Thermomechanical Modeling of the WAAM Process: Comparison
Please note this is a comparison between Version 2 by Jessie Wu and Version 1 by Fakada Dabalo Gurmesa.

Wire arc additive manufacturing (WAAM) is a type of additive manufacturing (AM) process that uses an electric arc welding technique with a wire feeding process and provides a high deposition rate. Previous investigations showed that the method is carried out using a layer-by-layer approach to build the part, and the corresponding heat input influences the mechanical properties of WAAM products. 

  • distortions
  • thermal
  • WAAM

1. Thermal Modeling

Thermal modeling in the wire arc additive manufacturing (WAAM) process is important to specify an optimum process parameter with geometrical consistency for the desired production by moderating the heat inputs. RSesidual stresses (RS) and distortions are considered to be the major obstacles against the more widespread application of WAAM. In particular, since WAAM involves significant amounts of heat input within the printed part, thermal management is the main important action needed to improve the quality of the part in cases of surface finish and induced internal voids [39,40][1][2]. For these defects, thermal modeling can provide the basis for the thermal properties of the RS remaining in the products. As the moving localized heat source causes steep temperature gradients, which are inevitable in this process, accurate prediction of the thermally induced RS and distortions is of paramount importance. The properties of the thermally affected build-up part are due to the layer-by-layer molten pool deposits and process conditions such as the deposition patterns, energy input, and heat conduction during the printing process [18,41][3][4].
Wire and arc-directed energy deposition (WADED) is among the metallic AM technologies that are most important in monitoring the temperature evolution since it directly affects the deposited quality of parts. The history of temperature in WADED can be attained using numerical simulations and/or experiments [42][5]. The thermal cycle in the WAAM was applied equally as a thermal load for each bead according to the welding process. The heat source parameters gained as a result are used for carrying out thermal and mechanical analysis. The influence of the heat input on the grain size, tensile properties, hardness, and impact toughness along the building direction are studied [29[6][7],43], which reveal superior tensile properties and hardness. However, variation in the mechanical properties are due to grain sizes and microstructural evolution evolved with the level of heat input. The heat input of the WAAM process arc melts the wire into the molten pool to make the parts. The heat loss modes in such processes are discussed and illustrated in several studies, such as [20,44,45][8][9][10]. The adjacent tool-path concentration makes fragments in the thermomechanical properties of continuous tool-paths that are likely to be poor, bringing about a degradation of performance and other defects [46][11].
Based on the heat input parameters, mathematical modeling [47,48][12][13] and management concepts are related to RS, tension, shrinkage, and deformation, which are critical and particularly covered in [49,50][14][15]. In such mathematical modeling works, a general equation of energy balance based on the First Law of Thermodynamics is often used, which is given in Equation (1):
Q L = Q C + Q C v + Q R  
where QL��  is the quantity of heat loss, QC��  is the conduction loss, QCv���  is the convection loss and QR��  is the radiation loss. The heat losses by radiation and convection from the surfaces of the substrate and deposit are applied as boundary conditions [9][16]. As reported in [10[17][18],51], the transient temperature fields over the part geometry in all directions were calculated from the 3D heat conduction equation, which is given in Equation (2). In this formulation, the thermal model was sequentially coupled to the mechanical model. Regarding the heat source model in the AM process, the heat transfer equation of the arc in the material and transient temperature fields is given by the Fourier equation for a transient and non-linear system, which is the heat-conduction equation during the process of deposition [10[17][18][19],14,51], as shown in Equation (2).
λ ( 2 T x 2 ) + λ ( 2 T y 2 ) + λ + q v = ρ c ( T t )
where T is the temperature, λ is the material’s thermal conductivity, qv�� v is the heat source or sink, ρ is the material density, c is the specific heat capacity of the materials, t is the time of heat transfer, and x, y, and z are coordinates in the space coordinate system.
The heat source model of a dual ellipsoid was used to simulate the heat input in the AM process [52][20] as shown in Figure 1a. The movement of the heat source simulation and the temperature change position are also shown in Figure 1b [11][21].
Figure 1. Illustration of (a) parameters of heat source volume with double ellipsoidal shape and (b) heat source movements and activation approach of mesh elements in the simulations (Reprinted from ref. [11][21]. Copyright 2019, Elsevier, in accordance with CC BY license, open access).
The front and rear halves of the heat flux densities are further discussed in the relevant literature [30,53,54[22][23][24][25],55], and the formulas are presented in Equations (3) and (4), respectively.
q f ( x , y , z ) = 6 3 f f Q π a f b c π e x p [ 3 ( x 2 a f 2 + y 2 b 2 + z 2 c 2 ) ]
q r ( x , y , z ) = 6 3 f r Q π a r b c π e x p [ 3 ( x 2 a r 2 + y 2 b 2 + z 2 c 2 ) ]
ff+ fr=2��+ ��=2
 
where af��  and ar��  are the lengths of the frontal ellipsoid and rear ellipsoid, respectively; b is the heat source width; c is the depth of the heat source, which is the heat source parameter; Q is the efficiency factor of energy input; and fr��  and ff��  are the distributing power factors for the rear and front heat source, respectively. The high-density electron beam heat source was modeled as a conical volumetric heat source with a Gaussian distribution or double ellipsoid volumetric heat source model, as given in Equation (5). The intensity distribution profile can be adopted and modified from the mathematical modeling [10,33][17][26].
Q=2h(1zh)3×η×Va×Ibπ×2eexp(2(x2+y2)2e)�=2ℎ(1−�ℎ)3×�×��×���×∅�2���(−2(�2+�2)∅�2) Q = 2 h ( 1 z h ) 3 × η × V a × I b π × e 2 e x p ( 2 ( x 2 + y 2 ) e 2 )
where Q is the power generated per unit volume, h is the penetration depth, η is the heat source efficiency, Va��  is the accelerating voltage, Ib��  is the beam current, e∅�  is the beam diameter, and x, y, and z are the local coordinates of a point on the part geometry.
Convective heat losses on the surface caused by the forced mechanism are ignored since the EBAM process is carried out in a vacuum. Radiation heat losses were considered from all outer free surfaces according to the Stefan–Boltzmann law as shown in Equation (6):
qrad=εσ(T4srfT4) ����=��(����4−�∞4)   q r a d = ε σ ( T s r f 4 T 4 )
where ε is the emissivity, σ is the Stefan–Boltzmann constant, Tsrf����  is the surface temperature of the part, and T�∞  is the ambient temperature (298 K).
Wu et al. [56][27] investigated the arc stability, bead formation, and metal transfer mechanisms during the fabrication of Ti6Al4V. From heat accumulation by gas tungsten WAAM, the effects of localized gas shielding and heat accumulation indicate the deposition constancy during the WAAM process control and component optimization. Others studied the indication of comprehensive thermal characteristics of a WAAM process and pointed out the thermal effects on the geometrical accuracy, process stability, and deposited part properties. The deposited beads’ temperature profile and the solidification parameters are examined using experimental and numerical simulation models that reflect the deposition process characteristics [22,57,58,59][28][29][30][31]. Inconsistent layer cross-sectional dimensions and undulated surface appearance are two types of defects observed in WAAM products. These are due to an inadequate heat input and nonlinear time-varying boundary conditions of the thermal effect in the molten pool [60][32]. Fluid flow and heat transfer models are usable to join process variables and parameters that affect the properties of printed parts. Both the arc pressure and the convective flow are the two main factors that govern the depth and width of penetration, respectively [61][33].
The summary of arc welding types and thermal distribution effects on the WAAM components is given in Table 1.
A high-temperature gradient near the melt pool leads to undesirable product dimensional distortion and deformation, which is the outcome of rapid thermal cycling. When the product is exposed to excessive strain due to thermal strain, it is more susceptible to fracture, which decreases the component life expectancy and increases the risk of early component failure [18][3]. The direction of deposition influences the residual strain and stress during the welding process of a single-pass multilayer [29,65][6][40]. The physical properties of the components made using WAAM depend on heat input during the printing process [48][13]. Similarly, physical properties such as the thermal conductivity and specific heat conductivity were increased with temperature while the density, viscosity, Young’s modulus, and yield strength decrease with increased temperature [66][41]. Moreover, thermomechanical simulation to predict the thermal flow and numerical analysis for metal addition during the deposition process are performed. The numerical analysis is validated with experimental investigation results by considering the effects of welding parameters such as the arc current, arc voltage, wire speed, welding speed, wire diameter, shielding gas, heat sources, and temperature fields, as studied and reported in [17,31,67][39][42][43].
In [51[18][44],68], it has been reported that a thin substrate and low interpass temperature are capable of minimizing the tensile RS in the WAAM products. The study also indicated that mechanical properties and microstructural characteristics are strongly dependent on the thermal history during printing. Likewise, significant temperature gradients, high heating and cooling rates, bulk temperature rises, and periodic thermal cycles cause variation in printed parts’ properties. One of the key factors in the WAAM process is thermal management to mitigate the accumulation of heat and cope with restrictions concerning geometry issues, the deposition cycle, and anisotropies in the mechanical properties [69,70][45][46]. Furthermore, a theoretical model was developed to optimize the heat input and the interlayer temperature for the deposition of each layer. In thin-walled structures built up in the WAAM process, molten pool solidification and bead geometry vary as wall height increases, mainly due to heat dissipation on the substrate. An active cooling system during the process using the technology of thermoelectric cooling is used to eliminate the variance in the dissipation of heat between the lower and upper layers [24,31,63][35][39][47]. The evolution of microstructure in the WAAM process depends on the material’s thermal history, which is estimated using FE modeling. Transformations of the microstructure are assessed with a diagram of continuous cooling transformation, and the microstructure is investigated with the use of scanning electron and optical microscopy. Similarly, the impact of temperature interpass on the hardness and microstructure was studied in [64,68,71][36][44][48] for AM80 HSLA steel, and the result revealed that a high interpass temperature leads to the development of martensite. During the WAAM process, heat accumulation results in the deposition of a multi-layer increase in the temperature preheat in the preceding built layer. These reasons lead to instabilities in the interlayer density, microstructure, mechanical properties, and geometry, which evolve differences in the material properties [72,73][49][50]. There are two types of temperature-measuring methods in the WAAM process: (1) contact (thermocouple) [10,30,33,74][17][22][26][51] and (2) non-contact measuring (infrared camera, pyrometrical methods, etc.,) [27,39,75,76][1][52][53][54].

2. Mechanical Modeling

Since the WAAM process has different thermal effects leading to varying mechanical properties of the printed parts, the process is examined in different building directions, which gave different results [18][3]. Additionally, variation in the substrate geometry has a significant effect on the molten pool geometry and heat transfer condition, which influence the welding for deposition of the WAAM process [1,77][55][56]. However, mechanical analysis to evaluate the stresses developed in the WAAM-printed products is carried out with a thermal load employed in finite element analysis (FEA) that caused RS and distortion. No significant variances were realized in either the mechanical or microstructural character succeeding heat treatments in air, in a vacuum, or in argon [78,79,80,81][57][58][59][60]. In the mechanical analysis, the clamping conditions (i.e., degrees of freedom constraints) have an important influence on the printed products [72][49].
The Lagrangian thermal model and nodal temperature results from the thermal analysis are used as thermal load inputs for the mechanical analysis [30,82][22][61]. Another investigation was performed in [30][22] using the Eulerian thermal model to calculate the steady-state temperature distributions, which are used as inputs to a 3D mechanical model for RS and distortion analysis. From the results of both the experimental and numerical studies, the temperature, distortion, and RS were compared and showed good agreement. Chen et al. [83][62] conducted a study of the superalloy (Ni-17Mo-7Cr) component fabricated with WAAM. The results showed stability in both tensile and hardness properties while a sharp reduction was perceived in the layers of the superficial surface. Thermal evolution is coupled to an elastoplastic mechanical boundary condition value problem that calculates the thermal stresses and distortions with Goldak thermal modeling approach. It was extensively modeled for accuracy by comparing the mechanical properties and thermal predictions with the equivalent experimental measurements with a non-linear transient function [41,53][4][23]. Hejripour et al. [84][63] studied on two parts built with WAAM, i.e., (1) wall and (2) tube from 2209 Duplex Stainless Steel. The results indicated that the slow rate of cooling for layers built at high temperatures caused austenite creation significantly in a ferrite matrix. The correlation between the developed thermal model for phase transformation and cooling rates is quite accurate to predict thermal cycles and weld zone profiles [7,69,85,86][45][64][65][66].
Even though WAAM technology is an energy-efficient manufacturing method for metal production, heat accumulation during deposition and associated mechanical changes result in RS, which was observed in the profiles across the cross-section of the printed components [25,87][67][68]. The mechanism of crack formation in a WAAM-manufactured Al-Zn-Mg-Cu alloy, which has high crack sensitivity, was studied using a combination of microstructure analysis and thermal-stress simulation [65,67][40][43]. The results indicated that increasing the deposition height increases the stress, which leads to crack propagation and the appearance of a macrocrack. In other studies, it has been reported that WAAM-manufactured super duplex stainless steel has both excellent corrosion resistance and mechanical properties [67,88][43][69]. In the studies, the microstructure of the wall deposited was carefully examined for variation in the mechanical properties. The results also revealed that the austenite/ferrite balance of phases in the wall body was fragmented by the overgrowing of the austenite phase. The anisotropic examination revealed that the ultimate tensile stress (UTS) and elongation appeared separately in the horizontal and vertical directions. The yield stress (YS) variables are rejected by the nitrogen work hardening result to a large extent.
WAAM technology with PAW has high deposition rates and can producing components of various sizes and yields with high mechanical performance. Similarly, two walls of Ti6Al4V were fabricated in a shielding argon atmosphere using WAAM-PAW to examine the deposition process, the growth in height per layer, the temperature deposition process, and the cooling times, which gave good results [69][45]. Three-dimensional thermoelastic–plastic transients and the thermomechanical multi-layer wall structure modeling approach are studied in [30][22]. The result of the study shows that temperature simulations and distortion expectations are proved by equating with the experimental results of laser scanners and thermocouples, while the RS is verified using neutron diffraction of strain scanner ENGIN-X. The response of the mechanical to the thermal history is examined by performing a 3D quasi-static incremental analysis.
The yield strength (σy��), elastic modulus (E), and coefficient of thermal expansion (α) are temperature-dependent, while other mechanical property values such as Poisson’s ratio are independent of temperature [89][70]. The physical and mechanical behavior of temperature-dependent properties such as the thermal conductivity, specific heat, and yield stress of SAE-AISI 1524 steel were studied in [90][71]. For materials such as SS 316 L, the behavior is supposed to be elastoplastic with a mix of kinematic and isotropic hardening (Chaboche-like model) [31][39]. Under the theory of small deformation, the total incremental strain can be disintegrated in elastic and inelastic forms. Equation (7) shows the formula for total deformation.
Δεtot=Δεel+Δεth+Δεp+ΔεvpΔ����=Δ���+Δ��ℎ+Δ��+Δ��� Δ ε t o t = Δ ε e l + Δ ε t h + Δ ε p + Δ ε v p
where Δεel, Δεth, Δεp, and ΔεvpΔ���, Δ��ℎ, Δ��, and Δ��� εvp are, respectively, the elastic, thermal, plastic, and viscoplastic total incremental strains. The thermal strain is expressed in the absence of metallurgical transformations by Equation (8):
εth=α(T)(TTref) ��ℎ=�(�)(�−����)   ε t h = α ( T ) ( T T r e f )
where α is the thermal expansion coefficient and Tref���� is a reference temperature. In all three directions, the strains measured were combined to analyze the stress, supposing the directions measured are given by Hooke’s law with principal strain directions [30][22], as Equation (9).
σx=E(1+υ)εx+Ev(1+υ)(12υ)(εx+εy+εz)��=�(1+�)��+��(1+�)(1−2�)(��+��+��) σ x = E ( 1 + υ ) ε x + E v ( 1 + υ ) ( 1 2 υ ) ( ε x + ε y + ε z )
where εx��, εy��, and εz��  are strains in the principal directions, E is the modulus of elasticity, and υ is Poisson’s ratio of the material.

References

  1. Carter, W.; Masuo, C.; Nycz, A.; Noakes, M.; Vaughan, D. Thermal process monitoring for wire-arc additive manufacturing using IR cameras. In Proceedings of the 30th Annual International Solid Freeform Fabrication Symposium—An Additive Manufacturing Conference Reviewed Paper, Austin, TX, USA, 8–10 August 2019; pp. 1812–1817.
  2. Israr, R.; Buhl, J.; Bambach, M. A study on power-controlled wire-arc additive manufacturing using a data-driven surrogate model. Int. J. Adv. Manuf. Technol. 2021, 117, 2133–2147.
  3. Köhler, M.; Hensel, J.; Dilger, K. Effects of thermal cycling on wire and arc additive manufacturing of al-5356 components. Metals 2020, 10, 952.
  4. Yang, Y.; Zhou, X.; Li, Q.; Ayas, C. A computationally efficient thermo-mechanical model for wire arc additive manufacturing. Addit. Manuf. 2021, 46, 102090.
  5. Bui, M.C.; Pham, T.Q.D.; Tran, H.S.; Van Tran, X. Efficient prediction of thermal history in wire and arc-directed energy deposition combining machine learning and numerical simulation. Res. Sq. 2022, 23, 0–23.
  6. Nagasai, B.P.; Malarvizhi, S.; Balasubramanian, V. Effect of welding processes on mechanical and metallurgical characteristics of carbon steel cylindrical components made by wire arc additive manufacturing (WAAM) technique. CIRP J. Manuf. Sci. Technol. 2022, 36, 100–116.
  7. Treutler, K.; Wesling, V. The current state of research of wire arc additive manufacturing (Waam): A review. Appl. Sci. 2021, 11, 8619.
  8. Ou, W.; Mukherjee, T.; Knapp, G.L.; Wei, Y.; DebRoy, T. Fusion zone geometries, cooling rates and solidification parameters during wire arc additive manufacturing. Int. J. Heat Mass. Transf. 2018, 127, 1084–1094.
  9. Srivastava, S.; Garg, R.K.; Sharma, V.S.; Sachdeva, A. Measurement and Mitigation of Residual Stress in Wire-Arc Additive Manufacturing: A Review of Macro-Scale Continuum Modelling Approach. Arch. Comput. Methods Eng. 2021, 28, 3491–3515.
  10. Cao, H.; Huang, R.; Yi, H.; Liu, M.L.; Jia, L. Asymmetric molten pool morphology in wire-arc directed energy deposition: Evolution mechanism and suppression strategy. Addit. Manuf. 2022, 59, 03113.
  11. Zhou, Z.; Shen, H.; Lin, J.; Liu, B.; Sheng, X. Continuous tool-path planning for optimizing thermo-mechanical properties in wire-arc additive manufacturing: An evolutional method. J. Manuf. Process. 2022, 83, 354–373.
  12. Derekar, K.S. Aspects of Wire Arc Additive Manufacturing (WAAM) of Alumnium Alloy 5183. Ph.D. Thesis, Coventry University, Coventry, UK, 2020; pp. 1–227.
  13. Kumar, N.; Bhavsar, H.; Mahesh, P.V.S.; Srivastava, A.K.; Bora, B.J.; Saxena, A.; Dixit, A.R. Wire Arc Additive Manufacturing—A revolutionary method in additive manufacturing. Mater. Chem. Phys. 2022, 285, 126144.
  14. Jafari, D.; Vaneker, T.H.J.; Gibson, I. Wire and arc additive manufacturing: Opportunities and challenges to control the quality and accuracy of manufactured parts. Mater. Des. 2021, 202, 109471.
  15. Singh, C.P.; Sarma, R.; Kapil, S. The qualitative analysis of warpage on residual stresses in wire arc additive manufacturing. Mater. Today Proc. 2022, 62, 6619–6627.
  16. Wu, Q.; Mukherjee, T.; De, A.; DebRoy, T. Residual stresses in wire-arc additive manufacturing—Hierarchy of influential variables. Addit. Manuf. 2020, 35, 101355.
  17. Ding, D.; Zhang, S.; Lu, Q.; Pan, Z.; Li, H.; Wang, K. The well-distributed volumetric heat source model for numerical simulation of wire arc additive manufacturing process. Mater. Today Commun. 2021, 27, 102430.
  18. Feng, G.; Wang, H.; Wang, Y.; Deng, D.; Zhang, J. Numerical Simulation of Residual Stress and Deformation in Wire Arc Additive Manufacturing. Crystals 2022, 12, 803.
  19. Li, R.; Xiong, J.; Lei, Y. Investigation on thermal stress evolution induced by wire and arc additive manufacturing for circular thin-walled parts. J. Manuf. Process. 2019, 40, 59–67.
  20. Gornyakov, V.; Sun, Y.; Ding, J.; Williams, S. Efficient determination and evaluation of steady-state thermal-mechanical variables generated by wire arc additive manufacturing and high pressure rolling. Model. Simul. Mater. Sci. Eng. 2022, 30, 014001.
  21. Oyama, K.; Diplas, S.; M’hamdi, M.; Gunnæs, A.E.; Azar, A.S. Heat source management in wire-arc additive manufacturing process for Al-Mg and Al-Si alloys. Addit. Manuf. 2019, 26, 180–192.
  22. Ding, J.; Colegrove, P.; Mehnen, J.; Ganguly, S.; Almeida, P.M.S.; Wang, F.; Williams, S. Thermo-mechanical analysis of Wire and Arc Additive Layer Manufacturing process on large multi-layer parts. Comput. Mater. Sci. 2011, 50, 3315–3322.
  23. Ahmad, S.N.; Manurung, Y.H.P.; Mat, M.F.; Minggu, Z.; Jaffar, A.; Pruller, S.; Leitner, M. FEM simulation procedure for distortion and residual stress analysis of wire arc additive manufacturing. IOP Conf. Ser. Mater. Sci. Eng. 2020, 834, 012083.
  24. Belhadj, M.; Werda, S.; Belhadj, A.; Kromer, R.; Darnis, P. Thermal analysis of wire arc additive manufacturing process. In Proceedings of the 24th International Conference on Material Forming 2021, Liège, Belgium, 14–16 April 2021; pp. 2018–2022.
  25. Montevecchi, F.; Venturini, G.; Scippa, A.; Campatelli, G. Finite Element Modelling of Wire-arc-additive-manufacturing Process. Procedia CIRP 2016, 55, 109–114.
  26. Sikan, F.; Wanjara, P.; Gholipour, J.; Kumar, A.; Brochu, M. Thermo-mechanical modeling of wire-fed electron beam additive manufacturing. Materials 2021, 14, 911.
  27. Wu, B.; Ding, D.; Pan, Z.; Cuiuri, D.; Li, H.; Han, J.; Fei, Z. Effects of heat accumulation on the arc characteristics and metal transfer behavior in Wire Arc Additive Manufacturing of Ti6Al4V. J. Mater. Process. Technol. 2017, 250, 304–312.
  28. Evjemo, L.D.; Langelandsvik, G.; Moe, S.; Danielsen, M.H.; Gravdahl, J.T. Wire-arc additive manufacturing of structures with overhang: Experimental results depositing material onto fixed substrate. CIRP J. Manuf. Sci. Technol. 2022, 38, 186–203.
  29. Wu, B.; Pan, Z.; van Duin, S.; Li, H. Thermal Behavior in Wire Arc Additive Manufacturing: Characteristics, Effects and Control; Springer: Singapore, 2019.
  30. Dharmendra, C.; Shakerin, S.; Ram, G.D.J.; Mohammadi, M. Wire-arc additive manufacturing of nickel aluminum bronze/stainless steel hybrid parts interfacial characterization, prospects, and problems. Materialia 2020, 13, 100834.
  31. Park, J.; Lee, S.H. Cmt-based wire arc additive manufacturing using 316l stainless steel (2): Solidification map of the multilayer deposit. Metals 2021, 11, 1725.
  32. Geng, H.; Li, J.; Xiong, J.; Lin, X. Optimisation of interpass temperature and heat input for wire and arc additive manufacturing 5A06 aluminium alloy. Sci. Technol. Weld. Join. 2017, 22, 472–483.
  33. Ou, W.; Knapp, G.L.; Mukherjee, T.; Wei, Y.; DebRoy, T. An improved heat transfer and fluid flow model of wire-arc additive manufacturing. Int. J. Heat Mass. Transf. 2021, 167, 120835.
  34. Rosli, N.A.; Alkahari, M.R.; bin Abdollah, M.F.; Maidin, S.; Ramli, F.R.; Herawan, S.G. Review on effect of heat input for wire arc additive manufacturing process. J. Mater. Res. Technol. 2021, 11, 2127–2145.
  35. Li, F.; Chen, S.; Shi, J.; Zhao, Y.; Tian, H. Thermoelectric cooling-aided bead geometry regulation in wire and arc-based additive manufacturing of thin-walled structures. Appl. Sci. 2018, 8, 207.
  36. Panchenko, O.; Kladov, I.; Kurushkin, D.; Zhabrev, L.; Ryl’kov, E.; Zamozdra, M. Effect of thermal history on microstructure evolution and mechanical properties in wire arc additive manufacturing of HSLA steel functionally graded components. Mater. Sci. Eng. A 2022, 851, 143569.
  37. Wu, B.; Pan, Z.; Ding, D.; Cuiuri, D.; Li, H. Effects of heat accumulation on microstructure and mechanical properties of Ti6Al4V alloy deposited by wire arc additive manufacturing. Addit. Manuf. 2018, 23, 151–160.
  38. Veiga, F.; Suárez, A.; Artaza, T.; Aldalur, E. Efect of the Heat Input on Wire-Arc Additive Manufacturing of Invar 36 Alloy: Microstructure and Mechanical Properties. Weld. World 2022, 66, 1081–1091.
  39. Cambon, C.; Rouquette, S.; Bendaoud, I.; Bordreuil, C.; Wimpory, R.; Soulié, F. Thermo-mechanical simulation of overlaid layers made with wire + arc additive manufacturing and GMAW-cold metal transfer. Weld. World 2020, 64, 1427–1435.
  40. Chen, S.; Xu, M.; Yuan, T.; Jiang, X.; Zhang, H.; Zheng, X. Thermal–microstructural analysis of the mechanism of liquation cracks in wire-arc additive manufacturing of Al–Zn–Mg–Cu alloy. J. Mater. Res. Technol. 2022, 16, 1260–1271.
  41. Sun, J.; Hensel, J.; Köhler, M.; Dilger, K. Residual stress in wire and arc additively manufactured aluminum components. J. Manuf. Process. 2021, 65, 97–111.
  42. Ding, D.; Pan, Z.; Cuiuri, D.; Li, H. Wire-feed additive manufacturing of metal components: Technologies, developments and future interests. Int. J. Adv. Manuf. Technol. 2015, 81, 465–481.
  43. Ríos, S.; Colegrove, P.A.; Martina, F.; Williams, S.W. Analytical process model for wire + arc additive manufacturing. Addit. Manuf. 2018, 21, 651–657.
  44. Derekar, K.; Lawrence, J.; Melton, G.; Addison, A.; Zhang, X.; Xu, L. Influence of Interpass Temperature on Wire Arc Additive Manufacturing (WAAM) of Aluminium Alloy Components. MATEC Web. Conf. 2019, 269, 05001.
  45. Scotti, F.M.; Teixeira, F.R.; da Silva, L.J.; de Araújo, D.B.; Reis, R.P.; Scotti, A. Thermal management in WAAM through the CMT Advanced process and an active cooling technique. J. Manuf. Process. 2020, 57, 23–35.
  46. Hönnige, J.; Colegrove, P.; Prangnell, P.; Ho, A.; Williams, S. The Effect of Thermal History on Microstructural Evolution, Cold-Work Refinement and {\alpha}/\b Growth in Ti-6Al-4V Wire + Arc AM. arXiv 2018, preprint. arXiv:1811.02903.
  47. Baier, D.; Wolf, F.; Weckenmann, M.; Zaeh, M.F. Thermal process monitoring and control for a near-net-shape Wire and Arc Additive Manufacturing. Prod. Eng. 2022, 16, 811–822.
  48. Kozamernik, N.; Bračun, D.; Klobčar, D. WAAM system with interpass temperature control and forced cooling for near-net-shape printing of small metal components. Int. J. Adv. Manuf. Technol. 2020, 110, 1955–1968.
  49. Singh, S.; Jinoop, A.N.; Tarun Kumar, G.T.A.; Palani, I.A.; Paul, C.P.; Prashanth, K.G. Effect of interlayer delay on the microstructure and mechanical properties of wire arc additive manufactured wall structures. Materials 2021, 14, 4187.
  50. Lopez, A.; Bacelar, R.; Pires, I.; Santos, T.G.; Sousa, J.P.; Quintino, L. Non-destructive testing application of radiography and ultrasound for wire and arc additive manufacturing. Addit. Manuf. 2018, 21, 298–306.
  51. Köhler, M.; Sun, L.; Hensel, J.; Pallaspuro, S.; Kömi, J.; Dilger, K.; <monospace> </monospace>Zhang, Z. Comparative study of deposition patterns for DED-Arc additive manufacturing of Al-4046. Mater. Des. 2021, 210, 110122.
  52. Pixner, F.; Buzolin, R.; Schönfelder, S.; Theuermann, D.; Warchomicka, F.; Enzinger, N. Contactless temperature measurement in wire-based electron beam additive manufacturing Ti-6Al-4V. Weld. World 2021, 65, 1307–1322.
  53. Jorge, V.L.; Teixeira, F.R. Pyrometrical Interlayer Temperature Measurement in WAAM of Thin Wall: Strategies, Limitations and Functionality. Metals 2022, 12, 765.
  54. Richter, A.; Gehling, T.; Treutler, K.; Wesling, V.; Rembe, C. Real-time measurement of temperature and volume of the weld pool in wire-arc additive manufacturing. Meas. Sensors 2021, 17, 100060.
  55. Ding, J.; Williams, S.W. Thermo-mechanical Analysis of Wire and Arc Additive Manufacturing Process. Ph.D. Thesis, Cranfield University, Bedford, UK, 2012; pp. 40–187.
  56. Mohebbi, M.S.; Kühl, M.; Ploshikhin, V. A thermo-capillary-gravity model for geometrical analysis of single-bead wire and arc additive manufacturing (WAAM). Int. J. Adv. Manuf. Technol. 2020, 109, 877–891.
  57. Artaza, T.; Suárez, A.; Veiga, F.; Braceras, I.; Tabernero, I.; Larrañaga, O.; Lamikiz, A. Wire arc additive manufacturing Ti6Al4V aeronautical parts using plasma arc welding: Analysis of heat-treatment processes in different atmospheres. J. Mater. Res. Technol. 2020, 9, 15454–15466.
  58. Srivastava, S.; Garg, R.K.; Sachdeva, A.; Sharma, V.S. A multi-tier layer-wise thermal management study for long-scale wire-arc additive manufacturing. J. Mater. Process. Technol. 2022, 306, 117651.
  59. Seow, C.E.; Coules, H.E.; Wu, G.; Khan, R.H.U.; Xu, X.; Williams, S. Wire + Arc Additively Manufactured Inconel 718: Effect of post-deposition heat treatments on microstructure and tensile properties. Mater. Des. 2019, 183, 108157.
  60. Vishnukumar, M.; Muthupandi, V.; Jerome, S. Effect of post-heat treatment on the mechanical and corrosion behaviour of SS316L fabricated by wire arc additive manufacturing. Mater. Lett. 2022, 307, 131015.
  61. Nadeem, M.F.; Ahmed, B. Thermo-Mechanical Modelling and Experimental Verification of Distortion and Residual Stress for S355J2G3 Plate Welded by using Different Filler. J. Sp. Technol. 2019, 9, 71–82.
  62. Chen, S.; He, T.; Wu, X.; Lei, G. Synergistic effect of carbides and residual strain on the mechanical behavior of Ni-17Mo-7Cr superalloy made by wire-arc additive manufacturing. Mater. Lett. 2021, 287, 129291.
  63. Hejripour, F.; Binesh, F.; Hebel, M.; Aidun, D.K. Thermal modeling and characterization of wire arc additive manufactured duplex stainless steel. J. Mater. Process. Technol. 2019, 272, 58–71.
  64. Mansoor, O.; Wahdatullah, N.; Harshavardhana, N. Wire Arc Additive Manufacturing (WAAM) of Inconel 625 Alloy and its Microstructure and Mechanical Properties. Int. Res. J. Eng. Technol. 2021, 8, 1517–1528.
  65. Bonifaz, E.A. Modelling of Thermal Transport in Wire + Arc Additive Manufacturing Process. Lect. Notes Comput. Sci. 2019, 11539, 647–659.
  66. da Silva, L.J.; Souza, D.M.; de Araújo, D.B.; Reis, R.P.; Scotti, A. Concept and validation of an active cooling technique to mitigate heat accumulation in WAAM. Int. J. Adv. Manuf. Technol. 2020, 107, 2513–2523.
  67. Srivastava, S.; Garg, R.K.; Sachdeva, A.; Sharma, V.S. Distribution of Residual Stress in Wire-Arc Additively Manufactured Small-Scale Component: Single- Versus Multi-Level Heat Input. J. Manuf. Sci. Eng. 2022, 145, 021008.
  68. Jimenez, X.; Dong, W.; Paul, S.; Klecka, M.A.; To, A.C. Residual Stress Modeling with Phase Transformation for Wire Arc Additive Manufacturing of B91 Steel. Jom 2020, 72, 4178–4186.
  69. Zhang, X.; Wang, K.; Zhou, Q.; Ding, J.; Ganguly, S.; Grasso, M.; Williams, S. Microstructure and mechanical properties of TOP-TIG-wire and arc additive manufactured super duplex stainless steel (ER2594). Mater. Sci. Eng. A 2019, 762, 138097.
  70. Norrish, J.; Polden, J.; Richardson, I. A review of wire arc additive manufacturing: Development, principles, process physics, implementation and current status. J. Phys. D. Appl. Phys. 2021, 54, 473001.
  71. Bonifaz, E.A.; Palomeque, J.S. A mechanical model in wire + Arc additive manufacturing process. Prog. Addit. Manuf. 2020, 5, 163–169.
More
Video Production Service