Next Article in Journal
Investigation of the Effect of Tool Rotation Rate in EDM Drilling of Ultrafine Grain Tungsten Carbide Using Predictive Machine Learning
Previous Article in Journal
Predicting the Relative Density of Stainless Steel and Aluminum Alloys Manufactured by L-PBF Using Machine Learning
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Processing, Microstructure, and Mechanical Behavior of Tungsten Heavy Alloys for Kinetic Energy Penetrators: A Critical Review

by
Rajneesh Patel
,
Gangaraju Manogna Karthik
* and
Pawan Sharma
Department of Mechanical Engineering, Indian Institute of Technology (BHU), Varanasi 221005, India
*
Author to whom correspondence should be addressed.
J. Manuf. Mater. Process. 2025, 9(6), 186; https://doi.org/10.3390/jmmp9060186
Submission received: 12 April 2025 / Revised: 21 May 2025 / Accepted: 28 May 2025 / Published: 4 June 2025

Abstract

:
Tungsten heavy alloys (WHAs) are two-phase composites known for their exceptional density, strength, hardness, and ductility, making them ideal for radiation shielding, kinetic energy penetrators, and aerospace components. Due to their high melting point, WHAs are primarily processed via powder metallurgy, with liquid-phase sintering (LPS). Spark plasma sintering (SPS) and microwave sintering are emerging as advanced consolidation techniques. Recent research has focused on improving WHA performance through microstructural manipulation, alloying with elements like Fe, Co, Mo, and Re; rare earth oxides like Y2O3, La2O3, and Ce2O3; and employing high-entropy alloys (HEAs) as matrix phase. Additionally, additive manufacturing (AM) techniques are increasingly being used to fabricate complex WHA components. Despite their advantages, WHAs still exhibit limitations in penetration performance, primarily due to their tendency to form mushroom-like heads upon impact rather than self-sharpening. Ongoing research seeks to enhance shear localization, refine grain structure, and optimize processing methods to improve the mechanical properties and impact resistance of WHAs. Furthermore, modeling and simulation approaches are being explored to understand the mechanical behavior of WHAs. This review comprehensively overviews the above aspects and presents recent advances in WHA processing.

1. Introduction

Kinetic energy penetrators (KEPs) are high-velocity projectiles designed to defeat armored targets using their mass and speed for penetration [1]. Modern KEPs, particularly Armor-Piercing Fin-Stabilized Discarding Sabot (APFSDS) rounds, require materials with high density and strength to maximize their impact effectiveness [2]. An ideal KEP material should exhibit low work-hardening rates and strain-rate hardening to facilitate flow softening and promote shear localization at lower strains. It should also have a low heat capacity to allow for a rapid temperature increase, along with high density and strength to maximize impact energy. Additionally, it should possess a high thermal-softening rate to ensure adiabatic shear banding initiation at lower temperatures, thereby enhancing penetration performance [3]. The performance of KEPs depends on various factors, including the penetrator’s material properties, geometry, mass, velocity, penetration depth, failure mechanisms, and target characteristics.
WHAs are two-phase composites made of W particles embedded in a ductile metallic matrix, typically containing 85–95 wt.% W [3]. W is a refractory metal with a melting point of 3420 °C [4], requiring extremely high temperatures for consolidation, and sintered pure W parts tend to be brittle [5]. To address these challenges, W powders are mixed with lower melting metals or metal combinations with some solubility in W [6]. The WHA matrix is a solid solution primarily of Ni, often with additives such as Fe, Co, Cu, Re, Ta, and Mo [6]. During sintering, the low-melting elements form a liquid phase, allowing rapid consolidation into a fully dense two-phase structure. Advanced sintering techniques, including SPS and microwave sintering, enable improved densification at lower temperatures with faster processing times [7].
WHAs are widely used for KEPs due to their high density (17–19 g/cm3), exceptional strength, and good ductility. They also offer reasonable toughness, radiation shielding, moderate electrical and thermal conductivity, effective vibration damping, and strong resistance to corrosion. Beyond KEP applications, WHAs are utilized in various industrial and military components, including high-performance weapon systems, rocket nozzles, oilfield rejuvenation projectiles, golf club weights, self-winding watch weights, aviation wing weights, and industrial vibrators [6,7,8,9,10]. Their high density enhances kinetic energy retention, leading to deeper armor penetration, while their toughness and resistance to deformation ensure effectiveness upon impact [11]. Importantly, WHAs serve as a viable alternative to depleted uranium (DU) KEPs, offering a non-radioactive and environmentally safer option [6].
The primary drawback of WHAs is their lower penetration capability compared to DU alloys [6]. The superior penetration ability of DU alloys is attributed to their self-sharpening behavior, which arises due to the formation of adiabatic shear bands (ASBs) upon impact. These bands promote localized shear failure, allowing the penetrator to maintain a sharp profile, thereby enhancing penetration depth [12]. In contrast, WHAs tend to form a mushroom-like tip, which reduces their overall penetration efficiency. Since ASBs play a critical role in penetration performance, key research focus for WHAs has been focused on enhancing their adiabatic shear localization tendency. This has been pursued through compositional modifications, such as adding Co [13] and Mo [14] to the Ni matrix or by replacing the Ni matrix with elements like Hf [15,16], Ti [17], and Ni3Al [18,19,20]. Additionally, novel processing approaches have been explored to enhance the mechanical properties of WHAs [13,21].
This review article summarizes recent advancements in WHA processing, strategies for producing high-strength WHA through microstructural manipulation, matrix composition modifications to promote ASB formation, thermo-mechanical processing techniques, and computational approaches for WHAs.

2. Processing of WHAs

WHAs are generally processed either by liquid-phase sintering (LPS) [22,23,24] or solid-state sintering (SSS) [25,26]. As previously discussed, WHAs consist of two components: the W phase and the matrix phase. The primary difference between LPS and SSS lies in the sintering temperature. In SSS, the sintering temperature remains below the melting point of both components, keeping the matrix phase in a solid state with no liquid formation. In contrast, LPS involves heating above the matrix phase melting point, allowing it to liquefy and bind the W particles.
Various techniques have been explored for sintering WHAs, including conventional high-temperature furnaces, SPS, and microwave sintering. These methods differ mainly in heating mechanisms (radiation/convection, microwave heating, and resistance heating) and heating rates. Conventional furnace sintering typically achieves heating rates of around 3–20 °C/min [27], while microwave sintering achieves rates around 50 °C/min [28], and SPS exceeds 100 °C/min [29]. Rapid heating and shorter dwell times in SPS and microwave sintering result in refined microstructures with enhanced mechanical properties [30]. Infiltration techniques have also been employed for WHA processing [31,32]. For instance, W-Cu [31] and W-Cu-Ag [33] alloys are fabricated through infiltration techniques, where a molten matrix phase permeates a sintered W skeleton [10]. Compared to conventional sintering, infiltration produces finer microstructures but increases W-W particle interfaces, leading to higher contiguity and reduced ductility.
While conventional sintering methods provide dense parts, they limit the fabrication of complex geometries. Powder injection molding (PIM) addresses this by enabling intricate shapes at lower machining costs [34,35,36]. In PIM, WHA powders are mixed with a polymer or wax binder to create a feedstock, which is then injection-molded into the desired shape [37,38]. The part then undergoes debinding to remove the binder, followed by sintering to achieve high density and desirable mechanical properties [39]. However, PIM faces challenges such as complex debinding procedures, dimensional instability, and microstructural defects, requiring careful optimization. Zu et al. [35] optimized the injection-molding of 90W-4.9Ni-2.1Fe, achieving a 1100 MPa tensile strength and 25% elongation, with properties influenced by the solvent-de-bonding temperature and environment [6]. AM techniques such as selective laser melting (SLM) [40,41,42,43], direct energy deposition (DED) [44,45,46], binder jet printing (BJP) [47], and bound metal deposition (BMD) [48,49] offer new possibilities for fabricating complex WHA components with minimal material waste and greater design flexibility. However, WHA processing via AM presents challenges due to W’s high melting point, low laser absorptivity, high thermal conductivity, and brittleness, leading to defects such as microcracking and poor densification [50].
Irrespective of the processing route, WHAs often exhibit poor mechanical properties due to the presence of a significantly weaker matrix phase than the W phase, brittle W/W interfaces. Defects such as porosity further degrade performance. To enhance strength and hardness, post-sintering thermo-mechanical treatments, such as hot isostatic pressing (HIP) [51], forging [52], rolling [53], extrusion [54], and swaging [22,55,56] have been employed. The following sections provide a comprehensive review of WHA processing techniques, highlighting their advantages, limitations, and potential for further development.

2.1. Sintering of WHAs

Extensive research has been conducted on the sintering of WHAs with diverse alloy compositions such as W-Ni-Cu [57,58,59], W-Ni-Fe [26,27,60,61,62,63,64,65,66,67,68,69], W-Ni-Fe-Co [13,27,56,58,70,71,72], W-Ni-Fe-Co-Mo [14,73,74,75,76,77,78], W-Ni-Co [79,80,81,82], W-Ni-Cu-Fe [14,83], W-Ni-Fe-Re [21,67], W-Ni-Fe-Mo [14], W-Ni-Mn [84], and W-Ni-Cu-Mn [85]. Recently, bulk metallic glass (BMG) W-ZrCuAlNi [86,87] and HEAs have been explored as alternative matrix materials, such as W-AlCrFeNiV [88], W-CoCrFeMnNi [89], W-FeNiCoCrCu [90], and W-CoCuFeNi [91].
Sintering is a thermal process that densifies powder compacts at elevated temperatures below their melting points. Since WHAs consist of a W phase and a matrix phase, the elemental powders must first be mixed to ensure uniformity. Various mixing techniques, including turbula mixing, attritor milling, planetary ball milling, high-energy vibratory milling, and mechanical alloying, are employed to achieve homogeneous powder blends [26,61,81,92]. Ball milling can function as either a high-energy mechanical mixing process or a low-energy blending technique. In high-energy milling, the powder particles undergo repeated cold welding, rupturing, and re-welding, altering their structure [93]. In contrast, low-energy milling improves homogeneity and particle bonding through controlled collision and deformation. Tumbler mixing, a low-energy approach (with or without milling media), ensures uniform elemental distribution without significant particle size reduction or structural changes.
Selecting the appropriate mixing method is essential for achieving high-density WHA samples. For 90W-7Ni-3Fe and 92.5W-5.25Ni-2.25Fe compositions, attritor milling enhances tensile properties and ductility over turbula mixing due to the improved powder dispersion [94]. Chen and Ma [95] investigated two milling methods for a W-Ni-Co-Y2O3 alloy. In the first, all elemental powders were ball milled for 24 h. In the second method, the Ni and Co powders were pre-milled separately for 24 h before being combined with W powder and Y2O3 nano powder for an additional 24 h. The latter approach produced a more homogeneously distributed matrix phase, resulting in improved microstructural uniformity.
The next step after mixing is compaction to the required shape. Compaction minimizes porosity, enhances densification, and ensures structural uniformity. Before compaction, WHA powders are typically reduced at 700 °C in a H2 atmosphere for 2 h to remove the oxide layer that develops during the mixing process [56,70]. This reduction step improves sintering behavior, removes impurities, and relieves stress from the mixing process. Panigrahi et al. [90] reported that mechanically milled W and HEA powders benefit from heat treatment at 600 °C for 2 h in H2 to relieve cold working stresses induced by mechanical milling. Generally, WHA powders are compacted using cold isostatic pressing (CIP) or hydraulic pressing. CIP applies uniform pressure in all directions, yielding higher green and sintered densities compared to uniaxial hydraulic pressing [96]. Table 1 summarizes the different WHA powder mixing and compaction methods.
Once compacted, WHAs undergo sintering, where process parameters, such as temperature, time, and atmosphere, play a critical role in determining microstructure and mechanical properties. Optimal sintering conditions promote densification, fine W phase dispersion, uniform matrix distribution, and minimized W/W interfaces [97]. Generally, WHAs are sintered in a reducing H2 atmosphere to prevent W oxidation at high temperatures. This effectively removes surface oxides, leading to cleaner metal surfaces and improved densification [14,22,70]. The W particle size and distribution also significantly affect sintered material properties. Fine W particles enhance densification and mechanical strength due to their larger surface-area-to-volume ratio, facilitating faster sintering [98]. However, finer W powders may also increase the risk of grain coarsening at higher sintering temperatures, which can reduce toughness. The following section presents the literature on the various sintering techniques used for WHA consolidation.
Table 1. Overview of the powder mixing and compaction methods used in the processing of WHAs.
Table 1. Overview of the powder mixing and compaction methods used in the processing of WHAs.
Composition (wt.%)Mixing Technique and DurationBall-to-Powder RatioCompaction MethodsRef.
93W-4.9Ni-2.1FeTumbler mixing for 30 h-CIP at 180 MPa[60]
92.5W-6.4Ni-1.1FeTurbula mixing for 30 min-Hydraulic pressing at 200 MPa[61]
90W-6Ni-2Fe-2CoBall milling for 24 h2:1, SS ballsCIP at 200 MPa[27]
93W-4.2Ni-1.2Fe-1.6CoBall milling for 48 h at 50 rpm2:1 CIP at 250 MPa[99]
90.5W-7.1Ni-1.65Fe-0.5Co-0.25MoBall milling at 35 rpm1:1, SS ballsCIP at 250 MPa[77]
90 W-6Ni-4CuTurbula mixing for 10 h at 96 rpm-Hydraulic pressing at 400 MPa[83]
91W-6Ni-3CoDrum mixing for 20 h-Hydraulic pressing at 200 MPa[82]
90W-5Ni-2Fe-3CoBall milling for 48 h at 40 rpm1:1CIP at 250 MPa[70]
90W-7Ni-3FeVibratory milling at 1000 rpm-Hydraulic press[26]
95W-3.5Ni-1.5FePlanetary ball milling for 5 h at 180 rpm5:1CIP at 200 MPa[92]

2.1.1. Solid-State Sintering of WHA

SSS of WHAs involves heating compacted powders below the melting point of both components (matrix and W). Since no liquid phase forms, densification is primarily driven by atomic diffusion, particle rearrangement, and grain boundary migration [100]. Additionally, plastic deformation mechanisms such as grain boundary sliding [101] and dislocation motion contribute to densification [102]. However, grain growth can occur as surface energy decreases, with larger grains forming at the expense of smaller ones, altering the material’s microstructure [102,103]. Ryu et al. [104] investigated the mechanical alloying of 93W-5.6Ni-1.4Fe and found that 48 h of milling produced a nanocrystalline W grain size of 16 nm. When sintered at 1300–1400 °C, the material achieved ultrafine W particles (~3 μm) and a high density (>97%). In contrast, liquid-phase sintering (LPS) at 1485 °C led to coarser particles (~27 μm) and a lower density due to swelling. Similarly, Gurwell [25] found that finer W powders (~1 μm) improved sinterability but reduced ductility, while coarser powders (5–10 μm) enhanced ductility but required longer sintering times. Optimal sintering near the matrix solidus temperature provides a balance between strength and ductility by incorporating both fine and coarse W powders.
Despite its advantages, SSS presents challenges. While SSS produces refined WHA microstructures, densification is slower due to its dependence on solid-state diffusion, often resulting in residual porosity. Additionally, maintaining a uniform distribution of the W and matrix phases is difficult due to atomic mobility and solubility differences, which can create weak regions and cause mechanical property variations. These limitations make SSS less widely used compared to LPS.

2.1.2. Liquid-Phase Sintering of WHAs

During LPS, the sintering temperature exceeds the liquidus temperature of the matrix phase. However, densification depends on selecting low-melting elements that can dissolve W, as only these elements effectively promote grain wetting and rearrangement. Pure Cu has a negligible solubility for W and fails to achieve significant densification, despite forming a liquid at sintering temperatures (Figure 1a). In contrast, adding Ni to Cu enhances wetting and enables partial W dissolution, resulting in near-complete densification [105]. As a result, early WHAs have Ni and Cu (90W-6Ni-4Cu and 90W-5Ni-5Cu, density ~16.5 g/cc) [106]. More recently, WHAs have evolved to include transition metals like Fe and Co, which offer superior solubility for W. Among these, W-Ni-Fe alloys have been extensively studied [105,106].
Unlike SSS, LPS significantly accelerates microstructural evolution by enhancing material transport through the liquid phase. Capillary forces pull W grains together, facilitating particle rearrangement [3]. Sharp W particle edges dissolve preferentially, leading to grain shape accommodation, while finer W grains dissolve into the molten matrix and later reprecipitate during cooling, a process governed by Ostwald ripening [3,107]. Sintering temperature and dwell time are critical in determining the final microstructure (Figure 1b). Suboptimal conditions result in excessive/insufficient liquid formation, causing inhomogeneity. Kuncicka et al. [81] investigated the effect of sintering temperature on 93W-6Ni-1Co and found that sintering below 1525 °C for 30 min produced an inhomogeneous structure. In contrast, sintering at or above 1525 °C or increasing the dwell time to 60 min led to improved homogeneity, enhanced mechanical properties, and greater W diffusion.
Higher W content increases the W volume fraction, contiguity, and grain size. Contiguity, which measures interphase contact, below 0.38 has been shown to prevent shape distortion, as confirmed by percolation theory and experimental findings [108]. Heating rates commonly used during the LPS of WHAs typically range from 5 to 10 °C/min. Bollina and German [109] analyzed densification at heating rates of 1, 5, 10, and 15 °C/min for WHAs with 83–93 wt.% W in a Ni-Fe (7:3) matrix. They found that the highest sintered density (95%) was achieved at 1 °C/min for 83 wt.% W.
Despite achieving near-theoretical densities [3], LPS WHAs are prone to shape distortion or slumping due to gravity [110]. This occurs because of density differences between solid W and liquid phases, leading to W particle segregation. LPS is typically restricted to materials with a high solid-phase fraction, and shorter sintering times are preferred [107]. Table 2 summarizes the sintering cycles and atmospheres used during the LPS of various WHA systems.

2.1.3. Spark Plasma Sintering of WHAs

SPS, also known as pulsed electric current sintering, is an advanced powder metallurgy technique that enables rapid densification through the simultaneous application of high temperature, pressure, and electric current [29]. In SPS, a graphite die containing the powder mixture is subjected to 30–70 MPa pressure, while an electric current generates uniform heating via resistance, enabling rapid densification at lower temperatures and shorter holding times [112]. Unlike LPS using a conventional furnace, which requires extended holding times, SPS minimizes the sintering time, limiting grain coarsening and leading to finer microstructures and improved mechanical properties. Hu et al. [113] investigated W-Ni-Fe alloys processed via SPS and found that slower heating rates promoted better densification and finer grain size, while higher rates accelerated transient liquid-phase formation, causing uneven microstructures and porosity. Pan et al. [114] studied mechanically milled 90W-6Ni-4Mn powders sintered at 1150 °C, achieving a high density (99.6%) and superior strength due to uniform matrix phase distribution and controlled grain growth. However, higher sintering temperatures increased porosity and reduced strength.
Using one-time spark plasma sintering on 90W-6Ni-4Mn powders (Figure 2a), W grain size was refined below 10 µm, achieving a maximum hardness of 69.1 HRA at 1150 °C [115]. Fan et al. [29] applied two-time spark plasma sintering (TTSPS) to the same alloy (Figure 2b), initially sintering at 1000 °C, followed by a second sintering at 1100–1200 °C with heating rates of 100 °C/min and 200 °C/min, respectively. This yielded finer W grains, improving hardness (76.6 HRA) and bending strength (785.3 MPa) due to the optimized grain structure and matrix phase distribution (Figure 2c). Singh et al. [116] analyzed an 84W-7Ni-3Fe-6Nb WHA sintered via SPS, observing that melt squeezing at 1215 °C indicated transient liquid-phase sintering due to eutectic reactions at powder contacts. The highest density (99%) was achieved at 1200 °C for 84W-7Ni-3Fe-6Nb, while 90W-7Ni-3Fe and 84W-7Ni-3Fe-6Mo reached 97.9% and 95.9% at 1150 °C, respectively. The 84W-7Ni-3Fe-6Nb alloy exhibited the highest hardness (678 HV at 1150 °C) due to the persistent liquid phase, whereas 84W-7Ni-3Fe-6Mo peaked at 589 HV at 1200 °C before declining. Coatings like boron nitride reduce carbon diffusion from graphite dies and prevent carbide formation on the sample surface (e.g., WC).
SPS enables rapid densification of WHAs at lower temperatures and shorter processing times (Table 3), improving density and mechanical properties. Controlled heating rates help prevent porosity and microstructural inconsistencies. However, SPS is mainly suitable for smaller components and is limited in producing large parts such as KEPs.

2.1.4. Microwave Sintering of WHAs

Microwave sintering is another advanced technique that consolidates materials using microwave radiation, in which heating occurs internally (inside-out), unlike conventional sintering, through dipolar polarization, conduction losses, and magnetic field interactions [119]. Microwaves consist of electric and magnetic components that influence the movement of dipoles, free electrons, domain walls, and electron spins during processing [119]. Solid metals and other conductors reflect microwaves, causing surface heating and plasma formation while microwave-absorbing materials convert radiation efficiently into heat [120]. Materials that directly absorb microwaves, such as ceramics and metallic powders, undergo direct heating. Bulk metals, which cannot absorb microwaves directly, require a susceptor, a microwave-absorbing material like graphite or silicon carbide (SiC), for hybrid heating [121]. Hybrid heating occurs in three stages: (1) The susceptor absorbs microwave energy and heats up, (2) it transfers heat via conduction and convection to the non-absorbing material, and (3) once the material reaches a critical temperature, it begins absorbing microwaves directly, leading to rapid heating [119]. Unlike conventional sintering, which transfers heat from the surface inward, microwave sintering generates heat volumetrically from the core outward, enhancing efficiency and uniformity [122]. Initially, heat transfer is driven by conduction and convection, with radiation becoming significant once the susceptor surpasses a threshold temperature.
Microwave sintering of WHAs has been carried out using a multi-mode cavity with a microwave frequency of 2.45 GHz [61,64,123]. SiC susceptors like dense SiC, oxide-bonded SiC, or nitride-bonded SiC are commonly used due to their high thermal stability (~1600 °C) and structural integrity, ensuring efficient and uniform heating [124]. Compared to conventional furnaces with slow heating rates (5 °C/min), microwave heating reduces the processing time and energy consumption by ~75% [125]. Zhou et al. [125] investigated the microstructural evolution of WHAs under different sintering conditions. Microwave sintering of 90W-6Ni-2Fe-2Co at 1470 °C for 60 min resulted in finer and more homogeneous microstructures compared to conventional sintering for 120 min at the same temperature, attributed to rapid densification and reduced grain growth (Figure 3a,b). Further, 90W-7Ni-3Fe alloys microwave-sintered at 1480 °C showed noticeable microstructural densification after just 10 min of holding time (Figure 3c,d). As shown in Figure 3e, increasing the heating rate promoted finer W grains and reduced contiguity, both of which are beneficial for mechanical performance.
Prabhu et al. [126] reported that microwave sintering significantly improved densification in pure W, achieving a 93% theoretical density with high-energy milled powder compared to 85% with as-received powder due to the finer particle size and higher surface energy. Microwave-sintered WHAs exhibit finer W particles and lower contiguity than conventionally sintered alloys, leading to superior mechanical properties [27]. For instance, a 90W-6Ni-2Fe-2Co alloy sintered via microwave reached a tensile strength of 989 MPa and impact strength of 203 J/cm2, while its conventionally sintered counterpart achieved only 816 MPa and 17 J/cm2. Microwave-sintered 90W-7Ni-3Cu samples also exhibited higher hardness than their conventionally sintered counterparts [59]. Liu et al. [127] studied the microstructural and mechanical behavior of a 93W-4.9Ni-2.1Fe alloy fabricated via microwave at 1500 °C. The process enhanced W dissolution and diffusion in the Ni-Fe matrix, yielding a relative density of 98.8%, tensile strength of 1185.6 MPa, elongation of 16.4%, and hardness of 42.1 HRC. As the sintering temperature increased, porosity decreased, W grains coarsened, and fracture modes shifted from intergranular to transgranular cleavage, improving mechanical strength. Additionally, W solubility in the matrix reached 35.79%, enhancing bonding strength while preventing brittle-phase formation due to rapid sintering and cooling.
Microwave sintering presents a cost-effective, energy-efficient alternative to conventional sintering, enabling faster heating, refined microstructures, and enhanced mechanical properties. However, MW heating develops non-uniform microstructures due to the skin effect, which causes surface-dominant heating and can lead to reduced electrical conductivity [7,128]. The skin effect occurs when alternating electromagnetic fields, like those used in microwave energy, induce electrical currents (called eddy currents) on the surface layer of conductive materials. These currents do not penetrate deeply into the bulk of the material, especially at high frequencies, and are confined to a shallow region near the surface, referred to as the “skin depth”. During microwave sintering, this concentration of current on the surface can lead to non-uniform heating, especially in materials like W, which is a good electrical conductor. While powdered metals can absorb microwave energy more effectively than bulk metals, surface-dominant heating still persists to some degree. As a result, there may be localized overheating and uneven microstructural evolution, which can introduce micro-defects, porosity, or grain boundary irregularities, all of which contribute to a drop in the material’s bulk electrical conductivity after sintering. Moreover, any phase transformations, impurity inclusion, or grain boundary oxidation that occur preferentially near the surface due to microwave-induced thermal gradients may further degrade the continuity of electron pathways, reducing overall conductivity. Strategies like using susceptors, hybrid sintering, and optimizing powder characteristics aim to improve heat distribution and minimize defects [7,128]. Table 4 summarizes the research on WHA consolidation via microwave sintering.
A summary of the microstructural and property benefits from different sintering methods is presented in Table 5.

2.2. Additive Manufacturing of WHAs

AM technologies have been an alternative for producing complex, customized WHA components, such as 90W-7Ni-3Fe [132,133,134], 95W-5Nb [40], 95W-5Ta [135], W-Cr [136], 75W-25Cu [137], W-2.5TiC [138], W-10 Cu [139], 90W-6Ni-2Fe-2Co [140], 80W-10Ni-10Cu [141], and pure W [142]. Powder bed fusion (PBF), a key AM method, uses either a laser (L-PBF) or electron beam (EBM) to selectively melt micron-sized powder layer by layer [143,144]. However, processing W is challenging due to its high ductile-to-brittle transition temperature [142]. Schwanekamp et al. [145] investigated the effect of L-PBF processing parameters on W-Ni-Fe alloys and found that optimal density was achieved at energy inputs of 500–1000 J/mm3, while excessive energy caused porosity due to evaporation. Preheating at 800 °C had minimal impact, but post-processing heat treatment at 1500 °C significantly improved the microstructure, making it comparable to conventionally sintered alloys. Similarly, Ivekovic et al. [146] reported that SLM-processed 90W-7Ni-3Fe reached ~95% density at 250–350 J/mm3, but excessive energy or preheating at 400 °C led to cracking. The as-built alloy exhibited brittle fracture with a UTS of 871 MPa.
In L-DED, WHAs such as W-Ni and W-Ni-Fe displayed layered microstructures with unmelted W and intermetallics, resulting in high strength but low ductility due to porosity and brittle fracture paths [147,148,149]. Wei et al. [150] studied the microstructure and fracture characteristics of a 90W-7Ni-3Fe alloy produced by L-DED. Compared to conventional LPS, L-DED resulted in a non-uniform microstructure with finer W particles and a higher W solubility in the matrix (Figure 4). Mechanical properties varied along the build direction, with the highest tensile strength in the middle section (Figure 4c). L-DED samples exhibited W particle cleavage and matrix rupture, while LPS samples showed more W-W interparticle fractures. Zhang et al. [132] reported that, during the laser sintering of W-Ni-Fe, the W particles did not fully melt or disperse evenly in the region with a cellular structure. In contrast, the Ni and Fe particles melted completely and fused with the unmelted W particles. EBM, with preheating up to 1400 °C, offered superior control over cracking and produced dense, crack-free structures with columnar grains and mixed textures [151,152,153,154,155]. Mechanical properties varied, with compressive strengths from 900 to 1523 MPa via alloying and oxide dispersion [151,152,153,154,155].
BJP is another viable AM method for producing WHA components with complex geometries. BJP offers fewer processing challenges than laser-based methods. The process involves inkjet printing a liquid binder onto a layer of metal powder to form a green part, which is then de-bonded and sintered to create a solid component [156]. Stawovy et al. [134] studied the BJP of W-Ni-Fe-Cu, achieving sintered densities of up to 99.7% of the theoretical value with 18–19% dimensional shrinkage. The tensile properties met ASTM standards, with an average UTS of 770 MPa and 8.6% elongation. Zhou et al. [136] developed a short-time laser powder sintering method that drastically reduced the liquid-phase duration, preventing excessive W grain growth and producing finer grains. This method resulted in WHAs with a 42% higher yield strength and improved ballistic performance. The challenges during AM can be mitigated by optimizing alloy compositions, refining powder selection, adjusting laser parameters, and improving processing environments.
Continued advancements in alloy design, microstructure control, and impurity mitigation are essential for producing high-performance WHAs via AM. Cracking is the most critical issue during the AM of WHAs due to W’s high ductile-to-brittle transition temperature (200–400 °C), thermal stresses, and grain boundary embrittlement from impurities such as oxygen [157,158,159]. Cracks predominantly form along high-angle grain boundaries (HAGBs) [157,158,160,161,162,163]. Although oxygen levels as low as 30 ppm have been reported, cracking persists, indicating that factors such as residual stress, grain morphology, and the intrinsic brittleness of W also contribute significantly [159,164]. Grain refinement, achieved through alloying additions (e.g., Ta, Nb, ZrC) or processing strategies (e.g., scan vector rotation, remelting), can reduce crack density but not fully eliminate it [159,165,166]. Powder characteristics (size, shape, distribution) and laser absorptivity also influence cracking behavior. W’s high thermal conductivity (173 W/m·K) accelerates heat dissipation and cooling, impairs melt pool wetting, and promotes balling defects [167]. Moreover, W’s high melting point, low laser absorptivity, and high viscosity hinder effective melting and fusion during printing. These challenges can be mitigated by using finer, spherical powders to enhance absorptivity and by optimizing process parameters to improve densification and microstructure control [168]. A systematic study on these parameters is still lacking.
Despite these inherent challenges, advances in laser processing, powder engineering, preheating, and hybrid manufacturing techniques have significantly improved build quality and mechanical performance. For example, Chen et al. [140] optimized SLM parameters (350 W laser power, 400 mm/s scan speed, 60 µm hatch spacing, 30 µm layer thickness) for W-18Ni-6Fe-6Co (wt.%), achieving 96.1% relative density, 1198 MPa ultimate tensile strength, and 9.5% elongation. Similarly, Wang et al. [147] optimized the process parameters for 90W-7Ni-3Fe (800 W laser power, 400 mm/min scan speed, 0.5 mm layer height, and a powder feed rate of 10–15 g/min under high-purity Ar shielding). The samples showed an alternating layered microstructure with reduced porosity (~0.87%), finer grains, increased W content in the matrix, and lower contiguity. These microstructural improvements, combined with supersaturated W in the matrix and enhanced interface bonding from rapid thermal cycling, yielded a high yield strength of 822 MPa and a tensile strength of 1037 MPa, surpassing conventional LPS WHAs. Hybrid AM approaches, such as combining AM with conventional techniques like sintering, hot isostatic pressing (HIP), or machining, are particularly beneficial for processing WHAs. One example is the BMD process, which integrates fused filament fabrication (FFF) and PIM to print bound metal rods into complex geometries, followed by debinding and sintering to produce dense metal parts. Bose et al. [48] used BMD to prepare 93 wt.% W, achieving a more than 96% sintered density.

2.3. Post-Processing of WHAs

For KEP applications, high strength and impact toughness are essential for optimal performance against target materials. However, WHAs, in their sintered state, often fall short of these requirements. To address this, post-sintering strategies such as heat treatment and thermo-mechanical processing (e.g., swaging and hot isostatic pressing) are employed to enhance mechanical properties. Various heat treatment methods have been explored, including vacuum annealing, solution treatment, cyclic heat treatment, and aging in different atmospheres (e.g., vacuum, H2, N2, Ar, or gas mixtures), followed by quenching in oil, water, or Ar. These treatments help prevent impurity segregation at the W/matrix interface and reduce hydrogen embrittlement, promoting a uniform distribution of elements within the matrix [58,169]. The heat treatment temperature is selected based on the solidus and solvus temperatures of the matrix.
For most WHAs, vacuum annealing at 1100–1250 °C followed by quenching can eliminate trapped hydrogen, remove W/matrix interfacial segregation, suppress intermetallic formation, and improve chemical homogeneity [76,79]. Das et al. [75] observed a decrease in W volume fraction (from 80% to 75%) and contiguity (from 0.7 to 0.3) in an LPS 90.5W-7.1Ni-1.65Fe-0.5Co-0.25Mo alloy after vacuum heat treatment at 1100 °C for 1 h. During this process, W partially dissolves into the matrix and remains in a solid solution after oil quenching. Solution heat treatment dissolves intermetallic compounds that may form during furnace cooling after sintering. Kiran et al. [73] found that in 90W-6Ni-2Fe-0.5Co-1.5Mo alloy, after three heat treatment cycles, contiguity and dihedral angle were reduced while matrix phase fraction was increased. This indication is relatively less prevalent in the case of single-cycle specimens.
Kumari et al. [71] reported that water quenching after solution treatment improved elongation and impact resistance in 89.6W-6.2Ni-1.8Fe-2.4Co alloys compared to Ar quenching. While Ar heat treatment removes hydrogen, water quenching prevents interstitial segregation and intermetallic precipitation at grain boundaries. During cyclic heat treatments, Kumari et al. [79] applied a repeated sequence of low-temperature holds (800–900 °C for 3 h) and high-temperature treatments (1100–1200 °C for 1.5 h), followed by vacuum heat treatment and water quenching. The low-temperature stage led to the formation of Co3W intermetallic compounds, while the high-temperature stage dissolved them. More cycles increased the volume of intermetallic precipitates. Das et al. [58] reported improved tensile strength (from 717 MPa to 1000 MPa) and elongation (from 5% to 20%) in a 91W-7Ni-1.5Fe-0.5Co alloy after heat treatment due to improved W/matrix bonding and a more uniform impurity distribution.
Grain refinement through alloying and deformation via swaging can further improve WHA strength. Despite the challenges posed by WHA hardness and brittleness, advancements in swaging technologies have enhanced processing efficiency and final product quality. Swaging after sintering also increases density and promotes strain hardening. Kiran et al. [169] studied the effects of swaging and heat treatment on an LPS 90W-6Ni-2Fe-2Co alloy. Heat-treated rods were swaged and aged at various temperatures for 1 h in N2. Swaging increased dislocation density, and subsequent aging significantly improved hardness. In a related study, Kiran et al. [73] found that cyclic heat treatment of 90W-6Ni-2Fe-0.5Co-1.5Mo reduced the contiguity and dihedral angle while increasing W solubility and the matrix volume fraction. Swaged specimens exhibited a lower work-hardening exponent than heat-treated ones. Using finite element analysis, Kocich et al. [30] used finite element analysis and experiments to assess sintered and cold-swaged W-Ni-Co alloys. Microhardness values for the sintered W particles and matrix were 389.2 and 296.8 HV, respectively. Swaging increased UTS significantly but reduced ductility. Caliskan et al. [69] investigated rotary swaging of a 90W-7Ni-3Fe alloy produced by vacuum LPS. They reported improved density, tensile strength, and ductility. Fractographic analysis revealed a shift from intergranular fracture in sintered alloys to a combination of transgranular fracture, matrix deformation, and W grain intergranular failure in swaged samples. The mechanical properties of WHAs after various post-processing treatments are summarized in Table 6.

3. Effect of Composition on the WHA Characteristics

To enhance the strength, hardness, and ductility of WHAs, strengthening the matrix phase through the addition of various alloying elements is one effective strategy. Common additions include Co, Mo, Ta, and Re. The W-Ni-Cu alloy was the first WHA developed. While it has a higher density than W-Ni-Fe alloys, its tensile strength is lower. Additionally, increasing Ni content in W-Ni-Cu alloys reduces the dihedral angle, further enhancing densification [24]. Das et al. [14] found that adding just 0.1 wt.% Fe to W-Ni-Cu alloys improved both the hardness and tensile strength. Shakunt et al. [83] reported that incorporating 1.0 wt.% Fe, while maintaining a Ni–Cu ratio of 6:4, increased the sintered density to 99.4%. However, a higher Fe content also increased the dihedral angle, indicating a more stable W/W interface than the W/liquid interface. This impeded liquid penetration and led to a higher compressive strength due to the reduced matrix volume and higher W/W contiguity. Wensheng et al. [170] observed that adding up to 1.0 wt.% Re to 93W-4.9Ni-2.1Fe increased the tensile strength from 997.2 MPa to 1161.2 MPa, though elongation dropped significantly from 26.4% to 8.6%. In contrast, Cr additions (0–1 wt.%) led to a reduction in tensile strength (from 997.2 MPa to 844.4 MPa) and elongation (from 26.4% to 7.7%). Cr tends to accumulate at interfaces, weakening interfacial cohesion and thereby degrading mechanical performance. Wang et al. [171] investigated W-Ta alloys with 5–20 wt.% Ta and found consistent grain size reduction compared to pure W. W-10Ta, prepared via SPS, demonstrated the best mechanical properties. Fractography revealed a transition from intergranular fracture in pure W to transgranular fracture in W-Ta alloys, indicating increased toughness with Ta addition. Chen et al. [172] reported that adding up to 10 wt.% Mo to W-Mo-Ni-Fe composites reduced the W grain size from 27.13 μm to 14.93 μm. Mo also improved the density and hardness, likely due to intermetallic compound formation.
Manganese (Mn) additions led to non-uniform elemental distribution in sintered compacts, reducing electrical conductivity, microhardness, and densification [85]. While traditional W-Ni-Fe WHAs perform well at low strain rates, their penetration capabilities are inferior to DU, likely due to a lower tendency to form ASBs. Since adiabatic shear is influenced by thermal conductivity and Mn has much lower thermal conductivity than Fe, replacing Fe with Mn may promote shear band formation. Hong et al. [173] studied a 90W-6Ni-4Mn alloy and emphasized the need to prevent manganese oxide formation during sintering to achieve full densification in Mn-containing WHAs.
The graph compiled from over 60 data points from the literature illustrates the effect of various alloying elements Fe, Cu, Mo, Re, Co, and Mn on the W content in the matrix phase of W-Ni-based WHAs (Figure 5). The summary of the effect of alloying elements on WNi-based WHAs is presented in Table 7. Among these, Co exhibits the highest W solubility, with both mean and median values clearly exceeding those of the other alloying additions. An increased W content in the matrix not only contributes to solid solution strengthening but also enhances the wetting and interfacial bonding between the matrix and W particles, which is critical for effective load transfer and improved performance under high-strain-rate conditions. In contrast, Cu shows the lowest matrix W content, reflecting its poor ability to dissolve W and limited wetting characteristics. This can result in weak particle/matrix interfaces and reduced mechanical properties. Mn shows a broad distribution in matrix W content, but its mean and median values are not significantly higher than those of the baseline W-Ni-Fe system. This suggests that, while Mn may influence local solubility or microstructural features, but due to its tendency to segregate, it does not substantially enhance overall W dissolution in the matrix. Re and Mo also exhibit moderate to low matrix W contents, with values close to or slightly below those of the Fe-containing baseline. Despite Re known advantages in high-temperature systems, its effect on W solubility in this context appears limited, possibly due to restricted solubility or kinetic factors during processing. Overall, matrix W content is a key factor influencing the strength, ductility, and interfacial integrity of WHAs. Among the elements studied, only Co clearly improves W solubility, making it particularly beneficial for high-performance applications such as KEPs.
The formation of fine W precipitates within the matrix of WHAs is governed by a complex interplay of alloy composition, thermal treatment, and microstructural evolution. Notably, alloys with a W-Ni-Co composition, such as 90W-7Ni-3Co, have shown a propensity to form fine W-rich precipitates within the matrix when subjected to specific cyclic heat treatments. These treatments typically involve an initial hold at 850 °C, where a W-rich intermetallic phase is formed in the Ni-Co matrix, followed by heating to 1150 °C to dissolve the intermetallic phase. Upon cooling, this process results in the nucleation of fine W precipitates (~1 μm), leading to a two-phase matrix microstructure that significantly enhances the alloy’s yield strength and strain-hardening response [174,175]. In contrast, WHAs containing Fe, such as 89.6W-6.2Ni-1.8Fe-2.4Co, tend not to form matrix precipitates under the same thermal conditions. This suppression of precipitation is attributed to the altered thermodynamic stability of the matrix phase caused by Fe addition, which discourages the formation and dissolution of the intermetallic phase necessary for precipitation [175]. Further studies have shown that the chemistry of the matrix itself plays a dominant role in determining precipitation behavior, where the W solubility and mobility in the matrix are key controlling factors [79]. Moreover, the presence of alloying elements like Mo has been shown to promote the precipitation of Mo-Ni intermetallic phases at W matrix interfaces when sufficient Mo is present in the liquid phase during sintering [172]. These findings underscore the importance of both thermodynamic driving forces and kinetic pathways in designing WHAs with tailored precipitation behavior and enhanced mechanical properties.
Additionally, rare earth oxides such as Y2O3, La2O3, and Ce2O3 have shown potential in refining the WHA microstructure and enhancing its mechanical performance, as fine oxide particles can improve WHA’s self-sharpening and penetration behavior. Dong et al. [176] synthesized W-5 wt.% Y2O3 composite powders via a wet chemical method, incorporating ultrasonic treatment and sodium dodecyl sulfate as a surfactant. They concluded that the traditional wet chemical process yielded high-quality composite precursors. Y2O3 additions improved hardness and thermal conductivity due to grain refinement. Park et al. [177] added 0.1 wt.% Y2O3 to a 93W-4.9Ni-2.1Fe powder mixture mechanically alloyed for 72 h, and LPS at 1493 °C for 1 h. Although oxide dispersion was achieved, the resulting specimens had a lower tensile strength and elongation than conventional LPS samples. Incomplete densification and micropores led to microcrack initiation under tensile loading, reducing ductility. These defects primarily formed at triple junctions and W/matrix interfaces. Xiao et al. [178] employed a hydrothermal method and mechanical alloying to fabricate nanoscale Zr(Y)O2-dispersed 93W-4.9Ni-2.1Fe alloys using molecular-level liquid–liquid doping combined with hot isostatic pressing. The resulting alloy showed a more uniform nanoparticle distribution than conventional powder processing methods. This combined approach effectively reduced particle size, contributing to improved dispersion strengthening. The brittleness of WHAs is highly sensitive to microstructural factors such as W grain size, porosity, and impurity segregation at grain boundaries. To achieve fine-grained WHAs, ultrafine composite powder precursors and rapid sintering are effective.
Table 7. Summary of the effect of alloying elements on WNi-based WHAs [61,85,170,179,180].
Table 7. Summary of the effect of alloying elements on WNi-based WHAs [61,85,170,179,180].
Feature/ElementsFeCoMoReMn
Effect on the matrixSolid solution strengtheningSolid solution strengtheningSolid solution strengtheningStrong solid solution strengtheningLimited solid solution strengthening
Grain size effectGrain refinement at moderate Fe content; excessive Fe can lead to grain coarseningGrain refinement at moderate Co contentGrain refinement and coarsening resistanceStrong grain refinementLimited grain refinement
Sintering behaviorImproves wettability up to moderate Fe content; reduces at high Fe contentImproves wettability up to moderate Co content; reduces at high Fe contentEnhances densificationEnhances sintering kinetics and diffusivityImproves wettability; assists in densification
DuctilityDecreases due to the formation of brittle intermetallic phasesDecreases above 6 wt.% due to brittle intermetallic phasesMaintains or slightly decreasesImproves both room- and high-temperature ductilitySlightly decreases
Intermetallic compoundsFe7W6Co3W, Co7W6MoNi-type intermetallic compoundNo intermetallicsSpecific intermetallic compound not reported; however, MnO can form during sintering and can reduce density
Precipitate formationFe-rich precipitate forms at the W/matrix interfaceCo-rich precipitates at the W/matrix interfaceMo-rich precipitates at grain boundariesNone, remains in solid solutionNone
Specific mechanical propertiesHigh strength and moderate ductilityHigh strength and moderate ductilityHigh-temperature strength and creep resistanceSuperior high-temperature strength and ductilityNot ideal for high-temperature applications

4. Microstructure of WHAs

The microstructure of WHAs undergoes significant transformations during LPS, which directly impacts their mechanical properties. This process involves densification, pore elimination, and microstructural coarsening, primarily driven by diffusion and capillary forces. Once the liquid phase forms, it infiltrates grain boundaries, facilitating particle rearrangement, enhancing densification, and causing localized softening [181]. The resulting microstructure features bonded W grains embedded in a solidified matrix. Key parameters such as grain size, W volume fraction, contiguity, and dihedral angle govern the alloy’s performance [182].
In the early stages of sintering, densification occurs through bulk transport mechanisms, including grain boundary diffusion and solid-state activated sintering [8]. Grain boundary diffusion is the dominant densification mechanism between 1300 and 1750 °C. Upon the formation of the liquid phase, pre-existing W-W sinter bonds dissolve due to a solubility shift, enabling liquid penetration into grain boundaries. As sintering progresses, coarsening occurs through a dissolution–reprecipitation mechanism, where smaller grains dissolve into the liquid and reprecipitate onto larger ones, reducing the grain count and increasing the average grain size [183]. This grain growth is described by the cubic model:
G 50 3 = G 0 3 + K t
where G0 is the initial grain size, G50 is the median grain size at the time t, and K is the temperature-dependent growth rate constant [184]. Pore elimination proceeds sequentially, with smaller pores shrinking first, promoting full densification [8]. In an 88W-10Ni-2Cu alloy sintered at 1412 °C, the median pore size initially increases due to grain restructuring and then decreases as full density is achieved [185]. The relationship between grain size and porosity is given by:
G 50 = θ G 0 ( 1 f ) 0.5
where f is fractional porosity [57]. As sintering continues, grains adopt polygonal shapes, aiding in pore elimination and enhancing mechanical strength [186]. Although the presence of a liquid phase accelerates diffusion, excessive liquid content can hinder grain growth by increasing diffusion distance [187]. The grain growth rate parameter K (μm3/s) depends on the liquid volume fraction (VL), as:
K = 1.1 + 1.9 V L 2 / 3
Experimental results reveal that higher liquid fractions reduce grain boundary interactions, moderating coarsening rates [188]. Additionally, liquid migration causes grain size variations along the sample height as heavier W grains settle and the liquid migrates upward. The transition from solid-state to liquid-assisted transport is critical to achieving optimal densification and mechanical integrity. Understanding these mechanisms is key to optimizing processing parameters for superior material performance.
The mechanical properties of WHAs are largely governed by microstructural features such as W grain size, contiguity, matrix volume fraction, and W solubility in the matrix [13]. The distribution, size, and morphology of these grains are influenced by sintering temperature, duration, and alloy composition. Grain size is commonly defined by the number of grains per unit area, and the yield strength is related to it via the Hall–Petch relationship [189].
σ y = σ 0 +   K d 1 / 2
where σy is the yield strength, σ0 the friction stress, K is the material constant, and d is the grain size. For WHAs, yield strength is also influenced by matrix volume fraction VM and W grain size D, as shown by [99]:
σ y = σ 0 +   G b 1 V M D V M 0.5
where G is the shear modulus and b is the Burgers vector. Li et al. [65] proposed another model considering the W-W contiguity CWW and matrix fraction, assuming that yielding begins in the Ni-Fe-W matrix:
σ y = σ 0 + K C W W 1 V M G W V M 0.5
where GW is the W grain size. This model suggests that yield strength increases with decreasing W grain size and matrix fraction. Ductility, according to Panchal and Nandy [30], depends on the matrix phase fraction and contiguity. A higher matrix fraction generally increases elongation due to the softer FCC structure of the matrix. W grain size also influences the failure mode during tensile testing. Large grains tend to fail via intergranular fracture—a lower energy process—whereas fine grains promote transgranular fracture due to the higher energy required to form new surfaces along W/matrix interfaces [14]. Humail et al. [66] found that increased W content reduces ductility due to higher W-W contiguity and fewer ductile matrix regions. The brittle W-W interfaces are more prone to microcrack formation and propagation, lowering the overall toughness. Contiguity (CWW), which describes the connectivity between W grains, is calculated as:
C W W = 2 N W W     2 N W W + N W M
where NWW and NWM are the counts of W-W and W/matrix interfaces, respectively, typically measured via line-intercept methods on SEM images [65]. Another critical microstructural parameter is the dihedral angle, the angle at the junction of two adjacent W grains. It impacts densification behavior: Low liquid and low dihedral lead to intermediate densification with distortion, while low liquid and high dihedral results in slow densification and minimal distortion. High liquid and low dihedral enable rapid densification with distortion, whereas high liquid and high dihedral causes slow densification with minimal distortion [8]. Optimizing liquid content and controlling the dihedral angle is essential for balancing densification, grain growth, and mechanical properties. Typically, the liquid phase content in LPS lies within the range of 5 to 15 vol.% [107].
The contiguity plots based on a comprehensive dataset compiled from 200 sources in the literature spanning a range of W contents underscore the critical influence of microstructural features on the mechanical behavior of WHAs fabricated through various sintering techniques (Figure 6a). The contiguity vs. W content plot reveals a consistent trend of increasing contiguity with increasing W content, reflecting the formation of a more interconnected W particle network. This effect is especially pronounced in conventionally sintered samples, where extended sintering durations and slower diffusion kinetics facilitate grain coarsening and promote extensive W-W contact. In contrast, SPS and microwave sintering result in lower contiguity at comparable W levels, primarily due to their rapid densification kinetics and the preservation of finer, more discrete W grains. Notably, SPS-processed samples exhibit slightly higher contiguity than the microwave-sintered counterparts at a similar W content. This distinction is attributed to the lower sintering temperatures typically employed in SPS, which limit diffusion and reduce W dissolution, thereby maintaining a greater degree of direct W-W contact. Post-sintering deformation processes, such as swaging or extrusion followed by heat treatment, further impact contiguity. These high-temperature mechanical and thermal treatments promote the dissolution of W into the matrix, disrupting the continuity of the W network and advancing the development of a more homogeneous microstructure. This reduction in contiguity enhances ductility and toughness by mitigating the brittleness associated with highly interconnected W frameworks while still preserving strength through improved matrix bonding and efficient load transfer.
Similarly, the W grain size vs. W content plot (Figure 6b) highlights the role of the sintering technique in controlling grain coarsening. Conventionally sintered WHAs show a clear increase in grain size with rising W content, primarily due to the extended thermal exposure that enables grain boundary migration and particle coalescence. In contrast, SPS and microwave sintering effectively suppress grain growth across the composition range due to the rapid densification and lower processing temperatures, resulting in finer and more uniform microstructures. Importantly, the data indicate that post-sintering deformation and subsequent heat treatments have minimal influence on W grain size, regardless of the initial sintering technique. This observation suggests that the inherent hardness and thermal stability of W grains limit further coarsening during mechanical processing and thermal exposure.
In addition, the dihedral angle vs. relative density plot (Figure 6c) reveals an inverse relationship, whereby the dihedral angle decreases with increasing relative density. This trend reflects enhanced densification during sintering, as higher relative densities are typically associated with the more effective liquid-phase wetting and capillary-driven rearrangement of W particles. Improved wetting behavior reduces the dihedral angle by strengthening solid–liquid interfacial interactions and minimizing pore volume at grain boundaries. As a result, lower dihedral angles at higher densities contribute to stronger interparticle bonding and a more cohesive microstructure, ultimately improving the mechanical performance.
Overall, these observations from the literature data highlight the role of sintering kinetics and densification mechanisms in controlling the key microstructural parameters, such as contiguity, grain size, and dihedral angle, that govern the mechanical response of WHAs.

5. Mechanical Properties of WHAs

WHAs’ mechanical properties depend on their W content (typically 80–97 wt.%), matrix composition, sintering parameters, and post-processing treatments. Increasing the W content enhances density and stiffness but can lead to grain coarsening and the formation of continuous W-W networks, which reduce ductility and promote brittleness. Achieving full densification is critical, as residual porosity significantly lowers both strength and ductility. In fully densified WHAs, hardness remains around 300 HV across a wide range of W contents but can reach ~428 HV with ceramic reinforcements (e.g., Y2O3, B4C) or post-sintering treatments [8]. However, over-sintering causes grain growth, which degrades mechanical performance. In W-Ni-Fe systems with a 7:3 Ni:Fe ratio, the post-sintered matrix typically consists of ~54Ni-23Fe-23W, achieving a ~500 MPa tensile strength and ~133 GPa elastic modulus. The testing method affects the reported strength values; tensile testing offers the most reliable data while bending and compression often overestimate performance [190,191,192]. Optimal strength and ductility are typically achieved near 93 wt.% W, particularly after deformation-aging treatments (e.g., swaging + aging), which can raise hardness (~330 HV), yield strength (~791 MPa), and tensile strength (~1039 MPa). Above 95 wt.% W, grain coarsening and increased contiguity degrade performance, and tensile strength becomes less consistent due to microstructural variability [193].
Processing techniques significantly influence WHA performance. Processing strongly influences mechanical behavior. Annealing can reduce strength by 15–20% but typically increases ductility by 20–30%. Deformation-aging (10–25% swaging + heat treatment) can increase tensile strength to 1375–1435 MPa, with an elongation of 7–10% [8]. Optimized sintering and deformation-aging are essential for balancing strength and ductility. Further, WHAs’ performance is governed by phase morphology, volume fraction, and interfacial characteristics. Weak W/W and W/matrix interfaces often cause premature failures in the matrix, while strong interfacial bonding shifts failure to W cleavage, improving strength. The first WHAs were based on W-Ni-Cu, and while they demonstrated higher density, they suffered from lower strength and corrosion resistance than W-Ni-Fe systems. W-Ni-Fe remains the dominant system, with 7:3 or 8:2 Ni–Fe ratios offering good mechanical balance. Excessive alloying may promote intermetallic phases and embrittlement [14]. Wu et al. [24] compared the LPS behavior of 80W-14Ni-6Cu vs. 80W-14Ni-6Fe alloys. While both reached full densification, the Ni-Cu alloy distorted significantly at higher sintering temperatures, whereas the Ni-Fe alloy retained dimensional stability up to 1500 °C. Kiran et al. [99] showed that adding 1.6 wt.% Co to 93W-4.9Ni-2.1Fe increased W solubility in the matrix and enhanced strength. Similarly, 0.2 wt.% Re addition further improved tensile strength.
Typically, sintered W-Ni-Cu/W-Ni-Fe often exhibit a strength range of 0.7–1 GPa, below the >1 GPa required for KEPs [70]. Das et al. [58] showed that a 91W-7Ni-1.5Fe-0.5Co alloy had superior room-temperature mechanical properties due to finer grains and lower porosity. Post-sintering heat treatment improved tensile behavior, and a 5Ni-5Fe matrix enabled >16% elongation without compromising strength. Kumari et al. [79] found that increasing the Ni/Co ratio in W-Ni-Co alloys suppressed fine W precipitates, softening the alloy. Swaged WHAs showed increased hardness with lower Ni–Co ratios. Panchal et al. [68] studied the flow and work-hardening behavior of 92W-5.5Ni-2.5Fe, noting that swaged specimens had higher strength, while heat-treated ones showed better elongation. Humail et al. [66] observed that 93 wt.% W alloys achieved the highest tensile strength due to optimized grain size, though ductility declined at higher W contents.
W-Ni-Co is a relatively newer ternary alloy system than W-Ni-Fe and W-Ni-Cu alloys. Co is commonly added to improve strength and ductility by reinforcing the binder and enhancing W grain wettability and interfacial bonding. Dincer et al. [194] investigated the microstructures of the W-Ni-Co system and found that microstructures consisted of rounded, almost pure W grains and a Ni-Co-W binder matrix phase. Adding Co led to an increase in W solubility in the matrix phase of up to 42 wt.%. However, higher Co content promoted the formation of brittle intermetallic phases (e.g., Ni2Co3W, Co7W6), reducing toughness [80]. Skoczylas et al. [82] found that a 91W-6Ni-3Co alloy reached theoretical density (~17.46 g/cm3), and that swaging from 15% to 20% increased the strength by ~70 MPa. Despite its mechanical benefits, Co poses health risks such as lung fibrosis and asthma [195], prompting efforts to develop Co-free WHAs. Cury et al. [196] demonstrated that Co-free alloys, with optimized Ni/Fe ratios and heat treatment, can match the mechanical performance of Co-containing variants. In ballistic testing, Co-free alloys achieved comparable penetration, confirming their suitability for KEPs. Further improvements have been made by refining W-Ni-Fe microstructures through rare earth additions (Y, La, oxides) and refractory metals (Re, Ta, Mo) [197]. Powder mixing methods also influence the mechanical properties. Eroglu et al. [94] compared wet attritor milling and dry turbula mixing for 90W-7Ni-3Fe and 92.5W-5.25Ni-2.25Fe alloys, finding that wet-milled powders produced denser, more ductile alloys for a given W powder size. Coarser W particles increased ductility in turbula-mixed alloys, whereas finer powders resulted in the lowest tensile strengths.
Figure 7 collectively illustrates the influence of different sintering techniques—conventional, microwave, and SPS—on the mechanical properties of WHAs, based on an extensive dataset compiled from over 250 sources in the literature, spanning a wide range of W contents. Both microwave and SPS methods generally produce higher tensile strengths compared to conventional sintering. This improvement is primarily attributed to their rapid heating rates and enhanced densification kinetics, which facilitate the formation of finer and more homogeneous microstructures. In contrast, WHAs processed via conventional sintering typically exhibit lower as-sintered strengths. However, post-sintering treatments such as swaging or extrusion followed by heat treatment play a critical role in improving tensile properties by refining the microstructure and reducing residual porosity, thereby enabling conventionally sintered alloys to reach performance levels comparable to those achieved through advanced sintering routes. Elongation data show greater variability, particularly in SPS-processed samples. This variability is likely due to the lower sintering temperatures associated with SPS, which make ductility more sensitive to microstructural fluctuations and W content. Conventionally sintered alloys generally exhibit lower elongation, often due to coarser W grains and higher porosity, unless improved through secondary processing. Similarly, both SPS and microwave sintering typically yield higher hardness values due to the improved densification and refined grain structures. Despite these general trends, all three datasets show substantial standard deviations, highlighting the complex interplay between processing conditions, alloy composition, and post-sintering treatments.
The contiguity plots further emphasize the role of microstructural connectivity in influencing mechanical behavior across different sintering techniques (Figure 7d,e). In the contiguity versus elongation plot, a clear negative correlation is observed: higher contiguity tends to reduce elongation, as increased W-W contact promotes early strain localization and brittle fracture. The contiguity vs. tensile strength relationship shows more complex behavior. While moderate contiguity contributes to strength through better load transfer across the W network, excessively high contiguity (common in high W alloys) leads to brittleness and premature failure. Additionally, the graph depicting the dihedral angle versus tensile strength reveals an inverse relationship: as the dihedral angle decreases, tensile strength increases (Figure 7f). This trend is primarily linked to the wetting behavior of the liquid binder phase (typically Ni-Fe or Ni-Co) during LPS. A lower dihedral angle reflects the improved wetting of W particles by the matrix, which strengthens the interfacial bonding between the hard W grains and the ductile binder. Stronger interfaces enable more effective load transfer and reduce the risk of intergranular fracture. Moreover, improved wetting promotes a more uniform and defect-free microstructure, further contributing to enhanced tensile strength. Conversely, a higher dihedral angle indicates poor wetting, resulting in weak interfacial bonding, increased porosity, and a higher susceptibility to crack initiation and propagation. Thus, optimizing sintering conditions to achieve lower dihedral angles is essential for improving the mechanical integrity of WHAs.
Summary of mechanical properties of WHAs with different sintering methods and with different alloying elements are presented in Table 8 and Table 9, respectively. Overall, while microwave and SPS techniques provide superior strength and hardness, conventionally sintered WHAs, when subjected to appropriate post-sintering treatments, can achieve comparable tensile and ductility characteristics. The significant data scatter highlights the need for careful optimization of processing routes to reliably tailor the mechanical performance of WHAs for critical applications.

6. Deformation Behavior of WHAs

The deformation behavior of WHAs is complex and governed by the interplay between W grains and the matrix phase. Key factors such as temperature, strain rate, and microstructural characteristics significantly affect their mechanical response. Understanding these aspects is essential for optimizing WHAs for various applications. Kim et al. [198] examined how the W particle shape affects the dynamic deformation and fracture behavior in WHAs using Kolsky bar torsional tests. Microstructural characterization (e.g., micro-CT, fractography), mechanical property measurements, and fractography results revealed that WHAs produced with coarse and irregularly shaped W particles contributed to a predominantly cleavage fracture mode. In contrast, WHAs with finer particles showed more localized central fractures with minor shear deformation, indicating different deformation and failure mechanisms that aligned closely with their observed impact behavior.
Park et al. [177] explored dynamic deformation in ODS WHAs, incorporating 0.1 wt.% Y2O3 via mechanical alloying. The mechanical alloying process refines the W particle size without increasing the interfacial area fraction, promoting interfacial de-bonding and ASB formation. The ODS WHAs exhibited lower maximum shear stress and fracture strain than conventionally sintered specimens, suggesting improved penetration performance through enhanced self-sharpening. Fang et al. [32] investigated the penetration and self-sharpening behavior of 80W-14Cu-6Zn, 90W-7Ni-3Fe, and 35CrMnSi alloys. Their results showed that the 80W-14Cu-6Zn penetrator benefited from high strength and an appropriate critical failure strain. They proposed that the melting and ejection of the Cu-Zn matrix during penetration acts as a lubricant, reducing lateral force and enhancing self-sharpening. Ryu et al. [199] studied 93W-5.6Ni-1.4Fe WHAs fabricated by mechanical alloying and sintering (1445–1485 °C, H2 atmosphere). They found that increasing the matrix content and reducing W/W contiguity enhanced the impact energy. The processing method also significantly affected the failure mode, which shifted from a brittle intergranular fracture in SSS WHAs to a ductile matrix shear fracture in LPS WHAs. These findings emphasize the importance of promoting a high matrix volume and low W contiguity for optimal mechanical performance [75].
Gong et al. [200] analyzed the dynamic deformation of a fine-grained 93W-4.9Ni-2.1Fe-0.03Y (wt.%) alloy with an average grain size of 6 µm at different strain rates. At higher strain rates, ASBs containing dense dislocations and twins were observed despite minimal thermal softening, suggesting that adiabatic shear failure can occur independently of a significant temperature rise. Grain refinement and Y-containing second-phase particles altered the deformation mechanisms of fine-grained WHAs relative to coarse-grained WHAs.
Self-sharpening in WHAs is critical for improving penetration performance and relies on localizing ASBs under high shear strain. U alloys, especially U-3/4Ti, historically excel due to ASB-induced self-sharpening, but environmental risks have shifted attention toward alternatives like WHAs and U-6Nb alloys. Despite WHAs’ high melting point and strain-rate sensitivity, they often suffer from mushrooming upon impact, which reduces penetration depth [3,201]. Mushrooming refers to the deformation mode where the leading end of a projectile flattens and expands radially upon high-velocity impact, resembling the cap of a mushroom. This phenomenon is commonly observed in penetrators made of pure W and WHAs, particularly under conditions where adiabatic shear localization is delayed or suppressed. W has a very high melting point (~3420 °C) and low thermal softening, meaning that even under intense plastic deformation, it resists flow softening. Moreover, W exhibits strong strain-rate sensitivity, so as the strain rate increases, the material’s resistance to deformation increases as well. These characteristics delay the onset of localized plastic flow (i.e., shear banding), and instead, the material deforms more uniformly and radially at the impact face. This deformation mode reduces the penetration efficiency, as energy is dissipated in a radial expansion rather than being concentrated along a narrow axial path. In contrast, uranium alloys, such as U-3/4Ti, exhibit early shear localization and fracture, producing a self-sharpening chiseled nose that maintains a narrow cross-section and penetrates deeper into the target [6].
Key strategies to mitigate this issue include reducing the W/W interfacial area and increasing the W grain size to lower the contiguity. Various testing methods, such as Hopkinson bar tests [202], normal plate impact tests [203,204,205], and oblique flyer-plate impact experiments [206], have been employed to study WHA behavior under dynamic loading. ODS-WHAs (0.1 wt.% Y2O3) demonstrate enhanced ASB formation due to improved interfacial bonding [177], while pre-twisted WHAs exhibit W grain aspect ratio and orientation-dependent ASB formation [207]. ASBs nucleate where maximum shear stress aligns with the major axis of W grains and propagate through the Ni-Fe matrix, which requires lower shear deformation energy.
Microstructural heterogeneity, particle shape, and matrix content significantly influence ASB formation [198,208]. Ultrafine-grained (UFG) W (~500 nm) exhibits lower strain-rate sensitivity and pronounced flow softening at ~103 s−1 strain rates, with peak stresses around 3 GPa. The strain-rate sensitivity of UFG W has been reported to be lower than that of coarse W [209]. WHA penetrators processed via hydrostatic extrusion and hot torsion (HE + HT) exhibit enhanced self-sharpening behavior, maintaining an acute shape upon ballistic impact due to localized ASBs, unlike as-sintered or as-extruded WHAs, which develop mushroomed heads [210]. Temperature and strain rate also play vital roles. The deformation behavior of W-Ni-Fe alloys has been studied across a range of temperatures (25–1100 °C) and strain rates (8 × 102–4 × 103 s−1) [211]. W-Ni-Fe WHAs shows increasing flow stress with strain rate and decreasing work hardening with temperature. At high shearing strain rates (~7 × 105 s−1), pressure-shear plate impact tests show thermal-softening-induced ASBs, demonstrating significant strain-rate sensitivity [212,213]. Using Taylor impact tests, ASB formation has been quantified as a function of impact velocity [214]. Reverse ballistic impact tests at velocities of 173 m/s and 228 m/s reveal a slight ellipticity in the mushroomed end, with low-speed impacts producing radial cracks and higher speed impacts leading to localized deformation along a curved path [215]. Numerical simulations using Johnson–Cook constitutive models provide equivalent failure strains for ASB formation [216].
Shear banding in WHAs often originates at the transition between the mushroomed region (deformed) and the undeformed regions. ASBs occur across various temperatures. At cryogenic temperatures, less shear strain occurs, indicating that temperature influences ASB formation. Shear banding has been observed in liquid-phase-sintered 90W-7Ni-3Fe alloys at strain rates of 5400 s−1 [217]. High-strain-rate Kolsky bar tests confirm that WHA deformation is primarily governed by W grain behavior [202].
The thermo-mechanical mismatch between W particles and ductility in WHAs promotes conditions favorable for shear localization. W particles exhibit higher flow stress, density, strain-rate sensitivity, and thermal conductivity, along with lower strain hardening and specific heat, compared to the matrix. As a result, the matrix deforms earlier under load, generating heat that is rapidly conducted into the W particles. This raises their temperature, reduces their flow stress, and promotes ASBs. This thermo-mechanical coupling explains why composite WHAs are more susceptible to shear localization than their individual constituents [218]. ASBs are narrow zones of intense plastic deformation caused by localized adiabatic heating at high strain rates, such as those experienced during ballistic impact. Under such conditions, the rate of heat generation exceeds the rate of thermal dissipation, triggering a rapid temperature rise and localized thermal softening. If the deformation fails to be controlled, it leads to a “mushrooming effect”, which increases the diameter of the penetrator’s head and reduces its ability to penetrate [3].
Gong et al. [200] observed <10 µm wide ASBs in fine-grained WHAs under a 1.9 × 103 s−1 strain rate. Wu et al. [219] demonstrated ASB formation in 93W-4.5Ni-1.5Fe-1Co under high-temperature split Hopkinson pressure bar and hypervelocity impacts into a concrete target test using a two-stage light gas gun. They demonstrated ASBs as fracture initiation sites and a key failure mode in WHAs under dynamic loading but also showed that they contribute to self-sharpening, a desirable behavior that enhances penetration efficiency in armor-piercing applications [220].
W grain size and matrix composition strongly influence ASB formation and penetration performance [200]. Luo et al. [221] showed that fine-grained 95W rods achieved greater penetration depth and produced narrower craters compared to coarse-grained rods, which exhibited mushrooming and wider craters. Kim et al. [63] revealed that ASB morphology was affected by W particle elongation, interfacial strength, and matrix composition. Cracks often initiated in shear bands but sometimes diverted along W/W interfaces due to their brittleness, which hindered shear band growth. Therefore, reducing W/W particle interfaces through microstructural control and optimized fabrication (e.g., heat treatment) was essential to promoting narrow, high-strain ASBs that facilitated crack propagation and self-sharpening.
An effective strategy to improve the ASB behavior in WHAs is to use alternative matrix materials with high susceptibility to shear localization. BMGs [86], HEAs [88,222], and systems like W-Ni-Mn [173] and W-Cu-Zn [223] have shown potential. Ma et al. [86] fabricated WHA-BMG composites using SPS with varying W volume fractions in a Zr55Cu30Al10Ni5 matrix. Under compressive loading, elastic modulus mismatch at the W/matrix interface generated stress concentrations that initiated shear bands. A higher W content promoted band formation, increasing free volume and facilitating localized plasticity. Hu et al. [87] demonstrated that W-reinforced, Zr-based BMGs (Zr55Cu30Al10Ni5) retained a sharp tip during penetration due to matrix softening and W plastic deformation, effectively preventing mushrooming.
Despite their advantages, BMGs suffer from limited macroscopic plasticity and poor fracture toughness [86]. Processing BMGs is also a challenge due to the critical cooling rate requirements to retain the amorphous phase; methods like spark plasma sintering are often required, limiting their scalability. Although high-density BMGs based on elements like Hf and Ta offer improved ductility and density, economic and processing constraints hinder their broader application in penetrators [224]. Conventional WHAs like W-Ni-Mn [173] and W-Cu-Zn [223] systems also enhance shear localization. Mn reduces the thermal conductivity of the WHAs and leads to enhance shear localization, while a Cu-Zn matrix facilitates shear failure. However, both systems suffer from drawbacks: Mn high vapor pressure increases porosity, and Cu-Zn soft matrix reduces dynamic strength, thereby necessitating the development of new matrix materials that overcome these limitations [225].
Recently, HEA as a matrix in WHAs has shown promise in improving penetration performance [210]. HEAs represent a novel category of materials that consist of several metallic elements, usually four or more, combined in approximately equal atomic ratios [226,227]. Liu et al. [228] developed a W-FeNiMo HEA-WHA featuring BCC, an FCC matrix, and μ-phase precipitates. These precipitates promoted dynamic recrystallization via localized strain gradients, sustaining sharp projectile heads and improving penetration depth by 10–20%. Zhou et al. [88] demonstrated that an AlCrFeNiV HEA matrix improved W/matrix interface properties. HEAs with low W solubility and diffusivity were particularly effective in refining W grains. Satyanarayana et al. [89] used CoCrFeMnNi HEAs as a crack-resistant WHA matrix. In addition to W grains and the HEA phase, consolidated WHA composites also showed the existence of a Cr-Mn-rich oxide phase. The Cr-Mn-rich oxide phase restricted the W grain growth. The SPS-processed composites exhibited higher compressive strength (~2041 MPa). Panigrahi et al. [90] reported improved densification, hardness, and strength in 90W-10FeNiCoCrCu HEA-WHAs. Anwar et al. [222] further showed that HEA-WHAs delivered 42% higher hardness and enhanced strain hardening, supporting superior penetration performance and positioning them as a viable alternative.

7. Modeling and Simulation of WHAs

Modeling and simulating WHAs requires a multidisciplinary approach that combines materials science, mechanical engineering, and computational techniques. Given their application in high-strain-rate environments, understanding the dynamic mechanical behavior of WHAs is crucial for optimizing performance. Sharma et al. [229] developed four constitutive models, the Modified Johnson–Cook (m-JC), Modified Zerilli–Armstrong (m-ZA), Modified Arrhenius (m-Arr), and Modified Khan–Huang–Liang (m-KHL) models, to predict the flow stress behavior of a 92.5W-5.25Ni-2.25Fe alloy. These models accounted for the coupled effects of strain, strain rate, and temperature. Processing maps were constructed at strains of 0.16, 0.18, and 0.20. Optimal processing conditions were identified at 0.16 strain, high strain rates (2500–4000 s−1), and temperatures between 427 and 627 °C, making this ideal for evaluating WHA penetration performance.
Finite element modeling (FEM) plays a pivotal role in simulating sintered component behavior, accounting for surface tension, gravity, grain growth, friction, and solid content. Park et al. [230] used FEM with constitutive equations for linear viscous compressive materials to simulate grain evolution, densification, and distortion. Li et al. [231] conducted detailed FEM simulations to investigate the penetration performance of W fiber/Zr-based metallic glass composite rods, comparing them to conventional WHA rods. The simulations incorporated a microstructure-based geometric model and a modified coupled thermo-mechanical constitutive model to accurately reflect the high strength and shear sensitivity of the metallic glass matrix under high-strain-rate conditions. Li et al. [232] created W/BMG composites using LS-DYNA, incorporating realistic particle distributions via a Monte Carlo approach. Couque et al. [214] examined how shear banding affects the penetration performance of four conventional WHAs, each containing 90 wt.% W, a ~17.1 g/cm3 density, and a ~1100 MPa yield stress, but differing in ductility and impact energy. Using symmetric Taylor impact tests (200–310 m/s) and ballistics tests (~1400 m/s against 303–330 HB steel) and AUTODYN-2D simulations with J-C models, they identified the critical impact velocities (199.8–223.2 m/s) for shear band initiation and the corresponding equivalent plastic strains.
Kennedy and Murr [233] provided a benchmark for simulation accuracy by comparing measured residual microhardness maps and penetration channel geometries with simulations generated via AUTODYN-2D. Hafizoglu and Durlu [234] highlighted that smoothed-particle hydrodynamics (SPH)-based simulations more accurately predicted the complex fracture shapes and dimensions observed in Taylor impact tests than mesh-dependent FEM approaches. FEM and SPH have proven essential for simulating high-strain-rate deformation and failure modes, including fragmentation and adiabatic shear banding. The Johnson–Cook model remains effective for WHAs, while modified coupled thermo-mechanical models address the shear localization and amorphous nature of metallic glasses [232,234]. Hafizoglu et al. [235] demonstrated that increasing the sintering temperature to 1480 °C and adjusting the Ni/Fe ratio to 4.67/2.33 optimized the plasticity and decreased contiguity, leading to improved fragmentation and energy transfer during impact. Simulation results correlated well with experimental data in terms of deformation at the impact and rear surfaces of targets, indicating that material design can be finely tuned to maximize ballistic effectiveness.
Lee et al. [211] examined the effect of strain rate and temperature on the impact deformation of a 92.5W-5.25Ni-2.25Fe LPS WHA. They employed the Zerilli–Armstrong dislocation mechanics-based model to describe the material’s rate-dependent behavior, which successfully captured the deformation response under high-rate and high-temperature conditions. This model is particularly valuable as it distinguishes between body-centered cubic (BCC) and face-centered cubic (FCC) metals. The mechanical performance of WHAs depends on the characteristics of each phase and the strength of interfacial bonding. Stress redistribution is limited under high-strain-rate loading, such as in explosions, leading to differences in stress and strain across phases. This mismatch contributes to damage mechanisms, including W grain cleavage, the ductile fracture of the matrix, and interfacial separations (W/W and W/matrix) [236]. Tang et al. [237] developed a two-phase constitutive model with damage for a 92.5W-4.9Ni-2.1Fe-0.5Co alloy. Using the Mori–Tanaka mean field theory and Eshelby’s equivalent inclusion theory, they incorporated damage evolution through continuum damage mechanics. The model showed good agreement with experimental data, except at extreme strain rates.
WHAs are modeled as two-phase composites consisting of rigid W particles dispersed within a ductile Fe-Ni-W matrix. Each phase is described using the Johnson–Cook (J-C) constitutive model, incorporating thermal softening based on the Zhou et al. formulation [238]. For the W particles, the material parameters are as follows: yield stress A = 730 MPa, hardening modulus B = 562 MPa, strain-rate sensitivity C = 0.029, strain-hardening exponent n = 0.0751, shear modulus μ = 155 GPa, bulk modulus K = 317 GPa, density ρ = 19,300 kg/m3, and specific heat capacity c = 138 J/kg·°C. The Fe-Ni-W matrix is characterized by the following: A = 150 MPa, B = 546 MPa, C = 0.0838, n = 0.208, μ = 98.84 GPa, K = 202.4 GPa, ρ = 9200 kg/m3, and C = 382 J/kg·°C. Taylor impact simulations conducted on a cylindrical rod measuring 6.35 mm in diameter and 25.4 mm in length at impact velocities between 173 and 228 m/s produce deformation results that closely align with experimental observations [238].
Sun et al. [239] evaluated a 90W-7Ni-3Fe WHA under both quasi-static (0.001–0.1 s−1) and dynamic (2000–6000 s−1) compression conditions, providing valuable experimental validation for simulation models. The constitutive behavior was accurately captured using the J-C model with the parameters A = 730 MPa, B = 1614 MPa, n = 0.16, and C = 0.109.
The J-C model remains widely used due to its simplicity and limited calibration requirements. It captures strain, strain rate, and temperature effects, and differentiates between BCC and FCC materials [240]
σ y = ( A + B ϵ n ) 1 + C l n ε ˙ ε ˙ 0 1 T T r T m T r m
where ε is equivalent plastic strain, ε ˙ is plastic strain rate, T is absolute temperature, T r is reference temperature, T m is melting temperatures, and A, B, n, C, and m are material constants.
The Zerilli–Armstrong (Z-A) model is based on dislocation mechanics and includes strain hardening, strain-rate sensitivity, and thermal softening. It also allows for discrimination between BCC and FCC materials and coupling between temperature and strain rate [240]:
  σ y = C 1   + C 2   e C 3   + C 4   l n ε ˙ T + C 5   ϵ n
Here, C1 relates to dislocation mechanisms and grain size, and C2-C5 and n are material constants. Scapian [240] applied both J-C and Z-A models to a W-Ni-Cu alloy with the following parameters: ε ˙ 0 = 1   s 1 , T r = 298   K , and T m = 1673   K . The optimized parameters were A = 647 MPa, B = 1337 MPa, n = 0.5616, C = 0.0640, and m = 0.4937. Both models showed excellent correlation with experimental flow stress data (R2 = 0.9883). The dynamic materials model treats the workpiece as an energy dissipation system. Under steady strain and temperature, the input power (P) during plastic deformation is split into two components: G (heat from plastic deformation) and J (internal energy change due to microstructural evolution) [229]:
P = σ ε ˙ = 0 ε ˙   σ ε ˙ + 0 σ   ε ˙ d σ = G + J
The strain-rate sensitivity (m) is defined as:
m = J G = ( l n σ ) ( l n ε ˙ )
m influences the energy distribution between G and J. BCC metals exhibit higher m values due to the limited mobility of screw dislocations, leading to increased flow stress [239]. Jinzhu et al. [241] conducted experimental and numerical studies on WHA projectile impacts against alumina armor. Their simulations accurately predicted penetration depth and failure patterns, aligning well with experimental results.
Crystal plasticity-based simulations offer critical insight into the role of microstructure in the dynamic response of WHAs. These models incorporate grain-level anisotropy and account for variations in grain orientation and boundary strength, both of which significantly influence damage localization and spall resistance. By integrating cohesive zone models at grain and phase boundaries along with rate-dependent constitutive laws, the simulations successfully replicate experimentally observed, free-surface velocity profiles and spall fracture morphologies [242,243,244]. Experimental data show the spall strength of WHAs between 1.7 and 2.0 GPa, reflecting the influence of factors such as grain structure, binder distribution, and impact conditions [242]. Fracture initiates once microcracks grow to a critical size, typically through a combination of cleavage within W grains and separation along grain boundaries. During plate impact tests, a shock wave travels through the material and reflects off the free surface as a rarefaction wave, generating intense tensile stresses. These stresses are especially high at interfaces, where spall fracture is likely to begin. The slope of the stress–time trajectory, from the shocked to the released state, serves as an indicator of spall strength, which increases with higher impact velocities [244]. Microstructural features and loading direction play key roles in determining how damage evolves. Cracks tend to initiate along weaker paths, such as grain boundaries or soft matrix regions, before coalescing into larger fracture zones. Post-impact analysis consistently reveals both intergranular cracking and cleavage, confirming the dual fracture mechanisms in WHAs. Compared to DU, WHAs generally exhibit more complex damage patterns and higher resistance to spall failure [242,243].
Collectively, these studies demonstrate the critical role of validated constitutive modeling and multi-scale simulation in predicting WHA performance under extreme conditions. Models that incorporate microstructural detail, interfacial behavior, strain rate, and temperature sensitivity provide a robust framework for designing next-generation KEPs.

8. Conclusions and Future Work

WHAs have emerged as critical materials for KEPs, primarily due to their high density, strength, and ductility. This review has summarized recent developments in the processing, microstructural control, mechanical property enhancement, and deformation behavior of WHAs. Among the various consolidation techniques, LPS remains the most widely adopted due to its ability to achieve near-theoretical densification. Nevertheless, challenges such as W grain coarsening, elevated W-W contiguity, and heterogeneous microstructures continue to limit the mechanical performance of WHAs.
Considerable efforts have been directed toward optimizing powder preparation, sintering parameters, and alloy chemistry to address these limitations. Alloying additions such as Co, Mo, Re, and rare earth oxides (e.g., Y2O3, La2O3, CeO2) have shown promise in enhancing matrix strength, refining microstructures, and promoting shear localization, an essential factor for improving the self-sharpening behavior of WHAs. Moreover, post-sintering treatments such as hot isostatic pressing, heat treatment, and rotary swaging have demonstrated significant improvements in tensile strength, hardness, and impact resistance by eliminating porosity and refining the grain structure.
Despite these advancements, WHAs still exhibit inferior penetration efficiency compared to DU alloys, primarily due to their mushrooming failure mode upon impact. While high-strain-rate ductility and strength have improved through microstructural engineering, the intrinsic limitations associated with W high melting point, brittleness, and interfacial cohesion remain critical challenges.
Future research should focus on advancing WHA systems through the following directions:
  • Development of matrix compositions and alloying strategies that enhance thermal softening and shear localization without compromising ductility.
  • Design and integration of HEA or multi-principal-element alloy matrices for improved bonding and thermal stability.
  • Optimization of AM parameters to overcome defects such as porosity and cracking, enabling the fabrication of net-shape, high-performance WHA components.
  • Use of ODS alloys to promote fine grain size, inhibit crack propagation, and enhance high-temperature performance.
  • Implementation of data-driven approaches and modeling frameworks to predict phase evolution, densification behavior, and dynamic failure mechanisms under impact conditions.
By addressing these challenges, WHAs can be further engineered to approach or even surpass the terminal ballistic performance of DU alloys, thereby offering a safer, sustainable, and high-performance solution for next-generation penetrator applications.

Author Contributions

Conceptualization, R.P. and G.M.K.; Resources, G.M.K. and P.S.; Data curation, R.P.; Writing—original draft preparation, R.P.; Writing—review and editing, G.M.K. and P.S.; Supervision, G.M.K. and P.S. All authors have read and agreed to the published version of the manuscript.

Funding

The authors received no specific funding for this work.

Data Availability Statement

The raw data will be made available by the authors on request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Magier, M. The Conception of the Segmented Kinetic Energy Penetrators for Tank Guns. J. Appl. Mech. Trans. ASME 2010, 77, 051802. [Google Scholar] [CrossRef]
  2. Chen, P.; Guo, L.; Wang, C.; Jin, S.; Gao, H.; Wang, G.; Ma, Y.; He, Y.; He, Y. Preparation and Penetration Behavior of the Reactive Fine-Grained Tungsten Heavy Alloy. Int. J. Refract. Met. Hard Mater. 2023, 115, 106306. [Google Scholar] [CrossRef]
  3. Upadhyaya, A. Processing Strategy for Consolidating Tungsten Heavy Alloys for Ordnance Applications. Mater. Chem. Phys. 2001, 67, 101–110. [Google Scholar] [CrossRef]
  4. Gu, D.; Guo, M.; Zhang, H.; Sun, Y.; Wang, R.; Zhang, L. Effects of Laser Scanning Strategies on Selective Laser Melting of Pure Tungsten. Int. J. Extrem. Manuf. 2020, 2, 025001. [Google Scholar] [CrossRef]
  5. Pan, S.H.; Yao, G.C.; Cui, Y.N.; Meng, F.S.; Luo, C.; Zheng, T.Q.; Singh, G. Additive Manufacturing of Tungsten, Tungsten-Based Alloys, and Tungsten Matrix Composites. Tungsten 2023, 5, 1–31. [Google Scholar] [CrossRef]
  6. Şahin, Y. Recent Progress in Processing of Tungsten Heavy Alloys. J. Powder Technol. 2014, 2014, 764306. [Google Scholar] [CrossRef]
  7. Manikandan, R.; Raja Annamalai, A. Tungsten Heavy Alloys Processing via Microwave Sintering, Spark Plasma Sintering, and Additive Manufacturing: A Review. Processes 2022, 10, 2352. [Google Scholar] [CrossRef]
  8. German, R.M. Sintered Tungsten Heavy Alloys: Review of Microstructure, Strength, Densification, and Distortion. Int. J. Refract. Met. Hard Mater. 2022, 108, 105940. [Google Scholar] [CrossRef]
  9. Mudda, S.; Annamalai, A.R.; Muthuchamy, A. Additive Manufacturing of Pure Tungsten and Tungsten Heavy Alloys-State-of-the-Art Processes. Met. Mater. Int. 2025, 1–32. [Google Scholar] [CrossRef]
  10. Senthilnathan, N.; Raja Annamalai, A.; Venkatachalam, G. Sintering of Tungsten and Tungsten Heavy Alloys of W–Ni–Fe and W–Ni–Cu: A Review. Trans. Indian Inst. Met. 2017, 70, 1161–1176. [Google Scholar] [CrossRef]
  11. Goroch, O.; Gulbinowicz, Z.; Magier, M.; Bednarczyk, E.; Skoczylas, P.; Pankowski, Z.; Sweklej, P.; Zochowski, P.; Jedrzejewski, W. Development and Experimental Verification of the New WHA Sinters Intended for Kinetic Energy Projectiles. Contin. Mech. Thermodyn. 2023, 35, 2193–2205. [Google Scholar] [CrossRef]
  12. Arora, A.; Gopal Rao, V.G. Tungsten Heavy Alloy for Defence Applications. Mater. Technol. 2004, 19, 210–216. [Google Scholar] [CrossRef]
  13. Senthilnathan, N.; Annamalai, A.R.; Venkatachalam, G. Microstructure and Mechanical Properties of Spark Plasma Sintered Tungsten Heavy Alloys. Mater. Sci. Eng. A 2018, 710, 66–73. [Google Scholar] [CrossRef]
  14. Das, J.; Kiran, U.R.; Chakraborty, A.; Prasad, N.E. Hardness and Tensile Properties of Tungsten Based Heavy Alloys Prepared by Liquid Phase Sintering Technique. Int. J. Refract. Met. Hard Mater. 2009, 27, 577–583. [Google Scholar] [CrossRef]
  15. Jacobs, J.; Haque, A.; Kulkarni, A.; Singh, J.; Matson, L. Microstructure of Tungsten Metal Alloys Produced by Field Assisted Sintering Technology (FAST). Int. J. Refract. Met. Hard Mater. 2019, 84, 104976. [Google Scholar] [CrossRef]
  16. Yin, Y.; Wang, T.; Qin, S.; Wang, W.; Shi, Y.; Yu, H. Research on Microstructure, Mechanical Properties, and High-Temperature Stability of Hot-Rolled Tungsten Hafnium Alloy. Materials 2024, 17, 3663. [Google Scholar] [CrossRef]
  17. Buzi, L.; Yeh, M.; Yeh, Y.W.; Donaldson, O.K.; Patino, M.I.; Trelewicz, J.R.; Yao, N.; Doerner, R.; Koel, B.E. Deuterium and Helium Ion Irradiation of Nanograined Tungsten and Tungsten–Titanium Alloys. Nucl. Mater. Energy 2019, 21, 100713. [Google Scholar] [CrossRef]
  18. Liu, B.; Dong, H.; Li, P.; Ai, T.; Li, W. Study on Microstructure and Mechanical Properties of W-Ni3Al Alloys Prepared by Ball Milling and Two-Step Sintering. Mater. Res. Express 2021, 8, 036505. [Google Scholar] [CrossRef]
  19. Mao, R.; Xiang, D. Significantly Elevated Strength of W-Ni3Al Alloy by Adding Trace Boron Element. Int. J. Refract. Met. Hard Mater. 2023, 111, 106091. [Google Scholar] [CrossRef]
  20. Hu, L.; Xiang, D.; Mao, R.; Gui, Z. Microstructure Evolution and Mechanical Property of W-Ni3Al Alloy by Adding Minor Co Element. J. Alloys Compd. 2024, 976, 173286. [Google Scholar] [CrossRef]
  21. Ravi Kiran, U.; Panchal, A.; Prem Kumar, M.; Sankaranarayana, M.; Nageswara Rao, G.V.S.; Nandy, T.K. Refractory Metal Alloying: A New Method for Improving Mechanical Properties of Tungsten Heavy Alloys. J. Alloys Compd. 2017, 709, 609–619. [Google Scholar] [CrossRef]
  22. Kiran, U.R.; Panchal, A.; Sankaranarayana, M.; Nandy, T.K. Tensile and Impact Behavior of Swaged Tungsten Heavy Alloys Processed by Liquid Phase Sintering. Int. J. Refract. Met. Hard Mater. 2013, 37, 1–11. [Google Scholar] [CrossRef]
  23. Liu, J.; German, R.M. Microstructure Effect on Dihedral Angle in Liquid-Phase Sintering. Metall. Mater. Trans. A 2001, 32, 165–169. [Google Scholar] [CrossRef]
  24. Wu, Y.; German, R.M.; Marx, B.; Suri, P.; Bollina, R. Comparison of Densification and Distortion Behaviors of W-Ni-Cu and W-Ni-Fe Heavy Alloys in Liquid-Phase Sintering. J. Mater. Sci. 2003, 38, 2271–2281. [Google Scholar] [CrossRef]
  25. Gurwell, W.E. Solid-State Sintering of Tungsten Heavy Alloys. Mater. Manuf. Process 1994, 9, 1115–1126. [Google Scholar] [CrossRef]
  26. Akhtar, F. An Investigation on the Solid State Sintering of Mechanically Alloyed Nano-Structured 90W-Ni-Fe Tungsten Heavy Alloy. Int. J. Refract. Met. Hard Mater. 2008, 26, 145–151. [Google Scholar] [CrossRef]
  27. Prabhu, G.; Arvind Kumar, N.; Sankaranarayana, M.; Nandy, T.K. Tensile and Impact Properties of Microwave Sintered Tungsten Heavy Alloys. Mater. Sci. Eng. A 2014, 607, 63–70. [Google Scholar] [CrossRef]
  28. Zhou, C.; Yi, J.; Luo, S.; Peng, Y.; Li, L.; Chen, G. Effect of Heating Rate on the Microwave Sintered W-Ni-Fe Heavy Alloys. J. Alloys Compd. 2009, 482, L6. [Google Scholar] [CrossRef]
  29. Fan, Z.S.; Xiang, D.P.; Pan, Y.L.; Jiang, H. Effect of Two-Time Spark Plasma Sintering on Microstructure and Mechanical Properties of W–6Ni–4Mn Alloy. Mater. Sci. Eng. A 2019, 745, 300–306. [Google Scholar] [CrossRef]
  30. Choi, J.; Sung, H.M.; Roh, K.B.; Hong, S.H.; Kim, G.H.; Han, H.N. Fabrication of Sintered Tungsten by Spark Plasma Sintering and Investigation of Thermal Stability. Int. J. Refract. Met. Hard Mater. 2017, 69, 164–169. [Google Scholar] [CrossRef]
  31. Chen, W.; Shi, Y.; Dong, L.; Wang, L.; Li, H.; Fu, Y. Infiltration Sintering of WCu Alloys from Copper-Coated Tungsten Composite Powders for Superior Mechanical Properties and Arc-Ablation Resistance. J. Alloys Compd. 2017, 728, 196–205. [Google Scholar] [CrossRef]
  32. Fang, X.; Liu, J.; Wang, X.; Li, S.; Guo, W. Investigation on the Penetration Performance and “Self-Sharpening” Behavior of the 80W-14Cu-6Zn Penetrators. Int. J. Refract. Met. Hard Mater. 2016, 54, 237–243. [Google Scholar] [CrossRef]
  33. Caceres, P.G. Effect of Microstructure on the Abrasive Wear Properties of Infiltrated Tungsten Alloys. Mater. Charact. 2002, 49, 1–9. [Google Scholar] [CrossRef]
  34. Mohsin, I.U.; Gierl, C.; Danninger, H.; Momeni, M. Thermal De-Binding Kinetics of Injection Molded W-8%Ni-2%Cu. Int. J. Refract. Met. Hard Mater. 2011, 29, 729–732. [Google Scholar] [CrossRef]
  35. Zu, Y.S.; Lin, S.T. Optimizing the Mechanical Properties of Injection Molded W 4.9%Ni 2.1%Fe in Debinding. J. Mater. Process Technol. 1997, 71, 337–342. [Google Scholar] [CrossRef]
  36. Loewenhoff, T.; Antusch, S.; Pintsuk, G.; Rieth, M.; Wirtz, M. High Pulse Number Thermal Shock Testing of Tungsten Alloys Produced by Powder Injection Molding. Nucl. Mater. Energy 2019, 20, 100680. [Google Scholar] [CrossRef]
  37. Ye, H.; Liu, X.Y.; Hong, H. Fabrication of Metal Matrix Composites by Metal Injection Molding—A Review. J. Mater. Process Technol. 2008, 200, 12–24. [Google Scholar] [CrossRef]
  38. Suri, P.; German, R.M.; de Souza, J.P. Influence of Mixing and Effect of Agglomerates on the Green and Sintered Properties of 97W–2.1Ni–0.9Fe Heavy Alloys. Int. J. Refract. Met. Hard Mater. 2009, 27, 683–687. [Google Scholar] [CrossRef]
  39. Loh, N.H.; Tor, S.B.; Khor, K.A. Production of Metal Matrix Composite Part by Powder Injection Molding. J. Mater. Process Technol. 2001, 108, 398–407. [Google Scholar] [CrossRef]
  40. Xue, J.; Feng, Z.; Tang, J.; Tang, C.; Zhao, Z. Selective Laser Melting Additive Manufacturing of Tungsten with Niobium Alloying: Microstructure and Suppression Mechanism of Microcracks. J. Alloys Compd. 2021, 874, 159879. [Google Scholar] [CrossRef]
  41. Ahn, S.Y.; Kim, E.S.; Karthik, G.M.; Ramkumar, K.R.; Jeong, S.G.; Kim, R.E.; Gu, G.H.; Kim, H.S. Thickness Effect on the Microstructures, Mechanical Properties, and Anisotropy of Laser-Powder Bed Fusion Processed 316L Stainless Steel. J. Mater. Sci. 2022, 57, 18101–18117. [Google Scholar] [CrossRef]
  42. Kwon, J.; Karthik, G.M.; Estrin, Y.; Kim, H.S. Constitutive Modeling of Cellular-Structured Metals Produced by Additive Manufacturing. Acta Mater. 2022, 241, 118421. [Google Scholar] [CrossRef]
  43. Karthik, G.M.; Sathiyamoorthi, P.; Zargaran, A.; Park, J.M.; Asghari-Rad, P.; Son, S.; Park, S.H.; Kim, H.S. Novel Precipitation and Enhanced Tensile Properties in Selective Laser Melted Cu-Sn Alloy. Materials 2020, 13, 100861. [Google Scholar] [CrossRef]
  44. Xie, J.; Lu, H.; Lu, J.; Song, X.; Wu, S.; Lei, J. Additive Manufacturing of Tungsten Using Directed Energy Deposition for Potential Nuclear Fusion Application. Surf. Coat. Technol. 2021, 409, 126884. [Google Scholar] [CrossRef]
  45. Kim, E.S.; Ahn, S.Y.; Jeong, S.G.; Go, E.H.; Bae, J.W.; Hong, S.J.; Wang, J.; Lee, B.J.; Karthik, G.M.; Kim, H.S. Cellular Structure Assisted Enhanced Deformation-Induced Martensitic Transformation in Laser-Directed Energy Deposited Ferrous Medium-Entropy Alloy. Mater. Sci. Eng. A 2025, 922, 147631. [Google Scholar] [CrossRef]
  46. Jeong, S.G.; Ahn, S.Y.; Kim, E.S.; Karthik, G.M.; Baik, Y.; Seong, D.; Kim, Y.S.; Woo, W.; Kim, H.S. Effect of Substrate Yield Strength and Grain Size on the Residual Stress of Direct Energy Deposition Additive Manufacturing Measured by Neutron Diffraction. Mater. Sci. Eng. A 2022, 851, 143632. [Google Scholar] [CrossRef]
  47. Nigarura, S.; Shah, R.; Karhumaa, T.; Trasorras, J.L. Manufacturing Process and Mechanical Properties of BJ3DP Tungsten Heavy Alloy Components. Int. J. Refract. Met. Hard Mater. 2022, 108, 105954. [Google Scholar] [CrossRef]
  48. Bose, A.; Reidy, J.P.; Tuncer, N.; Jorgensen, L. Processing of Tungsten Heavy Alloy by Extrusion-Based Additive Manufacturing. Int. J. Refract. Met. Hard Mater. 2023, 110, 106021. [Google Scholar] [CrossRef]
  49. Bose, A.; Schuh, C.A.; Tobia, J.C.; Tuncer, N.; Mykulowycz, N.M.; Preston, A.; Barbati, A.C.; Kernan, B.; Gibson, M.A.; Krause, D.; et al. Traditional and Additive Manufacturing of a New Tungsten Heavy Alloy Alternative. Int. J. Refract. Met. Hard Mater. 2018, 73, 22–28. [Google Scholar] [CrossRef]
  50. Zarinejad, M.; Tong, Y.; Salehi, M.; Mu, C.; Wang, N.; Xu, Y.; Rimaz, S.; Tian, L.; Kuah, K.X.; Chen, X. Advancements and Perspectives in Additive Manufacturing of Tungsten Alloys and Composites: Challenges and Solutions. Crystals 2024, 14, 665. [Google Scholar] [CrossRef]
  51. Hu, B.; Cai, G.; Fu, J.; Xin, Y.; Zhang, D.; Yuan, Y.; Guo, S. Densification Behavior of Tungsten Alloy Powders during Hot Isostatic Pressing. Mater. Today Commun. 2022, 31, 103576. [Google Scholar] [CrossRef]
  52. Kovářík, O.; Čech, J.; Čapek, J.; Hajíček, M.; Klečka, J.; Siegl, J. Mechanical Properties of Forged Tungsten Heavy Alloys. Acta Polytech. CTU Proc. 2020, 27, 149–154. [Google Scholar] [CrossRef]
  53. Alam, M.E.; Wang, J.; Henager, C.H.; Setyawan, W.; Odette, G.R. The Effect of Hot Rolling on the Strength and Fracture Toughness of 90W–7Ni3Fe Tungsten Heavy Metal Alloys. Mater. Sci. Eng. A 2021, 824, 141738. [Google Scholar] [CrossRef]
  54. Liu, J.; Shukui, L.; Xiaoqing, Z.; Zhaohui, Z.; Haiyun, Z.; Yingchun, W. Adiabatic Shear Banding in a Tungsten Heavy Alloy Processed by Hot-Hydrostatic Extrusion and Hot Torsion. Scr. Mater. 2008, 59, 1271–1274. [Google Scholar] [CrossRef]
  55. Panchal, A.; Reddy, K.V.; Azeem, P.A.; Nandy, T.K.; Singh, A.K. Microstructure, Texture, Tensile Flow and Work Hardening Behavior of Tungsten Heavy Alloys in Swaged and Swaged + Aged Conditions. Metallogr. Microstruct. Anal. 2020, 9, 438–456. [Google Scholar] [CrossRef]
  56. Ravi Kiran, U.; Venkat, S.; Rishikesh, B.; Iyer, V.K.; Sankaranarayana, M.; Nandy, T.K. Effect of Tungsten Content on Microstructure and Mechanical Properties of Swaged Tungsten Heavy Alloys. Mater. Sci. Eng. A 2013, 582, 389–396. [Google Scholar] [CrossRef]
  57. Shen, J.; Campbell, L.; Suri, P.; German, R.M. Quantitative Microstructure Analysis of Tungsten Heavy Alloys (W–Ni–Cu) during Initial Stage Liquid Phase Sintering. Int. J. Refract. Met. Hard Mater. 2005, 23, 99–108. [Google Scholar] [CrossRef]
  58. Das, J.; Appa Rao, G.; Pabi, S.K. Microstructure and Mechanical Properties of Tungsten Heavy Alloys. Mater. Sci. Eng. A 2010, 527, 7841–7847. [Google Scholar] [CrossRef]
  59. Mondal, A.; Upadhyaya, A.; Agrawal, D. Microwave and Conventional Sintering of Premixed and Prealloyed Tungsten Heavy Alloys. Mater. Sci. Technol. Oct. 2008, 4, 5–9. [Google Scholar]
  60. Yang, Y.; Lianxi, H.; Erde, W. Microstructure and Mechanical Properties of a Hot-Hydrostatically Extruded 93W-4.9Ni-2.1Fe Alloy. Mater. Sci. Eng. A 2006, 435, 620–624. [Google Scholar] [CrossRef]
  61. Upadhyaya, A.; Tiwari, S.K.; Mishra, P. Microwave Sintering of W-Ni-Fe Alloy. Scr. Mater. 2007, 56, 5–8. [Google Scholar] [CrossRef]
  62. Zhou, C.; Yi, J.; Luo, S. Sintering High Tungsten Content W-Ni-Fe Heavy Alloys by Microwave Radiation. Met. Mater. Trans. A Phys. Met. Mater. Sci. 2014, 45, 455–463. [Google Scholar] [CrossRef]
  63. Kim, D.-K.; Lee, S.; Baek, W.H. Microstructural Study of Adiabatic Shear Bands Formed by High-Speed Impact in a Tungsten Heavy Alloy Penetrator. Mater. Sci. Eng. A 1998, 249, 197–205. [Google Scholar] [CrossRef]
  64. Zhang, J.; Liu, W.; Ma, Y.; Ye, X.; Wu, Y. An Improved Two-Stage Sintering Method for Tungsten Heavy Alloys: Conventional Solidphase Sintering Followed by Microwave Heating. Int. J. Mater. Res. 2015, 106, 80–82. [Google Scholar] [CrossRef]
  65. Li, Y.; Hu, K.; Li, X.; Ai, X.; Qu, S. Fine-Grained 93W-5.6Ni-1.4Fe Heavy Alloys with Enhanced Performance Prepared by Spark Plasma Sintering. Mater. Sci. Eng. A 2013, 573, 245–252. [Google Scholar] [CrossRef]
  66. Humail, I.S.; Akhtar, F.; Askari, S.J.; Tufail, M.; Qu, X. Tensile Behavior Change Depending on the Varying Tungsten Content of W-Ni-Fe Alloys. Int. J. Refract. Met. Hard Mater. 2007, 25, 380–385. [Google Scholar] [CrossRef]
  67. Lakshmi Prasad, B.S.; Raja Annamalai, A. Effect of Rhenium Addition on Tungsten Heavy Alloys Processed through Spark Plasma Sintering. Ain. Shams. Eng. J. 2021, 12, 2957–2963. [Google Scholar] [CrossRef]
  68. Panchal, A.; Reddy, K.V.; Azeem, P.A.; Nandy, T.K.; Singh, A.K. On the Flow and Work Hardening Behavior of Tungsten Heavy Alloy 92W-5.5Ni-2.5Fe. Int. J. Refract. Met. Hard Mater. 2020, 88, 105203. [Google Scholar] [CrossRef]
  69. Çalişkan, N.K.; Durlu, N.; Bor, Ş. Swaging of Liquid Phase Sintered 90W-7Ni-3Fe Tungsten Heavy Alloy. Int. J. Refract. Met. Hard Mater. 2013, 36, 260–264. [Google Scholar] [CrossRef]
  70. Kiran, U.R.; Sankaranarayana, M.; Nageswara Rao, G.V.S.; Nandy, T.K. Effect of Cobalt Addition on Microstructure and Mechanical Properties of Tungsten Heavy Alloys. Trans. Indian Inst. Met. 2017, 70, 615–622. [Google Scholar] [CrossRef]
  71. Kumari, A.; Prabhu, G.; Sankaranarayana, M.; Nandy, T.K. Effect of Solution Treatment Temperature and Cooling Rate on the Mechanical Properties of Tungsten Heavy Alloy. Mater. Sci. Eng. A 2017, 688, 225–236. [Google Scholar] [CrossRef]
  72. Damodaram, R.; Karthik, G.M.; Lalam, S.V. Microstructure and Mechanical Properties of a Rotary Friction Welded Tungsten Heavy Alloy. Mater. Test. 2019, 61, 209–212. [Google Scholar] [CrossRef]
  73. Ravi Kiran, U.; Kumar, J.; Kumar, V.; Sankaranarayana, M.; Nageswara Rao, G.V.S.; Nandy, T.K. Effect of Cyclic Heat Treatment and Swaging on Mechanical Properties of the Tungsten Heavy Alloys. Mater. Sci. Eng. A 2016, 656, 256–265. [Google Scholar] [CrossRef]
  74. Das, J.; Sankaranarayana, M.; Nandy, T.K. Serrated Flow Behavior in Tungsten Heavy Alloy. Mater. Sci. Eng. A 2015, 646, 75–81. [Google Scholar] [CrossRef]
  75. Das, J.; Rao, G.A.; Pabi, S.K.; Sankaranarayana, M.; Nandy, T.K. Thermo-Mechanical Processing, Microstructure and Tensile Properties of a Tungsten Heavy Alloy. Mater. Sci. Eng. A 2014, 613, 48–59. [Google Scholar] [CrossRef]
  76. Panchal, A.; Nandy, T.K. Effect of Composition, Heat Treatment and Deformation on Mechanical Properties of Tungsten Heavy Alloys. Mater. Sci. Eng. A 2018, 733, 374–384. [Google Scholar] [CrossRef]
  77. Das, J.; Appa Rao, G.; Pabi, S.K.; Sankaranarayana, M.; Sarma, B. Deformation Behaviour of a Newer Tungsten Heavy Alloy. Mater. Sci. Eng. A 2011, 528, 6235–6247. [Google Scholar] [CrossRef]
  78. Kumari, A.; Sankaranarayana, M.; Nandy, T.K. On Structure Property Correlation in High Strength Tungsten Heavy Alloys. Int. J. Refract. Met. Hard Mater. 2017, 67, 18–31. [Google Scholar] [CrossRef]
  79. Kumari, A.; Prabhu, G.; Nandy, T.K. Effect of Nickel to Cobalt Ratio on Mechanical Properties of Tungsten Heavy Alloys. J. Alloys Metall. Syst. 2023, 3, 100030. [Google Scholar] [CrossRef]
  80. Skoczylas, P.; Kaczorowski, M. The Influence of Co Addition on the Structure and Mechanical Properties of Tungsten Heavy Alloys. Materials 2022, 15, 9064. [Google Scholar] [CrossRef]
  81. Kunčická, L.; Kocich, R.; Klečková, Z. Effects of Sintering Conditions on Structures and Properties of Sintered Tungsten Heavy Alloy. Materials 2020, 13, 2338. [Google Scholar] [CrossRef] [PubMed]
  82. Skoczylas, P.; Goroch, O.; Gulbinowicz, Z.; Penkul, A. The Effect of Cold Swaging of Tungsten Heavy Alloy with the Composition W91-6Ni-3Co on the Mechanical Properties. Materials 2021, 14, 7300. [Google Scholar] [CrossRef]
  83. Shakunt, N.S.; Gouthama; Upadhyaya, A. Effect of Fe Addition in W-Ni-Cu Heavy Alloy Processed through Powder Metallurgy on Microstructure and Mechanical Properties. J. Alloys Compd. 2024, 970, 172578. [Google Scholar] [CrossRef]
  84. Rajaei, M.; Raoufi, M.; Mostaan, H.; Beiranvand, Z.M.; Naeimi, A. The Effect of Pressure and Temperature on the Properties of 88 W-10Ni-2Mn Tungsten Heavy Alloy Produced by Cold Press Method. Int. J. Refract. Met. Hard Mater. 2024, 124, 106781. [Google Scholar] [CrossRef]
  85. Manikandan, R.; Raja Annamalai, A. Effect of Mn Addition in W-Ni-Cu Heavy Alloy Processed through Hot Press Sintering Techniques on Microstructure, Microtexture and Grain Boundary Character Evolution. Mater. Charact. 2024, 210, 113808. [Google Scholar] [CrossRef]
  86. Ma, Y.; Tang, X.; Wang, X.; Zhang, M.; Hu, H.; Gong, P.; Wang, X. Preparation and Mechanical Properties of Tungsten-Particle-Reinforced Zr-Based Bulk-Metallic-Glass Composites. Mater. Sci. Eng. A 2021, 815, 141312. [Google Scholar] [CrossRef]
  87. Hu, H.; Jin, H.; Chi, J.; Du, Y.; Ma, Y. Behavior of Tungsten-Particle-Reinforced Zirconium-Based Bulk Metallic Glass Composites When Penetrating a Semi-Infinite Target. Intermetallics 2025, 177, 108601. [Google Scholar] [CrossRef]
  88. Zhou, S.; Liang, Y.J.; Zhu, Y.; Jian, R.; Wang, B.; Xue, Y.; Wang, L.; Wang, F. High Entropy Alloy: A Promising Matrix for High-Performance Tungsten Heavy Alloys. J. Alloys Compd. 2019, 777, 1184–1190. [Google Scholar] [CrossRef]
  89. Satyanarayana, P.V.; Sokkalingam, R.; Jena, P.K.; Sivaprasad, K.; Prashanth, K.G. Tungsten Matrix Composite Reinforced with Cocrfemnni High-Entropy Alloy: Impact of Processing Routes on Microstructure and Mechanical Properties. Metals 2019, 9, 992. [Google Scholar] [CrossRef]
  90. Panigrahi, A.; Acharya, T.S.; Sengupta, P.; Kumar, D.; Sarangi, L.; Kumar, N.; Debasish, D.; Suwas, S.; Basu, S.; Debata, M. Microstructure and Mechanical Properties of Novel Tungsten Heavy Alloys Prepared Using FeNiCoCrCu HEA as Binder. Mater. Sci. Eng. A 2022, 832, 142451. [Google Scholar] [CrossRef]
  91. Ma, H.; Shao, Y.; Shek, C.H. Microstructure, Grain Growth Behavior and Mechanical Properties of W-CoCuFeNi Tungsten Heavy Alloys Prepared by Infiltration. Int. J. Refract. Met. Hard Mater. 2021, 98, 105572. [Google Scholar] [CrossRef]
  92. Gong, X.; Fan, J.L.; Ding, F.; Song, M.; Huang, B.Y. Effect of Tungsten Content on Microstructure and Quasi-Static Tensile Fracture Characteristics of Rapidly Hot-Extruded W-Ni-Fe Alloys. Int. J. Refract. Met. Hard Mater. 2012, 30, 71–77. [Google Scholar] [CrossRef]
  93. Suryanarayana, C. Mechanical Alloying and Milling. Prog. Mater. Sci. 2001, 46, 1–184. [Google Scholar] [CrossRef]
  94. Eroglu, S.; Baykara, T. Effects of Powder Mixing Technique and Tungsten Powder Size on the Properties of Tungsten Heavy Alloys. J. Mater. Process Technol. 2000, 103, 288–292. [Google Scholar] [CrossRef]
  95. Chen, C.L.; Ma, S.H. Study on Characteristics and Sintering Behavior of W-Ni-Co Tungsten Heavy Alloy by a Secondary Ball Milling Method. J. Alloys Compd. 2018, 731, 78–83. [Google Scholar] [CrossRef]
  96. Abdallah, A.; Fayed, A.; Abdo, G.; Sallam, M. Effect of Cold Isostatic Pressing On The Physical and Mechanical Properties of Tungsten Heavy Alloys. J. Eng. Sci. Mil. Technol. 2017, 17, 1–12. [Google Scholar] [CrossRef]
  97. Shao, Y.; Yu, W.; Wu, J.; Ma, H. Effect of Sintering Temperatures on Grain Coarsening Behaviors and Mechanical Properties of W-NiTi Heavy Tungsten Alloys. Materials 2022, 15, 8035. [Google Scholar] [CrossRef]
  98. Srikanth, V.; Upadhyaya, G.S. Effect of Tungsten Particle Size on Sintered Properties of Heavy Alloys. Powder Technol. 1984, 39, 61–67. [Google Scholar] [CrossRef]
  99. Ravi Kiran, U.; Panchal, A.; Sankaranarayana, M.; Nageswara Rao, G.V.S.; Nandy, T.K. Effect of Alloying Addition and Microstructural Parameters on Mechanical Properties of 93% Tungsten Heavy Alloys. Mater. Sci. Eng. A 2015, 640, 82–90. [Google Scholar] [CrossRef]
  100. Cheng, Q.; Wang, Y.; Zhang, J.; Conejo, A.N.; Liu, Z. The Grain Growth and Grain Boundary Migrations during Solid-Phase Sintering of Fe2O3: Experiments and Simulations. Chem. Eng. Sci. 2022, 262, 118038. [Google Scholar] [CrossRef]
  101. Shi, R.; Wood, M.; Heo, T.W.; Wood, B.C.; Ye, J. Towards Understanding Particle Rigid-Body Motion during Solid-State Sintering. J. Eur. Ceram. Soc. 2021, 41, 211–231. [Google Scholar] [CrossRef]
  102. German, R.M. Sintering Trajectories: Description on How Density, Surface Area, and Grain Size Change. JOM 2016, 68, 878–884. [Google Scholar] [CrossRef]
  103. Gouvêa, D. Thermodynamic of Solid-State Sintering: Contributions of Grain Boundary Energy. J. Eur. Ceram. Soc. 2024, 44, 116677. [Google Scholar] [CrossRef]
  104. Ryu, J.; Hong, S.H.; Baek, W.H. Mechanical Alloying Process of 93W-5.6Ni-1.4Fe Tungsten Heavy Alloy. J. Mater. Process. Technol. 1997, 63, 292–297. [Google Scholar] [CrossRef]
  105. Bose, A.; Sadangi, R.; German, R.M. A Review on Alloying in Tungsten Heavy Alloys. In TMS 2012 141st Annual Meeting and Exhibition, Materials Processing and Interfaces; John Wiley & Sons: Hoboken, NJ, USA, 2012; Volume 1, p. 455. [Google Scholar]
  106. Price, G.H.S. Copper-Nickel-Tungsten Alloys Sintered with a Liquid Phase Present. J. Inst. Met. 1938, 62, 239–264. [Google Scholar]
  107. German, R.M.; Suri, P.; Park, S.J. Review: Liquid Phase Sintering. J. Mater. Sci. 2009, 44, 1–39. [Google Scholar] [CrossRef]
  108. Liu, J.; German, R.M. Densification and Shape Distortion in Liquid-Phase Sintering. Metall. Mater. Trans. A 1999, 30, 3211–3217. [Google Scholar] [CrossRef]
  109. Bollina, R.; German, R.M. Heating Rate Effects on Microstructural Properties of Liquid Phase Sintered Tungsten Heavy Alloys. Int. J. Refract. Met. Hard Mater. 2004, 22, 117–127. [Google Scholar] [CrossRef]
  110. Olevsky, E.A.; German, R.M.; Upadhyaya, A. Effect of Gravity on Dimensional Change during Sintering—II. Shape Distortion. Acta Mater. 2000, 48, 1167–1180. [Google Scholar] [CrossRef]
  111. Li, Z.B.; Zhang, H.; Zhang, G.H.; Chou, K.C. Fabrication and Characterization of Tungsten Heavy Alloys with High W Content by Powder Metallurgy. Met. Mater. Trans. A Phys. Met. Mater. Sci. 2022, 53, 1085–1098. [Google Scholar] [CrossRef]
  112. Saiprasad, M.; Atchayakumar, R.; Thiruppathi, K.; Raghuraman, S. Consolidation of Copper and Aluminium Powders by Spark Plasma Sintering. In IOP Conference Series: Materials Science and Engineering; Institute of Physics Publishing: Bristol, UK, 2016; Volume 149. [Google Scholar]
  113. Hu, K.; Li, X.Q.; Yang, C.; Li, Y.Y. Densification and Microstructure Evolution during SPS Consolidation Process in W-Ni-Fe System. Trans. Nonferrous Met. Soc. China 2011, 21, 493–501. [Google Scholar] [CrossRef]
  114. Pan, Y.; Xiang, D.; Wang, N.; Li, H.; Fan, Z. Mechanical Milling-Assisted Spark Plasma Sintering of Fine-Grained W-Ni-Mn Alloy. Materials 2018, 11, 1323. [Google Scholar] [CrossRef] [PubMed]
  115. Pan, Y.-L.; Ding, L.; Xiang, D.-P. ping Microstructure Evolution and Mechanical Properties of Spark Plasma Sintered W–Ni–Mn Alloy. Trans. Nonferrous Met. Soc. China 2017, 27, 1588–1593. [Google Scholar] [CrossRef]
  116. Singh, I.; Dabhade, V.V. Transient Liquid Phase Sintering in Spark Plasma Sintered Tungsten Heavy Alloys with Nb and Mo Additives. Int. J. Refract. Met. Hard Mater. 2024, 124, 106849. [Google Scholar] [CrossRef]
  117. Chaurasia, J.; Ayyapan, M.; Patel, P.; Rajan, R.A.A. Activated Sintering of Tungsten Heavy Alloy. Sci. Sinter. 2017, 49, 445–453. [Google Scholar] [CrossRef]
  118. Chuvil’deev, V.N.; Nokhrin, A.V.; Boldin, M.S.; Baranov, G.V.; Sakharov, N.V.; Belov, V.Y.; Lantsev, E.A.; Popov, A.A.; Melekhin, N.V.; Lopatin, Y.G.; et al. Impact of Mechanical Activation on Sintering Kinetics and Mechanical Properties of Ultrafine-Grained 95W-Ni-Fe Tungsten Heavy Alloys. J. Alloys Compd. 2019, 773, 666–688. [Google Scholar] [CrossRef]
  119. Mishra, R.R.; Sharma, A.K. Microwave-Material Interaction Phenomena: Heating Mechanisms, Challenges and Opportunities in Material Processing. Compos. Part A Appl. Sci. Manuf. 2016, 81, 78–97. [Google Scholar] [CrossRef]
  120. Singh, S.; Gupta, D.; Jain, V.; Sharma, A.K. Microwave Processing of Materials and Applications in Manufacturing Industries: A Review. Mater. Manuf. Process 2015, 30, 1–29. [Google Scholar] [CrossRef]
  121. Srinath, M.S.; Sharma, A.K.; Kumar, P. Investigation on Microstructural and Mechanical Properties of Microwave Processed Dissimilar Joints. J. Manuf. Process 2011, 13, 141–146. [Google Scholar] [CrossRef]
  122. Sharma, A.K.; Mishra, R.R. Role of Particle Size in Microwave Processing of Metallic Material Systems. Mater. Sci. Technol. 2018, 34, 123–137. [Google Scholar] [CrossRef]
  123. Upadhyaya, A.; Sethi, G. Effect of Heating Mode on the Densification and Microstructural Homogenization Response of Premixed Bronze. Scr. Mater. 2007, 56, 469–472. [Google Scholar] [CrossRef]
  124. Singh, S.; Bhaskar, R.; Narayanan, K.B.; Kumar, A.; Debnath, K. Development of Silicon Carbide (SiC)-Based Composites as Microwave-Absorbing Materials (MAMs): A Review. J. Eur. Ceram. Soc. 2024, 44, 7411–7431. [Google Scholar] [CrossRef]
  125. Siores, E.; Do Rego, M.D. Materials Processing Technology Microwave Applications in Materials Joining. J. Mater. Process. Technol. 1995, 48, 619–625. [Google Scholar] [CrossRef]
  126. Prabhu, G.; Chakraborty, A.; Sarma, B. Microwave Sintering of Tungsten. Int. J. Refract. Met. Hard Mater. 2009, 27, 545–548. [Google Scholar] [CrossRef]
  127. Liu, W.; Ma, Y.; Zhang, J. Properties and Microstructural Evolution of W-Ni-Fe Alloy via Microwave Sintering. Int. J. Refract. Met. Hard Mater. 2012, 35, 138–142. [Google Scholar] [CrossRef]
  128. Roy, R.; Agrawal, D.; Cheng, J.; Gedevanishvili, S. Full Sintering of Powdered-Metal Bodies in a Microwave Field. Nature 1999, 399, 668–670. [Google Scholar] [CrossRef]
  129. Prabhu, G.; Sankaranarayana, M.; Nandy, T.K. Microwave Sintering of Tungsten Heavy Alloys. J. Appl. Res. Technol. 2018, 16, 115–125. [Google Scholar] [CrossRef]
  130. Ma, Y.; Zhang, J.; Liu, W.; Yue, P.; Huang, B. Microstructure and Dynamic Mechanical Properties of Tungsten-Based Alloys in the Form of Extruded Rods via Microwave Heating. Int. J. Refract. Met. Hard Mater. 2014, 42, 71–76. [Google Scholar] [CrossRef]
  131. Liu, W.; Ma, Y.; Cai, Q. Consolidation of MA W-Ni-Fe Alloyed Powder by Microwave-Assisted Sintering. Mater. Sci. Appl. 2011, 2, 609–614. [Google Scholar] [CrossRef]
  132. Zhang, D.; Cai, Q.; Liu, J.; Li, R. Research on Process and Microstructure Formation of W-Ni-Fe Alloy Fabricated by Selective Laser Melting. J. Mater. Eng. Perform. 2011, 20, 1049–1054. [Google Scholar] [CrossRef]
  133. Wang, X.; Wraith, M.; Burke, S.; Rathbun, H.; DeVlugt, K. Densification of W-Ni-Fe Powders Using Laser Sintering. Int. J. Refract. Met. Hard Mater. 2016, 56, 145–150. [Google Scholar] [CrossRef]
  134. Stawovy, M.T.; Myers, K.; Ohm, S. Binder Jet Printing of Tungsten Heavy Alloy. Int. J. Refract. Met. Hard Mater. 2019, 83, 104981. [Google Scholar] [CrossRef]
  135. Iveković, A.; Omidvari, N.; Vrancken, B.; Lietaert, K.; Thijs, L.; Vanmeensel, K.; Vleugels, J.; Kruth, J.P. Selective Laser Melting of Tungsten and Tungsten Alloys. Int. J. Refract. Met. Hard Mater. 2018, 72, 27–32. [Google Scholar] [CrossRef]
  136. Li, C.; Wang, Y.; Ma, S.; Yang, X.; Li, J.; Zhou, Y.; Liu, X.; Tang, J.; Wang, X.; Le, G. Densification, Microstructural Evolutions of 90W-7Ni-3Fe Tungsten Heavy Alloys during Laser Melting Deposition Process. Int. J. Refract. Met. Hard Mater. 2020, 91, 105254. [Google Scholar] [CrossRef]
  137. Su, S.; Lu, Y. Densified W[Sbnd]Cu Composite Fabricated via Laser Additive Manufacturing. Int. J. Refract. Met. Hard Mater. 2020, 87, 105122. [Google Scholar] [CrossRef]
  138. Gu, D.; Dai, D.; Chen, W.; Chen, H. Selective Laser Melting Additive Manufacturing of Hard-to-Process Tungsten-Based Alloy Parts with Novel Crystalline Growth Morphology and Enhanced Performance. J. Manuf. Sci. Eng. Trans. ASME 2016, 138, 081003. [Google Scholar] [CrossRef]
  139. Li, R.; Shi, Y.; Liu, J.; Xie, Z.; Wang, Z. Selective Laser Melting W-10 Wt.% Cu Composite Powders. Int. J. Adv. Manuf. Technol. 2010, 48, 597–605. [Google Scholar] [CrossRef]
  140. Chen, H.; Zi, X.; Han, Y.; Dong, J.; Liu, S.; Chen, C. Microstructure and Mechanical Properties of Additive Manufactured W-Ni-Fe-Co Composite Produced by Selective Laser Melting. Int. J. Refract. Met. Hard Mater. 2020, 86, 105111. [Google Scholar] [CrossRef]
  141. Dorow-Gerspach, D.; Kirchner, A.; Loewenhoff, T.; Pintsuk, G.; Weißgärber, T.; Wirtz, M. Additive Manufacturing of High Density Pure Tungsten by Electron Beam Melting. Nucl. Mater. Energy 2021, 28, 101046. [Google Scholar] [CrossRef]
  142. Vrancken, B.; King, W.E.; Matthews, M.J. In-Situ Characterization of Tungsten Microcracking in Selective Laser Melting. In Procedia CIRP; Elsevier, B.V.: Fürth, Germany, 2018; Volume 74, pp. 107–110. [Google Scholar]
  143. Kim, R.E.; Karthik, G.M.; Amanov, A.; Heo, Y.U.; Jeong, S.G.; Gu, G.H.; Park, H.; Kim, E.S.; Lee, D.W.; Kim, H.S. Superior Gradient Heterostructured Alloys Fabricated by Laser Powder Bed Fusion via Annealing and Ultrasonic Nanocrystal Surface Modification. Scr. Mater. 2023, 230, 115422. [Google Scholar] [CrossRef]
  144. Karthik, G.M.; Kim, H.S. Heterogeneous Aspects of Additive Manufactured Metallic Parts: A Review. Met. Mater. Int. 2021, 27, 1–39. [Google Scholar] [CrossRef]
  145. Schwanekamp, T.; Müller, A.; Reuber, M.; Gobran, H.; Gdoura, N.; von Cetto, S. Investigations on Laser Powder Bed Fusion of Tungsten Heavy Alloys. Int. J. Refract. Met. Hard Mater. 2022, 109. [Google Scholar] [CrossRef]
  146. Iveković, A.; Montero-Sistiaga, M.L.; Vanmeensel, K.; Kruth, J.-P.; Vleugels, J. Effect of Processing Parameters on Microstructure and Properties of Tungsten Heavy Alloys Fabricated by SLM. Int. J. Refract. Met. Hard Mater. 2019, 82, 23–30. [Google Scholar] [CrossRef]
  147. Wang, Y.P.; Ma, S.Y.; Yang, X.S.; Zhou, Y.Z.; Liu, X.; Li, J.F.; Zhang, J.J.; Li, C.; Wang, X.Y.; Le, G.M.; et al. Microstructure and Strengthening Mechanisms of 90W–7Ni–3Fe Alloys Prepared Using Laser Melting Deposition. J. Alloys Compd. 2020, 838, 155545. [Google Scholar] [CrossRef]
  148. Zhong, M.; Liu, W.; Ning, G.; Yang, L.; Chen, Y. Laser Direct Manufacturing of Tungsten Nickel Collimation Component. J. Mater. Process Technol. 2004, 147, 167–173. [Google Scholar] [CrossRef]
  149. Svetlizky, D.; Zheng, B.; Steinberg, D.M.; Schoenung, J.M.; Lavernia, E.J.; Eliaz, N. The Influence of Laser Directed Energy Deposition (DED) Processing Parameters for Al5083 Studied by Central Composite Design. J. Mater. Res. Technol. 2022, 17, 3157–3171. [Google Scholar] [CrossRef]
  150. Wei, C.; Ye, H.; Zhao, Z.; Tang, J.; Shen, X.; Le, G.; Ye, K.; Le, F. Microstructure and Fracture Behavior of 90W-7Ni-3Fe Alloy Fabricated by Laser Directed Energy Deposition. J. Alloys Compd. 2021, 865, 158975. [Google Scholar] [CrossRef]
  151. Plotkowski, A.; Ferguson, J.; Stump, B.; Halsey, W.; Paquit, V.; Joslin, C.; Babu, S.S.; Marquez Rossy, A.; Kirka, M.M.; Dehoff, R.R. A Stochastic Scan Strategy for Grain Structure Control in Complex Geometries Using Electron Beam Powder Bed Fusion. Addit. Manuf. 2021, 46, 102092. [Google Scholar] [CrossRef]
  152. Ellis, E.A.I.; Sprayberry, M.A.; Ledford, C.; Hankwitz, J.P.; Kirka, M.M.; Rock, C.D.; Horn, T.J.; Katoh, Y.; Dehoff, R.R. Processing of Tungsten through Electron Beam Melting. J. Nucl. Mater. 2021, 555, 153041. [Google Scholar] [CrossRef]
  153. Wright, J. Additive Manufacturing of Tungsten via Selective Laser Melting and Electron Beam Melting; University of Sheffield: Sheffield, UK, 2019. [Google Scholar]
  154. Cordero, Z.C.; Dinwiddie, R.B.; Immel, D.; Dehoff, R.R. Nucleation and Growth of Chimney Pores during Electron-Beam Additive Manufacturing. J. Mater. Sci. 2017, 52, 3429–3435. [Google Scholar] [CrossRef]
  155. Yang, G.; Yang, P.; Yang, K.; Liu, N.; Jia, L.; Wang, J.; Tang, H. Effect of Processing Parameters on the Density, Microstructure and Strength of Pure Tungsten Fabricated by Selective Electron Beam Melting. Int. J. Refract. Met. Hard Mater. 2019, 84, 105040. [Google Scholar] [CrossRef]
  156. Doddapaneni, V.V.K.; Lee, K.; Aysal, H.E.; Paul, B.K.; Pasebani, S.; Sierros, K.A.; Okwudire, C.E.; Chang, C.H. A Review on Progress, Challenges, and Prospects of Material Jetting of Copper and Tungsten. Nanomaterials 2023, 13, 2303. [Google Scholar] [CrossRef]
  157. Vrancken, B.; Ganeriwala, R.K.; Matthews, M.J. Analysis of Laser-Induced Microcracking in Tungsten under Additive Manufacturing Conditions: Experiment and Simulation. Acta Mater. 2020, 194, 464–472. [Google Scholar] [CrossRef]
  158. Sidambe, A.T.; Tian, Y.; Prangnell, P.B.; Fox, P. Effect of Processing Parameters on the Densification, Microstructure and Crystallographic Texture during the Laser Powder Bed Fusion of Pure Tungsten. Int. J. Refract. Met. Hard Mater. 2019, 78, 254–263. [Google Scholar] [CrossRef]
  159. Hu, Z.; Zhao, Y.; Guan, K.; Wang, Z.; Ma, Z. Pure Tungsten and Oxide Dispersion Strengthened Tungsten Manufactured by Selective Laser Melting: Microstructure and Cracking Mechanism. Addit. Manuf. 2020, 36, 101579. [Google Scholar] [CrossRef]
  160. Wang, D.-Z.; Li, K.-L.; Yu, C.-F.; Ma, J.; Liu, W.; Shen, Z.-J. Cracking Behavior in Additively Manufactured Pure Tungsten. Acta Metall. Sin. 2019, 32, 127–135. [Google Scholar] [CrossRef]
  161. Zhou, Z.; He, D.; Tan, Z.; Wang, Y.; Shao, W.; Guo, X.; Zhou, Z.; Wu, X.; Cui, L.; Wang, B. Grain Structure and Cracks Behaviors of Tungsten with Different Geometrical Shapes and Support Structure Prepared by Laser Powder Bed Fusion. J. Manuf. Process. 2022, 82, 253–264. [Google Scholar] [CrossRef]
  162. Chen, J.; Li, K.; Wang, Y.; Xing, L.; Yu, C.; Liu, H.; Ma, J.; Liu, W.; Shen, Z. The Effect of Hot Isostatic Pressing on Thermal Conductivity of Additively Manufactured Pure Tungsten. Int. J. Refract. Met. Hard Mater. 2020, 87, 105135. [Google Scholar] [CrossRef]
  163. Rebesan, P.; Bonesso, M.; Gennari, C.; Dima, R.; Pepato, A.; Vedani, M. Tungsten Fabricated by Laser Powder Bed FusionTungsten Hergestellt Durch Laserpulverbett-Fusion. BHM Berg. Hüttenmänn. Monatshefte 2021, 166, 263–269. [Google Scholar] [CrossRef]
  164. Tan, C.; Zhou, K.; Ma, W.; Attard, B.; Zhang, P.; Kuang, T. Selective Laser Melting of High-Performance Pure Tungsten: Parameter Design, Densification Behavior and Mechanical Properties. Sci. Technol. Adv. Mater. 2018, 19, 370–380. [Google Scholar] [CrossRef]
  165. Ren, X.; Liu, H.; Lu, F.; Huang, L.; Yi, X. Effects of Processing Parameters on the Densification, Microstructure and Mechanical Properties of Pure Tungsten Fabricated by Optimized Selective Laser Melting: From Single and Multiple Scan Tracks to Bulk Parts. Int. J. Refract. Met. Hard Mater. 2021, 96, 105490. [Google Scholar] [CrossRef]
  166. Wang, D.; Yu, C.; Zhou, X.; Ma, J.; Liu, W.; Shen, Z. Dense Pure Tungsten Fabricated by Selective Laser Melting. Appl. Sci. 2017, 7, 430. [Google Scholar] [CrossRef]
  167. Zhou, X.; Liu, X.; Zhang, D.; Shen, Z.; Liu, W. Balling Phenomena in Selective Laser Melted Tungsten. J. Mater. Process Technol. 2015, 222, 33–42. [Google Scholar] [CrossRef]
  168. Morcos, P.; Elwany, A.; Karaman, I.; Arróyave, R. Review: Additive Manufacturing of Pure Tungsten and Tungsten-Based Alloys. J. Mater. Sci. 2022, 57, 9769–9806. [Google Scholar] [CrossRef]
  169. Ravi Kiran, U.; Sambasiva Rao, A.; Sankaranarayana, M.; Nandy, T.K. Swaging and Heat Treatment Studies on Sintered 90W-6Ni-2Fe-2Co Tungsten Heavy Alloy. Int. J. Refract. Met. Hard Mater. 2012, 33, 113–121. [Google Scholar] [CrossRef]
  170. Wensheng, L.; Yunzhu, M.A.; Baiyun, H. Influence of Minor Elements Additions on Microstructure and Properties of 93W-4⋅9Ni-2⋅1Fe Alloys. Bull. Mater. Sci. 2008, 31, 1–6. [Google Scholar]
  171. Wang, Z.; Yuan, Y.; Arshad, K.; Wang, J.; Zhou, Z.; Tang, J.; Lu, G.H. Effects of Tantalum Concentration on the Microstructures and Mechanical Properties of Tungsten-Tantalum Alloys. Fusion. Eng. Des. 2017, 125, 496–502. [Google Scholar] [CrossRef]
  172. Chen, B.; Li, Z.B.; Liu, J.R.; Zhang, G.H. Effect of Molybdenum Addition on Microstructure and Mechanical Properties of 90% Tungsten Heavy Alloys. Int. J. Refract. Met. Hard Mater. 2022, 106, 105868. [Google Scholar] [CrossRef]
  173. Hong, M.-H.; Noh, J.-W.; Hyung Baek, W.; Kim, E.-P.; Song, H.-S.; Lee, S. A Study on the Improvement of the Sintered Density of W-Ni-Mn Heavy Alloy. Met. Mater. Trans. B 1997, 28, 835–839. [Google Scholar] [CrossRef]
  174. Prabhu, G.; Arockia Kumar, R.; Nandy, T.K. Fine Tungsten Precipitates in the Matrix Phase and Their Influence on the Mechanical Properties of a Tungsten Heavy Alloy. Philos. Mag. 2021, 101, 2004–2023. [Google Scholar] [CrossRef]
  175. Kumari, A.; Sarkar, R.; Panchal, A.; Prabhu, G.; Nandy, T.K. The Effect of Fine W Precipitates in Matrix Phase on Mechanical Properties of Tungsten Heavy Alloys. J. Appl. Res. Technol. 2022, 20, 487–501. [Google Scholar] [CrossRef]
  176. Dong, Z.; Liu, N.; Ma, Z.; Liu, C.; Guo, Q.; Alothman, Z.A.; Yamauchi, Y.; Md Hossain, S.A.; Liu, Y. Microstructure Refinement in W-Y2O3 Alloy Fabricated by Wet Chemical Method with Surfactant Addition and Subsequent Spark Plasma Sintering. Sci. Rep. 2017, 7, 6051. [Google Scholar] [CrossRef]
  177. Park, S.; Kim, D.-K.; Lee, S.; Kim, D.-K.; Ryu, H.J.; Hong, S.H.; Ryu, H.J. Dynamic Deformation Behavior of an Oxide-Dispersed Tungsten Heavy Alloy Fabricated by Mechanical Alloying. Metall. Mater. Trans. A 2001, 32, 2011–2020. [Google Scholar] [CrossRef]
  178. Xiao, F.; Barriere, T.; Cheng, G.; Miao, Q.; Zuo, S.; Wei, S.; Xu, L. Extremely Uniform Nanosized Oxide Particles Dispersion Strengthened Tungsten Alloy with High Tensile and Compressive Strengths Fabricated Involving Liquid-Liquid Method. J. Alloys Compd. 2021, 878, 160335. [Google Scholar] [CrossRef]
  179. Sarkar, R.; Singh, V.; Kumar, S.; Rao, Y.V.; Ghosal, P.; Nandy, T.K. Structure and Orientation of an Intermetallic Phase in a W-Ni-Co Alloy. Philos. Mag. 2019, 99, 1240–1258. [Google Scholar] [CrossRef]
  180. Hsu, C.-S.; Lin, S.-T. Effect of Molybdenum on Grain Growth of W-Mo-Ni-Fe Heavy Alloys. J. Mater. Sci. 2003, 38, 1543–1549. [Google Scholar] [CrossRef]
  181. German, R.M. The Microstructures of Liquid Phase Sintered Materials. In Sintering Technology; CRC Press: Boca Raton, FL, USA, 2020; pp. 213–220. [Google Scholar]
  182. Mao, R.; Xiang, D. Improved Mechanical Properties of Tungsten Alloy by Flaky Ni3Al and Trace B2O3 Synergistic Reinforcement. J. Alloys Compd. 2022, 923, 166467. [Google Scholar] [CrossRef]
  183. German, R.M. The Two-Dimensional Connectivity of Liquid Phase Sintered Microstructures. Metall. Mater. Trans. A 1987, 18, 909–914. [Google Scholar] [CrossRef]
  184. Lu, K.; German, R.M. Microstructure Analysis of Samples Sintered at Different Gravitational Conditions. J. Mater. Sci. 2010, 45, 4454–4461. [Google Scholar] [CrossRef]
  185. Liu, Y.; German, R.M.; Iacocca, R.G. Microstructure Quantification Procedures in Liquid-Phase Sintered Materials. Acta Mater. 1999, 47, 915–926. [Google Scholar] [CrossRef]
  186. German, R.M.; Bose, A.; Mani, S.S. Sintering Time and Atmosphere Influences on the Microstructure and Mechanical Properties of Tungsten Heavy Alloys. Metall. Trans. A 1992, 23, 211–219. [Google Scholar] [CrossRef]
  187. Fang, Z.; Patterson, B.R. Experimental Investigation of Particle Size Distribution Influence on Diffusion Controlled Coarsening. Acta Metall. Mater. 1993, 41, 2017–2024. [Google Scholar] [CrossRef]
  188. Yang, S.-C.; Mani, S.S.; German, R.M. The Effect of Contiguity on Growth Kinetics in Liquid-Phase Sintering. JOM 1990, 42, 16–19. [Google Scholar] [CrossRef]
  189. Greger, M.; Čížek, L.; Widomska, M. Structure and Mechanical Properties of Formed Tungsten Based Materials. J. Mater. Process. Technol. 2004, 157, 683–687. [Google Scholar] [CrossRef]
  190. Lai, X.; Wang, Z.; Qin, N. Comparative Studies of the Constitutive Models for Tungsten Heavy Alloy (95W-3.5Ni-1.5Fe) at High Strain Rates. Proc. Inst. Mech. Eng. C J. Mech. Eng. Sci. 2022, 236, 3636–3650. [Google Scholar] [CrossRef]
  191. Panchal, A.; Venugopal Reddy, K.; Azeem, P.A.; Sarkar, R.; Paradkar, A.; Nandy, T.K.; Singh, A.K. Effect of Ni/Fe Ratio on Microstructure, Tensile Flow and Work Hardening Behaviour of Tungsten Heavy Alloys in Heat Treated and Swaged Conditions. Philos. Mag. 2021, 101, 211–241. [Google Scholar] [CrossRef]
  192. Knipling, K.E.; Zeman, G.; Marte, J.S.; Kelly, S.M.; Kampe, S.L. Effect of Dissolved Tungsten on the Deformation of 70Ni-30Fe Alloys. Metall. Mater. Trans. A 2004, 35, 2821–2828. [Google Scholar] [CrossRef]
  193. Alam, M.E.; Odette, G.R. On the Remarkable Fracture Toughness of 90 to 97W-NiFe Alloys Revealing Powerful New Ductile Phase Toughening Mechanisms. Acta Mater. 2020, 186, 324–340. [Google Scholar] [CrossRef]
  194. Dinçer, O.; Pehlivanoğlu, M.K.; Çalişkan, N.K.; Karakaya, I.; Kalkanli, A. Processing and Microstructural Characterization of Liquid Phase Sintered Tungsten-Nickel-Cobalt Heavy Alloys. Int. J. Refract. Met. Hard Mater. 2015, 50, 106–112. [Google Scholar] [CrossRef]
  195. International Agency for Research on Cancer. Chlorinated Drinking-Water, Chlorination by-Products: Some Other Halogenated Compounds, Cobalt and Cobalt Compounds; IARC: Lyon, France, 1991; ISBN 9283212525. [Google Scholar]
  196. Cury, R.; Issartel, F.; Joubert, J.M.; Couque, H. Evolution of Cobalt-Free Tungsten Heavy Alloys for Kinetic Energy Penetrators. Powder Metall. 2013, 56, 347–350. [Google Scholar] [CrossRef]
  197. Natarajan, S.; Gopalan, V.; Annamalai, R.; Rajan, A.; Jen, C.-P. Materials Effect of Rare Earth Metals (Y, La) and Refractory Metals (Mo, Ta, Re) to Improve the Mechanical Properties of W-Ni-Fe Alloy—A Review. Materials 2021, 14, 1660. [Google Scholar] [CrossRef] [PubMed]
  198. Kim, D.-K.; Lee, S.; Song, H. Effect of Tungsten Particle Shape on Dynamic Deformation and Fracture Behavior of Tungsten Heavy Alloys. Metall. Mater. Trans. A 1998, 29, 1057–1069. [Google Scholar] [CrossRef]
  199. Ryu, H.J.; Hong, S.H.; Kim, D.-K.; Lee, S. Impact and Dynamic Deformation Behavior of Mechanically Alloyed Tungsten Heavy Alloy. Met. Mater. 1998, 4, 367–371. [Google Scholar] [CrossRef]
  200. Gong, X.; Fan, J.L.; Huang, B.Y.; Tian, J.M. Microstructure Characteristics and a Deformation Mechanism of Fine-Grained Tungsten Heavy Alloy under High Strain Rate Compression. Mater. Sci. Eng. A 2010, 527, 7565–7570. [Google Scholar] [CrossRef]
  201. Magness, L.S. High Strain Rate Deformation Behaviors of Kinetic Energy Penetrator Materials during Ballistic Impact. Mech. Mater. 1994, 17, 147–154. [Google Scholar] [CrossRef]
  202. Ramesh, K.T.; Coates, R.S. Microstructural Influences on the Dynamic Response of Tungsten Heavy Alloys. Metall. Mater. Trans. A 1992, 23, 2625–2630. [Google Scholar] [CrossRef]
  203. Bless, S.; Chau, R. Tensile Failure of Tungsten Alloys. In AIP Conference Proceedings; American Institute of Physics: College Park, MD, USA, 2006; Volume 845, pp. 603–606. [Google Scholar]
  204. Millett, J.C.F.; Bourne, N.K.; Rosenberg, Z.; Field, J.E. Shear Strength Measurements in a Tungsten Alloy during Shock Loading. J. Appl. Phys. 1999, 86, 6707–6709. [Google Scholar] [CrossRef]
  205. Dandekar, D.P.; Weisgerber, W.J. Shock Response of a Heavy Tungsten Alloy. Int. J. Plast. 1999, 15, 1291–1309. [Google Scholar] [CrossRef]
  206. Zhou, M.; Clifton, R.J. Dynamic Constitutive and Failure Behavior of a Two-Phase Tungsten Composite. J. Appl. Mech. 1997, 64, 487–494. [Google Scholar] [CrossRef]
  207. Wei, Z.; Yu, J.; Hu, S.; Li, Y. Influence of Microstructure on Adiabatic Shear Localization of Pre-Twisted Tungsten Heavy Alloys. Int. J. Impact Eng. 2000, 24, 747–758. [Google Scholar] [CrossRef]
  208. Bose, A.; German, R.N. Matrix Composition Effects on the Tensile Properties of Tungsten-Molybdenum Heavy Alloys. Metall. Trans. A Phys. Metall. Mater. Sci. 1990, 21, 1325–1327. [Google Scholar] [CrossRef]
  209. Wei, Q.; Jiao, T.; Ramesh, K.T.; Ma, E.; Kecskes, L.J.; Magness, L.; Dowding, R.; Kazykhanov, V.U.; Valiev, R.Z. Mechanical Behavior and Dynamic Failure of High-Strength Ultrafine Grained Tungsten under Uniaxial Compression. Acta Mater. 2006, 54, 77–87. [Google Scholar] [CrossRef]
  210. Xiaoqing, Z.; Shukui, L.; Jinxu, L.; Yingchun, W.; Xing, W. Self-Sharpening Behavior during Ballistic Impact of the Tungsten Heavy Alloy Rod Penetrators Processed by Hot-Hydrostatic Extrusion and Hot Torsion. Mater. Sci. Eng. A 2010, 527, 4881–4886. [Google Scholar] [CrossRef]
  211. Lee, W.-S.; Xiea, G.-L.; Lin, C.-F. The Strain Rate and Temperature Dependence of the Dynamic Impact Response of Tungsten Composite. Mater. Sci. Eng. A 1998, 257, 256–267. [Google Scholar] [CrossRef]
  212. Zhou, M.; Clifton, R.J.; Needleman, A. Shear Band Formation in a W--Ni--Fe Alloy Under Plate Impact. Tungsten Tungsten Alloys 1992, 343–356. [Google Scholar]
  213. Zhou, M. Finite Element Simulations of Shear Localization in Plane Impact. J. Mech. Phy. Solid 1975, 23, 239–264. [Google Scholar]
  214. Couque, H.; Nicolas, G.; Altmayer, C. Relation between Shear Banding and Penetration Characteristics of Conventional Tungsten Alloys. Int. J. Impact Eng. 2007, 34, 412–423. [Google Scholar] [CrossRef]
  215. Dick, R.D.; Ramachandran, V.; Williams, J.D.; Armstrong, R.W.; Holt, W.H.; Mock, W., Jr. Dynamic Deformation of W7Ni3Fe Alloy via Reverse-Ballistic Impact. In Tungsten and Tungsten Alloys; The Metallurgical Society Inc.: Warrendale, PA, USA, 1991. [Google Scholar]
  216. Johnson, G.R. A Constitutive Model and Data for Metals Subjected to Large Strains, High Strain Rates and High Temperatures. In Proceedings of the 7th International Symposium on Ballistics, The Hague, The Netherlands, 19–21 April 1983. [Google Scholar]
  217. Lankford, J.; Couque, H.; Bose, A.; German, R. Dynamic Deformation and Failure of Tungsten Heavy Alloys; The Metallurgical Society Inc.: Warrendale, PA, USA, 2025. [Google Scholar]
  218. Zhou, M. The Growth of Shear Bands in Composite Microstructures. Int. J. Plast. 1998, 14, 733–754. [Google Scholar] [CrossRef]
  219. Wu, Z.; Zhang, Y.; Yang, C.M.; Liu, Q.J. Mechanical Properties of Tungsten Heavy Alloy and Damage Behaviors after Hypervelocity Impact. Rare Met. 2014, 33, 414–418. [Google Scholar] [CrossRef]
  220. Li, Z.; Zhao, S.; Alotaibi, S.M.; Liu, Y.; Wang, B.; Meyers, M.A. Adiabatic Shear Localization in the CrMnFeCoNi High-Entropy Alloy. Acta Mater. 2018, 151, 424–431. [Google Scholar] [CrossRef]
  221. Luo, R.; Huang, D.; Yang, M.; Tang, E.; Wang, M.; He, L. Penetrating Performance and “Self-Sharpening” Behavior of Fine-Grained Tungsten Heavy Alloy Rod Penetrators. Mater. Sci. Eng. A 2016, 675, 262–270. [Google Scholar] [CrossRef]
  222. Anwer, Z.; Umer, M.A.; Nisar, F.; Hafeez, M.A.; Yaqoob, K.; Luo, X.; Ahmad, I. Microstructure and Mechanical Properties of Hot Isostatic Pressed Tungsten Heavy Alloy with FeNiCoCrMn High Entropy Alloy Binder. J. Mater. Res. Technol. 2023, 22, 2897–2909. [Google Scholar] [CrossRef]
  223. Fang, X.; Liu, J.; Wang, X.; Li, S.; Zheng, L. Study on Improving “Self-Sharpening” Capacity of W–Cu–Zn Alloy by the Pressureless Infiltration Method. Mater. Sci. Eng. A 2014, 607, 454–459. [Google Scholar] [CrossRef]
  224. Nagireddi, S.; Majumdar, B.; Bonta, S.; Diraviam, A.B. High-Density Bulk Metallic Glasses and Their Composites for Kinetic Energy Penetrator Applications: Process, Structure and Properties. Trans. Indian Inst. Met. 2021, 74, 2117–2134. [Google Scholar] [CrossRef]
  225. Zhou, S.; Jian, R.; Liang, Y.J.; Zhu, Y.; Wang, B.; Wang, L.; Wang, L.; Ren, Y.; Xue, Y. High Susceptibility to Adiabatic Shear Banding and High Dynamic Strength in Tungsten Heavy Alloys with a High-Entropy Alloy Matrix. J. Alloys Compd. 2021, 859, 157796. [Google Scholar] [CrossRef]
  226. Karthik, G.M.; Panikar, S.; Ram, G.D.J.; Kottada, R.S. Additive Manufacturing of an Aluminum Matrix Composite Reinforced with Nanocrystalline High-Entropy Alloy Particles. Mater. Sci. Eng. A 2017, 679, 193–203. [Google Scholar] [CrossRef]
  227. Karthik, G.M.; Kim, Y.; Kim, E.S.; Zargaran, A.; Sathiyamoorthi, P.; Park, J.M.; Jeong, S.G.; Gu, G.H.; Amanov, A.; Ungar, T.; et al. Gradient Heterostructured Laser-Powder Bed Fusion Processed CoCrFeMnNi High Entropy Alloy. Addit. Manuf. 2022, 59, 103131. [Google Scholar] [CrossRef]
  228. Liu, X.F.; Tian, Z.L.; Zhang, X.F.; Chen, H.H.; Liu, T.W.; Chen, Y.; Wang, Y.J.; Dai, L.H. “Self-Sharpening” Tungsten High-Entropy Alloy. Acta Mater. 2020, 186, 257–266. [Google Scholar] [CrossRef]
  229. Sharma, V.; Namburu, S.A.S.; Lalwani, P.; Sagar, C.K.; Gupta, A.K. Constitutive Modelling and Processing Map Analysis of Tungsten Heavy Alloy (92.5 W-5.25Ni-2.25Fe) at Elevated Temperatures. Int. J. Refract. Met. Hard Mater. 2018, 76, 168–179. [Google Scholar] [CrossRef]
  230. Park, S.J.; Chung, S.H.; Johnson, J.L.; German, R.M. Finite Element Simulation of Liquid Phase Sintering with Tungsten Heavy Alloys. Mater. Trans. 2006, 47, 2745–2752. [Google Scholar] [CrossRef]
  231. Li, J.C.; Chen, X.W.; Huang, F.L. FEM Analysis on the “Self-Sharpening” Behavior of Tungsten Fiber/Metallic Glass Matrix Composite Long Rod. Int. J. Impact Eng. 2015, 86, 67–83. [Google Scholar] [CrossRef]
  232. Li, J.-C.; Chen, X.-W.; Huang, F.-L. Ballistic Performance of Tungsten Particle/Metallic Glass Matrix Composite Long Rod. Def. Technol. 2019, 15, 132–145. [Google Scholar] [CrossRef]
  233. Kennedy, C.; Murr, L.E. Comparison of Tungsten Heavy-Alloy Rod Penetration into Ductile and Hard Metal Targets: Microstructural Analysis and Computer Simulations. Mater. Sci. Eng. A 2002, 325, 131–143. [Google Scholar] [CrossRef]
  234. Hafızoğlu, H.; Durlu, N. Effect of Sintering Temperature on the High Strain Rate-Deformation of Tungsten Heavy Alloys. Int. J. Impact Eng. 2018, 121, 44–54. [Google Scholar] [CrossRef]
  235. Hafizoglu, H.; Durlu, N.; Konokman, H.E. Effects of Sintering Temperature and Ni/Fe Ratio on Ballistic Performance of Tungsten Heavy Alloy Fragments. Int. J. Refract. Met. Hard Mater. 2019, 81, 155–166. [Google Scholar] [CrossRef]
  236. Gong, X.; Fan, J.; Ding, F. Tensile Mechanical Properties and Fracture Behavior of Tungsten Heavy Alloys at 25-1100 °C. Mater. Sci. Eng. A 2015, 646, 315–321. [Google Scholar] [CrossRef]
  237. Tang, J.J.; Liang, Z.F.; Zhang, Y.W. A Two-Phase Constitutive Model with Damage for Tungsten Heavy Alloy. Int. J. Multiphysics 2021, 15, 211–233. [Google Scholar]
  238. Zhou, M.; Needleman, A.; Clifton, R.J. Finite Element Simulations of Shear Localization in Plate Impact. J. Mech. Phys. Solids 1994, 42, 423–458. [Google Scholar] [CrossRef]
  239. Sun, J.; Zhang, L.; Huang, Y.; Chen, B.; Zhou, J.; Liu, W.; Ma, Y. Strain Hardening Rate and Strain Rate Sensitivity Behavior of Bcc/Fcc-Dual-Phase Tungsten Heavy Alloy. Int. J. Refract. Met. Hard Mater. 2023, 116, 106363. [Google Scholar] [CrossRef]
  240. Scapin, M. Mechanical Characterization and Modeling of the Heavy Tungsten Alloy IT180. Int. J. Refract. Met. Hard Mater. 2015, 50, 258–268. [Google Scholar] [CrossRef]
  241. Jinzhu, L.; Liansheng, Z.; Fenglei, H. Experiments and Simulations of Tungsten Alloy Rods Penetrating into Alumina Ceramic/603 Armor Steel Composite Targets. Int. J. Impact Eng. 2017, 101, 1–8. [Google Scholar] [CrossRef]
  242. Xiang, Y.; Chen, S.; Li, Z.; Yu, Y.; Hu, J. Spallation Damage of Tungsten Heavy Alloy under Shock Loading. Mater. Lett. 2024, 360, 135977. [Google Scholar] [CrossRef]
  243. Clayton, J.D. Modeling Dynamic Plasticity and Spall Fracture in High Density Polycrystalline Alloys. Int. J. Solids Struct. 2005, 42, 4613–4640. [Google Scholar] [CrossRef]
  244. Addessio, F.L.; Johnson, J.N. A Constitutive Model for the Dynamic Response of Brittle Materials. J. Appl. Phys. 1990, 67, 3275–3286. [Google Scholar] [CrossRef]
Figure 1. (a) Schematic representation of microstructural evolution during LPS, beginning with loosely packed powder particles separated by pores. Upon heating, solid-state diffusion initiates neck formation, followed by the formation and spreading of a liquid phase that promotes particle rearrangement and densification. Continued thermal exposure leads to grain growth and pore elimination, facilitated by enhanced atomic mobility within the liquid. (b) SEM micrograph of a W-5 wt.% Ni alloy sintered at 1550 °C for 180 min, demonstrating grain coalescence (as shown by arrow) [8].
Figure 1. (a) Schematic representation of microstructural evolution during LPS, beginning with loosely packed powder particles separated by pores. Upon heating, solid-state diffusion initiates neck formation, followed by the formation and spreading of a liquid phase that promotes particle rearrangement and densification. Continued thermal exposure leads to grain growth and pore elimination, facilitated by enhanced atomic mobility within the liquid. (b) SEM micrograph of a W-5 wt.% Ni alloy sintered at 1550 °C for 180 min, demonstrating grain coalescence (as shown by arrow) [8].
Jmmp 09 00186 g001
Figure 2. Microstructural comparison of W-6Ni-4Mn alloy, SPS at 1200 °C using (a) one-step SPS [115] and (b) two-step SPS [29]. (c) Variation in relative density with sintering temperature for 90W-6Ni-4Mn alloys processed by both OTSPS and TTSPS methods [29].
Figure 2. Microstructural comparison of W-6Ni-4Mn alloy, SPS at 1200 °C using (a) one-step SPS [115] and (b) two-step SPS [29]. (c) Variation in relative density with sintering temperature for 90W-6Ni-4Mn alloys processed by both OTSPS and TTSPS methods [29].
Jmmp 09 00186 g002
Figure 3. Microstructures of 90W-6Ni-2Fe-2Co alloys sintered at 1470 °C with a heating rate of 20 °C/min: (a) Microwave-sintered for 60 min and (b) conventionally sintered for 120 min [27]. Microstructures of 90W-7Ni-3Fe alloys microwave-sintered at 1480 °C with a heating rate of 65 °C/min: (c) 0 min and (d) 10 min. (e) Graph showing the evolution of W grain size and contiguity as a function of the heating rate [28].
Figure 3. Microstructures of 90W-6Ni-2Fe-2Co alloys sintered at 1470 °C with a heating rate of 20 °C/min: (a) Microwave-sintered for 60 min and (b) conventionally sintered for 120 min [27]. Microstructures of 90W-7Ni-3Fe alloys microwave-sintered at 1480 °C with a heating rate of 65 °C/min: (c) 0 min and (d) 10 min. (e) Graph showing the evolution of W grain size and contiguity as a function of the heating rate [28].
Jmmp 09 00186 g003
Figure 4. (a) SEM image of an L-DED sample showing alternating sublayers with varying W particle volume fractions along the deposition direction. The solid and dashed line in (a) indicates the presence of alternate layers of high and low W content in the DED processed samples. (b) SEM micrograph of a reference LPS sample, revealing a fully dense and uniform microstructure. In both images, W particles appear bright and the matrix dark. (c) Tensile stress–strain curves comparing the mechanical response of the L-DED and reference LPS samples [147].
Figure 4. (a) SEM image of an L-DED sample showing alternating sublayers with varying W particle volume fractions along the deposition direction. The solid and dashed line in (a) indicates the presence of alternate layers of high and low W content in the DED processed samples. (b) SEM micrograph of a reference LPS sample, revealing a fully dense and uniform microstructure. In both images, W particles appear bright and the matrix dark. (c) Tensile stress–strain curves comparing the mechanical response of the L-DED and reference LPS samples [147].
Jmmp 09 00186 g004
Figure 5. Summary of over 60 data points from the literature showing the influence of the various alloying elements Fe, Cu, Mo, Re, Co, and Mn on the W content in the matrix phase of W-Ni-based WHAs.
Figure 5. Summary of over 60 data points from the literature showing the influence of the various alloying elements Fe, Cu, Mo, Re, Co, and Mn on the W content in the matrix phase of W-Ni-based WHAs.
Jmmp 09 00186 g005
Figure 6. Summary of over 200 data points from the literature showing the influence of W grain size (a) and contiguity (b) on the tensile strength of WHAs, considering different processing routes. (c) shows the relationship between dihedral angle and relative density.
Figure 6. Summary of over 200 data points from the literature showing the influence of W grain size (a) and contiguity (b) on the tensile strength of WHAs, considering different processing routes. (c) shows the relationship between dihedral angle and relative density.
Jmmp 09 00186 g006
Figure 7. Summary of over 250 data points compiled from the literature, illustrating the influence of processing routes on the mechanical properties of WHAs with varying W contents. (ac) show the effect of W content and processing method on (a) tensile strength, (b) elongation, and (c) bulk hardness. (df) highlight the role of microstructural features, including contiguity and dihedral angle, on mechanical performance: (d) tensile strength vs. contiguity, (e) elongation vs. contiguity, and (f) tensile strength vs. dihedral angle.
Figure 7. Summary of over 250 data points compiled from the literature, illustrating the influence of processing routes on the mechanical properties of WHAs with varying W contents. (ac) show the effect of W content and processing method on (a) tensile strength, (b) elongation, and (c) bulk hardness. (df) highlight the role of microstructural features, including contiguity and dihedral angle, on mechanical performance: (d) tensile strength vs. contiguity, (e) elongation vs. contiguity, and (f) tensile strength vs. dihedral angle.
Jmmp 09 00186 g007
Table 2. Summary of sintering conditions used for LPS of WHAs.
Table 2. Summary of sintering conditions used for LPS of WHAs.
Composition (wt.%)Sintering Cycle and Sintering AtmosphereRef.
93W-4.9Ni-2.1Fe,
93W-4.2Ni-1.2Fe-1.6Co,
93W-4.9Ni-1.9Fe-0.2Re
Sintering temperature: 1480 °C
Holding time: 2 h in H2
Heating rate: 8 °C/min
Cooling rate: 1 °C/min
[99]
89.75W-6Ni-2Fe-2Co-0.25MoPre-sintering temperature: 1300 °C/1 h
Final sintering temperature: 1480 °C
Holding time: 75 min in H2
Heating rate: 3–5 °C/min
[78]
94.9W-3.4Ni-1.7CuSintered temperature: 1500 °C
Holding time: 1.5 h in H2
[14]
90W-5.7Ni-3.8Cu-0.5FeDegassing temperature: 1000 °C/20 min
Final sintering temperature: 1400 °C
Holding time: 1 h in H2
[83]
93W-6Ni-1CoSintering temperature: 1525 °C
Holding time: 30 min in H2
[81]
88W-xNi-yCu;
(x = 8.94–9.60, y = 2.4–3.06)
Debinding temperature: 500 °C/1 h/5 °C/min
Pre-sintering: 500–1000 °C/1 h/5 °C/min/AC
Final sintering temperature: 1412 °C
Holding time: 1 to 10 min in H2.
[57]
93W-4.9Ni-2.1FeSintering temperature: 1493 °C
Holding time: 3 h
[63]
91W-6Ni-3CoSintering temperature: 1520 °C
Holding time: 20 min in H2
[82]
90W-7Ni-3FePre-sintering: 1350 °C/15 °C/min/H2/15 min
Then, till 1350–1450 °C/10 °C/min
Final sintering: 1510 °C/5 °C/min/vacuum
[69]
88W-8.4Ni-3.6FeSintering: 1500 °C/30 min in H2/10 °C/min up to 800 °C, and 5 °C/min thereafter. During cooling, atmosphere changes from H2 to Ar.[66]
93W-4.9Ni-2.1FePre-sintering: 1380 °C/2 h under H2/AC, then microwave sintering: 1480 °C/10 min/10 at.% H2 + 90 at.% N2/15 °C/min[64]
97W-2.1Ni-0.9Fe,
99W-0.7Ni-0.3Fe
Sintering: 1600 °C/3h/Ar-20 vol.% and H2-200 mL/min[111]
89W-7Ni-3Fe-1ReSintering temperature: 1480 °C
Holding time: 2 h in H2
[21]
92W-7.2Ni-0.8Co,
92W-6.4Ni-1.6Co,
92W-5.6Ni-2.4Co,
92W-4.8Ni-3.2Co,
92W-4Ni-4Co
Sintering temperature: 1520 °C
Holding: 20 min in H2
[80]
90W-6Ni-4Co,
90W-7Ni-3Co,
90W-8Ni-2Co
Pre-sintering temperature: 1300 °C/1 h
Final sintering: Between 1480 and 1490 °C
Holding time: 75 min in H2
[79]
94.9W-3.4Ni-1.7Cu,
92.5W-5Ni-2.5Cu,
96.1–2.8Ni-1.1Cu,
96W-2.8Ni-1.1Cu-0.1Fe,
93.5W-4Ni-2Fe-0.5Co,
90.5W-7.2Ni-1.8Fe-0.45Co-0.05Mo,
74W-8Ni-2Fe-16Mo
Sintering temperature: 1500 °C
Holding time: 2 h in H2
[14]
Table 3. Summary of the literature data on SPS parameters used in the processing of WHAs.
Table 3. Summary of the literature data on SPS parameters used in the processing of WHAs.
Composition (wt.%)Sintering ConditionReference
93W-5.6Ni-1.4FeSintering temperature: 1050–1410 °C
Pressure: 30 MPa
Heating rate: 45–270 °C/min
[113]
99W-1Ni-1Co-1Fe-1Mo-1CuSintering temperature: 1200 °C
Pressure: 20 MPa
Heating rate: 1200 °C/min
[117]
90W-6Ni-4MnSintering temperature: 1000 °C to 1200 °C
Pressure: 30 MPa
Heating rate: 100 °C/min
[114]
93W-5.6Ni-1.4FeSintering temperature: 1360–1430 °C
Pressure: 30 MPa
Heating rate: 10–380 °C/min
[65]
95W-3.5Ni-1.5FeSintering temperature: 900–1300 °C
Pressure: 50–70 MPa
Heating rate: 50–300 °C/min
[118]
90W-5.6Ni-2.4Fe-2CoSintering temperature: 1400 °C
Pressure: 30 MPa
Heating rate: 100 °C/min
[13]
90W-6Ni-4MnFirst-step solid-phase-sintered at 1000 °C/100 °C/min/6 min, second-step sintering: 1000–1200/200 °C/min/3 min[29]
90W-7Ni-3Fe,
84W-7Ni3Fe-6Mo,
84W-7Ni-3Fe-6Nb
Sintering temperature: 1150–1275 °C
Pressure: 30 MPa
Heating rate: 100 °C/min
[116]
88W-2Re-7Ni-3FeSintering temperature: 1100 °C
Pressure: 50 MPa
Heating rate: 100 °C/min
[67]
Table 4. Summary of the microwave sintering conditions used for the processing of WHAs.
Table 4. Summary of the microwave sintering conditions used for the processing of WHAs.
Composition (wt.%)Sintering ConditionReference
92.5W-6.4Ni-1.1FeSintering temperature: 1500 °C
Holding time: 20 min in H2
Heating rate: 20 °C/min
[61]
90W-7Ni-3FeSintering temperature: 1460 °C
Holding time: 1 h in H2
Heating rate: 20 °C/min
Susceptor: SiC
[27]
90W-7Ni-3FeSintering temperature: 1480 °C
Holding time: 10 min in H2
Heating rate: 10–112 °C/min
Susceptor: SiC
[28]
90W-6Ni-2Fe-2CoSintering temperature: 1470 °C
Holding time: 60 min in H2
Heating rate: 20 °C/min
Susceptor: SiC
[27]
93W-4.9Ni-2.1FeSintering temperature: 1480 °C
Holding time: 10 min in H2
Heating rate: 15 °C/min
[64]
90W-7Ni-3CuSintering temperature: 1300–1450 °C
Heating rate: 22 °C/min
[59]
93W-4.9Ni-2.1FeSintering temperature: 1510 °C and 1520 °C
Holding time: 1 h in H2
Heating rate: 20 °C/min
[129]
93W-4.9Ni-2.1FePre-sintering temperature: 800 °C
Sintering temperature: 1250–1500 °C
Holding time: 5 min in 10 vol.% H2 and 90 vol.% N2
Heating rate: 30 °C/min
[127]
93W-4.9Ni-2.1FeExtruded-rod-sintered at 1550 °C
Holding time: 30 min in 10 vol.% H2 and 90 vol.% N2
[130]
93W-4.9Ni-2.1FeSintering temperature: 1250–1500 °C
Holding time: 5 min in 10 vol.% H2 and 90 vol.% N2
Heating rate: 30
[131]
Table 5. Comparative overview of microstructural features and mechanical property improvements associated with different sintering techniques used for WHAs.
Table 5. Comparative overview of microstructural features and mechanical property improvements associated with different sintering techniques used for WHAs.
PropertyMicrowaveSPSConventional
Process TimeVery LowVery LowHigh
Grain SizeFineVery FineCoarse
Density>98.5%95–98%~98%
W Solubility in the MatrixLowLowHigh
W-W ContiguityLow-ModerateLowHigh
Matrix UniformityHighHighModerate
Hardness (HV)HighHighModerate
Tensile/Impact StrengthSuperiorHighLow
Distortion RiskLow to ModerateLow to ModerateModerate to High
Table 6. Summary of mechanical properties of WHAs following various post-processing treatments.
Table 6. Summary of mechanical properties of WHAs following various post-processing treatments.
Composition (wt.%)Processing ConditionYS (MPa)UTS (MPa)Elong. (%)Impact Strength (J/cm2)Work-Hardening Exponent (n)Ref.
90W-6Ni-4CoSwaged (0.36 true strain) with six intermediate heat treatments (800–900 °C/3 h and 1100–1200 °C/4 h)133413724190.002[79]
90W-7Ni-3CoSwaged (0.36 true strain) with six intermediate heat treatments (800–900 °C/3 h and 1100–1200 °C/4 h)13001339102810.003
90W-8Ni-2CoSwaged (0.36 true strain) with six intermediate heat treatments (800–900 °C/3 h and 1100–1200 °C/4 h)12361278132790.003
90W-7Ni-3CoSwaged (0.36 true strain) with eight intermediate heat treatments (800–900 °C/3 h and 1100–1200 °C/4 h)13411410132270.004
90W-8Ni-2CoSwaged (0.36 true strain) with eight intermediate heat treatments (800–900 °C/3 h and 1100–1200 °C/4 h)12471311142580.002
90W-6Ni-2Fe-2CoSwaged (0.71 true strain) with two intermediate heat treatments (1100 °C/1.5 h)144214615710.05[78]
89W-6Ni-2Fe-3CoSwaged (0.71 true strain) with two intermediate heat treatments (1100 °C/1.5 h)154015824700.04
89.5W-6Ni-2Fe-2Co-0.5MoSwaged (0.71 true strain) with two intermediate heat treatments (1100 °C/1.5 h)152215516530.05
89.75W-6Ni-2Fe-2Co-0.25MoSwaged (0.71 true strain) with two intermediate heat treatments (1100 °C/1.5 h)152215475470.05
90W-6Ni-1.5Fe-2.5CoSwaged (0.71 true strain) with two intermediate heat treatments (1100 °C/1.5 h)156115945730.04
90W-6Ni-1Fe-3CoSwaged (0.71 true strain) with two intermediate heat treatments (1100 °C/1.5 h)1575161261210.05
93W-4.9Ni-2.1FeSwaging to a 44% reduction with two intermediate heat treatments (1100 °C/2 h)125512863130.018[99]
93W-4.2Ni-1.2Fe-1.6CoSwaging to a 44% reduction with two intermediate heat treatments (1100 °C/2 h)1219128012420.044
93W-4.9Ni-1.9Fe-0.2ReSwaging to a 44% reduction with two intermediate heat treatments (1100 °C/2 h)134713805260.026
90W-7Ni-2Fe-1CoTotal of 41% deformation with one intermediate heat treatment (1150 °C/2 h)136014001065-[56]
93W-4.9Ni-1.4Fe-0.7CoTotal of 41% deformation with one intermediate heat treatment (1150 °C/2 h)14101435745-
95W-3.5Ni-1.0Fe-0.5CoTotal of 41% deformation with one intermediate heat treatment (1150 °C/2 h)14181420414-
Table 8. Summary of mechanical properties of WHAs with different sintering methods.
Table 8. Summary of mechanical properties of WHAs with different sintering methods.
Composition
(wt.%)
Sintering MethodContiguityW Particle SizeUTS (MPa)Elong. (%)RemarksRef.
92.5W-6.4Ni-1.1FeConventional sintering at 1500 °C0.3217.36423.5Microwave-sintered specimens give fine W grains, resulting in higher strength.[61]
92.5W-6.4Ni-1.1FeMicrowave sintering at 1500 °C0.429.480511.2
90W-7Ni-3FeConventional sintering at 1480 °C, heating rate of 5 °C/min--86219.7[28]
90W-7Ni-3FeMicrowave sintering at 1480 °C heating rate of 80 °C/min--92219.7
90W-7Ni-3FeMicrowave sintering at 1460 °C 0.26217409[27]
90W–7Ni-3FeConventional sintering at 1460 °C 0.51326505
90W-6Ni-2Fe-2CoMicrowave sintering at 1470 °C 0.302198916Microwave-sintered specimens give lower contiguity and higher ductility.
90W-6Ni-2Fe-2CoConventional sintering at 1460 °C0.62476824
90W-7Ni-3FeSSS at 1350 °C-2.21204-SSS gives higher tensile strength because of fine W grains.[26]
90W-7Ni-3FeSPS at 1100 °C, heating rate of 100 °C/min, pressure of 50 MPa--9302.91SPS gives higher tensile strength but lower ductility.[67]
Table 9. Summary of mechanical properties of WHAs with different alloying elements.
Table 9. Summary of mechanical properties of WHAs with different alloying elements.
Composition
(wt.%)
Sintering methodContiguityW Particle SizeUTS (MPa)Elong. (%)RemarksRef.
92W-5.6Ni-2.4FeSPS, sintered at 1400 °C0.6412.397512Increasing the Co content increases the strength. A total of 1 wt.% Co gives the highest tensile strength.[13]
91.5W-5.6Ni-2.4Fe-0.5CoSPS, sintered at 1400 °C0.5411.5696116
91W-5.6Ni-2.4Fe-1.0CoSPS, sintered at 1400 °C0.479.48150820
90.5W-5.6Ni-2.4Fe-1.5CoSPS, sintered at 1400 °C0.439.68133021
90W-5.6Ni-2.4Fe-2.0CoSPS, sintered at 1400 °C0.5011.1125618
93W-4.9Ni-2.1FeConventional sintering at 1480 °C0.535685817Adding Co gives a slightly higher strength. Re refines the W grains.[99]
93W-4.2Ni-1.2Fe-1.6CoConventional sintering at 1480 °C0.435486918
93W-4.9Ni-1.9Fe-0.2ReConventional sintering at 1480 °C0.484987618
93.5W-4Ni-2Fe-0.5CoConventional sintering at 1500 °C--85314Adding Mo improves ductility due to a larger volume fraction of matrix. [14]
90.5W-7.2Ni-1.8Fe-0.45Co-0.05MoConventional sintering at 1500 °C--95021
95W-3.5Ni-1.5CuConventional sintering at 1510 °C0.6606603Tensile properties and hardness of the W-Ni-Cu alloy are inferior due to the coarser W grains.[58]
96W–3Ni–1CuConventional sintering at 1510 °C0.7706603
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Patel, R.; Karthik, G.M.; Sharma, P. Processing, Microstructure, and Mechanical Behavior of Tungsten Heavy Alloys for Kinetic Energy Penetrators: A Critical Review. J. Manuf. Mater. Process. 2025, 9, 186. https://doi.org/10.3390/jmmp9060186

AMA Style

Patel R, Karthik GM, Sharma P. Processing, Microstructure, and Mechanical Behavior of Tungsten Heavy Alloys for Kinetic Energy Penetrators: A Critical Review. Journal of Manufacturing and Materials Processing. 2025; 9(6):186. https://doi.org/10.3390/jmmp9060186

Chicago/Turabian Style

Patel, Rajneesh, Gangaraju Manogna Karthik, and Pawan Sharma. 2025. "Processing, Microstructure, and Mechanical Behavior of Tungsten Heavy Alloys for Kinetic Energy Penetrators: A Critical Review" Journal of Manufacturing and Materials Processing 9, no. 6: 186. https://doi.org/10.3390/jmmp9060186

APA Style

Patel, R., Karthik, G. M., & Sharma, P. (2025). Processing, Microstructure, and Mechanical Behavior of Tungsten Heavy Alloys for Kinetic Energy Penetrators: A Critical Review. Journal of Manufacturing and Materials Processing, 9(6), 186. https://doi.org/10.3390/jmmp9060186

Article Metrics

Back to TopTop