You are on page 1of 12

SUPPLEMENT TO THE WELDING JOURNAL, MARCH 1976

Sponsored by the American Welding Society and the Welding Research Council

("Rul

Deoxidation Practice for


Mild Steel Weld Metal
Quantitative metallography explains how weld composition controls ductile and brittle fracture resistance

BY D. J. WIDGERY

ABSTRACT. The effects of M n , Si, Ti


and Al on inclusion contents, microstructure, strength and toughness of
as-deposited, single run C 0 2 welds
made at 1.6 k J / m m (40 kJ/in) have
been studied. A statistically planned
series of experimental electrodes was
used and toughness was assessed
with the crack opening displacement
(COD) test.
Resistance to cleavage and to
fibrous fracture were differently affected by microstructure and c o m position. Good cleavage resistance
was achieved with acicular microstructures, which manganese c o n tents up to at least 1.6% in the deposit helped to promote, and with low
titanium levels to avoid dispersion
h a r d e n i n g . G o o d c l e a v a g e resistance was also achieved in a soft,
homogeneous structure with a yield
strength of 390 N/mm 2 (56 ksi) but
this might be difficult to guarantee in
commercial practice.
Resistance to ductile fracture by
microvoid coalescence was related to
inclusion content, yield strength and
strain hardening behavior of the weld.
Low inclusion contents were favored
by weld M n / S i ratios of 4-5 but it may
be more profitable to aim for soft
microstructures if good tearing resistance under highly triaxial stress is
required.

D. J. WIDGERY is associated with the


Welding Institute, Research Laboratory,
Abington Hall, Abington, Cambridge CBI
6AL, England.
Paper was presented at the AWS 56th
Annual Meeting held in Cleveland, Ohio,
April 21-25, 1975

Introduction
The important influence which weld
deoxidation practice can have on the
toughness of the deposit has long
been recognized, but only in recent
years have attempts been made to
quantify the v a r i o u s m e c h a n i s m s
through which this influence is exerted (Refs. 1-4). A recent program at
the Welding Institute set out to investigate how deoxidation practice affected inclusion contents and microstructure and thus how resistance to
fracture in the ductile and brittle
modes was controlled. The results of
this investigation are reported here.
Some oxidation almost invariably
accompanies the welding of steel. In
the slag shielded processes, the activity of oxygen in the slag is not negligible and perfect shielding from atmospheric oxidation difficult to achieve.
The G M A W process generally relies
on the use of a mildly oxidizing gas to
allow some oxidation of the workpiece and stabilize the arc. Since oxidation cannot be avoided, deoxidation procedures must be devised to
minimize its adverse effects on the integrity and performance of the weld
and the optimization of these procedures was the object of this program.
The solid wire COz (GMAW) process
was chosen for this because it permits a moderate degree of oxidation,
is readily mechanized and is u n c o m plicated by the presence of a flux.
Further incentives to the use of this
process were its increasing c o m m e r cial acceptance and the fact that
many reactions in the weld pool are
c o m m o n to welds made with basic
manual metal-arc electrodes or with

some submerged arc fluxes in which


carbonates are incorporated.
Aims of Deoxidation
In C 0 2 welding, droplets transferring between the wire and the pool
pick up oxygen as
C 0 2 ; C O + [0]

(D

The dissolved oxygen can react with


carbon in the steel to form CO,
:CO

[ C ] . + [0]
Fe

(2)

Fe

which gives rise to porosity if its equilibrium pressure p c o exceeds 1 a t m o sphere. To prevent this, it is necessary to limit the product [C] [O] in the
mass action equation
= constant.

(3)

[C] . [O]
In practice, except in welding pure
iron, the carbon content of the pool
will always be such that the oxygen
activity must be reduced by adding
metallic deoxidants if porous welds
are to be avoided (Ref. 5).
If the first aim of deoxidation is to
prevent porosity, the second is to
minimize the amount of oxygen remaining in the deposit after solidification. Since there is little evidence
to contradict Sims' assertion (Ref. 6)
that "for all practical purposes, the
solubility of oxygen [in solid steel] is
so low that it may be neglected," this
objective may be restated as the minimization of the amount of oxide inclusions in the weld.
In steelmaking practice, control of
inclusion contents is generally achiev-

WELDING RESEARCH SUPPLEMENT!

57-s

particular concern in the present investigation.

Table 1 Chemical Analysis (wt. %) of Experimental Electrodes


Electrode

0.04
0.13
0.055
0.055
0.075

0.011
0.009
0.012
0.017
0.012

0.008
0.004
0.006
0.006
0.006

0.25
0.44
0.40
0.19
0.14

0.20
0.12
0.10
0.15
0.12

0.012
0.010
0.012
0.009
0.010

0.006
0.003
0.004
0.017
0.018

0.43
1.05
0.43
0.43
1.11

0.09
0.08
0.23
0.09
0.23

0.08
0.07
0.06
0.21
0.08

0.008
0.010
0.009
0.009
0.009

0.018
0.015
0.023
0.020
0.021

0.98
1.98
1.62
2.27
0.63

1.07
0.45
0.91
0.78
0.74

0.24
0.20
0.08
0.14
0.15

0.19
0.14
0.15
0.13
0.14

0.008
0.008
0.010
0.008
0.011

0.017
0.022
0.020
0.018
0.016

1.42
1.47
1.46
1.51
1.41
1.38

1.51
<0.01
0.76
0.79
0.78
0.70

0.14
0.17
0.32
0.01
0.15
0.16

0.14
0.16
0.14
0.13
0.24
0.01

0.010
0.007
0.010
0013
0.007
0.011

0.014
0.015
0.018
0.013
0.011
0.018

Mn

Si

Ti

A
B
C
D
E

0.05
0.09
0.10
0.09
0.09

1.14
1.11
1.63
2.15
1.10

0.51
0.92
0.50
1.05
1.40

0.19
0.22
0.36
0.14
0.46

F
G
H
J1
J2

0.07
0.11
0.11
0.08
0.08

1.42
1.10
1.53
1.51
1.42

0.51
0.46
0.90
0.78
0.77

K
L
M
N
0

0.08
0.06
0.08
0.09
0.08

1.95
0.98
1.01
1.02
1.92

P
Q
R
S
T

0.08
0.09
0.08
0.09
0.07

U
V

0.07
0.08
0.10
0.07
0.09
0.07

Al

Experimental Program
The work was planned to study the
effects of four deoxidants, Mn, Si, Ti
and Al. The program centered on the
measurement of fracture toughness,
for which theoretical and practical
considerations led to the adoption of
the crack o p e n i n g d i s p l a c e m e n t
(COD) test (Ref. 8). Welding procedures were designed to provide COD
testpieces with the notch in a single
weld run. To provide the link between the composition and mechanical properties of the deposit, quantitative measurements of the microstructure and inclusion contents were
undertaken. The experimental plan
was based on the use of multiple
regression analysis to interpret the
results.
Experimental Electrodes

X
Y

\ ^ \/-3mm

mol md.

1 5mm mot md.

Fig. 1 Preparation
of
experimental
welds: (a) plate grooved, (b) buttered and
regrooved,
(c) experimental
weld completed

ed by ensuring a low oxygen solubility in the molten steel and relying on


the removal of most of the deoxidation products before solidification.
Much work has been devoted to evaluating oxygen solubility in the liquid
as a function of deoxidant additions
and the oxygen content of the solid
steel can closely approach the level
thus predicted.
In welds, on the other hand, the
final oxygen content is usually much
58-S I M A R C H

1976

greater than the solubility in the melt:


for example, the presence of 0.5% Mn
with 0.5% Si reduces the oxygen solubility in m o l t e n i r o n at 1600 C
(2912 F) to - 0 . 0 0 7 % (Ref. 7), whereas oxygen contents an order of magnitude higher may be encountered in
welds with comparable or greater
deoxidant contents (Ref. 2). Clearly,
the short time available for the removal of deoxidation products from
the pool means that the kinetics of
coalescence and flotation have more
influence on the final weld oxygen
content than the exact level of oxygen
solubility on the melt. This explains
the limited success of t h e r m o d y n a m ic arguments in predicting weld oxygen levels and shows the need to
concentrate further studies on those
properties of the deoxidation p r o d ucts themselves which affect their rate
of removal from the weld pool.
Since the elements used to deoxidize mild steel weld metals also c o n stitute the principal, often the only
alloying additions, a third aim of
deoxidation practice must be to ensure that the deoxidants which remain
in solution after all the oxygen is
chemically combined will promote the
formation of desirable microstructures and will not induce embrittlement through solid solution or dispersion hardening. It is through such
effects on the metal matrix, rather
than in their strictly d e o x i d i z i n g
capacity, that the reactive elements in
welding consumables exert most influence on the cleavage resistance of
the deposit, so these effects were of

A series of 26 vacuum melted


experimental electrodes was produced with varying amounts of M n , Si,
Ti and Al. The experimental design
had to take account of the likelihood
that the effects of the deoxidants
would be nonlinear, as shown by Sakaki (Ref. 9): in general, small additions are taken up in the reaction with
oxygen while successive larger additions have p r e d o m i n a n t l y m i c r o structural effects. A central, rotatable, composite design was adopted
since these were developed as an
economical means of investigation
nonlinear behavior (Ref. 10) and have
previously proved valuable in research on weld metals (Ref. 11). The
electrode compositions produced are
shown in Table 1. Electrode J was
duplicated by two manufacturers and
5 welds were made with the two filler
metals as a check on reproducibility.
The batch of electrodes constituting
the first half-factorial did not conform
well to the plan, but the experience
gained in their production allowed the
remainder to be made more accurately and the series as a whole was
satisfactory in respect of range and
orthogonality. Electrodes were drawn
to 1.2 m m (3/64 in.) diam without
c o p p e r c o a t i n g . Two c o m m e r c i a l
electrodes, one containing 0.48% Al,
were also included for comparison.
Welding Procedure
One of the advantages of the COD
testing technique is that since fracture initiation is controlled only by the
material close to the notch tip, it is
possible to sample predetermined regions of a single weld run rather than,
as in the Charpy test, a composite of
all the structures through which the
fracture passes as it crosses the
specimen. Previous work had strongly indicated that cleavage cracks ini-

tiated more readily in as-deposited


regions than in reheated regions of
multipass welds (Ref. 12), so a procedure was devised to produce COD
specimens with the notch tip along
the centerline of a single, as-deposited bead. A bead-in-groove specimen was found suitable for this purpose, but in order to achieve a wide
range of deposit compositions without changing the base material it was
necessary to reduce the effective dilution by means of a buttering technique. A groove somewhat larger than
that needed for the test bead was filled with several runs of the experimental weld metal and the groove for
the test bead was machined into this
deposit, as shown in Fig. 1. By this
means the dilution of the test bead
with the mild steel base metal was reduced below 10%.
The test bead was deposited at 30
V, 300 A, 330 m m / m i n (13 in./min),
giving spray transfer with an arc
energy of 1.6 k J / m m (40 kJ/in.), and
the C 0 2 flow rate was 0.25 l/s (30 cu.
ft/h). A preheat temperature of 100 C
(212 F) was used throughout. To help
in interpreting weld microstructures,
thermal analysis was carried out on
selected welds by inserting a thermocouple into the molten pool and
using a derivative processor to plot T
and d T / d t as the weld cooled.
Chemical Analysis

mm (0.005 in.) slitting wheel was extended in fatigue to a total depth of


about 6 mm (0.24 in.), so that an approximately square ligament 22 X 22
mm (0.87 x 0.87 in.) remained under
the notch. Testing was carried out in
3-point bending over a span of 127
mm (5 in.) and COD was monitored
using a double cantilever beam clip
gage mounted on knife edges at the
specimen surface. Readings were
converted to crack tip COD, <5, using
the relationship
V

& =-

cr = ke

(5)

was determined as the slope of the


log a vs log E plot over the range of
strain between e = 0.025 and the
maximum uniform strain.

(4)

1 + a+z

Metallography

r(W-a)
where V g is the knife edge displacement, a the total crack depth, z the
knife edge thickness, W the specimen width and r the rotational factor,
taken here as 1/3 in accordance with
the British S t a n d a r d s Institution's
Draft for Development on COD testing
(Ref. 14).
Tensile Testing

W^b~' r s , ! , , <*-^~> -fT\

The complexity of weld metal


microstructures confronts the investigator with a choice of classifying
them according to a simple system,
permitting measurements of good
reproducibility, or of devising a more
complex system which may describe
the transformation products more
accurately in a qualitative sense at the
Fdtigue c / w * 15mm
'5mm

Longitudinal all-weld-metal tensile


testpieces were extracted from each
weld. These had a gage diameter of
6.41 m m (0.253 in.) and a parallel and
gage length of 22.7 mm (0.894 in.).
Testing was carried out at room t e m perature. In addition to measure-

All welds were analyzed spectrographically using millings from the


final weld run, remelted under argon.
In addition, total nitrogen contents
were measured using the conventional indophenol blue finish method
and oxygen contents initially by the
vacuum fusion method. When difficulties arose in reconciling these oxygen measurements with the results
of inclusion volume fraction determinations, further oxygen analyses
were carried out using fast neutron
activation analysis (Ref. 13).

ments of yield or 0.2% proof stress,


tensile strength, elongation and
reduction of area, a simple computer
program was used to evaluate true
stress, a T , and true strain, e. The
strain h a r d e n i n g e x p o n e n t , n, as
defined by the Holloman equation
(Ref. 15)

V1

*p
&ep

3/ryn

X 3rmil\ S ^ r i y ! ,. 5mm.mot

md.orPmm

Fig. 2 Geometry ot COD specimen

Spsf1

COD Testing
In addition to the practical advantage that the COD test gives in sampling well defined microstructural regions, its theoretical basis, which
allows it to be used for the prediction
of critical or allowable defect sizes in
real fabrications, has led to its increased use in specifications for critical applications such as offshore
structures. A third benefit in using the
test is that since it measures resistance only to fracture initiation and
not to propagation, which may involve different mechanisms, metallurgical interpretation of the results
tends to be simpler than when the
Charpy test is used.
In the present work, specimens
were produced as shown in Fig. 2 with
the notch parallel to the welding
direction. A slot produced with a 0.13

.i

JU

JM :

iv:v-'X;^,-:~^:---'#'"' "Jf:W^

IP
-

'

'

>

.4-//.^:

Fig. 3 Microstructure classification used in point counting, x200, reduced 18%

W E L D I N G R E S E A R C H S U P P L E M E N T ! 59-s

Carbon replica

Fig. 4 Mechanism ot inclusion extraction tor counting in the electron microscope,


ing how inclusions of diameter x are extracted from a volume of depth x

expense of greater difficulty in achieving adequate reproducibility in quantitative work. For example, the former
approach might rely on light microscopy while the latter would use the
electron microscope. The decision
was made h e r e t o begin by using light
microscopy so that large numbers of
measurements could be made quickly and with minimal sampling errors.
The measurements were backed by
electron microscope studies on both
replicas and thin foils, but it was c o n cluded that quantitative data need not
be sought from these if it was possib l e to c o r r e l a t e t h e m e c h a n i c a l
p r o p e r t i e s of the w e l d s with the
results of light metallography, as was
f o u n d to be the case.
A Swift point counter* was used for
quantitative m i c r o s t r u c t u r a l m e a surements. This moves the specimen
under the microscope in a predetermined grid pattern, the investigator
identifying the structure and operating the appropriate counter at each
grid point. Transverse weld sections
were examined at a magnification of
x200 and 100 microstructural identifications were made at 0.05 m m
(0.002 in.) intervals along each of 10
lines, parallel to the plate surface and
0.2 mm (0.008 in.) apart, starting 1
m m (0.04 in.) below the plate surface.
Thus 1,000 counts were made in a 5
X 2 m m (0.2 X 0.08 in.) area in the
center of the weld, corresponding to
t h e n o t c h p o s i t i o n in t h e C O D
specimens.
T h e s y s t e m of c l a s s i f i c a t i o n
adopted is shown in Fig. 3, and represents an attempt to reconcile the requirements of easy recognition with
an accurate reflection of the transf o r m a t i o n process. To avoid the
semantic confusion apparent in the
literature on weld microstructures, the
products were initially identified only
by numbers: later, it was possible to
ascribe to them places in the overall
transformation scheme.
Inclusion Counting
Gloor et al showed in 1963 (Ref. 16)
that the majority of weld metal inclusions are too small, typically < 1 n m ,
"James Swift and Son Limited, Basingstoke, England.
60-s | M A R C H

1976

show-

to be resolved in the light microscope and most subsequent investigators have relied on transmission
or scanning electron m i c r o s c o p y
(TEM or SEM) for inclusion measurements. However, while either method
can provide a straightforward means
of measuring the size distribution of
inclusions on the surface of a fracture produced by microvoid coalescence, only indirect methods are
available to find the size and spatial
distribution of inclusions in the matrix.
These methods are based on two
models of specimen preparation, the
flat surface model and the deep etch
model.
In materials polished flat for o p tical metallography the area fraction
A A of any second phase on the surface may be equated to the volume
fraction V v of the phase in the matrix.
Against this ease of volume fraction
measurement must be set the need to
allow for sectioning effects in c o n verting the observed distribution of
inclusion section sizes to a volume
size distribution. Procedures for this,
based on a method proposed by
Scheil (Ref. 17), assume, reasonably
in the case of weld metals, that the inclusions are spherical. Although these
procedures are somewhat cumbersome and prone to cumulative errors,
a more important objection to the use
of the flat surface model is the difficulty of finding an imaging mechanism w h i c h the m o d e l accurately
represents. At first sight, it might
appear that the conditions of the
model would be fulfilled by the use of
polished specimens in the SEM and
trials were undertaken with this instrument.
As a "bench m a r k " weld, a manual
metal-arc deposit made with rutile
covered electrodes and containing
0.079% O (vacuum fusion determination) and 0.022% S was selected.
These figures could be converted to
an estimated inclusion volume fraction with reasonable confidence by
assuming that the deoxidation p r o d ucts were the same as those a n alyzed by Gloor et al (Ref. 16) and that
the sulfur was present as MnS. This
led to an estimated inclusion volume
fraction of 0.53% (Ref. 18). Since the
Al 2 0 3 and T i 0 2 contents of the inclusions were expected to be low, any

errors in oxygen measurement arising


from nonreduction of these oxides
should also have been low.
When a mechanically polished surface of this weld was photographed in
the SEM, the area fraction of inclusions was found to be 1.6 % or three
times the calculated volume fraction.
A similar result can be derived from
data presented by Ruckdeschel (Ref.
19) for SEM inclusion m e a s u r e ments. Investigation showed that at
the accelerating potential of 30 kV,
the penetration of the electron beam
was sufficient to image some inclusions which lay close beneath the polished surface and to increase the apparent sectioning diameter of others
which just broke the surface. With the
instrument then available at the W e l d ing Institute, attempts to reduce beam
penetration by lowering the accelerating potential were frustrated by a
simultaneous decline in resolution.
Newer instruments may overcome
this difficulty, as shown by an e x a m ination of results recently produced
by Hannerz and Lowery (Ref. 20), but
in the meantime an alternative a p proach has been developed.
A technique which has been used
for particle size measurement in the
TEM relies on etching or electropolishing the metal matrix so deeply that
all sectioned inclusions are removed
and those that remain are always
seen at their true diameter. These
may be extracted on replicas, Fig. 4. If
it is assumed that all particles which
touch the plane of polish are extracted, then particles of any given
diameter x are s a m p l e d f r o m a
volume of depth x and the number of
particles of this diameter per unit area
of replica, N s (x), is related to the
number per unit volume of the matrix,
N v (x), by
N(x) =

(1/xNs(x)

(6)

This is the basis of the analysis proposed by Ashby and Ebeling (Ref. 21)
and adopted in recent work at the
Welding Institute (Refs. 3, 22). The
use of an electropolishing technique
found to give an acceptably flat surface, together with shadowed replicas to show nonextracted inclusions,
allows v o l u m e f r a c t i o n m e a s u r e ments to be made without recourse to
the d o u b l e r e p l i c a t e c h n i q u e rep o r t e d elsewhere (Ref. 23). The
method readily yields all the important parameters describing the size
and spatial distributions of the particles using only simple computations, and the calculated inclusion
volume fractions are in acceptable
agreement with those estimated from
the weld chemistry. Although the
method used by Gloor et al (Ref. 16)
also produced credible results, this
was of much greater analytical c o m plexity, being based on a combination of the deep etch and flat surface
models.

In detail, the technique used in the


present work was as follows: after
rough polishing' on silicon carbide
papers to 600 grade, electropolishing was carried out in a 7% perchloric acid solution containing 2 I methanol, 60 ml ether, 25 ml distilled water
and 150 ml perchloric acid. The polishing potential was about 43 V, varying by a few volts for o p t i m u m results
on some specimens, the current d e n sity was about 10 m A / m m 2 and the
polishing time 30 s. The specimen
was then etched for 2 s in 2 % nital.
Carbon was deposited on the surface
at an angle of 45 deg and scribed into
2.5 m m (0.1 in.) squares. Unwanted
areas were lacquered off and the
replicas were removed electrolytically in 10% nital.
A typical micrograph of a replica
prepared in this way is shown in Fig.
5. Following Stumpf and Sellars (Ref.
23), unshadowed inclusions, X, were
considered to have floated into the
replica from some other part of the
s p e c i m e n and were not c o u n t e d .
Shadowing also revealed the sites of
lost inclusions and allowed extraction efficiencies to be calculated:
these generally exceeded 95%. The
magnification of the electron microscope was calibrated for each set of
micrographs, which were enlarged to

COD Tests

an exact magnification of x5,000 on


the print. Inclusion sizes were measured with a specially made vernier
device and recorded on punched
tape. A simple computer program
was then used to evaluate the frequency of the observed size distribution, the frequency and percentage
frequency of the volume size distribution and the inclusion volume
fraction, calculated after Ashby and
Ebeling (Ref. 21) as

Vv

The behavior of a material in the


COD test is most usefully characterized by the critical value, 8C, of COD at
which the specimen breaks. It is poss i b l e , h o w e v e r , at t e m p e r a t u r e s
above the ductile/brittle transition, for
a stable ductile tear to propagate so
far through the specimen that the load
passes through a m a x i m u m while the
specimen remains unbroken. In such
cases, it is conventional to measure
the COD at a point, 6m, where the load
first attains its m a x i m u m value and to
use this in plotting the transition
curve, Fig. 6.
However, the use of Sm values in
calculating critical defect sizes in real
structures has no theoretical basis
and moreover, in the present welds a
stable crack had often propagated
out of the weld metal and into the
base metal before maximum load was
attained. The results in such a case
simply indicate that at the temperature concerned, cracks were u n likely to become unstable at low COD
values, such as might be developed at
very small defects under service
loading.
The main concern of fabricators is
to avoid materials which are sensitive
to small defects, which for mild steel

( T T / 6 ) S " X 3 N V (x) r)x =

>/6)2"x2Ns(x)5x

(7)

E x p e r i m e n t a l Results
Thirty-one welds were made and
their chemical analyses are given in
Table 2. Sulfur and phosphorus c o n tents were less than 0.02% in all the
welds made with experimental electrodes. Carbon contents averaged
0.09% and showed a very low degree
of correlation with the levels of metallic deoxidants, though in the single
case of weld V a low carbon content of
0.03% was associated with a low total
deoxidant content: this weld was too
porous to be tensile tested or COD
tested.

Table 2 Chemical Analysis (wt %) of Weld Deposits Made with Experimental and Two
Commercial Electrodes
Mn

Weld

Ti

Al

-.(a)

(b)

A
B
C
D
E

0.10
0.11
0.11
0.09
0.10

0.77
1.33
1.10
1.56
0.82

0.27
0.54
0.26
0.58
0.61

0.04
0.04
0.07
0.04
0.12

0.008
0.013
0.013
0.018
0.025

0.017
0.017
0.012
0.013
0.014

0.018
0.018
0.012
0.015
0.012

0.010
0.010
0.010
0.011
0.007

0.055
0.055
0.048
0.055
0.047

F
G
H
J1
J2

0.08
0.09
0.11
0.09
0.07

0.96
0.74
1.11
1.13
1.05

0.23
0.28
0.49
0.60
0.57

0.06
0.10
0.11
0.07
0.05

0.046
0.023
0.024
0.06
0.04

0.013
0.011
0.013
0.010
0.012

0.013
0.014
0.010
0.016
0.016

0.012
0.009
0.008
0.004
0.012

0.067
0.058
0.057
0.060
0.056

0.088

J1R
J2R
J2RR
K
L

0.08
0.09
0.08
0.07
0.05

1.10
1.06
1.00
1.13
0.72

0.50
0.48
0.53
0.17
0.86

0.06
0.05
0.05
0.02
0.03

0.06
0.05
0.04
0.02
0.03

0.007
0.011
0.013
0.008
0.007

0.018
0.015
0.017
0.012
0.007

0.009
0.007
0.007
0.010
0.003

0.054
0.063
0.063
0.064
0.063

M
N
O
P
Q

0.07
0.08
0.08
0.09
0.10

0.53
0.73
1.43
0.77
1.46

0.16
0.26
0.92
0.92
0.31

0.04
0.03
0.08
0.09
0.08

0.014
0.06
0.03
0.065
0.07

0.009
0.007
0.009
0.011
0.008

0.015
0.011
0.013
0.014
0.013

0.004
0.006
0.010
0.008
0.006

0.048
0.053
0.045
0.048
0.042

R
S
T
U
V

0.08
0.09
0.07
0.08
0.03

1.27
1.60
0.45
1.06
0.68

0.76
0.57
0.52
1.20
<0.01

0.03
0.04
0.05
0.05
<0.01

0.055
0.045
0.045
0.045
<0.005

0.008
0.008
0.011
0.012
0.008

0.013
0.017
0.014
0.013
0.014

0.005
0.014
0.009
0.009
0.004

0.047
0.040
0.052
0.043
0.080

0.09
0.08
0.08
0.08
0.09
0.10

1.06
1.12
1.15
0.93
1.09
0.84

0.58
0.58
0.66
0.45
0.58
0.28

0.11
<0.01
0.06
0.03
<0.01
<0.01

0.05
0.04
0.12
<0.005
<0.005
0.19

0.010
0.010
0.010
0.010
0.011
0.013

0.014
0.020
0.015
0.014
0.021
0.015

0.007
0.009
0.005
0.016
0.006
0.011

0.049
0.041
0.046
0.029
0.034
0.036

X
Y

Com. 1
Com. 2

0.094
0.097

0.12

0.10
0.095
0.10

0.10
0.12
0.10

(a) Vacuum fusion


(b) activation analysis

WELDING

RESEARCH SUPPLEMENT!

61-S

Table 3 0.1 mm COD Transition Temperatures of Experimental and Commercial


Weld Metals <">
To, C
T, C
Weld
To,C
Weld
Weld

)
I

A
B
C
D
E
F
G
H
J1
J2

&

-^-

COO a fracture
o COO at maximum toad

f*

0-6

05

03

0-2

,.^
or

I
A

" " " ^

-100

-SO

-60

-AO

remperattire.

-20

.1

'C

Fig. 6 COD test results for weld J1

weld metals generally means that interest is concentrated at the lower end
of the transition range. To compare
this aspect of the behavior of the experimental welds, a transition temperature Tc corresponding to 5c = 0.1
mm (0.004 in.) is tabulated for all the
welds in Table 3. This may be regarded as a measure of their cleavage resistance.
The results of a regression analysis of Tc on weld composition are
shown in Table 4. These results may
be presented graphically by omitting
each term successively f r o m the
regression equation and using the decompleted equation to predict T c . The
error in prediction, ATC, is then related to the level of the variable
omitted. Thus Fig. 7 effectively shows
the variation in Tc with Mn content,
other elements being corrected to a
constant level.
An unsatisfactory feature of Fig. 7 is
the presence of a single point, indicating the greatest overestimate of T c ,
but which by appearing at the low Mn
end of the scale seems to reverse the
general trend to lower transition temperatures with increasing Mn. This
62-s | M A R C H

J1R
J2R
J2RR
K
L
Wl
N
0
P
Q

63
47
46
94
30
-141
67
75
25
54

R
S
T
U
W
X
Y
Z
Com. 1
Com. 2

88
-113
11
31
18
79
90
77
65
55

(a) Weld V was too porous to test

Fig. 5 Transmission electron micrograph of inclusions on a replica, as used in


inclusion measurements. Unshadowed inclusions, x, are not counted. x5,000,
reduced 46%

0-7

-39
-51
-28
-94
- 1
-38
+ 16
-19
-76
-49

1976

point represents weld M. If similar


plots are made to show the effects of
Si, Ti and Al, the results for weld M
again accord poorly with the rest of
the data. For reasons w h i c h will
appear below, it was later considered
justifiable to repeat the regression
analysis without including the result
for weld M and the results of this are
also shown in Table 4.
No very convenient measure of
tearing resistance emerged from the
COD tests. Examination of fractured
COD specimens in the SEM showed
that small amounts of ductile tearing
preceded cleavage in all specimens
broken in the transition range, the
COD at the initiation of tearing, <5>,
being typically of the order of 0.1 m m
(0.004 in.), rising to 0.25 mm (0.010
in.) for weld M only. Precise definition of Si, was hindered by the fact that
in the early stages of tearing only a
few isolated microvoids could be seen
at high magnification, as compared
with the fibrous thumbnail visible to
the naked eye at higher values of <5C.
Potential drop (Ref. 24) and stress
wave emission (Ref. 25) techniques
did not reliably detect the appearance of the first microvoids. Since h< is
not a basis for practical design c r i teria, the ductile behavior of the experimental welds was mainly inferred
from tensile test results.
Tensile Tests
Tensile test results are given in
Table 5. The range of yield or 0.2%
proof strengths was f r o m 388 to 582
N / m m 2 and of tensile strengths from
486 to 734 N/mm 2 . Regression analysis showed that Mn, Si and Ti had a
significant effect on both yield and
tensile strengths, Table 6: weld M was
included in this analysis. Carbon did
not have a detectable effect in yield
strength and its effect on tensile
strength was not significant at the
95% probability level, but it seems
possible that this reflects the small
variation in
-^thpr
than the lack of real effect
:. ^CI.Icients for Al were large and positive
but not quite significant at the 95%

level unless results for the c o m m e r cial electrodes were included. Average RA values varied between 50 and
70%: individual results below 50%
were always associated with a small
pore. Tensile ductility was not d i rectly correlated with composition in
any simple way.
Microstructure
Table 7 gives the results of classifying the weld microstructures according to the simple scheme of Fig.
3. Thin foil and replicas studied in the
TEM and thermal analysis of the
welds as they cooled allowed the
structures seen to be rationalized as
follows.
In low dilution situations, mild steel
weld metals solidify as columnar 5
ferrite grains and austenite nucleates
at the boundaries of these as they
cool. If an austenite nucleus grows
into only one of the adjacent 5 ferrite
grains, the f> boundary will be preserved as an austenite boundary: if
the nucleus grows both ways, the
boundary will disappear. In fact, the
position of the original ferrite b o u n daries can often be revealed by solute
sensitive etches such as brominemethanol and this enables the relationship between the 5 ferrite and
austenite structures to be inferred,
since the austenite boundaries subsequently act as nucleation sites for
pro-eutectoid ferrite. In the present
work the austenite boundaries were
found to be largely independent of the
solidification boundaries, although
the structure remained columnar.
Structure I in the classification
scheme was pro-eutectoid ferrite,
which in addition to forming thin veins
at the austenite boundaries was also
seen as polygonal grains within the
columnar structure in weld metals of
low hardenability. Structure II, mainly
Widmanstatten ferrite side plates,
often grew continuously from the
massive grain boundary ferrite with
no intermediate sub-boundaries. Any
upper bainite which formed would
have been included as structure II on
the basis of its appearance in the light

microscope, but there was no evidence on replicas or thin foils e x a m ined in the TEM of the sheaf structure
of independently nucleated sub-units
which has been cited as a distinguishing feature of upper bainite.
Structure III may be described as
acicular ferrite and was the only one
of the four whose incidence in the experimental welds was negatively correlated with transformation temperature. Weld S, containing 8 8 % of acicular ferrite, transformed over the approximate range 630-450 C (1170840 F), while at the other extreme of
hardenability weld V yielded only 8%
of acicular ferrite after transforming in
the range 710-545 C (1310-1010 F).
The grain size of the acicular ferrite
was represented by a mean linear intercept in the range 1-2 (tm, the finer
structures corresponding to the lower
transformation temperatures.
The fourth key on the point c o u n ter was reserved for all structures
which could not readily be assigned to
the other three classes, but only one
type of structure made an important
contribution to this g r o u p in quantitative terms. This was a sub-pearlitic
aggregate of ferrite and carbides, recognizable but not usually resolved in
the light microscope. Electron microscopy showed the carbides to be
aligned in rows, w h i c h often a p peared parallel to a y-t-ct transformation boundary, suggesting an interphase precipitation mechanism. The
structure was found in welds of low
hardenability and strength but no a d v e r s e e f f e c t s on t o u g h n e s s w a s
detected.
Other microstructural constituents,
pearlite, retained austenite and martensite, were also observed, though
as these were not only low in total
quantity, but also so finely dispersed
that they were difficult to resolve and
identify in the light microscope, they
did not significantly influence the
overall microstructural assessment.
X-ray diffraction measurements of the
retained austenite are included in
Table 7. Since these were independent of the metallographic measurements the figures are not included in

Table 4 Regression Coefficients of 0.1 mm COD Transition Temperature T, on


Weld Metal Chemical Analysis
Regression coefficient (partial correlation coefficient)

Constant

Data set
A - U , W-Z
(V too porous)
A-Z, excluding M , V
A-Z, excluding M,V

-51
-

Mn

Si

Ti

Al

-45*
(-0.446)
-69***
(-0.786)
-68***
(-0.781)

28
(0.253)
10
(0.149)
not
included

703***
(0.632)
627***
(0.784)
630***
(0.782)

-301
(-0.270)
-390**
(-0.535)
-370**
(-0.520)

Multiple
correlation
coefficient
0.72
0.89
0.89

Significance levels:' 95%, " 9 9 % , *"99.9%

Table 5

- Tensile Test Results


Strain
hardening
exponent,
n

Yield or 0.2%
proof stress,
N/mm2

Tensile
str,
N/mm2

Elongation,

RA,

A
B
C
D
E

479
470
537
563
547

608
586
648
696
693

26
24
24
28
23

60
60
64
56
50

0.16
0.15
0.12
0.14
0.12

F
G
H
J1
J2

514
520
532
536
539

611
633
711
661
630

22
20
24
27
28

58
60
57
57
58

0.12
0.12
0.13
0.11
0.13

J1R
J2R
J2RR
K
L

537
503
504
453
453

660
604
608
565
559

24
27
28
26
34

60
60
60
55
69

0.13
0.16
0.15
0.13
0.16

M
N
O
P
Q

388
451
569
538
582

486
555
689
674
710

34
32
30
28
24

59
63
61
60
58

0.17
0.14
0.13
0.13
0.09

R
S
T
U
W

547
569
465
573
578

650
734
557
714
708

30
26
29
27
24

57
52
65
54
59

0.13
0.13
0.14
0.14
0.10

X
Y
Z
Com 1
Com 2

535
560
533
477
468

630
692
623
618
601

30
26
31
28
27

59
56
66
55
58

0.15
0.12
0.14
0.13
0.13

Weld

Table 6 Results of Regression Analysis of Weld Metal Yield and Tensile Strengths on
Electrode and Weld Composition, for Strengths in N/mm2 and Composition in Wt. %<a)
Regression coefficient (partial correlation coefficient)
Dependent
variable

Source
of analysis

Const.

Electrode

347

<7V

Weld

331

UTS

Electrode

400

UTS

Weld

358

<*Y

Mn

Si

Ti

Al

not
included
192
(0.07)
not
included
725
(0.26)

66***
(0.69)
93***
(0.66)
88***
(0.74)
111***
(0.71)

45*
(0.43)
56*
(0.44)
62**
(0.50)
78**
(0.55)

127"
(0.49)
601**
(0.48)
215***
(0.64)
865***
(0.60)

181
(0.34)
305
(0.27)
221
(0.36)
333
(0.28)

Multi
correl.
coeff.
0.77
0.85
0.88
0.90

(a) Significance levels: "95%; " 9 9 % ; ""99.9%

W E L D I N G R E S E A R C H S U P P L E M E N T ! 63-

the 100% sum of the latter results:


electron microscopy showed the austenite to be mainly present in the i n terstices of the acicular ferrite grains.
Inclusion Measurements
Table 8 lists the parameters describing the inclusion distributions in
selected experimental welds. Inclusions were initially measured in welds
A-H, which f o r m e d a half-factorial
section of the experimental plan. Attempts were made to relate inclusion
volume fraction to a series of c o m positional factors, among them the
M n / S i ratio. This latter was included
because a number of investigators
(Refs. 26-28) have shown an apparent connection between M n / S i ratio
and weld inclusion c o n t e n t s , f o l lowing a suggestion by Sekiguchi
(Ref. 29) that the fluidity of deoxidation products based on the M n - S i - 0
system, at temperatures just above
the steel freezing point, depends critically on this ratio. When it was found
that the M n / S i ratio, alone of the c o m positional parameters tried, gave a
useful correlation with the inclusion
content of welds A - H , a second set of
8 welds with a wider range of M n / S i

ratios was chosen for further measurements.


The results of all the inclusion measurements are plotted against the
M n / S i ratio of the electrodes in Fig. 8.
Although the correlation for the c o m plete set of results remained statistically significant at the 95% level, the
regression line shown is that originally derived for welds A - H . Other
authors (Ref. 28) have found that the
relationship breaks down at high
M n / S i ratios, which may explain the
high inclusion content of weld K.
Rather than attempting to represent
this by a second order regression
equation, it was considered that the
acceptable fit of the rest of the data
warranted the use of the original
equation. It is interesting to note that
this accurately predicted the inclusion contents of weld L and T on the
basis that both had a M n / S i ratio of
0.9 although the absolute levels of Mn
and Si were 50% higher in weld L;
however, the prediction broke down
for weld S which had the highest Mn
level of all. Similar results were o b tained plotting weld composition in
place of electrode composition.
Although the present work produced no grounds for disputing the

conventional view of the importance


of the M n / S i ratio in determining weld
inclusion contents, energy-dispersive x-ray analysis of inclusions in the
SEM showed that most contained
large amounts of Al and Ti in addition
to Mn and Si. The results are thus not
readily explained by the simple model
of Sekiguchi, and the case for regarding the M n / S i ratio as the main c o n trolling influence on weld metal inclusions remains a phenomenological
one.
Oxygen Analysis
The negligible solubility of oxygen
in solid steel implies a close relationship between oxygen and inclusion
contents in steel. The main reason for
embarking on time consuming direct
m e a s u r e m e n t of inclusions, however, was the lack of general success,
with certain exceptions (Ref. 20), in
correlating weld oxygen contents with
mechanical properties. One likely explanation for this was that not only the
volume fraction, but also the size and
spatial distribution of inclusions were
important in determining fracture behavior. The possibility of significant
errors in oxygen analysis seemed

Table 7 Microstructure of Welds as Determined by Point Counting, with Retained


Austenite from X-ray Diffraction. Columns I-IV Sum to 100% and the Austenite is Dispersed Through these Structures
I
Proeutectoid
ferrite

II
Lamellar
components

III
Acicular
ferrite

IV
Other
structures

Retained
austenite

A
B
C
D
E

37.2
26.5
21.4
16.7
27.3

10.1
2.8
1.9
0.5
5.5

41.3
64.6
65.1
78.7
49.9

11.4
6.1
11.6
4.1
17.3

F
G
H
J1
J2

26.3
36.4
23.7
22.4
24.2

3.6
6.3
5.5
5.1
1.5

55.6
20.2
57.8
58.7
67.3

14.5
37.1
13.0
13.8
7.0

J1R
J2R
J2RR
K
L

21.7
29.1
24.3
18.9
59.4

1.8
4.1
1.8
20.3
2.6

67.0
62.0
72.6
16.4
32.3

9.5
4.8
1.3
44.4
5.7

2.3
2.3
2.3
1.0
1.0

M
N
O
P
Q

45.5
25.2
15.7
28.9
2.1

11.3
13.5
2.0
11.3
2.2

10.2
22.4
78.0
30.4
92.7

33.0
38.9
4.3
29.4
3.0

n.d.
1.0
3.3
3.3
1.8

R
S
T
U
V

20.8
8.9
41.6
28.5
29.6

3.7
1.6
10.0
1.3
8.3

71.5
87.6
36.8
67.6
5.7

4.0
1.9
11.6
2.6
56.4

2.8
4.8
1.0
3.7
n.d.

10.4
23.2
11.9
24.0
21.4
27.2

3.7
0.8
0.2
4.1
23.3
21.2

63.7
74.3
82.9
65.6
13.9
6.8

22.0
1.7
5.0
6.3
41.4
44.8

1.7
1.8
1.4
2.2
2.6
1.9

Weld

X
Y
Z
Com 1
Com 2
n.d.: not detected

64-s I M A R C H

1976

Fig. 7 Partial regression of 0.1 mm COD


transition temperature on weld metal manganese content for welds A-Z

4
Mn/Si. wire

Fig. 8 Variation of inclusion volume fraction, Vy, in the deposit with Mn/Si ratio
of the wire

much more remote in view of the extensive d e v e l o p m e n t of s t a n d a r d


techniques.
In the attempt to relate metallographically determined inclusion contents to weld composition, the degree
of correlation with oxygen contents
d e t e r m i n e d by a c o n v e n t i o n a l
vacuum fusion method was very low.
This at first caused little concern since
the type of deoxidation product varied
from weld to weld, but it was then
realized that the density of the oxides
involved is so closely linked to the
equivalent weight of the deoxidants
that a given weight of oxygen produces quite similar volumes of MnO,
S i 0 2 , AI2O3, T i 0 2 and intermediate
products. When the estimated volume
fraction of MnS was included in the
calculations, the chemically determined inclusion contents remained
consistently lower than the metallographic figures by up to 50%.
Towards the end of the work an o p portunity arose to check some of the
oxygen contents by fast neutron activation analysis (Ref. 13). The amount
of material remaining was not sufficient to do this for all welds, but the
results of 10 analyses are shown in
Table 2. Given that both the activation analysis and the inclusion counts
had a standard error of about 10%,
there was now an acceptable agreement between the two sets of figures.
Discussion
The aims of weld deoxidation practice were defined at the outset as the
prevention of porosity, the minimization of weld inclusion contents and the
provision, through the deoxidants remaining in solution at the end of the
deoxidation reaction, of adequately
strong and tough weld microstructures. The first of these requirements
was the most easily satisfied, since
only in the Si-free weld V was serious
porosity encountered. Isolated pores
were found in weld M, but the addition of alloying elements required to
give a yield strength of 450 N/mm 2 (65
ksi) sufficed in all cases to produce
sound welds. In practice, deoxidant
additions somewhat higher than the
minimum would be used to provide a
margin of protection when welding in
drafts or on dirty plate.
Weld Cleanliness
It has been concluded that the weld
metal oxygen contents measured by
the vacuum fusion method cannot at
this stage be regarded as reliable
indicators of inclusion contents. A
tentative explanation for this is based
on the supersaturation of CO necessary to generate bubbles on very
small nuclei. The analytical method
u s e d r e q u i r e s the r e d u c t i o n of
a l u m i n a by c a r b o n at 1750 C
(3180 F), which will only proceed at
CO partial pressures of a few mi Hi-

Table 8 -- Parameters Describing the Size Distribution and Spatial


Distribution of Inclusions in Some Experimental Weld Metals
Inclusion
volume
fraction.

Geometric
mean
inclusion
diam, nm

Avg. vol.
nearest
neighbour
spacing,
A3, jim

No. of
inclusions
per mm 3 ,
N U X10 8

A
B
C
D
E

0.49
0.61
0.32
0.55
0.75

0.055
0.06
0.03
0.06
0.07

1.36
0.96
0.54
0.98
1.74

0.27
0.33
0.36
0.30
0.26

1.1
1.2
1.5
1.2
1.0

F
G
H
J1
K

0.43
0.41
0.50
0.41
0.51

0.03
0.05
0.06
0.03
0.03

0.45
0.52
1.03
0.29
0.65

0.41
0.35
0.28
0.50
0.42

1.6
1.5
1.2
1.8
1.4

L
N
S
T
U
Y

0.69
0.48
0.65
0.69
0.63
0.57

0.06
0.04
0.04
0.07
0.08
0.08

0.50
0.36
0.64
0.45
0.40
0.40

0.49
0.49
0.44
0.46
0.50
0.47

1.5
1.7
1.4
1.6
1.6
1.6

Weld

Est. std.
error in
vol. f r a c ,

meters of mercury. However, pressures much greater than this are required to sustain bubbles of the size
of weld metal inclusions: 1 atmosphere sustains a bubble of about 60
Mm (0.0024 in.) diam in molten steel
(Ref. 30). If the inclusions do not
coalesce into particles larger than
this, it seems possible that some may
not be reduced. Further investigations into the discrepancy between
different analytical methods are now
in progress, but in the meantime both
sets of oxygen figures have been
treated with caution.
It n e v e r t h e l e s s s e e m s c l e a r ,
whether we consider the chemical,
physical or metallographic estimates
of oxygen level, that this is independent, over a wide range, of the total
amount of deoxidants present and
their capacity to reduce oxygen activity in the molten steel. Only in the
porous weld V was there any indication of a high oxygen level associated
with a low total deoxidant level. The
only compositional parameter which
was significantly correlated with inclusion volume fraction was the Mn/Si
ratio. This has been assumed in the
past to influence the fluidity of the
deoxidation products and thus their
ability to coagulate with each other
and with the top slag. A simple calculation (Ref. 18) shows that flotation
according to Stokes' law is inadequate to explain inclusion removal, a
1 itim inclusion moving only about 1.7
Mm under its own buoyancy in the 5 s
or so for which the pool is molten, but
the concept of inclusions in a turbulent pool impinging on and being
held by the pool surface is attractive
and invites a reappraisal, in future
work, of the role of surface tension in
addition to that of fluidity of the deoxidation products.
WELDING

Fig. 9 Variation of tensile fracture


strain, e", with inclusion volume faction

Ductile Behavior
Given that it may be possible to use
the Mn/Si ratio to control weld cleanliness, it becomes useful to know how
this is related to the mechanical properties of the deposit. It appeared from
the inclusion counts on welds A-H that
a useful correlation existed between
tensile fracture strain e", calculated
as In [100/(100-%RA)] and inclusion
volume fraction V v , as work on other
materials has shown (31). However,
when further welds were examined to
see whether this c o r r e l a t i o n was
maintained over a wider range of
properties, it became clear that the
earlier view was oversimplified. The
complete set of results is shown in
Fig. 9 with the original regression
equation indicated by a dashed line.
The two welds which performed significantly better than predicted from
the inclusion volume fraction alone,
welds L and T, were characterized by
RESEARCH

SUPPLEMENT!

65-3

150

V*

50

40

30

20

o
oo

10

<?%

o
o

i
0

__

oo

-20

30
1

-40

01

02

03

Ot, 05

06 07
Si. V.

06

09

10

11

12

Fig. 10 Partial regression of 0.1 mm


transition temperature on weld metal
manganese content, excluding weld M

Fig. 11 Partial regression of 0.1 mm


transition temperature on weld metal
silicon content, excluding weld M

low yield strength, tTY, and a high


strain hardening exponent, n. The
complete set of results suggested that
both parameters played a part in
determining fracture strains, though
there was insufficient data to allow e*
to be expressed quantitatively as a
function of V v , "y and n.
Models of the microvoid coalescence process predict the d e p e n dence of fracture strain hardening behavior: if yield strength also has an
effect, this implies that the void nucleation process is critical. The fact that
this is so for mild steel weld metals is
confirmed by the observation that not
all the inclusions in the weld metal are
capable of nucleating microvoids: indeed, it has previously been shown
(Ref. 32) that in the relatively soft (<TY =
460 N/mm 2 = 67 ksi) "bench mark"
weld metal used in developing the inclusion counting technique, the size
distribution of inclusions taking part
in the fracture of a Charpy specimen
at 0 C (32 F) peaked at 0.7 pm,
whereas the v o l u m e d i s t r i b u t i o n
peaked at 0.3 um and only 15% of the
inclusions had a diameter of 0.7 um or
more. Examination of fractured tensile specimens in the SEM confirmed
this trend for the experimental welds
and a similar "cutoff" effect has been
reported by Hill and Passoja (Ref. 2).
It must be concluded that the practical benefits of producing cleaner
welds by suitable deoxidation practice are probably not extensive for
welds of relatively low strength. All the
present welds achieved tensile elongations of more than 20%, which in
the absence of stress concentrators
should be sufficient guarantee against
all failures other than those due to
gross overloading. It is more interesting to consider the resistance to
fracture initiating at a stress concentration or defect in a structure and
the simulation of this situation by
means of a notched test. Both experimental tests (Ref. 33) and theoretical models (Ref. 34) have d e m o n strated the strong interaction be-

tween strain hardening behavior and


triaxiality of stress. As the triaxiality
increases, the rate of strain hardening plays an increasing part in
determining the fracture strain and
the possibility of effecting useful improvements by producing cleaner
materials is correspondingly reduced.
In the welds studied here, the volume fraction of inclusions was reduced by raising the M n / S i ratio. Lowering the Si level, as far as was consistent with the prevention of porosity,
carried no penalty in terms of tearing
resistance but the addition of Mn refined the structure of the deposit and
regression analysis showed that this
raised its yield strength and lowered
its strain hardening exponent, n. Because the tensile fracture strain is relatively weakly dependent on n, some
improvement can result from the i n creased cleanliness as Mn is added,
but if the triaxiality of stress is increased to the point where unstable
fracture is likely to be of practical c o n cern, the improvement in cleanliness
is outweighed by the fall in n. Thus
there was no indication, in the present tests, that the COD at tearing initiation increased as the inclusion content fell over the range studied, and
only in the case of weld M, with a very
low strength and high n value, was <5i
significantly greater than the average.
Of course, where it is possible to improve cleanliness without major
microstructural changes, for e x a m ple by reducing sulfur contents or the
oxygen content of the shielding gas,
more unequivocal benefits may be
obtained (Refs. 20, 22). In multipass
welds also, the effects of weld cleanliness may be less masked by microstructural variations.

66-s I M A R C H

1976

C l e a v a g e Resistance
In order to explain the variations in
cleavage resistance among the experimental welds, the strengthening
mechanisms at work were first investigated. The good combination of

strength and toughness possible in


mild steel welds is known to result at
least in part from the development of
fine, acicular microstructures and the
first hypothesis tested was that the
proportion of acicular ferrite (Structure III) present directly controlled the
yield strength. A significant correlation was found, but the correlation
coefficient at 0.78 left room for improvement and other strengthening
mechanisms were considered.
While the strengthening effects of
Mn and possibly Si may be supposed
to result at least partly from changes
in the nature of the austenite d e c o m position products, that of Ti as shown
in Table 6 is so large in relation to any
effect of this element on hardenability either detected here or reported
elsewhere (Ref. 35) that only a dispersion hardening effect seems able
to explain it. This leads to the idea that
it might be possible to represent the
mechanisms controlling weld metal
strength by two factors, one related to
the nature and morphology of the
t r a n s f o r m a t i o n p r o d u c t s and the
other to matrix strengthening by solid
solution and dispersion hardening.
The simplest way of doing this would
be to take, for example, the proportion of acicular ferrite as the first factor and to ignore all elements but t i tanium in the second. This hypothesis was tested by regression analysis and led to the equation
n Y = 390 + 1.69 (% acicular ferrite)
+ 644 Ti

/g)
2

in which strength is in N/mm , both


terms are significant at the 99.9%
level and the correlation coefficient is
0.90.
Applying the same approach to the
prediction of COD transition temperature, simple regression of T c on the
various microstructural components
did not produce significant results but
with the inclusion of titanium the
equation giving the best correlation
was
T c = - 1 3 2 + 1.24 (% pro-eutectoid
ferrite) + 807 Ti
(9)
with a correlation coefficient of 0.70.
However, inspection of the data showed that the behavior of weld M was out
of line with the general trend.
In most of the experimental welds,
the soft pro-eutectoid ferrite formed
narrow veins separating blocks of
stronger acicular ferrite. When such
structures are tested, strain is c o n centrated in the pro-eutectoid ferrite
and it is here that cleavage cracks first
initiate, as was shown by metallographic examination of fracture profiles. Thus the embrittling effect of the
pro-eutectoid ferrite results, not from
any inherent brittleness of this c o m ponent, but from its "mortar-in-brickwork" morphology.

When the p r o p o r t i o n of softer


phases reaches 80 or 90%, as in weld
M, this analogy breaks down and
strain concentration may increase as
the proportion of acicular ferrite, now
the minority phase, increases. Edelson and Baldwin (Ref. 31) have remarked on the similarities in the
effect, as stress concentrators, of
hard and soft second phases.
In interpreting the present experimental results, two approaches are
possible. If it is allowed that weld M
differs qualitatively from the other
welds, data for this weld may legitimately be excluded from the regression analyses. Alternatively, second
order terms may be incorporated in
the regression equations so that these
fit the whole data set including weld
M, whose behavior is only a n o m alous in the sense that it is not predicted by the first order equations. In
view of the fact that the coefficients of
the second order terms could not be
determined with good precision, while
all the data except that for weld M
could be satisfactorily fitted to first
o r d e r e q u a t i o n s , the f o r m e r a p proach was adopted. The data set
without weld M now gave the regression equation
T c = - 1 4 2 + 1.87 (% pro-eutectoid
ferrite) + 771 Ti
/10)
whose correlation coefficient at 0.87
was little short of that obtained using
the full chemical analysis.
Consumable Design
A rational basis for weld deoxidation practice has now begun to
e m e r g e . This a c k n o w l e d g e s that
different factors control fracture in the
ductile and brittle modes and that in
formulating welding consumables,
priority must be assigned to performance in one or other mode. If the
ductile behavior is important, for
example Charpy upper shelf level,
there may be some advantage in using a Mn/Si ratio approaching 3-4 in
the electrode or 4-5 in the deposit to
limit inclusion contents. It is more
profitable to achieve this by limiting
the Si content, provided porosity is
avoided, than by raising the Mn content: the latter increases strength and
lowers the rate of work hardening,
both of which can offset the benefits
of improved cleanliness.
If, as is more usually the case,
cleavage resistance is the more important criterion, the effects of the
deoxidants are those shown in Figs.
10-13. These have been derived in the
same way as Fig. 7, but without using
the results from weld M. For the remaining welds, that is, all those with
yield strengths above 450 N / m m 2 (65
ksi), manganese had a u n i f o r m l y
beneficial effect within the range
studied, that is, from 0.45 to 1.6% in
the deposit. This is achieved through

30

'

s
0

50

40

30

20

o
o

0
j<

0
10
i

o
i

'

'

0 001 002 003 004 005 0 06007 006 009 0 10 0 ti 0 12

001 002 003 001. 005 006 007 0 08 009 0 10 On 012


r,. /.

Fig. 12 Partial regression of 0.1 mm


transition temperature on weld metal titanium content, excluding weld M

replacement of pro-eutectoid ferrite


by fine, interlocking acicular ferrite in
the weld structure and is accompanied by an increase in yield strength
of some 66 N / m m 2 (10 ksi) per 1%Mn
in the deposit.
No statistically significant effect of
silicon could be detected over the
range from 0.17 to 1.2% in the deposit. Titanium increased the transition temperature uniformly, within the
limits of experimental accuracy, as its
level in the deposit rose from 0 to
0.12%: it should be noted, however,
that these were single run welds and
some of the r e p o r t e d beneficial
effects of Ti in multipass welds could
result from a refinement of the reheated structure. There is also scope
for further investigation of Ti levels in
the range 0.01-0.015% which recent
work has shown to be important, (Ref.
36). Aluminum lowered the transition temperature over the range of 00.12% in the deposit, though this
result was rather heavily weighted by
a single point at 0.12%, and the 0.19%
Al in the second commercial weld did
not produce any detectable effect on
cleavage resistance: f u r t h e r work
would be needed to confirm this evidence of non-linearity in the effect of
Al.
Over the range of composition covered by the experimental weld metals
with strengths above 450 N/mm 2 (65
ksi) the above results are embodied in
the equation

Fig. 13 Partial regression of 0.1 mm


transition temperature on weld metal aluminum content, excluding weld M

-120-HO-100-90-80 -70-60 -50-40 -30-20 -10


Predicted T C .'C

10

Fig. 14 Comparison of observed 0.1 mm


COD transition temperature Tc with value
predicted from the regression equation Tc
= -1-68 Mn ' 630 Ti-370 Al. The results
for weld M and the commercial welds 1
and 2 were not used in deriving the regression equation
ture. Some additional deoxidation by
non-hardening elements such as Al
would have to be provided since weld
M was barely adequately deoxidized
and c o n t a i n e d o c c a s i o n a l p o r e s :
careful consideration would also have
to be given to the possible effects of
dilution in preventing achievement of
the desired structure.
Conclusions

-1 - 6 8 Mn + 630 Ti
370 Al

(11)

which is shown graphically in Fig 14.


If a low strength weld metal is
acceptable, with yield and tensile
strengths below about 400 and 500
N/mm 2 (60 and 75 ksi) respectively,
an alternative design philosophy is
suggested by the excellent toughness of weld M. This would involve reducing the amount of acicular ferrite
in the weld to very low levels and producing a homogeneous, soft struc-

1. Deoxidation
practice
affects
both weld inclusion contents and
microstructure and so can influence
resistance to both microvoid coalescence and cleavage fracture.
2. In as-deposited single pass C 0 2
welds with strengths above 450
N/mm 2 (65 ksi), good cleavage resistance is promoted by a fine acicular
ferrite structure and an absence of
d i s p e r s i o n h a r d e n i n g effects. For
welds made at 1.6 k J / m m (60 kJ/in.)
the 0.1 mm (0.004 in.) COD transition
temperature was related to c o m p o s i -

W E L D I N G R E S E A R C H S U P P L E M E N T ! 67-S

tion and microstructure by


Tc

Tc

- 1 - 6 8 M n + 6 3 0 Ti

3 7 9 Al
-142+1.87

(% p r o -

eutectoid ferrite)
+ 7 7 1 Ti
w h e r e T c is m e a s u r e d in d e g r e e s C.
3. T h e s e r e l a t i o n s h i p s d i d n o t a p p ly t o a w e l d of 3 9 0 N / m m 2 ( 5 6 k s i )
yield strength, which derived g o o d
t o u g h n e s s f r o m a relatively soft struct u r e w i t h o n l y 1 0 % of a c i c u l a r f e r r i t e .
4 . I n c l u s i o n c o n t e n t s in C 0 2 w e l d s
a r e n o t r e l a t e d t o t h e t o t a l a m o u n t of
deoxidants present unless these are
at s u c h l o w l e v e l s t h a t p o r o s i t y is i m minent. There a p p e a r e d to be a relationship between inclusion content
a n d M n / S i r a t i o in t h e e x p e r i m e n t a l
welds, the cleanest welds being prod u c e d w i t h M n / S i r a t i o s of 3 - 4 in t h e
e l e c t r o d e o r 4 - 5 in t h e d e p o s i t .
5. T h e r e l a t i o n s h i p b e t w e e n w e l d
cleanliness a n d resistance to f i b r o u s
f r a c t u r e is n o t s i m p l e b u t d e p e n d s o n
the strength and strain hardening b e h a v i o r of t h e m a t r i x . A s t h e t r i a x i a l i t y
of s t r e s s i n c r e a s e s , m a t r i x p r o p e r t i e s
b e c o m e increasingly important and
high yield strengths and low strain
hardening exponents reduce tearing
resistance.
Acknowledgments
The author is indebted to Dr. R. Week,
Director General of the Welding Institute,
for permission to present this paper. His
thanks are also due to his colleagues and
to Dr. J. F. Knott for many helpful discussions and to Mr. B. G. Ginn and all those
who helped with the experimental work.
References
1 . C h i n , L. L - J . , " A M o d e l f o r
Toughness Studies of Welds,"
Welding
Journal,
48 (8), Aug. 1969, Research
Suppl., 290s-294s.
2. Hill, D. C. and Passoja, D. E.,
"Understanding the Role of Inclusions and
M i c r o s t r u c t u r e in D u c t i l e F r a c t u r e , "
Welding Journal, 53 (11), Nov. 1974, Research Suppl., 481s-485s.
3. Widgery, D. J., "Deoxidation Practice
and Toughness of Mild Steel Weld Metal,"
Welding Research International, 4 (2), Feb.
1974, 54-80.
4. Widgery, D. J., "Deoxidation Practice
and Toughness of Mild Steel Weld Metal Report 2," Welding Institute Members'
Report M/78/74, August 1974.
5. Pollard, B. and Milner, D. R., "Gasmetal Reactions in C 0 2 Welding," Journ.
Iron and Steel Inst., 209 (4), Apr. 1971,
291-300.

6. Sims, C. E., "The Behaviour of Gases


in Solid Iron and Steel," Gases in Metals,
A S M , Cleveland, 1953, 119-198.
7. Hilty, D. C. and Crafts, W., Trans.
Met. Soc. AIME, 188, 1950, 424-436.
8. Burdekin, F. M. and Stone, D. E. W.,
"The Crack Opening Displacement A p proach to Fracture Mechanics in Yielding
Materials," Journal of Strain Analysis, 1
(2), Feb. 1966, 145-153.
9. S a k a k i , H., " E f f e c t of A l l o y i n g
Elements on Notch Toughness of Basic
W e l d M e t a l s , " Journal
of the
Japan
Welding Society, 28 (12), Dec. 1959. 858863; 29 (2), Feb. 1960, 102-105; 29 (6),
June, 474-484; 29 (7), July, 539-544; 29
(12), D e c , 940-945; 30 (4), April 1961, 242247; 30 (6), June, 394-397: Welding Institute Translation.
10. Cochran, W. G. and Cox, G. M., Experimental
Designs, John Wiley & Sons,
New York, 1957, 342-347.
11. Gross, J. H., "The New Development of Steel Weldments," Welding Journal, 47 (6), J u n e 1968, Research Suppl.,
241s-270s.
12. Widgery, D. J., "The Influence of
Microstructure on Fracture Initiation in
Mild Steel Weld Metals," Welding Institute
Members' Report M/46/69, May 1969.
13. A r m s o n , F. J. and Bennett, H. L.,
"Determination of Oxygen in Steel by
Neutron Activation Using the 'Analox',"
Journal of the Iron and Steel Institute, 208,
(8), Aug. 1970, 748 to 7 5 1 .
14. M e t h o d s f o r C r a c k
Opening
Displacement (COD) Testing: Draft for
Development 19, British Standards Institution, 1972.
15. Hollomon, J. H., "Tensile Deformation," Trans. Met. Soc. AIME, 162 1945,
268-290.
16. Gloor, K., Christensen, N., Maehle,
G. and Simonson, T., Welding in the
World, 4 (2), Feb. 1966, 70-85: ex IIW Doc.
11-281-63, 1963.
17. Scheil, E., "Die Berechnung der A n zahl und Grossenverteilung kugelformige
Kristalle in undurchsichtigen Korpen mit
Hilfe der d u r c h einen ebenen Schnitt
erhaltenen Schnittkreise," Zeitschrift
fur
Anorganische
und Allgemeine
Chemle,
201, 1931, 259-264.
18. Widgery, D. J., Deoxidation Practice
and the Toughness of Mild Steel Weld
Metal, Ph.D Thesis, University of C a m bridge, November 1974.
19. Ruckdeschel, W. E. W., "Inclusions
in Mild Steel MAG Weld Metal their
Composition, Size and Number," IIW Doc.
XII-B-174-74, 1974.
20. Hannerz, N. E. and Lowery, J. F.,
"Influence of Micro-Slag Distribution on
M I G - M A G W e l d Metal I m p a c t P r o p erties," Metal Construction,
7 (1), Jan.
1975, 21-25.
21. Ashby, M. F. and Ebeling, R., "On
the Determination of the Number, Size,
Spacing and Volume Fraction of Spherical
Second-Phase Particles from Extraction
Replicas," Trans. Met. Soc. AIME, 236,

1966. 1396-1404.
22. Steel, A. C , "The Effects of Sulphur
and Phosphorus on the Toughness of Mild
Steel Weld Metal," Welding Institute M e m bers' Report M / 6 4 / 7 1 , July 1971.
23. Stumpf, W. E. and Sellars, C. M.,
"Measurement of Particle Density and
Volume Fraction from
Extraction
Replicas," Metallography,
1 (1), Jan. 1968,
25-34.
24. M c l n t y r e , P. and Priest, A. H.,
" M e a s u r e m e n t of S u b - C r i t i c a l Flaw
Growth by the DC Electrical Resistance
Technique," British Iron and Steel Research A s s o c i a t i o n Report M G / 5 4 / 7 1 ,
June 1971.
25. M i r a b i l e , M. a n d P a l o m b i , E.,
"Dependence of Stress Wave Emission
upon Brittle and Ductile Fracture Mechanisms," Third International Conference
on Fracture, Munich, 1973, Vol III, Paper II234.
26. Born, K., "Zusammensetzung und
Form nichtmetallischer EinschlCisse in
L i c h t b o g e n s c h w e i s s u n g e n , " Stahl
und
Eisen, 74 (13), 1954, 822-831.
27. North, T. H. and Wallace, E. P.,
" P r o p e r t i e s of 2 ' / 2 % C r - 1 % M o Welds,"
W e l d i n g of C r e e p R e s i s t a n t S t e e l s ,
Welding Institute Conference, Newcastle
upon Tyne, 1970, 115-125.
28. Tuliani, S. S., Boniszewski, T. and
E a t o n , N. F., " N o t c h T o u g h n e s s of
Commercial S u b m e r g e d - A r c Weld Metal,"
Welding and Metal Fabrication,
37 (8),
Aug. 1969, 327-339.
29. Sekiguchi, H., Theory and Proposal
on Steel
Fusion
Welding
and
their
Applications,
Nikkan Kogyo Shinbun,
Tokyo, 1964.
30. Darken, L. S. and Gurry, R. W.,
Physical Chemistry
of Metals, McGrawHill, p 4 9 1 .
31. Edelson, B. I. and Baldwin, W. M.,
ASM Transactions
Quarterly,
35, 1962,
230-250.
32. Dolby, R. E., Saunders, G. G. and
Widgery, D. J., "Metallurgical Factors C o n trolling the Fracture Toughness of Weld
Microstructures, with Particular Reference to the Micromechansim of Fracture
Initiation," The Practical Implications of
Fracture Mechanisms, Conf. P r o c , Institution of Metallurgists, London. 1973, 54-62.
33. Schwartzbart, H. and Brown, W. F.
Jr., " N o t c h - B a r Tensile P r o p e r t i e s of
Various Materials and their Relation to the
Unnotch Flow Curve and Notch Sharpness," ASM Transactions
Quarterly, 46,

1954, 998-1020.
34. McClintock, F. A., "A Criterion for
Ductile Fracture by the Growth of Holes,"
Journal of Applied Mechanics, 35 (2), Feb.
1968, 363-371.
35. Comstock, G. F., Titanium in Iron
and Steel, John Wiley and Sons, New York,

1955, 103-112.
36. Sawhill, J. M. Jr. and Wada, T.,
"Properties of Welds in Low Carbon M n Mo-Cb Line Pipe Steels," Welding Journal,
54(1), Jan. 1975, Research Suppl.. 1s-11s.

The American Welding Society publishes over a hundred different Codes, Standards,
Recommended Practices, Guides, Handbooks and other books related to welding. If you
would like a copy of our Publications List, simply circle No. 150 on our Reader Info-Card
and mail it.

68-s I

MARCH

1976

You might also like