. 3.20 the growth rate of the
[100] and the [010] deridrites is given by:
and
Fig. 3.18. Spatial misorientation between two columnar grains growing in the <100> direction (sche-
matic).
Heat source
Weld
Fig. 3.19. Orientation of the single crystal Fe-15Ni-15Cr disk with respect to beam travel direction (Ex-
ample (3.2)).
Welding direction
Fig. 3.20. Schematic diagram showing the pertinent orientation relations between the fusion boundary
interface normal and the dendrite growth directions (Example (3.2)).
From this it is seen that the velocity of the [100] dendrites is always equal to that of the heat
source v. In contrast, the growth rate of the [010] dendrites depends both on v and a, and will
therefore vary with position along the fusion boundary. It follows from minimum velocity
criterion that the [100] dendrites will be selected when the interface normal angle a is less than
45°, while the [010] dendrites will develop at larger angles. This is shown graphically in Fig.
3.21.
At pseudo-steady state the fusion boundary can be calculated from the Rosenthal thin plate
solution (equation (1-83)) according to the procedure shown in Example (1.10). If we include
dendrites
/V
hk ,
R
a, degrees
Fig. 3.21.Normalised minimum dendrite tip velocity vs interface normal angle a (Example 3.2)).
a correction for the latent heat of melting, the QbIn3 ratio at the melting point becomes:
Substitution of this value into equation (1-83) gives the fusion boundary contour shown in
Fig. 3.22(a). Included in Fig. 3.22 are also schematic drawings of the predicted solidification
microstructure in different sections of the weld.
The results in Fig. 3.22 should be compared with the reconstructed 3-D image of the solidi-
fication microstructure in Fig. 3.23, taken form Rappaz et al.17 Due to partial heat flow in the
z-direction, [001] dendrite trunks will also develop. Nevertheless, these data confirm the gen-
eral validity of equations (3-8) and (3-10) relating crystal growth rate to welding speed and
weld pool shape.
dendrites +x (mm)
dendrites
+y (mm)
Fusion boundary
Base plate
Fig. 3.22. Schematic representation of the weld metal solidification micro structure (Example 3.2));
(a) Top view of fusion zone, (b) Transverse section of fusion zone.
Example (3.3)
Consider a curved columnar grain of iridium which grows from the fusion boundary towards
the weld centre-line along a circle segment of length L, as shown schematically in Fig. 3.24.
Based on the assumption that the bowing is accommodated solely by branching of [010] dendrites
in the [100] direction, calculate the maximum local growth rate of the crystal during solidifica-
tion.
y
X
2
Fig. 3.23. Reconstructed 3-D image of solidification microstructure in an electron beam welded Fe-15Ni-
15Cr single crystal. The letters (a), (b) and (c) refer to [100], [010] and [001] type of dendrites, respec-
tively. After Rappaz et al.17
Weld centre-line
Fusion line
Solution
In principle, the solution to this problem is identical to that presented in Example (3.2). Refer-
ring to Fig. 3.24 the growth rate of [100] and the [010] dendrite stems is given by:
and
It follows from Fig. 3.21 that growth will occur preferentially in the [010] direction as long
as the interface normal angle a is larger than 45°, while the [100] direction is selected at
smaller angles. This means that the local growth rate of the dendrites, in practice, never will
exceed the welding speed v.
The former mechanism is of particular importance in welding, since the weld metal often
contains a high number of second phase particles which form in the liquid state. These parti-
cles can either be primary products of the weld metal deoxidation or stem from reactions
between specific alloying elements which are deliberately introduced into the weld pool through
the filler wire. The latter process is also known as inoculation.
(3-11)
Under such conditions, the nucleus will readily grow from the liquid on the substrate. Un-
fortunately, data for interfacial energies are scarce and unreliable, which makes predictions
based on equation (3-11) rather fortuitous.18
In pure metals, experience has shown that the solid/liquid interfacial energies are roughly
proportional to the melting point, as shown by the data in Fig. 3.25. On this basis, it can be
expected that the higher melting point phases will reveal the highest ySL values, and thus be
nucleants for lower melting phases. A similar situation also exists in the case of non-metallic
inclusions in liquid steel, where the high-melting point phases are seen to exhibit the highest
solid/liquid interfacial energies (see Fig. 3.26).
In contrast, very little information is available on the substrate/embryo interfacial energy
yES. For fully incoherent interfaces, yES would be expected to be of the order of 0.5 to 1 J m~2.5
However, this value will be greatly reduced if there is epitaxy between the inclusions and the
nucleus, which results in a low lattice disregistry between the two phases. In general, assess-
ment of the degree of atomic misfit between the nucleus n and the substrate s can be done on
Melting point, K
Interfacial energy, J / m 2
Fig. 3.25. Values of solid/liquid interfacial energy ySL of various metals as function of their melting points.
Data from Mondolfo.18
lnterfacial energy, J/m2
Melting point, 0 C
Fig. 3.26. Values of interfacial energy 7 5L for different types of non-metallic inclusions in liquid steel at
16000C as function of their melting points. Data compiled from miscellaneous sources.
the basis of the Bramfitt's planar lattice disregistry model :19
(3-12)
« „ « >
Fig. 3.27. Relationship between planar lattice disregistry and undercooling for different nucleants in
steel. Data compiled from miscellaneous sources.
Example (3.4)
In low-alloy steel weld metals, titanium nitride can form in the melt due to interactions be-
tween dissolved titanium and nitrogen. Assume that the TiN particles are faceted and delim-
ited by {100} planes. Calculate on the basis of equation (3-12) the minimum planar lattice
disregistry between TiN and the nucleating delta-ferrite phase under the prevailing circum-
stances. Indicate also the plausible orientation relationship between the two phases. The
lattice parameters of delta ferrite and TiN at 15200C may be taken equal to 0.293 and 0.43 lnm,
respectively.
Solution
Titanium nitride has the NaCl crystal structure, while delta ferrite is body-centred cubic, as
shown in Fig. 3.28(a) and (b). It is evident from Fig. 3.29(a) that a straight cube-to-cube ori-
entation relationship between TiN and 8-Fe will not result in a small lattice disregistry.
However, the situation is largely improved if the two phases are rotated 45° with respect to
each other (see Fig. 3.29(b)), conforming to the following orientation relationship:
A comparison with the data in Fig. 3.27 shows that the calculated lattice disregistry con-
forms to an undercooling of about 1 to 2°C. This value is sufficiently small to facilitate hetero-
geneous nucleation of new grains ahead of the advancing interface during solidification.
Considering other inclusions with more complex crystal structures, the chances of obtain-
ing a small planar lattice disregistry between the substrate and the delta ferrite nucleus are
Fe- atoms
N-atoms
Ti- atoms
(a) (b)
Fig. 3.28. Crystal structures of phases considered in Example (3.4); (a) Titanium nitride, (b) Delta ferrite.
TiN
(a)
TiN
(C)
(b)
Fig. 3.29. Possible crystallographic relationships between titanium nitride and delta ferrite (Example (3.4));
(a) Straight cube-to-cube orientation, (b) Twisted cube-to-cube orientation, (c) Details of lattice arrange-
ment along coherent TiN/d-Fe interface.
rather poor (see Fig. 3.27). Nevertheless, such particles can act as favourable sites for hetero-
geneous nucleation if 7 ^ is sufficiently large compared with yEL and 7 ^ . This is illustrated by
the following example:
Example (3.5)
In low-alloy steel weld metals 7-Al2O3 inclusions can form during the primary deoxidation
stage as discussed in Section 2.12.4.2 (Chapter 2). Based on the classic theory of heteroge-
neous nucleation, evaluate the nucleation potency of such inclusions with respect delta ferrite.
Solution
It is readily seen from Fig. 3.27 that the planar lattice disregistry between delta ferrite and
Al2O3 is very large, which indicates of a fully incoherent interface (i.e. yES « 0.75 J m" 2 ).
Moreover, readings from Figs. 3.25 and 3.26 give the following average values for the delta fer-
rite/liquid and the inclusion/liquid interfacial energies:
and
According to equation (3-11) complete wetting is achieved when ySL > yES + yEL. This
requirement is clearly met under the prevailing circumstances.
Similar calculations can also be performed for other types of non-metallic inclusions in
steel weld metals. The results are presented graphically in Fig. 3.30. It is evident that the
nucleation potency of the inclusions increases in the order SiO2-MnO, Al 2 O 3 -Ti 2 O 3 -SiO 2 -
MnO, Al2O3, reflecting a corresponding increase in the inclusion/liquid interfacial energy ySL.
The resulting change in the weld metal solidification microstructure is shown in Fig. 3.31,
from which it is seen that both the average width and length of the columnar grains decrease
with increasing Al2O3-contents in the inclusions. This observation is not surprising, consider-
ing the characteristic high solid/liquid interfacial energy between aluminium oxide and steel
(see Fig. 3.26). The important effect of deoxidation practice on the weld metal solidification
microstructure is well documented in the literature.320"22
p (degrees)
No wetting
Complete wetting
G*he/AGhom
Embryo'
A
Inclusion
(Y sf Y ES )/7 EL
Fig. 3.30. Nucleation potency of different weld metal oxide inclusions with respect to delta ferrite.
(a) AI2O3 content (wt%)
Pure AI2O3
(A%AI)weld/[%O]anaL
Pure AI2O3
<A%AIW[%0W
Fig. 3.31.Effect of deoxidation practice on the columnar grain structure in low-alloy steel weld metals;
(a) Average width of columnar grains, (b) Average length of columnar grains. Data from Kluken et al.22
Nhet(i.e. number of nuclei which form per unit time and unit volume of the melt) is interre-
lated to the energy barrier AG^ through the following equation:5
(3-13)
whereZ1 is a frequency factor, Nv is the density of nucleation sites per unit volume of the melt,
AGD is the activation energy for diffusion of atoms across the interface, and k is the Boltzmann
constant.
Since AGD is often negligible compared with AG^et in liquids, equation (3-13) reduces to:
(3-14)
Equation (3-14) shows that the nucleation rate Nhet depends both on Nv and &Ghe{. Hence,
under full wetting conditions (AGhet ~ 0), the number of nuclei which form per second and mm3
ahead of the advancing interface is directly proportional to the instantaneous concentration of
catalyst particles in the melt. Examples of such particles are TiAl3 in aluminium and TiN/Al2O3
in steel. The important effect of controlled titanium additions and subsequent TiAl3 precipi-
tation on the columnar grain structure in 1100 aluminium welds is illustrated in Fig. 3.32.
(3-15)
where co is the total grain rotation angle, and / is the average length of the columnar grains (in
mm).
Average width of grains, (x m
Fig. 3.32. Effect of titanium on the columnar grain structure in 1100 aluminium welds. The value Y is
the fractional distance from fusion line to top surface of weld metal. Data from Yunjia et alP
Weld metal
Fig. 3.33. Characteristic growth pattern of columnar grains in bead-on-plate welds (schematic).
By introducing reasonable average values for co and / in the case of SA welding of low-
alloy steel,22 we obtain:
(3-16)
Calculated values for 4>* in steel weld metals are presented in Fig. 3.34, using data from
Kluken et al.22 An expected, the critical cell/dendrite alignment angle in fully aluminium
deoxidised steel welds is seen to be very small (of the order of 2°), reflecting the fact that
nucleation of delta ferrite occurs readily at Al2O3 inclusions. The value of 4>* increases gradu-
ally with decreasing Al2O3 contents in the inclusions and reaches a maximum of about 4° for
Si-Mn deoxidised steel weld metals. This situation can be attributed to less favourable nucle-
ating opportunities for delta ferrite at silica-containing inclusions, which reduces the possibili-
ties of obtaining a change in the crystal orientation during solidification through a nucleation
and growth process.
Dendrite fragmentation
In principle, nucleation of new grains ahead of the advancing interface can also occur from
random solid dendrite fragments contained in the weld pool. Although the source of these
solid fragments has yet to be investigated, it is reasonable to assume that they are generated by
some process of interface fragmentation due to thermal fluctuations in the melt or mechanical
disturbances at the solid/liquid interface.3 At present, it cannot be stated with certainty whether
grain refinement by dendrite fragmentation is a significant process in fusion welding.26
Grain detachment
Since the partially melted base metal grains at the fusion boundary are loosely held together by
liquid films between them, there is also a possibility that some of these grains may detach
themselves from the base metal and be trapped in the solidification front.26 Like dendrite
fragments, such partially melted grains can act as seed crystals for the formation of new grains
in the weld metal during solidification if they are able to survive sufficiently long in the melt.
Next Page
R L / v cos a
Calculated from
equation (3-10)
So far, we have discussed growth of columnar grains without considering in detail the weld
metal solidification microstructure. In general, each individual grain will exhibit a substruc-
ture consisting of a parallel array of dendrites or cells. This substructure can readily be re-
vealed by etching, also in cases where it is masked by subsequent solid state transformation
reactions (as in ferrous alloys).2224
R L / v cos a
Calculated from
equation (3-10)
So far, we have discussed growth of columnar grains without considering in detail the weld
metal solidification microstructure. In general, each individual grain will exhibit a substruc-
ture consisting of a parallel array of dendrites or cells. This substructure can readily be re-
vealed by etching, also in cases where it is masked by subsequent solid state transformation
reactions (as in ferrous alloys).2224
Fig. 3.36. Optical micrograph showing the characteristic cellular-dendritic substructure in a low-alloy
steel weld. The metallographic section is normal to the columnar grain growth direction. After Kluken
etal.22
(a)
(b)
Fig. 3.39. Ion (SIMS) micrograph showing evidence of phosphorus segregations at primary solidifica-
tion (cell) boundaries in a low-alloy steel weld. After Kluken et al22
(3-17)
Provided that equilibrium exists at the solid/liquid interface, i.e. (Cs)t = IcJC1J1, equation
(3-17) can be rewritten as:
(3-18)
where DL is the diffusivity of the solute in the liquid, and ko is the equilibrium partition coeffi-
cient (Jco < 1).
(a)
Temperature
Concentration
(b)
Liquid composition
Distance (x)
(C)
Temperature
Distance (x)
Fig. 3.40. Constitutional undercooling in alloy solidification; (a) Schematic representation of binary phase
diagram, (b) Build-up of solute-enriched layer in front of solid/liquid interface, (c) Undercooled region
ahead of solid/liquid interface.
With the aid of the phase diagram in Fig. 3.40(a) it is easy to verify that the equilibrium
liquidus temperature increases with distance from the interface because of the lower solute
content. Referring to Fig. 3.40(c) the latent heat of melting AHm will diffuse away from the
interface (thereby stabilising possible interface protuberances) if the actual temperature gradi-
ent in the liquid (dTA/dx)t is less than the equilibrium liquidus temperature gradient (3T1ZdX)1.
The latter gradient can be expressed as:
(3-19)
where (dTL/dCL)i denotes the slope of the liquidus curve in the phase diagram (designated mL).
By combining equations (3-18) and (3-19), we obtain the following criterion for the inter-
face stability:
(3-20)
Taking (C1). = C0Ik0 and Rj = RL this equation can alternatively be written as:
(3-21)
Example (3.6)
In aluminium welding, binary Al-Si and Al-Mg alloys are frequently used as filler metals.
Consider stringer bead deposition with an Al-4.5wt%Si filler wire at a constant welding speed
of 1.5mm s"1. Based on equation (3-21) calculate the critical (minimum) temperature gradient
Alloying level (C )
Planar
GL/RL
Fig. 3.41. Schematic representation of the combined effect of crystal growth rate RL and melt thermal
gradient GL on the weld metal solidification microstructure.
Fusion zone
in the weld pool which gives a planar solidification front. Relevant physical data for the Al-Si
system are given below:2
Solution
When mL, ko and D1 are known the critical temperature gradient (G1)cr can readily be calcu-
lated from equation (3-21). Taking RL ~ v, we obtain:
The above calculations show that the temperature gradient in the weld pool must be ex-
tremely large in order to promote a planar solidification front. The calculated value for (GL)cr
corresponds to a cooling rate of about 92 4000C s"1 (CR. = G1R1). Besides low-heat input
electron beam and laser welding, such high cooling rates are rarely observed in fusion welding
(see Fig. 3.3). On this basis it is not surprising to find that the weld metal solidification micro-
structure is normally of the cellular or the dendritic type.
For a rough evaluation of the weld metal solidification microstructure the diagram in Fig.
3.43 can be used. This diagram summarises the various microstructures which can be obtained
(using a typical alloy with a melting range of 500C) when the imposed temperature gradient GL
and crystal growth rate RL are varied. Moving from the lower left to the upper right along the
lines at 45° leads to a refinement of the structure without changing the morphology (constant
GL/RL ratio). Crossing these lines by passing from the upper left to the lower right leads to
changes in the morphology from planar to cellular or dendritic growth, while the scale of the
microstructure remains essentially the same. The gray bands define the regions over which
one structure changes into another. An example of the application of this diagram is given
below.
Example (3.7)
Consider GTA butt welding of a 2mm thin sheet of aluminium (Al-Mg alloy) under the fol-
lowing conditions:
Based on the diagram in Fig. 3.43, estimate the weld metal solidification microstructure in
different positions from the fusion boundary. In these calculations we shall assume that the
actual thermal gradient in the weld pool is equal to the average thermal gradient within the
temperature interval from 680 to 5500C, as evaluated from the Rosenthal thin plate solution
(equation (1-83)). Relevant thermal data for the Al-Mg alloy are given in Table 1.1 (Chapter
1).
Solution
The contours of the fusion boundary and the 680 and the 5500C isotherms can be calculated
according to the procedure shown in Example (1.10). If we neglect the latent heat of melting,
the corresponding 68//?3 ratios at these temperatures become:
GL,°C/mm
RL, mm/s
Fig. 3.43. Variation of weld metal solidification microstructure with GL and RL. The diagram is based on
the ideas of Kurz and Fisher.2
Substitution of the above values into equation (1-83) gives the isothermal contours shown
in Fig. 3.44. It follows from Fig. 3.44 that the thermal gradient decreases from about 25°C
mnr 1 at the fusion boundary to approximately 100C mm"1 close to the weld centre-line. This
occurs parallel with an increase in the nominal crystal growth rate from 0.8 to 4mm s"1.
According to Fig. 3.43 the calculated values of GL and RL conform to a cellular-dendritic
solidification microstructure within the central regions of the weld and an equiaxed-dendritic
microstructure close to the weld centre-line. Both types of substructures are commonly ob-
served in aluminium weldments.2728
Weld pool (680 0C)
Fusion boundary (650 0C)
HAZ (550 0C)
+x (mm)
+Y (mm)
Fig. 3.44. Predicted shape of fusion boundary and neighbouring isotherms during thin plate aluminium
welding (Example (3.6)).
3.5.3 Dendrite morphology
Within the fully dendritic region of the weld, the dendrite morphology remains largely un-
changed over a wide range of cooling rates. Nevertheless, it will become finer as the heat is
extracted at greater rates.
(3-22)
In general, the half width b of the dendrite stem is proportional to the primary arm spacing
X1. In cases where the arrangement of the dendrite trunks can be represented by a simple
close-packed hexagonal array, the value of b is exactly equal to X1 IV3 -2 Similarly, the total
length of the dendrite stem g can be calculated by considering the difference between the tip
temperature Ttip and the root temperature Troot, as shown in Fig. 3.48. Taking the thermal
gradient in the mushy zone equal to GL, we obtain:
(3-23)
This gives the following relationship between the primary dendrite arm spacing X1, the tip
radius rd and the thermal gradient GL:
(3-24)
rd
b
Fig. 3.46. Definition of primary X1 and secondary X2 dendrite tig. 3.47.Geometrical relationship
arm spacings. between the radius of curvature rd
and the width to length ratio of an
elliptical dendrite.
Ttip
Troot
9 X
Heat flow
Cell/dendrite
Fig. 3.48. Definition of the dendrite tip Ttip and root Troot temperatures.
If we also take into account that the dendrite tip radius rd is inversely proportional to the
square root of the crystal growth rate RL within the central range of RL (see Fig. 3.45), equation
(3-24)reduces to:
(3-25)
where c{ is a kinetic constant which is characteristic of the alloy system under consideration.
It is evident from the above analysis that the primary dendrite arm spacing cannot readily
be characterised by one single parameter (e.g. the cooling rate), since its dependence on GL
and R1 have different exponents.
Example (3.8)
Based on equation (3-25), show that the following relationships exist between the primary
dendrite arm spacing X1, the net arc power qo, and the weld travel speed v during thick plate
and thin plate welding, respectively:
The thermal gradient GL in the mushy zone close to the weld centre-line can be obtained
by differentiating of the above equation with respect to the x-coordinate:
(3-28)
By inserting the appropriate expressions for GLandRL into equation (3-25) (noting that RL=v
at the weld centre-line), we arrive at an expression for X1 which is identical with the one
presented in equation (3-26).
Similarly, under 2-D heat flow conditions, the pseudo-steady state temperature distribution
is given by the Rosenthal thin plate solution (equation 1-81). For points located on the weld
centre-line behind the heat source y = 0 and r = -x = vt. If IxI is sufficiently large, it is a fair
approximation to set Ko(u) ~ exp(-u)^n!2u . Hence, equation (1-81) reduces to:
(3-29)
From this we see that the primary dendrite arm spacmg X1 during thin plate welding is
interrelated to qo and v through a relationship of the type shown in equation (3-27).
In Fig. 3.49 the validity of equation (3-26) has been checked against the experimental data
of Jordan and Coleman,27 who measured the primary dendrite arm spacing in different GMA
Al-Mg-Mn welds. It is evident from this plot that their data-points can approximately be
represented by straight lines passing through the origin, as required by the theory. Moreover,
a closer inspection of the figure shows that the dendrite arm spacing close to the centre-line
varies systematically from the bottom to the top of the weld. This observation is not surpris-
ing, considering the fact that the rate of solidification increases progressively from the toe to
the surface of the plate.
Example (3.9)
Consider GTA butt welding of a 2mm thin aluminium sheet (Al-Mg alloy) under conditions
similar to those employed in Example (3.7). Based on equation (3-25) calculate the relative
change in the primary dendrite arm spacing X1 from the fusion boundary to the weld centre-
line during solidification.
GMAW (Al-Mg-Mn alloys)
Just above toe of weld
Level with surface of
Primary dendrite arm spacing, p,m
plate
Parallel to plate
surface
Just above toe.
of weld
[(qo)1/2/(v)1/4],(WS1/2/mm1/2)1/2
Fig. 3.49. Experimental verification of equation (3-26). Data from Jordan and Coleman.27
Solution
As shown in Fig. 3.44, the thermal gradient, GL, decreases from about 25°C mm"1 at the
fusion boundary to approximately 10°C mm"1 close to the weld centre-line. This occurs
parallel with an increase in the nominal crystal growth rate from 0.8 to 4mm s"1. Taking the
primary dendrite arm spacing at the fusion boundary equal to X1*, the ratio X1 /X1* in different
positions of the weld becomes:
Weld centre-line:
In contrast to that predicted above, the smallest dendrite arm spacing is normally observed at
the weld centre-line.26 This has to do with the fact that the constant C1 in equation (3-25), in
practice, decreases with increasing distance from the fusion boundary due to solute segrega-
tion, which gradually reduces the coarsening rate of the dendrites.
(3-30)
(3-31)
The parameter Q in equation (3-31) refers to the Gibbs-Thomson coefficient and is given as:
(3-32)
where a is the solid/liquid interfacial energy, and AS, is the entropy of fusion.
In practice, the value of M can easily vary by an order of magnitude. Nevertheless, its effect
on the secondary dendrite arm spacing is rather weak, since X2 is proportional to the cube root
of M. Hence, a plot of the secondary dendrite arm spacing vs the local solidification time in a
logarithmic diagram will normally reveal a straight-line relationship between X2 and to with a
slope close to 0.33, as shown in Fig. 3.50.
Slopes 0.33
to'5
Fig. 3.50. Relation between secondary dendrite arm spacing X2 and local solidification time to in type
AISI 310 stainless steel welds. Data from Kou and Lee.31
Example (3.10)
Based on equation (3-30), derive a relationship between the secondary dendrite arm spacing
X2, the net arc power qo, and the weld travel speed v during thick plate welding. Calculate then
the secondary dendrite arm spacing in the centre of a thick GTA Al-Si weld deposited under
the following conditions:
Relevant physical data for the Al-Si system are given below:2
Solution
The local solidification time to is the time for the dendrite array to pass an arbitrary point in the
weld and is therefore a measure of the solidification rate. Referring to Fig. 3.48, the local
solidification time is defined as:
(3-33)
(3-34)
(3-35)
Although reliable experimental data are not available for a direct comparison, the calcu-
lated value for X2 is considered reasonably correct. In Al-Cu castings a secondary dendrite
arm spacing of about 4(im corresponds to a cooling rate of the order of 103 °C s"1, as shown in
Fig. 3.51. In the present example, the cooling rate at the solid/liquid interface is close to:
Example (3.11)
The operating parameters given in Example (3.10) are also applicable to single pass butt weld-
ing of thin aluminium plates. Based on equation (3-30), estimate the secondary dendrite arm
spacing in the centre of a 2mm thick Al-Si butt weld deposited under such conditions.
Solution
In thin plate welding the local solidification time to is obtained by combining equations (3-29)
and (3-33). Noting that RL = v at the weld centre-line, we get:
(3-36)
AI-4.5 wt% Cu
X 2 ,nm
Fig. 3.57.Relation between secondary dendrite arm spacing X2 and cooling rate CR. in Al-Cu castings.
Data compiled by Munitz.32
The secondary dendrite arm spacing is thus given as:
(3-37)
3
Taking pc equal to 0.0027 J mm C ] in the case of Al-Si alloys, we obtain:
It is evident from the above calculation that the secondary dendrite arm spacing is signifi-
cantly coarser in thin plate welding compared with thick plate welding. This observation is not
surprising, considering the pertinent difference in the heat transfer mode and thus the cool-
ing rate between these two types of weldments. In the former case, we have:
A comparison with the data in Fig. 3.51 shows that a cooling rate of about 78°C s"1 is
compatible with a secondary dendrite arm spacing of 13 jim.
As already stated in the introduction of the chapter, an equiaxed zone is often observed close to
the weld centre-line. This zone can in certain cases be very dominating and completely over-
ride the columnar grain zone, as shown in Fig. 3.52.
As long as growth occurs in a columnar manner the crystals are in contact with solid metal.
The heat will therefore be conducted through the crystals in a direction which is parallel and
opposite to their growth direction, as illustrated in Fig. 3.53(a). In equiaxed dendritic growth
the situation is different. Here the melt will be slightly undercooled because of extensive seg-
regation of solute to the weld centre-line where the solidification fronts growing from each
side impinge. This makes the crystals hotter than the liquid and gives rise to a radial heat flow
away from the crystals in the same direction as that of growth (see Fig. 3.53(b)).
FL CL
Fig. 3.52.Example of grain refinement in an Al-Mg-Si plasma arc weld due to TiAl3-precipitation (FL:
fusion line, CL: centre-line). Courtesy of M.I. Onsoien, SINTEF, Trondheim, Norway.
transition occurs when the GJR1 ratio drops below a certain critical value fcn. Hence, we may
write:
(3-38)
During thick plate welding (3-D heat flow) the temperature gradient at the weld centre-line
is given by equation (3-28). Taking RL = v, a combination of equations (3-28) and (3-38) gives:
(3-39)
where c4 is a kinetic constant which is characteristic of the alloy system under consideration.
Similarly, for welding of medium thick plates (mixed heat flow), we get:
(3-40)
where c5 is a new kinetic constant, and n is an exponent which varies between zero and unity
depending on the mode of heat flow (i.e. thin, medium thick or thick plate welding).
(a) (b)
Heat flow
Fig. 3.53. Schematic diagrams showing different dendrite growth morphologies in castings; (a) Colum-
nar dendritic growth, (b) Equiaxed dendritic growth. The diagrams are based on the ideas of Kurz and
Fisher.2
Equation (3-40) predicts that the columnar to equiaxed transition in fusion welds occurs at
critical combinations of the net arc power qo and the welding speed v. Thus, a decrease in qo
must always be compensated by a corresponding increase in v in order to maintain equiaxed
dendritic growth at the weld centre-line. This is also in agreement with general experience
(see Fig. 3.54).
Example (3.12)
Based on the experimental data in Fig. 3.54, calculate the critical G1IR1 ratio which provides
equiaxed dendritic growth during thick plate welding of Al-Mg alloys (5083 series). Relevant
thermal data for the Al-Mg system are given in Table 1.1 (Chapter 1).
Plate thickness: 10-12 mm
Fig. 3.54. Effect of net arc power qo and welding speed v on the columnar to equiaxed transition in
different aluminium welds. Data from Matsuda et al.33
Solution
The critical G1IR1 ratio is obtained by rearranging equation (3-39):
By multiplying fcr with the welding speed we see that the calculated value corresponds to
a critical temperature gradient of about 928°C mm"1. This means that the thermal conditions
existing in welding will favour growth of equiaxed dendrites close to the weld centre-line. A
requirement is, of course, that the melt contains a sufficient number of seed crystals to facili-
tate heterogeneous nucleation of new grains ahead of the advancing solid/liquid interface.
Next Page
The former mechanism is particularly relevant to welding, since the weld metal often con-
tains a high number of second phase particles which form in the liquid state. As already
mentioned in Section 3.4.4.2, these particles can either be primary products of the weld metal
deoxidation or stem from reactions between specific alloying elements which are deliberately
introduced into the weld pool through the filler wire. The important effect of deoxidation
practice (inclusions) on the columnar to equiaxed transition in ferritic stainless steel GTA
welds is shown in Figs. 3.55 and 3.56.
During solidification of fusion welds, alloying and impurity elements tend to segregate exten-
sively to the centre parts of the intercellular or interdendritic spaces under the conditions of
rapid cooling. This, in turn, alters the kinetics of the subsequent solid state transformation
reactions.
(i) Uniform liquid composition (i.e. complete mixing in the liquid state).
(ii) A flat solid/liquid interface.
(iii) Local equilibrium at the solid/liquid interface (ko is constant).
(iv) Negligible solid-state diffusion.
(v) Equal solid and liquid densities.
Under such conditions it is fairly simple to derive an expression for the solute concentration
in the metal as a function of the fraction solidified. Referring to Fig. 3.57, a mass balance
gives:
(3-41)
The former mechanism is particularly relevant to welding, since the weld metal often con-
tains a high number of second phase particles which form in the liquid state. As already
mentioned in Section 3.4.4.2, these particles can either be primary products of the weld metal
deoxidation or stem from reactions between specific alloying elements which are deliberately
introduced into the weld pool through the filler wire. The important effect of deoxidation
practice (inclusions) on the columnar to equiaxed transition in ferritic stainless steel GTA
welds is shown in Figs. 3.55 and 3.56.
During solidification of fusion welds, alloying and impurity elements tend to segregate exten-
sively to the centre parts of the intercellular or interdendritic spaces under the conditions of
rapid cooling. This, in turn, alters the kinetics of the subsequent solid state transformation
reactions.
(i) Uniform liquid composition (i.e. complete mixing in the liquid state).
(ii) A flat solid/liquid interface.
(iii) Local equilibrium at the solid/liquid interface (ko is constant).
(iv) Negligible solid-state diffusion.
(v) Equal solid and liquid densities.
Under such conditions it is fairly simple to derive an expression for the solute concentration
in the metal as a function of the fraction solidified. Referring to Fig. 3.57, a mass balance
gives:
(3-41)
(a)
(b)
Fig. 3.55. Effect of deoxidation practice (aluminium additions) on the columnar to equiaxed transition in
ferritic stainless steel GTA welds; (a) Average fraction of equiaxed grains observed at the surface of the
welds vs aluminium content, (b) Surface equiaxed grain size vs aluminium content. Data from Villafuerte
etal21
(3-42)
(3-43)
(a)
(b)
Average grain size, JI m
Fig. 3.56. Effect of deoxidation practice (titanium additions) on the equiaxed transition in ferritic stain-
less steel GTA welds; (a) Average fraction of equiaxed grains observed at the surface of the weld vs
titanium content, (b) Surface equiaxed grain size vs titanium content. Data from Villafuerte et al.21
from which
(3-44)
and
(3-45)
Distance, z
Fig. 3.57. Solute redistribution during non-equilibrium freezing according to the Scheil equation.
Equation (3-44) is valid up to CL = Ceut where the remaining melt solidifies in the form of
intercellular or interdendritic eutectics. The eutectic fraction feut is, in turn, given as:
(3-46)
Figure 3.58 shows how the Scheil equation can be used for an evaluation of the
microsegregation pattern in binary alloy systems by considering a small volume element of
length L* which solidifies perpendicular to the cell/dendrite growth direction.
Since the Scheil equation does not allow for solid state diffusion during solidification, a
slightly refined version of this equation also exists in the literature:1
(3-47)
(3-48)
Solid
Liquid
Equation (3-47) has been used by Brody and Flemings34 to evaluate the effect of solid
diffusion on the amount of eutectic in different cast structures. The extent of this diffusion
depends on the dimensionless product a* ko and it becomes significant only for values of
a* Jc0 greater than about O.I.1 Note that in cellular or cellular-dendritic growth the primary
dendrite arm spacing X1 provides a measure of the diffusion length. Taking L* = X ,/2, equation
(3-48) can be rewritten as:
(3-49)
In contrast, during equiaxed dendritic growth the secondary dendrite arm spacing X2 is a
more appropriate dimension for the solidification microstructure, since the back diffusion process
here occurs mainly between secondary arms and not between primary trunks.2 In such cases
we may write:
(3-50)
Example (3.13)
Consider GTA welding of an Al-2wt%Cu alloy under the following solidification condi-
tions:
Welding speed
Cooling rate
Primary dendrite arm spacing
Estimate on the basis of the Scheil equation the degree of microsegregation occurring dur-
ing weld metal solidification. Relevant physical data for the Al-Cu system are given below:
Solution
The local solidification time can be calculated from equation (3-33). Taking RL = v, we get:
Since the solidification conditions in this case facilitate the formation of a cellular-dendritic
type of substructure close to the weld centre-line (see Fig. 3.43), the characteristic diffusion
length L* is determined by the primary dendrite arm spacing X1. The product a* ko is then
given as:
Because the numerical value of a* ko is very small, the contribution from diffusion in the
solid state can be neglected. Hence, the extent of microsegregation occurring during solidifi-
cation can be evaluated from equation (3-45). Taking C0 = 2wt% and ko = 0.17, we obtain:
These results should be compared with the experimental data of Brooks and Baskes35
replotted in Fig. 3.60. It is evident that the measured copper concentration profile for the Al-
2wt%Cu GTA weld is similar to that inferred from the Scheil equation, although the observed
dendrite core concentration lies significantly above the predicted one. Consequently, the Scheil
equation gives a reasonable description of the segregation pattern during weld metal solidifi-
cation, in spite of the simplifying assumptions inherent in the model.
z.^m
Cu concentration, wt%
Position, ]im
Fig. 3.60. Electron microprobe analysis of Cu across primary solidification (cell) boundaries in an Al-
2wt% Cu GTA weld. Data from Brooks and Baskes.35
3.61). Although the origin of the phenomenon is not yet fully understood, it is reasonable to
assume that the pertinent fluctuations in the solidification rate occurs as a result of frequent
variations in the heat flux during welding. Direct experimental evidence for such a correlation
can be obtained from the data of Garland and Davis.36
Moreover, there is a pronounced tendency for alloying and impurity elements to segregate
to the weld centre-line where the columnar grains growing from each side impinge. This, in
turn, may produce hot tearing as a result of the formation of low-melting eutectics between the
dendrite arms. In general, the risk of hot tearing decreases with increasing width to depth ratio
of the weld because of a more favourable crystal growth mode, as illustrated in Fig. 3.62.
Width Width
Depth
Depth
Fig. 3.62. Effect of weld width to depth ratio on the tendency to centre-line cracking; (a) Correct width to
depth ratio, (b) Incorrect width to depth ratio.
Fig. 3.63. Example of gas porosity in a GTA 7106 aluminium weld. After D'annessa.39
Solid
Fig. 3.64. Growth of a gas bubble due to diffusion in the liquid phase (schematic).
a*, the metallostatic pressure head/?m, and the ambient pressure pa. Thus, for a stable bubble,
we may write:1
(3-51)
where pg is the total gas pressure inside the bubble, and rg is the radius of the gas bubble.
For shallow welds, the contribution from the metallostatic pressure head pm can be ig-
nored. Hence, equation (3-51) reduces to:
(3-52)
Since a* is typically of the order of 1 J m 2 (9.87 atm |im) for most gas-metal systems, we
may write:
(3-53)
It is evident from equation (3-53) that the interfacial energy term is negligible at large
values of r . However, if the radius of curvature becomes sufficiently small, extremely large
pressures are required to maintain a stable bubble. Thus, there is a bubble nucleation prob-
lem, which is formally similar to that of nucleation of a solid from a liquid (discussed in
Section 3.3.1). In fact, it can be shown on the basis of classic nucleation theory that the driving
force normally associated with rejection of dissolved gases in liquid metals is by far too small
to allow for homogeneous nucleation of gas bubbles in the weld pool during cooling. This, in
turn, implies that solid particles (e.g. inclusions) entrained in the liquid metal will be the most
probable sites for gas bubble formation in fusion welds.
(3-54)
(3-55)
where CL is the molar concentration of solute in the supersaturated liquid, C6 is the equilibrium
molar concentration of the solute at the gas/liquid interface, and p^ is the gas density (in the
same units as CL and C6).
Equations (3-54) and (3-55) may be used to estimate the growth rate of a bubble while it is
still attached to the solid/liquid interface. The bubble becomes detached when the buoyancy
force, which is pushing it upwards, exceeds the surface tension force, which tends to keep it
attached to the solid surface. The bubble radius at which detachment occurs is given by the so-
called Fritz equation:38
(3-56)
where gc is the gravity constant, and P is the wetting angle (in degrees).
Based on equations (3-54) and (3-56) it is possible to evaluate the conditions for growth
and detachment of gas bubbles during weld metal solidification. This is a subject of consider-
able importance in welding, since the pore formation will inevitably affect the mechanical
integrity of the weldment.
Example (3.14)
Consider GTA butt welding of a 3mm thin Al-Mg sheet under the following conditions:
Suppose that gas bubbles form at the solid/liquid interface during solidification due to re-
jection of dissolved hydrogen from a supersaturated liquid. Based on equations (3-54) and
(3-56) estimate the maximum theoretical radius of the gas bubbles and the critical radius at
which the bubbles detach themselves from the solid/liquid interface during welding. Relevant
physical data for the Al-Mg system are given below:
Solution
Since particles located at the solid/liquid interface are the most probable sites for hydrogen
gas evolution, the local solidification time t0 provides a conservative estimate of the growth
time t in equation (3-54). From equation (3-36), we have:
In order to calculate the growth constant from equation (3-55), it is necessary to convert the
concentration driving force to molar units:
The molar density of the gas p is obtained from the ideal gas law:
The maximum theoretical radius of the gas bubbles can now be evaluated from equation
(3-54) by inserting the appropriate values for Q*, DH, and to:
Similarly, the critical radius at which the hydrogen bubbles become detached may be esti-
mated from equation (3-56). Since PM > > PH 2 , we obtain:
By inserting this value into equation (3-54) it is also possible to estimate the average bubble
detachment frequency under the prevailing circumstances:
Since r (crit.)« r (max), the maximum pore radius will probably be closer to 0.7mm than
30mm in a real welding situation. This is also in agreement with practical experience (see Fig.
3.65).
GTAW (Aluminium)
Diameter of pores, mm
Fig. 3.65. Measured distribution of pore diameters in some GTA aluminium welds deposited with differ-
ent hydrogen-containing shielding gases. Data from Tomii et al.40
(3-57)
where dg is the diameter of the gas bubbles, and jLl is the viscosity of the liquid.
Larger gas bubbles (characterised by a bubble Reynold number between 2 and 400) will
also rise in a rectilinear manner, but their terminal velocity may be as much as 50% greater
than that predicted from Stokes law.38 Depending on their flotation rate, such ascending gas
bubbles will either escape to the weld surface or be trapped in the weld metal solidification
front in the form of macroscopic gas porosity (see Fig. 3.63).
Example (3.15)
Based on Stokes law (equation (3-57)) calculate the rising velocity of a 0.2mm large hydrogen
bubble ascending in liquid aluminium. Relevant physical data for liquid aluminium are given
below:
Solution
Since p z » p , we may write:
It is evident from the above calculations that the flotation rate of such gas bubbles is quite
high and of the same order of magnitude as the weld pool fluid flow velocity (discussed in
Section 2.11.2). Hence, the buoyancy force would be expected to play a significant role in the
separation process of detached gas bubbles in the weld pool. On this basis it is not surprising to
find that a change in the welding position (e.g. from flat to overhead) results in a dramatic
increase in the volume of porosity during GTAW of aluminium alloys (see Fig. 3.66).
Although a great deal has been reported on the causes and effects of porosity in weld metals
(see Ref. 37 for an excellent discussion), little is known about the mechanism of pore forma-
tion relative to solidification mechanics, nucleation, growth and transport of gas bubbles in the
weld pool. Consequently, a more fundamental approach to the porosity problem in fusion
welding (along the lines indicated above) is necessary in order to obtain a verified, quantitative
understanding of the phenomenon.
GTAW
Overhead position
Volume of pores, ml/10Og
Flat position
Fig. 3.66. Porosity in GTA welds deposited on 2mm sheets of aluminium at various orientations from flat
to overhead. Data compiled by Devletian and Woods.37
3.7.4 Removal of microsegregations during cooling
As shown in Section 3.7.1, the characteristic growth pattern of cellular and dendritic solidifi-
cation, in combination with the rapid cooling rates normally associated with fusion welding,
lead to extensive segregation of alloying and impurity elements to the intercellular or
interdendritic spaces. Segregation produced by this means is remarkably persistent, and can in
certain cases only be eliminated by prolonged high-temperature heat treatment. A simplified
analysis of homogenisation of microsegregations in fusion welds is given below.
where C and Cmax are as indicated in Fig. 3.67, lavg is the average distance between adja-
cent maxima and minima, and Ds is the diffusivity of the solute in the solid.
Equation (3-58) states that the concentration remains constant and equal to Cavg at positions
JC = 0, JC = lavg, x = 2/ etc., while the peak of the sine wave is attained at distances x = lavg /2,
x = 5lavg /2 etc. during the decay. If only the peak concentration is considered, the sine term
becomes equal to unity and equation (3-58) reduces to:
(3-59)
from which the homogenisation time during isothermal heat treatment thom can be obtained:
(3-60)
Since the diffusion length I , in practice, is equal to the half dendrite arm spacing (X1 /2 or
X2/2), equation (3-60) predicts that the homogenisation time is proportional to the square OfX1
or X2. The latter parameters are, in turn, determined by the thermal conditions existing within
the mushy zone during weld metal solidification, and are therefore sensitive to variations in
welding variables such as the net arc power qo and the travel speed v.
(3-61)
Dendrite arm spacing
Concentration
The integral on the right-hand side of equation (3-61) represents the kinetic strength of the
weld thermal cycle with respect to homogenisation (solid state diffusion), and can be deter-
mined by means of numerical methods when the cooling programme is known. The extent of
solute diffusion may then be evaluated from equation (3-60) by inserting representative
values for t and / :
T UY'g.
(3-62)
Example (3.16)
[n low-alloy steel, both carbon, phosphorus, and manganese are known to segregate to the
interdendritic spaces during solidification. Consider SA welding on a thick plate of steel
under the following conditions:
Based on equations (3-61) and (3-62) calculate the extent of homogenisation occurring
within the solid weld metal during cooling in the austenite regime (i.e. from 1520 to 7000C).
Relevant data for the diffusivity of carbon, phosphorus, and manganese in austenite are given
below:
Solution
During thick plate welding, the cooling programme can be calculated from equation (1-45).
For points located on the weld centre-line behind the heat source y = z = 0, and R* = -x = vt.
Hence, equation (1-45) reduces to:
From this it is seen that the actual transformation temperatures (i.e. 1520 and 7000C) are
reached after 20.9 and 46.2s, respectively. If 13500C is used as a reference temperature, the
kinetic strength of the cooling cycle with respect to solute diffusion can be expressed as:
Numerical integration of this equation over the weld cooling cycle gives:
and
from which
This value is reasonably close to that obtained during high-heat input SA welding
(0.036 s jirrr2), as shown in the above example.
Fig. 3.68. Banding of martensite (M) along primary solidification boundaries in an AISI 4340 SA steel
weld. After Burck.42
3.8 Peritectic Solidification
Referring to the schematic phase diagram in Fig. 3.69, crystallisation within peritectic systems
starts with primary precipitation of the 8p-phase from the liquid. At the peritectic temperature
T a new solid phase 7 forms, according to the reaction:
5p+ liquid -> yp <3"63)
Because of the nature of this reaction, there is a strong tendency for the secondary phase 7
to grow along the 8^/liquid interface and, thus, to isolate the primary phase from contact with
the liquid.43 Depending on the growth mode, the decomposition of the primary 8^-phase is
said to occur either through a peritectic reaction or by a peritectic transformation,44 as
shown schematically in Fig. 3.70 and 3.71, respectively.
Growth direction
Liquid Diffusion of
solute atoms
Fig. 3.70.Peritectic reaction by which the secondary 7p-phase grows along the surface of the primary
8p-phase (schematic).
Liquid
Fig. 3.7L Peritectic transformation involving long-range diffusion of solute atoms through the secondary
7^-phase (schematic).
Delta ferrite
Liquid
Solute segregations
Austenite
Austenite grain
boundary
Fig. 3.72. Primary delta ferrite solidification with subsequent growth of austenite along the boundaries of
the primary 8Fe-phase (schematic).
Austenite
Columnar
grain
Liquid
at prior solidification (cell) and austenite grain boundaries*, respectively in a SA steel weld
metal deposited under the shield of a basic flux. It is obvious from Fig. 3.75 that there is no
matching between the two types of boundaries in this particular case.
It follows from the analysis of Kluken et al?2 that the observed shift in the mechanism of
the peritectic transformation can probably be attributed to heterogeneous nucleation of
austenite at inclusions (e.g. Al2O3), which is energetically more favourable than nucleation at
8Fe/8Fe grain boundaries. Under such conditions, the austenite is not bound by an orientation
relationship with the delta ferrite and is thus free to grow across the original delta ferrite
columnar grain boundaries, as shown schematically in Fig. 3.76. The austenite grains will
therefore adopt a morphology which is different from the columnar one.245556
References
1. M.C. Flemings: Solidification Processing, 1974, New York, McGraw-Hill Book Company.
2. W. Kurz and DJ. Fisher: Fundamentals of Solidification, 3rd Edn, 1989, Aedermannsdorf
(Switzerland), Trans. Tech. Publications.
3. GJ. Davis and J.G. Garland: Int. MetalL Rev., 1975, 20, 83-106.
4. S.A. David and J.M. Vitek: Int. Mater. Rev., 1989, 34, 213-245.
5. D. A. Porter and K.E. Easterling: Phase Transformations in Metals and Alloys, 1981, Wokingham
(England), Van Nostrand Reinhold Co. Ltd.
6. K.E. Easterling: Introduction to the Physical Metallurgy of Welding, 1983, London, Butterworths
& Co (Publisher) Ltd.
7. 0. Grong, TA. Siewert and G.R. Edwards: Weld. J., 1986, 65, 279s-288s.
8. G.M. Oreper, TW. Eagar and J. Szekely: Weld. J., 1983, 62, 307s-312s.
9. Y.H. Wang and S. Kou: Proc. Int. Conf. on Advances in Welding Science, Gatlinburg, TN,
May, 1986, 65-69, Publ. ASM International.
*Because of a similar atomic weight of BO2~ and AlO~, both types of ions will be sampled in the SIMS analysis.
Consequently, the white spots observed within the interior of the grains are probably traces of aluminium oxide
inclusions.
(a)
(b)
Fig. 3.75. Ion (SIMS) micrographs showing austenite grain boundaries crossing primary delta ferrite
solidification boundaries in a low-alloy SA steel weld; (a) Phosphorus segregations at primary solidifica-
tion (cell) boundaries, (b) Boron segregations at prior austenite grain boundaries (same area as in Fig.
3.75(a)). After Kluken et alP
Delta ferrite.
Liquid
Solute segregations
Fig. 3.76. Primary delta ferrite solidification with subsequent nucleation of austenite at inclusions (sche-
matic).
10. K.C. Mills and BJ. Keene: Int. Mater. Rev., 1990, 35, 185-216.
11. T. Zacharia, S.A. David, J.M. Vitek and H.G. Kraus: Metall. Trans., 1991, 22B, 243-257.
12. CR. Heiple and J.R. Roper: Weld. J., 1982, 61, 97s-102s.
13. CR. Heiple and P. Burgardt: Weld. J., 1985, 64, 159s-162s.
14. T. Habrekke, H.O. Knagenhjem and J.O. Berge: Proc. 9th. Int. Conf. on Offshore Mech. and
Arctic Engin., Houston, TX, Feb., 1990, 511-515, Publ. ASME.
15. H. Biloni: Physical Metallurgy, 3rd Edn (Eds R.W. Chan and P. Haasen), 1983, Amsterdam,
North-Holland Physics Publ., 478-579.
16. T. Senda, F. Matsuda and M. Kato: Techn. Reports of the Osaka University (Japan), 1970,20,
527-558.
17. M. Rappaz, S.A. David, J.M. Vitek and L.A. Boatner: Metall. Trans. A, 1989,2OA, 1125-1138.
18. L.F. Mondolfo: Mat. ScL and TechnoL, 1989, 5, 118-122.
19. B.L. Bramfitt: Metall. Trans., 1970, 1, 1987-1995.
20. G.N. Heintze and R. McPherson: Weld. /., 1986, 65, 71s-81s.
21. J.C. Villafuerte, E. Pardo and H.W. Kerr: Metall. Trans. A, 1990, 21A, 2009-2019.
22. A.O. Kluken, 0. Grong and G. R0rvik: Metall. Trans. A, 1990, 21 A, 2047-2058.
23. H. Yunjia, R.H. Frost, D.L. Olson and G.R. Edwards: Weld. J., 1989, 68, 280s-289s.
24. CE. Cross, 0. Grong, S. Liu and J.F. Capes: Applied Metallography, (Ed. G.F. Vander Voort),
1986, New York, Van Nostrand Reinhold Inc., 197-210.
25. A.O. Kluken and 0 . Grong: Proc. Int. Conf. on Recent Trends in Welding ScL and TechnoL
(TWR 89), Gatlinburg, TN, May, 1989, 781-786, Publ. ASM International.
26. S. Kou: Welding Metallurgy, 1987, New York, John Wiley & Sons.
27. M.F. Jordan and M.C. Coleman: Brit. Weld. J., 1968,15, 553-558.
28. CE. Cross and G.R. Edwards: Treatise on Mat. ScL TechnoL, 1989, 31, 171-187.
29. W. Kurz, B. Giovanola and R. Trivedi: Acta Metall., 1986, 34, 823-830.
30. R. Trivedi and W. Kurz: Acta MetalL, 1986, 34, 1663-1670.
31. S. Kou and Y. Lee: MetalL Trans., 1982,13A, 1141-1152.
32. A. Munitz: MetalL Trans., 1985,16B, 149-161.
33. F. Matsuda, K. Nakata, K. Tsukamoto and K. Arai: Trans. JWRI, 1983,12, 81-87.
34. H.D. Brody and M.C. Flemings: Trans. AIME, 1966, 236, 615-623.
35. J.A. Brooks and M.I. Baskes: Proc. Int. Conf. on Trends in Welding Research (TWR 86),
Gatlinburg, TN, May, 1986, 93-99, Publ. ASM International.
36. J.G. Garland and GJ. Davis: Brit. Weld. J., 1970,17, 171-175.
37. J.H. Devletian and W.E. Woods: Weld. Res. Council Bull., 1983, 290, 1-18.
38. J. Szekely and NJ. Themelis: Rate Phenomena in Process Metallurgy, 1971, New York, John
Wiley & Sons, Inc.
39. A.T. D'annessa: Weld. J., 1967, 46, 491s-499s.
40. Y. Tomii, A. Sakaguchi and M. Mizuno: Proc. 4th Int. Conf. on Aluminium Weldments, Tokyo
(Japan), 1988, 46-60, Publ. Japan Light Metal Welding and Construction Association.
41. J.D. Verhoeven: Fundamentals of Physical Metallurgy, 1975, New York, John Wiley & Sons,
Inc.
42. P. Burck: M.Sc. thesis, 1984, Colorado School of Mines, Golden, Colorado (USA).
43. M. Hillert: Solidification and Casting of Metals, 1979, London, The Metals Society, 81-87.
44. H. W. Kerr, J. Cisse and G.F. Boiling: Acta MetalL, 1974, 22, 677-686.
45. J. Cisse, G.F. Boiling and H.W. Kerr: /. Cryst. Growth, 1972,13/14, 777-781.
46. J. Cisse, H.W. Kerr and G.F. Boiling: MetalL Trans., 1974, 5, 633-641.
47. H. Fredriksson: Met. ScL, 1976,10, 77-86.
48. H. Fredriksson and J. Stjerndahl: MetalL Trans., 1977, 8A, 1107-1115.
49. H. Fredriksson and J. Stjerndahl: Met. ScL, 1982,10, 575-585.
50. N.S. Pottore, CL Garcia and AJ. DeArdo: MetalL Trans., 1991, 22A, 1871-1880.
51. A.O. Kluken, 0. Grong and J. Hjelen: MetalL Trans., 1991, 22A, 657-663.
52. N. Suutala, T. Takalo and T. Moisio: MetalL Trans., 1979,10A, 1183-1190.
53. S.A. David: Weld. J., 1981, 60, 63s-71s.
54. A.A.B. Sugden and H.K.D.H. Bhadeshia: MetalL Trans., 1987,19A, 669-674.
55. J.G. Garland and RR. Kirkwood: Proc. Int. Symp. on Welding of Line Pipe Steels, St. Louis
(USA), 1977, 176-227, Publ. Welding Research Council (New York).
56. R.C. Cochrane: Weld, in the World, 1983, 21, 16-29.
Appendix 3.1
Nomenclature
4.1 Introduction
Referring to Fig. 4.1, particle coarsening occurs typically at temperatures well below the
equilibrium solvus Te of the precipitates, while particle dissolution is the dominating mecha-
nism at higher temperatures. On the other hand, there exists no clear line of demarcation
between these two processes, which means that particle coarsening can take place simultane-
ously with reversion in certain regions of the weld where the peak temperature of the thermal
cycle falls within the 'gray zone' in Fig. 4.1. Nevertheless, it is important to regard them as
separate processes, since the reaction kinetics are so different (coarsening is driven by the
surface energy alone, whereas dissolution, which involves a change in the total volume frac-
tion, is driven by the free energy change of transformation).
The symbols and units used throughout this chapter are defined in Appendix 4.1.
The solubility product is a basic thermodynamic quantity which determines the stability of
the particles under equilibrium conditions. Because of its simple nature, the solubility product
is widely used for an evaluation of the response of grain size-controlled and dispersion-hard-
ened materials to welding and thermal processing.23
'Grey zone1
Particle coarsening |
%B
Fig. 4.1. Schematic diagram showing the characteristic temperature ranges where specific physical
reactions occur during reheating of grain size-controlled and dispersion-hardened materials.
(4-1)
At equilibrium, we have:
(4-2)
where AH° and AS° are the standard enthalpy and entropy of reaction, respectively. The other
symbols have their usual meaning (see Appendix 4.1).
When pure An Bm is used as a standard state, the activity of the precipitate {aAn Bm) is equal
to unity. In addition, for dilute solutions it is a fair approximation to set aA~[%A] and aB~[%B],
where the matrix concentrations of elements A and B are either in wt% or at%*. Hence, the
solubility product can be written as:
(4-3)
*For the solute, the standard state is usually a hypothetical 1 % solution. This implies that the activity coefficient is
equal to unity as long as Henry's law is obeyed.
Table 4.1 gives a summary of equilibrium solubility products for a wide range of precipi-
tates in low-alloy steels and aluminium alloys.
In addition to the compounds listed in Table 4.1, different types of mixed precipitates may
form within systems which contain more than two alloying elements.3"6 However, since the
presence of such multiphase particles largely increases the complexity of the analysis, only
pure binary intermetallics will be considered below.
(4-4)
where [%A]o and [%B]o refer to the analytical content of elements A and B in the base metal,
respectively.
Equation (4-4) shows that the equilibrium dissolution temperature increases with increas-
ing concentrations of solute in the matrix. This is in agreement with the Le Chatelier's princi-
ple.
Table 4.1 Equilibrium solubility products for different types of precipitates in low-alloy steels and
aluminium alloys. Data compiled from miscellaneous sources.
Calculate on the basis of the reported solubility products in Table 4.1 the equilibrium disso-
lution temperature of each of the following three nitride precipitates, i.e. NbN, AlN, and TiN.
Solution
The equilibrium dissolution temperature of the precipitates can be computed from equation (4-
4) by inserting the correct values for C* and D* from Table 4.1:
It is evident from these calculations that precipitates of the NbN and the AlN type will
dissolve readily at temperatures above 1050 to 11000C, while TiN is thermodynamically sta-
ble up to about 14500C. In practice, however, a certain degree of superheating is always
required to overcome the inherent kinetic barrier against dissolution, particularly if the heating
rate is high. Consequently, in a real welding situation the actual dissolution temperature of the
precipitates may be considerably higher than that inferred from simple thermodynamic calcu-
lations based on the solubility product (to be discussed later).
(4-5)
Equation (4-5) describes the solvus surface within the solvent-rich corner of the phase dia-
gram. However, when a pure binary compound dissolves the concentration of elements A and
B in solid solution is fixed by the stoichiometry of the reaction. The following relationship
exists between [%B] and [%A]:
(4-6)
or
[%A]0
[%B]
[%B]o
Excess B
[%A]
Fig. 4.2. Concentration displacements during dissolution of binary intermetallics (equilibrium condi-
tions).
I Increased dissolution temperature
Temperature
Increased additions
Nominal alloy composition
_of element B
Reduced solid solubility
of element A
Concentration of element A
Fig. 4.3. Factors affecting the solid solubility of a binary intermetallic compound in a multi-component
alloy system (schematic).
crease with increasing temperature when AH° is positive. This type of behaviour is character-
istic of intermetallics in metals and alloys, since the dissolution process in such systems is
endothermic.7 As a result, increased additions of a second alloying element B will also reduce
the solubility of the first alloying element A by shifting the 'solvus boundary' towards higher
temperatures when an intermetallic compound between A and B is formed.
With the aid of Fig. 4.3 it is easy to verify that the equilibrium volume fraction of the
precipitates/^ at a fixed temperature is given by:
(4-7)
where fmax is the maximum possible volume fraction precipitated at absolute zero.
Equation (4-7) provides a basis for estimating the equilibrium volume fraction of binary
intermetallics in complex alloy systems at different temperatures in cases where the concentra-
tion of element B is sufficiently high to tie-up all A in the form of precipitates. Similarly, if A
is present in an overstoichiometric amount with respect to B, we may write:
(4-8)
Example (4.2)
In Al-Mg-Si alloys the equilibrium Mg2Si phase may form during prolonged high tempera-
ture annealing. Consider a pure ternary alloy which contains 0.75 wt% (0.83 at.%) Mg and 1.0
wt% (0.96 at.%) Si. Estimate on the basis of the solubility product the equilibrium volume
fraction of Mg2Si at 4000C. Make also a sketch of the Mg2Si solvus in a vertical section
through the ternary Al-Mg-Si phase diagram. Relevant physical data for the Al-Mg-Si sys-
tem are given below:
Solution
The maximum possible volume fraction of Mg2Si precipitated at absolute zero (fmax) can be
estimated from a simple mass balance by considering the stoichiometry of the reaction:
Moreover, the solubility product [at.% Mg]2 [at.% Si] at 4000C (673K) can be obtained
from equation (4-3) by utilising data from Table 4.1:
from which
If we also take into account the stoichiometry of the reaction, the solubility product can be
expressed solely in terms of the Mg-concentration. Substituting
into the above equation gives [at.% Mg] ~ 0.20. The equilibrium volume fraction OfMg2Si at
4000C is thus:
Similarly, the equilibrium Mg2Si solvus can be calculated from the solubility product by
substituting
into equation (4-3). By inserting data from Table 4.1 and rearranging this equation, we get:
It is seen from the graphical representation of the above equation in Fig. 4.4 that the Mg2Si
compound is thermodynamically stable up to about 3000C. At higher temperatures the phase
will start to dissolve until the process is completed at 5600C.
It is obvious from these calculations that the microstructure of overaged Al-Mg-Si alloys
should be very persistent to the heat of welding processes. In practice, only a narrow solutionised
zone forms adjacent to the fusion boundary. However, within this zone significant strength
recovery may occur after welding due to reprecipitation of hardening phases from the
supersaturated solid solution. Consequently, in such weldments the ultimate HAZ strength
level is usually higher than that of the base metal, as illustrated in Fig. 4.5.
at% Mg
at%Mg Si
spherical precipitate is acted on by an external pressure of say 1 atm, the same precipitate is
also subjected to an extra pressure AP due to the curvature of the particle/matrix interface, just
as a soap bubble exerts an extra pressure on its content (see Fig. 4.6(a)). The pressure AP is
given as:8
(4-9)
where 7 is the particle/matrix interfacial energy, and r is the radius of the precipitate.
Because of this extra pressure, the Gibbs energy of a small precipitate will be higher than
that of a large one, which, in turn, increases its solubility (see Fig. 4.6(b)). The important
influence of particle curvature on the solid solubility has been extensively investigated and
reported in the literature.18 Usually, the phenomenon is referred to as the capillary or the
Gibbs-Thompson effect.
In the following we shall assume that the thermodynamic and crystallographic properties
of the metastable precipitates are similar to those of the equilibrium phase and that the re-
duced thermal stability is only associated with capillary effects. For single phase precipitates
in binary alloy systems, it is fairly simple to show that the concentration of solute across
a curved interface, [%A]r, is interrelated to the equilibrium concentration of solute across a
planar interface, [%A], through the following equation:8
(4-10)
where Vn is the molar volume of the precipitate (in m3 mol"1), and Q is the contribution of the
interface curvature to the reaction enthalpy (equal to IyVJr).
(a)
Matrix
Atmospheric
pressure
(b)
Small precipitate
Gibbs energy
Matrix
Large precipitate
[%A] [%A] r
Concentration
Fig. 4.6. Effect of interfacial energy on the solubility of small particles; (a) Schematic representation of
spherical particles embedded in a metal matrix, (b) Integral molar Gibbs energy of matrix and precipi-
tates at a constant temperature.
Assuming that this relationship also holds in the case of binary intermetallics, a combina-
tion of equations (4-5) and (4-10) gives:
(4-11)
where
(4-12)
or
Alternatively, we can express T as a function of the product [%A]rn [%B]rm. This gives the
following expression for the solvus temperature of metastable precipitates T'eq:
(4-13)
It is evident from the graphical representation of equation (4-13) in Fig. 4.7 that the solid
solubility at a given temperature is significantly increased at small particle radii. Taking as an
example 7 = 0.5 J n r 2 , Vm = 10~5 m3 moH, R = 8.314 J Kr1 moH, T = 500 K, we obtain from
equation (4-10):
or
Concentration
Solution
First we estimate the molar volume of the precipitate:
The metastable [3"(Mg2Si) solvus can now be calculated from the solubility product by
substituting
into equation (4-13). By inserting data from Table 4.1 and rearranging this equation, we get:
It is evident from the graphical representation of the above equation in Fig. 4.8 that the
particle curvature has a dramatic effect on the solid solubility. A comparison with Fig. 4.4
shows that the dissolution temperature drops from about 5600C in the case of the equilibrium
Mg2Si phase to approximately 225°C for the metastable |3"(Mg2Si)-phase. On this basis it is
not surprising to find that artificially aged (T6 heat treated) Al-Mg-Si alloys suffer from se-
vere softening in the HAZ after welding, as shown schematically in Fig. 4.9. Moreover, it is
Dissolution temperature: 225 0C
Temperature, 0C
composition A
Nominal alloy
Metastable solvus
boundary
at% Mg
at% Mg 2 Si
HAZ
Fig. 4.9. Response of artificially aged Al-Mg-Si alloys to welding and subsequent heat treatment (sche-
matic).
evident that the characteristic low dissolution temperature of the precipitates also gives rise to
the formation of a heat affected zone which is significantly wider than that observed during
welding of overaged Al-Mg-Si alloys.9 This shows that the response of age-hardenable alu-
minium alloys to welding and thermal processing depends strongly on the initial base metal
temper condition.
With the aid of equation (4-11) it is also possible to calculate an average (apparent) metastable
solvus boundary enthalpy for hardening |3"(Mg2Si)-precipitates in Al-Mg-Si alloys. A closer
evaluation of the exponent gives:
This value is in close agreement with the reported solvus boundary enthalpy for (3"(Mg2Si)-
precipitates in 6082-T6 aluminium alloys. 910
When dispersed particles have some solubility in the matrix in which they are contained, there
is a tendency for the smaller particles to dissolve and for the material in them to precipitate on
larger particles. The driving force is provided by the consequent reduction in the total interfa-
cial energy and ultimately, only a single large particle would exist within the system.
(4-14)
where ro is the initial particle radius, 7 is the particle-matrix interfacial energy, Dm is the ele-
ment diffusivity, Cm is the concentration of solute in the matrix, Vm is the molar volume of the
precipitate per mole of the diffusate, and t is the retention time.
Although the classic Lifshitz-Wagner theory suffers from a number of simplifying assump-
tions, experimental observations usually reveal a cubic growth law of the form given by equa-
tion (4-14).13
(4-15)
where c{ is a kinetic constant, and Qs is the activation energy for the coarsening process (for
binary intermetallics Qs may be taken equal to the activation energy for diffusion of the less
mobile constituent atom of the precipitates in the matrix).
(4-16)
The integral on the right-hand side of equation (4-16) represents the kinetic strength of the
thermal cycle with respect to particle coarsening, and can be determined by means of numer-
ical methods when the weld thermal (T-t) programme is known. The resulting radius of the
precipitates may then be evaluated from equation (4-16) by inserting representative values
for the constants ro and C1 (e.g. obtained from quantitative particle measurements).
Example (4.4)
Consider stringer bead deposition (GMAW) on a thick plate of a Ti-microalloyed steel under
the following conditions:
Assume that the base metal contains a fine dispersion of TiN precipitates in the as-received
condition. Calculate on the basis of equation (4-16) and the Rosenthal thick plate solution
(equation (1-45)) the extent of particle coarsening occurring within the fully transformed heat
affected zone during welding. Relevant physical data for titanium-microalloyed steels are
given below:
Solution
In the present example the problem is to calculate the size of the TiN precipitates in different
positions from the fusion boundary. This requires detailed information about the weld thermal
programme, as shown in Fig. 4.10(a). By substituting the appropriate values for qo, X, a and v
into the Rosenthal thick plate solution, the governing heat flow equation becomes:
where /?* refers to the three-dimensional radius vector in the moving coordinate system (desig-
nated R in equation (1-45)), while x is the welding direction (equal to vt at pseudo-steady
state).
Since titanium nitride is thermodynamically stable up to the melting point of the steel,
equation (4-16) can be used to calculate the extent of particle coarsening occurring within the
transformed parts of the HAZ. In the present example, we may write:
During welding, the thermal pulse experienced by the heat affected zone adjacent to the fusion
boundary can result in complete dissolution of the base metal precipitates. Since this may give
rise to subsequent strength loss and grain growth, it is important to understand how variations
in welding parameters and operational conditions affect the dissolution rate. In the following,
the kinetics of particle dissolution will be discussed from a more fundamental point of view.
Weld metal
HAZ
Temperature
Y~ regime
Time
(b)
(i/r)exp(-Qs/RT)
Time
Fig. 4.10. Kinetic strength of weld thermal cycle with respect to particle coarsening (Example (4.4)); (a)
HAZ temperature-time programme (schematic), (b) Numerical integration procedure (schematic).
Fully transformed HAZ •
Weld metal
Peak temperature, 0C
Particle radius, nm
Fig. 4.12. Measured size distribution of TiN before (broken lines) and after (full lines) weld thermal
simulation. Operational conditions as in Example (4.4). Data from Ion et al.14
ing precipitates. Nevertheless, it will be shown below that at least some of them are suffi-
ciently accurate to capture the essential physics of the problem and to give valuable quantita-
tive information on the extent of particle dissolution occurring during the weld thermal cycle.
Annealing temperature: 1350 0C
Particle radius, nm
Steel A
Steel B
Steel C
Annealing time, s
Fig. 4.13. Effects of annealing time and steel chemical composition on the mean particle size of TiN.
Data from Matsuda and Okumura.15
(4-17)
where r is the particle radius, a is the dimensionless supersaturation (defined in Fig. 4.14), and
Dm is the element bulk diffusivity.
The term Hr on the right-hand side of equation (4-17) stems from the steady-state part of
the diffusion field, while the (1 A/7) term arises from the transient part. Because of the com-
plex form of equation (4-17) it cannot be integrated analytically and hence, numerical
methods must be applied. However, if the transient part of equation (4-17) is neglected (con-
forming to the solution after long times), it is possible to obtain a simple expression for the
particle radius as a function of time:
(4-18)
Distance
Fig. 4J4. Schematic representation of the concentration profile around a dissolving spherical particle in
an infinite matrix.
ently by Aaron et al.16 and is valid after a certain period of time, provided that there is no
impingement of diffusion fields from neighbouring precipitates. As shown in Fig. 4.15, this
simplified solution gives a reasonable description of the dissolution kinetics of small spherical
precipitates in steel during reheating above the AC1 -temperature.
Following the treatment of Agren,18 the time required for complete dissolution of a spher-
ical precipitate td can be obtained from equation (4-18) by setting r = 0:
(4-19)
Moreover, the volume fraction of the precipitates/as a function of time is given by:
(4-20)
Example (4.5)
The following example illustrates the direct application of equations (4-18) and (4-19). Con-
sider a niobium-microalloyed steel which contains a fine dispersion of NbC precipitates. Pro-
vided that impingement of diffusion fields from neighbouring particles can be neglected, cal-
culate the total time required for complete dissolution of a 100 nm large NbC precipitate at
Particle radius, jam
Simplified analytical
solution (Whelan)
Time, s
Fig. 4.15. Dissolution kinetics of spherical cementite particles in austenite at 8500C. Data from Agren.18
135O°C. Data for the steel chemical composition and the diffusivity of Nb in austenite at
13500C are given below:
Solution
In the present example it is reasonable to assume that the dissolution rate of the precipitate is
controlled by diffusion of Nb in austenite. For a single NbC precipitate embedded in a Nb-
depleted matrix, the dimensionless supersaturation becomes:
This gives:
The dissolution time td can now be calculated from equation (4-19) by inserting the appro-
priate values for ro, aNt)J and Dm\
A comparison with Fig. 4.16 shows that the predicted value is off by a factor of about 4
compared with that obtained from more sophisticated numerical calculations. This degree of
accuracy is acceptable and justifies the use of equation (4-18) for prediction of the dissolution
rate of spherical precipitates under different thermal conditions provided that the model is
calibrated against experimental data points.
4.4.1.2 Application to continuous heating and cooling
Application of the model to continuous heating and cooling requires numerical integration of
equation (4-18) over the weld thermal cycle:
(4-21)
Under such conditions the volume fraction of the precipitates is given by:
(4-22)
Equations (4-21) and (4-22) provide a basis for predicting the extent of particle dissolution
occurring within the HAZ during welding in the absence of impingement of diffusion fields
from neighbouring precipitates.
Dissolution time, s
Particle diameter, nm
Fig. 4.16. The dissolution time of NbC in austenite at 13500C as function of initial particle diameter lro
for different Nb and C levels (numerical solution). Data from Suzuki et al.6
Example (4.6)
Consider stringer bead deposition (SAW) on a thick plate of a Nb-microalloyed steel (0.10
wt% C - 0.03 wt% Nb) under the following conditions:
Assume that the base metal contains a fine dispersion of NbC precipitates in the as-received
condition. Calculate on the basis of equation (4-22) and the Rosenthal thick plate solution
(equation (1-45)) the extent of particle dissolution occurring within the fully transformed HAZ
during welding. Relevant physical data for Nb-microalloyed steels are given below:
Solution
In the present example the problem is to calculate the variation in the//fo ratio across the fully
transformed HAZ. By substituting the appropriate values for qo, X, a, and v into the Rosenthal
thick plate solution, the governing heat flow equation becomes:
Since it is reasonable to assume that the dissolution rate of the precipitates is controlled by
diffusion of Nb in austenite, the dimensionless supersaturation reduces to:
This gives:
By substituting the appropriate expressions for aNb and DNb into equation (4-22), we ob-
tain:
Here the lower and upper integration limits refer to the total time spent in the thermal cycle
from Ac3 to T and down again to Ac3.
The extent of particle dissolution occurring within the HAZ during welding can now be
calculated in an iterative manner by numerical integration of the above equation over the weld
thermal cycle. The results from such computations are presented graphically in Fig. 4.17.
It is evident from these data that NbC starts to dissolve when the peak temperature of the
thermal cycle T exceeds the equilibrium dissolution temperature Td of the precipitate. The
process is completed when T approaches 13300C, conforming to a temperature interval of
1900C. This shows that considerable superheating is required in order to overcome the inher-
ent kinetic barrier against particle dissolution under the prevailing circumstances.
No dissolution-
f/fo
Complete dissolution
Peak temperature, 0C
(4-23)
(a)
(b)
Concentration
Distance
Fig. 4.18. Numerical model for dissolution of rod-shaped particles in a finite, depleted matrix; (a) Disso-
lution cell geometry, (b) Particle/matrix concentration profile (moving boundary).
For a specific alloy, the ratio between ro and L (the mean interparticle spacing) can be
calculated from a simple mass balance, assuming that all solute is tied-up in precipitates. Tak-
ing this ratio equal to 0.06 for rod-shaped precipitates in diluted alloys,9 the kinetics of particle
dissolution during isothermal heat treatment have been examined for a wide range of opera-
tional conditions. These results are presented in a general form in Fig. 4.19 by the use of the
following groups of dimensionless parameters:
Dimensionless time (4-24)
where c2 is a kinetic constant, and n{ is a time exponent (assumed constant and equal to 0.5
under the prevailing circumstances).
The rate of particle dissolution will gradually decline with increasing values of T as a result
of impingement of diffusion fields from neighbouring precipitates which reduces a. In prac-
tice, this is seen as a continuous decrease in the slope of the flfo-x curves in Fig. 4.19 (nx < 0.5).
In such cases equation (4-26) will only be valid within small increments of X.
log <1 - f/fo)
log T
Fig. 4.19. Dissolution kinetics of rod-shaped particles in a finite, depleted matrix. Data from Myhr and
Grong.9
4.4.2.2 Generic model
Myhr and Grong9 have shown how this model can be applied to specific alloy systems.
From equation (4-26) we have:
(4-27)
(4-28)
For isothermal heat treatment at a chosen reference temperature (Tn), the rate of particle
dissolution is determined by the retention time tr{. Let f* denote the maximum hold time
required for complete dissolution of the precipitates. It follows that equation (4-28) can be
written in a general form by normalising tn with respect to t*x. The parameter t*n is obtained
by setting flfo = 0:
(4-29)
(4-30)
By inserting the approximate expressions for C1 and Dm into equation (4-30) (see previous
examples), and rearranging equation (4-28), we obtain:
(4-31)
and
(4-32)
where Q'apP. is the apparent (metastable) solvus boundary enthalpy (defined in Section 4.2.3.2),
and Qs is the activation energy for diffusion of the less mobile constitutive atom of the precipi-
tates.
Equations (4-31) and (4-32) exploit some good modelling techniques. For example, the use
of a dimensionless time eliminates an unknown kinetic constant which premultiplies t and ^1*
in the derivation of equation (4-32). Moreover, by raising the dimensionless time to a power
H1 means that the premultiplying constant, here unity, is independent of the value of nv and is
itself also dimensionless. Finally, the form of equation (4-31) eliminates further unknown
kinetic constants, and may readily be calibrated using an experimental time r* at a reference
temperature.
Figure 4.20 shows the variation inflfo with time (on log axes), from a range of isothermal
experiments carried out on 6082-T6 aluminium alloys, using hardness (or electrical conduc-
tivity) measurements to evaluate/7/\ The curve (equation (4-32)) extrapolates back to a slope
of 0.5 (the exponent n{) for the case of the early stages of dissolution before impingement of
the diffusion fields. The exponent nx is seen to fall to lower values when the proportion dis-
solved is higher, in agreement with the theoretical curves in Fig. 4.19.
log (t/t*)
(4-34)
and
(4-35)
Since dfldlx and t\ are unique functions of/and T, respectively, the additivity condition is
satisfied. Consequently, when the temperature varies with time, we replace the term 11 t*x in
equation (4-34) by dt I t*x and integrate over the thermal cycle, giving:
(4-36)
This integral is called the kinetic strength of the thermal cycle with respect to reversion.
The resulting volume fraction of the precipitates following a heating cycle is then found by
evaluating the integral Z1 numerically (e.g. by utilising input data from Table 4.2) and replac-
ing 111* with Z1 in equation (4-32), yielding a value for flfQ from the master curve of Fig. 4.20.
Case Study (4.1)
By utilising equation (4-36) and the general heat flow model for welding on medium thick
plates (i.e. equation (1-104)), it is possible to calculate the variation in the f/fo ratio (i.e. the
solute distribution) across the HAZ of single pass 6082-T6 aluminium weldments for a wide
range of operational conditions (see Table 4.3). The results from such computations are pre-
sented graphically in Fig. 4.21 -4.23.
When stringer bead deposition is carried out on a plate of medium thickness, the solute
distribution in the transverse y direction is expected to vary with distance from the plate sur-
face due to a continuous change in the heat flow conditions. A closer inspection of Figs. 4.21
and 4.22 shows that this is correctly accounted for in the present model. In contrast, a full
penetration butt weld will always reveal a similar solute distribution in the transverse direction
of the weld, as shown in Fig. 4.23. This situation arises from the lack of a temperature gradient
in the through-thickness z direction of the plate.
Table 4.2 Basic input data in dissolution model for hardening p" (Mg2Si)- precipitates in 6082-T6
aluminium alloys. Data from Myhr and Grong.9
Parameter Q'app Qs nx t\
(starting value) (375°C)
Value 3OkJm0I-1 13OkJm 0 I- 1 0.5 600 s
(a)
Peak temperature, 0C
Wo
Comolete dissolution
Aym ,mm
(b)
Peak temperature, 0C
f/fo
Complete dissolution
ym,mm
Fig. 4.21. Dissolution of p"(Mg2Si)-precipitates during aluminium welding (Weld 1); (a) Upper plate
surface, (b) Lower plate surface. Operational conditions as in Table 4.3.
Table 4.3 Operational conditions used in aluminium welding experiments (Case study 4.1).
Peak temperature, 0C
Mo
Complete dissolution
Aym,mm
(b)
Peak temperature, 0C
f/fo
Complete dissolution
ym,mm
Fig. 4.22. Dissolution of (3"(Mg2Si)-PrCCiPiIaIeS during aluminium welding (Weld 2); (a) Upper plate
surface, (b) Centre of plate. Operational conditions as in Table 4.3.
Complete dissolution
Ay ,mm
m
Fig. 4.23. Dissolution of p"(Mg2Si)-precipitates during aluminium welding (Weld 3). Operational con-
ditions as in Table 4.3.
Figure 4.24 shows plots of the variation in the flfo ratio across the HAZ of 6082-T6 alu-
minium weldments for different values of qo Ivd. It follows that a narrow width of the dissolu-
tion zone requires the use of a low energy per mm2 of the weld. In practice, this can be
achieved by the use of an efficient welding process (e.g. electron beam or laser welding) which
facilitates deposition off a full penetration butt weld without employing a groove preparation
(i.e. eliminates the need for filler metals).
f/fo
Scale:
Fig. 4.24. Process diagram showing the solute distribution within the HAZ of single-pass 6082-T6 alu-
minium butt welds for different values of qo Ivd.
References
1. R.D. Doherty: Physical Metallurgy, 3rd Edn (Eds R.W. Chan and R Haasen), 1983, Amster-
dam, North-Holland Physics Publ., 934-1030.
2. K.E. Easterling: Introduction to the Physical Metallurgy of Welding, 1983, London, Butterworths
& Co., Ltd.
3. H. Adrian and RB. Pickering: Mater. ScL TechnoL, 1991, 7, 176-182.
4. B. Loberg, A. Nordgren, J. Strid and K.E. Easterling: MetalL Trans., 1984,15A, 33-41.
5. J. Strid and K.E. Easterling: Acta MetalL, 1985, 33, 2057-2074.
6. S. Suzuki, G.C. Weatherly and D.C. Houghton: Acta. MetalL, 1987, 35, 341-352.
7. J.L. Petty-Galis and R.D. Goolsby: J. Mater. ScL, 1989, 24, 1439-1446.
8. D. A. Porter and K.E. Easterling: Phase Transformations in Metals and Alloys, 1981, Wokingham
(England), Van Nostrand Reinhold Co. Ltd.
9. O.R. Myhr and 0. Grong: Acta MetalL Mater., 1991, 39, 2693-2702; ibid., 2703-2708.
10. H.R. Shercliff and M.F. Ashby: Acta MetalL Mater., 1990, 38, 1789-1802; ibid., 1803-1812.
11. J.M. Lifshitz and V.V. Slyozov: /. Phys. Chem. Solids, 1961,19, 35-50.
12. C. Wagner: Z. Electrochem., 1961, 65, 581-591.
13. L.C. Brown: Acta MetalL Mater., 1992, 40, 1293-1303.
14. J.C. Ion, K.E. Easterling and M.F. Ashby: Acta MetalL, 1984, 32, 1949-1962.
15. S. Matsuda and N. Okumura: Trans. ISIJ, 1978,18, 198-205.
16. H.B. Aaron, D. Fainstein and G.R. Kotler: J. Appl. Phys., 1970, 41, 4404-4410.
17. MJ. Whelan: Metal ScL J., 1969, 3, 95-97.
18. J. Agren: Scand. J. MetalL, 1990,19, 2-8.
19. R.A. Tanzilli and R.W. Heckel: Trans. Met. Soc. AIME, 1968, 242, 2313-2321.
20. H.B. Aaron and R. Kotler: MetalL Trans., 1971, 2, 393-408.
21. R. Asthana and S.K. Pabi: Mat. ScL Eng., 1990, A128, 253-258.
22. U.H. Tundal and N. Ryum: MetalL Trans., 1992, 23A, 433-444; ibid., 445-449.
23. H.R. Shercliff, 0. Grong, O.R. Myhr and M.F. Ashby: Proc. 3rd Int. Conf. on Aluminium
Alloys — Their Physical and Mechanical Properties, Trondheim, Norway, June 1992, Vol. Ill,
pp. 357-369, The University of Trondheim, The Norwegian Inst. of Technol.
24. J.W. Christian: Phase Transformations in Metals and Alloys, 1975, Oxford, Pergamon Press.
Appendix 4.1
Nomenclature
arbitrary function of X
universal gas constant artificially aged condition
(8.314JK-1InOl-1)
voltage (V)
universal gas constant multi-
plied by InIO (19.14J Kr1 welding speed (mm s"1)
mol"1)
welding direction (mm)
three-dimensional radius
vector in Rosenthal equation state variable
(mm)
transverse direction (mm)
standard entropy of reaction width of HAZ referred to fu-
(J K mol"1) sion boundary (mm)
5.1 Introduction
Grain growth is an important aspect of welding metallurgy. Normal grain growth in metals
and alloys is a thermally activated process driven by the reduction in the grain boundary en-
ergy. Physically, it occurs by growth of the larger grains at the expense of the smaller ones
which tend to shrink.
Under isothermal heat treatment conditions, normal grain growth is well described by the
following empirical equation:1
(5-1)
where D is the mean grain size (diameter), D0 is the initial grain size, n is the time exponent,
/ is the isothermal annealing time, Qapp. is the apparent activation energy for grain growth, and
C1 is a kinetic constant. The other symbols have their usual meaning.
For most metals and alloys the time exponent n in equation (5-1) varies typically in the
range from 0.1 to 0.4, as shown in Fig. 5.1. Only in the case of ultrapure metals annealed at
very high temperatures the time exponent may approach a constant value of 0.5. This corre-
sponds to the limiting case where the grain boundary migration rate is directly proportional to
the driving pressure 7 /D (7 denotes the grain boundary interfacial energy per unit area, while
D is the grain size).
It is well recognised that alloying and impurity elements both in the dissolved state and in
the form of inclusions or second phase particles will retard grain growth.2"* Consequently, a
comprehensive theoretical treatment of grain growth in welds must include a consideration of
such effects. The present analysis will therefore start with a closer examination of factors
affecting the grain boundary mobility in metals and alloys under conditions applicable to weld-
ing.
The symbols and units used throughout this chapter are defined in Appendix 5.1.
•Brass
Z.R.AI
measurements. Different parameters are used to describe the size and the shape of individual
grains. In three dimensions, individual grain volumes cannot be determined directly from
measurements made in single cross sections through the structure. Therefore, certain geomet-
ric assumptions must be employed to obtain these quantities.
Since most grain size measurements seek to correlate the interaction of grain boundaries
with specific properties (e.g. the transformation behaviour), an estimate of the grain boundary
area per unit volume Sv is often required. This parameter can be calculated without assump-
tions concerning grain shape and size distribution from measurements of the mean linear grain
intercept D*.5
(5-2)
If the mean grain diameter D is required from 5V, this may be obtained by assuming a
spherical grain shape. Noting that each boundary is shared by two adjacent grains, we obtain:
(5-3)
from which
(5-4)
It follows that the mean linear grain intercept D* is always smaller than the actual grain size
D.
A common observation in metals and alloys is that the size distributions of grain aggregates
at different annealing times become equivalent when the measured grain size parameter, D, is
normalised (scaled) by the time-dependent average of this metric, D (see Fig. 5.2). This
means that grain structures can be completely characterised, in a statistical sense, by simple
probability functions of the standard deviation of the distribution together with the time de-
pendence of the average size scale D.
(5-5)
where y is the grain boundary interfacial energy, and p* is the radius of curvature.
It is conventional practice to replace p* in equation (5-5) with some measure of the average
grain size, such as the mean linear grain intercept D*, or with the diameter of the equivalent
spherical volume of some geometrically modelled average grain size. Experimental measure-
ments performed on high purity aluminium indicate that p* ~ 3.23 D*.7 This observation is
consistent with the model of Hellman and Hillert,8 which predicts that the curvature of the
most critical element of grain boundary that must be stabilised is p* = 3 D . Under such condi-
tions, the driving pressure becomes:
(5-6)
Rayleigh distribution
Log-normal
distribution
D/D
Fig. 5.2. Comparison of measured grain size data in iron with the Rayleigh and log-normal distribution
functions F. The similarity of the size distributions at different annealing times illustrates the self-similar
scaling behaviour of normal grain growth. After Pande.6
In practice, the numerical constant in equation (5-6) can vary by, at least, a factor of three,
depending on the assumptions of the models. Consequently, in the general case the average
driving pressure is given by:
(5-7)
(i) A low velocity limit, where the rate of grain boundary migration is controlled by diffu-
sion of impurity atoms perpendicular to the boundary.
(ii) A high velocity limit, where the grain boundary migration process is mainly governed
by the diffusion of solvent atoms across the boundary (i.e. controlled by the rate of boundary
self diffusion).
The low velocity limit is associated with either a low driving pressure PG or a high impurity
level C0, and is characterised by a linear type of relationship between the grain boundary
migration rate Va and PG:2
(5-8)
Here e denotes the intrinsic drag coefficient, while 1F is a parameter depending on the
diffusivity and the interaction energy between the grain boundary and the impurity atoms.
For the other extreme (i.e. the high velocity limit), the grain boundary migration rate Vb is
described by the relationship:2
(5-9)
where 4Vp2 denotes another complex function of the impurity diffusivity and the interaction
energy between the grain boundary and the impurity atoms.
In the case of spherical grains, the classic grain growth equation predicts that the grain
boundary migration rate V is a power function of the driving pressure (y /D). This is readily
seen by differentiating equation (5-1) and inserting 7 (the grain boundary interfacial energy)
into the resulting expression:
(5-10)
where c3 is a new kinetic constant.
Note that at very high driving pressures or low impurity levels, equation (5-9) approaches
the limiting case where the migration rate Vb becomes directly proportional to PG, correspond-
ing to a time exponent n = 0.5 in the grain growth equation. At this point the grain boundary
will break away from its surrounding impurity atmosphere and migrate at a rate close to the
rate of boundary self diffusion. In most cases, however, the relationship between V and PG
derived from equation (5-10) will be different from the theoretical one due to the empirical
nature of the grain growth equation.
For high driving pressures and intermediate impurity concentrations, the classic impurity
drag theories predict a discontinuous transition from the high to the low velocity limit. It has,
however, been argued by Vandermeer9 that the observed transition may be considered as con-
tinuous. Thus, in a log V vs log PG plot it would appear as a steep curve connecting the lines
for the high and low velocity extremes together, as illustrated schematically in Fig. 5.3.
(5-11)
Transition region
log PG
Fig. 5.3. Schematic variation of grain boundary migration rate V with driving pressure PQ according to
the classic impurity drag theories of Cahn2 and Liicke and Stiiwe.3 The diagram is based on the ideas of
Vandermeer.9
If the number of interacting particles per unit area of the grain boundary is taken equal to na,
the resulting retardation pressure becomes:
(5-12)
Assuming that only one half of the particles which touches the grain boundary will interact
with a maximum force, na is related to Nv (the number of particles per unit volume) through the
following equation:410
(5-13)
Given that the particles are spherical and of uniform size, Nv can be expressed as:
(5-14)
where/is the particle volume fraction.
A combination of equations (5-12), (5-13), and (5-14) leads to the well-known expression
for the so-called Zener drag (or Zener retardation pressure):410
(5-15)
In practice, the numerical constant in equation (5-15) can vary by, at least, a factor of five,
depending on the assumptions of the models.11 Consequently, in the general case the Zener
retardation pressure is given by:
(5-16)
(5-17)
It follows from equation (5-18) that the grain boundary migration rate V becomes inversely
proportional to the average grain size D when n = 0.5 a n d / = 0. This corresponds to the
limiting case where the grain boundary will break away from its surrounding impurity atmos-
phere and migrate at a rate which is controlled by the diffusion of solvent atoms across the
boundary. In most cases, however, the observed relationship between V and APG will be
different from the theoretical one due to drag from second phase particles (f> 0) or impurity
elements in solid solution (n < 0.5).
(5-19)
(5-20)
From this it is seen that the parameters M0 and k are true physical constants which are
related to the grain boundary mobility and the pinning efficiency of the precipitates, respec-
tively.
(5-21)
The parameter k (which in the following is referred to as the Zener coefficient) is defined as
the ratio between the numerical constants in equations (5-7) and (5-16), respectively. In the
original Zener's model k = 4/3, while other investigators have arrived at different results.811"14
As shown in Fig. 5.4, the limiting grain size may vary by over one order of magnitude, depend-
ing upon the assumptions of the models. This makes it difficult to apply equations (5-20) and
(5-21) for quantitative grain size analyses without further background information on the Zener
coefficient.
Diim.nm
Gladman
r/f, um
Fig. 5.4. Relation between limiting grain size Dum., particle radius r, and volume fraction/predicted by
different models.
Example (5. J)
Consider multipass GMA welding on a thick steel plate under the following conditions:
Based on the models of Zener,10 Hellman and Hillert,8 and Gladman13 estimate the limiting
austenite grain size Dlim in the transformed parts of the weld HAZ when the oxygen and
sulphur contents of the as-deposited weld metal are 0.04 and 0.01 wt%, respectively.
Solution
As shown in Chapter 2 of this textbook the volume fraction of oxide and sulphide inclusions
can be calculated from equation (2-75):
Similarly, the average radius of the grain boundary pinning inclusions can be obtained from
equation (2-79):
This gives the following values for the limiting austenite grain size:
Zener:
Hellman and Hillert:
Gladman:
and
As expected, the limiting austenite grain size is seen to vary by more than one order of
magnitude, depending on the assumptions of the models. In practice, the Zener coefficient in
low-alloy steel weld metals falls within the range from 0.32 to 0.93, as shown in Fig. 5.5. The
average value of A: is close to 0.52, which is the same as that inferred from the Gladman model
(upper limit). When it comes to intermetallic compounds such as titanium nitride, the Zener
coefficient varies typically between 0.75 and 0.25 during grain growth in the austenite re-
gime. 1617 This suggests that k ~ 0.50 is a reasonable estimate of the grain boundary pinning
efficiency of oxides and nitrides in steel.
(5-22)
(a)
SA steel weld metal
Austenite grain size, Jim
Annealing temperature, 0 C
(b)
GMA and SA steel
weld metals
D||m. Hm
r/f, urn
Fig. 5.5. Evaluation of the Zener coefficient in steel weld metals containing stable oxide and sulphide
inclusions; (a) Determination of Dum. from isothermal grain growth data (holding time: 30 min),
(b) Variation in Dum. with the inclusion rlf ratio. Data from Skaland and Grong.15
(b)
Steel A
Steel B
LogD7
Fig. 5.6. Evaluation of the time exponent n and the activation energy Q for austenite grain growth in
steel under thermal conditions applicable to welding; (a) Time exponent n, (b) Activation energy Qapp.
Data from Akselsen et ah18
(5-22) can be solved explicitly for different values of DUm, n, and Z1. The results may then be
presented in the form of novel diagrams which show the competition between the various
processes that lead to grain growth during heat treatment of metals and alloys.
A more thorough documentation of the predictive power of the model and its applicability
to welding is given in Section 5.4.
(5-24)
Referring to Fig. 5.7, the average grain size D becomes a simple cube root function of Z1
when n = 0.5 and D 0 = 0. In other situations (n < 0.5), the grains will coarsen at a slower rate
due to drag from alloying and impurity elements in solid solution. This is seen as a general
reduction in the slope of the D-Ix curves in Fig. 5.7.
The important austenite grain growth inhibiting effect of phosphorus and free nitrogen in
steel following particle dissolution is shown in Fig. 5.8.
(5-25)
D, ^m
I1W"
Fig. 5.7. Predicted variation in average grain size D with /, and n f o r / = 0 and D0 = 0 ('free' grain
growth).
(a)
Steel A
D Y ,fim
Number of cycles
(b)
Steel B
D y ,um
Number of cycles -
Fig. 5.8. Illustration of the austenite grain growth inhibiting effect of phosphorus and free nitrogen in
low-alloy steel during reheating above the Ac^ temperature (multi-cycle weld thermal simulation);
(a) Steel A (50ppm P, 20ppm N), (b) Steel B (180ppm P, 80ppm N). Data from Akselsen et a/.18
from which the average grain size D is readily obtained. In other cases, numerical methods
must be employed to evaluate D.
It is evident from the graphical representation of equation (5-25) in Fig. 5.9 that the grain
growth inhibiting effect of the precipitates is very small during the initial stage of the process
when D « D lim. Under such conditions the grains will coarsen at a rate which is comparable
with that observed for free grain growth (n = 0.5,/= 0). The grain coarsening process becomes
gradually retarded as the average grain size increases because of the associated reduction in
the effective driving pressure APG until it comes to a complete stop when AP0 = 0 (i.e. D =
D Hm)-
D, jim
I 1 ^m 2
Fig. 5.9. Predicted variation in average grain size D with Z1 and Dnm. for n = 0.5 and D0 = 0 (stable
precipitates).
D, (im
I 1 , [nm]1/n
Fig. 5.10. Predicted variation in average grain size D with Z1 and n for Dum. = 250|Jin and D0 = 0
(stable precipitates). Dotted curves correspond to grain growth in the absence of pinning precipitates.
If grain growth at the same time occurs under the action of a constant drag from impurity
elements in solid solution, the situation becomes more complex. As shown in Fig. 5.10, a
decrease in the time exponent from say 0.5 to 0.2 gives rise to a marked reduction in the slope
of the D-I 1 curves, similar to that observed in Fig. 5.7 for particle-free systems (/= 0). How-
ever, the predicted grain coarsening rate is lower than that evaluated from equation (5-24) due
to the extra drag exerted by the grain boundary pinning precipitates. This leads ultimately to a
stabilisation of the microstructure when D = DUm.
(5-26)
where Qs is the activation energy for the coarsening process, C5 is a kinetic constant, and I2 is
the kinetic strength of the thermal cycle with respect to particle coarsening. The other symbols
have their usual meaning.
If the base metal contains particles of an initial radius ro and volume fraction/^, the limiting
grain size at I2 = 0 (D° lim) can be defined as:
(5-27)
from which
(5-28)
(5-29)
By combining equations (5-26), (5-28), and (5-29), we arrive at the following relationship
between (D Um ) and I2:
(5-30)
It is seen from equation (5-30) that the limiting grain size in the presence of growing
particles depends on the product (k/fo)3I1. In practice, the grain boundary pinning effect of the
precipitates is determined by the relative rates of particle coarsening and grain growth in the
material, i.e. whether the grain boundary mobility is sufficiently high to keep pace with the
increase in DUm during heat treatment. Generally, the pinning conditions are defined by the
(k/fo)3 I1IIx ratio, which after substitution and rearranging yields:
(5-31)
In cases where the parameters c5 ,Qs, M0*, and Qapp are known, the average grain size D
can readily be evaluated from equations (5-22), (5-30), and (5-31) by utilising an appropriate
integration procedure. However, since Qs normally differs from Qapp^ the (klfo)3I1I Ix ratio
will depend on the thermal path during continuous heating and cooling. Consequently, solu-
tion of these coupled equations generally requires stepwise integration in temperature-time
space via a fourth heat flow equation. This problem will be dealt with in Section 5.4.
The situation becomes much simpler if heat treatment is carried out isothermally. Under
such conditions the product (k/fo)311 will only differ from Ix by a proportionality constant m,
which is characteristic of the system under consideration. Accordingly, equation (5-30) can be
rewritten as:
(5-32)
From this we see that the two coupled equations (5-22) and (5-32) can be solved explicitly
for different values of D°nm., n, m, and Z1. Hence, it is possible to present the results in the form
of novel 'mechanism maps' which show the competition between particle coarsening and
grain growth during isothermal heat treatment for a wide range of operational conditions.
Examples of such diagrams are given in Figs. 5.11 and 5.12.
It is evident from these figures that the grain coarsening behaviour during isothermal heat
treatment is very sensitive to variations in the proportionality constant m. For large values of
m, the matrix grains will coarsen at a rate which is comparable with that observed in Fig. 5.7
for particle-free systems (f = 0). This corresponds to a situation where the grain boundary
pinning precipitates will completely outgrow the matrix grains. It is interesting to note that
particle outgrowing is more likely to occur if the time exponent n is small, as shown in Fig.
5.12, because of the associated reduction in the grain boundary mobility in the presence of
impurity elements in solid solution. In other systems, where the proportionality constant m is
closer to unity, the reduced coarsening rate of the precipitates gives rise to a higher Zener
retardation pressure and ultimately to a stagnation in the matrix grain growth. In the limiting
case, when m = 0, the grain growth behaviour becomes idential to that observed in Figs. 5.9
and 5.10 for stable precipitates.
Time exponent n = 0.5
D.jim
I 1 ,^m 2
Fig. 5.11. Predicted variation in average grain size D with Ix and m for D°um. = 50jLim, n = 0.5, and
D0 =0 (growing precipitates).
I1^m1'"
Fig. 5.12. Predicted variation in average grain size D with I1 and m for D°um. - 50|im, n = 0.3, and
D0 = 0 (growing precipitates).
Example (5.2)
Consider a titanium-microalloyed steel with the following chemical composition:
Assume that the base metal contains an uniform dispersion of TiN precipitates in the as-
received condition, conforming to a limiting austenite grain size T>°um. of 50 |iim. Provided that
boundary drag from impurity elements in solid solution can be neglected (i.e. n ~ 0.5), esti-
mate on the basis of Fig. 5.11 the average austenite grain size D1 in the material after 25 s of
isothermal annealing at 13000C. Relevant physical data for titanium-microalloyed steels are
given below:
Solution
The initial volume fraction of TiN in the material can be estimated from simple stoichiometric
calculations by considering the difference between total and soluble titanium. Taking the
atomic weight of Ti and N equal to 47.9 and 14.0 g mol"1, respectively, we obtain:
From this we see that the initial radius of the TiN precipitates in the base metal is close to:
Since heat treatment is carried out under isothermal conditions, the parameters m and Z1 can
be obtained directly from equations (5-31) and (5-22) without performing a numerical integra-
tion:
Similarly, in the case of Z1 we get:
The average austenite grain size can now be read from Fig. 5.11 by linear interpolation
between the curves for m = 10 and 100 Jim. This gives:
Although experimental data are not available for a direct comparison, the predicted grain
size is of the expected order of magnitude. From this it is obvious that considerable austenite
grain growth may occur in titanium-microalloyed steels because of particle coarsening, in
spite of the fact that TiN, from a thermodynamic standpoint, is stable up to the melting point of
the steel. The process can, to some extent, be counteracted by the use of a finer dispersion of
TiN precipitates in the material. For example, if the initial particle radius is reduced by a factor
of five (conforming to a change in I W from 50 to 10 Jim), the austenite grain size of the
annealed material decreases from 75 to 65 jLim, as shown in Fig. 5.13. Nevertheless, since
particle coarsening is a physical phenomenon occurring during high temperature heat treat-
ment of metals and alloys, austenite grain growth cannot be avoided. This explains why, for
instance, conventional titanium-microalloyed steels are not suitable for high heat input weld-
ing due to their tendency to form brittle zones of Widmanstatten ferrite and upper bainite in the
coarse grained HAZ region adjacent to the fusion boundary.22
Stable particles
I1-Hm2
Fig. 5.13. Predicted variation in average grain size D with Z1 and m for D°Hm. = 10 um, n = 0.5, and
Do = 0 (growing precipitates).
5.3.3.5 Grain growth in the presence of dissolving precipitates
Little information is available in the literature on the matrix grain growth behaviour of metals
and alloys in the presence of dissolving precipitates. As shown in Chapter 4, the model of
Whelan23 provides a basis for calculating the dissolution rate of single precipitates embedded
in an infinite matrix. If the transient part of the diffusion field is neglected, the variation in the
particle radius r with time t at a constant temperature is given by equation (4-18):
(5-33)
where a is the dimensionless supersaturation (defined in Fig. 4.14), and Dm is the element
diffusivity.
Application of the model to continuous heating and cooling requires numerical integration
of equation (5-33) over the weld thermal cycle:
(5-34)
where I3 is the kinetic strength of the thermal cycle with respect to particle dissolution.
From this relation the following expression for the particle volume fraction can be derived
(see equation (4-22), Chapter 4):
(5-35)
(5-36)
It is seen from equation (5-36) that the limiting grain size increases from D°um. at I3 = 0 to
infinite when I3 = (fo Ik)2 (D°um. )2 • Since the magnitude of the Zener drag, in practice, depends
on the relative rates of grain growth and particle dissolution in the material, the pinning condi-
tions are defined by the (klfo)2131 Ix ratio:
(5-37)
Equation (5-37) shows that the (k/fo)2I3111 ratio is contingent upon the thermal path during
continuous heating and cooling. Consequently, application of the model to welding generally
requires numerical integration of the coupled equations (5-22), (5-36), and (5-37) over the
weld thermal cycle.
However, the integration procedure is largely simplified if heat treatment is carried out
isothermally. In such cases the product (k/fo)213 will only differ from Ix by a proportionality
constant m*, which is characteristic of the system under consideration. By substituting m*Ix
into equation (5-36), we obtain:
(5-38)
From this we see thaUhe two coupled equations (5-22) and (5-38) can be solved explicitly
for different values of Dun., n,m*, and I1. Hence, it is possible to present the results in the
form of novel 'mechanism maps' which show the competition between particle dissolution
and grain growth during isothermal heat treatment for a wide range of operational conditions.
Examples of such diagrams are given in Figs. 5.14 and 5.15.
As expected, the stability of the second phase particles is sensitive to variations in the
proportionality constant m*. Normally, the precipitates will exert a drag on the grain bounda-
ries as long as they are present in the metal matrix. However, when the dissolution process is
completed, the matrix grains are free to grow without any interference from precipitates. This
Stable particles
I 1 ,nm 2
Fig. 5.14. Predicted variation in average grain size D with Z1 and ra* for D°um. = 50 um, n = 0.5, and
D0 =0 (dissolving precipitates).
Time exponent n = 0.3
Complete particle dissolution
D,jim
I 1 ^m 1 ' 0
Fig. 5.15. Predicted variation in average grain size D with I1 and m* for D°um. - 50 Jim, n - 0.3, and
Do =0 (dissolving precipitates).
means that the grains, after prolonged high temperature annealing, will coarsen at a rate which
is comparable with that observed in Fig. 5.7 for particle-free systems. In the limiting case,
when m* = 0, the grain growth behaviour becomes identical to that shown in Figs. 5.9 and 5.10
for stable precipitates.
Example (5.3)
Consider a niobium-microalloyed steel with the following composition:
Assume that the base metal contains a fine dispersion of NbC precipitates in the as-received
condition, conforming to a limiting austenite grain size Dnm. of 50 jim. Provided that the
boundary drag from impurity elements in solid solution can be neglected (i.e. n ~ 0.5), esti-
mate on the basis of Fig. 5.14 the average austenite grain size D 7 in the material after 25 s of
isothermal annealing at 13000C. Relevant physical data for niobium-microalloyed steels are
given below:
Solution
The initial volume fraction of NbC in the material can be estimated from simple stoichiometric
calculations by considering the difference between total and soluble niobium. Taking the
atomic weight of Nb and C equal to 92.9 and 12.0 g mol"1, respectively, we obtain:
From this we see that the radius of the NbC precipitates in the base metal is close to:
By substituting this value into the expression for the proportionality constant m*, we obtain:
The average austenite grain size can now be read from Fig. 5.14 by interpolation between
the curves for m* = 1 and/= 0 (free grain growth). This gives:
Since the calculated value of D1 is reasonably close to that observed for a particle-free
system, it means that the presence of a fine dispersion of NbC in the base metal has no signifi-
cant effect on the resulting austenite grain size under the prevailing circumstances. Other
Next Page
types of niobium microalloyed steels may reveal a different grain coarsening behaviour, de-
pending on the chemical composition, size distribution, and initial volume fraction of the base
metal precipitates. However, the pattern remains essentially the same, i.e. the growth inhibi-
tion is always succeeded by grain coarsening as long as the precipitates are thermally unstable.
In welding the temperature will change continuously with time, which makes predictions of
the HAZ grain coarsening behaviour rather complicated. The method adopted from Ashby et
^ 24,25 j s b asec i o n m e jd e a o f integrating the elementary kinetic models over the weld thermal
cycle where the unknown kinetic constants are determined by fitting the integrals at certain
fixed points to data from real or simulated welds. Although the introduction of the Zener drag
in the grain growth equation largely increases the complexity of the problem, the methodology
and calibration procedure remain essentially the same. This means that the results from such
complex computations can be presented in the form of simple grain growth diagrams which
show contours of constant grain size in temperature-time space.
(5-39)
types of niobium microalloyed steels may reveal a different grain coarsening behaviour, de-
pending on the chemical composition, size distribution, and initial volume fraction of the base
metal precipitates. However, the pattern remains essentially the same, i.e. the growth inhibi-
tion is always succeeded by grain coarsening as long as the precipitates are thermally unstable.
In welding the temperature will change continuously with time, which makes predictions of
the HAZ grain coarsening behaviour rather complicated. The method adopted from Ashby et
^ 24,25 j s b asec i o n m e jd e a o f integrating the elementary kinetic models over the weld thermal
cycle where the unknown kinetic constants are determined by fitting the integrals at certain
fixed points to data from real or simulated welds. Although the introduction of the Zener drag
in the grain growth equation largely increases the complexity of the problem, the methodology
and calibration procedure remain essentially the same. This means that the results from such
complex computations can be presented in the form of simple grain growth diagrams which
show contours of constant grain size in temperature-time space.
(5-39)
(5-41)
where the limits tx and t2 refer to the total time spent in the thermal cycle from the chosen
reference temperature Tc to the peak temperature Tp and down again to Tc.
(5-42)
(5-43)
for the same pairs of values of Tp and Ar875 as above. The differential grain growth equation
can now be solved by selecting an appropriate starting value for M0* and evaluating the corre-
sponding Q-value which conforms to a mean grain size of D\ and D2, respectively. The
computations are repeated by adjusting M0* until a contour in Mo*-Q space is built up for each
grain size. The accepted values of M* and Q are then found by considering the intersection
point between the two curves, as shown schematically in Fig. 5.16(a).
(a) MS.
Accepted,
"" value
" ~vaTue~
Accepted^
a
(b) MJ
Accepted.
~ value" "
Acce|3ted_
value" "
Q*
Fig. 5.16. Method for calibrating unknown kinetic constants to experimental grain growth data;
(a) M*o-Q. (coarsening model), (b) M*o-Q* (dissolution model).
(5-44)
where a° and D°m include all constants entering the expressions for the dimensionless
supersaturation a and the element diffusivity Dm, respectively.
Under such conditions equation (5-37) becomes:
(5-45)
where AH* is the standard enthalpy of the dissolution reaction per mole of the diffusate (de-
fined by the solubility product in equation (4-5), Chapter 4), and Qd is the activation energy for
diffusion of the less mobile constituent atom of the precipitates in the matrix.
By calculating the integral in equation (5-45) for the same pairs of values of Tp and Af8/5 as
above and selecting an appropriate starting value for Mo* in equation (5-41), it is possible to
build up a contour in Mo*-£T space for each grain size that satisfies the differential grain
growth equation. The accepted values of M* and Q* are then found by considering the inter-
section point between the curves representing D1 and D2 in Fig. 5.16(b).
(5-46)
(5-47)
(5-48)
(5-49)
Table 5.1 Chemical composition of Ti-microalloyed steel used by Ion et al.25 (in wt%).
C Si Mn P S Al Ti N
Table 5.2 Data used to construct welding maps for Ti-microalloyed steel (compiled from miscellane-
ous sources).
i
Calibration point
The response of the base material to welding under 2-D and 1-D heat flow conditions is
shown in Fig. 5.18(a) and (b), respectively. As expected, the presence of TiN particles is seen
to retard austenite grain growth within the heat affected zone during welding. However, since
particle coarsening is a physical process occurring at temperatures well below the equilibrium
solvus of the precipitates, the problem cannot be eliminated. This means that a coarse grained
region will always form adjacent to the fusion boundary, even at very low heat inputs, as
indicated by the nomograms in Fig. 5.18(a) and (b).
Example (5.4)
Consider SA welding on a thick plate of a titanium-microalloyed steel under the following
conditions:
Evaluate on the basis of the nomograms in Fig. 5.18(a) the variation in the austenite grain
size across the fully transformed HAZ after welding. Estimate also the total width of the HAZ
(referred to the fusion boundary) under the prevailing circumstances.
Solution
First we calculate the net heat input per unit length of the weld:
Readings from Fig. 5.18(a) give the HAZ austenite grain size profile shown in Fig. 5.19.
(a) Thick plate welding (2-D heat flow)
Prior austenite grain size (^m)
Relative size of pinning precipitates (r/ro)
Net heat input (qo/v), kJ/mm
Peak temperature, 0C
FZg. 5.18. HAZ grain growth diagrams for titanium-microalloyed steel; (a) Thick plate welding (2-D heat
flow), (b) Thin plate welding (1-D heat flow). No preheating (T0 = 200C).
SAW (Ti-microalloyed steel)
Peak temperature, 0C
Fig. 5.19. Predicted variation in austenite grain size across the fully transformed HAZ of a Ti-microalloyed
steel weld (Example 5.4).
The total width of the fully transformed HAZ can now be estimated from equation (5-47)
by using data from Table 1.1 (Chapter 1). Taking the Ac3 -temperature equal to 9100C, we
obtain:
Based on the quoted value of £2 in Table 5.2, it is also possible to estimate the initial volume
fraction of TiN in the base metal. Taking C5 « 6.67 X 104 |im3 K s"1 and k « 0.5 for titanium-
microalloyed steels, we obtain from equation (5-42):
A comparison with the experimental data of Ringer et aill shows that a volume fraction of
2 X 10"4 is reasonably close to that measured by microscopic assessment methods.
Example (5.5)
In Table 5.4 the quoted value for the effective grain boundary pinning constant Q* is 5.0 X
1018 |im 2 s"1. Based on equations (5-27) and (5-44), estimate the initial volume fraction/, and
radius ro of NbC in the parent material under the prevailing circumstances.
Table 5.3 Chemical composition of Nb-microalloyed steel used by Ion et al.25 (in wt%).
C Si Mn P S Al Nb N
Nb-microalloyed steel
SAW
GMAW
Laser welding
Weld simulation
Measured grain size, urn
Calibration point
Fig. 5.20. Comparison between measured and predicted HAZ austenite grain sizes after calibration of
model to data reported by Ion et al.25 for Nb-microalloyed steel (real and simulated thick plate welds).
(a)
Thick plate welding (2-D heat flow)
Prior austenite grain size (urn)
Relative volume fraction of pinning precipitates (f/fo)
Net heat input (qo/v), kJ/mm
Peak temperature, 0C
Fig. 5.21. HAZ grain growth diagrams for niobium-microalloyed steel; (a) Thick plate welding (2-D heat
flow), (b) Thin plate welding (1-D heat flow). No preheating (T0 = 200C).
Table 5.4 Data used to construct welding maps for Nb-microalloyed steel (compiled from miscellane-
ous sources).
Reasonable average values for a 0 , D°m, k, and D°nm. are given below:
Solution
The initial volume fraction of NbC can be estimated from equation (5-44). After rearranging
this equation, we obtain:
The corresponding radius of the precipitates can now be obtained from equation (5-27):
Although experimental data are not available for a direct comparison, the calculated values
for/o and ro are reasonable and of the expected order of magnitude (see Example 5.3).
Base plate
Peak temperature
Fig. 5.22. Schematic diagram illustrating the effect of non-metallic inclusions on the weld metal grain
coarsening behaviour.
Table 5.5 Chemical composition of C-Mn steel weld metal used by Kluken et al.29 (in wt%).
C O Si Mn P S N Nb V Al Ti
0.09 0.034 0.48 1.86 0.01 0.01 0.005 0.004 0.02 0.018 0.005
Table 5.6 Data used to construct welding maps for C-Mn steel weld metal (compiled from
miscellaneous sources).
" I ~pp. I K
1
I bHm. I K?
(kJmol" ) (|xm) (|xm) (JJLm2S-1)
dispersed inclusions within the weld metal gives rise to a strong austenite grain boundary
pinning effect, similar to that documented for TiN in microalloyed steels. However, since no
particle coarsening occurs in the present case, it means that a small austenite grain size
(< 95jnm) is preserved at all relevant peak temperatures, irrespectively of the applied heat
input. Accordingly, the weld metal grain growth behaviour is seen to be quite different from
that of the base metal, even when the nominal chemical composition has not been significantly
changed by the welding process.
Austenite grain size, jim
Cooling time, A t . . s
8/5
Fig. 5.23. Comparison between measured and predicted austenite grain sizes after calibration of model
to data reported by Kluken et al.29 for C-Mn steel weld metal (simulated thick plate heat cycles).
Example (5.6)
Consider deposition of a cap layer (GMAW) on the top of a thick multipass Nb-microalloyed
steel weld under the following conditions:
Estimate on the basis of the nomograms in Figs. 5.21(a) and 5.24(a) the variation in the
prior austenite grain size across the fully transformed HAZ at different locations along the
periphery of the weld after arc extinction. The situation is illustrated in Fig. 5.25.
Solution
First we calculate the net heat input per unit length of the weld:
Readings from Figs. 5.21(a) and 5.24(a) give the HAZ austenite grain size profiles shown
in Fig. 5.26. From this we see that the prior austenite grain size adjacent to the fusion bound-
ary of a cap layer will vary significantly with position along the periphery of the weld, depend-
ing on the type of material sampled (i.e. base plate or weld metal).
Peak temperature, 0C
Fig. 5.24. Grain growth diagrams for reheated C-Mn steel weld metal; (a) Thick plate welding (2-D heat
flow), (b) Thin plate welding (1-D heat flow). No preheating (T0 = 200C).
HAZ Base plate
Dissolution of NbC
Peak temperature, 0C
Fig. 5.26. Predicted variation in austenite grain size across the fully transformed HAZ at different loca-
tions along the periphery of the weld (Example 5.6).
and their high creep strength. They are readily weldable, although reheat cracking and cold
cracking in the HAZ and fusion zone may be a problem.30 In practice, these difficulties can be
overcome by the choice of an appropriate welding procedure (e.g. preheating in combination
with a low heat input), which reduces austenite grain growth and maximises the proportion of
the HAZ refined by subsequent weld passes.3031
The grain growth data of Miranda and Fortes31 provide a basis of calibrating the kinetic
constants in equations (5-22) and (5-45), as shown in Fig. 5.27. Information about steel chemical
composition and parameters used to construct the maps are contained in Table 5.7 and 5.8,
respectively.
It is seen from the nomograms in Fig. 5.28(a) and (b) that considerable austenite grain
growth occurs in Cr-Mo low-alloy steel weldments due to dissolution of molybdenum carbide
(Mo2C) at elevated temperatures. However, the maximum HAZ austenite grain size is signifi-
Table 5.7 Chemical composition of Cr-Mo steel used by Miranda and Fortes31 (in wt%).
C Si Mn P S Cr Mo Ni Cu Al V
0.10 0.24 0.46 0.01 0.01 2.19 0.94 0.23 0.1 0.01 0.01
Calibration points
Symbol * w s >
Table 5.8 Data used to construct welding maps for Cr-Mo low-alloy steel (from Refs. 24 and 31).
cantly smaller than that observed during welding of Nb-microalloyed steels, in spite of the fact
that Mo2C is less stable than NbC. This situation can probably be attributed to drag from
alloying and impurity elements in solid solution, which retard grain growth through elastic
attraction of the atoms towards the open structure of the grain boundary (indicated by a time
exponent n = 0.32 in Table 5.8).
Peak temperature, 0C
Fig. 5.28. HAZ grain growth diagrams for Cr-Mo low-alloy steel; (a) Thick plate welding (2-D heat
flow), (b) Thin plate welding (1-D heat flow). No preheating (T0 = 200C).
Table 5.9 Chemical composition of type 316 austenitic stainless steel used by Ashby and Easterling24
(in wt%).
C Mn Si Cr Ni Mo P S
Table 5.10 Data used to construct welding maps for type 316 austenitic stainless steel (compiled from
miscellaneous sources).
n I Qapp. 1
I Atf*+ I Q/ I D°lm.
1 1
I K I M
o* I ^*
(kJmol- ) (kJmol- ) (kJ mol" ) (|jim) (fxm) (1JLm2S-1) (1Xm2S-1)
!Estimated from data quoted by Kou32 for Cr23C6 in 304 stainless steel.
^Activation energy for diffusion of Cr in austenite.
Measured grain size, u.m
Calibration point
Fig. 5.29. Comparison between measured and predicted HAZ austenite grain sizes after calibration of
model to data reported by Ashby and Easterling24 for type 316 stainless steel (simulated thick plate
welds).
It follows from the nomograms in Fig. 5.30(a) and (b) that the presence of Cr23C6 has little
influence on the HAZ grain coarsening behaviour because it dissolves at a fairly low tempera-
ture. This results in a rather coarse austenite grain structure adjacent to the fusion boundary,
which in certain cases may exceed 100 jim.
Thick plate welding (2-D heat flow)
Austenite grain size ^x m)
Relative volume fraction of pinning precipitates (f/fo)
Net heat input (qo/v), kJ/mm
Peak temperature, 0C
Fig. 5.30. HAZ grain growth diagrams for type 316 austenitic stainless steel; (a) Thick plate welding
(2-D heat flow), (b) Thin plate welding (1-D heat flow). No preheating (T0 = 200C).
Example (5.7)
Consider GTA butt welding of a 2mm thin sheet of type 316 austenitic stainless steel under the
following conditions:
Provided that the conditions for one dimensional (1-D) heat flow are met, estimate on the
basis of the nomograms in Fig. 5.30(b) the variation in the austenite grain size across the HAZ
after welding. Estimate also the total width of the HAZ (referred to the fusion boundary)
under the prevailing circumstances.
Solution
First we calculate the net heat input per mm2 of the weld:
Readings from Fig. 5.30(b) give the HAZ austenite grain size profile shown in Fig. 5.31.
The total width of the grain growth zone can be estimated from equation (5-49) by using
data from Table 1.1 (Chapter 1). Taking the peak temperature for incipient dissolution of
Cr23C6 equal to 9000C, we obtain:
Austenite grain size, jim
Peak temperature, 0C
Fig. 5.31. Predicted variation in austenite grain size across the HAZ of a 316 austenitic stainless steel
weld (Example 5.7).
5.5 Computer Simulation of Grain Growth
The multivariable characteristics of grain growth impose several restrictions on the use of
analytical modelling techniques for a mathematical description of the grain evolution during
welding. However, many of these restrictions can be relaxed by employing novel computa-
tional techniques in mainframe facilities which enable the modelling of exact grain shapes in
topological connected microstructures.
Historically, the application of the computer simulation technique to poly cry stalline micro-
structures has evolved along two different paths. The first approach treats the grain boundaries
as continuous interfaces, which are governed by determistic equations of motion.33"37 The
other approach is a Monte Carlo-based technique that takes into account explicitly the interac-
tions among individual grains by discretising the microstructure to construct an image of the
grain aggregates in the computer.38"43 Both approaches have been applied to the modelling of
grain growth phenomena in welding 374445 with the objective of incorporating important side-
effects which cannot readily be accounted for in a simple analytical treatment of the process.
D x 10 (arbitrary units)
(b)
D x 10 (arbitrary units)
(C)
D x 10 (arbitrary units)
Fig. 5.32. Effect of 'thermal pinning' on grain growth; (a) Plot of D vs t for different simulation condi-
tions, (b) Coarsening kinetics of different thermal regions for a mobility gradient of 5:1, (c) Same as in
(b) for a mobility gradient of 10:1. Data from Fortes and Soares.37
5.5.2 Free surface effects
In a weld HAZ the fusion line represents a physical barrier against grain growth which cannot
be exceeded. In principle, it can be regarded as a free surface, which means that the grain
boundaries must meet the fusion line at right angles.
The simulation results of Saetre and Ryum36 provide a basis for evaluating to what extent
grain growth under isothermal conditions is affected by the presence of a free surface. Figure
5.33 shows the evolution of a 2-D Voronoi structure with time. In Fig. 5.33(a) all grain boundaries
are straight lines, but the triple line junctions are not in equilibrium. However, adjustments of
the triple line junctions into equilibrium positions lead to local curvatures of the grain bounda-
ries near the junctions, and the grain growth process is thus initiated. Initially, the Voronoi
structure contained 485 grains, but this number is gradually reduced during the coarsening
process (Figs. 5.33(b), (c) and (d)). A qualitative inspection of the diagrams reveals, on the
other hand, no clear difference in the grain size in the radial position, which indicates that the
constrain provided by the free surface is only of minor importance in the present context. It
should be noted that this does not exclude the possibility that the HAZ grain size is influenced
by the presence of a fusion boundary, since other effects such as surface grooving and solute
segregation (not studied here) can impose additional restrictions on the system by contributing
to physical pinning of the grain boundaries. Consequently, further modelling work is required
to explore these possibilities.
References
(C) (d)
Fig. 5.33. The evolution of 2-D grain structures during normal grain growth; (a) Initial Voronoi structure,
485 grains, (b) 456 grains, (c) 349 grains, (d) 251 grains. Data from Saetre and Ryum.36
Appendix 5.1
Nomenclature
end temperature of ferrite to austenite element diffusivity (jum2 s"1, mm2 s"1
transformation (0C or K) or m2 s"1)
kinetic strength of thermal cycle with initial particle radius at t - 0 (nm, jam
respect to particle coarsening (fxm3) or m)
kinetic constant in expression for M cooling time from 800 to 5000C (s)
(variable units)
temperature (0C or K)
modified kinetic constant in expression
for M(J^m1711S-1) reference temperature (0C or K)
arc efficiency
6
Solid State Transformations in Welds
6.1 Introduction
The majority of phase transformations occurring in the solid state take place by thermally
activated atomic movements. In welding we are particularly interested in transformations that
are induced by a change in temperature of an alloy with a fixed bulk composition. Such
transformations include precipitation reactions, eutectoid transformations, and massive trans-
formations both in the weld metal and in the heat affected zone.
Since welding metallurgy is concerned with a number of different alloy systems (including
low and high alloy steels, aluminium alloys, titanium alloys etc.), it is not possible to cover all
aspects of transformation behaviour. Consequently, the aim of the present chapter is to pro-
vide the background material necessary for a verified quantitative understanding of phase
transformations in weldments in terms of models based on thermodynamics, kinetics, and
simple diffusion theory. These models will then be applied to specific alloy systems to illumi-
nate the basic physical principles that underline the experimental observations and to predict
behaviour under conditions which have not yet been studied.
(6-1)
Here AGy (the volume free energy change associated with the transformation) and AGD
(free energy donated to the system when the nucleation takes place heterogeneously) are nega-
tive, since they assist the transformation, while AGS (increase in surface energy between the
two phases) and AGE (increase in strain energy resulting from lattice distortion) are both posi-
tive because they represent a barrier against nucleation. It follows that the transformation
Molar free energy Stable (B) Stable (a)
Temperature
Fig. 6.1. Schematic representation of the molar free energies of two solid a and P phases as a function of
temperature (allotrophic transformation — no compositional change).
reaction can proceed when the driving force AG becomes greater than the right-hand side of
equation (6-1).
For an allotropic transformation, in which there is no compositional change, AG will be a
simple function of temperature, as illustrated in Fig.6.1. For alloys the situation is slightly
more complex, since there is an additional variable, i.e. the composition. In such cases the
temperature at which the a-phase becomes thermodynamically unstable (Teq) corresponds to a
fixed point on the a-(3 solvus boundary in the equilibrium phase diagram, as shown
schematically in Fig. 6.2. Since phase diagrams are available for many of the important indus-
trial alloy systems, it means that the driving force for a transformation reaction can readily be
obtained from such diagrams in the form of a characteristic undercooling (AT).
Temperature
%B
Fig. 6.2. Schematic representation of the a-(3 solvus boundary in a simple binary phase diagram.
6.2.2 Heterogeneous nucleation in solids
In general, solid state transformations in metals and alloys occur heterogeneously by nuclea-
tion at high energy sites such as grain corners, grain boundaries, inclusions, dislocations and
vacancy clusters. The potency of a nucleation site, in turn, depends on the energy barrier
against nucleation (AG*^) which is a function of the 'wetting' conditions at the substrate/
nucleus interface.1
It can be seen from Fig. 6.3 that nucleation at for instance inclusions or dislocations is
always energetically more favourable than homogeneous nucleation (AG*het < tsG*hom ) but
less favourable than nucleation at grain boundaries or free surfaces. As a result, the transfor-
mation behaviour is strongly influenced by the type and density of lattice defects and second
phase particles present within the parent material.
(6-2)
Grain boundary
Free
surface Inclusion
horn.
Dislocations/stacking faults
AG* /AG*
Vacancy clusters
het.
Inclusions
boundaries
Grain
corners
Grain
surfaces
Free
Nucleation site
Fig. 6.3 Schematic diagram showing the most potent sites for heterogeneous nucleation in metals and
alloys.
It follows from the graphical representation of equation (6-2) in Fig. 6.4 that the nucleation
rate Nhet is highest at an intermediate temperature due to the competitive influence of
undercooling (driving force) and diffusivity on the reaction kinetics. This change in Nhet with
temperature gives rise to corresponding fluctuations in the transformation rate, as shown
schematically in Fig. 6.5. Note that the peak in transformation rate is due to two functions,
growth and nucleation (which peak at different T) whereas peak in Nhet is due to nucleation
only.
(6-3)
(6-4)
T T
Low undercooling
High diffusivity
High undercooling
Low diffusivity
%B Nhet.
Fig. 6.4. Schematic diagram showing the competitive influence of undercooling (driving force) and
diffusivity on the heterogeneous nucleation rate.
T
logt
Fraction transformed
logt
Fig. 6.5. Fraction transformed as a function of time referred to the C-curve (schematic).
(6-5)
At the chosen reference temperature Tr the time difference between the real C-curve and
the extension of the lower asymptote amounts to (see Fig. 6.6):
(6-6)
from which
(6-7)
It follows that equations (6-5) and (6-7) provide a systematic basis for obtaining quantita-
tive information about Qd and AGhet from experimental microstructure data through a simple
graphical analysis of the shape and position of the C-curve in temperature-time space.
1/r
T
lnt
Fig. 6.6. Determination of AG*het and Qd from the C-curve (schematic).
(6-8)
where TV4 is the Avogadro constant, ^ n is the interfacial energy per unit area between the nu-
cleus and the matrix, AGV is the driving force for the precipitation reaction (i.e. the volume free
energy change associated with the transformation), and 5(0) is the so-called shape factor which
takes into account the wetting conditions at the nucleus/substrate interface.
For a particular alloy, AGV is for small Ar proportional to the degree of undercooling:l
(6-9)
where C3 is a kinetic constant. This equation follows from the definition of AGv in diluted alloy
systems and the mathematical expression for the solvus boundary in the binary phase dia-
gram.
By substituting equation (6-9) into equation (6-8), we get:
(6-10)
It follows that A0 is a characteristic material constant which is related to the potency of the
heterogeneous nucleation sites in the material. The value of A0 is, in turn, given by equations
(6-7) and (6-10):
(6-11)
In cases where A0 is known, it is possible to obtain a more general expression for t* by
substituting equation (6-10) into equation (6-3):
(6-12)
Equation (6-12) can further be modified to allow for compositional and structural varia-
tions in the parent material by using the calibration procedure outlined in Fig. 6.7. Let tr
denote the time taken to precipitate a certain fraction of (3 at a chosen reference temperature
T= Tr in an alloy containing Nv nucleation sites per unit volume. If we take the corresponding
solvus temperature of the (3-phase equal to T*q , the expression for t* becomes:
(6-13)
(6-14)
Equation (6-14) provides a basis for predicting the displacement of the C-curve in tempera-
ture-time space due to compositional or structural variations in the parent material. In gen-
T
C-cun/e(Nv=N^)
logt
Fig. 6.7. Method for eliminating unknown kinetic constant in expression for t*.
eral, an increase in Nv will shift the nose of the C-curve to the left in the diagram (i.e. towards
shorter times), as shown schematically in Fig. 6.8, because of the resulting increase in the
nucleation rate. Moreover, in solute-depleted alloys the critical undercooling for nucleation
will be reached at lower absolute temperatures where the diffusion is slower. This results in a
marked drop in Nhet, which displaces the C-curve towards lower temperatures and longer times
in the IT-diagram, as indicated in Fig. 6.9.
Example (6.1)
Isothermal transformation (IT) or continuous cooling transformation (CCT) diagrams are avail-
able for many of the important alloy systems.4 In the case of aluminium, so-called tempera-
ture-property diagrams exist for different types of wrought alloys.45 Suppose that the C-curve
in Fig. 6.10 conforms to incipient precipitation of [3'(Mg2Si) particles at manganese-contain-
ing dispersoids in 6351 extrusions. Use this information to estimate the values of A0 and Qd in
equation (6-3) when the solvus temperature of (3'(Mg2Si) is 5200C.
Solution
The parameters A0 and Qd can be evaluated from the C-curve according to the procedure
shown in Fig. 6.6. Referring to Fig. 6.11, the value of AG^ at the chosen reference tempera-
ture Tr = 35O°C (623K) is equal to:
When AGhet is known, the parameter A0 can be obtained from equation (6-11):
iogt
Fig. 6.8. Effect of Nv on the shape and position of C-curve in temperature-time space (schematic).
T T T
%B \e, logt
Fig. 6.9. Effect of solute content on the shape and position of C-curve in temperature-time space (sche-
matic).
Similarly, Qd can be read from Fig. 6.11 by considering the slope of the lower asymptote:
This value is in good agreement with the reported activation energy for diffusion of magne-
sium in aluminium.6
AA 6351 - T6
Temperature, 0C
Time, s
Fig. 6.10. C-curve for 99.5% maximum yield strength of an AA6351-T6 extrusion. After Staley.5
Solvus temperature: 520 0C
103/T, K"1
lnt
Fig. 6.11. Determination of kG*heU and Qd from the C-curve in Fig. 6.10 (Example 6.1).
(6-15)
(a)
(b)
Fig. 6J2. Schematic illustration of atom transfer across different kinds of interfaces; (a) Incoherent
interface, (b) Coherent interface.
U
a/p
Lateral move-
ment of incoherenf
interface
(6-16)
(6-17)
Liquid
Temperature
%B
Concentration
Diffusion of solute
Distance
Fig. 6.14. Schematic representation of concentration profile ahead of advancing interface during precipi-
tation of (B from a supersaturated a-phase.
Temperature
%B
Concentration
Diffusion of solute
Distance
Fig. 6.15. Schematic representation of concentration profile ahead of advancing interface during growth
of solute-depleted P into a metastable a-phase.
(6-18)
where Dm is the diffusivity of the solute in the matrix, and erfc(u) is the complementary error
function (defined previously in Appendix 1.3, Chapter 1).
Similarly, for growth of spherical precipitates, the variation in the radius r with time can be
written as:8
(6-19)
The parabolic relations in equations (6-17) and (6-19) imply that the growth rate slows
down as the (3-phase grows. This is due to the fact that the total amount of solute partitioned
during growth decreases with time when the diffusion distance increases. Moreover, the form
of equations (6-18) and (6-20) suggests that the maximum in the growth rate is achieved at an
intermediate temperature because of the competitive influence of undercooling (driving force)
and diffusivity on the reaction kinetics. Consequently, a plot of E1 or £2 vs temperature will
reveal a pattern similar to that shown in Fig. 6.4 for the nucleation rate, although the thicken-
ing constants generally are less temperature-sensitive.
In addition to the models presented above for plates and spheres, approximate solutions
also exist in the literature for thickening of needle-shaped precipitates, based on the Trivedi
theory for diffusion-controlled growth of parabolic cylinders.9 However, because of space limi-
tations, these solutions will not be considered here.
where X is the fraction transformed, n is a time exponent, and k is a kinetic constant which
depends on the nucleation and growth rates.
The exponential growth law summarised in the Avrami equation is valid for linear growth
under most circumstances, and approximately valid for the early stages of diffusion-controlled
growth.10 Table 6.1 gives information about the value of the time exponent for different ex-
perimental conditions.
In general, the value of n will not be constant, but change due to transient effects until the
steady-state nucleation rate is reached and n attains its maximum value. Subsequently, the
nucleation rate starts to decrease as the sites become filled with nuclei and eventually ap-
proach zero when complete saturation occurs. This is because the heterogeneous nucleation
sites are not randomly distributed in the volume, but are concentrated near other nucleation
sites leading to an overall reduction in n. From then on, the transformation rate is solely
controlled by the growth rate.
kinetic theory, using the classic models of nucleation and growth described in the previous
sections. In practice, however, this is a rather cumbersome method, particularly if the base
metal is of a heterogeneous chemical nature. Alternatively, we can calibrate the Avrami equa-
tion against experimental microstructure data, e.g. obtained from generic IT-diagrams. A con-
venient basis for such a calibration is to write equation (6-21) in a more general form:
(6-22)
where k* is a new kinetic constant (equal to kr1/n). In the latter equation the parameter k* can be
regarded as a time constant, which is characteristic of the system under consideration. Note
that this form of the Avrami equation is mathematically more appropriate, as the dimensions of
the k* constant are not influenced by the value of the time exponent n.
During the early stages of a transformation reaction, the reaction rate is controlled by the
nucleation rate. Let f denote the time taken to precipitate a certain fraction of P (X = Xc) at
an arbitrary temperature T (previously defined in equation (6-14)). It follows from equation
(6-22) that the value of k* in this case is given as:
(6-23)
(6-24)
from which
(6-25)
(6-26)
By using the standard Johnson-Mehl correction for physical impingement of adjacent trans-
formation volumes, we may write in the general case:
(6-27)
This specific form of the Avrami equation is valid under conditions of early site saturation
where the a/p-interface is completely covered by P nuclei at the onset of the transformation.
Fig. 6.16, Schematic illustration of the planar geometry assumed in the site saturation model.
metals and alloys, it appears that the bulk of the research has been concentrated on modelling
of microstructural changes under predominantly isothermal conditions.1"411 In contrast, only
a limited number of investigations has been directed towards non-isothermal transforma-
tions. 51012 " 18 However, these studies have clearly demonstrated the advantage of using ana-
lytical modelling techniques to describe the microstructural evolution during continuous cool-
ing, instead of relying solely on empirical CCT-diagrams.
Cooling curve
logt
Fig. 6.17. Schematic illustration of the Scheil theory.
6.2.5.2 Isokinetic reactions
The concept of an isokinetic reaction has previously been introduced in Section 4.4.2.3 (Chap-
ter 4). A reaction is said to be isokinetic if the increments of transformation in infinitesimal
isothermal time steps are additive, according to equation (6-29). Christian10 defines this math-
ematically by stating that a reaction is isokinetic if the evolution equation for some state vari-
able X may be written in the form:
(6-30)
(6-31)
(6-32)
In equation (6-32) Z1 represents the kinetic strength of the thermal cycle with respect to P-
precipitation. This parameter is generally defined by the integral:
-2 -1
Temperature, 0C
Fig. 6.18. Predicted variation in N*het and E1 with temperature during the austenite to ferrite transfor-
mation in a C-Mn steel (0.15 wt% C, 0.40 wt% Mn). Data from Umemoto et al.19
(6-33)
where dt is the time increment at T, and f is the corresponding hold time required to reach Xc
at the same temperature (given by equation (6-14)). The derivation of equation (6-32) is shown
in Appendix 6.2
The principles of additivity are also applicable under conditions of early site saturation. If
only U^p varies with temperature, it is possible to rewrite equation (6-28) in an integral form:
(6-34)
This equation can readily be integrated by numerical methods when the temperature-time
programme is known.
tions. Solution of the differential equation then requires stepwise integration in tempera-
ture-time space, using an appropriate numerical integration procedure. As already pointed
out, this will generally be the case for diffusion-controlled precipitation reactions, since the
evolution parameter X is a true function of temperature. Under such conditions, experimen-
tally based continuous cooling transformation (CCT) diagrams must be employed.
The microstructural constituents listed above are indicated in Fig. 6.19, which shows
photomicrographs of typical regions within low-alloy steel weldments.
tions. Solution of the differential equation then requires stepwise integration in tempera-
ture-time space, using an appropriate numerical integration procedure. As already pointed
out, this will generally be the case for diffusion-controlled precipitation reactions, since the
evolution parameter X is a true function of temperature. Under such conditions, experimen-
tally based continuous cooling transformation (CCT) diagrams must be employed.
The microstructural constituents listed above are indicated in Fig. 6.19, which shows
photomicrographs of typical regions within low-alloy steel weldments.
(C) (d)
Fig. 6.19. Optical micrographs showing various microstructural constituents commonly found in low-
alloy steel weldments; (a) Coarse grained HAZ (low heat input welding), (b) Coarse grained HAZ (high
heat input welding), (c) As-deposited weld metal (low heat input welding), (d) Reheated weld metal (low
heat input welding). Letters in micrographs are defined in the text.
discretions. This controversy in terminology has been a source of confusion, and the work by
Sub-Commission IXJ of the International Institute of Welding (HW)23 for developing guide-
lines for quantification of microstructures is, therefore, an important step towards a standard-
ised system of nomenclature.
The IIW recommendations are based on the scheme originally proposed by Abson and
Dolby.23 The IIW system involves a simplified classification procedure compared with the
outline used in Fig. 6.19, since the distinction between acicular ferrite and the various sideplate
structures is based on features such as aspect ratio, relative lath size, and number of parallel
laths. This has led to the introduction of the FS-constituent (ferrite with aligned second phase),
which, in principle, comprises both Widmanstatten ferrite and upper bainite.
In contrast to the HW approach to classifying microstructural elements based on their ap-
pearance in the optical microscope, other investigators rank the various constituents solely in
terms of their transformation behaviour, according to the scheme originally proposed by Dube
et ai26 From a scientific point of view, this classification system is more correct, since it does
not violate common terminology based on thermodynamics and kinetics of transformation
reactions. However, with the omission of the FS-constituent grouping utilised by the HW, the
Dube system is more inconvenient to use in practice because the different transformation prod-
ucts often cannot readily be identified on the basis of their transformation characteristics. Con-
sequently, both classification systems appear to have their weaknesses, which, in turn, limit
their applicability.
This orientation relationship, which lies within the so-called Bain orientation region,27 is
adopted in order to minimise the increase in the strain energy resulting from lattice distortion
AGEby formation of a low-energy interface between the ferrite nucleus and the parent austenite
phase.1 Subsequent growth of the ferrite may then occur into the adjacent austenite grain with
which the ferrite has a random orientation relationship,28 since a disordered (incoherent) inter-
face generally has a higher mobility than an ordered (coherent/semi-coherent) interface at low
undercoolings.
(6-35)
Edge
Corner
Fig. 6,20. Sketch of an austenite grain showing different sites for ferrite nucleation.
Based on equation (6-35) it is possible to predict the displacement of the C-curve in tem-
perature-time space due to structural or compositional variations in the parent material. As an
illustration, we shall assume that the parameters listed in Table 6.2 are representative of nu-
cleation of grain boundary ferrite in a low-alloy steel with an initial austenite grain size of 10
jim. In addition, we need information about the Ae3-temperature in the equilibrium phase
diagram. This temperature can readily be obtained from thermodynamic calculations, even for
multicomponent systems.29 Alternatively, we can use the empirical relationship quoted by
Leslie:30
(6-36)
Parameter I T I T I T*eq I T0 I Q] I Dy
(K) (s) (K) (Jmol"1) (kJmol-1) (jim)
Evaluate on the basis of the grain growth diagram in Fig. 5.21 (a) (Chapter 5) and equations
(6-32), (6-33), and (6-35) the conditions for ferrite formation at two different positions within
the HAZ corresponding to a peak temperature Tpof 13500C and 10000C, respectively. As-
sume in these calculations that the ferrite may form within the temperature range from 800 to
6000C, and that the equilibrium volume fraction of ferrite (f^qFe) in the fully transformed
steel is 0.9.
Solution
First we calculate the net heat input per unit length of the weld:
Readings from Fig. 5.21 (a) give the following HAZ grain sizes:
As expected, the theoretical C-curves in Fig. 6.21 reveal a strong effect of the austenite
grain size on the HAZ transformation kinetics. By considering the superimposed weld cool-
ing curve, it is possible to estimate the volume fraction of ferrite/""7^ which forms in each case
from equations (6-32) and (6-33). Taking the time exponent n in the Avrami equation equal
to 5/2 for nucleation of ferrite at austenite grain boundaries10 and 1 — Xc = 0.98, we obtain:
Time, s
Fig. 6.21. Effect of austenite grain size on the HAZ transformation kinetics (Example 6.2). The superim-
posed cooling curve corresponds to a cooling time, A%5, of 21s.
and
and
From this we see that polygonal ferrite dominates the microstructure within the grain re-
fined region, whereas ferrite hardly forms within the grain coarsened HAZ under the prevail-
ing circumstances. Although experimental data are not available for a direct comparison, the
predicted effect of the austenite grain size on the HAZ transformation kinetics is reasonable
and consistent with general experience (e.g. see experimental CCT-diagrams in Fig. 6.22).
(i) To ensure the desired strength level by solid solution or precipitation strengthening;
(ii) To control the microstructure through modification of the nucleation and growth rates of
proeutectoid ferrite.
Temperature, 0C
Time, s
Fig. 6.22. CCT-diagrams for a low-carbon Cu-Ni containing steel. Superimposed on the CCT-diagrams
are two cooling curves corresponding to Af875 of 10 and 100 s, respectively. Austenitising conditions;
Heavy solid lines: 9000C for 5 min, Heavy broken lines: 13000C for 5 s. Data from Cross et al?x
In the latter case the effect is related to a shift in the Ae3-temperature of the steel, which
alters the undercooling and hence, the driving force for the austenite to ferrite transformation.
This point is illustrated by the following example.
Example (6.3)
Consider SA welding on a thick plate of a Nb-microalloyed steel under conditions similar to
those employed in Example (6.2). Based on equations (6-32), (6-33), and (6-35) estimate the
volume fraction of grain boundary ferrite in the grain coarsened HAZ (Tp = 13500C) after
welding when the A^-temperature of the steel is 863°C (1136K).
Solution
By substituting data from Table 6.2 into equation (6-35), we arrive at the following expression
for r *50:
It is evident from the graphical representation of the above equation in Fig. 6.23 that an in-
crease in the v4e3-temperature (e.g. from 1108 to 1136K) displaces the C-curve towards higher
temperatures and shorter times in the IT-diagram. This, in turn, gives rise to improved con-
ditions for ferrite nucleation. Taking the time exponent n in the Avrami equation equal to 5/2
and (1 — Xc) = 0.98 as in the previous example, we obtain after integration:
Temperature, 0C
Time, s
Fig. 6.23.Effect of steel chemical composition (Ae3-temperature) on the HAZ transformation kinetics
(Example 6.3). The superimposed cooling curve corresponds to a cooling time, Ar875, of 21s.
and
Although the calculated value offa'Fe is rather uncertain, the trends predicted in the present
example are reasonable and consistent with general experience.
*Based on classic nucleation theory it can be argued that a decrease in the y-y grain boundary energy due to boron
segregations will suppress the formation of ferrite at these sites. Further enhancement of the energy barrier against
ferrite nucleation through an increase in the total strain energy of the embryo is possible if the free grain boundary
volume becomes filled with either boron atoms or borocarbide precipitates (e.g. Fe23(B9C)6).
During welding, quantitative information about the extent of boron segregation which oc-
curs to the austenite grain boundaries under various thermal programmes can be obtained on
the basis of a well established theoretical model for quench-induced segregation of boron in
steel.3839 At peak temperatures above 1000 to 11000C, borocarbides and -nitrides present in
the base plate will rapidly dissolve in the matrix,4041 leading to a significant increase in the
amount of free diffusible boron. Generally, solute atoms in a crystal lattice will have an as-
sociated strain energy,38 which implies that it is energetically feasible to pair the solute boron
atom with a vacancy. Since the formation of vacancies is a thermally activated process, it fol-
lows that the fraction of boron occupying such sites, [B]v, increases exponentially with tem-
perature:39
(6-37)
Here m contains various geometric and entropy terms, Zy is the vacancy formation energy,
Eb is the vacancy-boron binding energy, and [B] is the bulk concentration of free boron. By
substituting reasonable average values for ra, Ep and Eb into equation (6-37), we arrive at the
following expression for [B]V:39A2
(6-38)
when Tp » Ae3.
In addition to the diffusible boron content, the grain size is also an important variable in
steel hardenability. A quantitative estimate of the combined effect of boron segregations and
austenite grain size on the HAZ transformation kinetics can be obtained by assuming that the
ferrite nucleates primarily on grain faces. In such cases the total number of heterogeneous
nucleation sites per unit volume Nvis given as:
(6-40)
where na is the number of nucleation sites per unit grain boundary area, and Sv is the grain
boundary surface area per unit volume (equal to 2/Dy).
If we, as a second approximation, assume that na is inversely proportional to the amount of
boron which diffuses to the grain boundaries on cooling, the N* / Nv ratio in equation (6-14)
can be written as:
(6-41)
Equation (6-41) predicts that the position of the ferrite C-curve in temperature-time space
depends on the l[B]gb Z)7) I l[B]*gb Z>y J ratio, as shown schematically in Fig. 6.24. Consequently,
this ratio can be regarded as a measure of the HAZ hardenability during welding of boron-
containing steels.
In Fig. 6.25 the microstructure data of Akselsen et al.42 have been replotted vs the
hardenability parameter [B]gbDy, taking the product [B]*gbDy in the reference steel equal to
unity for a direct comparison between theory and experiments. It is evident from the graph
that the HAZ martensite content of the two boron-containing steels can be represented by one
single curve under the prevailing circumstances. This result is to be expected if the displace-
ment of the ferrite C-curve in temperature-time space is determined by a relationship of the
type shown in equation (6-41).
log time
Fig. 6.24. Effect of boron alloying on the shape and position of ferrite C-curve in temperature-time space
according to the site blocking mechanism (schematic).
Martensite, vol%
Fig. 6.25. Relation between the volume fraction of martensite and the hardenability parameter
[B]gb Dy for simulated weld HAZs ([B]gb in ppm, Dy in urn, [B]gbDy = 1). Data from Akselsene£#/.42
Example (6.4)
Consider GMA welding on a thick plate of a boron-containing steel under the following con-
ditions:
/=300A, U = 30V, v = 4mm s-1, TI = 0.8, T0 = 200C
Suppose that the free (diffusible) boron content of the steel at elevated temperatures is 40
ppm. Estimate on the basis of the theory outlined in the previous sections the conditions for
ferrite/martensite formation in the grain refined HAZ (Tp = 11000C) when the austenite grain
size is 15|Lim. In these calculations we shall assume that the [B]*gb Dy product in the reference
steel (with thermodynamic properties as in Table 6.2) is close to unity.
Solution
First we estimate the amount of boron which segregates to the austenite grain boundaries
during cooling from equation (6-39). When Tp = 11000C (1373K), we get:
This gives the following value of the Af* / Nv ratio (equation 6-41):
The resulting displacement of the ferrite C-curve can now be calculated from equation (6-
14), using input data from Table 6.2:
As expected, the theoretical C-curves in Fig. 6.26 reveal a strong effect of boron alloying
on the HAZ transformation kinetics. By considering the superimposed weld cooling curve, it
is possible to estimate the volume fraction of ferrite f^~Fe which forms from equations (6-32)
and (6-33). Taking the time exponent n in the Avrami equation equal to 5/2 for nucleation of
ferrite at austenite grain boundaries and (1 - Xc) = 0.98 as in the previous examples, we ob-
tain after integration within the temperature range from 800 to 600° C:
and
If the same calculations are performed for the reference steel in Table 6.2 (characterised by
[*]gb = iBfgb )> we get:
and
From this we see that boron, even in small quantities, can have a dramatic effect on the
HAZ transformation kinetics by promoting the formation of bainite and martensite at the ex-
pense of grain boundary ferrite. This is in good agreement with general experience (see ex-
perimental CCT-diagrams in Fig. 6.27).
Time, s
Fig. 6.26. Predicted displacement of ferrite C-curve in temperature-time space due to segregation of
boron to prior austenite grain boundaries (Example 6.4). The superimposed cooling curve corresponds
to a cooling time, Ats/5, of 9.2s.
to enhance the nucleation rate of grain boundary ferrite at these sites due to the associated
increase in the Ae3-temperature. The above phenomenon should not be confused with equilib-
rium segregation of phosphorus to austenite grain boundaries during heat treatment of steel,
which stems form attraction of the atoms towards the open structure of the boundary. In the
latter case phosphorus may act as a hardenability element by occupying favourable sites for
ferrite nucleation at the austenite grain boundaries analogous to that documented for boron in
steel.43'44
Example (6.5)
Suppose that the local phosphorus content adjacent to the columnar austenite grain boundaries
in a low-alloy steel weld metal is 500 ppm, whereas the bulk concentration of phosphorus is
100 ppm. Estimate on the basis of the theory developed in the previous sections the resulting
displacement of the ferrite C-curve in temperature-time space when the columnar austenite
grain size is 80|jun. In these calculations we shall assume that the transformation characteristic
of the bulk metal is similar to that of the reference steel in Table 6.2.
Solution
First we estimate the actual Ae3-temperature within the phosphorus-rich region adjacent to the
columnar austenite grain boundaries from equation (6-36). Taking the A<?3-temperature of the
bulk phase equal to 835°C (1108K), the local phase boundary temperature becomes:
By substituting data from Table 6.2 into equation (6-35), we arrive at the following expres-
sion for ^ 0 :
(a)
Temperature, 0C
Steel A (11 ppmB)
Fig. 6.27. CCT-diagrams for boron-containing steels; (a) Steel A (llppm B), (b) Steel B (26ppm B).
Data from Akselsen et al.42
It is evident from the graphical representation of the above equation in Fig. 6.28 that the
observed increase in the yl^-temperature from 1108 to 1136K displaces the ferrite C-curve
towards higher temperatures and shorter times in the IT-diagram. The resulting effect on the
weld metal transformation kinetics is obvious, since an increase in the nucleation rate of fer-
rite will favour early site saturation at the austenite grain boundaries. On this basis it is not
surprising to find that allotriomorphic ferrite in low-alloy steel weld metals tends to form con-
tinuous veins of blocky ferrite along the columnar austenite grain boundaries, as shown in Fig.
6.19(c). Outside the fusion zone the conditions for early site saturation are less favourable,
since modern steelmaking practice implies that solidification-induced segregations are re-
moved by prolonged high-temperature annealing prior to the welding operation. Hence, fer-
rite veining of the type shown in Fig. 6.19(c) is not commonly observed within the reheated
regions of the base plate, unless the heat input is extremely large (see Fig. 6.19(a) and (b)).
Temperature, 0C
Time, s
Fig. 6.28. Effect of solidification-induced phosphorus segregations on the austenite to ferrite transforma-
tion in low-alloy steel weld metals (Example 6.5).
(6-42)
(6-43)
where C4 is a kinetic constant which is characteristic of the alloy system under consideration.
An indication of the applicability of equation (6-43) to welding can be obtained from the
micro structure data of Evans,45 reproduced in Fig. 6.29. In this plot, the reported heat inputs
have been converted into an equivalent cooling rate at 7000C via equation (1-71) (Chapter 1).
It is evident from the figure that the observed variation in the ferrite grain size with cooling
rate is consistent with calculations based on equation (6-43). Hence, the model of Umemoto et
Ferrite grain size,jim
Equation (6-43)
Example (6.6)
Consider multipass welding with covered electrodes (SMAW) on a thick plate of low-alloy
steel under the following conditions:
Previous experience has shown that polygonal ferrite forms within the low-temperature
reheated regions of the weld metal, typically 1.0 to 1.7 mm beneath the surface (fusion bound-
ary) of subsequent weld passes. Estimate on the basis of the grain growth diagram in Fig.
5.24(a) (Chapter 5) and equation (6-43) the maximum variation in the ferrite grain size across
the weld HAZ under the prevailing circumstances. Thermal data for low-alloy steels are given
in Table 1.1 (Chapter 1).
Solution
First we need to convert the depths at which polygonal ferrite appears beneath the fusion
boundary to an equivalent (characteristic) peak temperature range. If we neglect the contribu-
tion from heat flow in the welding direction, this conversion can be done on the basis of
equation (5-47) in Chapter 5:
Taking Ar*m equal to 1 and 1.7mm, respectively, we get:
and
The prior austenite grain size Dy at these two locations can now be read from Fig. 5.24(a)
and Table 5.6, respectively:
Since the cooling rate (CR.) at a given temperature is essentially the same across the weld
HAZ, the maximum variation in the ferrite grain size can be evaluated directly from equation
(6-43) without further background information:
From this we see that the variation in the ferrite grain size is significantly smaller than the
corresponding change in the prior austenite grain size. This result is in good agreement with
general experience.
(6-44)
where E1 is the one-dimensional parabolic thickening constant (defined in equation (6-18)).
The corresponding growth rate of allotriomorphic ferrite Ua is then obtained by differenti-
ating equation (6-44) with respect to time:
(6-45)
From a stereological standpoint, the grain boundary surface area per unit volume Sv cannot
be calculated without further assumptions regarding the shape of the columnar austenite grains.
Bhadeshia et al.31*46*49 solved this problem by representing the grain morphology by a uniform,
space-filling array of hexagonal prisms. However, with the precision aimed at here, it is suffi-
cient to assume that Sv is equal to the surface area per unit volume of an inscribed cylinder
whose volume is equivalent to that of the hexagonal prisms (see Fig. 6.30). Noting that each
grain boundary is shared by two adjacent grains, the expression for Sv (in the absence of end
effects) becomes:47'48
(6-46)
where dy and I1 are the diameter and length of the inscribed cylinder, respectively.
By substituting the above expressions for Ua and Sv into equation (6-34), we obtain:
(6-47)
Fig. 6.30. Hexagonal prism model for the columnar austenite grain morphology in low-alloy steel weld
metals. The inscribed cylinder in the figure has approximately the same surface to volume ratio as the
hexagonal prisms.
Parabolic thickening constant (E1), ^m/s1/2
Temperature, 0C
Fig. 6.31. Effect of temperature and steel chemical composition on the one-dimensional parabolic thick-
ening constant for allotriomorphic ferrite. (1): 0.03 wt% C, (2): 0.06 wt% C, (3): 0.08 wt% C and (4):
0.10 wt% C. Data from Bhadeshia et al.46
(6-48)
from which
(6-49)
Equation (6-49) predicts that the volume fraction of grain boundary ferrite in the as-depos-
ited weld metal depends on the combined action of the following three main variables:
(i) The one-dimensional parabolic thickening constant ei which is determined by the weld
metal chemical composition (i.e. the content of austenite and ferrite stabilising elements).
(ii) The columnar austenite grain size dy which is controlled by the weld metal solidification
microstructure (i.e. the weld metal chemistry, the weld pool geometry, and the thermal
conditions under which solidification occurs).
(iii) The retention time within the critical transformation temperature range for allotriomorphic
ferrite, as determined by the applied heat input and the mode of heat flow (i.e. thick plate,
medium thick plate, or thin plate welding, respectively).
It can be seen from the microstructure data of Grong et al.50 reproduced in Fig. 6.32 that the
influence of these variables are adequately accounted for in the present model. However, the
calculated volume fractions of grain boundary ferrite are consistently lower than the measured
ones. This discrepancy can probably be attributed to the use of constant values for the one-
dimensional parabolic thickening constant. Consequently, if proper corrections are made for
the inherent variation in E1 with temperature and steel chemical composition, the agreement
between theory and experiments is significantly improved, as shown by the data of Bhadeshia
et al.46 reproduced in Fig. 6.33.
Example (6.7)
Consider GMA welding on a thick plate of a low-alloy steel under the following conditions:
Suppose that the grain boundary allotriomorphs form within the temperature range from
750 to 6000C at a constant rate ei of 3 jam s~1/2. Estimate on the basis of the theory developed
in the previous section the volume fraction of allotriomorphic ferrite in the weld deposit when
the columnar austenite grain size is 80|Lim. In these calculations we shall assume that the
equilibrium volume fraction offerritef®q~Fe in the fully transformed steel is 0.9. Thermal data
for low-alloy steels are given in Table 1.1 (Chapter 1).
Solution
The situation is described in Fig. 6.34. In this case the problem is to estimate the retention time
/Str within the critical temperature range for ferrite formation from the Rosenthal thick plate
solution (Chapter 1). From equation (1-66), we have:
Regression line
Volume fraction of GF
x * . ^ )
Fig. 6.32. Experimental verification of equation (6-49). Data from Grong et al.5Q
C-Mn steel weld metals
Calculated volume
fraction of GF
Fig. 6.33. Comparison between measured and predicted volume fractions of grain boundary ferrite in
C-Mn steel weld metals. Data from Bhadeshia et al.46
Temperature
log time
Fig. 6.34. Conditions for allotriomorphic ferrite formation in low-alloy steel weld metals (Example 6.6).
By inserting this value into equation (6-49), we obtain:
Although experimental data are not available for a direct comparison, the calculated value
of fa'Fe is reasonable and of the expected order of magnitude.
*A different view is suggested by Bhadeshia et al.31'46'51'52 who claim that growth of Widmanstatten ferrite occurs in
a displacive manner analogous to that documented for martensite in steel, with the exception that carbon must diffuse
during growth.
Next Page
a b
Growth rate of WF, u,m/s
Temperature, 0C
Fig. 6.35. Predicted growth rates of Widmanstatten ferrite in C-Mn steel weld metals; (a) Growth rate
calculations for weld compositions listed in Fig. 6.31, (b) Growth rate calculations after modifying com-
positions to allow for carbon enrichment due to grain boundary ferrite formation. Data from Bhadeshia
etai46
a b
Growth rate of WF, u,m/s
Temperature, 0C
Fig. 6.35. Predicted growth rates of Widmanstatten ferrite in C-Mn steel weld metals; (a) Growth rate
calculations for weld compositions listed in Fig. 6.31, (b) Growth rate calculations after modifying com-
positions to allow for carbon enrichment due to grain boundary ferrite formation. Data from Bhadeshia
etai46
Calculated volume
fraction of WF
Fig. 6.36. Comparison between measured and predicted volume fractions of Widmanstatten ferrite in C-
Mn steel weld metals. Data from Bhadeshia et al.46
*The orientation relationship can alternatively be described by the Nishiyama-Wasserman (N-W) correspondence
which also lies within the Bain orientation region. However, since the K-S and the N-W orientation relationships
only differ from each other by a 5.26° rotation of the close packed planes, they can be regarded as equivalent.
(a) (b)
Fig. 6.37. The development of transformation textures in as-deposited steel weld metals containing acicular
ferrite; (a) (200) pole figure showing the crystallographic orientations of acicular ferrite referred to the
original cell/dendrite growth direction, (b) Backscattered electron channeling contrast image of delta
ferrite/austenite columnar grain (the metallographic section is normal to the cell/dendrite growth direc-
tion). After Kluken et al.63
parallel with the cell/dendrite growth direction (Fig. 6.38(a)). The other plates have a <111>
direction aligned in the same crystal growth direction (Fig. 6.38(b)). These data can be repre-
sented by two sub-components which are displaced with respect to each other by a 60° rotation
about a common <111> axis.
It is evident from the measurements of Kluken et al.63 that the acicular ferrite plates in as-
deposited steel weld metals exhibit an orientation relationship with both the austenite and the
prior delta ferrite columnar grains in which they grow. This 'memory' effect arises from the
characteristic solidification pattern and transformation behaviour of low-alloy steel welds (e.g.
see discussion in Section 3.8.2, Chapter 3). As shown schematically in Fig. 6.39, the columnar
grain region will exhibit a sharp <100> solidification texture which has its origin in the phe-
nomenon of preferred crystal growth. At the onset of the peritectic reaction, the austenite
adopts a K-S type of orientation relationship with the delta ferrite in order to minimise the
energy barrier against nucleation.64 The austenite subsequently grows around the periphery of
the primary phase until impingement occurs on neighbouring columnar grain boundaries. During
the yFe to aFe transformation, this memory effect gives rise to the formation of acicular ferrite
plates which have a <100> direction approximately parallel with the original cell/dendrite
growth direction. The presence of the two other texture components within the weld metal is
thus a result of complementary crystal rotations taking place within the same orientation re-
gion.
The proposed sequence of reactions is in excellent agreement with the texture analysis of
Hu65 who made theoretical calculations of the resulting orientations of iron after §Fe to yFe and
yFe to aFe transformations in succession according to the scheme outlined in Fig. 6.39.
(a) (b)
Fig. 6.38. Schematic diagrams showing the three main texture components in acicular ferrite according
to the Kurdjumow-Sachs orientation relationship; (a) The <100> texture component, (b) The two com-
plementary <111> texture components. After Kluken et a/.63
fully understood. However, detailed TEM studies performed by Bhadeshia et aL37>59>62>66 have
clearly demonstrated that acicular ferrite is a form of intragranularly nucleated bainite. In
practice, this means that the microconstituent may be present either as 'upper' or 'lower'
acicular ferrite in the weld deposit (depending on the carbon concentration), as shown
schematically in Fig. 6.40.
In general, the ferrite component of upper bainite is composed of groups of thin parallel
laths (subplates) with a well-defined crystallographic habit.2061 Although the growth mecha-
nism of upper bainite is still a subject of considerable controversy, it has been postulated that
the subplates advance into the austenite with their own tip configurations. One model is shown
in Fig. 6.41, where each subplate forms as a ledge upon the adjacent subplates through a
nucleation and growth process. These ferrite laths possess the same variant of the K-S orien-
tation relationship, which means that they are separated by low-angle grain boundaries. A
typical austenite grain will contain numerous sheaves of bainitic ferrite exhibiting different
variants of the K-S orientation relationship. This implies that the boundary between adjacent
plates of acicular ferrite should alternately be of the low-angle and high-angle type, a feature
which also has been observed experimentally (see data in Fig. 6.42). Hence, both the mor-
phology and the crystallography of acicular ferrite bear a close resemblance to upper bainite.
Columnar grain
Fig. 6.39. Schematic diagram showing the sequence of reactions occurring during cooling of a low-alloy
steel weld through the critical transformation temperature ranges. After Kluken et al.63
Carbide precipitation
from austenite
Fig. 6.40. Schematic illustration of the transition from 'upper' to 'lower' acicular ferrite in low-alloy
steel weld metals. The diagram is based on the ideas of Bhadeshia and Svensson.37
Successive
nucleation and
growth of
parallel plates
Fig. 6.41. Proposed model for nucleation and growth of upper bainite in steel (schematic). After
Verhoeven.61
Misorientation, degrees
Plate number
Fig. 6.42. Measured spatial misorientation between adjacent plates of acicular ferrite in a low-alloy steel
weld. Data from Kluken et a/.63
A more realistic approach would be to estimate the volume fraction of acicular ferrite
via the equation:37
(6-50)
where f%FFe and f%FFe are the corresponding volume fractions of grain boundary ferrite and
Widmanstatten ferrite, respectively (note that in equation (6-50) the formation of microphases
has been disregarded).
The method outlined above has shown to work well for numerous welds (e.g. see Fig. 6.43),
but fails when the primary microstructure consists of a mixture of acicular ferrite and martensite,
as is the case in high strength steel weld deposits6768 In spite of this shortcoming, equation (6-
50) expresses in an explicit manner the real essence of the problem, namely that the evolution
of the acicular ferrite microstructure depends on the interplay between several competing nu-
cleation and growth processes which occur consecutively during cooling from the Ae3-tem-
perature. This important point is often overlooked when discussing the conditions for acicular
ferrite formation in low-alloy steel weld metals.
Allotriomorphic
ferrite
Volume fraction
Widmanstatten
ferrite
Acicular
ferrite
Inclusion
het.
, Austenite
grain boundary
is within the typical size range of most weld metal inclusions (see Figs. 2.57-2.61 in Chapter
2). This finding is in excellent agreement with the results of Barbaro et al.70 reproduced in Fig.
6.45, showing that a certain minimum inclusion size (say 0.2-0.3 jam) is required for acicular
ferrite nucleation in steel weld deposits.
It should be noted, however, that Ricks et al69 omitted a consideration of the effects of
plastic strain produced as a result of differences in thermal contraction between the austenite
and the particles as well as the possibility for the ferrite to adopt reasonable orientation rela-
tionships with both the austenite and the catalyst particles. Based on nucleation theory it can
be argued that these factors will influence the transformation process.1 This, in turn, may
explain why certain types of inclusions appear to be more favourable nucleation sites for acicular
ferrite than other (to be discussed below).
(i) Nucleation resulting from a small lattice disregistry between the inclusions and the
ferrite.
(iii) Nucleation in the vicinity of inclusions resulting from favourable strain or dislocation
arrays due to differences in the thermal contraction between the particles and the
matrix.
Probability of nucleation
(6-51)
Imagine now that the inclusion lattice is rotated through 360° about the <100>raxis which
lies within a <100> y . Fe -region. The other two <100> r poles will then lie within
<110>7.Fe-regions for at the most 4 X 22° = 88° of this rotation, since the diameter of the
<100>7_Fe-regions is 22°. Hence, the probability P2 that, if one <100> r pole lies within a
<100>7.Fe-region, the other two will lie within <110>7.Fe-regions is given in the upper limit
by:
(6-52)
The total probability that a given orientation relationship lies within the Bain region purely
by chance is thus:
(6-53)
Therefore, assuming random orientation, about 4% of the weld metal inclusions would lie
within the Bain region purely by chance if they were single phase cubic crystals. In practice,
however, inclusions commonly found in low-alloy steel weld metals are of a very heterogene-
ous chemical and crystalline nature. As shown in Fig. 2.72 in Chapter 2, a typical inclusion
may contain up to six different constituent phases, including the three cubic phases 7-Al2O3,
MnOAl2O3, and TiN. This implies that at least 12% of the inclusions may contain a cubic
phase which lies within the Bain orientation region.
Measurements of orientation relationships between specific inclusion constituent phases
(i.e. 7-Al2O3, MnOAl2O3, and TiN) and contiguous acicular ferrite plates performed by Grong
et a/.85 support the above interpretation (see Table 6.3). Referring to the standard stereographic
projections in Fig. 6.47, a very high proportion of the ferrite/inclusion orientations falls within
the Bain region. In addition, two other variants (i.e. No. 5 and 6) are indicated for TiN which
do not meet this requirement. They are therefore regarded as spurious (in the sense that the
observed orientation relationships do not stem from a catalyst nucleation event) and should be
ignored. Hence, it may be concluded that the observed orientation relationships between acicular
ferrite and specific inclusion constituent phases are not fully reproducible in the true meaning
of the word, since only those combinations which satisfy the inherent crystallography of the
acicular ferrite microstructure are acceptable.
An interesting observation from the data in Table 6.3 is that nucleation of acicular ferrite on
inclusions is always associated with low-index planes of the {100} or the {110} type, which
indicates a faceted growth morphology of the inclusions. Faceted growth may occur as a result
of anisotropy in the growth rates between high-index and low-index crystallographic planes,
and can in the extreme case lead to a morphology of the type shown in Fig. 6.48. Conse-
quently, formation of faceted inclusions in the liquid steel during deoxidation appears to be an
intrinsic feature of low-alloy steel weld metals.
Simple verification on the basis of classic nucleation theory shows that the associated re-
duction of the energy barrier to nucleation, AGlet , is the primary cause for the ferrite nucleus
to develop epitaxial orientation relationships with the substrate and the austenite.8687 Refer-
ring to Fig. 6.49, a qualitative ranking of the different inclusion constituent phases with respect
to nucleation potency of acicular ferrite can be made from the data presented in Table 6.3. It is
evident that both 7-Al2O3, MnOAl2O3, and TiN reveal a good lattice matching with the ferrite
phase in one crystallographic direction. In addition, nucleation of acicular ferrite at TiN offers
Table 6.3 Observed orientation relationships between acicular ferrite and different inclusion constitu-
ent phases in a SA low-alloy steel weld. Data from Grong et al.85
Substrate Orientation Relationship Interplanar Spacing
(s) Variant No. Plane Combinations Ratio^
the advantage of partial lattice coherence in a second (independent) direction, which further
contributes to a reduction of AG^e/ through a minimisation of the interfacial energy between
the two phases. This makes TiN an extremely efficient nucleant for acicular ferrite.
Microstructure data available for submerged arc (SA) steel weld deposits clearly support
the above findings that nucleation of acicular ferrite occurs preferentially at inclusions which
contain aluminium or titanium. As shown in Fig. 6.50, a high volume fraction of acicular
ferrite is always achieved when sufficient amounts of titanium are added either through the
filler wire or the flux, irrespectively of the aluminium and oxygen concentrations. This is in
sharp contrast to welds produced with welding consumables containing low levels of titanium,
where the acicular ferrite content drops rapidly with decreasing [A%Al]weld/[%O]anal ratios
Fig. 6.48, Example of a faceted crystal delimited by {100} and {110} planes (schematic).
MnOAI2O3
T-Al2O3
TiN
Nucleation site
Fig. 6.49. Qualitative ranking of different inclusion constituent phases with respect to nucleation po-
tency of acicular ferrite.
due to the presence of lower fractions of 7-Al2O3 and MnOAl2O3 in the inclusions (see Fig.
2.72 in Chapter 2). Similar observations have also been made by other investigators.3672"
76,78,79
It should be noted that the weld metal transformation behaviour in practice depends on
complex interactions between a number of important variables, including alloying and
deoxidation practice, the solidification microstructure, the prior austenite grain size, and the
weld thermal cycle.36'5358 This means that the presence of 7-Al2O3, MnOAl2O3 or TiN at the
surface of the inclusions is perhaps a necessary but not a sufficient criterion for formation of
acicular ferrite in steel weld metals.
Example (6.8)
Consider a partly Ti-Al deoxidised steel weld metal which contains a total number of 4 X 107
inclusions per mm 3 . Based on Fig. 2.72 in Chapter 2 and the theory developed above, estimate
an upper limit for the volume of a typical plate of acicular ferrite when the weld metal [A%Al]weld/
\y°O\anaL r a t i o i s °- 80 -
Vol% AF
[A%A|lweld/[%o]anal
Fig. 6.50. Effect of deoxidation practice (inclusion chemistry) on the acicular ferrite transformation in
low-alloy steel weld metals. Data compiled from miscellaneous sources.
Solution
From Fig. 2.72 it is seen that the total number of constituent phases in the inclusions is six,
including the three cubic phases 7-Al2O3, MnOAl2O3, and TiN. If we assume a random ori-
entation and only one nucleation event per inclusion, the following upper limit for the acicular
ferrite plate volume is obtained:
The above volume corresponds to an acicular ferrite plate which has the shape of a square
lath of side lOjim and thickness 2|im. Although this estimate is in reasonable agreement with
experimental observations,37 the prediction is conservative in the sense that it assumes only
one nucleation event per inclusion. In practice, an oxide inclusion which is orientated within
the Bain region has the capability of nucleating several acicular ferrite plates, as shown by the
SEM micrograph in Fig. 6.51. In addition, the acicular ferrite plates may nucleate
autocatalytically at aFe /yFe boundaries, a process which also is referred to as sympathetic
nucleation in the literature.36'37'7083 At present, it is not clear to what extent autocatalytic
nucleation plays a role in the development of the acicular ferrite microstructure.
Hardenability effects
Since acicular ferrite is one of the last phases to form after the growth of allotriomorphic and
Widmanstatten ferrite, it is bound to be influenced by the prior transformation products, as
indicated by equation (6-50). The strong dependence of the acicular ferrite content on the
austenite grain size must therefore be understood on this basis. 3747 ' 48 ' 7088 " 90 It is evident from
the data of Barbaro et al.70 reproduced in Fig. 6.52 that a coarse austenite grain size favours
intragranular nucleation of acicular ferrite at the expense of formation of allotriomorphic and
Fig. 6.51. SEM micrograph of a carbon extraction replica showing evidence of multiple nucleation of
acicular ferrite at a weld metal inclusion.
Widmanstatten ferrite. This effect is most pronounced during slow cooling, since the combi-
nation of a small austenite grain size and a slow cooling rate implies that much of the yFe-
phase already has transformed to allotriomorphic ferrite before the temperature for intragranular
nucleation of acicular ferrite is reached.
Example (6.9)
Consider a low-alloy steel weld metal which contains a total number of 5 X 107 inclusions per
mm3 with an average radius of 0.25 Jim. Use this information to evaluate the conditions for
acicular ferrite formation within the as-deposited weld metal and the low-temperature reheated
region of the weld when the austenite grain size is 100 and 10 |iim, respectively. In these
calculations we shall assume that the equilibrium volume fraction of ferrite ff~Fe in the fully
transformed steel is 0.9.
Solution
First we need to estimate the total surface area per unit volume available for ferrite nuclea-
tion at austenite grain boundaries, Sv (GB), and non-metallic inclusions, Sv (/), respectively.
(as before)
Volume fraction of AF
Fig. 6.52. Effect of austenite grain size on the acicular ferrite transformation in low-alloy steel weld
deposits. Data from Barbara et al.70
From the above calculations it is apparent that the conditions for intragranular nucleation of
acicular ferrite at inclusions are particularly favourable within the as-deposited weld metal,
since SV(I) > SV(GB), whereas nucleation and growth of ferrite at austenite grain boundaries
will dominate within the reheated region of the weld (SV(GB)» SV(I)). This conclusion is also
consistent with predictions based on the Avrami equation (equation (6-49)). If we, as an illus-
tration, assume that the volume fraction of grain boundary ferrite in the as-deposited weld
metal is 0.3, the corresponding fraction of GF in the reheated weld metal becomes:
On this basis it is not surprising to find that the microstructure within the grain refined
region of low-alloy steel welds is usually polygonal ferrite, while the as-deposited weld metal
also contains high proportions of acicular ferrite (see Fig. 6.19(c) and (d), respectively).
It is important to realise, however, that the presence of allotriomorphic ferrite at the austenite
grain boundaries has the benefical effect of suppressing the formation of bainitic sheaves at the
austenite grain boundaries, which, in turn, allows the acicular ferrite to develop on intragranular
nucleation sites.9091 Consequently, due to the number of competing nucleation and growth
processes involved, the volume fraction of acicular ferrite is often seen to pass through a maxi-
mum when the weld metal hardenability is successively increased by additions of alloying
elements, as shown in Fig. 6.53.
Microstructural component, %
GF and PF
WF Fertile with aligned
second phase (FS)
63.7 Bainite
Bainitic microstructures (besides acicular ferrite) are frequently observed during welding, par-
ticularly in the HAZ of low-carbon microalloyed steels,22'93"95 but also within the fusion re-
gion of the weld if the nucleation conditions are favourable.37'55'5991 Two main forms can be
identified, i.e. upper and lower bainite, as indicated in Fig. 6.19(a).
WF/UB
HAZ
Fig. 6.54. Schematic illustration of the HAZ transformation behaviour during high heat input welding;
(a) Ti-oxide containing steel, (b) Conventional Al-Ti microalloyed steel. The diagrams are based on the
ideas of Homma et al.92
the classic theory of martensite nucleation assuming a pure invariant plane strain deforma-
tion.1'20'61
Ferrite with
iUBj aligned second - WF
dhase (FS)
Mill
Cooling time, At 8/5 , s
Fig. 6.55. Effect of cooling time, Af8/5, on the grain coarsened HAZ transformation behaviour (simulated
thick plate heat cycles, Tp « 13500C). Data from Akselsen et al96
Fig. 6.56. Optical micrograph showing formation of upper bainite within the columnar grain region of a
SA steel weld.
shift in the weld metal transformation behaviour is related to a change in the deoxidation
practice which alters the kinetics of the subsequent solid state transformation reactions through
a modification of the solidification microstructure.64 Solidification induced phosphorus
segregations are of particular interest in this respect, since previous examinations have shown
that phosphorus can strongly enhance the formation of grain boundary ferrite by raising the
local Ae3 temperature of the steel97 (see also Example 6.5).
Following the discussion in Section 6.3.5.4, the evolution of allotriomorphic ferrite at the
austenite grain boundaries has the beneficial effect of suppressing the formation of bainitic
sheaves at these sites, which, in turn, allows the acicular ferrite to develop on intragranular
Microstructural component, %
Carbides
Lower bainite
Martensite
Fig. 6.58. TEM micrograph showing the formation of lower bainite within the HAZ of a low-carbon
microalloyed steel.
6.3.8 Martensite
At very high undercoolings, the austenite decomposes to martensite by means of an invariant
plane strain deformation, which implies that there is no change in the steel chemical composi-
tion. The reaction product will either be lath or plate (twinned) martensite, depending on the
alloy level. Lath martensite is commonly found in plain carbon and low-alloy steels up to
about 0.5wt% C, and is formed by a slip mechanism, as shown schematically in Fig. 6.59(a).
When the carbon content exceeds this threshold, the martensite transformation occurs rather
by formation of deformation twins (Fig. 6.59(b)). The crystal structure of plate martensite is
bet (body-centred tetragonal), while lath martensite reveals a bcc (body-centred cubic) struc-
ture,61 which becomes increasingly distorted with increasing steel carbon contents.20 Both
transformation products exhibit the characteristic Kurdjumow-Sachs orientation relationship
with the austenite, but this relationship tends to be less precise at high carbon levels.1'20
a b
Fig. 6.59. Mechanisms of martensite formation in steel (schematic); (a) Slip along parallel planes, (b)
Generation of deformation twins. The diagram is based on the ideas of Verhoeven.61
Steel A
Steel B
Martensite content, vol%
Steel C
(i) Rapid growth of austenite into pearlite until the dissolution process is completed.
(ii) Slower growth of austenite into ferrite at a rate which is either controlled by carbon
diffusion in the austenite or by diffusion of substitutional elements such as manga-
nese in the ferrite, depending on the applied annealing temperature.
T T
t wt% C
Fig. 6.61. Schematic illustration of the formation of plate (twinned) martensite within the intercritical
HAZ of low-carbon microalloyed steels; Heating leg of thermal cycle: Pearlite -> -yFe (>0.5 wt% C),
Cooling leg of thermal cycle: yFe(> 0.5 wt% C) -> a'Fe (twinned martensite).
(iii) Very slow final equilibration of solute concentration gradients within the austenite or
the ferrite through diffusion.
At temperatures above say 7700C, the rate of pearlite decomposition is sufficiently high for
complete dissolution, even during low-heat input welding. Hence, in hot rolled and normal-
ised steels all pearlite will transform to austenite when the peak temperature of the thermal
cycle exceeds this threshold. However, if the starting microstructure is a mixture of martensite
and upper bainite (as often will be the case in multi-pass welding), the austenite will nucleate
both at prior austenite grain boundaries and along the interfaces between laths of bainite or
martensite, as shown by the optical micrographs in Fig. 6.62. A similar pattern has also been
observed during intercritical annealing of dual-phase steels.104'105 Considering the kinetics,
carbides precipitated within autotempered low-carbon martensite or between the laths of bainite
would be expected to dissolve at a rate comparable with that of pearlite.105 Hence, it is reason-
able to assume that no retained carbides will be present intragranularly after reheating to say
Tp « 7700C when the starting microstructure is a mixture of martensite and bainite.
Further growth of the austenite into the ferrite requires, however, that the peak temperature
of the thermal cycle is raised significantly above 7700C in order to reach the kinetic
(paraequilibrium) stage where the reaction is controlled by diffusion of carbon in austenite."
Consequently, the possibilities of obtaining growth of the austenite colonies within the low
peak temperature regions of the intercritical HAZ are strongly limited, which implies that the
carbon content of the austenite islands in these regions should be close to the saturation level
of carbon in austenite at the temperature of dissolution.
(b)
Fig. 6.62. Optical micrographs showing favourable sites for austenite formation during two-pass weld
thermal simulation. (First cycle: Tp « 13500C, Atm « 12 s, Second cycle: Tp « 775°C, Atm « 12 s);
(a) Intergranular, (b) Intragranular. After Akselsen et aim
metal pearlite bands or within isolated pearlite colonies at high cooling rates, as shown by the
TEM micrographs in Fig. 6.63(a) and (b). With decreasing cooling rates, however, the trans-
formation product shifts from twinned martensite to predominatly pearlite (see Fig. 6.63(c)
and (d)) in the absence of strong hardenability elements such as molybdenium and boron which
can stabilise the M-A constituent.102'107"109
In controlled rolled and accelerated cooled steels, where the carbides are mainly present in
the form of submicroscopic colonies or isolated particles located at ferrite/ferrite grain boundaries
(see Fig. 6.64(a)), the situation becomes slightly more complex. Under such conditions, the
majority of the austenite islands formed within the low peak temperature region of the
intercritical HAZ will be of a size below 1 jam. It has been verified experimentally that such
(a) (b)
(C) (d)
Fig. 6.63. Examples of transformation products formed after intercritical thermal cycling of a ferritic/
pearlitic starting microstructure to a peak temperature of 775°C; (a) M-A islands (black areas) sur-
rounded by ferrite (At6/4 ~ 5 s), (b) Close-up of twinned martensite within a M-A island (At6/4 ~ 5 s),
(c) Isolated pearlite colony formed at an intermediate cooling rate (Ar674 ~ 12 s), (d) Pearlite colonies
formed during slow cooling (At614 ~ 35 s). After Akselsen et al.103
small particles do not readily transform to martensite, but will largely remain in the steel in the
form of retained austenite.100'103'110'111 This can be explained by the lack of nucleation oppor-
tunities for martensite or by the volume restraint provided by the surrounding ferrite matrix.1 x x
Consequently, twinned martensite in controlled rolled and accelerated cooled steel is seen to
form within a few, relatively large austenite islands which stem from decomposition of single
carbide colonies, as illustrated by the TEM micrograph in Fig. 6.64(b).
During multipass welding, the initial carbide distribution in the base plate will be of less
importance because of the transformations imposed by the heat of previous weld passes. Hence,
after full reaustenitising of the steel the M-A constituent may form both intergranularly and
intragranularly, depending on the starting microstructure, as shown previously in Fig. 6.62(a)
and (b). An example of intergranularly nucleated twinned martensite is contained in Fig. 6.65.
(a)
(b)
Fig. 6.64. Conditions for martensite formation within the intercritical HAZ of a controlled rolled and
accelerated cooled steel containing copper and nickel; (a) TEM micrograph of the initial base plate car-
bide distribution, (b) TEM micrograph showing evidence of twinned martensite within a l(im large
austenite colony formed during intercritical thermal cycling to T « 775°C (Ar674 « 35 s). After Akselsen
etal^
Simple austenitic stainless steels generally contain between 18 and 30 wt% chromium, 8 to 20
wt% nickel, and between 0.03 to 0.1 wt% carbon.20 Since the solubility of carbon decreases
rapidly with temperature (see Fig. 6.67), reheating of the steels within the temperature range
from 500 to 9000C will lead to the rejection of carbon from solid solution, usually by chro-
mium carbide precipitation (Cr23C6). These carbides nucleate preferentially at the austenite
grain boundaries, which, in turn, results in depletion of the regions adjacent to the grain bounda-
ries with respect to chromium, as shown schematically in Fig. 6.68. The presence of such
chromium-depleted regions within the steel makes it sensitive to intergranular corrosion in
service.20'112
Previous Page
(a)
(b)
Fig. 6.64. Conditions for martensite formation within the intercritical HAZ of a controlled rolled and
accelerated cooled steel containing copper and nickel; (a) TEM micrograph of the initial base plate car-
bide distribution, (b) TEM micrograph showing evidence of twinned martensite within a l(im large
austenite colony formed during intercritical thermal cycling to T « 775°C (Ar674 « 35 s). After Akselsen
etal^
Simple austenitic stainless steels generally contain between 18 and 30 wt% chromium, 8 to 20
wt% nickel, and between 0.03 to 0.1 wt% carbon.20 Since the solubility of carbon decreases
rapidly with temperature (see Fig. 6.67), reheating of the steels within the temperature range
from 500 to 9000C will lead to the rejection of carbon from solid solution, usually by chro-
mium carbide precipitation (Cr23C6). These carbides nucleate preferentially at the austenite
grain boundaries, which, in turn, results in depletion of the regions adjacent to the grain bounda-
ries with respect to chromium, as shown schematically in Fig. 6.68. The presence of such
chromium-depleted regions within the steel makes it sensitive to intergranular corrosion in
service.20'112
Fig. 6.65. TEM micrograph showing evidence of intergranularly nucleated twinned martensite in dou-
ble-cycled specimens (First cycle: Tp ~ 13500C, A;6/4 -12s, Second cycle: Tp - 775°C, A;6/4 - 12s) After
Akselsen et al.103
Content of M-A constituent, vol%-
Fig. 6.66. Effect of base plate pearlite content on the volume fraction of M-A constituent in thermally
cycled specimens (simulation conditions: Tp -775°C, Atm « 5s). Data from Akselsen et al.m
Temperature, 0C
Fig. 6.67. Solvus temperatures for different types of carbides in 304 austenitic stainless steel as a func-
tion of carbon content. Data from Kou.112
Cr carbide precipitate
Fig. 6.68. Schematic illustration OfCr23C6 precipitation at grain boundaries in austenitic stainless steels.
The diagram is based on the ideas of Kou.112
6.4.1 Kinetics of chromium carbide formation
Precipitation of Cr23C6 in austenitic stainless steels is a typical nucleation and growth process.
The temperature-time transformation curve is therefore C-shaped, and can be modelled ac-
cording to the general principles outlined in Section 6.2.2.3. As an illustration, we shall as-
sume that the parameters listed in Table 6.4 are representative of nucleation of Cr23C6 in a 304
type austenitic stainless steel containing 0.07 wt% C. If we at the same time allow for the
variation in austenite grain size with distance from the fusion boundary, the shape and position
of the C-curve in temperature-time space at different locations within the HAZ is given by
equation (6-35):
(6-54)
A graphical representation of equation (6-54) in Fig. 6.69 shows that the nose of the
C-curve for the reference steel (characterised by Teq = T*eq and D1 = D 7 ) is strongly shifted to
the left in the IT-diagram, thereby providing favourable conditions for Cr23C6 precipitation in
the heat affected zone during welding. The problem becomes less imminent if the base plate
carbon content is reduced from say 0.07 to 0.04 wt%. In that case the associated reduction in
the solvus temperature Teq from 920 to 8000C will displace the C-curve towards longer times
in the diagram which by far exceed the duration of a typical weld thermal cycle, as shown in
Fig. 6.69.
Example (6.10)
Consider single pass butt welding of 2mm sheets of 304 austenitic stainless steels with cov-
ered electrodes under the following conditions:
Evaluate on the basis of equations (6-32), (6-33), and (6-54) in combination with the grain
growth diagram in Fig. 5.30(b) (Chapter 5) the conditions for chromium carbide formation
within the HAZ during welding when the base plate carbon content is 0.07 and 0.04 wt% C,
respectively. Thermal data for austenitic stainless steels are given in Table 1.1 (Chapter 1).
Temperature, 0C
Time, s
Fig. 6.69. Effect of carbon content on the isothermal precipitation of C ^ C 6 in 304 austenitic stainless
steels. Broken curve: 0.07 wt% C, Solid curve 0.04 wt% C. The diagrams are constructed on the basis
of equation (6-54).
T
Table 6.4 Input data used to construct C-curve for Cr23C6 precipitates in reference steel.
Age-hardenable Al-Mg-Si alloys are widely used as structural components in welded assem-
blies. They offer tensile strength values higher than 350 MPa in the artificially aged T6 condi-
tion owing to the presence of very fine, needle shaped (3"(Mg2Si) precipitates along <100>
directions in the aluminium matrix.113 Although Al-Mg-Si alloys are readily weldable, they
suffer from severe softening in the heat affected zone (HAZ) because of reversion (dissolu-
tion) of the 3"(Mg2Si) precipitates during the weld thermal cycle. 6112 This type of mechanical
impairment represents a major problem in engineering design.114
Unaffected base metal
Peak temperature, 0C
Fig. 6.71. Conditions for Cr23C6 precipitation within the HAZ of a single pass austenitic stainless steel
butt weld (Example 6.10).
6.5.1 Quench-sensitivity in relation to welding
High strength alloys such as AA 6082 contains manganese in addition to magnesium and
silicon. Manganese is added to control recrystallisation and grain growth in the material dur-
ing hot forming. The disadvantage is that it increases the quench sensitivity of the alloy.115
The reason for this is that the Mn-bearing dispersoids (which form during homogenisation)
provide favourable nucleation sites for the non-hardening (3'(Mg2Si) phase, as shown by the
TEM micrograph in Fig. 6.72. The resulting reduction in the solute content leads to a reduced
HAZ strength both in the naturally aged TA and peak aged T6 conditions. 56116
Fig. 6.72. TEM micrograph showing nucleation of non-hardening (3'(Mg2Si) precipitates at Mn-bearing
dispersoids in an AA 6082 alloy.
Table 6.5 Input data used to construct C-curve for p' (Mg2Si) precipitates in reference aluminium alloy.
Parameter Tr t\ T*q. A0 Qj S^
A graphical representation of equation (6-55) in Fig. 6.73 shows that the nose of the C-
curve for the reference alloy (characterised by Teq = r and Sv = S^ is strongly shifted to
the left in the IT-diagram, thereby providing favourable conditions for (3'(Mg2Si) formation
dunng welding. In general, an increase in T or Sv will enhance the quench-sensitivity of the
material because of the resulting increase in the nucleation rate. This will be the case if the
alloy contams large amounts of excess silicon in solid solution or is homogenised at a tem-
perature lower than 5800C.118119
Example (6.11)
Consider plasma-MIG butt welding of a 10 mm thick Al-Mg-Si plate under the following
conditions:
qo = 10 kW, v = 10 mm s~\ T0 = 200C
Evaluate on the basis of equations (6-32), (6-33), and (6-55) the conditions for P'(Mg,Si)
precipitation within the high peak temperature region of the HAZ during welding (T7 > T )
In these calculations we shall assume that the transformation behaviour of the base'metafis
similar to that of the reference alloy in Table 6.5. Relevant thermal data for Al-Mg-Si alloys
are given in Table 1.1 (Chapter 1).
Solution
Referring to Fig. 1.43 in Chapter 1, the mode of heat flow becomes essentially one-dimen-
sional if the net arc power is kept sufficiently high compared with the plate thickness (i.e.
Temperature,°C
Time, s
Fig. 6.73. C-curve for precipitation of 3'(Mg2Si) in the reference AA 6082 alloy. The diagram is con-
structed on the basis of equation (6-55).
qold>0.05 kW mm"1). Since this requirement is met in the present case, the HAZ temperature-
time pattern is given by equation (1-100).Taking n = 0.75 and (1 - Xc) = 0.84 for precipitation
of p'(Mg2Si) in AA6082 aluminium alloys,118 we obtain after integration of equations (6-32)
and (6-33) over the weld cooling cycle:
and
The above calculations show that precipitation of (3'(Mg2Si) particles at dispersoids is, in-
deed, a significant process under the prevailing circumstances. Since the mode of heat flow
during single pass butt welding of aluminium plates is essentially one-dimensional, it is possi-
ble to construct general transformation diagrams which give the fraction transformed as a
function of the applied heat input. An example of such a diagram is contained in Fig. 6.74.
logt
Fig. 6.74. Conditions for 0'(Mg2Si) precipitation within the HAZ of single pass AA 6082 butt welds
(Tp>Teq).
Water-quenched
specimens
Hardness (VPN)
5-7 days
Log time
Fig. 6.75. Typical ageing curve for an AA 6082 aluminium alloy at room temperature (schematic).
2-D kinetic (cell) model, assuming that the reaction is interface-controlled. Let r denote the
radius of the growing precipitates (defined in Fig. 6.76(a)). Since we are only interested in the
terminal value of r at a fixed temperature, the time t in the expression for r can be regarded as
constant. Hence, we may write (when C0 » Ca and Cp » C0):
(6-56)
where C4 is a kinetic constant.
If the distribution of the precipitates is approximated by that of a 2-D face-centered cubic
space lattice (see Fig. 6.76(b)), the parameter, AXp, is simply given as:
(6-57)
where rm denotes the maximum particle radius which forms within the system if all alloying
elements are present in solid solution at the onset of the ageing reaction (C0 = C*).
Because of the stoichiometry of the precipitation reaction, C0 and C* in the expression for
AXp may be taken proportional to the magnesium concentration in solid solution. Hence, we
may write:
(6-58)
(6-59)
where pm is the measured resistivity, pss is the resistivity in the as-quenched condition, while p0
is the corresponding resistivity in the fully annealed condition (i.e. when all Mg and Si are tied
up in precipitates).
A comparison between equation (6-58) and the electrical resistivity data in Fig. 6.77 con-
firms the relevance of this power-law-relationship, although the deviation in certain cases in
admittedly large.
Example (6.12)
Consider plasma-MIG butt welding of a 10 mm thick Al-Mg-Si plate under conditions similar
to those employed in Example 6.11. Estimate on the basis of equation (6-58) the relative
(a) (b)
Concentration
Distance
Fig. 6.76. Simplified 2-D kinetic (cell) model for precipitation of hardening particles in Al-Mg-Si alloys
during natural ageing; (a) Particle/matrix concentration profile, (b) Cell geometry.
AXp
C
Mg /C*Mg
Fig. 6.77. Experimental verification of equation (6-58).
fraction of hardening precipitates which forms within the fully reverted region of the HAZ
after prolonged room temperature ageing.
Solution
Referring to Example (6.11), the relative fraction of (3'(Mg2Si) precipitates which forms at
dispersoids during the weld cooling cycle amounts to:
from which
This gives:
Since the resulting precipitation strength increment is directly proportional to AXp,ul loss
of solute in the form of (3'(Mg2Si) particles during the weld cooling cycle will inevitably lead
to a reduced HAZ strength in the naturally aged (T4) condition.6 We shall return to this ques-
tion in Chapter 7 (Section 7.4.3).
(6-60)
Partly deformed Fully plasticized Partly deformed
region region region
(a)
(b)
Fig. 6.78. Micrographs showing the subgrain structure within the fully plasticised and partly deformed
region of a friction welded Al-Mg-Si alloy; (a) Overview, (b) Close-up of the subgrain structure (EBSP
image).
where
(6-61)
The peak temperature T and the strain rate i distributions within the fully plasticised re-
gion of a friction weld can be computed on the basis of the generic models developed by
Midling and Grong.122 Examples of such calculations are shown in Fig. 6.79(a). Plots of the re-
sulting Zener-Hollomon parameter and subgrain diameter at different locations within the
HAZ are contained in Fig. 6.79(b).
Fully plasticized region Partly deformed region
Radial position:
Peak temperature, 0C
Strain rate, s"1
Axial distance, mm
l
Zener-Hollomon parameter, s
Axial distance, mm
Fig. 6.79. Modelling of the subgrain evolution in the fully plasticised region of a friction welded AA
6082 aluminium alloy; (a) Predicted peak temperature Tp and strain rate (e) distributions, (b) Plots of
the Zener-Hollomon Zh parameter and resulting subgrain diameter ds at different locations within
the HAZ. Data from Midling and Grong.122
It is evident from the graphs that the value of the Zener-Hollomon parameter is of the order
of 1014-1012 s"1 within the fully plasticised region of the HAZ during continuous drive friction
welding of Al-Mg-Si alloys. The predicted range in Zh corresponds to a subgrain size of 2-3
jam, in agreement with experimental observations (see EBSP image in Fig. 6.78(b)). Outside
the fully plasticised region the Zener-Hollomon parameter drops from about 1012 to 1010 s"1
due to a sudden change the strain rate from 102 to 10° s"1 as the contribution from the material
flow field in the rotational direction ceases. This value is outside the validity range of equation
(6-60), and leads to the formation of coarse subgrains at the boundary between the fully plas-
ticised and the partly deformed region, as shown by the TEM micrographs in Fig. 6.80.
Fig. 6.80. TEM micrographs showing evidence of coarse subgrains at the boundary between the fully
plasticised and partly deformed region of a friction welded AA 6082 aluminium alloy.
References
1. D. A. Porter and K.E. Easterling: Phase Transformations in Metals and Alloys, 1981, Wokingham
(England), Van Nostrand Reinhold.
2. M.E. Fine: Phase Transformations in Condensed Systems, 1964, New York, The Macmillian Com-
pany.
3. N. Ryum: Compendium in Physical Metallurgy, 1973, Trondheim (Norway), The University of
Trondheim, The Norwegian Institute of Technology.
4. G.F. Vander Voort (ed.): Atlas of Time-Temperature Diagrams for Nonferrous Alloys, 1991, Ma-
terials Park (Ohio), ASM International.
5. J.T. Staley: Mater. ScL TechnoL, 1987, 3, 923-935.
6. O.R. Myhr and 0. Grong: Acta Metall Mater., 1991, 39, 2693-2702; ibid., 2703-2708.
7. R.D. Doherty: In Physical metallurgy, 3rd Edn (Eds R.W. Cahn and P. Haasen), 1983, Amster-
dam, 934-1030, North-Holland Physics Publishing.
8. H.B. Aaron, D. Fainstein and G.R. Kotler: J. Appl. Phys., 1970, 41, 4404-4410.
9. R. Trivedi: Metall. Trans., 1970, 1, 921-927.
10. J. W. Christian: The Theory of Phase Transformations in Metals and Alloys -Part I, 1975, Oxford
(England), Pergamon Press.
11. Atlas of isothermal transformation and cooling transformation diagrams, 1977, Materials Park
(Ohio), ASM International.
12. E. Scheil: Arch. Eisenhuttenwes., 1934/35, 8, 565-571.
13. J.S. Kirkaldy: Metall. Trans., 1973, 4, 2327-2333.
14. M.B. Kuban, R. Jayaraman, E.B. Hawbolt and J.K. Brimacombe: Metall Trans., 1986, 17A,
1493-1503.
15. LA. Wierszykkowski: Metall Trans., 1991, 22A, 993-999.
16. R.G. Kamat, E.B. Hawbolt, L.C. Brown and J.K. Brimacombe: Metall. Trans., 1992, 23A, 2469-
2480.
17. M. Avrami: J. Chem. Phys., 1939, 7, 1103-1112; ibid, 1940, 8, 212-224; ibid, 1941, 9, 177-184.
18. J.W. Chan: Acta Metall, 1956, 4, 572-575.
19. M. Umemoto, Z.H. Guo and I. Tamura: Mat. ScL Technol, 1987, 3, 249-255.
20. R.W.K. Honeycombe: Steels — Microstructure and Properties, 1980, London, Edward Arnold
(Publishers) Ltd.
21. H. Suzuki: Weld. World, 1982, 20, 121-148.
22. 0. Grong and O.M. Akselsen: Metal Construction, 1986, 18, 557-562.
23. R.C. Cochrane: HW Doc. IXJ-102-85 (1985).
24. DJ. Abson and R.E. Dolby: HW Doc. IXJ-29-80 (1980).
25. CA. Dube, H.I. Aaronson and R.F. Mehl: Rev. Met, 1958, 55, 201-210.
26. H.K.D.H. Bhadeshia: Progress in Mater. ScL, 1985, 29, 321-386.
27. E.C. Bain: Trans. Am. Inst. Min. Eng., 1924, 70, 25-46.
28. S.S. Babu and H.K.D.H. Bhadeshia: Mater. ScL Eng., 1991, A142, 209-219.
29. H.K.D.H. Bhadeshia and D.V. Edmonds: Acta Metall, 1980, 28, 1265-1273.
30. W.C. Leslie: The Physical Metallurgy of Steels, 1981, London (England), McGraw-Hill.
31. CE. Cross, 0 . Grong, S. Liu and J.F. Capes: In Applied Metallography, (Ed. G.F. Vander Voort),
1986, New York, Van Nostrand Reinhold, 197-210.
32. J.E. Morral and T.B. Cameron: Proc. Int. Symp. on Boron in Steels, Milwaukee (Wisconsin),
September, 1979, 19-32, Publ. The Metall. Soc. AIME (1980).
33. S. Pakrasi, E. Just, J. Betzold and F. Hollrigl-Rosta: ibid, pp. 147-164.
34. L. F. Porter: ibid., pp. 199-211.
35. J.H. Devletian: Weld. J., 1976, 55, 5s-12s.
36. 0. Grong and D.K. Matlock: Int. Met. Rev., 1986, 31, 27-48.
37. H.K.D.H. Bhadeshia and L.E. Svensson: In Mathematical Modelling of Weld Phenomena, (Eds
H. Cerjack and K.E. Easterling), 1993, London, The Institute of Materials, 109-180.
38. E.D. Hondros and M.P. Seah: In Physical metallurgy, (Eds R.W. Cahn and P. Haasen), 1983,
Amsterdam, North Holland Physics Publishing B.V., 855-931.
39. T.M. Williams, A.M. Stoneham and D.R. Harris: Met. ScL J., 1976,10, 14-19.
40. R.W. Fountain and J. Chipman: Trans. AIME, 1962, 224, 599-606.
41. Ph. Maitrepierre, D. Thivellier and R. Tricot: Metall Trans., 1975, 5A, 287-301.
42. O.M. Akselsen, 0 . Grong and PE. Kvaale: Metall Trans., 1986,17A, 1529-1536.
43. CJ. McMahon: Metall. Trans., 1980, HA, 531-535.
44. FA. Jackobs and G. Krauss: J. Heat Treating, 1981, 2, 139-146.
45. G.M. Evans: HW Doc. II-A-469-79 (1979).
46. H.K.D.H. Bhadeshia, L.E. Svensson and B. Gretoft: Acta Metall., 1985, 33, 1271-1283.
47. S. Liu and D.L. Olson: Weld. J., 1986, 65, 139s-15Os.
48. N.A. Fleck, 0. Grong, G.R. Edwards and D.K. Matlock: Weld. J., 1986, 65, 113s-121s.
49. H.K.D.H. Bhadeshia, L.E. Svensson and B. Gretoft: J. Mat. ScL, 1986, 21, 3947-3951.
50. 0. Grong, T.A. Siewert and G.R. Edwards: Weld. J., 1986, 65, 279s-288s.
51. H.K.D.H. Bhadeshia: Mater. ScL Forum, 1990, 56-58, 263-274.
52. A. AIi and H.K.D.H. Bhadeshia: Mater. ScL Technol., 1990, 6, 781-784.
53. DJ. Abson and RJ. Pargeter: Int. Met. Rev., 1986, 31, 141-194.
54. J.R. Yang and H.K.D.H. Bhadeshia: Weld. /., 1990, 69, 305-307.
55. RL. Harrison and R.A. Ferrar: Int. Mater. Rev., 1989, 34, 35-51.
56. R.C. Cochrane: Weld, in the World, 1983, 21, 16-29.
57. J.M. Dowling, J.M. Corbett and H.W. Kerr: Metall. Trans., 1986,17A, 1611-1623.
58. DJ. Abson: Weld, in the World, 1989, 27, 11-28.
59. M. Strangwood and H.K.D.H. Bhadeshia: Proc. Int. Conf. on Advances in Welding Science and
Technology, Gatlinburg, TN, May 1986, pp. 209-213. Publ. ASM INTERNATIONAL, Metals
Park, OH, 1986.
60. R.F. Hehemann, K.R. Kinsman and H.I. Aaronson: Metall. Trans., 1972, 3, 1077-1094.
61. J.D. Verhoeven: Fundamentals of Physical Metallurgy, 1975, New York, John Wiley & Sons.
62. H.K.D.H. Bhadeshia and J.W. Christian: Metall. Trans., 1990, 21A, 767-797.
63. A.O. Kluken, 0. Grong and J. Hjelen: Metall. Trans., 1991, 22A, 657-663.
64. A.O. Kluken, 0 . Grong and G. R0rvik: Metall. Trans., 1990, 21A, 2047-2058.
65. H. Hu: Trans. AIME, 1965, 233, 1071-1075.
66. A.A.B. Sugden and H.K.D.H. Bhadeshia: Metall. Trans., 1989, 2OA, 1811-1818.
67. P. Deb, K.D. Challenger and A.E. Therrien: Metall. Trans., 1987,18A, 987-999.
68. H.K.D.H. Bhadeshia and L.E. Svensson: J. Mater. ScL, 1989, 24, 3180-3188.
69. R.A. Ricks, G.S. Barritte and PR. Howell: Proc. Int. Conf. on Solid-Solid Phase Transformations,
Pittsburg (Pennsylvania), August 1981, pp. 463-468. Publ. The Metall. Soc. AIME (1982).
70. FJ. Barbaro, P. Krauklis and K.E. Easterling: Mater. ScL Technol., 1989, 5, 1057-1068.
71. S. Liu: In Ferrous Alloy Weldments, (Eds D.L. Olson and T.H. North), 1992, Zurich (Switzer-
land), Trans. Tech. Publications, pp. 1-18.
72. 0. Grong and A.O. Kluken: ibid., pp. 21 -46.
73. L. Devillers, D. Kaplan, B. Marandet, A. Ribes and PV. Ribound: Proc. Int. Conf on Effects of
Residual, Impurity and Alloying Elements on Weldability and Weld Properties, London, Novem-
ber 1983, Paper 1, Publ. The Welding Institute (England).
74. C. Bonnet and F.P. Charpentier: ibid., Paper 8.
75. M.E. Saggese, D.N. Hawkins and J.A. Whiteman: ibid., Paper 15.
76. A.R. Mills, G. Thewlis and J.A. Whiteman: Mater. ScL Technol., 1987, 3, 1051-1062.
77. R.A. Farrar and PL. Harrison: J. Mat. ScL, 1987, 22, 3812-3820.
78. G.M. Evans: Weld. J., 1992, 71, 447s-454s.
79. M. Es-Souni, PA. Beaven and G.M. Evans: Mater. ScL Eng., 1990, A130, 173-184.
80. S. Liu and D.L. Olson: /. Mater. Eng., 1987, 9, 237-251.
81. S. Liu: PhD thesis, 1984, Dept. of Metall. Eng., Colorado School of Mines, Golden, Colorado
(USA).
82. 0 . Grong and A.O. Kluken: Unpublished work.
83. R.A. Ricks, PR. Howell and G.S. Barritte: J. Mat. Sci, 1982,17, 732-740.
84. PL. Ryder, W. Pitsch and R.F. Mehl; Acta Metall., 1967,15, 1431-1440.
85. 0 . Grong, A.O. Kluken, H.K. Nylund, A.L. Dons and J. Hjelen: Metall. Mater. Trans., 1995,26A,
525-534.
86. WC. Johnson, C. L. White, RE. Marth, RK. Ruf, S.M. Tuominen, K.D. Wade, K.C. Russell and
H.I. Aaronson: Metall Trans., 1975,6A, 911-919.
87. RE. Marth, H.I. Aaronson, G.W. Lorimer, T.L. Bartel and K.C. Russell: Metall. Trans., 1976, 7A,
1519-1528.
88. CM. Dallam and D.L. Olson: Weld. J., 1989, 68, 198s-2O5s.
89. G. Thewlis: HW Doc. IXJ-165-90 (1990).
90. R.A. Farrar, Z. Zhang, S.R. Bannister and G.S. Barrite: J. Mater. ScL, 1993, 28, 1385-1390.
91. S.S. Babu and H.K.D.H. Bhadeshia: Mater. ScL Technol, 1990, 6, 1005-1020.
92. H. Homma, S. Ohkita, S. Matsuda and K. Yamamoto: Weld. J., 1987, 66, 301s-309s.
93. R.E. Dolby: Weld. J., 1979, 58, 225s-238s.
94. B.C. Kim, S. Lee, NJ. Kim and D.Y. Lee: Metall Trans., 1991, 22A, 139-149.
95. R.H. Phillip: Weld. J., 1983, 62, 12s-18s.
96. O.M. Akselsen, J.K. Solberg, G. R0rvik and AJ. Paauw: Technical Report STF34 F87013, 1987,
Trondheim (Norway), Sintef-Division of Metallurgy.
97. A.O. Kluken and 0. Grong: Proc. Int. Conf. on Recent Trends in Welding Science and Technology,
Gatlinburg, TN, May, 1989, pp. 781-786. Publ. ASM International, Materials Park, OH, 1990.
98. O.M. Akselsen, G. R0rvik, M.I. Ons0ien and 0. Grong: Weld. J., 1989, 68, 356s-362s.
99. G.R. Speich, V.A. Demarest and R.L. Miller: Metall. Trans., 1981,12A, 1419-1428.
100. C.A.N. Lanzillotto and FB. Pickering: Metal ScL, 1982, 16, 371-382.
101. J.H. Chen, Y. Kikuta, T. Araki, M. Yoneda and Y. Matsuda: Acta Metall, 1984, 32, 1779-1788.
102. M. Ramberg, O.M. Akselsen and 0 . Grong: Proc. Int. Conf. on Advances in Welding Science and
Technology, Gatlinburg, TN, May, 1986, pp. 679-684. Publ. ASM International, Materials Park,
OH, 1986.
103. O.M. Akselsen, 0. Grong and J.K. Solberg: Mater. ScL Technol, 1987, 3, 649-655.
104. X, -L. Cai, AJ. Garratt-Reed and W.S. Owen: Metall. Trans., 1985, 16A, 543-557.
105. JJ. YI, LS. Kim and H.S. Choi: Metall. Trans., 1985, 16A, 1237-1245.
106. O.M. Akselsen, J.K. Solberg and 0 . Grong: Stand. J. Metall, 1988,17, 194-200.
107. A.R. Marder: Metall. Trans., 1981, 12A, 1569-1579.
108. X.P Shen and R. Priestner: Metall Trans., 1990, 21 A, 2547-2553.
109. NJ. Kim and YG. Kim: Mater. ScL Eng., 1990, A129, 35-44.
110. K.R. Kinsman, G. Das and R.F Hehemann: Acta Metall, 1977, 25, 359-365.
111. N.K. Balliger andT. Gladman: Met. ScL, 1981, 15, 95-108.
112. S. Kou: Welding Metallurgy, 1987, Toronto (Canada), John Wiley & Sons.
113. J.E. Hatch (ed.): Aluminium — Properties and Physical Metallurgy, 1984, Metals Park (Ohio),
American Society for Metals.
114. FM. Mazzolani: Aluminium Alloy Structures, 1985, Boston (USA), Pitman Publish. Inc.
115. O. Lohne and A.L. Dons: Scand. J. Metall, 1983, 12, 34-36.
116. R.P Martukanitz: Proc. Int. Conf. on Advances in Welding Science and Technol, Gatlinburg
(Tenn.), May, 1986, pp. 193-201. Publ. ASM International, Materials Park (OH).
117. H.R. Shercliff, 0. Grong, O.R. Myhr and M.F Ashby: Proc. 3rd. Int. Conf. on Aluminium Alloys
—Their Physical and Mechanical Properties, Trondheim (Norway), June, 1992, Vol. Ill, pp. 357-
369. Publ. The Norwegian Institute of Technology, Trondheim, Norway.
118. D.H. Bratland, 0. Grong, O.R. Myhr and H.R. Shercliff: Proc. Int. Conf on Computer-Assisted
Materials Design and Process Simulation, Tokyo (Japan), September, 1993, pp. 135-141. Publ.
The Iron and Steel Institute of Japan.
119. D.H. Bratland, 0 . Grong, H.R. Shercliff, O.R. Myhr and S.Tjotta:,4cta Metall Mater., Overview
No. 124,1997,45,1-22.
120. CM. Sellars: Proc. 3rd Int. Conf. on Aluminium Alloys—Their Physical and Mechanical
Properties, Trondheim (Norway), June, Vol. Ill, pp. 89-106. Publ. The Norwegian Institute of
Technology, Trondheim, Norway.
121. HJ. McQueen and JJ. Jonas: Treatise on Mater. ScL Technol, 1975,6, 393-493.
122. O.T. Midling and 0 . Grong: Acta Metall Mater., 1994,42,1595-1609: ibid., 1611-1622.
Appendix 6.1
Nomenclature
In order to prove that equation (6-32) is the isokinetic version of the Avrami equation, we first
need to rewrite it in a differential form. From equation (6-25), we have:
(A6-1)
Substituting
(A6-2)
(A6-3)
(A6-4)
into equation (A6-1) then gives:
(A6-5)
Provided that equation (A6-5) contains separable variables of X and t* (T), it can be inte-
grated as follows:
(A6-6)
(A6-9)
7.1 Introduction
Weldments are prime examples of components where the properties obtained depend upon the
characteristics of the microstmcture. Since failure of welds often can have dramatic conse-
quences, a wealth of information is available in the literature on structure-property relation-
ships. However, in order to fit some of the apparently conflicting results into a more consistent
picture, a theoretical approach is adopted here rather than a review of the literature. This
procedure also involves the use of phenomenological models for the quantitative description
of structure-property relationships in cases where a full physical treatment is not possible.
The symbols and units used throughout the chapter are defined in Appendix 7.1.
The major impetus for developments in high-strength low-alloy (HSLA) steels has been
provided by the need for: (i) higher strength, (ii) improved toughness, ductility, and formability,
and (iii) increased weldability. In order to meet these contradictory requirements, the steel
carbon contents have been progressively lowered to below 0.10 wt% C. The desired strength
is largely achieved through a refinement of the ferrite grain size, produced by the additions of
microalloying elements such as aluminium, vanadium, niobium, and titanium in combination
with various forms for thermomechanical processing.1 This procedure has made it possible to
improve the resistance of steels to hydrogen-assisted cold cracking, stress corrosion cracking,
and brittle fracture initiation in the weld heat-affected zone (HAZ) region, without sacrificing
base metal strength, ductility, or low-temperature toughness.2 Controlled rolled HSLA steels
are currently produced with a minimum yield strength in the range from 350-550 MPa. Above
this strength level, quenched and tempered steels are commonly employed.
Although the microstructural changes taking place within the weld metal on cooling through
the critical transformation temperature range in principle are the same as those occurring dur-
ing rolling and heat treatment of steel, the conditions existing in welding are significantly
different from those employed in steel production because of the characteristic strong non-
isothermal behaviour of the arc welding process. For example, in steel weld deposits the
volume fraction of non-metallic inclusions is considerably higher than that in normal cast steel
products because of the limited time available for growth and separation of the particles. Oxy-
gen is of particular interest in this respect, since a high number of oxide inclusions is known to
influence strongly the austenite to ferrite transformation both by restricting the growth of the
austenite grains as well as by providing favourable nucleation sites for various types of
microstructural constituents (e.g. acicular ferrite). Moreover, during solidification of the weld
metal, alloying and impurity elements tend to segregate extensively to the centre parts of the
interdendritic or intercellular spaces under the conditions of rapid cooling.67 The existence of
extensive segregations further alters the kinetics of the subsequent solid state transformation
reactions. Accordingly, the weld metal transformation behaviour is seen to be quite different
from that of the base metal, even when the nominal chemical composition has not been signifi-
cantly changed by the welding process.3"5 This, in turn, will affect the mechanical integrity of
the weldment.
The relative contribution from each is determined by the steel chemical composition and
the weld thermal history. Because of the number of variables involved, a full physical treat-
ment of the problem is not possible. Consequently, the simplified treatment of Gladman and
Pickering8 has been adopted here.
Figure 7.1 shows the individual strength contributions in low-carbon bainite, which is the
dominating microconstituent in as-deposited steel weld metals (includes both upper and lower
bainite as well as acicular ferrite). Firstly, there are the solid solution strengthening increments
from alloying and impurity elements such as manganese, silicon and uncombined nitrogen,
which in the present example correspond to a matrix strength of about 165 MPa. Secondly, the
grain size contribution to the yield stress is shown as a very substantial component, the magni-
tude of which is determined by the bainite lath size. Finally, a typical increment for dispersion
strengthening is indicated. This contribution is negligible at large lath sizes typical of upper
bainite, but becomes significant at small grain sizes because of a finer intralath carbide disper-
sion.8 Hence, in steel weld deposits containing high proportions of acicular ferrite or lower
bainite carbides will make a direct contribution to strength, even at relatively low carbon lev-
els.
The results in Fig. 7.1 are of significant practical importance, since they show the inherent
limitations of the system with regard to the maximum strength that can be achieved through
control of the microstructure. As shown in Section 6.3.5.4 (Chapter 6), the typical lath size
(width) of acicular ferrite in low-alloy steel weld metals is about 2 jam. According to Fig. 7.1,
this corresponds to a maximum yield strength of approximately 650 MPa, which is in good
agreement with the observed threshold strength of acicular ferrite containing steel weld depos-
its.9 If higher strength levels are required, it is necessary to decrease the grain (lath) size
through a refinement of the microstructure, i.e. by replacing acicular ferrite with either lower
p 1/4 -f/o
Number of carbides per mm Nv (mm" )
Yield strength, MPa
Dispersion
Grain size
Matrix strength
-1/2
Bainitic ferrite grain size, mm
Fig. 7.1. Contributions to strength in low-carbon bainite. Data from Gladman and Pickering.8
bainite or martensite. Development along these lines has led to the introduction of a new
generation of high strength steel weld metals with a yield strength in the range from 650 to 900
MPa.10'11
Details of these three stages may vary widely in different materials and with the state of
stress existing during deformation. Similarly, the fractographic appearance of the final frac-
ture surface is also influenced by the same factors.
(7-1)
where c\ is an empirical constant.
The tensile test data of Widgery12 reproduced in Fig. 7.2 reveal a strong dependence of £/
on Vv, but the relationship appears to be linear rather than non-linear, as predicted by equation
(7-1). Due to a similar fracture mechanism, a correlation also exists between the Charpy V-
notch (CVN) upper shelf energy and the true fracture strain in tensile testing, as shown in Fig.
7.3. For this reason, the weld metal impact properties are normally seen to decrease with
increasing oxygen concentrations when testing is performed in the upper shelf region. From
Fig. 7.4 we see that the CVN upper shelf energy is a linear function of the weld metal oxygen
content. This observation is not surprising, considering the fact that the inclusion volume
fraction is directly proportional to the oxygen level (see equation (2-75) in Chapter 2).
(7-2)
where 7^ is the surface energy of the particle, Ep is the Young's modulus of the particle, A is the
stress concentration factor at the particle, and dv is the particle diameter.
GMAW
(E=1.6kJ/mm)
Fig. 7.3. Correlation of CVN upper shelf energy with true fracture strain in tensile testing. Data from
Akselsen and Grong.20
Equation (7-2) predicts that large inclusions will tend to form voids first as the stress re-
quired for initiation is proportional to (l/dv )1/2. This result is also in agreement with experi-
mental observations. As shown in Fig. 7.5, the size distribution of inclusions located in the
centre of voids at the fracture surface is significantly coarser than the corresponding particle
size distribution in the material. In particular, large, angular shaped aluminium oxide (AI2O3)
SAW
Fig. 7.4. Correlation of CVN upper shelf energy with analytical weld metal oxygen content. Data from
Devillers et aL 13
inclusions appear to be preferential nucleation sites for microvoids in low-alloy steel weld
metals (see Fig. 7.5(b)). Although the combined effect of particle size and local stress concen-
tration on the ductile fracture behaviour cannot readily be accounted for in a mathematical
simulation of the process, the CVN data in Fig. 7.6 suggest that the content of large inclusions
(e.g. of a diameter greater than about 1.5 Jim) should be minimised in order to maintain a high
resistance against dimpled rupture. In practice, this requires careful control of the weld metal
aluminium-oxygen balance and the heat input applied during welding (see Section 2.12 in
Chapter 2).
Inclusions associated
with dimples
Frequency, %
Frequency, %
Frequency, %
Inclusions associated
with dimples
Fig. 7.5. Histograms showing the size distribution of non-metallic inclusions in the weld metal and in the
centre of microvoids at the fracture surface, respectively; (a) Low aluminium level (Al-containing man-
ganese silicate inclusions), (b) High aluminium level (AI2O3 inclusions). Data from Andersen.18
High Ti levels
Medium Ti levels
CVN upper shelf energy, J
Low Ti levels
SAW
Nv (d v >1.5 um)-105
Fig. 7.6. Correlation of CVN upper shelf energy with number of particles per mm3 greater than 1.5 urn,
Nv(dv > 1.5 um). Data from Grong and Kluken.15
Fig. 7.7. Correlation of true fracture strain with ultimate tensile strength (low-alloy steel weld metals).
Data from Akselsen and Grong.20
(a)
(b)
Fig. 7.8. Schematic diagrams showing cleavage crack deflection at interfaces; (a) High angle ferrite-
ferrite grain boundaries, (b) High angle packet boundaries (bainitic microstructures).
SAW (E = 5.2-6.2 kJ/mm) Al: 0.018-0.062 wt%
Ti: 0.005 - 0.065 wt%
Transition temperature, 0C
O: 0.018-0.058 wt%
(7-5)
where En is the Young's modulus of the matrix, ye^ is the effective surface energy (equal to
the sum of the ideal surface energy and the plastic work), and c is the half crack length.
Since c is proportional to the particle diameter dv, equation (7-5) predicts that welds con-
taining large inclusions should be more prone to cleavage cracking than others. This result is
also in agreement with general observations. For example, in self-shielded FCA steel weld
metals it has been demonstrated that cleavage crack initiation is usually associated with large
aluminium-containing inclusions which form in the molten pool before solidification (see Fig.
7.10). Consequently, control of the inclusion size distribution is essential in order to ensure an
adequate low-temperature toughness.
(b)
(C)
Fig. 7.10. Initiation of cleavage fracture in a self-shielded FCA steel weld from an aluminium-containing
inclusion; (a) Initiation site short distance ahead of the notch, (b) Detail of initiation site showing cracked
inclusion, (c) Detail of cracked inclusion (remnants of particle are left in the hole).
Example (7.1)
Consider multipass FCA steel welding with two different electrode wires, one with titanium
additions and one without. Table 7.1 contains a summary of weld metal chemical composi-
tions. Provided that the microstructure and the inclusion size distribution are similar in both
cases, use this information to evaluate the low-temperature toughness of the welds, as revealed
by CVN testing.
Solution
Since the nitrogen content of both welds is quite high (0.011 wt%), the risk of a toughness
deterioration due to strain ageing is imminent, particularly at low Ti levels. Taking the atomic
weight of titanium and nitrogen equal to 47.9 and 14.0 g mol"1, respectively, the stoichiometric
amount of titanium that is necessary to tie-up all nitrogen as TiN can be calculated as follows:
WeIdA
In weld A most of the nitrogen is free (uncombined) due to an unbalance in the titanium con-
tent. This means that the risk of a toughness deterioration due to strain ageing is high.
(a)
Tensile strength: 600 MPa
Absorbed energy, J
Vol% acicular ferrite
35 Joules
Test temperature, 0C
(b)
Tensile strength: 800 MPa
Absorbed energy, J
35Joules.
Test temperature, 0C
Fig. 7.11. Predicted effect of weld metal acicular ferrite content on the CVN transition curve at two
different tensile strength levels; (a) Rm = 600 MPa, (b) Rm = 800 MPa. Data from Akselsen and Grong.20
WeIdB
Weld B contains 0.030 wt% Ti, which is not far from the stoichiometric amount of titanium
necessary to tie-up all nitrogen. Although some titanium also is bound as Ti2O3, it is reason-
able to assume that the free nitrogen content in this case is sufficiently low to eliminate prob-
lems with strain ageing. Consequently, weld B would be expected to exhibit the highest tough-
ness (i.e. the lowest CVN transition temperature) of the two, as indicated in Fig. 7.14.
(a)
25 vol% acicular ferrite
35 Joules
Test temperature, 0C
(b)
75 v o l % acicular ferrite
Absorbed energy, J
UTS
-..35J.Q.ute$-
Test temperature, 0 C
Fig. 7.12. Predicted effect of weld metal ultimate tensile strength (UTS) on the CVN transition curve at
two different volume fractions of acicular ferrite; (a) 25 vol% AF, (b) 75 vol% AF. Data from Akselsen
and Grong.20
Table 7.1 Chemical composition of FCA steel weld metals considered in Example (7.1).
Element
Weld wt% C wt% Si wt% Mn wt% Al wt% Ti wt% S wt% N wt% O
(b)
Test temperature, 0C
Fig. 7.13. Effect of impurities on weld metal CVN toughness; (a) Nitrogen content, (b) Inclusion level.
Data from ESAB AB (Sweden) and Grong et al. 22
(i) Enhancing the fatigue strength through a general reduction of welding residual
stresses.
Absorbed energy
WeIdB WeIdA
35 J
Test temperature, 0C
Fig. 7.14. Schematic drawings of the CVN transition curves for welds A and B (Example (7.1)).
(a)
(b)
Fig. 7.15. Typical low-temperature fracture modes of Ti-B containing steel weld metals; (a) Quasi-
cleavage (as-welded condition), (b) Intergranular fracture (after PWHT).
(ii) Increasing the toughness by recovery (i.e. removal of strain-aged damage) and
martensite tempering.
For these reasons local PWHTs are commonly required for all structural parts above a
specified plate thickness (e.g. 50 mm according to current North Sea offshore specifications).
Post-weld heat treatment is usually carried out in the temperature range from 550 to 6500C.
In practice, however, the toughness achieved will depend on the weld metal chemical com-
position, and in some cases deterioration rather than improvement of the impact properties is
observed after PWHT. In such cases the reduction in toughness can be ascribed to:3'4
(i) Precipitation hardening reactions. Present experience indicates that elements such
as vanadium, niobium, and presumably titanium can produce a marked deteriora-
tion in toughness because of precipitation of carbonitrides in the ferrite, provided
that these elements are present in the weld metal in sufficiently high concentra-
tions.
(ii) Segregation of impurity elements (e.g. P, Sn, Sb and As) to prior austenite grain
boundaries, which, in turn, can give rise to intergranular fracture. The indications
are that this type of embrittlement is strongly enhanced by the presence of second
phase particles at the grain boundaries.
Experience shows that Ti-B containing steel weld metals often fail by intergranular frac-
ture in the columnar grain region after PWHT,23 as evidenced by the SEM fractographs in Fig.
7.15. The observed shift in the fracture mode is associated with a significant drop in toughness
(Fig. 7.16) and arises from the combined action of solidification-induced phosphorus
segregations and borocarbide precipitation along the prior columnar austenite grain bounda-
Base line
ries (Fig. 7.17). Since borocarbides are brittle and partly incoherent with the matrix, they can
be regarded as microcracks (of length dp) ready to propagate. In such cases there is virtually
no plastic deformation occurring before crack propagation, which implies that the intergranular
fracture stress is given by the Griffith's equation:24
(7-6)
where 7 ^ is again the effective surface energy (equal to the sum of the ideal surface energy
and the plastic work), and dp is the particle diameter.
Although the value of yeg_ would be expected to be low in the presence of solidification-
induced phosporus segregations,24 this alone is not sufficient to initiate intergranular fracture
in the weld metal. However, during PWHT the borocarbides will start to grow from an ini-
tially small value to a maximum size of about 0.1 to 0.2jim (Fig. 7.17), following the classic
growth law for grain boundary precipitates dpatl/4?5 This implies that the intergranular frac-
ture stress, Oj(I), will gradually decrease with increasing annealing times, as indicated in Fig.
7.18. When the matrix fracture strength, Cj(M), is reached, the fracture mode shifts from
predominantly quasi-cleavage in the as-welded condition (Fig. 7.15(a)) to intergranular rup-
ture after PWHT (Fig. 7.15(b)). This is observed as a marked reduction in the CVN toughness,
as shown previously in Fig. 7.16.
Quasi-cleavage
fracture mode
Particle diameter
[Annealing time]174
Fig. 7.18. Schematic illustration of the mechanisms of temper embrittlement in Ti-B containing steel
weld metals (Gf(M): matrix fracture strength, (*/(/): intergranular fracture strength).
p0.2> R nv M P a
HV 5 ,kp/mm 2
Cooling time, At 8 / 5 , s
Fig. 7.19. Structure-property relationships in the grain coarsened HAZ of low-carbon microalloyed steels
(vol% M: martensite content, Rp : 0.2% proof stress, Rm: ultimate tensile strength, HV5: Vickers hard-
ness, A%5.* cooling time from 800 to 5000C). Data from Akselsen et al.26
A number of different empirical models exist in the literature for prediction of HAZ peak
hardness and strength.26"31 However, the aptness of some of these models is surprisingly
good, which justifies construction of iso-hardness and iso-strength diagrams for specific grades
of steels.32 Examples of such diagrams are given in Fig. 7.20. It is evident from Fig. 7.20 that
the HAZ peak strength is controlled by two main variables, i.e. the steel chemical composition
and the weld cooling programme. The compositional effect is allowed for by the use of an
empirical carbon equivalent, which ranks the influence of the various alloying elements on the
steel hardenability. According to Yurioka et al.,2* the CEn-equivalent is given as:
(7-7)
Ordinate:
(7-8)
Abscissa:
(7-9)
The different parameters in equations (7-8) and (7-9) are defined in Appendix 7.1.
The results in Fig. 1.49 are interesting, since they show that the cooling time, A%5, depends
on the mode of heat flow during welding. In this case the transition from 'thick' to 'thin'
plates, corresponding to an abscissa of about 0.64, is clearly not represented by a single plate
thickness d, but will be a function of both the net heat input r\E and the ambient temperature T0.
Accordingly, the HAZ strength level is seen to vary between wide limits, depending on the
steel chemical composition and the operational conditions applied (Fig. 7.20).
Example (7.2)
Consider stringer bead deposition (GMAW) on two low-alloy steel plates of similar composi-
tion but different thickness under the following conditions:
According to the steel mill certificate the CEn carbon equivalent is equal to 0.46 wt%. Use
this information together with the diagrams in Figs. 1.49 and 7.20 to estimate the peak HAZ
strength level when the plate thickness is 10 and 30 mm, respectively.
(a)
CEn, wt%
Cooling time, A t 8 / 5 , s
(b)
CE||f wt%
Cooling time, At 8 7 5 , s
Fig. 7.20. HAZ iso-property diagrams for HSLA steels; (a) Iso-hardness contours, (b) Iso-yield strength
contours. Data from Kluken et al.32
Solution
First we calculate the net heat input per unit length of the weld r\E:
d = 10 mm:
from which
d = 30 mm:
from which
We can now use the diagrams in Fig. 7.20(a) and (b) to obtain the peak HAZ hardness and
yield strength, respectively. This gives:
d = 10 mm:
d = 30 mm:
It is evident from the above calculations that the HAZ strength level is sensitive to varia-
tions in the welding conditions. Normally, the HAZ hardenability is high enough to facilitate
a local strength increase adjacent to the fusion boundary, as shown in Fig. 7.21. An exception
is high heat input welding on quenched and tempered steels (Fig. 7.2l(b)), where the presence
of large amounts of Widmanstatten ferrite and polygonal ferrite within the grain coarsened and
grain refined region, respectively can lead to a severe HAZ softening. This type of mechanical
impairment represents a problem in engineering design, since it puts a restriction on the use of
high strength steels in welded structures.
R p02 and R m ,
Gf[R.
GCR
GCR
GRR
BM"
BM
JfL
IR"
GCR
"GRR
TR"
SR"
BM"
JfL,
,BM
[SR
Fig. 7.21. Effects of steel chemical composition and welding conditions on the HAZ strength level (BM:
base metal, SR: subcritical region, IR: intercritical region, GRR: grain refined region, GCR: grain coars-
ened region); (a) Low heat input, (b) High heat input. Data from Akselsen and R0rvik.34
7.2.2.2 Tempering of the heat affected zone
Certain regulations for offshore structures require that no part of the welded joint shall be
harder than a specified limit, e.g. 280, 300 or 325 VPN, to reduce the risk of hydrogen crack-
ing. Such requirements cannot always be met by a suitable choice of preheating and welding
conditions.
In practice, a reduction in the HAZ strength level can be achieved by applying a PWHT.
The tempering effect of different temperature-time combinations can be described by the
Hollomon-Jaffe parameter:35
(7-10)
Filled symbols:
t = 10 seconds
Vickers hardness, VPN
Fig. 7.22. Hollomon-Jaffe type plot of isothermal hardness data. After Olsen et al.36
Temper bead
Fusion line
Last Ac3 line
filler pass Ac1 line
(7-11)
where Qapp. is the apparent activation energy for the controlling diffusion reaction.
The Dorn parameter has proved useful to compare isothermal and pulsed tempering data on
the assumption that the kinetics of softening, in the actual range of hardness, are controlled by
diffusion of carbon in ferrite. Qualitatively, the aptness of equation (7-11) can be illustrated in
a plot of measured hardness against the diffusional parameter P^ ( s e e Fig- 7.24). It is evident
from Fig. 7.24 that the isothermal data points can be represented by a smooth curve which
coincides with the upper boundary of the scatter band obtained in pulsed tempering. The
slightly higher hardness observed after isothermal tempering arises probably from a brief pe-
riod of heating that makes the effective time somewhat less than 10 s.
1 2
^s '
Fig. 7.24. Measured hardness ratio HVIHVmax. vs the Dorn parameter P2 (Qapp. = 83.14 kJ mol *). Data
fromOlsentf/tf/.36
beads is clearly the same. In Fig. 7.25 an estimate has been based on the simplified Rykalin
thick plate solution, which applies to a fast moving high power source on a semi-infinite body
(see equation (1-73) in Chapter 1). At T-T0 ~ 15000C, a fusion line radius of about 4.4 mm is
obtained for a net heat input of 0.8 kJ mm"1. The corresponding Ac\ radius is 6.5 mm.
The temperature-time pattern is shown in the lower left graphs of Fig. 7.25 for three differ-
ent positions in the HAZ, i.e. y = 0 (former fusion line), y = 1 mm, and y = 2 mm (z = 0). The
corresponding plots of dP2 ldt vs t are shown to the right. Taking the area P2 under each curve
and reading the hardness ratio at TJP^ from Fig. 7.24, an expected hardness profile is ob-
tained, as shown in the upper diagram of Fig. 7.25. The expected effect of tempering is seen to
range from a hardness of about 65% (HV « 265 VPN) at the fusion line to about 80% (HV «
340 VPN) close to the outer boundary of the HAZ (y = 2 mm). If the centre-line displacement
had been different from the chosen optimum of 2.1 mm (e.g. say 3 mm), the predicted hardness
curve would be shifted to about 75% and 90% of the peak hardness at y = 0 and y = 2 mm,
respectively. On the other hand, if the centre-line distance had been shorter, say 1 mm, a
narrow zone of the original HAZ would be re-austenitised and therefore about as hard as
before deposition of the temper bead.
The results from the above modelling exercise show that the HAZ hardness of weld toes
and cap layers can be reduced by applying an appropriate temper bead technique. However,
this requires an extremely good process control, since the temper beads must be positioned
very precisely for a successful result. Consequently, the use of temper beads for improvement
of the HAZ properties has not found a wide application in the industry.3641
HA2
We d metal Parent plate
T, 0C
106exp(-10000/T)
t,s t,s
Fig. 7.25. Application of Dorn parameter to weld bead tempering (Case Study (7.1)).
form brittle microstructures within specific thermal regions of the weld. 4142 Moreover, im-
provement of the HAZ toughness through PWHT is sometimes found to be difficult in contrast
to experience with more traditional C-Mn steels.41'43 Consequently, the increasing use of low-
carbon microalloyed steels in various welded structures has introduced new problems related
to the HAZ brittle fracture resistance which formerly did not appear to be of particular con-
cern.44
(D
c
O
I Sn
igel cycel
0
Peak e
tmperau
tre,C
Fig. 7.26. Effects of peak temperature on the CVN energy absorption at -400C (SR: subcritical region,
IR: intercritical region, GRR: grain refined region, GCR: grain coarsened region). Data from Akselsen et
a/.45
to the fusion boundary where the peak temperature of the thermal cycle has been above about
12000C. The problem can mainly be ascribed to the presence of low-toughness microstruc-
tures such as upper bainite and Widmanstatten ferrite which form typically at intermediate and
slow cooling rates (see Fig. 6.55 in Chapter 6). In contrast, the grain refined region will almost
always exhibit a satisfactory low-temperature toughness owing to the characteristic fine po-
lygonal ferrite microstructure.41 An exception is low heat input welds produced from steels
with a heavily banded pearlite/ferrite microstructure, where the risk of a toughness deteriora-
tion is imminent due to martensite formation along the prior base metal pearlite bands.45'46
In recent years a new class of low-carbon microalloyed steels has emerged which is charac-
terised by an excellent low temperature HAZ toughness, even at high heat inputs (see Fig.
7.27). This particular grade is frequently referred to as Ti-O steels due to their content of
indigenous titanium oxide inclusions (presumably Ti2O3). Although the mechanisms involved
are not yet fully understood, it is reasonable to assume that the improved toughness at high
heat inputs arises from a refinement of the HAZ microstructure, as discussed previously in
Section 6.3.6 (Chapter 6). It is interesting to note that the major effect of the titanium oxide
inclusions in this case appears not to be control of the austenite grain size (which in some cases
can exceed 500 |im at the fusion boundary), but is rather to act as favourable nucleation sites
for acicular ferrite intragranularly.4748 Similar phenomena are well known from transforma-
tion kinetics of low-alloy steel weld deposits, where non-metallic inclusions play an important
role in the development of the acicular ferrite microstructure.3"5
Intercritical region
The microstructural evolution in the intercritical HAZ of low-carbon steels has previously
been discussed in Section 6.3.8.2 (Chapter 6).
In order to understand the origin of embrittlement in the intercritical region, consideration
must be given to the stress fields and the transformation strains developed in the ferrite matrix
Ti-O steel
Ti-N steel
Transition temperature, 0C
Peak temperature, 0C
Fig. 7.27. Response of modern Ti-O steels and traditional Ti-N steels to CVN testing following weld
thermal simulation. Data from Homma et al.41
as a result of the martensite formation.49 It follows from Fig. 7.28 that the hard martensite-
austenite (M-A) islands will give rise to significant stress concentrations at the martensite/
ferrite interface owing to the pertinent difference in the yield strength (stiffness) between the
two phases. At the same time, the volume expansion associated with the austenite to martensite
transformation leads to significant elastic and plastic straining of the ferrite.50 At moderately
high temperatures and deformation strains, many of the matrix dislocations will be mobile,
which means that the ferrite will maintain its ductility, while the stiffer M-A islands are ex-
posed to cracking and debonding. With increasing strain, the cracks can grow into voids and
further develop into deep holes, until final rupture occurs by hole/void coalescence due to
internal necking.49 However, when mechanical testing is performed at subzero temperatures
under high strain rate conditions (> 102 s"1 for CVN testing), the flow strength of the ferrite
increases significantly because of the reduced mobility of the screw dislocations.51 In addition,
strain partitioning between the M-A islands and the ferrite may also occur, which further
enhances the stress concentrations at the M-A/ferrite interface.52 Accordingly, the local stress
level at the interface will eventually exceed the cleavage strength of the ferrite, with conse-
quent initiation of brittle fracture. This conclusion is consistent with observations made from
tensile testing of dual-phase steels, showing that failure of dual-phase microstructures often is
caused by fracture in the ferrite region.52"54
Because the intercritical HAZ toughness is closely related to the volume fraction of the
M-A constituent in the matrix, 4 5 5 1 5 5 embrittlement can normally be avoided by decreasing
the cooling rate through the critical transformation temperature range to facilitate pearlite for-
mation (see Fig. 7.29). An exception is boron-containing steels, where the HAZ hardenability
is high enough to stabilise the M-A constituent, even at slow cooling rates (see CVN data for
steel B in Fig. 7.29).
Normalized stress
Stiff
particle
Normalized distance
Fig. 7.28. Stress distribution in matrix caused by stiff inclusion (or: radial stress, (5$: tangential stress,
tmax.' maximum shear stress). Data from Chen et al.49
Open symbols:
Filled symbols:
Steel A
Absorbed energy, J
(T-L)
(L-T)
Steel B
(T-L)
Cooling time, At 6 / 4 , s
Fig. 7.29. Effect of cooling time Ar674 on the intercritical HAZ toughness at -20 0 C (thermally cycled
specimens). Steel A: 11 ppm B, Steel B: 26 ppm B. Data from Ramberg et al.55
Effect of PWHT
Considering the intercritical HAZ, a significant improvement of the CVN toughness can be
achieved by applying a PWHT, as shown by the data of Akselsen et al51 This effect arises
partly from a reduction of the stress concentrations at the M-A/ferrite interface as a result of
martensite tempering and partly from relaxation of transformation strains within the ferrite
matrix.51 Such recovery reactions will start to occur when the temperature is raised above
about100 0 C.
(a)
(b)
CTOD at -1O0C, mm
P content, wt%
Fig. 7.30. Effects of PWHT on the grain coarsened HAZ toughness; (a) Example of intergranular frac-
ture along prior austenite grain boundaries after PWHT (6000C - 1 h), (b) Measured CTOD vs base
plate phosphorus content for post weld heat treated specimens (6000C - 4 h). Data quoted by Grong and
Akselsen.41
In contrast to the behaviour described above for the intercritical HAZ, the reported effect of
PWHT on the grain coarsened HAZ toughness is much more complicated and rather confus-
ing. However, experience has shown that particularly niobium-vanadium containing steels
are sensitive to PWHT due to the strong precipitation hardening potential of Nb(C,N) and
V(C,N).43'56 In addition, a toughness deterioration may occur as a result of segregation of
impurity elements such as phosphorus, tin, and antimony to prior austenite grain boundaries.
This, in combination with a tempered martensitic microstructure, can lead to intergranular
fracture when testing is performed at subzero temperatures (see Fig. 7.30(a)). The detrimental
effect of phosphorus on the HAZ toughness of low carbon microalloyed steels after PWHT is
shown in Fig. 7.30(b).
Example (7.3)
Consider procedure test SA welding on a thick plate of a Nb-microalloyed steel under the
following conditions:
Table 7.2 contains data from CVN testing of the base plate and thermally cycled specimens.
The weld thermal simulation experiments were carried out at three different peak tempera-
tures (i.e. 13500C, 10000C, and 7800C) under cooling conditions similar to those employed in
the SA welding trial. Based on the data in Table 7.2 and the simplified Rykalin thick plate
solution (equation (1-73) in Chapter 1), estimate the locations of the brittle zones (referred to
the fusion boundary) within the HAZ of the SA procedure test weld considered above.
Solution
It is evident from the CVN data in Table 7.2 that the HAZ toughness would be expected to be
low in positions of the weld where the peak temperature has been close to 780 and 13500C,
conforming to the intercritical and grain coarsened region, respectively. Based on the simpli-
fied Rykalin thick plate solution, the following expression can be derived for an arbitrary
isothermal zone width, Ar*m, referred to the fusion boundary (see equation (5-47) in Chapter
5):
Taking pc and Tm equal to 0.005 J mm"3 0C"1 and 15200C, respectively for low-alloy steels
(from Table 1.1 in Chapter 1), we obtain:
Table 7.2 Results from CVN testing of base metal, thermally cycled specimens, and procedure test
weld (Example (7.3))
1
GCR: grain coarsened region; GRR: grain refined region; IR: intercritical region.
Next Page
From this we see that the brittle zones are located 3.5 and 0.5 mm from the fusion bound-
ary, respectively. A comparison with the procedure test results in Table 7.2 shows that the
measured CVN toughness after welding at these locations is slightly higher than that inferred
from the weld thermal simulation experiments. This observation is not surprising, considering
the fact the CVN specimens extracted from the procedure test weld, in practice, include a
wide spectrum of thermal regions which have undergone highly different temperature-time
programmes, whereas the microstructure within the thermally cycled CVN specimens is more
homogeneous due to a similar temperature-time pattern across the whole gauge length (see
Fig. 7.31). Hence, weld thermal simulation cannot replace procedure testing carried out on
real welds. Nevertheless, it is a useful method of evaluating the microstructural stability and
mechanical response of materials to reheating, as experienced in welding.
From this we see that the brittle zones are located 3.5 and 0.5 mm from the fusion bound-
ary, respectively. A comparison with the procedure test results in Table 7.2 shows that the
measured CVN toughness after welding at these locations is slightly higher than that inferred
from the weld thermal simulation experiments. This observation is not surprising, considering
the fact the CVN specimens extracted from the procedure test weld, in practice, include a
wide spectrum of thermal regions which have undergone highly different temperature-time
programmes, whereas the microstructure within the thermally cycled CVN specimens is more
homogeneous due to a similar temperature-time pattern across the whole gauge length (see
Fig. 7.31). Hence, weld thermal simulation cannot replace procedure testing carried out on
real welds. Nevertheless, it is a useful method of evaluating the microstructural stability and
mechanical response of materials to reheating, as experienced in welding.
Specimen holder
[Homogeneous zone
CVN-specimen
Notch location
CVN-
specimen Base metal
Fig. 7.31. Methods for evaluation of HAZ toughness (schematic); (a) Weld thermal simulation, (b) Weld
procedure testing.
(a) (b)
Toe
Root irack
Toe crack Underbead
crack crack
Underbead HAZ
crack
Fig. 7.32. Schematic diagrams showing hydrogen-induced cracks in different types of welds; (a) Fillet
weld, (b) Butt weld. The diagrams are based on the ideas of Coe.57
5%/min 104%/min
Uncharged
Uncharged
Charged
True fracture strain
Charged
5x105%/min 1.9x106%/min
Uncharged
Charged
Charged
Uncharged
Test temperature, 0C
Fig. 7.33. Variation of true fracture strain with nominal strain rate and test temperature for charged and
uncharged specimens. Data from Brown and Baldwin.59
Over the years a number of mechanisms have been proposed to explain the origin of hydro-
gen embrittlement. The three most important are:
Fracture
Resistance change x 10 ,Q.
-8
Fracture
Applied stress:
1240MPa
Applied stress:
1100MPa
Time, min
Fig. 7.34. Example of stepwise crack propagation in notched tensile specimens, as inferred from electri-
cal resistivity measurements. Data from Steigerwald et a/.60
(a) The hydrogen gas pressure model, originally proposed by Zapfee and Sims,61 which
postulates that atomic hydrogen will diffuse to microvoids where it recombines to
form molecular hydrogen. In ferritic steels the equilibrium H2(g) pressure within
the microvoids is typically of the order of 106 to 107 atm, which is more than
sufficient to bring about a local fracture development.
(b) The surface energy model (Petch62). According to this model hydrogen will re-
duce the effective surface energy of the crack. Under such conditions the crack
can propagate at a lower nominal stress in the presence of hydrogen, in agreement
with the Griffith's theory (equation (7-5)).
(c) The slip softening model of Beachem,63 which accounts for the experimental ob-
servation that hydrogen-charged specimens generally exhibit a lower flow stress
than hydrogen-free specimens. This suggests that hydrogen interfers with dislo-
cations in a manner which facilitates different types of fracture, including micro void
coalescence (or dimpled rupture), quasicleavage fracture, and intergranular frac-
ture.64
Currently, it cannot be stated with certainty which of these three mechanisms that are opera-
tive under the conditions existing in welding. However, this question is of minor importance
in the present context, since we here are mainly concerned with the factors responsible for
hydrogen cracking in steel weldments.
7.2.3.2 Solubility of hydrogen in steel
Since hydrogen is the smallest of all atoms, it is readily soluble in iron. In general, both
octahedral and tetrahedral lattice sites are potential traps for interstitials, as indicated in Fig.
7.35. In the case of hydrogen it is believed that the dissolved atoms are mainly present in
tetrahedral positions in the form of protons.65 Because of the pertinent difference in the size of
the fee and the bcc interstices (see Fig. 7.35), the hydrogen solubility in iron will change
stepwise with temperature following the bfe —> yFe and yFe —> aFe transformations, as shown
previously in Fig. 2.7(c) (Chapter 2).
(a)
(b)
Fig. 7.35. Schematic representation of octahedral and tetrahedral lattice sites in different crystal struc-
tures; (a) Face-centred cubic (fee) structures, (b) Body-centred cubic (bcc) structures.
In addition to the interstitial fraction, hydrogen may be present in the form of molecular
(gaseous) hydrogen trapped in micro voids or plane lattice defects. This amount is frequently
referred to as residual hydrogen, and can in many cases overshadow the equilibrium hydrogen
content. For example, at room temperature the maximum solubility of atomic hydrogen in the
iron lattice is estimated to be 0.001 to 0.01 ppm, while the analytical hydrogen content of
steels varies typically from 1 to 10 ppm. This supersaturation is formidable and provides the
necessary driving force for trapping of gaseous hydrogen in the microstructure.
(7-12)
where D^ is the lattice diffusion coefficient for hydrogen in bcc iron, K is the density of trap
sites (i.e. number of trap sites per number of lattice sites), and EB is the binding energy be-
tween hydrogen and the trap site.
A graphical representation of equation (7-12) is shown in Fig. 7.36. A closer inspection of
the graph reveals that the predicted temperature dependence of the apparent diffusion coeffi-
cient is in fair agreement with the reported diffusivity data for hydrogen in steel. Moreover, it
is interesting to note that the hydrogen diffusion coefficient in austenite is nearly two orders of
magnitude lower than the corresponding value for the ferrite phase at a given temperature.
This observation is not surprising, considering the pertinent difference in the packing density
between the fee iron lattice and the bcc iron lattice (74% and 68%, respectively). Thus, for
diffusion of hydrogen in austenite, we have:57
(7-13)
Ferritic steels
Trapping
theory
Austenitic
steels
1000/T1K'1
Fig. 7.36. Summary of reported diffusion coefficients of hydrogen in iron and steel. Data compiled by
Coe57 and Yurioka and Suzuki.58
Several successful attempts have been made in the past to model hydrogen diffusion in
welds by means of numerical methods.68"70 Unfortunately, none of these solutions are simple
enough to get a good overall indication of the hydrogen redistribution during cooling and
subsequent PWHT. As an illustration of principles, we shall therefore present a simplified
analytical solution to the hydrogen diffusion problem in welding, based on an analogy be-
tween diffusion and heat conduction.
Diffusion model
The idealised model considers a butt weld of uniform hydrogen concentration in the longitudi-
nal direction, as shown in Fig. 7.39. The width of the fusion zone is 2L, while the initial
hydrogen concentration at the time of solidification (i.e. at t = 0) is Q. The hydrogen con-
centration in the base metal outside the fusion zone is C0. If element losses to the surroundings
are neglected, the problem can be treated as uniaxial diffusion in an isotropic solid analogous
to that described in Section 1.7 (Chapter 1) for heat conduction in thermit welding. Thus, in
the limiting case where the diffusion coefficient can be regarded as constant, the hydrogen
concentration ( Q as a function of time (t) and distance (y) is given by equation (1-22):
Deposited metal
ml H 2 /10Og
Fused metal
Position x, mm
ml H 2 /100g
Mean value
Diffusible
Residual
10 mm
Fig. 7.37. Measured longitudinal and lateral distributions of hydrogen in a single pass SMA weld quenched
right after welding. Data from Christensen et al.61
(7-14)
where D** is the hydrogen diffusivity, and erf(u) is the Gaussian error function (defined previ-
ously in Appendix 1.3, Chapter 1).
In practice, it is necessary to rewrite equation (7-14) in a differential form to allow for the
variation in the hydrogen diffusion coefficient with temperature. After some manipulation, we
obtain:
(7-15)
H I/cm3
15 mm
a)
p. I/cm3
b)
Fig. 7.38. Redistribution of hydrogen following welding (numerical calculations); (a) Right after weld-
ing, (b) After 12 h at room temperature (ljil cm"3 = 0.0115 ppm). Data from Christensen.68
Fusion zone
when when
when when
(b)
Temperature-time Solid curves: Centre - line
programme Broken curves: HAZ
T0 = 200C
(C-C0)Z(C1-C0)
Temperature, 0C
Hydrogen
concentration
Time, s
(C)
Temperature-time Solid curves: Centre - line
programme Broken curves: HAZ
T0 = IOO0C
(C-C0)Z(C1 -C 0 )
Temperature, 0C
Hydrogen
concentration1
Time, s
Fig. 7.40. Computed temperature and hydrogen concentration profiles during GTA butt welding of a
2mm thin steel sheet (Case Study (7.2)); (a) Sketch of weld, (b) Redistribution of hydrogen in the ab-
sence of preheating, (c) Redistribution of hydrogen after preheating to 1000C.
veloped over the years to study the mechanisms of hydrogen cracking in weldments (e.g. see
the review of Yurioka and Suzuki58). Broadly speaking, the cold cracking tests fall into either
one of the two categories, i.e. self-restrained tests or externally loaded tests. Examples of the
former type are the Tekken (oblique Y-groove) cracking test, the CTS (controlled thermal
severity) cracking test, and the cruciform cracking test. Well-known externally loaded tests
are the implant cracking test, the TRC (tensile restraint cracking) test, and the RRC (rigid
restraint cracking) test.
Test weld
Tensile load
The CEW parameter in equation (7-16) refers to the so-called HW carbon equivalent, origi-
nally developed for C-Mn steels:
(7-17)
Moreover, A/333 is a hydrogen diffusional parameter which takes into account variations in
the measured implant rupture strength after various thermal treatments (including preheating
and PWHT). According to Christensen and Simonsen,71 the extent of hydrogen diffusion
which occurs in the low-temperature regime can be reported in the form of an equivalent
isothermal hold time at 6O0C (or 333K), defined as:
(7-18)
where Tc refers to the local HAZ temperature at the moment of quenching (usually taken as
1000C).
It follows from equation (7-16) that the two first members reflect the influence of micro-
structure upon the implant rupture strength, and is therefore related to the HAZ peak hardness.
The two last members take into account the effect of analytical HFM and local hydrogen con-
centrations. As shown in to Fig. 7.42, the numerical values of the partial derivatives dRIR/
dCEw, dRIR/dAts/5, 3RIR/3 log HFM, and dRIR I B^At333 may vary within relatively wide limits,
depending on the steel chemical composition and the operational conditions applied. Never-
theless, the concept is still useful for quantitative predictions of the HAZ cracking resistance,
as illustrated below.
Example (7.4)
Experience has shown that conventional pipeline steels with carbon equivalent CEW up to
0.4% can be welded with basic electrodes (Af8/5 ~ 8-9 s) without the use of external preheating,
provided that the weld metal hydrogen content is kept sufficiently low (HFM ~ 4 ppm). Sup-
pose that the same procedure shall be employed in hyperbaric welding of pipeline steels at a
depth of 320 m (33 bar total pressure). Based on the implant test data in Fig. 7.42, estimate the
minimum reduction in the steel carbon equivalent (ACEW) which must be incorporated in the
specifications to compensate for the increased hydrogen absorption observed at such depths
(#FM-10ppm).
Solution
The concept of partial derivatives implies that we will have the same safety against hydrogen
cracking if there is no net change in the implant rupture strength (i.e. ARm = 0). Since the weld
(a)
HSLA
R|R,MPa
steels
CEW,%
(b)
HSLA steel
R|R, MPa
At8/5, S
Fig. 7.42. Examples of implant test results: (a) Rm vs CEW, (b) RIR vs Ar8/5.
cooling programme is similar in both cases, the variation in A/8/5 and ^At333 can be neglected.
Hence, equation (7-16) reduces to:
(C)
HSLA steel
R1R, MPa
HFM, ppm
(d)
R|R, MPa
^ • • "
Fig. 7.42. Examples of implant test results (continued); (c) RIR vs HFM, (d) RIR vs -^Ar333 . Data from
Christensen and Simonsen.71
In the present example the total change in the weld metal hydrogen content between 1 and
33 bar is equal to:
Moreover, the numerical values of dRIR/dCEw and dRIR/d log HFM can be read from Fig.
7.42(a) and (c), respectively. When A/8/5 ~ 8.6 s, we obtain:
and
This gives:
The above calculations suggest that the CEW carbon equivalent of pipeline steels should not
exceed 0.35% if hydrogen cracking is to be avoided under hyperbaric welding conditions.
(7-19)
Equation (7-19) predicts that the threshold stress (and thus the steel cracking resistance)
passes through a local maximum as the yield strength increases. The locus of this peak stress
is obtained by setting dath/dRpo2 = 0, which gives Rpo2 ~ 600MPa and crth (max) ~ 360MPa.
In practice, a hardness criterion rather than a yield strength criterion is used for ranking of
steels with regard to H2S stress corrosion cracking resistance. According to Dieter,19 the fol-
lowing relation exists between Rpo and HV:
(7-20)
Stress
Anode:
Cathode:
Stress
Fig. 7.43. Mechanisms of hydrogen absorption in cathodic stress corrosion cracking (schematic).
corresponds to a hardness of about 250VPN. This value should be compared with the maxi-
mum hardness level of 22HRC (Rockwell C) or 248VPN incorporated in many offshore speci-
fications.
Ultra-low-carbon steel
Low-carbon steel
C-Mn steel
A
WS
Fig. 7.44. Computed a f/l -A% 5 profiles for selected steels.
Normalized threshold stress
Ultra-low-carbon steel
Low-carbon steel
C-Mn steel
A s
w
Fig. 7.45. Effect of peak hardness on the HAZ stress corrosion cracking resistance.
Previous Page
Stainless steels are widely used in various industries where corrosion is of particular concern.
These materials can be classified into four main categories, based on their microstructure:
As shown in Table 7.3, the welding of stainless steels is encumberred by a number of differ-
ent metallurgical problems, including solidification cracking, hydrogen cracking, precipita-
tion reactions, and grain growth. Some of these problems will be discussed below in the light
of information available in recent literature.75"78
Chemical Composition
Material Major Elements Minor Elements Welding Problems
(a)
Dissolution of
TiC or NbC
Temperature
1st thermal
cycle
Time
(b)
Temperature
2nd thermal
cycle
Time
Fig. 7.46. Mechanisms of knife-line corrosion attack in austenitic stainless steel weldments (schematic);
(a) Dissolution of TiC or NbC during the initial weld thermal cycle, (b) Precipitation of Cr23C6 in the
low peak temperature region of the weld HAZ following deposition of the second layer. (The
corresponding C-curves for precipitation of TiC/NbC and Cr23C6 in the grain growth zone are displaced
far to the right in the diagram.)
overlaps the lower part of the fusion boundary of the first (root) pass, as indicated in Fig.
7.47(a). In practice, the problem can be eliminated by simply reversing the welding sequence
or by changing the welding procedure so that the second pass overlaps the middle rather than
the lower part of the root pass. The latter point is illustrated in Fig. 7.47(b).
(7-21)
where ct is the friction stress (representing the overall resistance of the crystal lattice to dislo-
cation movement), k is the locking parameter (which measures the relative hardening contri-
bution of the grain boundaries), and D is the average grain diameter.
It follows from the analysis in Section 5.4.2.5 (Chapter 5) that the grain size across the HAZ
of austenitic stainless steel weldments may vary by a factor of three to five, depending on the
(a)
Cr-depleted
region
(b)
Cr-clepleted
region
No corrosion attack
Fig. 7.47. Effect of welding performance on the corrosion resistance of Ti/Nb-stabilised austenitic stain-
less steels; (a) Low resistance against knife-line corrosion attack, (b) High resistance against knife-line
corrosion attack. The diagrams are based on the ideas of Kou.76
applied heat input. This means that a permanent soft zone will form adjacent to the fusion
boundary after welding, which may reduce the overall load-bearing capacity of the joint.
Example (7.5)
Consider plasma arc butt welding of a 5mm thick plate of type 316 austenitic stainless steel
under the following conditions:
Provided that the conditions for one dimensional heat flow are met, estimate on the basis of
the nomograms in Fig. 5.30(b) (Chapter 5) the variation in the austenite grain size across the
HAZ after welding. Calculate then via the Hall-Petch relation (equation (7-21)) the expected
reduction in the HAZ strength level due to this change in the microstructure.
Input data:
Base metal yield strength: 300 MPa
Locking parameter in Hall-Petch relation: k = 227 MPa (im1/2
Solution
First we need to calculate the net heat input per mm2 of the weld:
Readings from the nomograms in Fig. 5.30(b) give the HAZ grain size profile shown in Fig.
7.48. In order to obtain the resulting HAZ strength distribution, it is necessary to fix the value
of the friction stress, Oi, in the Hall-Petch relation. In the present example, we have:
from which
It follows from the graphical representation of the Hall-Petch relation in Fig. 7.48 that the
variation in the yield strength across the HAZ is rather small under the prevailing circum-
stances. In fact, the maximum HAZ strength reduction which may occur in such materials
because of grain growth is about 18%, corresponding to high heat input welding conditions.
This implies that welding does not impose severe restrictions on the design stress as long as the
steel is used in the fully annealed condition.
Fig. 7,48. Computed HAZ grain size and yield strength profiles in a 5mm thick butt weld of type 316
austenitic stainless steel (Example (7.5)).
Absorbed energy
Test temperature
steels unsuitable for many structural applications where the HAZ toughness is of particular
concern.
In contrast, medium- and high-strength fee metals have usually such high toughness that
brittle fracture is not a problem at low temperatures, unless there is some special reactive
chemical environment. Austenitic stainless steels and aluminium alloys fall within this cat-
egory. When it comes to duplex stainless steel weldments, the situation is more complex.
Here the HAZ toughness is determined by the austenite/ferrite balance in the weld, which, in
turn, depends on the steel chemical composition and the operational conditions applied.78"80
As shown in Fig. 7.50, complete ferritisation is normally achieved in regions close to the
fusion boundary during the initial heating leg of the thermal cycle. Provided that the cooling
rate through the critical transformation temperature range is kept reasonably low, a significant
proportion of the ferrite may retransform back to austenite on cooling (see Fig. 7.51). This will
contribute to a high HAZ toughness, even at subzero temperatures, as indicated by the CVN
transition curves in Fig. 7.52. In general, an austenite content of about 30 vol% is sufficient to
avoid problems with the HAZ toughness in duplex stainless steel welds.
9
Partly transformed zone §
I
Unaffected base metal
I
Fig. 7.50. Schematic diagram defining different thermal regions within the HAZ of a single pass duplex
stainless steel weld.
30-45sat1300°C Fe-26.0 wt% Cr - 6.9 wt% Ni
Temperature, 0C A4- temperature: 1260 0C
% 5 - ferrite at RT
Time, s
Fig. 7.51. CCT-diagram for a duplex stainless steel. Data from Mundt and Hoffmeister.79
energy interface between yFe and o>e, which prevents spreading of the liquid along the grain
boundaries.82 For these reasons, a minimum weld metal delta ferrite content of about 5 to 10
vol% is usually specified for austenitic stainless steels.
The quantitative relationship between the delta ferrite content and the weld metal chemical
composition in austenitic stainless steels has been determined first by Schaeffler83 and later by
Delong et a/.84'85 The constitution diagram of Delong is shown in Fig. 7.53. Here the alloying
elements are grouped into ferrite formers (i.e. Cr, Mo, Si, and Nb) and austenite formers (i.e.
Ni, C, N, and Mn) to determine the corresponding chromium and nickel equivalents for a
given alloy. The Delong diagram differs from the Schaeffler diagram in that the important
nitrogen contribution also is included in the former, thus allowing a more accurate prediction
of the weld metal delta ferrite content.
Example (7.6)
Consider plasma arc butt welding of a 5mm thick plate of austenitic stainless steel under con-
ditions similar to those employed in Example (7.5). Data for the base metal (BM) and the
filler wire (FW) chemical compositions are given in Table 7.4. Use this information together
with the Delong diagram in Fig. 7.53 to determine which of the two filler wires (I or II) that
provides the highest resistance against weld metal solidification cracking under the prevailing
circumstances. In the present example we shall assume that the mixing ratio B/(B+D) is 0.57
(the mixing ratio is defined in Section 1.10.8, Chapter 1).
Solution
In the absence of oxidation losses, the weld metal chemical composition is given by the 'rule
of mixtures':
3-D heat flow Peak temperature: 13000C
conditions
Absorbed energy, J
Test temperature, 0C
Fig. 7.52. CVN transition curves for a duplex stainless steel after weld thermal simulation. Data from
Videm.81
Table 7.4 Chemical composition of base plate and filler wires used in Example (7.6).
This leads to the weld metal chemical compositions shown in Table 7.5. We can now
calculate the Cr- and Ni-equivalents for both welds:
Nickel equivalent (%Ni+30x%C+30x%N+0.5x%Mn)
Austenite
Austenite+
martensite Austenite+ferrite
Fig. 7.53. The Delong diagram85 showing the relationship between delta ferrite content and weld metal
chemical composition for stainless steels.
From this we see that wire I provides the highest resistance against weld metal solidifica-
tion cracking for the combination of steel and operational conditions considered above.
It should be emphasised that the Delong diagram gives no information about the real solidi-
fication microstructure, since it is based on measurements of retained delta ferrite at room
temperature. Also, the important effect of cooling rate on the weld metal transformation be-
haviour is neglected in the present analysis. Consequently, the use of such empirical diagrams
for selection of steel and welding consumables is a keenly debated question in the scientific
literature.
7.4 Aluminium Weldments
Aluminium alloys are to an increasing extent used as structural components in welded assem-
blies because of their high strength, low density, and good resistance against general corro-
sion. In certain cases the application of aluminium is restricted by a low HAZ strength level
due to softening reactions occurring during welding. In other cases the cracking resistance or
the fatigue strength becomes the limiting factor, depending on the design criterion. Table 7.6
summarises typical problems associated with welding of aluminium and its generic alloys. In
the following, we shall focus on the structural and mechanical response of age-hardenable
aluminium alloys to the heat of welding processes, with particular emphasis on Al-Mg-Si
alloys and Al-SiC metal matrix composites.
Chemical Composition
Material Major Elements Minor Elements Welding Problems
Al-Mg system
Example (7.7)
Consider plasma arc butt welding of a 10mm thick aluminium plate of type AA 6082-T6
(containing 0.7 wt% Mg and 0.9 wt% Si) under the following conditions:
Experience shows that the bead reinforcement amounts to 10% of the groove cross section
(details of the groove geometry are given in Fig. 7.56). Two different filler wires are available,
wire I with 5 wt% Si and wire II with 5 wt% Mg. Use this information along with the diagrams
in Fig. 7.54 to determine which of the two filler wires (I or II) that provides the highest resist-
ance against weld metal solidification cracking under the prevailing circumstances. In these
calculations we shall assume that the temperature field around the heat source is given by the
simplified Rykalin thin plate solution (equation (1-100) in Chapter 1). Thermal data for Al-
Mg-Si alloys are contained in Table 1.1.
Solution
First we need to estimate the mixing ratio BI(B+D). The total width of the fusion zone (2ym)
can be obtained from equation (1-100):
(a)
(b)
Fig. 7.55. Examples of solidification cracking in aluminium weldments (Varestraint test coupons);
(a) Al-I wt% Si, (b) Al-I wt% Mg. After Cross.87
Fig. 7.56. Groove geometry for a single pass Al-Mg-Si butt weld (Example (7.7)).
This gives:
and
from which
In the absence of oxidation losses, the weld metal chemical composition is given by the
'rule of mixtures':
A comparison with Fig. 7.54 shows that wire I provides the highest safety against weld
metal solidification cracking under the prevailing circumstances, while wire II is unaccept-
able.
The lower cracking resistance of the Al-Mg wire compared with the Al-Si wire at high
dilution ratios arises from the pertinent difference in the fraction of eutectic liquid which forms
during weld metal solidification. For pure binary alloys the eutectic fraction feuL is given by
equation (3-46) in Chapter 3:
where CeuL is the eutectic concentration, and ko is the equilibrium partitioning coefficient (given
by the binary Al-Si and Al-Mg phase diagrams).
If the contribution from the accompanying alloying element is neglected, the values of feuL
become:
From this we see that the fraction of eutectic liquid in weld I is so abundant that it backfills
and 'heals' all incipient cracks, while feut in weld II is just large enough to form continuous
films at the columnar grain boundaries which, in turn, promotes solidification cracking.
Liquid (L)
Temperature
Composition
Fig. 7.57. Schematic binary phase diagram defining the equilibrium conditions for partial melting during
reheating.
higher than Cmax^ and the alloy is heated to a temperature above TeuU partial melting will
occur. Since the eutectic phase is usually located at the grain boundaries, these sites become
immediately covered with liquid films if the wetting conditions are favourable.
Liquid
Liquid+solid
Temperature, 0C
ss
Al+Si
Fig. 7.58. Section of the Al-rich comer of the binary Al-Si phase diagram. Data from Ref. 95.
(a) (b)
(C) (cO
Fig. 7.59. Optical micrographs showing the microstructural evolution during homogenisation heat treat-
ment of binary Al-Si alloys; (a) Isolated Si particle embedded in a matrix of Al, (b) Partially melted Si
particle surrounded by Al-Si eutectic (25s at 582°C), (c) Globular Al-Si eutectic structure formed after
complete dissolution of the Si particle (60s at 582°C), (d) Spreading of eutectic liquid along a grain
boundary (25s at 582°C). Courtesy of O. Lohne, Sintef - Division of Metallurgy, 7034 Trondheim,
Norway.
Further heating to a higher temperature (T > 593°C) provides additional time for dissolu-
tion of Mg2Si and formation of more liquid of variable composition. In practice, reaction (7-
Temperature, 0C
Fig. 7.60. Quasi-binary section of the Al-rich corner of the ternary Al-Mg-Si phase diagram. Data from
Reiso et al.96 and Phillips.97
22) may also proceed below the quasi-binary eutectic temperature Teut, since the diffusion rate
of Mg is higher than that of Si.96 By considering the kinetics it can be shown that the interface
concentration will gradually move towards the silicon side of the quasi-binary line in the phase
diagram as the Mg2Si particles dissolve, thereby reducing the temperature at which liquation
occurs (from 593 down to about 5800C). Moreover, in alloys containing excess amounts of
silicon, two other side reactions can take place:96
liquid (7-23)
and
liquid (7-24)
The former is analogous to the melting reaction in binary Al-Si alloys (Fig. 7.58), while the
latter corresponds to the eutectic reaction in the ternary Al-Mg-Si system (Teut varies from
555 to 559°C, depending on the source). Since this eutectic temperature represents the lowest
temperature at which a melt can exist within the system, it means that local melting of second
phase particles cannot take place below, say, 555 to 559°C during welding of Al-Mg-Si al-
loys.
7.4.2.3 Factors affecting the hot cracking susceptibility
In addition to the thermodynamic and kinetic effects mentioned above, there are several other
factors, some interrelated, which play an important part in the formation of hot cracks in
Al-Mg-Si weldments. These are:76
(i) The number density and size distribution of Mg2Si and Si particles in the base
metal.
(ii) The total grain boundary area per unit volume available for absorption of eutectic
liquid (determined by the HAZ grain size).
(iv) The local tensile stress level in the partially melted region.
It follows that the hot cracking susceptibility may be significantly altered by a change in
one of these parameters, but, in practice, there is very little that can be done to prevent grain
boundary liquation if the base material already contains second phase particles. In fact, the
only useful way of eliminating the cracking problem is to reduce the tensile stress in the par-
tially melted HAZ through proper selection of welding consumables. This is because the
composition of the weld metal can be adjusted so that solidification is completed first in the
partially melted region and then in the fusion zone, thus avoiding hot cracking in the former.
As an illustration of principles, the diagrams of Gittos and Scott98 (reproduced in Fig. 7.61)
will be considered. These diagrams show the variation of the weld metal solidus temperature
with base metal dilution for two commercial filler wires (wire I: Al-5 wt% Si, and wire II: Al-
5 wt% Mg). It is evident that the risk of hot cracking is highest when wire II is used, particu-
larly at high B/(B+D) ratios, since solidification occurs first in the weld metal and then in the
partially melted region. In contrast, wire I provides a good HAZ cracking resistance over the
whole composition range because of the resulting lower solidus temperature. In the latter case
the weld metal solidification and thermal contraction stresses are imposed on the HAZ at a
stage where liquid no longer exists at the grain boundaries.
Example (7.8)
Consider plasma arc butt welding of a 10mm thick aluminium plate of AA6082-T6 (contain-
ing 0.7 wt% Mg and 0.9 wt% Si) under conditions similar to those employed in Example (7.7).
Use the diagrams in Fig. 7.61 to determine which of the two filler wires (I or II) that provides
the highest resistance against hot cracking in the partially melted region under the prevailing
circumstances.
Solution
In the previous example the base metal dilution ratio, BI(B+D), was found to be 0.78. Under
such conditions wire I provides the highest resistance against hot cracking, since solidification
first occurs in the partially melted region and then in the weld metal. This explains why Al-Si
based filler wires are usually recommended for single pass butt welding of Al-Mg-Si extru-
sions if the strength level is not of particular concern.
(a)
Solidus temperature, 0C
liquid
B/(B+D), %
(b)
Solidus temperature, 0C
liquid
B/(B+D), %
Fig. 7.61. Variation of weld metal solidus temperature with dilution for 6082 aluminium alloys (Al-0.7
wt% Mg-0.9 wt% Si); (a) Wire I (Al-5 wt% Si), (b) Wire II (Al-5 wt% Mg). The melting temperature
for different base metal constituent phases are indicated by the horizontal broken lines in the diagrams.
Data from Gittos and Scott98 and Reiso et al.96
(a) Peak temperature
HAZ-
(b)
C-curve for
Temperature
precipitation
of (3'(Mg2 Si)
Time
Fig. 7.62. Schematic diagrams showing the sequence of reactions occurring in the HAZ of 6082-T6
aluminium welds; (a) Hardness distribution following p"(Mg2Si) dissolution; (b) Precipitation of P'(Mg2Si)
at dispersoids during the weld cooling cycle.
7A3 HAZ microstructure and strength evolution during fusion welding
The age-hardenable Al-Mg-Si alloys have been widely studied, to the extent that most of the
underlying physical processes are well established. They offer tensile strength values higher
than 350 MPa in the artificially aged (T6) condition owing to the presence of very fine, needle-
shaped 3"(Mg2Si) precipitates along <100> directions in the aluminium matrix." Although
Al-Mg-Si alloys are readily weldable, they suffer from severe softening in the heat affected
zone (HAZ) because of reversion (dissolution) of the (3"(Mg2Si) precipitates during the weld
thermal cycle.76100"103 This type of mechanical impairment represents a major problem in en-
gineering design, since it reduces the load-bearing capacity of the joint.104
Contribution;
from natural:
ageing \
Hardness
Resulting hardness
profile
Fig. 7.62. Schematic diagrams showing the sequence of reactions occurring in the HAZ of 6082-T6
aluminium welds (continued); (c) Hardness distribution after prolonged room temperature ageing.
7.4.3.2 Strengthen ing mechan isms in A l-Mg-Si alloys
Due to the lack of experimental evidence of coherency strains around (3"(Mg2Si) precipitates
in Al-Mg-Si alloys," it has been suggested that the increased resistance to dislocation motion
accompanying the presence of these structures arises from the high energy required to break
Mg-Si bonds in the particles as dislocations shear through them. Assuming that this strength-
ening effect is associated with order hardening, the net precipitation strength increment, Aop,
can be calculated from the equation originally derived by Kelly and Nicholson:105
(7-25)
where 7/ is the internal interface (or antiphase) boundary energy, b is the Burgers vector,/is
the particle volume fraction, and c$ is a kinetic constant.
By introducing the relative particle volume fraction, f/fo, we obtain:
(7-26)
where fo is the initial volume fraction of (3"(Mg2Si) precipitates in the alloy, and C4 is a new
kinetic constant (equal to c 3 / o ).
It is evident from equation (7-26) that Aop = Aap (max) = c4 when/7/ o = 1. Hence, this
equation can be rewritten as:
(7-27)
Here amin denotes the intrinsic matrix strength after complete particle dissolution, while
dmax is the original base metal strength in the artificially aged (T6) condition.
Provided that a linear relationship exists between yield strength and hardness, equation (7-
27) can be rewritten as:102
(7-28)
Equation (7-28) provides a basis for assessing the reaction kinetics through simple hardness
measurements. Typical values for amax, HVmax, omin, and HVmin are given in Table 7.7.
(7-29)
where t\ is the maximum hold time required for complete particle dissolution at a given tem-
perature (defined by equation (4-31) in Chapter 4), and n is a time exponent (< 0.5). The
variation of n with/// o is shown in Fig. 4.20.
(7-30)
where
(7-31)
Here O is a material constant (defined in Table 7.7), and t\ is the critical time required to
precipitate a certain fraction of $'(X = Xc) at an arbitrary temperature (T). The variation of
t*2 with temperature is given by equation (6-55) in Chapter 6.
The net precipitation increment following natural ageing (a2) can therefore be written as:
(7-32)
Coupling of models
Based on equations (7-29) and (7-32) it is possible to calculate the HAZ strength distribution
after welding and subsequent natural ageing when the weld thermal programme is known.
Figure 7.63 shows a sketch of the superimposed hardness profiles, as evaluated from these
equations. Since the resulting strength level in the partly reverted region depends on the inter-
play between two competing processes (i.e. dissolution and reprecipitation), it is convenient to
define the 'boundary' between the two models on the basis of the intersection point in Fig.
7.63 where a\ = a 2 , i.e.:
(7-33)
and
(7-34)
Peak temperature
Partly reverted
Fully reverted
region
region
Reversion model
Hardness
Intersection
point
Accuracy of predictions
Examples of measured and predicted HAZ hardness profiles are shown in Figs. 7.64 and 7.65.
When stringer bead welding is carried out on a plate of medium thickness, the hardness
distribution in the transverse y direction will vary with distance from the plate surface due to a
continuous change in the heat flow conditions. A comparison between observed and predicted
hardness profiles in Fig. 7.64 shows that such effects are readily accounted for in the present
model. In contrast, a full penetration butt weld will always reveal a similar HAZ hardness
distribution in the transverse section of the weld, as shown in Fig. 7.65. This situation arises
from the lack of a temperature gradient in the through-thickness z direction of the plate. Moreo-
ver, it is evident from Figs. 7.64 and 7.65 that the final dimensions of the HAZ are strongly
influenced by variations in welding parameters and operational conditions. Hence, it is diffi-
cult to justify the use of a constant safety factor for the width of the HAZ as recommended in
current design rules for welded Al-Mg-Si alloys.104
Aptness of models
Based on the kinetic models described in the proceding sections, it is possible to construct
two-dimensional (2-D) maps which show characteristic hardness and peak temperature con-
tours in the HAZ of 6082-T6 aluminium weldments. Examples of such diagrams are given in
(a)
HV (predicted)
Peak temperature, 0C
Hardness, VPN
Aym, mm
(b)
HV (predicted)
Peak temperature. 0C
Hardness, VPN
ym,mm
Fig. 7.64. Comparison between measured and predicted HAZ hardness profiles in a stringer bead GMA
weld; (a) Upper plate surface, (b) Lower plate surface. The peak temperature distribution is indicated by
the broken lines in the graphs. (Operational conditions: qo = 9.1 kW, v = 5.1 mm s"1, d = 15mm). Data
from Myhr and Grong.102
Fig. 7.66. Included is also a 3-D plot of the HAZ hardness distribution in the transverse
section of the weld.
The results in Fig. 7.66 reveal a direct relationship between the HAZ isothermal contours
on the one hand and the resulting HAZ hardness/strength distribution on the other. In this
particular example the soft zone closely follows the contour of the 4000C isotherm. This, in
turn, implies that the minimum HAZ strength level is fairly constant and virtually independent
of choice of welding parameters (i.e. close to 60 VPN for single pass welds).
HV (predicted)
Peak temperature. 0C
Hardness, VPN
A y m , mm
Fig. 7.65. Comparison between measured and predicted HAZ hardness profiles in a single pass plasma
arc butt weld. The peak temperature distribution is indicated by the broken line in the graph. (Opera-
tional conditions: qo = 14.0 kW, v = 5.8 mm s"1, d = 13mm). Data from Myhr and Grong.102
In practice, the HAZ hardness can be converted into an equivalent yield or ultimate tensile
strength through the following regression formulae:102
(7-35)
and
(7-36)
From equation (7-35) we see that a minimum HAZ hardness of about 60 VPN corresponds
to a strength reduction factor of:
This value is in good agreement with the recommended strength reduction factor of 0.49
incorporated in many welding specifications and standards.104'107
Hardness code
z, mm
Fusion zone
(a)
(b)
Fig. 7.66. Computed HAZ hardness and peak temperature contours in the transverse section of a stringer
bead GMA weld; (a) 2-D graphical representation, (b) 3-D graphical representation. (Operational condi-
tions as in Fig. 7.64). Data from Myhr and Grong.102
Figure 7.67 shows plots of the HAZ hardness/strength profiles for different values of qo Ivd.
It follows that a narrow width of the HAZ requires the use of a low energy input per mm2 of the
weld. In practice, this can be achieved by the choice of an efficient welding process (e.g.
electron beam or laser welding) which allows deposition of a full penetration butt weld with-
out employing a groove preparation (i.e. eliminates the need for filler metals).
Multipass welding
The present process model can also be extended to multipass welding if it is assumed that
reversion of indigenous (3'(Mg2Si) precipitates occurs instantaneously on reheating above the
phase boundary solvus temperature (here taken equal to 5200C).
Hardness, VPN V2[MPa] Rm [MPa]
Scale:
10 mm
Example (7.9)
Consider plasma arc butt welding of a 10mm thick aluminium plate of type AA 6082-T6 under
conditions similar to those employed in Example (7.7). Estimate on the basis of the process
diagram in Fig. 7.67 and the simplified Rykalin thin plate solution (equation (1-100) in Chap-
ter 1) both the minimum HAZ strength level, the total width of the reduced strength zone after
welding, as well as the lower temperature limit for dissolution of the (3"(Mg2Si) precipitates
during the weld thermal cycle. Thermal data for Al-Mg-Si alloys are given in Table 1.1
(Chapter 1).
Solution
First we need to calculate the net heat input per mm2 (q ol vd). In the present example, we
have:
y, mm
Hardness code
z, mm
Fusion zone
(a)
(b)
Fig. 7.68. Computed HAZ hardness and peak temperature contours for a simulated two-pass butt weld
(the second pass is deposited immediately after cooling of the first pass); (a) 2-D graphical representa-
tion, (b) 3-D graphical representation. (Operationalconditions: qo = 9A kW, v = 5.1 mms~ l ,d- 15mm).
Data from Myhr and Grong.102
and
Similarly, by considering the extension of the HAZ and the corresponding length of the
scale bar in Fig. 7.67, the total width of the reduced strength zone becomes:
The relationship between peak temperature T- Tp and distance y = ym from the heat source
can now be obtained by differentiating equation (1-100) with respect to time. After some ma-
nipulation, we obtain:
By substituting qo/vd = 200 J ram"2, pc = 0.0027 J mm"3 0C"1, and Tm = 652°C into the
above equation, the following temperature for incipient dissolution of the p"(Mg2Si) precipi-
tates is obtained:
It is obvious from the above calculations that the degree of HAZ softening occurring during
welding is substantial under the prevailing circumstances. This explains why, for instance,
high heat input deposition is usually not recommended for Al-Mg-Si alloys.
(7-37)
where M is the interfacial torque, R* is the surface radius, and P(r) is the pressure distribution
across the interface (here assumed constant and equal to P).
Pressure Pressure
If all the shearing work at the interface is assumed to be converted into frictional heat, the
average heat input per unit area and time becomes:114
(7-38)
where qo is the net power (in W), P is the friction pressure (in N mm"2), A is the cross section
(in mm2), and wmax. is the maximum surface velocity at the outer edge (in m s"1).
Equation (7-38) provides a basis for estimating the heat generation at the interface during
continuous drive friction welding in the absence of asperity melting.
(i) The fully plasticised region, ZpL, where the material is able to accommodate the
plastic strain by dynamic recovery (or recrystallisation) of the microstructure.
Contact section
Z
Fig. 7.70. Schematic diagram showing the three main reaction zones within a friction welded component
(Zpi/. fully plasticised region, Zpd\ partly deformed region, Zud: undeformed region).
(ii) The partly deformed region, Zpd., where the degree of plastic deformation is ac-
commodated by an increase in the dislocation density of the matrix grains. In this
region the temperature is sufficiently high to facilitate dissolution of the base metal
hardening precipitates.
(iii) The undeformed region, Z1^., characterised by partial reversion of the base metal
precipitates.
Aspects of HAZ subgrain evolution during continuous drive friction welding have been
described in Section 6.5.2 (Chapter 6). In the following, the structural and mechanical re-
sponse of T6 heat treated Al-Mg-Si alloys and Al-SiC metal matrix composites to the im-
posed heating and plastic deformation will be considered more in detail.
(a)
Reversion model
Hardness
(b)
Reversion model
Hardness
Axial distance, Z
Fig. 7.71. Schematic representation of the HAZ hardness distribution after friction welding and subse-
quent natural ageing; (a) Short duration thermal cycle, (b) Long duration thermal cycle. The parameters
Zp/., Zp(L and Zud% are defined in Fig. 7.70.
7.4.4.5 Prediction of the HAZ hardness distribution
The predictions are based on computer programmes which utilise the heat and material flow
models described in Ref.113 in combination with the constitutive equations given above to
calculate the HAZ hardness distribution for specific welding conditions.
Accuracy of predictions
Examples of measured and predicted hardness profiles are given in Figs. 7.72 and 7.73.
A closer inspection of the graphs reveals a good agreement between theory and experi-
ments in all three cases. It is interesting to note that there is no clear distinction in the shape of
the HAZ hardness profiles between friction welded Al-Mg-Si alloys and Al-SiC metal matrix
composites when comparison is made on the basis of a similar temperature-time pattern (see
Fig. 7.73). However, large diameter weld components will normally reveal a different hard-
ness distribution, as shown in Fig. 7.72, because longer welding times will increase the total
heat input. Under such conditions the contribution from the plastic deformation becomes
negligible, which means that the resulting HAZ hardness profile will closely resemble that
observed during conventional gas metal arc (GMA) and plasma arc welding of Al-Mg-Si
alloys (see Figs. 7.64 and 7.65).
Process diagrams
Based on the above process model, it is possible to construct a series of diagrams which sum-
marise information about the effect of important welding variables in a systematic and illustra-
tive manner. Examples of such diagrams for 6082-T6 aluminium alloys and T6 heat treated
Al-SiC metal matrix composites are given in Fig. 7.74(a) and (b), respectively.
Hardness, VPN
Predicted
Measured
Unaffected base
material
Axial distance, mm
Fig. 7.72.Comparison between measured and predicted HAZ hardness profiles in a <£26mm Al-Mg-Si
weld component. (Assumed input data: <E> = *F = 0.56, HVmax= 110, HVmin = 42). Operational condi-
tions: qJA = 17W m m 2 and ts = 6s. Data from Midling and Grong.113
(a)
Hardness, VPN
Predicted
Measured
Axial distance, mm
(b)
Hardness, VPN
Predicted
Measured
Axial distance, mm
Fig. 7.73.Comparison between measured and predicted HAZ hardness profiles; (a) 4> 16mm Al-Mg-Si
weld component. (Operational conditions: qJA - 25W mrrr2 and ts = 0.9s), (b) <l>16mm friction welded
Al-SiC metal matrix composite. (Operational conditions: qJA = 25W mirr 2 and ts= 3.8s). Data from
Midling and Grong.113
(a) Hardness, VPN
Axial distance, mm •
(b)
Hardness, VPN
Axial distance, mm
Fig. 7.74. Process diagrams for friction welding; (a) 6082-T6 aluminium alloys. (Operational condi-
tions: umax_ = 2.5m s"1 and [i = 0.5), (b) Al-SiC-T6 metal matrix composites. (Operational conditions:
Umax. = 2.5m s"1 and \x = 0.5). Data from Midling and Grong.113
It is evident from these diagrams that the HAZ hardness distribution depends on the total
heat input applied during friction welding. Although the controlling parameters qo IA and ts
(welding time), in practice, are kept within relatively narrow limits, it is obvious that a small
width of the HAZ requires the use of a high specific power (qol A) in combination with a short
duration heating cycle (ts < 2 s). This is also in agreement with general experience.109"112
Example (7.10)
Consider continuous drive friction welding of a T6 heat treated Al-SiC metal matrix compos-
ite under the following conditions:
Use the process diagram in Fig. 7.74(b) to estimate the minimum HAZ hardness level as
well as the total width of the strength reduced zone after welding. In this example we shall
assume that the friction coefficient Ji is equal to 0.5.
Solution
First we need to calculate the frictional heat per unit area of the weld. From equation (7-38),
we have:
A comparison with Fig. 7.74(b) shows that a specific power of 25 W mm"2 corresponds to
a minimum HAZ hardness of about 90 VPN, i.e. a reduction of 45 VPN compared with the
base material. At the same time the total width of the reduced strength zone is seen to be
12 mm.
It should be emphasised that the observed strength loss is not permanent, since the resulting
HAZ strength level is mainly controlled by dissolution reactions taking place within the alu-
minium matrix during the weld thermal cycle. Consequently, a full HAZ strength recovery
can be achieved by the use of an appropriate post weld heat treatment, as shown by the tensile
test data in Table 7.8.
Table 7.8 Mechanical properties of friction welded Al-SiC metal matrix composites. Data from
Midling and Grong109'113
HV R R e
p0 2 m B
Material [VPN] [MPa] [MPa] [%]
Solution heat treated at 535°C for 3 h followed by water-quenching and artifical ageing at 1600C for 10 h.
References
1. R. W.K. Honeycombe: Steels — Micro structure and Properties, 1980, London, Edward Arnold
(Publishers) Ltd.
2. H. Suzuki: Weld. World, 1982, 20, 121-148.
3. 0 . Grong and D.K. Matlock: Int. Met. Rev., 1986, 31, 27-48.
4. DJ. Abson and RJ. Pargeter: Int. Met. Rev., 1986, 31, 141-194.
5. RL. Harrison and R.A. Ferrar: Int. Met. Rev., 1989, 34, 35-51.
6. GJ. Davis and J.G. Garland: Int. Met. Rev., 1975, 20, 83-106.
7. S.A. David and J.M. Vitek: Int. Mater. Rev., 1989, 34, 213-245.
8. T. Gladman and RB. Pickering: In Yield, Flow and Fracture of Poly crystals (Ed. T.N. Baker),
1983, London, Applied Science Publishers, 141-198.
9. A.O. Kluken, M.I. Ons0ien, O.M. Akselsen and G. R0rvik: Joining ScL, 1991,1, 14-22.
10. R Deb, K.D. Challenger and A.E. Therrien: Metall. Trans., 1987,18A, 987-999.
11. RT. Odland, CW. Ramsay, D.K. Matlock and D.L. Olson: WeIdJ., 1989, 68, 158s-168s.
12. DJ. Widgery: Weld J, 1975, 54, 57s-68s.
13. L. Devillers, D. Kaplan, B. Marandet, A. Ribes and RV. Riboud: Proc. Int. Conf. on Effects of
Residual, Impurity and Microalloying Elements on Weldability and Weld Properties, London,
Nov. 1983, Paper 1, Publ. The Welding Institute.
14. R.A. Farrar: Welding and Metal Fabr., 1976, 44, 578-581.
15. 0. Grong and A.O. Kluken: In Ferrous Alloy Weldments, (Eds D.L. Olson and T.H. North),
1992, Zurich (Switzerland), Trans. Tech. Publications, 21-46.
16. R.H. Van Stone, T.B. Cox, J.R. Low and J.A. Psioda: Int. Met. Rev., 1985, 30, 157-179.
17. A.W Thompson: Acta Metall, 1983, 31, 1517-1523.
18. I. Andersen: MSc Thesis, 1989, Division of Metallurgy, The Norwegian Institute of Technol-
ogy, Trondheim, Norway.
19. G.E. Dieter: Mechanical Metallurgy, 3rd edn, 1986, New York, McGraw-Hill Book Company.
20. O.M. Akselsen and 0. Grong: Mater. ScL Eng., 1992, A159, 187-192.
21. D.E. McRobie and J.F. Knott: Mater. ScL TechnoL, 1985,1, 357-365.
22. 0. Grong, A.O. Kluken and B. Bj0rnbakk: Joining and Materials, 1988, l, 164-169.
23. A.O. Kluken and 0 . Grong: Proc. 3rd Int. Conf. on Trends in Welding Research, Gatlinburg,
TN, June, 1992, pp. 569-574. Publ. ASM International, Materials Park, Ohio (1993).
24. J.I. Ustinovshchikov: Acta Metall., 1983, 31, 355-364.
25. AJ. Ardell: Acta Metall., 1972, 20, 601-609.
26. O.M. Akselsen, G. R0rvik, M.I. Ons0ien and 0 . Grong: Weld J, 1989, 68, 356s-362s.
27. H. Suzuki: Trans. Jap. Weld. Soc, 1981,16, 25-32.
28. N. Yurioka, S. Ohshita and H. Tamehiro: Proc. Int. Symposium on Pipeline Welding in the
'80s, March, 1981, pp. 1-15, Publ. Australian Welding Research Association.
29. C.L.M. Cottrell; Metal Constr., 1984,16, 740-744.
30. HJ.U. Cotton: Metal Constr., 1987,19, 217R-223R.
31. T. Kasuya and N. Yurioka: WeIdJ, 1993, 72, 263s-268s.
32. A.O. Kluken, S. Ibarra, S. Liu and D.L. Olson: Proc. 11th Int. Conf. on Offshore Mechanics
and Arctic Engineering, 1992, Publ. ASME, Book No. H0744A-92.
33. O.R. Myhr and 0 . Grong: Acta Metall. Mater., 1990, 38, 449-460.
34. O.M. Akselsen and G. R0rvik: Mater. ScL TechnoL, 1990, 6, 383-389.
35. J.H. Hollomon and L.D. Jaffe: Trans. AIME, 1945,162, 223-249.
36. K. Olsen, D.L. Olson and N. Christensen: Scand. J. Metall, 1982,11, 163-168.
37. PJ. Alberry: Weld J, 1989, 68, 410s-417s.
38. R Ravi Vishnu and K.E. Easterling: In Mathematical Modelling of Weld Phenomena (Eds H.
Cerjak and K.E. Easterling), 1993, London, The Institute of Materials, pp. 241-299.
39. O.D. Sherby and J.E. Dorn: Trans. AIME, 1953,197, 324-330.
40. T. Reti, M. Gergely and R Tardy: Mater. ScL Technol., 1987, 3, 365-371.
41. 0 . Grong and O.M. Akselsen: Metal Constr., 1986,18, 557-562.
42. C. Thaulow, AJ. Paauw, A. Gunleiksrud and OJ. Naess: Metal Constr, 1985,17, 94R-99R.
43. O.M. Akselsen, 0. Grong and G. R0rvik: Scand. J. MetalL, 1990,19, 258-264.
44. C. Thaulow, AJ. Paauw and K. Guttormsen: Weld J, 1987, 66, 266s-279s.
45. O.M. Akselsen, J.K. Solberg and 0. Grong: Scand. J. MetalL, 1988,17, 194-200.
46. O.M. Akselsen, 0 . Grong and P.E. Kvaale: MetalL Trans., 1986,17A, 1529-1536.
47. H. Homma, S. Ohkita, S. Matsuda and K. Yamamoto: Weld J, 1987, 66, 301s-309s.
48. Y-T. Pan and J-L. Lee: Proc. 3rd. Int. Conf. on Trends in Welding Research, Gatlinburg (Ten-
nessee), June, 1992, pp. 539-543. Publ. ASM International, Materials Park, Ohio (1993).
49. J.H.Chen,Y.Kikuta,T.Araki,M.YonedaandY.Matsuda: ActaMetalL, 1984,32,1779-1788.
50. C.A.N. Lanzillotto and RB. Pickering: Met. ScL, 1982,16, 371-382.
51. O.M. Akselsen, 0 . Grong and J.K. Solberg: Mater. ScL Technol., 1987, 3, 649-655.
52. H.P. Shen, T.C. Lei and J.Z. Liu: Mater. ScL Technol., 1986, 2, 28-33.
53. NJ. Kim and G. Thomas: MetalL Trans., 1981,12A, 483-489.
54. A.F. Szewezyk and J. Garland: MetalL Trans., 1982,13A, 1821-1826.
55. M. Ramberg, O.M. Akselsen and 0. Grong: Proc. 1st Int. Conf. on Trends in Welding Re-
search, Gatlinburg, TN, May, 1986, pp. 679-685. Publ. ASM International, Metals Park, Ohio
(1987).
56. N.E. Hannertz: Schweissen & Schneiden, 1976, 28, 379-382.
57. RR. Coe: Welding Steels Without Hydrogen Cracking, 1973, Abington (Cambridge), The
Welding Institute.
58. N. Yurioka and H. Suzuki: Int. Mater. Rev., 1990, 35, 217-249.
59. LT. Brown and W.M. Baldwin: Trans. AIME, 1954, 200, 298-303.
60. E.A. Steigerwald, RW. Shaller and A.R. Troiano: Trans. AIME, 1959, 215, 1048-1052.
61. CA. Zapffe and CE. Sims: Trans. AIME, 1941,145, 225-237.
62. NJ. Petch: Phil. Mag., 1956,1, 331-337.
63. CD. Beachem: MetalL Trans., 1972, 3, 437-451.
64. A.W. Thompson and LM. Bernstein: In Effects of Hydrogen on the Behaviour of Metals (Eds
LM. Bernstein and A.W. Thompson), 1980, Warrendale, PA, Metallurgical Society of AIME,
pp. 291-308.
65. J.D. Fast: Gases in Metals, 1976, London, Macmillan Press Ltd.
66. R.A. Oriani: Ada MetalL, 1970,18, 147-157.
67. N. Christensen, K. Gjermundsen and R. Rose: Brit. WeIdJ., 1958, 5, 272-281.
68. N. Christensen: Svetsen, 1975, 34, 22-32.
69. N. Yurioka and S. Ohshita: HW Doc. IX-1161-80 (1980).
70. C Zhang and J.A. Goldak: HW Doc. IX-1662-92 (1992).
71. N. Christensen and T. Simonsen: Scand. J. MetalL, 1981,10, 120-126.
72. NACE Standard MR-01-75 (1975).
73. M.I. Ons0ien, O.M. Akselsen, 0. Grong and P.E. Kvaale: WeIdJ., 1990, 69 (No. 1), 45-51.
74. H. Suzuki: Trans. Jap. Weld. Soc, 1981,16, 25-32.
75. D.L. Olson: Weld J, 1985, 64, 281s-295s.
76. S. Kou: Welding Metallurgy, 1987, New York, John Wiley & Sons, Inc.
77. R.D. Campbell: In Ferrous Alloy Weldments (Eds D.L. Olson and T.H. North), 1992, Zurich,
Trans. Tech. Publications Ltd., pp. 167-216.
78. I. Varol, J.C Lippold and W.A. Baeslack: ibid., pp. 217-252.
79. R. Mundt and H. Hoffmeister: Arch. Eisenhuttenwes., 1983, 54, 253-256; ibid. 333-336.
80. S. Atamert and J.E. King: Mater. ScL Technol., 1992, 8, 896-911.
81. M. Videm: MSc. Thesis, 1985, Department of Metallurgy, The Norwegian Institute of Tech-
nology, Trondheim, Norway.
82. J.A. Brooks, A.W. Thompson and J.C. Williams: WeldJ, 1984, 63, 71s-83s.
83. A.L. Schaeffler: Metal Prog., 1949, 56, 680-680B.
84. W.T. Delong, G. Ostrom and E. Szumachowski: WeIdJ., 1956, 35, 526s-533s.
85. CJ. Long and W.T. Delong: WeIdJ., 1973, 52, 281s-297s.
86. J.H. Dudas and RR. Collins: Weld J, 1966, 45, 241s-249s.
87. CE. Cross: Ph.D Thesis, 1986, Colorado School of Mines, Golden, Colorado, USA.
88. JJ. Pepe and W.F. Savage: Weld J, 1967, 46, 41 ls-422s.
89. JJ. Pepe and W.F. Savage: Weld J, 1970, 49, 545s-553s.
90. O. Reiso: Proc. 3rd. Int. Conf. on Aluminium Extrusion Technology, Atlanta, GA, 1984, vol. 1,
pp. 31-40. Publ. Aluminium Association (1984).
91. O. Reiso: Proc. 4th Int. Aluminium Extrusion Technology Seminar, Chicago, IL, 1988, vol. 2,
pp. 287-295. Publ. Aluminium Association (1988).
92. O. Reiso, H.G. 0verlie and N. Ryum: Metall. Trans., 1990, 21A, 1689-1695.
93. H. Gjestland, A.L. Dons, O. Lohne and O. Reiso: In Aluminium Alloys — Their Physical and
Mechanical Properties, 1986, Warley (UK), Engineering Materials Advisory Service Ltd., pp.
359-370.
94. O. Lohne and N. Ryum: Proc. 4th Int. Aluminium Extrusion Technology Seminar, Chicago, IL,
1988, vol. 2, pp. 303-308. Publ. Aluminium Association (1988).
95. Metals Handbook, 8th Edition (vol. 8).
96. O. Reiso, N. Ryum and J. Strid: Metall. Trans. A, 1993, 24A, 2629-2641.
97. H.W.L. Phillips: Annotated Equilibrium Diagrams of Some Aluminium Alloy Systems, 1959,
London, The Institute of Metals, pp. 65-71.
98. N.F. Gittos and M.H. Scott: WeIdJ, 1981, 60, 95s-103s.
99. J.E. Hatch (Ed.): Aluminium — Properties and Physical Metallurgy, 1984, Ohio (USA), Ameri-
can Society for Metals.
100. T. Enjo and T. Kuroda: Trans. JWRI, 1982,11, 61-66.
101. S.D. Dumolt: Ph.D Thesis, 1983, Carnegie-Mellon University, USA.
102. O.R. Myhr and 0. Grong: Ada Metall. Mater., 1991, 39, 2693-2702; ibid, 2703-2708.
103. 0 . Grong and O.R. Myhr: In Mathematical Modelling of Weld Phenomena, (Eds H. Cerjak
and K.E. Easterling), 1993, London, The Institute of Materials, pp. 300-311.
104. FM. Muzzolani: Aluminium Alloy Structures, 1985, Boston (USA), Pitman Publishing Inc.
105. A. Kelly and R.B. Nicholson: Progr. Mat. ScL, 1963,10, 151-156.
106. O.R. Myhr, Ph.D Thesis, 1990, Division of Metallurgy, The Norwegian Institute of Technol-
ogy, Trondheim, Norway.
107. European Recommendations for Aluminium Alloy Structures, 1978.
108. G. Steidl and R. Mossinger: Aluminium, 1977, 53, 199-203.
109. O.T. Midling, 0. Grong and M. Camping: Proc. 12th Riso Int. Symp. on Materials Science:
Metal Matrix Composites—Processing, Micro structure and Properties, Roskilde, Denmark,
1991, pp. 529-534. Publ. Riso National Laboratory (1991).
110. O.T. Midling, 0. Grong and D.H. Bratland: Proc. 3rd Int. Conf. on Aluminium Alloys — Their
Physical and Mechanical Properties, Trondheim, Norway, 1992, pp. 99-105. Publ. The Nor-
wegian Institute of Technology, Department of Metallurgy (1992).
111. O.T. Midling and 0. Grong: Proc. 3rd Int. Conf. on Trends in Welding Research, June, 1992,
Gatlinburg, TN, pp. 1147-1151. Publ. ASM International (1993).
112. O.T. Midling and 0. Grong: Proc. Int. Conf Advanced Composites '93, Wollongon, Australia,
February 1993, pp. 1221-1226. Publ. The Minerals, Metals & Materials Society (1993).
113. O.T. Midling and 0. Grong: Acta Metall. Mater., 1994, 42, 1595-1609; ibid, 1611-1622.
114. N.N. Rykalin, A.I. Pugin, V.A. Vasil'eva: Weld. Prod, 1959, 6, 42-52.
115. B. Crossland: Cont. Phys., 1971,12, 559-574.
Appendix 7.1
Nomenclature
martensite-austenite constituent
welding time (s)
metal matrix composite
cooling time from 600 to 4000C (s)
time exponent
cooling time from 800 to 5000C (s)
dimensionless operating parameter in
cooling time from 1200 to 8000C (s)
heat flow model
equivalent isothermal hold time at
friction pressure (N mm"2, MPa) 333K (s)
Hollomon-Jaffe parameter temperature (0C, K)
area under stress-strain curve (J m~3) critical stress for particle cracking
(MPa)
ultimate tensile strength
radial stress (MPa)
welding speed (mm s"1)
threshold stress for H2S stress corrosion
inclusion volume fraction cracking (MPa)
partly deformed region in friction weld- dimensionless cooling time from 800
ing model (mm) to 5000C
8.1 Introduction
This chapter contains a collection of different exercise problems which the author has
adopted in his welding metallurgy course for graduate (mature) students. They illustrate how
the models described in the previous chapters can be used to solve practical problems of more
interdisciplinary nature. Each of them contains a 'problem description' and some background
information on materials and welding conditions. The exercises are designed to illuminate the
microstructural connections throughout the weld thermal cycle and show how the properties
achieved depend on the operating conditions applied. Solutions to the problems are also pre-
sented. These are not complete or exhaustive, but are just meant as an aid to the reader to de-
velop the ideas further.
Problem description
Consider gas metal arc (GMA) welding of low allow steels under the following conditions:
(i) Tack welding of a T-joint (Fig. 8.1)
(ii) Root pass deposition in a single V-groove (Fig. 8.2)
(iii) Root pass deposition in a X-groove (Fig. 8.3)
(iv) Deposition of cap layer during multipass welding (Fig. 8.4)
The materials to be welded are a C-Mn steel and a Nb-microalloyed low carbon steel with
chemical compositions and properties as listed in Tables 8.1 and 8.2. Details of welding par-
ameters and operational conditions are given in Table 8.3 and 8.4, respectively.
Table 8.1 Exercise problem I: Base plate chemical compositions (in wt%).
Steel C Si Mn P S Nb Al
1
C-Mn 0.20 0.35 1.46 0.003 0.002 - 0.037
1
LC-Nb 0.08 0.26 1.44 0.003 0.003 0.020 0.025
1
Ti: -0.008, N: 0.0027, Ca: 0.0040, B: 0.0002.
Table 8.4 Exercise problem I: Operational conditions and filler wire characteristics K
Analysis:
The students should work in groups (3 to 4 persons) where each group select a specific com-
bination of base material and welding conditions (e.g. deposition of a cap layer on the top of
a thick multipass C-Mn steel weld). The problem here is to evaluate the response of the base
material to heat released by the welding arc. The analysis should be quantitative in nature and
based on sound physical principles. The following points shall be considered:
(a) Select an appropriate heat flow model for the system under consideration.
(b) Estimate the minimum bead length which is required to achieve pseudo-steady state (i.e.
a temperature field that does not vary with position when observed from a point located
in the heat source).
(c) Estimate the value of the deposition coefficient kx (in gA " 1 S" 1 ), the weld cross section
areas D and B (in mm2), and the mixing ratio DI(B + D) during welding.
(d) Estimate the weld metal chemical composition. Calculate then the following quantities:
(e) Carry out a total oxygen balance for the system, and estimate the resulting CO content
in the welding exhaust gas.
(f) Estimate the chemical composition, volume fraction, and mean size (diameter) of the
oxide inclusions which form in the cold part of the weld pool. Calculate then the follow-
ing inclusion characteristics:
(g) Estimate the weld metal solidification mode and the resulting columnar grain mor-
phology. Indicate also the type of substructure which form at different positions from the
weld centre line.
(h) Evaluate the thermal stability of the base metal grain boundary pinning precipitates. At
which temperature will these precipitates dissolve?
(i) Calculate the austenite grain size profile across the HAZ. Estimate also the size of the
columnar austenite grains in the weld metal.
(j) Estimate the primary reaction products which form in the weld metal and the HAZ after
the austenite to ferrite transformation.
(k) Estimate the maximum hardness in the HAZ after welding. Use this information to eval-
uate the risk of hydrogen cracking and H2S stress corrosion cracking during service.
(1) Estimate the CVN toughness both in the weld metal and the HAZ after welding.
(m) Based on the results obtained explain why the carbon content of modern structural steels
has been gradually lowered to values below 0.1 wt% in step with the progress in steel
manufacturing technology.
Solution:
In all cases we can use stringer bead deposition on thick plates as a model system. It follows
from the analysis in Section 1.10.7 (Chapter 1) that the pertinent difference in the effective
heat diffusion area between a bead-on-plate weld and a groove weld may conveniently be ac-
counted for by introducing a correction factor/, which depends on the geometry of the groove
(see Fig. 1.68). Thus, in the general case the net (effective) power of the heat source can be
written as:
In the following, we shall only consider deposition of a cap layer on a thick plate where
/ = 1, but the analysis can readily be applied to other combinations of steels and welding con-
ditions as well (e.g./< 1). In the former case, we get:
Table 1.1 (Chapter 1) contains relevant input data for the steel thermal properties.
(a) The problem of interest is whether we must use the general (but complex) Rosenthal
thick plate solution (equation (1-45)) or can adopt the simplified solution for a fast moving
high power source (equation (1-73)). Fig 1.24 provides a basis for such an evaluation. The
most critical position will be the fusion line. If we neglect the latent heat of melting, the QJn3
ratio at the melting point becomes:
Readings from Fig. 1.24 suggest that the error introduced by neglecting the contribution
from heat flow in the welding direction is sufficiently small that it can be disregarded in the
calculations of the HAZ thermal programme. This means that equation (1-73) can be used in
replacement of equation (1-45) if that is desirable.
(b) The duration of the transient heating period depends on the actual point of observation
(i.e. the distance from the heat source). If we, as an illustration of principles, would like to
apply the pseudo-steady state solution down to a peak temperature of, say, 7000C, the corre-
sponding nJQ ratio at that temperature becomes:
From Fig. 1.21 we see that this ratio corresponds to a dimensionless radius vector a3m of
about 5. The duration of the transient heating period may now be read from Fig. 1.18. A crude
extrapolation gives:
from which
(c) The value of the deposition coefficient may be estimated from the data in Table 8.4.
The corresponding area of fused parent metal is most conveniently read from Fig. 1.21.
Taking the n3/Q ratio at the melting point equal to (1/0.22) ~ 4.5, we obtain:
from which
This value is somewhat lower than the expected mixing ratio, which for low heat input
welding is close to 0.67.
(d) The composition data in Table 8.4 refer to all weld metal deposit. Since the dilution with
respect to the base material in this case is small, the weld metal composition would be ex-
pected to be close to that given in Table 8.4.
An estimate of the total burn-off of alloying elements during welding can be obtained by
considering the difference in chemical composition between the filler wire and the weld
metal. In the present case we get:
Loss of silicon
As shown in Section 2.10.1.3 (Chapter 2), the silicon loss can partly be ascribed to SiO(g) for-
mation in the arc column (with consequent fume formation), and partly to reactions with oxy-
gen in the weld pool during the deoxidation stage (with consequent silicate slag formation).
The former loss can be estimated from the fume formation data presented in Table 2.6. Taking
the fume formation rate (FFF) of silicon equal to 63 mg min"1, the total loss of silicon in the
arc column amounts to:
Loss of manganese
As shown in Section 2.10.1.4 (Chapter 2), manganese is partly lost in the arc column due evap-
oration and partly in the weld pool due to deoxidation reactions. Taking the fume formation
rate of manganese equal to 14 mg min"1 (from Table 2.6), the total loss of Mn in the arc col-
umn amounts to:
The corresponding oxidation loss of manganese in the weld pool is thus:
Reading from Fig. 2.56 gives a residual oxygen content of about 0.07 wt%. The total oxy-
gen pick-up in the weld pool is thus:
From this we see that most of the oxygen which is picked up at elevated temperatures is
rejected again during cooling in the weld pool due to deoxidation reactions and subsequent
phase separation.
A comparison with Fig. 2.35 shows that the calculated weight of slag is in reasonable agree-
ment with experimental observations.
(e) The oxygen balance is carried out in accordance with the procedure outlined in Section
2.10.1.7 (Chapter 2). First we need to estimate the total mass of weld metal produced per unit
time:
The total CO2 consumption is thus:
Oxidation of carbon:
Oxidation of silicon:
Oxidation of manganese:
A comparison with the experimental data in Table 2.2 shows that the calculated CO con-
tent is of the expected order of magnitude.
(f) The deoxidation model in Section 2.12.4.1 (Chapter 2) can be used to estimate the inclu-
sion composition. From Fig. 2.68 we see that the inclusions are essentially pure manganese sil-
icates with an overall composition close to MnSiO3.
When the inclusion composition is known, it is possible to convert the residual weld metal
oxygen content into an equivalent inclusion volume fraction according to the procedure out-
lined in Section 2.12.1.Taking the stoichiometric conversion factor equal to 5.0 X 10~2 for
manganese silicate slags, we obtain:
Moreover, we can use equation (2-79) in Section 2.12.2.2 to calculate the mean diameter
of the inclusions:
The different inclusion characteristics may now be estimated from equations (2-80) to (2-83):
Number of particles per mm3:
A comparison with Table 2.11 shows that the calculated inclusion characteristics are in
reasonable agreement with those reported for C-Mn steel weld metals.
(g) The characteristic growth pattern of columnar grains in bead-on-plate welds is shown
schematically in Fig. 3.33. The first phase to form will be delta ferrite which subsequently de-
composes to austenite via a peritectic transformation (see Fig. 3.72). The important question is
whether re-nucleation of the grains will occur during solidification. In practice, this depends on
the interplay between a number of variables which cannot readily be accounted for in a sim-
plified analysis, including the weld pool geometry, the cooling rate and the nucleation potency
of the non-metallic inclusions. Broadly speaking, the energy barrier associated with nucleation
of delta ferrite at manganese silicates is rather high (e.g. see Fig. 3.30), which suggests that for-
mation of new grains ahead of the advancing solid/liquid interface is not very likely under the
prevailing circumstances. Hence, the columnar grain zone would be expected to extend entirely
from the fusion line towards the centre of the weld, as frequently observed in this type of welds.
Moreover, Fig. 3.43 provides a basis for estimating the substructure of the weld metal
columnar grains. Close to weld centre-line the local crystal growth rate will approach the
welding speed (i.e. RL ~ 4 mm s"1). At the same time a simple analytical solution exists for the
thermal gradient in the weld pool (equation (3-28)):
From this we see that a cellular-dendritic type of substructure is likely to form within the
central parts of the fusion zone, in agreement with general experience (see Fig. 3.36).
(h) Fig. 5.25 shows the location of the cap layer. Since the base plate is a Nb-microalloyed
steel, the important grain boundary pinning precipitates within the HAZ are either NbC, NbN
or a mixture of these. In the former case the equilibrium dissolution temperature may be es-
timated from the solubility product of the pure binary compounds. From equation (4-4) and
Table 4.1, we have:
and
This shows that NbC is thermodynamically more stable than NbN. In practice, the real dis-
solution temperature may be significantly higher than that predicted from equation (4-4) be-
cause of the kinetic superheating (see discussion in Section 4.4, Chapter 4). The grain growth
diagram in Fig. 5.21 (a) provides a basis for estimating the effect of heating rate (heat input)
on the dissolution kinetics. Taking the ordinate qo /v equal to 2678/4000 = 0.67 kJ mm"1, we
obtain:
In the HAZ on the weld metal side (see Fig. 5.25), oxide inclusions may act as effective
grain boundary pinning precipitates. These will be thermodynamically stable up to the melt-
ing point of the steel.
(i) The austenite grain size profile across the base plate HAZ can be read from Fig. 5.21(a).
Taking the ordinate q/v equal to 0.67 kJ mm"1, we see that the maximum austenite grain size
at the fusion boundary will exceed 100 /mm because of dissolution of the base metal grain
boundary pinning precipitates. In the HAZ on the weld metal side, the situation is different.
Here the stable weld metal oxide inclusions will impede austenite grain growth to a much
larger extent.The limiting austenite grain size may be calculated from equation (5-21).Taking
the Zener coefficient equal to 0.5 for oxide inclusions in steel (Fig. 5.4), we obtain:
Because of the phenomenon of epitaxial grain growth (see Section 3.3, Chapter 3), the in-
itial size of the weld metal delta ferrite/austenite columnar grains would be expected to be
comparable to the size of the HAZ austenite grains adjacent to the fusion boundary. Since the
latter varies along the periphery of the fusion boundary at the same time as competitive grain
growth leads to a general coarsening of the solidification microstructure with increasing dis-
tance from the fusion boundary, an average columnar austenite grain size of about 50 /mm
seems reasonable under the prevailing circumstances.
(j) As an illustration of principles, we shall assume that the CCT diagram in Fig. 6.27(a) pro-
vides an adequate description of the base plate transformation behaviour during welding. The
cooling time from 800 to 500 0C can be calculated from equation (1-67):
from which
Readings from Fig. 6.27(a) give the following microstructures within the grain coarsened
and grain refined region of the HAZ, respectively:
It follows that the observed difference in the HAZ transformation behaviour can mainly
be attributed to a corresponding difference in the prior austenite grain size, which according
to Fig. 5.21(a) is about 50 /im at Tp « 1350 0C and below 10 ^m at Tp « 10000C.
In addition, small islands of plate martensite will form within the intercritical (partly trans-
formed) HAZ, where the peak temperature of the thermal cycle has been between Ac1 and
Ac3 (see discussion in Section 6.3.8.2, Chapter 6). Just above the Ac1 temperature the volume
fraction of the M-A (martensite-austenite) constituent is approximately equal to the base
plate pearlite content (Fig. 6.66), which in the present case is about 8 vol%, as judged from the
steel carbon content.
Considering the weld metal, the situation is different. Here the oxide inclusions will
strongly affect the microstructure evolution by promoting intragranular nucleation of acicu-
lar ferrite (see discussion in Section 6.3.5, Chapter 6). In practice, the role of inclusions in weld
metal transformation kinetics is difficult to assess and hence, we will take a more simplistic
(pragmatic) approach to this problem by just comparing the total surface area available for
nucleation of ferrite at prior austenite grain boundaries and inclusions, respectively (SJGB)
versus SJI)). The following three regions of the weld are considered:
From the above calculations it is apparent that the conditions for acicular ferrite formation
are particularly favourable within the as-deposited weld metal (Sx(I) > SJGB)), and some-
what less favourable within the high peak temperature region of the weld HAZ (SJGB) >
SJI)). In contrast, acicular ferrite would not be expected to form within the low peak tem-
perature region of the HAZ, since nucleation of ferrite at austenite grain boundaries in this
case will completely override nucleation at inclusions (SJGB) » SJI)).This is also in agree-
ment with general experience (e.g. see photographs of typical microstructures in Fig. 6.19(c)
and (d)).
(k) The maximum hardness/strength level within the grain coarsened region of the HAZ can
be estimated from the diagrams presented in Section 7.2.2 (Chapter 7) if the steel composition
and welding parameters fall within the specified range. Alternatively, we can use Fig. 7.19,
which applies to low carbon microalloyed steels. Taking the cooling time from 800 to 500 0C,
Ar8/5, equal to 3.3 s, we obtain:
In general, a hardness rather than a strength criterion is used as a basis for evaluation of the
risk of hydrogen cracking and H2S stress corrosion cracking during service. In the former case
an upper limit of about 300 to 325 VPN is incorporated in many welding specifications to avoid
problems with hydrogen cracking, but this restriction can be relaxed if specific precautionary
actions are taken during the welding operation to reduce the supply of hydrogen as shown in
Section 7.2.3 (Chapter 7). Considering the H2S stress corrosion cracking resistance a maximum
hardness level of 248 VPN is strictly enforced in many welding specifications, as discussed pre-
viously in Section 7.2.4 (Chapter 7). Hence, significant tempering of the martensite would be
required if the weldment is going to be used in environments containing sour oil or gas.
(1) In general, the toughness requirements vary with the type of application, but for offshore
structures a minimum CVN toughness of 35J at — 400C is frequently specified. From the CVN
data in Tables 8.2 and 8.4 it apparent that both the base plate and the weld metal meet this re-
quirement. Moreover, auto-tempered low carbon martensite and polygonal ferrite, which
form within the grain coarsened and grain refined region of the HAZ, respectively are known
to have an adequate cleavage resistance.This means that the intercritical HAZ is the most
critical region of the joint when it comes to toughness due to the presence of high carbon plate
martensite within the ferrite matrix (see Figs. 6.61 through 6.65 and discussion in Section
7.2.2.3, Chapter 7). In practice, the problem may be solved by applying an appropriate post
weld heat treatment (PWHT).
(m) Since the properties of martensite depend on the carbon content, C-Mn steel weldments
will generally be more prone to hydrogen cracking, H2S stress corrosion cracking and brittle
fraction initiation in the HAZ than low carbon microalloyed steel weldments. This explains
why the base plate carbon content has been gradually lowered to values well below 0.1 wt%
in step with the progress in steel plate manufacturing technology.
Problem description:
Consider GTA welding of 2 mm thin sheets of AISI 316 austenitic stainless steel with chemi-
cal composition as listed in Table 8.5. The base plate has an average grain size of 18 /xm in the
fully annealed condition, which conforms to a tensile yield strength of about 300 MPa. The
sheets shall be butt welded in one pass, using a simple I-groove with 3 mm root gap. In this
case the addition of filler wire is adjusted so that the area of the weld reinforcement amounts
to 50% of the groove cross section. Details of welding parameters and operational conditions
are given in Table 8.6 and 8.7, respectively.
Table 8.5 Exercise problem II: Base plate chemical composition (in wt%).
Steel C Mn Cr Ni
AISI316 0.03 2.0 16 12
Table 8.7 Exercise problem II: Operational conditions and filler wire characteristics1.
Shielding gas: Argon
Wire composition:
Weld metal*
properties:
Analysis:
The problem here is to evaluate the response of the base material to welding under the con-
ditions described above. The analysis should be quantitative in nature and based on sound
physical principles. The following input data are recommended:
Specific questions:
(a) Select an appropriate heat flow model for the system under consideration.
(b) Estimate the minimum bead length which is required to achieve pseudo-steady state
down to a peak temperature of 1000 0C.
(c) Estimate the deposition rate (in gA^s" 1 ), the weld cross section areas D and B (in
mm2), and the dilution ratio B/(B + D) during welding.
(d) Estimate the weld metal chemical composition for the given combination of base
plate, filler wire and dilution ratio.
(e) Sketch the contour of the weld pool and the resulting columnar grain morphology in
the x-y plane after solidification. Estimate also the weld metal delta ferrite content.
(f) Evaluate the risk of solidification cracking during welding.
(g) Calculate the austenite grain size profile across the HAZ. Estimate also the size of the
columnar grains in the weld metal.
(h) Evaluate the risk of chromium carbide formation in the HAZ during welding.
(i) Estimate on the basis of the Hall-Petch relation the maximum load bearing capacity of
the joint during service.
Solution:
(a) The problem of interest is whether we must use the general (but complex) Rosenthal thin
plate solution (equation (1-81)) or can adopt the simplified solution for a fast moving high
power source (equation (1-100)). Fig 1.43 provides a basis for such an evaluation. The most
critical position will be the fusion line. If we neglect the latent heat of melting, the BJn^ ratio
at the melting point becomes:
Readings from Fig. 1.43 show that we are outside the validity range of the simplified 1-D
model close to the fusion line, but that this solution is a good approximation within the low
peak temperature region of the HAZ. Here equation (1-100) may be used in replacement of
equation (1-81).
(b) The duration of the transient heating period depends on the actual point of observation
(i.e. the distance from the heat source). If we would like to apply the pseudo-steady state sol-
ution down to a peak temperature of 1000 0C, the corresponding nJ8B ratio becomes:
From Fig. 1.31 we see that this ratio corresponds to a dimensionless radius vector a5m of
about 5. The duration of the transient heating period may now be read from Fig. 1.28. A crude
extrapolation gives:
from which
The total area of fused metal can be read from Fig. 1.31. At the melting point the n3/0p8
ratio is close to 2, which gives:
and
This gives:
Note that in these calculations we have assumed that A2 is equal to the sum of (B+D) in
order to achieve realistic numbers.
(d) The weld metal composition can be calulated from a simple 'rule of mixtures':
(f) Normally, a minimum delta ferrite content of about 5 to 10 vol% is specified to avoid
problems with solidification cracking in the weld metal (see discussion in Section 7.3.4,
Chapter 7). This requirement is clearly met under the prevailing circumstances.
(g) The HAZ austenite grain size in different positions from the fusion boundary can be read
from Fig. 5.30(b). In the present example the net heat input per mm2 of the weld is equal to:
This corresponds to a maximum austenite grain size of about 60/mi close to the fusion
boundary, which also is a reasonable estimate of the weld metal columnar grain size.
(h) The most critical position is the low peak temperature region of the weld HAZ where Tp
is between 800 and 1000 0C, as shown in Section 6.4.2 (Chapter 6). However, it is evident from
Fig. 6.69 that the risk of chromium carbide formation in this case is negligible because of the
low base plate carbon content. Hence, the corrosion resistance will not be significantly affec-
ted by the welding operation.
(i) The minimum HAZ strength level may conveniently be calculated from equation (7-21),
using input data from Example 7.5 (page 530):
This gives the following strength reduction factor for the joint:
8.4 Exercise Problem III: Welding of Al-Mg-Si Alloys
Problem description:
Table 8.10 Exercise problem III: Operational conditions and filler wire characteristics1.
Shielding gas: Argon
Gasflowrate: 20 Nl per min
Wire diameter: 1.6 mm
Wire feed rate: 5.5 m per min
Wire composition: Wire I : Al + 5 wt% Si
Wire II: Al +5 wt% Mg
Weld metal* Wire I:
properties: Rp02 : 55 MPa, Rn; 165 MPa, El.: 18%
Wire II:
Rp02 : >130 MPa, Rn;. >280 MPa, El.: >17%, CVN+20: >30 J
Data compiled from dedicated filler wire catalogues and welding manuals.
Values refer to all weld metal deposit.
Analysis:
The problem here is to evaluate the response of the base material to welding under the con-
ditions described above. The analysis should be quantitative in nature and based on sound
physical principles. The following input data are recommended:
Specific questions:
Atomic percent silicon
Temperature, 0C
(i) Estimate for each combination of filler wire and parent material an overall strength re-
duction factor which determines the load bearing capacity of the joint.
(j) Imagine now that the same extrusion instead is used in the fully annealed (O- temper)
condition with a Vickers hardness and tensile yield strength of 50 VPN and 100 MPa, re-
spectively. To what extent will the temper condition affect the microstructure and strength
evolution during welding?
Solution:
(a) The problem of interest is whether we must use the general (but complex) Rosenthal thin
plate solution (equation (1-81)) or can adopt the simplified solution for a fast moving high
power source equation (1-100)). Fig 1.43 provides a basis for such an evaluation. The most
critical position will be the fusion line. If we neglect the latent heat of melting, the 6 In3 ratio
at the melting point becomes:
(b) The duration of the transient heating period depends on the actual point of observation
(i.e. the distance from the heat source). If we would like to apply the pseudo-steady state sol-
ution down to a peak temperature of 200 0C, the corresponding n/86p ratio becomes:
From Fig. 1.31 we see that this ratio corresponds to a dimensionless radius vector <r5m of
about 5. The duration of the transient heating period may now be read from Fig. 1.28. A crude
extrapolation gives:
from which
It follows that the minimum bead length required to achieve pseudo-steady state during
aluminium welding is much longer than in steel welding due to the pertinent differences in the
heat flow conditions (e.g. see Example 1.5, Chapter 1).
(c) The value of the deposition coefficient may be estimated from the data in Table 8.10:
This value corresponds to a A: Vp ratio of about 0.92 mm3A 1S \ which is in excellent agree-
ment with the data quoted in Table 1.7. The area D of deposited metal thus becomes (see
equation (1-120)):
The total area of fused metal can be read from Fig. 1.31. At the melting point the nJ0p8
ratio is close to 0.93, which gives:
and
This gives:
Note that in these calculations we have assumed that A2 is equal to the sum of (B + D) in
order to achieve realistic numbers.
(d) The weld metal composition can be calulated from a simple 'rule of mixtures':
Wire I:
Wire II:
(e) The bead morphology can be read from Fig. 1.29. Taking the 68In3 ratio at the melting
point equal to 1, it is easy to verify that the shape of weld pool in this case is elliptical. The
columnar grain structure is therefore similar to that shown in Fig. 3.11 (a).
Moreover, Fig. 3.43 provides a basis for estimating the substructure of the weld metal
columnar grains. Close to weld centre-line the local crystal growth rate will approach the
welding speed (i.e. RL ~ 10 mm s"1). At the same time a simple analytical solution exists for
the thermal gradient in the weld pool (equation (3-29)):
From this we see that a cellular-dendritic type of substructure is likely to form within the
central parts of the fusion zone, in agreement with general experience.
If we only consider the contribution from the major alloying element in each case, the Scheil
equation (equation (3-46)) may be used for an analysis of the segregation pattern during sol-
idification. By using input data from the binary phase diagrams in Figs. 8.5 and 8.6, we get:
Wire I:
Wire II:
From this we see that the amount of eutectic liquid which forms during solidification is sen-
sitive to variations in the filler wire chemical composition (i.e. the Si or Mg content).
(f) Fig. 7-54 provides a basis for evaluation of the hot cracking susceptibility.
Wire I
In this case the fraction of eutectic liquid is so abundant that it backfills and 'heals' all incipi-
ent cracks. Hence, the hot cracking susceptibility is low.
Wire II
When the Al-Mg filler wire is used the fraction of eutectic liquid is just large enough to form
continuous films at the columnar grain boundaries. Hence, the hot cracking susceptibility is
high.
(g) Liquation cracking arises from melting of specific phases present within the base material
(e.g. Mg2Si and Si), as discussed in Section 7. 4.2.1 (Chapter 7). Fig. 7.61 provides a basis for
evaluating the HAZ cracking susceptibility:
Wire I
In this case the risk of liquation cracking is small because the solidus temperature of the weld
metal is lower than the actual melting temperatures of the base metal constituent phases.
Wire II
Due to the high Bi(B + D) ratio involved, the solidus temperature of the weld metal will ex-
ceed the actual melting temperatures of the base metal constituent phases. This may lead to
liquation cracking in parts of the HAZ where the peak temperature is greater than, say, 555
to 559 0C.
(h) The sequence of reactions occurring within the HAZ during welding of AA 6082-T6 alu-
minium alloys is shown in Fig. 7.62. In the present case we can use Fig. 4.24 for a quantitative
analysis of the /3" dissolution kinetics. Taking the net heat input per mm2 qjvd equal to 0.08
kJmm~2, we obtain:
First we need to estimate the corresponding if/m -coordinate in the HAZ from Fig. 1.31:
from which
First we need to estimate the corresponding if/m -coordinate in the HAZ from Fig. 1.31:
from which
A comparison with the phase diagram in Fig. 4.8 shows that the calculated temperature for
incipient dissolution of /3" is in good agreement with that obtained from the solubility prod-
uct.
(i) The yield strength in HAZ and the weld metal can be obtained from Fig. 7.67 and Table
8.10, respectively:
Wire I:
HAZ: Rp02 (min) « 130 MPa, Weld metal: Rp02 « 55 MPa , Base metal: Rp02 « 280 MPa
Wire II:
HAZ: Rp02 (min) - 130 MPa7WeId metal: Rp02 > 130 MPa , Base metal: Rp02 « 280 MPa
From this we see that the Al-Mg filler wire (wire II) yields the best weld metal mechanical
properties and should therefore be used, unless the cracking resistance is of particular con-
cern.
(j) When the material is present in the O-temper condition, it will contain an appreciable
amount of the equilibrium /3-Mg2Si phase. This will tend to accelerate the problem with li-
quation cracking within the HAZ during welding.
In addition, it is evident from Figs. 4.4 and 4.8 that the equilibrium /3-Mg2Si phase is ther-
modynamically much more stable than the metastable /3" phase. In practice, this means that
only a narrow solutionised zone will form adjacent to the fusion boundary. However, within
this zone significant strength recovery may occur after welding due to natural ageing effects
(see Fig. 4.5), which may result in a HAZ hardness and tensile yield strength level of about 80
VPN and 190 MPa, respectively. Hence, for the O-tempered material, we get:
Wire I
HAZ: Rp02 (max) - 190 MPa, Weld metal: Rp02^ 55 MPa, Base metal: Rp02 « 100 MPa
/=55/100 = 0.55
Wire II
HAZ: Rp()2 (max) - 190 MPa, Weld metal: Rp02 > 130 MPa, Base metal: Rp02 - 100 MPa
/ = 100/100 = 1
Author Index
A
absorption of elements see hydrogen, nitrogen, oxygen
acicular ferrite in low-alloy steels 428
crystallography of 428
nature of 430
nucleation and growth of 432
texture components of 429
acicular ferrite in wrought steels 444
aluminium as alloying element in steel
effect on inclusion composition 202 206
effect on solidification microstructure 246 272 293
effect on weld properties 481 486
solubility product of precipitates 303
aluminium weldments 458 536
age-hardenable alloys 458
quench sensitivity 459
precipitation conditions during cooling 459
strength recovery during natural ageing 461
subgrain evolution in friction welding 464
characteristics 536
constitutional liquation
in Al-Mg-Si alloys 542
in Al-Si alloys 541
example (7.9) – minimum HAZ strength level 554
example (7.8) – weld metal hot cracking 544
example (7.7) – weld metal solidification cracking 537
example (7.10) – minimum HAZ hardness level 562
HAZ microstructure evolution 547
This page has been reformatted by Knovel to provide easier navigation. 595
596
Index terms Links
aluminium weldments (Continued)
constitutive equations 548 558
during friction welding 555
during fusion welding 547
hot cracking 540
factors affecting 544
solidification cracking 536
strength evolution during welding 547
constitutive equations 548 558
during friction welding 555
hardness and strength distribution 550 560
strengthening mechanisms in alloys 547
amplitude of weaving – definition 80
arc atmosphere composition 132
see also shielding gases
arc efficiency factors 26
definition 26
selected values 27
arc welding 24
definition of processes 24
austenite
grain size in low-alloy steels 409
primary precipitation in fusion welds 292
austenite formation in low-alloy steels 449
conditions for 450
austenitic stainless steels 453
see also stainless steel weldments
characteristics of 527
chromium carbide formation 456
grain growth diagrams for steel welding 375
weld decay area 456
B
Bain orientation region 436
bainite in low-alloy steels 444
lower 447
upper 444
bead morphology 96
bead penetration 99
deposit and fused parent metal 96
example (1.16) – SA welding of steel 97
example (1.17) – SMAW welding of steel 98
example (1.18) – Jackson equation 99
Bessel functions – modified 46 47 49
boron in steel
effect on transformation behaviour 413
segregation of 294
weld properties 493 505
bowing of crystal 240
Bramfitt’s planar lattice disregistry model 244
see also solidification of welds
C
carbon equivalents 496 521
carbon as alloying element in steel
austenitic stainless steels 453
weld deposits 424
carbon-manganese steel weld metals, grain growth in 370
casting, structural zones 221
D
Delong diagram 535
delta ferrite, primary precipitation of 290 292
dendrite arm spacing 261
primary 261
secondary 264
dendrite fragmentation 250
see also solidification of welds
E
energy barrier to solidification 225
see also solidification of welds
enthalpy of reaction 302
definition of 302
values 303
entropy of reaction 302
definition 302
values 303
epitaxial solidification 222
equiaxed dentritic growth 268
equilibrium dissolution temperature of precipitates 303
see also solidification of welds
error functions see Gaussian error functions
F
fluid flow pattern in weld pools 186 228
flux basicity index 171
friction welding 18
see also aluminium weldments
dimensionless temperature 20
dimensionless time 20
dimensionless x-coordinate 21
example (1.4) – peak temperature distribution 23
heat flow model 18
temperature-time pattern 23
Fritz equation 281
G
gas absorption, kinetics of 120
rate of element absorption 121
thin film model 120
gas desorption, kinetics of 123
rate of element desorption 123
Sievert’s law 124
gas porosity in fusion welds 279
growth and detachment of gas bubbles 281
nucleation of gas bubbles 279
separation of gas bubbles 283
Gaussian error functions, definition 112
Gaussian heat distribution 112
see also distributed heat sources
Gibbs-Thomson law 309
Gladman equation 344
grain boundary ferrite 408
crystallography of 408
growth of 422
nucleation of 408
grain detachment 250
see also solidification of welds
grain growth 337
computer simulation 380
diagrams
construction of 360
I
implant testing 520
see also hydrogen cracking
inclusions in welds – origin 192
constituent elements and phases in inclusions 202
example (2.10) – computation of inclusion volume fraction 194
example (2.12) – computation of total number of constituent phases in
inclusions 211
prediction of inclusion composition 204
size distribution of inclusions 195
coarsening mechanism 196
effect of heat input 196
example (2.11) – computation of number density and size
distribution of inclusions 201
volume fraction 193
stoichiometric conversion factors 194
instantaneous heat sources 5
line source 5
plane source 5
point source 5
interface stability 254
K
Kurdjumow-Sachs orientation relationship 408 427 429
444 448
L
latent heat of melting 3
lattice disregistry
see Bramfitt’s planar lattice disregistry model
local fusion in arc strikes 7
dimensionless operating parameter 7
dimensionless radius vector 7
dimensionless temperature 7
dimensionless time 7
example (1.1) – weld crater formation and cooling conditions 9
heat flow model 7
temperature-time pattern 8
low-alloy steel weldments 477
acicular ferrite in 428
crystallography of 428
nature of 430
nucleation and growth in 432
texture components of 429
austenite formation in 449
conditions for 450
bainite in 444
lower 447
M
magnesium in aluminium alloys
solubility product of precipitates 303
martensite
in low-alloy steels 447
austenite formation, kinetics of 449
lath 447
M-A formation, conditions for 450
plate (twinned) 447
martensitic stainless steels, characteristics of 527
mass transfer in weld pool, overall kinetic model of 124
medium thick plate solution 59
see also heat flow models
dimensionless maps for heat flow analysis 61
case study (1.1) – temperature distribution in steel and aluminium
weldments 69
construction of maps 61
N
non-isothermal transformations
additivity principle 403
and Avrami equation 404
O
operating parameter, dimensionless
point and line heat source models 31
weaving model 82
Ostwald ripening see particle coarsening
oxygen, absorption of 148
classification of shielding gases 166
overall oxygen balance 166
content in welds 148
covered electrodes 173
R
reversion see particle dissolution
example (1.15) – cooling conditions during root pass welding 95
heat flow model 95
Reynold number
definition 187
of gas bubbles 284
of particles 187
root pass welding, thermal conditions in 95
Rosenthal equations
see thick and thin plate solutions
S
Scheil equation 272
modified 276
original 272
separation of gas bubbles in fusion welds 283
shielding gases see oxygen, hydrogen and nitrogen, absorption of
CO-evolution 166
T
texture in welds
solidification 221 290
solid state 429
V
volume of weld metal 36
volume fraction of inclusions 193
volume heat capacity 3
W
Wagner-Lifshitz equation 196 314 351
water content 137
in electrode coating 137
in welding flux 138
see also hydrogen absorption
Z
Zener drag, definition of 341
in grain growth 341
Zener equation 342 344
Zener-Hollomon parameter 465
zinc in aluminium
solubility product of precipitates 303