Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Journal of Materials Processing Technology 214 (2014) 2739–2747

Contents lists available at ScienceDirect

Journal of Materials Processing Technology


journal homepage: www.elsevier.com/locate/jmatprotec

Real time monitoring of phase transformation and strain evolution in


LTT weld filler material using EDXRD
J. Gibmeier a,∗ , E. Held a , J. Altenkirch b , A. Kromm c , Th. Kannengiesser c , Th. Buslaps d
a
Institute for Applied Materials, Karlsruhe Institute for Technology (KIT), Campus South, Kaiserstr. 12, 76131 Karlsruhe, Germany
b
Siemens AG, Rheinstr. 100, 45478 Mülheim an der Ruhr, Germany
c
BAM Federal Institute for Materials Research and Testing, Unter den Eichen 87, 12205 Berlin, Germany
d
European Synchrotron Radiation Facility, Rue Jules Horowitz, 38000 Grenoble, France

a r t i c l e i n f o a b s t r a c t

Article history: For a newly developed 10% Cr and 10% Ni low transformation temperature (LTT) weld filler material, the
Received 24 January 2014 local phase transformation kinetics and the strain evolution during gas tungsten arc welding (GTAW)
Received in revised form 6 June 2014 under real welding conditions was studied. An experimental set-up and a measuring and evaluation
Accepted 7 June 2014
strategy are presented to gain a real time insight into the welding process. The experiments were carried
Available online 16 June 2014
out at the beam line ID15@ESRF using a two detector EDXRD (energy dispersive X-ray diffraction) set-up
and high energy synchrotron X-rays. The time-resolved diffraction analysis during welding was carried
Keywords:
out locally throughout the weld in longitudinal as well as in transverse direction to the weld line to
In situ synchrotron X-ray diffraction
Low transformation temperature
examine the interdependence of the strain state and the transformation kinetics. This comprehension is
Welding crucial for the optimization of the weld process, and thus for the tailoring of the resulting residual stress
states, which is one of the main issues for the application of LTT alloys. Using the herein proposed approach
EDXRD diffraction pattern can be monitored during real welding with a counting rate of 5 Hz. By means
of the time resolved diffraction data the local transformation temperatures and times were determined
and the local phasespecific strain evolutions are discussed with respect to the transformation rates and
the time-delayed phase transformations.
© 2014 Elsevier B.V. All rights reserved.

1. Introduction means of low transformation temperature (LTT) filler materials. The


application of low transformation temperature (LTT-) filler materi-
Tensile residual stresses generated in welds are of major con- als allows controlling the transformation such that the associated
cern for structural integrity assessment in industrial components as strain arising from the martensite transformation compensates for
residual stresses superimpose onto the in service load and reduce the thermal contraction strains. LTT-alloys can be applied to miti-
the fatigue limit. This was emphasized, e.g. by Williams and Starke gate detrimental tensile residual stress generally arising around the
(2003) and Withers (2007), who both studied the impact of residual weld-line or (depending on the LTT-alloy and the welding parame-
stresses on the failure behavior of welded components. Conse- ter) to generate even compressive residual stress in the weld zone
quently, it is desired to reduce tensile welding stresses or introduce as shown, e.g. by Dai et al. (2008), who carried out neutron diffrac-
compressive stresses in the vicinity of the weld line. In order to tion stress analysis for LTT welds. The development of special weld
generate beneficial compressive residual stresses at the surface of filler materials, which exhibit a low phase transformation temper-
a welded component, various post-weld treatment procedures are ature, was reported by Ohta et al. (1999a,b). The decrease of the
available like shot peening, ultrasonic peening, etc. However, these martensite start temperature Ms was mainly achieved by the allo-
post-weld treatments are time and cost extensive and depending ying elements nickel and chromium. In Wang et al. (2002) a concept
of the geometry of the component often impractical. An attrac- for an appropriate alloying concept and for the design of the elec-
tive alternative is to generate compressive residual stresses over trodes was presented and the impact of the Ms temperature on the
the complete weld joint in the course of the welding procedure by compressive straining of the weld was illustrated. In Moat et al.
(2011) thermodynamic modeling was carried out to design and
to optimize low Ms temperature ferritic alloys with respect to the
∗ Corresponding author. Tel.: +49 721 608 42675; fax: +49 721 608 48044. toughness of the material and the process induced residual stresses.
E-mail address: jens.gibmeier@kit.edu (J. Gibmeier). In previous work on 10% Cr/10% Ni LTT filler the large impact of

http://dx.doi.org/10.1016/j.jmatprotec.2014.06.008
0924-0136/© 2014 Elsevier B.V. All rights reserved.
2740 J. Gibmeier et al. / Journal of Materials Processing Technology 214 (2014) 2739–2747

the alloying elements on the transformation temperature for LTT-


alloys was indicated as e.g. presented in Kromm and Kannengiesser
(2009).
Various studies on the improvement of mechanical properties,
e.g. the fatigue performance or crack initiation and their propa-
gation behavior in the component can be found in literature. As
an example Bhadeshia (2004) demonstrated the improvement of
the fatigue behavior of welded martensitic and bainitic steels by
an appropriate tailoring of the microstructure such that the defor-
mation caused by transformation eliminates residual stresses. In
Ohta et al. (1999a,b) and in Ohta et al. (2003) the fatigue behavior
of LTT welds for a butt weld and for a lap joint was investigated,
respectively. In all cases the fatigue behavior of the LTT joint was
found to be superior compared to conventional filler materials. The
explanation of this gain in mechanical integrity is that the com-
pressive residual stresses induced in the weld sustainably support
the crack closure effect and hence the fatigue behavior. However,
the mechanical approach for residual stress analysis only give an
integrative value of the residual stresses released during the cutting Fig. 1. Micrograph taken below the surface of a weld with LTT-alloy containing
10 wt% Ni showing the cellular arrangement of. Etching agent: Lichtenegger and
process and do not account for the local residual stress distribution.
Bloech LB1. Welding direction is normal to the imaged plane.
An understanding of the phase specific transformation kinet-
ics of LTT alloys is important for the further development of
these alloys and especially for understanding the generation of is essential with respect to tailoring the respective local residual
residual stresses during welding and the complex interactions of stress distribution, e.g. to minimize welding induced distortion. By
phase transformation, thermal strains, local stress distribution and this means straitening operations after welding of large compo-
restraints from the surrounding material. Thus, it is essential to nents can be reduced.
carry out studies to improve the fundamental understanding of the
local (micro-) structural evolutions as well as the development of 2. Experimental
stress and strain and to be able to optimize the chemical composi-
tion of the LTT-alloy and of the welding process. 2.1. Material
Dilatometer experiments give only limited information about
the complex interactions during welding. Using Synchrotron X-ray The LTT-filler material studied in this investigation is designed
diffraction methods, the phase specific transformation kinetics of according to the 10% Cr–10% Ni alloy concept presented, e.g. by
LTT materials have been studied by several authors as a function of Ohta et al. (1999a,b) for a high-strength filler material with lowered
time, temperature and composition. E.g. Stone et al. (2008) pro- Ms -temperature. The detailed composition is displayed in Table 1.
posed a set-up to study the phase transformations of LTT weld The calculation of Ms from the chemical composition of the alloy
metals being reheated by means of a electro-thermal mechani- after Steven and Haynes (1956) reveals a martensite start tem-
cal test rig using high energy synchrotron radiation. However, to perature of 179 ◦ C. Somewhat lower transformation temperatures
date the reported investigations were performed under conditions were determined by Altenkirch et al. (2011) in in situ oven exper-
dissimilar to real welding processes. iments for the identical LTT-alloy studied here. Here, the phase
Altenkirch et al. (2011) carried out in situ furnace experiments transformation was observed by soft synchrotron X-ray diffrac-
at the PDIFF beam line at ANKA (KIT, Karlsruhe) to determine tion at the surface near region. The sample geometry prevented
transition temperatures and phase specific thermal expansion a free shrinkage, which certainly has a significant influence on the
coefficients. In furnace experiments the experimental conditions transformation kinetics.
are far from real welding experiments, especially regarding tem- The weldability of LTT-filler materials was proven by Kromm
perature gradients, mechanical constraints and heating and cooling (2011). The mechanical properties of the LTT alloy, the yield
rates. Further, when using the furnace set-up and the soft syn- strength, the ultimate tensile strength and the elongation to frac-
chrotron X-rays the information depth is limited to the near ture are given in Table 1.The base metal for the current investigation
surface region while for the discussion of welding applications is the fine grained construction steel S690Q). The chemical compo-
diffraction information from the bulk of the weld line is essen- sition and the mechanical properties yield strength  ys , ultimate
tial. Kannengiesser et al. (2009) presented in situ experiments to tensile strength UTS and elongation at fracture are also given in
monitor the phase transformations during welding. However, the Table 1.
experimental set-up did not allow a sufficient time resolution to Fig. 1 shows a micrograph taken below the weld surface of
observe phase transitions properly and to determine specific trans- the LTT-weld. The weld microstructure consists of a martens-
formation temperature. Results from LTT welding experiments ite (˛ )/austenite () duplex structure with a varying ˛ /-phase
closer to the actual welding process were reported by Gibmeier fraction which depends on the local Ms -temperature and the solid-
et al. (2014) Here, ex situ results (subsequent to the welding) for ification conditions. The solidification can be described as cellular
welding under an additional influence of the welding parameters with martensite in the cell cores and austenite in the intercel-
and under mechanical constraints were presented. lular regions. The reason for the existence of austenite at room
In the present project a set-up for fast in situ synchrotron X- temperature is the pronounced segregation of Cr and Ni during
ray diffraction studies, which allows monitoring diffraction spectra the primary austenitic solidification. Intercellular areas contain-
in real time under conditions similar to welding, is established. ing higher Cr and Ni contents solidify later than the cell cores.
A method is presented for studying local transformation kinetics An increased amount of alloying elements suppresses the mar-
and evolution of strains during welding of a LTT weld filler mate- tensite formation even at ambient temperatures leading to local
rial throughout the weld to gain a fundamental understanding of varying amounts of retained austenite. Regarding the Schaeffler
the residual stress built up during welding using LTT-fillers. This diagram small amount of austenite would be expected at room
J. Gibmeier et al. / Journal of Materials Processing Technology 214 (2014) 2739–2747 2741

Table 1
Chemical composition in wt% of the LTT-alloy and the base material S690Q as well as selected material properties with indication of the references from which some of the
data are taken from. The martensite start temperature was calculated according to Steven and Haynes (1956).

C Ni Cr Mn Si P S Mo Nb V B Fe

LTT alloy 10% Ni 0.04 10 10 0.7 0.4 Bal.


S690Q 0.116 0.481 0.498 1.52 0.402 0.0017 0.001 0.111 0.005 0.054 0.001 Bal.

LTT-alloy 10% Ni S690Q (data from Kromm and Kannengiesser, 2009)

Ms (calculated) = 179 ◦ C  ys ∼ 340 MPa (Kromm, 2011)  ys ∼ 690 MPa


Ms (Altenkirch et al., 2011) = 82 ◦ C UTS > 800 MPa UTS ∼ 770–940 MPa
(Kromm, 2011) Elongation of 14%

temperature for the chosen LTT-alloy. Irrespective of this item, 7.6 mm in beam direction is available. Strain measurements were
austenite observed at room temperature is considered as retained performed in transmission such that the strain was determined in
austenite in the following. the weld longitudinal (Ge-det 0) and weld transverse (Ge-det 1)
direction simultaneously within one exposure. With the described
∗ = 3◦ away
set-up, the strain is actually measured in a direction QL/T
2.2. Thermal exposure for in situ welding
from the weld longitudinal (QL ) and transverse (QT ) directions:
The thermal cycle for the in situ diffraction experiment was however, the difference in strain between the measured and the
applied by online gas tungsten arc welding (GTAW). A special actual in-plane directions is negligible.
designed online welding device, which was first presented by To provide fast data analysis a special readout configuration of
Kannengiesser and Kromm (2009) as applied for comparable stud- the multi-channel (MCA) electronics developed by the ESRF BLISS
ies, is depicted in Fig. 2a) with two parallel translation axes, was group was provided. This high speed software/hardware architec-
used with the welding torch mounted on one axis and the weld ture consists of a commercial MCA acquisition card (XIA LLC) which
sample on the second one. can be read out in ∼20 ms and a multi-purpose digital I/O NIM
This set-up allows translating the welding torch along the static module built around a programmable logic device and a micro-
sample to monitor the material behavior at a fixed weld line posi- controller (=OPIOM; developed by the Instrumentation Services &
tion. For simulation of the welding process, a S690Q base plate Development Division at ESRF). The Opium NIM module was pro-
(100 mm × 80 mm × 6 mm) with a layer of LTT material welded in grammed to synchronize the acquisition of the MCA data and the
multiple beads onto the long edge of the base metal and subse- operation of the welding unit as firstly reported by Altenkirch et al.
quently machined to 100 mm × 10 mm × 5 mm was used. By this (2012). In order to observe the phase kinetics and strain evolution
means, the welding process monitored during the in situ diffrac- during welding throughout the sample, diffraction spectra were
tion experiments is a re-fusion of the pre-deposited weld made collected every 200 ms (5 Hz) at 10 lateral distances (0.5–5 mm
from the LTT-filler material, which leads to similar results of phase distance from the top of the weld line in 0.5 mm steps by shif-
transformation kinetics and strain evolution in case of solidification ting the sample along the transverse direction). For each run the
without constraint as for the we prior welding process. A schematic weld tool was traversed along the pre-welded LTT-material. Data
illustration of the plate is shown in Fig. 2b). GTAW was performed were collected until the sample reached ambient temperatures. In
at a feed rate of 80 mm/min over a 65 mm long weld path with a between runs a 5 min delay was included to ensure the sample
welding current of 80 A and argon as shielding gas. The tungsten was at the same starting temperature for each measurement. The
electrode was positioned approximately 1 mm above the surface. welding cycle start was triggered electronically by the diffraction
Surface temperature measurements as a function of time and dis- measurement.
tance to the weld pool were carried out using an infrared camera
during offline welding under identical conditions. The temperature
data were confirmed using thermocouples. 2.4. Data analysis

The EDXRD data were normalized against the effective storage


2.3. Time-resolved energy dispersive X-ray diffraction
ring current and the energy-intensity distribution function of the
synchrotron source. Moreover, a correction for the spurious peak
Time resolved energy dispersive synchrotron X-ray diffraction
shifts due to dead time variations was carried out according to
(EDXRD) measurements were carried out at the multi-purpose
the procedure reported by Altenkirch et al. (2012) for the identical
X-ray diffraction beamline ID15A@ESRF, which is dedicated to
EDXRD set-up applied here. For the investigation of the transfor-
applications using high flux and high energy synchrotron X-rays.
mation kinetics the temperature at which a diffraction peak (dis-)
The polychromatic beam is generated by a 1.6 m long asymmetric
appeared was determined. Single peak fitting by using a Pearson
wiggler and the photon energy available reaches 500 keV, which
VII-function was applied for the ˛ -phase {2 0 0}, {2 1 1}, {2 2 0},
together with the high flux available provides penetration depths
{3 1 0}, {2 2 2} and {3 2 1} as well as -phase {2 2 0}, {3 1 1}, {2 2 2}
of several centimeters in most engineering materials. For precise
and {4 2 0} diffraction lines respectively and the energy Ehkl for each
details of the measurement arrangement it is referred to the stud-
line was converted into a lattice specific unit cell parameter ahkl
ies of Altenkirch et al. (2008), who used a similar set-up for their
using:
studies.
The set-up consists of two energy-discriminating TRP-type Ge- 
detectors optimized for high count-rates, which were placed under hP c · (h2 + k2 + l2 )
ahkl = (1)
a scattering angle 2␪ = 6◦ such that the scattering vectors are 2E ∗ sin 
orthogonal (vertical and horizontal) to each other (see Fig. 3). With
the primary set of slits set to 400 ␮m × 400 ␮m and a secondary set with hP being the Planck’s constant and c the speed of light. Assum-
of collimating slits (100 ␮m) in front of each detector a diamond ing a texture factor FT = 1 for this first approach (texture free grain
shaped gauge volume of 0.61 mm3 with an elongation of approx. orientation) and using the lattice plane multiplicity factor mhkl as
2742 J. Gibmeier et al. / Journal of Materials Processing Technology 214 (2014) 2739–2747

Fig. 2. Illustration of the welding device used for online welding (a) and a schematic illustration of the test samples (b). 10 layers off the LTT alloy were pre-welded onto the
long edge of a 6 mm thick base plate (S690Q) and subsequently machined into shape.

well as the hkl plane specific Young’s modulus Ehkl a {h k l} plane NIST (Chantler, 1995) the phase fraction of retained austenite was
specific weighting factor m∗hkl was calculated using: calculated according to the procedure given in the textbook of
Macherauch (1992):
FT · mhkl · Ehkl
m∗hkl = (2)
E
which was presented by Daymond (2004) for the analysis of mul- 100 vol%
v =       (4)
tiple diffraction peaks of one phase, where E is the macroscopic m  m 
˛i ·I˛i i ·Ii
phase specific and temperature dependent Young’s modulus. The 1+ R˛i
/ m˛i / Ri
/ mi
variation of Ehkl and E as a function of temperature was calculated i=hkl i i=hkl i
based on data taken from Davis (1996) for the ˛ - and -phase,
respectively. The unit cell parameter a for each phase was calcu-
lated using: Assuming 1 = v + v˛ the phase fraction of the ˛ -phase can be

a · m∗hkl
hkl hkl
determined.
a=  ∗
(3) For the calculation of the lattice strain the stress free unit cell
hkl
mhkl
parameter a0 has to be known as a function of the temperature for
This determination of the unit cell parameter is valid for the both the ˛- and the -phase. For the -phase it was assumed that
austenite and for the martensite phase. Due to the high amount of the first solidified grains are stress free at high temperatures. The
nickel and chromium in the LTT alloy a cubic body centered mar- stress free unit cell parameter a0 of the -phase was determined at
tensite lattice structure exist that corresponds to the findings of the solidification temperature and afterwards calculated for lower
Reed (1962) for studies on a comparable alloy. temperatures via the coefficient of thermal expansion (CTE). The
The phase fraction of retained austenite was calculated accord- temperature measurement, which is the basis of this approach only
ing to the procedure defined in Laine (1978) who proposed an account to 2% in maximum, hence, the effect will be neglected in
evaluation procedure to determine the volume fractions of ferrite in the following. The stress free unit cell parameter of the ˛ -phase
austenitic steels using EDXRD and Faninger and Hartmann (1972) was determined at room temperature using the comb-method in
who discussed the physical bases of quantitative X-ray phase anal- V-groove weld joints as reported by Gibmeier et al. (2014) for the
ysis. Using the integrated peak intensity Ihkl of the ˛ - and -phase identical welding samples to a0,˛ = 2.871 Å. The CTE was measured
peaks and the multiplicity factor m∗hkl as well as the temperature earlier by Altenkirch et al. (2011) to ˛˛ = 11.39 K\10−6 for the ˛ -
dependent R-factor Rhkl (calculation is based on data taken from phase and to ˛ = 28.39 K\10−6 for the -phase.

Fig. 3. Schematic illustration of the EDXRD diffraction experiment set-up for monitoring the phase transformation kinetics and the strain evolution using an online welding
rig, which allows for automatically traversing the GTAW weld torch along the pre-welded layer of LTT material. As a result of the set-up, the scattering vectors are slightly
rotated by  out of the in-plane directions.
J. Gibmeier et al. / Journal of Materials Processing Technology 214 (2014) 2739–2747 2743

The temperature distribution is superimposed. While Fig. 4


shows the complete welding cycle with heating up, melting, solidi-
fication and cooling down, the analyzed {h k l} diffraction peaks are
indexed.
In addition to the expected diffraction line from the ˛ - and the
-phase of the welded steel plate fluorescence lines of tungsten (W)
and lead (Pb) arise, which can be assigned to interactions from the
synchrotron beam with the materials used for the slits and for the
shielding. Those interference lines near the WK˛1 , WK˛2 , PbK˛1 and
PbK˛2 fluorescence lines or those where the signal-to-background
ratio was too low for proper evaluation are denoted in prentices
and were not used for data analysis. The {2 2 2} ˛ - and {4 2 0}
-peaks were not used because their energies were to close for
accurate peak separation, which is especially important near the
transformation temperatures.

3.1. Phase transformation kinetics

The phase transformation kinetic was analyzed by interpreta-


tion of the diffraction line intensities. For the phase transformation
only results from the detector in the longitudinal direction are
shown here, since this detector shows a much higher efficiency and
therefore interference lines during transformation can be detected
at earlier stages of the transformation process. The considerable dif-
ferent efficiencies between the two detectors are due to the linear
polarization of the synchrotron radiation in the horizontal plane as
well as the different voltage settings, which must be applied to the
identical detectors.
After around 21 s into the weld cycle the intensities of the (partly
already existing) -diffraction peaks increase indicating the start of
the martensite-austenite transformation as shown in Fig. 5. This
results in an Ac1 -temperature of approximately 580 ◦ C in 3 mm
distance from the top of the weld line. With the ˛ -phase being
completely dissolved and the diffraction peaks disappearing after
22 s and at approximately Ac3 = 675 ◦ C, the martensite–austenite
transformation is finished. As a result of the high heating rates up to
238 K/s the -peaks were apparent for approx. 2.5 s only before no
detectable peak was apparent due to the high temperature starting
at T = 1175 ◦ C. However, even after the disappearance of continues
diffraction lines around maximum temperatures of 1338 ◦ C sporad-
ically some transient interference line signal occurs indicating that
Fig. 4. Density plots of the diffraction energy vs. intensity (logarithmic scale) and
some fraction of crystal order remains at these temperatures.
time in the weld longitudinal and transverse direction obtained 3 mm away from the
weld surface at 5 Hz count rate in the LTT alloy. The local temperature distribution
Upon cooling the {3 1 1} -peak appeared after 27.6 s into
is superimposed by the red curves. (For interpretation of the references to color in the weld cycle for approx. 2.6 s at a temperature ranging from
this figure legend, the reader is referred to the web version of the article.) 1325 ◦ C to 1175 ◦ C in the longitudinal direction indicating spo-
radic grain formation and orientation (Fig. 6b). On the other
hand, in the longitudinal direction the {4 2 2} -peak appeared
Using the strain free unit cell parameter the elastic strains were
continuously starting at 1251 ◦ C after 29.2 s (Fig. 6a). After 63 s
calculated.
into the weld cycle at 403 ◦ C the {3 1 1} -peak is constantly
adet 0 (T ) − a∗0 (T ) adet 1 (T ) − a∗0 (T ) visible in both detectors, while other -peaks appear sporad-
εel,long = , εel,trans = (5)
a∗0 (T ) a∗0 (T ) ically, e.g. {2 0 0}, {4 0 0},{4 2 0}. Differences in the intensities
recorded in the two solid state detector are due to differences
Local change in the chemical position might affect the diffraction in their sensitivity and particularly due to the development
data and thus the determined elastic strain data, but due to the of local crystallographic textures during cooling down. During
size of the used gauge volume we expect that this effect can be further cooling down the austenite-martensite transformation
neglected. For a residual stress determination in the welding too can be monitored and the Ms temperature was determined
many restrictive assumption have to be made regarding the three by the first continuously appearing ˛ -peak, which in the cur-
dimensional stress state in the welding. Therefore in the following rent case is the {2 1 1} interference line at 150 ◦ C (171.2 s)
only (elastic) strains are discussed in the longitudinal direction. Except for the {3 1 1} -peak,
which increases in intensity for a short time, from this point
3. Results and discussion onwards the ˛ -peak intensities steadily increase while the remain-
ing -peaks reduce in intensity with decreasing temperature,
Fig. 4 displays the density plots of the raw energy spectra vs. indicating a progressively martensite transformation during con-
time obtained at 5 Hz count rate for the longitudinal (det 0) and tinuous cooling down. This continuous increase indicates that the
transverse (det 1) weld direction (see Fig. 3) for the examined LTT austenite–martensite transformation is not finished until room
alloy as an example at 3 mm distance from the weld line. temperature and thus that the martensite finish temperature is
2744 J. Gibmeier et al. / Journal of Materials Processing Technology 214 (2014) 2739–2747

Fig. 5. Time and temperature evolution of the martensite (˛ ) and austenite () phase fraction in the LTT test weld at 3 mm from the weld center line in longitudinal direction.

not reached. Even though the martensite finish temperature is weld cycle (b) as a function of distance from the weld surface. The
expected above room temperature, the ongoing increase in ˛ - Ac3 -temperature is relatively constant at around 700 ◦ C in the com-
phase fraction may be attributed to local chemical variations plete welding zone and also in the transition zone to the base metal.
causing for a martensite formation at even lower temperatures The comparison of the Ms -temperature as a function of lateral dis-
(Fig. 7). tance from the weld surface reveals only a slight increase of the
Fig. 6 displays the Ms - and Ac3 -temperatures (a) as well as the value for Ms up to the depth of 4 mm from about 110 ◦ C to 150 ◦ C.
starting time of the martensite transformation after start of the Beyond that distance Ms increases strongly. As soon as the transi-
tion zone to the base metal is reached the effect of the increased
Ni-content on the transformation temperatures vanishes and the
Ms -temperatures increase to values which are more typical for low
alloyed steel alloys. Hence, the strong effect might be due to local
demixing of the filler material. Moreover, upon cooling and mar-
tensite formation a complex local stress field may arise, which can
lead to further deformation induced martensite formation. Fig. 6b
clearly shows that the martensite transformation starts at the bot-
tom of the weld at the transition to the base material. Since in this
region the local heat input is smaller and the heat conductance
through the base material is higher than in the regions closer to
the weld surface, the martensite start temperature is reached ear-
lier with a higher distance to the weld surface. The larger delay
between the transformation start at 3.5 and 4 mm is due to the
increase of the Ms temperature (Fig. 5a) that results from the mixing
of the LTT filler material with the S690 base material that contains
less nickel. Thus, the martensite start temperature is reached even
earlier in distances higher than 4 mm. An additional influence can
also be assigned to the locally varying (residual) stress states. The

Fig. 6. (a) Martensite start Ms and Ac3 temperature determined in longitudinal direc-
tion (b) begin of the martensite transformation as a function of the distance from Fig. 7. Amount of retained austenite after cooling down to room temperature as a
the weld line. function of the distance from the weld line. Results averaged from the two detectors.
J. Gibmeier et al. / Journal of Materials Processing Technology 214 (2014) 2739–2747 2745

This can also be observed in transversal direction (Fig. 8). Once


the martensite transformation starts the measured strains remain
constant in both phases. In the longitudinal direction, where the
constraint from the surrounding material is higher, a compressive
stress state develops, whereas in the transversal direction tensile
strains develop. This might indicate that the prestraining of the
austenite prior to the martensite transformation is decisive for the
strain development in the ˛ -phase. However, at the data recorded
for a distance of 3 mm to the weld surface a slight tendency of
the phasespecific strain can be observed in the further course of
the cooling down, hence at rather low cooling rates that might be
due to the mismatch in thermal expansion. In longitudinal direc-
tion the compressive strains in martensite slightly increase, while
the tensile strains in austenite show an increasing trend, which is
almost similar in transverse direction. Here, the tensile strains in
martensite show a slight decrease.
However, comparing the local strain at the different positions
of the weld, it has to be kept in mind that the temperature dur-
Fig. 8. Development of phase specific strain during cooling down in the longitudinal
ing the weld cycle varies with the considered position in the weld.
(top) and transversal (bottom) direction of the weld at 3 mm distance from the weld
surface. The maximum temperature varies from 1850 ◦ C in a 1 mm distance
from the weld line to 1170 ◦ C in a distance of 4 mm from the weld
line.
stresses due to martensite transformation and the resulting volume Hence, for better comparison of the strain development at the
expansion will also have an influence on Ms and the transformation different positions of the weld line, the strains are plotted for selec-
time and kinetics as shown by Patel and Cohen (1953), who stud- tive positions as a function of (a) the welding time and (b) the local
ied the effect of external loads on the martensite transformation temperature. Fig. 9 displays the strain in longitudinal direction in
for iron-nickel and iron-nickel-carbon alloys. Accordingly, uniaxial the ˛ -phase for different distances from the weld surface z. The
stresses will increase Ms while hydrostatic stress states will lead to plotted strain is the elastic strain resulting from the phase trans-
a decreasing Ms . formation and the constraint of the surrounding material during
Regarding the amount of retained austenite, apart from the local quenching/cooling down.
variations, a tendency can be observed which indicates a decrease During re-melting the martensite transformation starts from the
of the amount of retained austenite with an increasing distance bottom of the weld and moves successively up to the weld surface.
from the weld line. The decrease might again be due to the changes The supplementary results indicate that in larger distances from
in the chemical composition as well as to further strain induced the surface the martensite transformation starts earlier (Fig. 6b).
martensite transformation due to the transformation in the above Furthermore, the transformation starts at higher temperatures.
laying regions that transforms later. The relative strong variation (Fig. 6a). A difference of 40 ◦ C in the Ms -temperature was observed
in the data at 3 mm may be attributed to texture and grain size between the weld surface and a distance of 4 mm.
effects, which cause variations in the diffraction peak intensities. As stated before, the differences in the transformation behav-
Further, as depicted in Fig. 1 the amount of retained austenite may ior and the Ms -temperatures might be explained by the different
vary significantly between measurement points in the weld cross local heat input, the heat transfer through the base material, vary-
section due to the local microstructural inhomogeneity within the ing chemical composition and the local stress states, which result
weld line. from the cooling down in combination with the local constraint and
further from transformation induced strain.
3.2. Evolution of residual strain Moreover, the results displayed in Fig. 9 indicate that the ini-
tial strain state in martensite clearly depends on the position in the
Since the cooling phase of the welding cycle is of more weld. At a distance of 4 mm the initial strain state is a compres-
importance for the strain and stress development and for the under- sive, whereas at points closer to the weld surface tensile strains are
standing of the austenite to martensite phase transformation in the observed at the beginning of the martensite formation.
following the focus is set to the cooling part only. Fig. 8 shows The closer the examined point is to the weld surface, the more
the elastic strain development during the cooling phase for the tensile is the initial strain state, irrespective of the direction in the
austenite and the martensite in the longitudinal (Fig. 8 top) and the weld, i.e. for strain in longitudinal and in transverse direction. This
transversal (Fig. 8 bottom) direction as an example for a distance might be explained by local strain states in adjacent regions. In the
of 3 mm to the weld surface. base material the heat input is not high enough for the austeniza-
After solidification tensile strains develop almost linearly with tion of the material. Hence, the different coefficients of thermal
the continuing cooling in the austenite phase, both in longitudinal expansion (CTE) of the - and ˛ -phase (a˛ < a ) should favor a by
and in transverse direction. Right before the start of the martensite far more less tensile stress state. In the course of the material expan-
formation the tensile strains in the -phase increase more quickly, sion during the martensite transformation of the first regions, i.e.
due to the martensite formation in the underlying material, which close to the interface to the base steel plate, the austenite in zones
transforms earlier (see Fig. 6b). With start of the martensite trans- above that region has not transformed yet will be stretched. As
formation at 3 mm distance to the weld surface the tensile strains a consequence more tensile strain is determined. Thus, the strain
in the -phase decrease in the longitudinal direction and therefore state in the austenite might not only influence the stress state in the
effectively counteract the strains due to thermal shrinkage. martensite but might also shift the Ms -temperature to higher val-
At the beginning of martensite transformation the strains ues. Furthermore, transformation induced plasticity might occur,
remain relatively constant both in the - and in the ˛ -phase, which can explain the locally different strain behavior during the
indicating that the continuing martensite transformation is coun- transformation. When the transformation finished, approximately
teracting the thermal shrinkage both in the - and in the ˛ -phase. the same strain state is observed in all examined positions of the
2746 J. Gibmeier et al. / Journal of Materials Processing Technology 214 (2014) 2739–2747

Fig. 9. Strain (elastic plus thermal expansion strain) development in the ˛ -phase in longitudinal direction at different distances z from the weld top surface (left) as a function
of time and (right) as the function of the temperature.

weld. Apart of the amount of retained austenite, which varies with (DFG) is acknowledged for financial support through project GI
the location in the weld, the phasespecific strains in martensite do 376/4-1 and RE 1648/2-1.
not show a significant dependency on the local austenite content.
However, the variation in retained austenite will certainly affect the
resulting macrostress state and must be considered for calculation References
of macro (residual) stresses.
Altenkirch, J., Steuwer, A., Peel, M., Richards, D.G., Withers, P.J., 2008. The effect of
tensioning and sectioning on residual stresses in aluminum AA7749 friction stir
4. Conclusions welds. Mater. Sci. Eng. A 488, 16–24.
Altenkirch, J., Gibmeier, J., Kostov, V., Kromm, A., Kannengiesser, T., Doyle, S., 2011.
A method was presented that allows real time monitoring of Time- and temperature-resolved synchrotron X-ray diffraction: observation of
phase transformation and strain evolution in novel low temperature transfor-
the local phase transformation and strain evolution during welding mation weld filler materials. J. Strain Anal. Eng. Des. 46, 563–579.
under real welding conditions using time resolved energy disper- Altenkirch, J., Gibmeier, J., Buslaps, T., Honkimäki, V., 2012. EDXRD setup for real
sive synchrotron X-ray diffraction (EDXRD). time observation of a gas tungsten arc (GTA) welding process. Mater. Sci. Forum
706–709, 1655–1660.
Bhadeshia, H.K.D.H., 2004. Developments in martensitic and bainitic steels: role of
• Measurements at a rate of 5 Hz were carried out successfully the shape deformation. Mater. Sci. Eng. A 378, 34–39.
Chantler, C.T., 1995. Theoretical form factor, attenuation, and scattering tabulation
at ID15A@ESRF, which give a very good resolution in terms of
for Z = 1–92 from E = 1–10 eV to E = 0.4–1.0 MeV. J. Phys. Chem. Ref. Data 24,
resolved temperatures. 71–643.
• Local information about the start and the end of phase transfor- Dai, H., Francis, J.A., Stone, H.J., Bhadeshia, H.K.D.H., Withers, P.J., 2008. Character-
mations was extracted together with the transformation rate. izing phase transformations and their effects on ferritic weld residual stresses
with X-rays and neutrons. Metall. Mater. Trans. A 39, 3070–3078.
Davis, J.R., 1996. ASM Specialty Handbook: Carbon and Alloy Steels. ASM Interna-
We focused on phase transformation kinetics in a LTT weld filler tional, Materials Park, Ohio, USA.
Daymond, M.R., 2004. The determination of a continuum mechanics equivalent
material with 10%Cr and 10%Ni. The results determined for the LTT- elastic strain from the analysis of multiple diffraction peaks. J. Appl. Phys. 96,
weld can be summarized as follows: 4263–4272.
Faninger, G., Hartmann, U., 1972. Physikalische Grundlagen der quantitativen
röntgenographischen Phasenanalyse (RPA). HTM Härterei-Techn. Mitt. 27,
• the delayed martensitic transformation successfully counteracts 233–244.
the thermal strain resulting in a compressive stress state at room Gibmeier, J., Obelode, E., Altenkirch, J., Kromm, A., Kannengiesser, T., 2014. Residual
temperatures. stress in steel fusion welds joined using low transformation temperature (LTT)
filler material. Mater. Sci. Forum 768–769, 620–627.
• not only the local chemical constitution is important for Kannengiesser, T., Kromm, A., 2009. Formation of welding residual stresses
the austenite-martensite transformation, but also the local in low transformation temperature (LTT) materials. Soldagem Inspeção 14,
stress/strain field has a significant influence on the transforma- 74–81.
Kannengiesser, T., Kromm, A., Rethmeier, M., Gibmeier, J., Genzel, Ch., 2009.
tion strain and the martensite start temperature. Residual stresses and in-situ measurement of phase transformation in low
• the time-resolved and local resolved results are essential for the transformation temperature (LTT) welding materials. Adv. X-ray Anal. 52,
understanding of the interdependence of transformation induced 755–762.
Kromm, A., Kannengiesser, T., 2009. In-situ-phase analysis using synchrotron radi-
strains and materials straining that results from local constraints ation of low transformation temperature (LTT) welding material. Soldagem
due to local differing cooling rates and time-delayed cooling Inspeção 14, 82–88.
down. Kromm, A., 2011. Umwandlungsverhalten und Eigenspannungen beim Schweißen
neuartiger LTT-Zusatzwerkstoffe. BAM-Dissertationsreihe. Band 72.
Laine, E.S.U., 1978. A high-speed determination of the volume fraction of ferrite in
The data monitored in real-time during welding are valuable austenitic stainless steel by EDXRD. J. Phys. F: Met. Phys. 8, 1343–1348.
for a proper validation of simulations of LTT welds and hence for Macherauch, E., 1992. Praktikum in Werkstoffkunde: Skriptum für Inge-
nieure, Metall-Werkstoffkundler, Werkstoffwissenschaftler, Eisenhüttenleute,
an effective improvement of the existing simulation models. Fertigungs- und Umformtechniker. 10. verbesserte Auflage. Vieweg + Teubner
Verlag.
Moat, R.J., Stone, H.J., Shirzadi, A.A., Francis, J.A., Kundu, S., Mark, A.F., Bhadeshia,
Acknowledgements H.k.d.h., Karlsson, L., Withers, P.J., 2011. Design of weld fillers for mitigation of
residual stresses in ferritic and austenitic steel welds. Sci. Technol. Weld. Joining
The authors would like to thank the European Synchrotron Radi- 16, 279–284.
Ohta, A., Matsuoka, K., Nguyen, N.T., Maeda, Y., Suzuki, N., 1999a. Fatigue strength
ation Facility (ESRF) in Grenoble for granting beam time through improvement by using newly developed low transformation temperature weld-
proposal number MA847. Further, German Research Foundation ing material. Weld. World 46, 38–42.
J. Gibmeier et al. / Journal of Materials Processing Technology 214 (2014) 2739–2747 2747

Ohta, A., Suzuki, N., Maeda, Y., Hiraoka, K., Nakamura, T., 1999b. Superior fatigue Stone, H.J., Bhadeshia, H.K.D.H., Withers, P.J., 2008. In situ monitoring of weld
crack growth in newly developed weld metal. Int. J. Fatigue 21, 113–118. transformations to control weld residual stresses. Mater. Sci. Forum 571–572,
Ohta, A., Matsuoka, K., Nguyen, N.T., Maeda, Y., Suzuki, N., 2003. Fatigue strength 393–398.
improvement of lap joints of thin steel plate using low-transformation- Williams, J.C., Starke Jr., E.A., 2003. Progress in structural materials for aerospace
temperature welding wire. Weld. Res. 04, 78–83. systems. Acta Mater. 51, 5775–5799.
Patel, J.R., Cohen, M., 1953. Criterion for the action of applied stress in the martensitic Wang, W., Huo, L., Zhang, Y., Wang, D., Jing, H., 2002. New developed welding elec-
transformation. Acta Metall. 1, 531–538. trode for improving the fatigue strength of welded joints. J. Mater. Sci. Technol.
Reed, R., 1962. The spontaneous martensitic transformations in 18% Cr, 8% Ni steels. 18, 527–531.
Acta Metall. 10, 865–877. Withers, P.J., 2007. Residual stress and its role in failure. Rep. Progress Phys. 70,
Steven, W., Haynes, A.G., 1956. The temperature of formation of martensite and 2211–2264.
bainite in low alloy steels. J. Iron Steel Inst. 183, 349–359.

You might also like