Thermal Behavior in WAAM
Thermal Behavior in WAAM
https://doi.org/10.1007/s11665-024-10441-6 1059-9495/$19.00
Wire arc additive manufacturing (WAAM) is a directed energy metal additive manufacturing technology
that features high-speed production of large and dense parts. The process is exposed to thermal shocks due
to a high heat input. These thermal shocks may lead to various issues, including distortions, residual
stresses, surface morphology defects, and irregular macro- and microstructures. This study proposes
adopting infrared heaters in the WAAM process. Thus, it presents a preliminary investigation to promote
further studies of the effects of infrared heating on the WAAM process and the quality of the produced
parts. An experimental study was carried out to observe the effects of infrared heaters on part quality in
WAAM. Two walls were built using ER70S-6 welding wire to compare the conventional WAAM process
with the WAAM process that used an infrared heater. The results showed that the wall produced by WAAM
with infrared heaters had more regular geometries, refined internal structures, and stable mechanical
properties. The optic and SEM images showed that this method reduced the acicular and polygonal ferrite
structures. Notably, bainite formation was not observed, and there was a decrease in the formation of
defects, such as pores, lack of fusion, and interlayer discontinuities. The specimen subjected to the external
infrared heater exhibited approximately 17% higher tensile strength than the standard WAAM specimen.
Moreover, the WAAM with infrared heater procedure reduced the microhardness variations along the
build direction by 12% compared to the 30% reduction achieved by the conventional WAAM process.
and is especially a disadvantage of high deposition rates. (Ref A more active approach for thermal control is to use a
19). Zhao et al. (Ref 20), Yang et al. (Ref 21), and Wang et al. cooled platform, as explained by Lu et al. (Ref 26). This
(Ref 22) indicated that when the component height rises, heat method, whether used with small preforms or materials with
accumulation intensifies. In such cases, excess heat can cause high thermal diffusivity, has been demonstrated to be effective
distortion, high residual stresses, geometric irregularities, only for the first layer. Da Silva et al. (Ref 27) used close-
excessive oxidation, and coarse microstructure. Therefore, immersion active cooling (NIAC) to control heat during part
different approaches have been investigated to reduce heat production. In this method, metal layers were deposited into the
build-up and to control heat from production. For example, tank with gradually rising water. The NIAC technique was
Alhakeem et al. (Ref 23) developed an algorithm that generates found to be efficient in reducing heat build-up in the WAAM of
deposition paths based on thermal variance analysis. Thus, they aluminum, and the author claimed that the wall width remained
controlled the process during production and reduced the almost constant, with the preforms becoming thin and long. Wu
distortion in part. According to their findings, the suggested et al. (Ref 28) used forced interlayer cooling with compressed
method decreased thin-walled overflow by 21% and surface CO2 on the upper surface of a Ti-6Al-4V flat-wall geometry
deformation by 37%. (150 mm height). They indicated that WAAM production with
One of the most widespread methods for controlling heat is CO2 gas interlayer cooling dramatically reduced dwell time
the inter-pass cooling time approach. Yang et al. (Ref 21) between deposited layers and significantly increased production
investigated the thermal behavior during the production of the efficiency. Moreover, they expressed that forced interlayer
WAAM part with infrared thermography and indicated that as cooling using compressed CO2 gas can be easily applied and
the production height increased, the heat accumulation useful for additively produced Ti-6Al-4V components and
increased drastically. By extending the interlayer cooling time, provides an attractive surface finish with less visible surface
the maximum and average temperatures of the thin-walled parts oxidation, a refined microstructure, improved hardness, and
decreased. By contrast, the temperature gradient and cooling increased strength. Rodrigues et al. (Ref 29) offered a novel
rates increased. However, the authors asserted that the exag- approach named ultracold wire and arc additive manufacturing
gerated interlayer cooling time could not effectively reduce the (UC-WAAM). They compared this WAAM-based method with
heat build-up of the deposited parts, making the deposition a traditional GWAM. By removing the electric current flowing
process time-consuming. Lei et al. (Ref 24) built finite element through the substrate and rerouting the electric arc from the
models to simulate the role of interlayer dwell time in thermal melt pool, this novel WAAM version lowers process temper-
behavior during gas metal arc welding (GMAW)-based additive atures and boosts cooling rates without affecting the integrity of
manufacturing and conducted an experiment to validate the the produced parts or the deposition rate. It also has strong arc
modelsÕ validity. They showed that as the interlayer residence stability and avoids spatter production. Therefore, it is a useful
time increased, the cooling rate of the melt pool decreased in substitute for arc-based additive manufacturing. The layers are
the first layer and then increased in the fifth and tenth layers. also taller and thinner because of faster solidification.
Denlinger et al. (Ref 25) observed a reduction in residual stress Li et al. (Ref 30) took the interesting approach of heating the
and distortion levels with increasing dwell times from 0 to 40 s wire, delivering the hot wire arc additive manufacturing
for Inconel 625. However, the dwell time had the opposite (HWAAM) method. Here, the wire was heated with resistance
effect for Ti-6Al-4V, with shortened or even zero dwell time heating before entering the weld pool. The morphology of the
reducing distortion. The authors further stated that shorter dwell specimens underwent major changes with an increase in the hot
times provided parts with significantly lower residual stress. wire current. The width of the thin walls ranged from 8.5 to
Fig. 2 Illustration of the external heat model. The part is surrounded by heaters to minimize thermal shocks
duction. AISI 1045 material was chosen as a substrate for tanks, shafts, car bodies, pipes, steel castings, and construction
reasons such as easy accessibility, compatibility with the work (Ref 47, 48). Table 3 shows the standard properties of the
material produced, and its welding properties. base material (substrate) and the deposited material. Whereas
As the deposited material (AWS A5.18 ER70S-6), 1.2-mm- the mechanical properties of the deposited material were
diameter copper-plated mild steel wire was used. This material superior, the density of the two materials was nearly the same.
is used in many areas, such as general workshop fabrication,
The chemical composition of the welding wire is shown in 3.3 Mechanical and Metallurgical Tests of Parts
Table 4.
The same mechanical and metallurgical tests were applied to
the walls. First, the wallsÕ dimensions were measured. Second,
3.2 The Infrared Heater System Setup
they were removed from the substrate. Third, to obtain samples
The deposited materials, substrates, and production param- from the walls, they were cut in the directions shown in Fig. 6
eters were the same as in the conventional WAAM setup in using a liquid-cooled abrasive cutting machine with the
section 3.1. In this part of the study, in addition to the systems discharge technique. Four samples were taken from each wall
given in WAAM, an infrared heating system was used. The for more consistent results.
system consists of three different subsystems. 3.3.1 Microhardness and Tensile Tests. Microhardness
measurements were made with an HWMMT-X3B HIGH-
• Infrared heaters: The system consisted of two infrared WOOD microhardness device. Measurements were done at 500
ceramic heaters. The heaters had a wavelength range of gf, 15 s dwell time, and 100-lm intervals in the sub-, middle-,
2–10 lm and a power of 400 W. They were used to pro- top-, and top-two-layer regions of the samples.
vide heat to the process. Infrared heaters were chosen be- A tensile test was carried out with a Shimadzu brand AGS-X
cause of their technological advantages, such as high heat model device. Tensile test specimens were prepared according
transfer capabilities, heating only the closest point directly to the ASTM E8m-04 standard with a thickness of 7 mm, a
seen, and not damaging other production equipment. gauge length of 32 mm, and a total length of 112 mm. Tensile
• Control unit: This system consisted of one controller used tests were performed according to the DIN EN ISO 6892-1
to control ceramic infrared heaters. standard at room temperature at a tensile speed of 1 mm/min.
• Monitoring system: The system consisted of thermocou- 3.3.2 Macrostructure and Microstructure Analy-
ples used to monitor temperature. In this study, a K-type sis. The etching process was performed by swabbing with
thermocouple was used with an accuracy of ± 2.2 C. 4% nital reagent applied to the samples for 25–30 s to achieve
macro-etching and for 55–60 s for micro-etching. Macrostruc-
In this study, two infrared heaters were placed facing the long tures of the upper and lower regions of the walls produced by
sides of the wall (Fig. 5). The distance between the heaters and WAAM and WAAM with infrared heater were examined with a
the part was 5 cm. The heaters were connected to a temperature stereo-zoom microscope. The microstructure of the upper and
control device. The system temperature was achieved with a lower regions of the walls was examined using a light optical
thermocouple connected to the controller. The thermocouple microscope (OM) (Leica, CTR6000) and a scanning electron
was on the substrate and between the workpiece and the microscope (FEI, Quanta FEG 450). ImageJ software was used
infrared heater. to analyze the images. The chemical elements were verified
The infrared heating system was activated before production using energy distribution spectroscopy (EDS). The resulting
and remained active throughout the process. A thermal couple microstructure properties and grain size were correlated with
was placed between the wall and the heater ceramic plate, as the thermal behavior and mechanical properties of the walls
shown in Fig. 5. Thus, the heaters were monitored and produced by WAAM and WAAM with infrared heaters.
controlled via this thermocouple. The maximum safety value 3.3.2.1 Electron Backscatter Diffraction Analysis. The
for the thermocouple was set at 150 C. At the beginning of the microstructure and crystallographic texture of walls produced
production, the substrate temperature was measured as 110 C by WAAM and WAAM with infrared heaters were further
with a laser infrared temperature meter. When the part examined using EBSD characterization, providing helpful
production ended, the thermocouple controlling the heaters insights into their composition. Given the high sensitivity of
was set to 100 C, and the heaters were turned off after running EBSD patterns to surface defects (Ref 49), meticulous sample
for another 10 min. The heaters and the wall were then allowed preparation was conducted to ensure that the examined surfaces
to cool. were completely devoid of visible scratches or contamination,
a b c
Wall 200 7 38
Substrate 300 150 30
as verified under an optical microscope (Ref 50). EBSD statistical description of the microstructure, the EBSD mea-
analyses were performed using a Zeiss Gemin 2 scanning surements involved sampling large areas of 320 lm 9 240 lm
electron microscope integrated with a Hikari EBSD camera and to capture enough grains. The resulting EBSD data were then
EDAX/TSL EBSD system. To provide a comprehensive analyzed using TSM OIM Analysis v7.2.
wt.% 0.06-0.15 1.40-1.85 0.0035 max 0.15 max 0.03 max 0.15 0.50 0.80-1.15 0.025 0.15 max
Fig. 7 Temperature measurement point: (a) WAAM, (b) WAAM with infrared heater
though the inter-pass cooling times were equal for both cases, described, rapid cooling caused the final layers of the part to
no overflow was observed. deviate from the center relative to the build direction. This
In Fig. 10, the comparison between the manufactured deviation is likely attributed to the rapidly changing surface
walls and the CAD data includes an assessment of the temperature and grain growth in the core, which the cooling
stability and accuracy of the build direction. As previously process could not adequately accommodate. In this context,
we compared the wall samples produced by WAAM with the greater stability in the build direction and exhibited less
infrared heater to the CAD data. The comparison revealed deviation from the CAD data. This finding suggests that the
that the walls produced with the infrared heater demonstrated use of an infrared heater aids in maintaining the desired
geometry and can mitigate distortions typically induced by of the nearby regions. Additionally, in Fig. 12(d), a magnified
rapid cooling. image of grain size changes taken from different regions of the
sample is presented. Lastly, Fig. 11(e) shows another image of
4.3 Macrostructure and Microstructure Analysis internal structure changes due to uncontrolled thermal behavior.
It should be noted that the microstructural features men-
Figure 11 shows the macrostructures of the walls produced
tioned above were observed in all layers of the wall along the
by WAAM and WAAM with infrared heaters. The deposited
build direction. Therefore, a degree of inhomogeneity was
layers were easily distinguished. The beads appeared to have
detected in each bead deposited from the center of the melt pool
fused neatly. There were no visible defects in the weld zones.
toward its boundary. It is well known that in metal additive
Consequently, the macrostructures of the two walls appeared
manufacturing processes, the presence of multiple and complex
similar. Although the macrostructures were similar, the cross-
thermal cycles is the primary factor leading to the non-
sectional images of the walls revealed differences in the internal
uniformity of microstructure along the build direction (Ref 51,
structures. The deformations seen in the WAAM wall are the
52). During the manufacturing process, repeating heating and
natural product of uncontrolled cooling. These are defects
cooling cycles for each new layer introduces significant thermal
caused by the core region that cannot accompany rapid cooling.
gradients within the material. These thermal cycles cause
Rapid cooling can cause undesirable stresses to be generated in
thermal stresses and gradients across the HAZ, the melt pool
the part; as a result, some parts of the part may shrink. These
boundary, and the melt pool center of the previously deposited
shrinkages result in distortion of the part geometry.
layer. As shown in Fig. 12(b), these thermal cycles caused more
Figure 12 clearly shows the negative effect of thermal
pronounced segregation of these zones. The observable effect
instability on grain size homogeneity. Figure 12(a), (b) depicts
on grain size distribution was particularly striking: repeated
the two consecutively deposited beads in the build direction,
heating and cooling cycles led to a marked change in grain size
with three recognizable regions with different microstructures:
along the affected regions. This phenomenon not only affects
melt pool center, boundary, and heat-affected zone (HAZ).
the microstructural properties but also has a considerable
Every new layer deposited caused grain growth in the HAZ
influence on the mechanical properties of the material. Fluc-
region of the previous layer. Further, a periodic non-uniform
tuation in grain size distribution in these regions can lead to
microstructure with various interlayer discontinuities, such as
changes in material strength, ductility, and fracture toughness,
pores and lack of fusion, was seen. Here, the defects were
affecting the overall mechanical behavior of the WAAM
mainly along the interlayer regions perpendicular to the build
component.
direction. Figure 12(c) shows different grain structure regions
Controlling the heat via the WAAM with infrared heater The microstructure images of the final layers clearly show
system resulted in a more regular internal structure. Figure 13 the effect of cooling on the microstructure. Figure 14(a) shows
shows less variation in grain size in the build direction. Since the microstructure of the top two layers of the wall produced by
the thermal fluctuation was reduced, the effect of the cooling– WAAM. The classical dendritic structure appears. Rafieazad
heating cycle on grain size was more limited. Contrary to the et al. (Ref 53) emphasize that rapid cooling rates near the melt
WAAM wall, structural defects, such as interlayer errors, lack pool boundaries can lead to martensite formation, especially in
of fusion, and porosity, were almost absent in the wall produced regions with higher cooling rates, but fine acicular ferrite and
using WAAM with infrared heaters. Overall, controlling bainite are also observed. This martensitic structure is known to
thermal behavior contributes to structural integrity by mini- be a product of rapid cooling. On the contrary, in Figure 14(b),
mizing differences in microstructure at the lower to the upper ferrite and pearlite structures were formed for the WAAM with
regions. The use of an external heater in the WAAM process infrared heater wall. Further, the resulting microstructure
prevented uncontrolled cooling of the part during layer pattern was more homogeneous throughout the layers. The
production, dwell time, and cooling time at the end of effect of thermal behavior on the microstructure was apparent
production. Consequently, all layers experienced approximately in the whole piece. To further analyze this outcome, the average
consistent heating and cooling cycles, resulting in a uniform grain sizes in the two pieces were measured using the ImageJ
microstructure along the build direction. The addition of an open-source program. Here, the average grain size of the
external heater to the WAAM process was an intervention to the WAAM wall was 10.8 lm, which is consistent with the
conventional heating and cooling regime typically associated literature (Ref 54). In the wall produced by WAAM with
with the deposition of each new layer. As shown in Fig. 13(b), infrared heaters, we measured 12.62 lm. It is known that
this change was manifested by the reduction in the apparent higher cooling rates lead to a decrease in ferrite grain size (Ref
boundary between critical regions, such as the HAZ, the 55). Moreover, hard fine phases with dislocations form because
boundary of the melt pool, and the central region of the melt they suppress atomic diffusion. By contrast, slow cooling rates
pool from the previous layer. This phenomenon affected the cause phases such as polygonal ferrite to transition into soft,
grain structure and mechanical properties of the components. coarse, and less dislocated states. The ultimate mechanical
The blurring observed at the boundaries is indicative of an characteristics are influenced by the size and percentage
altered thermal history to which the material was exposed distribution of ferrite and pearlite inside the microstructure
during deposition. This thermal history may make it possible to (Ref 56, 57).
interfere with material properties. In addition to the optical microscopy investigation, SEM
and EDS analyses (Fig. 15 and 16) were performed for detailed
microstructural characterization of the walls. The EDS results higher hardness and toughness than ferrite but lower than
revealed that the chemical composition of the walls was pearlite (Ref 60).
consistent with the chemical composition of the welding wire. Figure 16(b1) shows the ferrite structure, pearlite formed at
The micrograph in Fig. 15 confirms the presence of pearlite in the grain boundaries, and a few defects. Perlite and defects
the ferrite matrix. The white areas are ferrite, and the dark areas formed at the grain boundaries were lower compared to the
are pearlite formed at the grain boundaries. Figure 15b1 shows WAAM wall. The infrared heat sources seemed to have reduced
the ferrite structure, the perlite formed at the grain boundaries, the formation of polygonal and acicular ferrites in the internal
and the incomplete and flawed internal structures. Polygonal structure. Thus, infrared heaters contributed to the homogeneity
and acicular ferrite structures formed due to high cooling are of the internal structure by causing the growth of ferrite grains,
shown in Fig. 15(b2). Further, bainite formation was observed cleaner grain boundaries, and even grain distribution. The size
(Ref 47). The formation of bainite is essential evidence that and distribution of ferrite are shown in Fig. 16(b) and b2.
shows uncontrolled cooling, as the formation of bainite is Notably, bainite formation was not observed in the structure.
affected by the cooling rate. As indicated by the continuous
cooling transformation (CCT) diagram produced using ther- 4.4 EBSD Analysis
modynamic simulation for the ER70S-6 material, moderate
To supplement the SEM analysis, an EBSD analysis was
cooling rates ranging from 0.1 to 1 C/s result in a ferritic–
conducted to gain additional insights into the crystallographic
pearlitic microstructure, while severe cooling rates in the field
texture and orientation characteristics of the walls. Image
of 10–100 C/s lead to non-equilibrium phases such as bainite
quality (IQ) map imaging was used to visualize the EBSD
(Ref 58, 59). Bainite is a microstructural phase that forms at
patterns and microstructural morphologies obtained from
lower temperatures than pearlite and ferrite. It typically has
different points on the specimen surface during EBSD analysis. coarse grain region with an average grain size of
Figure 17 and 18 shows typical IQ map images taken for the 11.41 ± 2.16 lm. In contrast, the region inside the melt pool
samples from WAAM and WAAM with infrared heaters and the exhibited a significantly smaller grain size of 8.72 ± 1.24 lm.
corresponding inverse pole figure (IPF) map, where colors Considering the inherent layer-by-layer deposition process
indicate crystal orientations. involved in the WAAM method, each layer deposited reheats
Figure 17 presents EBSD images from the sample identified the preceding layer. Consequently, the previously solidified
as zone 3 in Fig. 6(a). Images were taken from the middle part layer experienced elevated temperatures, leading to grain
of the specimen in the interlayer zone. Randomly oriented boundary migration and subsequent grain growth near each
coaxial grains with an average grain size of 10.8 lm were melt pool boundary. These microstructural irregularities within
seen. Thus, the microstructure in Fig. 10, 11, 12, and 15 the produced componentÕs structure have negative effects on its
appears to be in good agreement with the OM and SEM mechanical properties and contribute to the development of
observations. The grain coarsening in the HAZ of the anisotropic mechanical behavior (Ref 61).
previously deposited layer, as shown in Fig. 11, was promi- Figure 18 displays EBSD images of the wall produced by
nently evident in the IPF map, indicated by the presence of a the WAAM with infrared heaters corresponding to Zone 3 in
Fig. 6(a) The images were captured from the central region of different heights and cross sections for the walls produced by
the specimen, precisely at the interlayer zone. The average WAAM and WAAM with infrared heaters and the average
grain size of the WAAM with the infrared heater sample was deviation values in the measurement region. The mean
12.62 lm along the wall, which is consistent with the OM microhardness values of the WAAM wall were 182 ± 10 HV
images. In contrast with the WAAM sample, the average grain for the sub-region, 160 ± 5 HV for the mid-region,
size had the same values along the production direction. Thus, 180 ± 6 HV for the top-region, and 200 ± 10 HV for the
it was complex to identify regions such as the HAZ and melting top two layers; these results are consistent with the literature
pool in the sample. The external heaters reduced the temper- (Ref 6, 47, 54). Although the same deposition parameters were
ature difference between the part surface and the core region, consistently applied, the local thermal gradients inherent in the
providing the time necessary for the grains to grow. This WAAM process can result in variable cooling rates across
observation indicates that external heat reduces grain size different regions of the sample. Variations in cooling rates can
variability, thus indicating that anisotropy can be reduced with lead to significant differences in the microstructure, particularly
this approach. in grain size and phase distribution, which in turn may affect
Figure 19 presents the phase maps of specimens from the microhardness of the material. For the WAAM with infrared
WAAM and WAAM with infrared heaters. While both heater case, the mean microhardness values were 166 ± 3 HV
materials were composed mainly of the ferrite phase, some for the sub-region, 162 ± 4 HV for the mid-region,
austenite retained during solidification was distributed mainly 166 ± 3 HV for the top-region, and 175 ± 3 HV for the top
along the ferrite grain boundaries. Alpha represents the ferrite two layers. The decrease in hardness is associated with the size
phase (BCC), and gamma represents the austenite phase (FCC) and amount of ferrite grains in the structure (Ref 63). In this
(Ref 62). Since austenite is a harder structure than ferrite, it context, the microhardness measurements were consistent with
increased the hardness of the WAAM sample. The external the microstructural differences in the SEM and microstructure
heating caused the austenite structure to transform into ferrite. images.
Tensile tests were performed on specimens produced using
4.5 Microhardness and Tensile Test Analysis WAAM and WAAM with an infrared heater. The tensile
specimens before and after testing are shown in Fig. 21. The
The microhardness of four different wall sections (see
specimens were extracted from the lower regions of the walls
Fig. 6(a)) at four different heights (see Fig. 6(b)) along the build
using waterjet technology. The specimens from the same
direction was measured. The first two layers have been
regions produced by WAAM and WAAM with an infrared
removed, and the remaining layers are labeled: approximately,
heater exhibited different fracture zones. As shown in Fig. 12,
layers 3 to 7 are called the lower zone, layers 8 to 12 the middle
changing the grain structure in the direction of the structure also
zone, layers 13 to 18 the upper layer, and the top two layers 18
altered the fracture dynamics. When comparing the WAAM
to 20. Figure 20 shows the average microhardness values at
specimen to the infrared-heated WAAM specimen, a fracture
occurred in a region different from the specimen center. The tensile test results for the WAAM and the WAAM with
Additionally, the external heater caused a fracture to occur at infrared heaters revealed significant differences in their
the specimen center. mechanical properties. The WAAM with infrared heaters
Both a control specimen and a test specimen comparatively exhibited a higher maximum elongation force (11,428 N
analyzed. The data obtained from the tensile tests are presented compared to 9742 N) and maximum elongation stress
in Table 5, and the stress–strain graph is shown in Fig. 22. (253 N/mm2 compared to 216 N/mm2), indicating that the
material can withstand greater loads before deforming. Addi- maximum force (19,448 N compared to 16,687 N) and
tionally, both the maximum elongation (135 mm compared to maximum stress (432 N/mm2 compared to 370 N/mm2) were
133 mm) and percentage elongation (42% compared to 41%) significantly higher in the infrared heater specimen.
were slightly higher for the infrared heater specimen, suggest- The tensile test results shown in the stress–strain curve
ing enhanced ductility. The fracture force (11,747 N compared (Fig. 22) compare the mechanical properties of WAAM and
to 10,117 N) and fracture stress (261 N/mm2 compared to 224 WAAM with infrared heaters. The infrared heater specimen
N/mm2) are also higher for the infrared heater specimen, shows superior tensile properties as evidenced by the higher
indicating better resistance to breaking under stress. Further- maximum stress of 432.18 MPa compared to 370.83 MPa for
more, fracture elongation (135 mm compared to 133 mm) was the standard WAAM specimen. This significant difference
marginally greater in the infrared heater specimen. The indicates that the material treated with the infrared heater can
Fig. 18 EBSD analysis image for WAAM with infrared heaters: (a) IQ map, (b) IPF map and details
withstand larger tensile loads before failure. Also, the overall concentrations and microcracks, which can weaken the mate-
shape of the stress–strain curves indicates improved ductility rial. This results in improved fracture resistance and overall
for the infrared heater specimen as it maintains higher stress mechanical performance. The infrared heater also promotes the
levels over a wider stress range. formation of equiaxed grains rather than columnar grains,
The application of an infrared heater likely results in a more which are generally associated with better mechanical proper-
refined and homogeneous microstructure by reducing thermal ties, including higher tensile strength and better elongation.
gradients. This refined microstructure enhances both the Furthermore, the additional heat provided by the infrared heater
strength and ductility of the material. Also, infrared heating enhances atomic diffusion during the deposition process,
can alleviate residual stresses that typically form during rapid leading to better bonding between layers and reduced porosity,
cooling in the WAAM process, reducing the likelihood of stress which contributes to the material’s overall strength and
Fig. 20 Average VickerÕs microhardness for the walls produced by WAAM and WAAM with infrared heaters
toughness. Finally, the use of an infrared heater ensures a more 4.6 XRD Analysis
uniform temperature distribution across the specimen during
The XRD model of ER70S-6 walls in region 2 (given in
the WAAM process, minimizing the formation of thermal
Figure 6a) measurements was made for phase identification
hotspots and cold spots and resulting in more consistent and
from the middle zones along the build direction. The obtained
predictable material behavior under tensile loads, as shown in
spectra are given in Fig. 23 for walls produced by WAAM and
Table 6.
WAAM with infrared heaters. As can be seen, the WAAM wall
Fig. 22 Comparison of stress–strain curves for WAAM and WAAM with infrared heaters
Table 6 Tensile test results for WAAM and WAAM with infrared heaters
Parameter WAAM WAAM with ınfrared heater
was composed of a-Iron (BCC, ferrite) at 2h at 45.66, 65.91, Here, we observed that the intensity obtained for the WAAM
and 83.15, according to the JCPDS 98-000-9982 patterns (Ref case was lower and that there was a
3, 53, 64, 65). The WAAM with infrared heater wall was Luo et al. (Ref 66) found that the induced thermal stress in
45.48, 65.67, and 82.93, respectively, for the same pattern. the sample was attributable to the higher temperature and larger
pool size. They considered this condition a higher energy and more consistent microhardness values. The findings can be
density resulting from grain coarsening. Thus, this was likely to summarized as follows:
result in a change in the diffraction peak intensity. Conse-
quently, there might be a reduction in the full width at half the • The wall structure produced by WAAM with infrared hea-
maximum. The authors further stated that the diffraction pattern ters had fewer flaws.
shift to the right indicates that the crystal lattice parameter • Optical images showed that the infrared heaters changed
becomes smaller. In addition, they showed that, according to the wall microstructure. For example, a reduction in the
BragÕs law (2dsinh = nk), the right shift of the diffraction acicular and polygonal ferrite structures was observed. In
pattern causes the lattice parameter of the crystal to shrink when contrast with WAAM, no bainite formation was observed
the wavelength k remains constant. They supposed that the in the SEM images. The infrared heaters appeared to re-
shrinkage of the cage parameters may be due to residual duce the formation of irregular microstructures with vari-
compressive stress inside the imprinted 316L stainless steels. ous interlayer discontinuities, such as pores and a lack of
We observed a similar right shift in the diffraction pattern in the fusion.
WAAM case. This shift to the right in the WAAM fragment • The dendritic structure formed in the top two layers was
indicates that the defects in the crystal structure were excessive. decomposed by the infrared heaters, giving the structure a
Thus, the defects in the crystal structure produced dislocations. relatively more homogenous appearance.
Optical and SEM images and hardness measurements agree • EBSD analyses showed that anisotropy could be reduced
with the XRD results and Lou et al.Õs interpretation of residual by infrared heaters.
stress. • The microhardness values of the wall produced using
WAAM with infrared heaters were significantly lower.
The difference between the top two layers was approxi-
mately 35 HV. The measured microhardness variation
5. Conclusion along the structure direction decreased from 30 to 12%
compared to WAAM.
In this study, the effects of infrared heaters on parts • The specimen subjected to the external infrared heater
produced by WAAM were investigated. This study aimed to exhibited a higher tensile strength value compared to the
control the thermal behavior of the WAAM process. Infrared standard WAAM specimen. The increase in tensile
heaters provide additional heat that helps maintain a more strength was calculated to be approximately 17%.
homogeneous temperature distribution between layers during • The XRD results supported the other findings. For exam-
deposition and dwell time. This mechanism addresses some of ple, the theta (2h) angle shifted to the right in the case of
the defects found in conventional WAAM, such as rapid the WAAM wall, which might indicate more dislocations.
cooling–heating and thermal gradients, which can lead to
residual stresses, distortions, and uneven microstructural fea- Considering these findings, the use of infrared heaters seems
tures. We experimented to analyze and compare two walls promising in terms of reducing some defects in WAAM.
fabricated WAAM with and without infrared heaters. The Moreover, this approach may even allow for microstructure
results showed that the use of infrared heaters resulted in better tailoring. Further studies are required to develop this novel
stability in the build direction, less deviation from CAD data, method.