Next Article in Journal
The Influence of Thermal and Mechanical Aging on the Flexural Properties of Conventional and 3D-Printed Materials Used in Occlusal Splints Manufacturing
Previous Article in Journal
Properties of Loose-Fill Insulation Made of Leaves
Previous Article in Special Issue
Effect of Initial Relative Density on Liquid-Phase Sintering Behaviors of Al Powder Using Al–Cu Eutectic Alloy Aid: In Situ Observations Using Tomography and Microscopy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Advanced Numerical Modeling of Powder Bed Fusion: From Physics-Based Simulations to AI-Augmented Digital Twins

Faculty of Metals Engineering and Industrial Computer Science, AGH University of Krakow, al. Mickiewicza 30, 30-059 Krakow, Poland
*
Author to whom correspondence should be addressed.
Materials 2026, 19(2), 426; https://doi.org/10.3390/ma19020426
Submission received: 4 December 2025 / Revised: 23 December 2025 / Accepted: 14 January 2026 / Published: 21 January 2026

Abstract

Powder bed fusion (PBF) is a widely adopted additive manufacturing (AM) process category that enables high-resolution fabrication across metals, polymers, ceramics, and composites. However, its inherent process complexity demands robust modeling to ensure quality, reliability, and scalability. This review provides a critical synthesis of advances in physics-based simulations, machine learning, and digital twin frameworks for PBF. We analyze progress across scales—from micro-scale melt pool dynamics and mesoscale track stability to part-scale residual stress predictions—while highlighting the growing role of hybrid physics–data-driven approaches in capturing process–structure–property (PSP) relationships. Special emphasis is given to the integration of real-time sensing, multi-scale modeling, and AI-enhanced optimization, which together form the foundation of emerging PBF digital twins. Key challenges—including computational cost, data scarcity, and model interoperability—are critically examined, alongside opportunities for scalable, interpretable, and industry-ready digital twin platforms. By outlining both the current state-of-the-art and future research priorities, this review positions digital twins as a transformative paradigm for advancing PBF toward reliable, high-quality, and industrially scalable manufacturing.

1. Introduction

Additive manufacturing (AM) has transitioned over the past three decades from a rapid prototyping tool to a disruptive manufacturing technology capable of producing functional, high-performance components across aerospace, biomedical, automotive, and energy sectors [1]. Unlike subtractive processes, AM builds parts layer by layer directly from digital models, enabling unprecedented design freedom, near-net-shape fabrication, and efficient material utilization. Within the family of AM processes, powder bed fusion (PBF) has emerged as one of the most advanced and industrially relevant methods for producing metallic components with exceptional geometric precision and tailored microstructures [2].
PBF encompasses laser-based processes such as selective laser melting (SLM) and selective laser sintering (SLS), grouped recently in the laser beam PBF (PBF-LB or LPBF) or electron beam-based methods like electron beam melting (EBM or PBF-EB). In these processes, thin layers of metallic powder are selectively melted or sintered using a focused energy source under a controlled atmosphere or vacuum. The layer-by-layer consolidation enables the fabrication of complex geometries—including lattice structures, internal channels, and graded materials—that are often impossible to achieve through conventional methods. Furthermore, PBF is increasingly used for processing advanced materials such as titanium alloys (Ti-6Al-4V), stainless steels (316L), nickel-based superalloys (Inconel 718, Hastelloy X), and high-entropy alloys, where microstructural control directly influences the final part’s mechanical and functional performance [3].
Despite these advantages, PBF remains a complex, highly nonlinear process governed by the interplay of multi-scale and multiphysics phenomena. Localized energy input results in steep thermal gradients, rapid solidification, and transient melt pool dynamics, while layer-wise powder deposition introduces stochastic variability in packing density and morphology. These factors can lead to undesirable defects, such as porosity, lack of fusion, keyholing, cracking, and anisotropic mechanical properties [4]. The strong coupling between process parameters, transient thermal history, microstructural evolution, and residual stress development makes process optimization challenging [1].
Modeling and simulation have become indispensable for understanding, predicting, and controlling these complex interactions. Early modeling efforts—based on empirical correlations and analytical heat transfer equations—provided valuable but limited insights, often applicable only under narrow process conditions. As computational capabilities advanced, physics-based numerical models emerged, offering deeper process fidelity. Techniques such as the finite element method (FEM) have been used for thermal and mechanical analysis, computational fluid dynamics (CFD) for melt pool flow and keyhole formation, discrete element method (DEM) for powder spreading and packing behavior, and phase-field and cellular automata (CA) models for predicting microstructure and texture formation during solidification [2]. These methods have been instrumental in elucidating the process–structure–property (PSP) relationships in PBF, enabling the virtual exploration of parameter space and targeted optimization of part performance.
However, even the most advanced physics-based simulations face significant practical limitations. They are computationally expensive, often requiring high-performance computing resources, and may be difficult to generalize across materials, machines, and building geometries. Furthermore, their predictive power is limited by uncertainties in material property data at extreme processing conditions and by the lack of integration with real-time manufacturing data [5].
The emergence of digital twins—dynamic virtual replicas integrating simulation, in situ monitoring, and machine learning (detailed in Section 6)—represents a paradigm shift toward real-time process control [6]. Unlike static models, digital twins enable adaptive parameter adjustment during fabrication, bridging the gap between prediction and manufacturing to achieve "first-time-right" production and accelerate qualification cycles [7]. This integration of modeling, sensing, and control unlocks new capabilities for microstructure tailoring and functional optimization of the final product [6].
This review traces the evolution of PBF modeling from classical physics-based simulations to AI-augmented digital twins, emphasizing how multi-scale frameworks, hybrid approaches, and sensor fusion are reshaping the understanding of process–structure–property. By systematically evaluating advances, limitations, and open challenges, we establish a roadmap for scalable, interpretable digital twins capable of driving high-quality PBF manufacturing and accelerating integration into next-generation additive manufacturing.

1.1. Novelty and Significance

The modeling of powder bed fusion (PBF) processes has been the subject of numerous recent reviews, each addressing specific aspects of the field but leaving critical gaps in integration and scope. Bayat et al. [8] provided comprehensive coverage of multi-scale and multiphysics simulations emphasizing finite element methods and thermal–mechanical coupling, yet offered limited treatment of data-driven approaches and digital twin integration. Zinovieva et al. [9] focused specifically on microstructure-informed mechanical modeling with detailed crystal plasticity frameworks, but did not encompass the complete process chain from powder spreading through solidification or address optimization and digital twin architectures. Chowdhury et al. [10] and Sarkar et al. [2] delivered extensive surveys of numerical techniques, including finite element, computational fluid dynamics, and discrete element methods, though both preceded the recent surge in hybrid physics–machine learning approaches and industrial digital twin implementations. In the digital twin domain, Ahsan et al. [11] and Ben Amor et al. [12] cataloged applications and architectural frameworks across additive manufacturing, but lacked deep integration with underlying physics-based modeling foundations and specific optimization strategies essential for PBF deployment. This review addresses these gaps by synthesizing the complete PBF modeling ecosystem within an integrated, process-chain-oriented framework that systematically connects powder spreading, laser powder interaction, melt pool dynamics, solidification, thermomechanical response, and defect formation. The work establishes explicit cross linkages between physics-based simulations, in situ monitoring and sensor fusion, digital twin architectures, and optimization frameworks—revealing how advances in one domain enable progress in others. A multi-dimensional evaluation framework is introduced and applied uniformly across all modeling approaches, assessing physics fidelity, validation level, computational cost, real-time capability, transferability, and industrial readiness, thereby transforming the review from a descriptive catalog into a decision-support tool for method selection. A comprehensive treatment of emerging hybrid physics–machine learning methodologies, multi-sensor data fusion, and adaptive control strategies is provided with a critical assessment of current limitations and industrial deployment pathways. The review concludes with a time-bounded research roadmap distinguishing near-term priorities (validation benchmarks, open-source frameworks, GPU-accelerated simulations) from mid-term goals (industrial-scale digital twins, multi-material modeling, uncertainty quantification for certification), explicitly connecting research advances to industrial needs and providing actionable guidance for advancing PBF toward certifiable, production-ready manufacturing.

1.2. Review Method

A systematic literature review was conducted targeting peer-reviewed publications primarily from 2020 to 2025, with the selective inclusion of foundational published works before 2020 (8% of the corpus). Five databases were queried (Scopus, Web of Science, ScienceDirect, IEEE Xplore, SpringerLink) using Boolean search terms combining PBF process variants (“powder bed fusion” OR “LPBF” OR “L-PBF” OR “selective laser melting”) AND modeling approaches (“simulation” OR “machine learning” OR “digital twin” OR “physics-informed” OR “process monitoring”).
Inclusion criteria: Peer-reviewed journal articles, conference proceedings, and reviews in English focused on PBF physics-based modeling, data-driven/ML frameworks, hybrid approaches, digital twins, or in situ sensing.
Exclusion criteria: Non-peer-reviewed content (preprints, patents, theses), studies on other AM processes (DED, binder jetting), and insufficient methodological detail.
After duplicate removal and screening, 406 publications were retained, representing current PBF modeling, AI integration, and digital twin research.
The screening and eligibility steps for study inclusion are detailed in the PRISMA flow diagram (Figure 1), which illustrates the number of records identified, screened, excluded, and finally included in the qualitative synthesis.
The analyzed corpus includes 408 peer-reviewed studies (2015–2025) that cover physics-based modeling, machine learning, and digital twin approaches in PBF. As shown in Table 1, LPBF dominates with titanium and nickel alloys being studied most frequently (87%). FEM and ML methods prevail, and nearly 70% of works include experimental validation.

2. Powder Bed Fusion in Materials Processing

The history of PBF began with the development of selective laser sintering (SLS) by Deckard at the University of Texas [13]. PBF technologies are classified into several categories: selective laser sintering (SLS), selective laser melting (SLM), direct metal laser sintering (DMLS), and electron beam melting (EBM) [14]. These processes are generally meant for rapid prototyping and rapid tooling. SLS was also known as solid freedom fabrication, layer manufacturing technology, selective laser reactive sintering, etc. [10].
Table 2 presents the major PBF variants used in additive manufacturing. Each technology offers distinct advantages and limitations based on its energy sources, processing environments, and material compatibilities.
Kruth et al. [15,16] classified PBF (SLS/SLM) processes based on the mechanisms that occur in the process into four categories:
  • Solid-state sintering: Mainly applied for consolidating ceramic powders;
  • Chemically induced binding: Not commonly used in commercial equipment, but it can be a feasible consolidation mechanism for polymers, metals, and ceramics;
  • Liquid phase sintering: Partial melting, the main mechanism in SLS for glasses, polymers, and ceramics;
  • Full melting: The basic mechanism of the SLM.
EBM differs from other LPBF processes by the energy source (electron beam vs. laser beam) applied to metal (primarily titanium alloys) to fuse metals by melting in a vacuum environment. The basic concept of PBF is similar to that of SLA. PBF uses a moving laser or electron beam to trace and selectively sinter/melt powdered polymer and/or metal composite materials into the three-dimensional part of the final product. The product is constructed in a chamber located on a movable platform that is adjusted in height equal to the thickness of the forming layer. Powder for PBF is added from the powder feed supply, for example, with a roller. The powder is deposited on top of the powder bed in the build chamber, which contains a previously sintered or melted–solidified layer of the product and the unprocessed powder. Then the new layer is melted or sintered with a laser/electron beam. Cycles are repeated until the final product is obtained (Figure 2). SLM (and EBM) consists of four basic processes, which take place sequentially or simultaneously. They include powder deposition; heating by a laser beam and melting; free flow of molten material among unmolten and semi-melted particles and over the solid basement; and solidification. The last three processes are accompanied by heat transfer among solids, liquids, and gases, and by heat conduction.
A wide range of materials are processed in PBF [17]: aluminum alloys [19,20,21], stainless steel [22,23,24,25], Ti-6Al-4V [18,26,27,28], nickel superalloys (e.g., Inconel 718/625) [29,30,31,32], multiple materials [33,34,35], functionally graded materials (FGMs) [36,37,38], high-entropy alloys [39], and biomaterials [40,41,42]. Materials produced by SLM and EBM processing exhibit excellent static mechanical properties, equivalent to or superior to those of traditionally manufactured counterparts. However, SLS demonstrates unprocessed particles, porosity, and weakness [43,44], warping and distortion [45], limited material properties, and residual stresses [46].
Table 3 summarizes the key physical and mechanical properties (e.g., melting point, thermal conductivity, particle morphology, etc.) that influence the processability and performance of materials in PBF methods.
PBF requires numerous post-processing/finishing operations, including conventional machining, laser polishing, and electropolishing. Manufactured parts generally require material removal from surfaces to establish dimensional tolerances and achieve the desired surface properties [50].
PBF technology has several significant advantages over other additive manufacturing technologies:
  • High precision and excellent surface finish: PBF processes produce parts with fine details and smoother surfaces compared to FDM or other methods without extensive post-processing, and are capable of achieving tight tolerances and intricate geometries;
  • Complex geometries and internal structures: Processes are highly complex, organic, and has hollow internal features, supporting the creation of lightweight structures such as lattices and topology optimizations;
  • Material diversity: Processes include high-performance alloys such as titanium, nickel-based superalloys, aluminum, and steels, enabling the production of functional, load-bearing, and wear-resistant parts in metals, as well as polymers, ceramics, glasses, and composites;
  • Suitable for functional and end-use parts: The technology has high mechanical properties, enabling the rapid production of tooling, aerospace parts, medical implants, and customized components;
  • Less support material required: Unlike FDM or SLA, PBF generally does not require extensive support structures because the powders themselves support overhangs and complex features during build;
  • The process goes through layer-by-layer material consolidation, especially in metal PBF processes such as SLM and EBM;
  • The process is capable of producing multiple parts simultaneously in a single build cycle;
  • Reduced waste and material efficiency: Only the material in the powder bed is melted or sintered, so excess powder can often be recycled, reducing material waste.
PBF is a multifunctional technology that is mainly used in applications where complex geometry, material performance, and customization are crucial, such as aerospace, medical implants, automotive components, tools, jewelry, and energy. Flexibility in complex geometries and experimental manufacturing designs enables material testing, design validation, and innovative structural concepts.
In aerospace and defense applications, quality, reliability, material consumption, production costs, manufacturing time, energy consumption, occupational health, waste, impact on life cycles, capital investment, and ultimately, political issues must be considered [51]. PBF is used to manufacture engine components, turbine blades, structural components, and light lattice structures. It reduces weight and obtains complex geometry for optimized aerodynamics and a high strength-to-weight ratio. In the automotive industry, PBF is used for custom prototype parts, tools, and final parts, such as frames, exhaust components, and lightweight structures. It can be used for rapid prototyping, custom components, and lightweight components to improve efficiency [52].
In medicine and dentistry, PBF overcomes the most economic, scientific, and technical challenges. The high customization, biocompatible materials, and complex internal features allow production of personalized implants (hip, knee, dental crowns and bridges), surgical guides, and instrumentation [53,54].
In energy, PBF is utilized in energy storage and conversion systems, as well as in hydrogen reduction reactions, fuel cells, batteries, and supercapacitors [55]. It is also used in components for oil and gas, as well as parts of nuclear reactors and turbines, utilizing complex high-temperature materials and intricate cooling channels [56].
PBF is widely used for tooling and molds. Short lead times, ability to produce conformal cooling channels, and reduced costs for small batches allow the manufacturing of complex molds, inserts, and fixtures. PBF is also used in jewelry and the fashion industries. Due to the great detail resolution, complex geometries, rapid prototyping, and intricate jewelry designs, PBF produces customized accessories.

3. Physics-Driven Simulation Throughout the PBF Workflow

3.1. Powder Spreading and Packing Dynamics

Powder spreading is the foundational step in PBF that critically influences layer quality, powder bed density, and downstream defect formation. Discrete element method (DEM) simulations have emerged as the primary computational tool for modeling particle–scale interactions during recoater blade or roller spreading [57,58,59]. Advanced DEM models now incorporate realistic particle shape effects, including spherical, satellite-decorated, and irregular morphologies obtained from X-ray-computed tomography scans, which significantly affect packing density (typically 55–65% for spherical powders versus 50–58% for irregular particles) and surface roughness [57,58]. Spherical powders exhibit superior flowability characterized by Hausner ratios of 1.1–1.2, while irregular particles show Hausner ratios exceeding 1.3, indicating poor flowability and increased susceptibility to powder bed defects [58,59].
Key spreading parameters systematically investigated through DEM include recoater speed (50–500 mm/s), layer thickness (20–100 μm), blade geometry (rigid versus flexible, angle of attack), and powder flowability metrics [57,58,60]. Recent studies demonstrate that spreading speed–layer thickness combinations create distinct packing regimes: slow spreading (50–150 mm/s) with thin layers (30–50 μm) yields dense, uniform beds (ρ = 0.62–0.64) with surface roughness Ra < 3 μm but reduces productivity, while fast spreading (>300 mm/s) causes particle ejection, streaking, and density gradients (ρ < 0.55) with roughness Ra > 6 μm [57,58]. The optimal range of 100–250 mm/s balances quality and throughput for typical metal powders with particle size d50 = 25–45 μm [57,59].
Temperature effects introduce additional complexity—preheating the powder bed (100–200 °C for Ti-6Al-4V, 700–850 °C for electron beam melting) alters inter-particle friction coefficients, cohesion forces, and spreading dynamics through thermal expansion and incipient sintering between particles [57,59]. Coupled DEM-CFD models further capture gas flow effects during spreading, revealing that high recoater speeds generate turbulent wakes that entrain particles and create voids [57,60]. Contact mechanics parameters critical for DEM accuracy include particle–particle friction (μpp = 0.2–0.5), particle–wall friction (μpw = 0.15–0.35), and coefficient of restitution (e = 0.3–0.6), typically calibrated against experimental measurements of angle of repose, flow rate, and packing density [57,58,59].
DEM challenges include computational cost (106–108 particles for realistic domains), parameter calibration from limited experimental data, and validation scarcity [57,58]. Multi-layer simulations incorporating thermal histories remain limited to millimeter-scale domains [57,61]. GPU acceleration, ML surrogates, and integration with melt pool models represent emerging directions for scalable predictions [57,58,59,61].
Figure 3 demonstrates a DEM simulation of powder spreading and packing behavior for Ti-6Al-4V powder. Particle velocity field (Figure 3a) during recoater blade advancement at 200 mm/s shows active spreading zone with velocities 0.1–0.5 m/s transitioning to the settled powder bed. Packing density distribution (Figure 3b) reveals typical values of ρ = 0.60–0.63 in well-packed regions with local voids (ρ ≈ 0.51) caused by particle agglomeration or turbulent flow effects. Particle morphology (Figure 3c) influences flowability and density (spherical powders: ρ = 0.62–0.64, Hausner ratio 1.15 indicating good flowability vs. irregular particles with satellites: ρ = 0.53–0.58, Hausner ratio 1.35 indicating poor flowability). The effect of recoater speed on surface roughness and packing quality (Figure 3d) shows surface roughness increasing from Ra ≈ 2.2 μm at 50 mm/s to Ra ≈ 8.8 μm at 500 mm/s, while packing density exhibits threshold behavior, maintaining ρ ≈ 0.64 below 150 mm/s then decreasing to ρ < 0.55 above 300 mm/s due to particle ejection, with an optimal range of 100–250 mm/s for 50 μm layer thickness [57,58,59].

3.1.1. Spreading Strategy Comparisons

Recoater mechanism selection—rigid blade, flexible blade, or roller—fundamentally influences powder bed quality through distinct particle interaction regimes [59,62]. Rigid blades achieve moderate packing densities (ρ = 0.58–0.62) with a surface roughness Ra = 4–7 μm and enable high speeds (>300 mm/s) but generate streaking defects on irregular substrates [63,64]. Flexible blades (elastomeric or compliant V-geometries) improve uniformity (ρ = 0.60–0.64, Ra = 3–5 μm) by conforming to substrate topography, though limited-to-medium speeds (150–250 mm/s) [62]. Rollers achieve comparable densities (ρ = 0.60–0.63, Ra = 3–6 μm) through rotational shear; counter-rotation effectively densifies cohesive powders by disrupting agglomerates, while co-rotation promotes ejection [62,65,66]. Optimal roller speeds (100–200 mm/s) balance shear-induced rearrangement against particle disturbance [67].
Key spreading mechanisms and their influence on powder layer quality are presented in Table 4.
Cohesive fine powders (d50 < 20 μm, Hausner > 1.3) benefit from counter-rotating rollers, while free-flowing coarse powders (d50 > 40 μm, Hausner < 1.2) perform adequately with rigid blades at high throughput [64,66].

3.1.2. Environmental Effects: Gravity and Temperature

Gravitational acceleration profoundly affects powder dynamics, which is particularly relevant for space-based manufacturing [68]. Reducing gravity from 9.81 m/s2 (Earth) to 1.62 m/s2 (lunar, 0.17 g) decreases packing density by 8–12% and increases surface roughness by 15–20% due to reduced compaction [68]. Under microgravity (g ≈ 106 m/s2), cohesive forces dominate gravitational effects, causing increased particle dispersion, prolonged airborne times, and thickness variation (±30–40% versus ±5–10% at 1 g) [69]. Space applications require modified strategies: increased recoater forces, controlled gas environments, and potential electrostatic guidance [68].
Temperature significantly alters spreading through competing mechanisms [70]. Preheating from ambient (20 °C) to 100–200 °C (laser PBF) or 700–850 °C (EBM) reduces inter-particle friction (μ: 0.30–0.35 → 0.15–0.25) via surface softening, enhancing flowability [70]. However, elevated temperatures simultaneously increase cohesion through enhanced van der Waals forces, moisture condensation (<150 °C), and incipient sintering (>500–600 °C for Ti-6Al-4V, ~0.4–0.5 Tm) [70]. Non-monotonic behavior results: moderate preheating (100–150 °C) can improve spreadability, while higher temperatures (>200 °C) degrade uniformity through agglomeration, increasing density heterogeneity (CV: 6–10% → 10–15%) [70,71].

3.1.3. Non-Spherical Particles and Recoater Effects

Particle morphology critically affects spreading through geometric interlocking and altered contact mechanics [72]. Spherical powders (sphericity ψ > 0.95) achieve superior flowability (Hausner 1.1–1.2), high packing density (ρ = 0.62–0.64), and low surface roughness (Ra < 3 μm), permitting aggressive spreading (250–300 mm/s) [73,74]. Irregular particles (ψ < 0.85) exhibit poor flowability (Hausner > 1.3), reduced density (ρ = 0.50–0.58), and increased roughness (Ra = 5–8 μm) due to mechanical interlocking and inefficient space-filling [72,75].
Recoater–morphology interactions amplify these effects: rigid blades encounter 2–3× higher forces with irregular powders, causing ejection and jamming [64,76]. Flexible blades partially mitigate issues through compliance but morphology effects persist [62]. Rollers show particular sensitivity—irregular particles experience increased pinching risk and fragmentation [67]. Counter-rotation partially overcomes interlocking via controlled shear but requires elevated forces [66]. Process optimization demands morphology specifications, emphasizing sphericity, parameter adjustments (reduced speed, increased force) for irregular powders, and potential conditioning (tumbling, spheroidization) [72,74].

3.1.4. Packing Density Metrics and Dependencies

Packing density (solid/total volume ratio) constitutes the primary spreading quality metric, influencing melting, porosity, and mechanical properties [77,78]. Global density for spherical powders under optimal conditions ranges from 0.60 to 0.64 (approaching random close packing limit 0.64); non-spherical particles yield 0.50–0.58 [72,73]. Local density heterogeneity (CV = 5–12% for well-spread layers) increases with suboptimal parameters: high speeds (>300 mm/s) elevate CV to 15–20% via ejection/streaking [79], and cohesive powders exhibit CV 12–18% from agglomeration [71].
Measurement approaches include DEM spatial binning (high-resolution 3D mapping), transmission X-ray imaging (~10 μm resolution, 2D projection), X-ray CT (5–20 μm voxels, 3D reconstruction), and gravimetry (global only) [78,80]. Optimal density window (0.58 < ρ < 0.64) balances material availability (ρ < 0.50 causes lack of fusion) against gas permeability (ρ > 0.68 promotes gas porosity) [77,81].
The effects of powder morphology, recoating, and process conditions on layer density are presented in Table 5.

3.1.5. DEM Fundamentals

The discrete element method (DEM) simulates granular systems by resolving the motion of individual particles through Newton’s equations of motion [59,73]:
Translational motion:
m i d v i d t = b F i j + F g , i + F d , i ,
Rotational motion:
I i d ω i d t = b T i j ,
where m i is the mass of particle i; v i and ω i are its translational and angular velocity vectors; F i j is the contact force exerted by particle j; F g , i is the gravitational force; F d , i is the fluid drag; I i is the moment of inertia; T i j is the torque due to contact interactions.
Contact forces are decomposed into normal and tangential components using the Hertz–Mindlin contact model [83]:
Normal force:
F n = 4 3 E R δ n 3 / 2 γ n v n , rel ,
Tangential force:
F t = min μ F n , k t δ t ,
F n and F t denote the normal and tangential contact forces, respectively; E and R are the effective Young’s modulus and effective radius of contact; δ n and δ t are the normal and tangential displacements; γ n is the normal damping coefficient; v n , r e l is the relative normal velocity; k t is the tangential stiffness; and μ is the coefficient of friction.
The numerical timestep is limited by the Rayleigh stability criterion:
Δ t π R 0.163 ν + 0.877 ρ G ,
where Δ t is the integration timestep, R is the particle radius, ρ is the material density, G is the shear modulus, and ν is Poisson’s ratio. For metallic powders, this typically yields Δ t values on the order of 10 8 10 7 s, requiring millions of timesteps to simulate milli-second-scale dynamics [59,61]. GPU acceleration achieves 100–1000× speed-up, enabling 107–108 particle simulations approaching realistic domains [61].

3.2. Laser–Powder Interaction and Energy Absorption

Accurate prediction of energy absorption during laser–powder bed fusion requires sophisticated ray-tracing models that account for the complex optical behavior of powder beds. The absorptivity of metal powders during LPBF processes is significantly higher (0.3–0.7 for common alloys) than that of solid surfaces due to multiple reflections and scattering between particles [84,85]. Ray-tracing simulations explicitly model laser beam propagation through the powder bed, tracking individual photon paths as they interact with particle surfaces, accounting for reflection, refraction, and absorption at each interface [84,85,86]. These models reveal that energy absorption is strongly influenced by powder morphology, particle size distribution, and packing density, with smaller particles and irregular morphologies generally increasing absorptivity through enhanced light trapping [87,88,89].
The temperature-dependent nature of absorptivity introduces additional complexity, as material optical properties evolve during heating and melting. Studies demonstrate that absorptivity increases with temperature due to changes in surface oxidation and electronic structure, necessitating dynamic updating of absorption coefficients in high-fidelity simulations. Furthermore, the transition from powder to melt pool significantly alters the absorption mechanism, with molten material exhibiting reduced absorptivity compared to the powder bed, though still higher than that of bulk solid due to surface roughness and Fresnel absorption effects. Ray-tracing models coupled with thermal simulations enable prediction of these transitions and their impact on melt pool formation, though computational cost remains a limiting factor for part-scale applications.
Absorptivity exhibits temperature dependence due to surface oxidation and changes in electronic structure, requiring dynamic updating in high-fidelity simulations [86,87]. The powder-to-melt pool transition alters absorption mechanisms, with molten material showing reduced absorptivity compared to powder beds [84,86]. Ray-tracing models coupled with thermal simulations capture these transitions, although computational cost limits part-scale applications [84,85]. Quantitative relationships between absorptivity variation and thermomechanical response are detailed in Section 3.5.2.
The powder bed characteristics exert a profound influence on laser–powder interaction and subsequent melt pool behavior. Discrete element method (DEM) simulations of powder spreading reveal that packing density varies between 0.55 and 0.65 for metal powders, with spatial heterogeneities arising from particle size distribution, spreading velocity, and layer thickness [87,88,89]. These packing variations directly affect local absorptivity and heat transfer, creating non-uniform thermal fields that influence melt pool stability and defect formation [87,88]. Particle morphology—including sphericity, surface roughness, and satellite particles—further modulates energy absorption through altered scattering patterns and inter-particle contact mechanics [87,88,89].
Experimental and computational studies demonstrate that powder bed surface roughness, typically ranging from 10 to 50 μm for common LPBF powders, affects the effective beam–material interaction area and absorption efficiency [88,89]. Irregular or agglomerated particles increase surface roughness, enhancing absorptivity but potentially introducing process instabilities through non-uniform melting [87,89]. The coupling between powder bed structure and laser absorption is particularly critical at low energy densities, where incomplete melting of poorly packed regions leads to lack-of-fusion defects [87,88]. Advanced ray-tracing models integrated with DEM-predicted powder bed geometries provide insights into these phenomena, although validation remains challenging due to the difficulty of in situ measurement of powder-scale absorption [84,85,87].
The modeling of laser–powder interaction spans multiple computational frameworks, each with distinct capabilities and limitations. Particle-based methods such as smoothed particle hydrodynamics (SPH) and the lattice Boltzmann method (LBM) naturally capture discrete powder behavior and dynamic free surfaces during melting, enabling detailed prediction of absorption and heat transfer at the particle scale [84,90,91,92]. However, these approaches face a significantly high computational cost when extended to multi-layer or part-scale simulations, often requiring simplifications in optical models or powder bed resolution [84,91]. Mesh-based finite element and finite volume methods offer computational efficiency for continuum-scale thermal analysis but struggle to represent discrete powder morphology accurately, typically resorting to effective medium approximations that may overlook local absorption heterogeneities [2,93,94].
Hybrid modeling strategies that combine ray-tracing for absorption with continuum thermal solvers represent a pragmatic compromise, enabling prediction of temperature fields with reasonable fidelity at manageable computational costs [84,92]. Model validation remains challenging due to the difficulty in directly measuring absorptivity and energy distribution during processing [86,87]. Validation typically relies on indirect metrics: melt pool dimensions, pyrometry/thermography, or post-process fusion characterization [85,86,95]. Limited experimental datasets encompassing powder bed structure, absorption, and thermal histories constrain confidence across diverse materials [84,85,86].
Table 6 highlights that model selection must balance physical fidelity with computational tractability based on the specific application. Ray-tracing models provide the highest accuracy for absorption predictions but are computationally prohibitive for routine process optimization. Effective medium approximations enable part-scale simulations but sacrifice local accuracy, making them suitable for thermal management and distortion control rather than defect-critical applications. DEM-coupled approaches bridge the powder bed structure and absorption, offering insights into the effects of powder quality but requiring careful calibration. Hybrid particle-based methods represent the state-of-the-art for fundamental studies but remain impractical for industrial deployment due to computational demands. Future advances in reduced-order modeling and machine learning surrogates may enable more efficient high-fidelity predictions, though experimental validation across diverse materials and conditions remains a persistent challenge [84,85,86]. The interplay between absorptivity variation and residual stress development is examined in Section 3.5.2, where thermomechanical coupling is addressed.

3.3. Melt Pool Formation and Dynamics

Melt pool formation in laser–powder bed fusion is governed by thermal–fluid interactions spanning multiple scales. High-fidelity CFD simulations incorporating energy conservation, momentum transport, and phase change provide detailed predictions [91,93,113,114]. Temperature gradients (104–106 K/m) and rapid solidification (104–106 K/s cooling rates) create transient conditions challenging numerical accuracy [113,114,115]. Surface tension gradients (∂γ/∂T ≈ −0.3 mN/m·K) generate a Marangoni convection that drives fluid from hot to cold regions, enhancing penetration and width; neglecting Marangoni effects causes up to 50% underprediction of melt pool depth [113,116,117]. At high laser power densities, recoil pressure from evaporation exceeds surface tension forces, driving keyhole formation [92,113,116,117]. While controlled keyhole melting enhances penetration, unstable keyholes collapse and trap vapor bubbles, creating porosity [86,92,113]. Advanced models incorporating these coupled phenomena successfully predict melt pool morphology and defect formation across stainless steels, titanium alloys, and nickel-based superalloys [113,115,116,117].
Analytical models (Rosenthal point/line source, Goldak double-ellipsoid) enable rapid thermal field estimation [2,114,118,119,120]. Rosenthal provides efficiency (<1 s per track) but limited accuracy (±30–50% error) by neglecting fluid flow and temperature-dependent properties [118,119,120]. Goldak improves fidelity through volumetric energy distribution (±20–30% accuracy, ~10 min per track) [2,114]. Both fail to capture keyhole formation, Marangoni flow, and defect mechanisms, requiring high-fidelity CFD with recoil pressure and vapor dynamics [92,113,116].
High-fidelity multiphysics models integrate thermal, fluid, and phase change phenomena using finite volume, smoothed particle hydrodynamics (SPH), or lattice Boltzmann methods (LBMs) [84,90,91,92,113,116]. These approaches explicitly resolve the melt pool free surface and capture Marangoni convection, recoil pressure, and evaporation [91,92,113,116]. Despite detailed physics, many models simplify or omit powder-scale effects and dynamic powder bed interactions [91,121]. The treatment of vaporization and keyhole dynamics remains challenging due to complex multi-phase interactions and rapid transient phenomena [84]. The trade-off between computational efficiency and physical realism (Figure 4) guides model selection: Rosenthal/Goldak is needed for parameter screening and process mapping across large design spaces; CFD is needed for defect-critical applications requiring ±10–20% accuracy in melt pool predictions [2,8,9,114].
Extreme conditions and rapid time scales complicate a validation of melt pool models. High-speed imaging, pyrometry, and X-ray radiography provide valuable data but face limitations in spatial and temporal resolution [86,95,122]. High-speed cameras capture melt pool surface dynamics at frame rates exceeding 104 Hz, enabling observation of keyhole oscillations and spatter ejection [86,115]. Pyrometry and infrared thermography measure surface temperature distributions but are affected by emissivity variations and calibration challenges [95,123,124]. Synchrotron X-ray imaging offers unprecedented access to subsurface melt pool dynamics and keyhole formation, but is limited to specialized facilities [86]. Post-process characterization via optical microscopy and X-ray computed tomography provides detailed information on melt pool geometry and defect distribution but lacks temporal resolution [115,125]. The scarcity of comprehensive validation datasets encompassing multiple materials, process parameters, and measurement techniques limits confidence in model predictions [1,8,9].
Table 7 emphasizes that no single modeling approach satisfies all requirements for accuracy, computational efficiency, and physical completeness. Analytical models enable rapid parameter screening but lack predictive capability for defects. Continuum FEM/FVMs offer a practical balance for part-scale predictions but may miss critical fluid flow effects. High-fidelity CFD offers the most accurate melt pool predictions but remains computationally prohibitive for routine use. Particle-based methods excel at capturing powder-scale physics but face severe scalability limitations. Reduced-order models show promise for bridging fidelity and efficiency, although their reliability depends critically on the quality of training data. Future progress requires multi-scale frameworks that strategically deploy high-fidelity simulations where needed while leveraging efficient approximations elsewhere, guided by rigorous experimental validation [8,9,113,114].

3.4. Solidification and Microstructure Evolution

PBF processes are complex AM techniques that require precise modeling of microstructure evolution, grain growth dynamics, and phase transitions. Two prominent computational models used for these purposes are the phase-field (PF) and cellular automaton (CA) models.
Phase-field models are highly versatile and accurate for simulating microstructure evolution during PBF processes. They solve physical equations that govern phase transitions and grain growth, making them particularly effective in capturing detailed microscopic characteristics such as dendritic growth, stability of the solid–liquid interface, and grain nucleation [142,143,144]. For example, a three-dimensional phase-field model was used to simulate grain evolution during a three-layer, three-track PBF process, reproducing nucleation, growth, and coarsening processes in melting pools and heat-affected zones [144]. Similarly, phase-field models have been applied to study competitive growth and nucleation in alloys with spatially varying initial compositions, revealing the influence of composition on dendrite morphology [142].
Phase-field models are widely used in both academic research and industrial applications. They provide insight into the relationship between process parameters (e.g., laser power and scan speed) and microstructure evolution, allowing optimization of PBF processes [142,145]. For example, a nonisothermal phase-field model was used to generate a densification map for 316L stainless steel, classifying the process parameters based on the resulting morphologies [145]. A transient 3D multi-scale phase-field model was developed to simulate microstructure evolution in Ti-6Al-4V components produced by LPBF. The model accounted for rapid solidification conditions and the impact of process parameters (e.g., laser power, speed, beam shape) on grain morphology and texture. The results were validated against experimental data, demonstrating the ability of the model to predict grain size and texture evolution [49]. In addition, phase-field models have been coupled with machine learning techniques to accelerate simulations, allowing large-scale predictions of the microstructure evolution in PBF processes [146,147].
Despite their accuracy, phase-field models suffer from high computational costs, limiting their application to large-scale simulations relevant to AM processes. To address this, researchers have developed surrogate machine learning models trained in small-scale phase-field simulations, reducing computational time by orders of magnitude while maintaining high accuracy [146,147].
Cellular automaton (CA) models are computationally efficient tools for simulating microstructure evolution and grain growth in PBF processes. They use a grid-based approach with predefined rules to model nucleation and growth, making them suitable for large-scale simulations [148,149,150]. For example, a coupled CA-FE model was developed to predict grain texture evolution during the solidification of Alloy 625, capturing competitive grain growth and columnar/equiaxed grain structures [149]. Similarly, a 3D CA model was used to simulate the evolution of grain structure in multi-layer, multi-track LPBF components, considering factors such as thermal gradients and epitaxial growth [151,152].
Cellular automaton models are particularly useful for studying the effects of process parameters on grain morphology and texture. For example, a CA model was used to investigate the dependence of microstructure on the printing areas of Inconel 625, revealing a weaker texture intensity and more heterogeneous grain structures in broader sections [148]. Additionally, CA models have been coupled with finite element models to simulate the thermal history of PBF processes, allowing predictions of grain nucleation and growth based on temperature fields [149,153].
Although CA models are computationally efficient, they may lack the resolution to capture detailed microscopic features such as transient growth conditions and stability of the solid–liquid interface, which are critical for accurate microstructure prediction [144]. However, refinements to the CA grid and the introduction of additional physical considerations, such as thermal gradients and solid–solid interface growth, have improved the flexibility and accuracy of these models [148].
Figure 5 illustrates the complementary roles of phase-field and cellular automaton methods in predicting microstructure evolution, validated against experimental EBSD measurements. Phase-field simulations (Panel A) resolve dendritic solidification at the interface scale, capturing primary dendrite arm spacing (λ1 = 3–5 μm) and secondary arm spacing (λ2 ≈ 1 μm) characteristic of LPBF’s high cooling rates (104–106 K/s). These fine-scale features contrast sharply with conventional casting (λ1 = 50–100 μm at ~102 K/s), reflecting the relationship λ1 ∝ (cooling rate)−0.5. Cellular automaton models (Panel B) operate at the mesoscale to predict the evolution of the grain structure, epitaxial growth from the substrate, and the location of the columnar-to-equiaxed transition (CET). The CET typically occurs at 200–400 μm above the substrate, where the thermal gradient-to-solidification velocity ratio (G/R) decreases sufficiently to favor equiaxed nucleation. EBSD validation (Panel C) confirms strong ⟨001⟩ fiber texture in the columnar region (red IPF-Z coloring) due to preferential grain growth along the thermal gradient direction, with texture strength diminishing in the equiaxed region. Quantitative comparisons show that phase-field predictions achieve ±15–25% accuracy for dendrite arm spacing, while CA models predict grain size distributions within ±20–30%, of which both are validated with experimental measurements [139,142,143,148,149,151,152,153,154,155].

3.5. Thermomechanical Response and Residual Stress

Thermomechanical modeling constitutes a critical component of powder bed fusion (PBF) process simulation, enabling prediction of residual stresses, thermal gradients, and part distortion that directly influence dimensional accuracy and mechanical integrity [2,156]. Finite element method (FEM) simulations incorporating sequentially coupled thermal–mechanical analyses capture the complex interplay between rapid heating and cooling cycles, temperature-dependent material properties, and evolving stress states across multiple length scales [94,157]. These models typically employ element activation techniques to simulate layer-by-layer material deposition, nonlinear thermoelastoplastic constitutive laws to represent material behavior under extreme thermal gradients (exceeding 106 K/s), and adaptive meshing strategies to balance computational efficiency with spatial resolution requirements [158,159].
Inherent strain methods offer computationally efficient alternatives by extracting effective plastic strains from detailed process simulations and applying them at the part scale, achieving computational speed-ups of 6–100× while maintaining prediction accuracy within 10–15% for residual stress and distortion [160,161]. Modified inherent strain approaches incorporate temperature-dependent mechanical properties, anisotropic effects induced by scanning strategies, and shear strain components to enhance fidelity [162,163,164]. Validation against experimental measurements, including synchrotron diffraction, neutron diffraction, X-ray diffraction, and optical three-dimensional scanning, consistently demonstrates that both FEM and inherent strain frameworks capture residual stress distributions and distortion patterns with errors typically below 15%, although computational cost–accuracy trade-offs remain significant [165,166,167].

3.5.1. Influence of Preheating on Residual Stress

Substrate and powder bed preheating represents a widely adopted strategy for mitigating residual stress and distortion in PBF by reducing thermal gradients during processing [166]. Thermomechanical simulations demonstrate that preheating Ti-6Al-4V substrates from ambient temperature (20 °C) to elevated temperatures (200–300 °C) reduces peak thermal gradients by 30–50%, corresponding to proportional reductions in maximum residual stress magnitudes of 40–60% [156,168]. The mechanism underlying stress reduction involves decreased temperature differentials between molten material and surrounding solid regions, resulting in lower thermal strain accumulation during solidification and reduced driving force for plastic deformation [166].
The temperature-dependent yield strength governs stress evolution during thermal cycling and can be approximated by the following:
σ y T = σ y 0 k T
where σ y 0 is the yield strength at the reference temperature (Pa), k is the thermal softening coefficient (typically 0.5–1.5 MPa·K−1 for Ti–6Al–4V), and T is the temperature rise above the reference state (K) [156]. At elevated preheating temperatures, reduced yield strength facilitates stress relaxation through plastic flow, preventing accumulation of elastic residual stresses that would otherwise persist upon cooling to ambient conditions [169].
Preheating also redistributes tensile and compressive stress zones: without preheating, high tensile stresses concentrate near scan track edges and top surfaces (often exceeding 500–700 MPa for Ti-6Al-4V), while compressive stresses develop in substrate regions adjacent to the build [157,170]. With 200 °C preheating, peak tensile stresses reduce to 300–400 MPa, and the compressive zone extends deeper into the substrate, improving stress balance and reducing distortion tendency [156,168]. However, excessive preheating (>400 °C for Ti alloys) can promote grain coarsening, alter phase transformation kinetics, and reduce mechanical property uniformity, necessitating optimization of preheating temperature for specific material–geometry combinations [166].

3.5.2. Role of Absorptivity Variation

Laser absorptivity variation (A = 0.3–0.7 for metallic powders, detailed in Section 3.2) critically influences energy coupling efficiency, melt pool dimensions, and thermal gradients driving residual stress formation [156,171,172]. Temperature-dependent absorptivity (increasing with electron–phonon coupling) is calculated as follows [156,173]:
A T = A 0 + β T ,
where A0 is the baseline absorptivity; β is the temperature coefficient of absorptivity (typically 1–3 × 10−4 K−1), quantifying the thermal sensitivity of absorptivity; and T is the temperature (or temperature rise) in kelvin (K).
Absorptivity variations of ±20% translate to melt pool depth fluctuations of ±20–30% and width variations of ±10–15%, directly affecting fusion quality and residual stress [171,172]. Higher absorptivity enlarges melt pools, intensifies thermal gradients, and increases residual stress by 15–25%; lower absorptivity reduces penetration, risking lack-of-fusion defects [156,171].
The sensitivity of residual stress predictions to absorptivity underscores the importance of accurate material characterization and considering surface condition evolution during multi-layer builds [172]. Powder layers exhibit higher absorptivity (0.4–0.6) than solidified surfaces (0.3–0.45) due to multiple scattering and cavity effects, creating layer-dependent energy coupling that influences thermal history and stress accumulation [156,173]. Advanced thermomechanical models incorporate temperature- and state-dependent absorptivity to capture these effects, improving prediction accuracy for residual stress and distortion [171].

3.5.3. Combined Process Parameter Interactions

Laser power (P), scan speed (v), and hatch spacing (h) constitute the primary process parameters governing energy density (E = P/(v·h·t), where t is layer thickness) and resulting in the thermal–mechanical response in PBF [156,174]. These parameters exhibit strong nonlinear interactions: increasing laser power from 150 W to 250 W while maintaining constant scan speed (800 mm/s) and hatch spacing (0.1 mm) elevates peak temperatures by 200–300 K, expands melt pool volume by 40–60%, and increases maximum residual stress by 25–35% [156,168]. However, simultaneous increases in scan speed proportionally reduce energy density, partially offsetting power effects and creating complex parameter coupling [174].
Response surface methodology (RSM) and factorial design studies reveal that scan speed exerts the strongest influence on residual stress magnitude (accounting for 40–50% of variance), followed by laser power (30–35%) and hatch spacing (15–20%) [172,174]. Interaction effects between parameters account for an additional 10–15% of variance, emphasizing the inadequacy of one-factor-at-a-time optimization approaches [174]. For example, the stress-reducing benefit of increased scan speed diminishes at low laser powers where insufficient fusion occurs, while high-power, high-speed combinations can induce keyhole porosity that locally concentrates stresses [156].
Hatch spacing influences residual stress through its effect on thermal overlap between adjacent scan tracks: smaller spacing (0.08–0.10 mm) increases re-melting and reheating of previously solidified material, promoting stress relaxation but also elevating cumulative heat input and part-scale thermal gradients [168,172]. Larger spacing (0.12–0.15 mm) reduces inter-track thermal interaction, potentially decreasing global residual stress but risking lack-of-fusion defects between tracks that act as stress concentrators [172,174]. Optimal parameter combinations identified through thermomechanical simulations typically balance energy density near the material-specific threshold for full melting (60–80 J/mm3 for Ti-6Al-4V) while minimizing thermal gradients through moderate power (180–220 W) and speed (900–1100 mm/s) selections [156,168].

3.5.4. Laser Operation Mode Effects

Continuous wave (CW) versus pulsed laser operation modes fundamentally alter temporal energy delivery, resulting in a thermal–mechanical response in PBF [169,171]. CW lasers maintain constant power output throughout scanning, generating steady-state melt pools with relatively uniform thermal gradients along the scan direction. Pulsed mode operation introduces periodic energy modulation (pulse duration 0.1–10 ms, duty cycle 50–90%), creating transient melt pool dynamics with alternating heating and partial cooling phases [171].
Thermomechanical simulations demonstrate that pulsed laser operation with optimized parameters (pulse duration 2–5 ms, duty cycle 70–80%, peak power 1.5–2× average CW power) reduces the maximum residual stress by 20–40% compared to equivalent average-power CW processing [169]. The stress reduction mechanism involves intermittent heating that allows partial stress relaxation during pulse-off intervals, reducing peak thermal gradients (by 15–25%) and limiting plastic strain accumulation [171]. Additionally, pulsed operation promotes finer microstructures through repeated thermal cycling, potentially enhancing mechanical properties [169].
However, pulsed mode introduces fusion quality risks: insufficient pulse overlap causes periodic lack-of-fusion defects between pulses, while excessive peak powers during pulse-on phases can induce keyholing and porosity [171]. Optimal pulsed operation requires careful matching of pulse frequency to scan speed to ensure adequate melt pool overlap (typically 60–80% overlap between consecutive pulses) [169]. Modulated power strategies employing gradual power ramping at scan track start/end or layer transitions offer intermediate approaches that reduce residual stress (by 10–20%) while maintaining fusion integrity superior to aggressive pulsing [158,169].
The computational modeling of pulsed laser operation demands finer temporal resolution (timesteps 0.1–1 ms versus 5–10 ms for CW) to capture transient thermal dynamics, increasing simulation cost by factors of 5–10 [171]. Hybrid models employing detailed pulsed simulations for calibration followed by effective continuous heat source representations at the part scale offer practical compromises for design optimization [169].
The influence of key process parameters on residual stress and distortion is summarized in Table 8.
Figure 6 presents an integrated view of validation metrics across the entire PBF modeling chain, illustrating how accuracy and efficiency evolve from powder to part scale. Starting with DEM simulations, the models achieve 5–10% accuracy in predicting powder packing density for well-characterized morphologies, forming a reliable basis for subsequent melt pool modeling. Melt pool simulations reveal the steep computational trade-offs among analytical, finite element, and high-fidelity CFD approaches—spanning sub-second to hour-long runtimes and errors from ±30–45% down to ±10–15%. At the part scale, thermomechanical predictions align with synchrotron and neutron diffraction data within ±10–15% for key alloys such as Ti-6Al-4V, 316L, and IN718. Finally, microstructure modeling shows phase-field methods attaining 70–82% accuracy for grain and dendrite features, while cellular automaton models achieve 70–78% with greater efficiency. Together, these results underscore the central challenge of balancing physical realism with computational tractability in advancing PBF modeling toward industrial application.

4. Evolution of Modeling Approaches

The modeling of powder bed fusion (PBF) has undergone a remarkable evolution, reflecting the increasing complexity and industrial relevance of the process. Early work relied on empirical correlations and analytical heat transfer models, which provided computationally efficient tools for mapping process parameters and defect regimes, but suffered from oversimplifications that limited predictive accuracy. With advances in computational power, physics-based numerical simulations emerged, enabling multiphysics and multi-scale insights into thermal fields, melt pool dynamics, powder spreading, and residual stress formation. More recently, hybrid modeling strategies have sought to combine the rigor of physics-based approaches with the adaptability of data-driven and machine learning methods, improving both predictive fidelity and computational efficiency. Complementing these developments, the integration of in situ monitoring data has opened new possibilities for real-time calibration, adaptive control, and the eventual realization of digital twin frameworks. Together, these advances chart a trajectory from simple analytical tools toward intelligent, data-integrated models capable of guiding defect mitigation, microstructure control, and performance optimization in PBF.
Figure 7 outlines the progression of PBF modeling methods, from early empirical approaches to advanced digital twin frameworks.

4.1. Physics-Based Modeling

The integration of empirical data with analytical solutions enhanced the robustness of process maps and improved defect prediction frameworks [119,178,179]. Dimensionless parameters and scaling laws were introduced to generalize processing windows, yet the static nature of many early models hindered their application in dynamic process control and real-time optimization [47,180]. Critical assumptions—including uniform powder bed properties, simplified heat source representations, and steady-state conditions—reduced physical realism and limited accurate prediction of process instabilities under practical manufacturing conditions [47,181,182].
Physics-based numerical simulations address the limitations of analytical models through detailed multiphysics representations. Mesh-based methods (FEM, FVM) provide validated accuracy for temperature distributions, residual stresses, and distortions [1,2,93], though challenges persist in handling large deformations and fine-scale melt pool dynamics [2,183].
Particle-based methods—including discrete element method (DEM), smoothed particle hydrodynamics (SPH), and lattice Boltzmann method (LBM)—offer flexible frameworks for modeling discrete powder behavior, melt pool fluid dynamics, and complex free surface phenomena [1,59,60,84,91,184]. GPU acceleration and adaptive refinement techniques have enabled mesoscale simulations with improved resolution, though computational intensity remains a barrier to routine application [61,90,185]. Coupled approaches, such as DEM-CFD, DEM-LBM, and DEM-SPH, have demonstrated enhanced predictive capabilities for powder–fluid interactions, vapor–powder–melt dynamics, and microstructure evolution [60,101,186,187]. Cellular automata (CA) coupled with LBM have effectively simulated powder dynamics and thermal–fluid interactions [188].
Despite these advances, multiphysics integration increases model complexity and computational cost, often necessitating simplifications that may compromise accuracy [1,106]. The validation of coupled models is challenging due to limited comprehensive experimental data spanning all relevant phenomena [101,187]. Parameter sensitivity, numerical stability, and scalability issues further constrain the robustness and industrial applicability of advanced numerical frameworks [60,186]. Mesh-based methods struggle with accurate representation of discrete powder behavior and dynamic free surfaces, while mesh-free methods face challenges in numerical stability and parameter calibration [1,59,102,106].
The transition from empirical to physics-based modeling reflects a broader shift toward higher-fidelity, multi-scale approaches capable of capturing the intricate interplay of thermal, fluid, and mechanical processes in PBF.
Table 9 synthesizes the principal capabilities, computational efficiency, and limitations across this modeling spectrum, highlighting the critical trade-offs that continue to shape model selection and development priorities.
The synthesis reveals that while empirical and analytical models remain valuable for rapid process design and parameter screening, the increasing demand for predictive accuracy and physical realism has driven adoption of computationally intensive physics-based simulations. The ongoing challenge lies in balancing model fidelity with computational feasibility, particularly for industrial-scale applications requiring real-time or near-real-time process optimization and control.

4.2. Hybrid Physics–Data-Driven Approaches

The convergence of physics-based simulations and data-driven techniques addresses a fundamental tension in PBF modeling: physics-based models require prohibitive computational resources for real-time control, while purely data-driven approaches lack physical interpretability and generalizability beyond training domains [204,205]. Hybrid frameworks leverage complementary strengths—physics components encode governing equations ensuring thermodynamic consistency, while machine learning accelerates computations and identifies latent patterns in complex process signatures [204,205,206]. This synergy is critical for PBF, where rapid thermal cycles, stochastic powder bed heterogeneity, and transient melt pool dynamics generate high-dimensional process spaces, challenging both purely analytical and empirical methods [205].
Recent hybrid efforts demonstrate substantial progress in melt pool prediction and thermal history estimation. Integration of CFD with machine learning improves melt pool width prediction accuracy while reducing computational time [204]. Physics-informed neural networks (PINNs) embed governing equations into neural network loss functions, enabling temperature field prediction with limited training data [205]. These frameworks consistently outperform standalone approaches in predicting melt pool dynamics, thermal profiles, and defect formation [111,204,205], with experimental validations supporting reliability [205,206]. Concurrent modeling of microstructural evolution and porosity enhances practical relevance [111], while techniques such as equivalent boundary condition methods optimize computational efficiency [158].
Hybrid model applicability spans stainless steels, titanium alloys, aluminum alloys, and multi-material systems [206,207,208]. Integration with digital twin frameworks supports defect prediction, parameter optimization, and adaptive control in aerospace, biomedical, and defense sectors [7,209,210,211]. Hybrid manufacturing strategies that combine LPBF with milling or direct energy deposition expand application scope [212,213,214], demonstrating improvements in surface finish, dimensional accuracy, and mechanical performance [214,215].
Despite progress, significant challenges constrain broader adoption. Integration methods lack standardization, and the fusion of heterogeneous data sources introduces biases [204,205]. Heavy reliance on extensive datasets or high-fidelity simulations limits generalizability across alloys and process regimes [204,205,207,208,216]. Validation remains constrained by material-specific campaigns, restricting transferability to novel systems [208,216]. Computational efficiency gains have not yet enabled real-time control for full-component simulations with detailed multiphysics [111,205]. Scalability challenges persist in extending models to industrial applications, requiring real-time data acquisition and integration with manufacturing execution systems [7,217]. Translation to industrial practice is hindered by challenges in model updating and integration with legacy systems [7,217], while optimization of hybrid process parameters remains complex and often empirical [207,208,214].
Table 10 synthesizes principal advances and persistent challenges.
Hybrid physics–data-driven approaches represent a promising yet still-maturing paradigm. While substantial progress has been achieved in prediction accuracy and computational efficiency, realizing full potential for industrial-scale, real-time process control requires continued advances in data integration, model standardization, multi-scale coupling, and experimental validation across diverse material systems and manufacturing conditions.

4.3. Integration of In Situ Monitoring Data

The literature shows notable progress in leveraging in situ thermal, optical, and multi-sensor data for calibrating PBF models, enhancing process insight, and defect prediction. Many studies demonstrate the potential of machine learning techniques to enhance defect detection and process control, while others focus on the challenges of data registration and sensor fusion. However, limitations persist in the robustness of methodologies, data quality, and the generalizability of models across different geometries and process conditions. The synthesis underscores the need for improved calibration protocols, comprehensive multi-modal data integration, and real-time adaptive control strategies to fully leverage in situ monitoring for PBF model calibration. Table 11 provides a systematic, evaluative assessment of in situ monitoring approaches for powder bed fusion, addressing sensor capabilities, validation rigor, computational performance, transferability, and industrial readiness. Each entry identifies which methods are most credible for specific purposes.

5. Digital Twins and AI Integration for PBF

Digital twins represent the convergence of physics-based modeling, real-time sensing, and artificial intelligence into integrated frameworks for PBF process control and optimization. This section examines the architectural foundations of PBF digital twins, the multi-modal data streams that enable them, the AI methods that power real-time inference, and industrial implementations that demonstrate their capabilities.

5.1. Digital Twin Concept and Architecture

A digital twin (DT) for PBF comprises three interconnected architectural layers that enable real-time process monitoring and control [251].
The data layer integrates multi-modal sensor fusion, combining optical, thermal (infrared), acoustic emission, and pyrometry data through feature-level or decision-level fusion strategies. High-speed multi-modal systems achieve 95–98.5% accuracy in flaw detection with temporal resolution up to 20 kHz, proposed for intra-layer closed-loop control [252,253].
At the foundational level, the digital twin begins with high-fidelity data acquisition from the physical PBF system. The PBF process involves intricate interactions among laser energy input, powder feedstock behavior, thermal gradients, and environmental conditions. Capturing these phenomena requires a robust sensor suite, including photodiodes, infrared pyrometers, high-speed cameras, and thermocouples [254]. These sensors monitor key parameters, including melt pool geometry, temperature distributions, powder flow consistency, and layer integrity. The sensor data is typically generated at high frequencies and in large volumes, which necessitates efficient edge computing strategies for real-time pre-processing. Utilizing edge devices mitigates latency and data bottlenecks while ensuring time-critical information is delivered to higher digital layers with minimal delay [255].
Once collected, sensor data is channeled through standardized communication protocols, such as OPC-UA or MQTT, into centralized data repositories. These repositories organize information as time-series records and metadata registries, capturing everything from machine configurations to build plans and environmental variables. This layer supports data provenance and traceability—an essential requirement for certification in regulated industries. Furthermore, the integration of machine and process metadata into a unified data framework enables correlation of in-process signals with final part quality metrics.
The physics layer employs reduced-order models, surrogate models, or hybrid computational fluid dynamics (CFD)/finite element method (FEM) simulations to predict thermal history, melt pool dynamics, and microstructure evolution. Bayesian calibration frameworks enable probabilistic DTs that quantify uncertainty and support layer-wise parameter adjustment based on real-time sensor feedback [256].
The modeling and simulation layer builds on this data foundation by providing a virtual mirror of the PBF process. High-fidelity finite element and computational fluid dynamics (CFD) models simulate heat transfer, melt pool behavior, phase transitions, and residual stress evolution during fabrication. These physics-based models are invaluable for predicting defect formation, including porosity, keyholing, and delamination. However, due to their computational intensity, they are often supplemented by surrogate models—typically built using machine learning techniques—that approximate system behavior with significantly lower latency. Notably, deep neural operators and Fourier neural operator frameworks have been proposed to emulate temperature fields and melt pool dynamics in real time, allowing predictive control and uncertainty quantification [224].
The AI layer provides rapid inference through trained ML models, with potential for predictive maintenance, defect prognosis, and adaptive control at latencies approaching process time scales under controlled experimental conditions [8].
The analytics and inference layer acts as the digital twin’s decision engine. This component employs machine learning models, statistical algorithms, and signal processing tools to extract actionable insights from process data. It identifies anomalies, estimates quality outcomes, and recommends corrective measures. For instance, supervised learning algorithms can classify melt pool instabilities, while unsupervised models detect deviations from baseline operation. The control module then used these predictions to enact dynamic adjustments in process parameters. Closed-loop feedback control, based on such analytics, has been demonstrated to stabilize the process by actively adjusting laser power or scan speed in response to emerging anomalies in controlled experimental studies on specific materials and machine configurations [257,258].
Real-time synchronization between physical and virtual models is achieved through bidirectional data exchange, where sensor measurements continuously update the virtual representation while simulation predictions guide the adjustments to the physical process [7].
Human interaction with the digital twin is facilitated through advanced visualization and user interface systems. Real-time dashboards and 3D process visualization tools offer intuitive representations of part progress, thermal maps, and defect probabilities. Augmented reality (AR) interfaces further enhance operator situational awareness by overlaying digital process data onto physical environments, streamlining maintenance and inspection workflows [259,260]. These visualizations are not merely cosmetic; they are critical for diagnostics, decision-making, and collaboration across multidisciplinary teams.
For comprehensive functionality, the digital twin must also be seamlessly integrated into enterprise IT systems. Integration with product lifecycle management (PLM), manufacturing execution systems (MESs), and enterprise resource planning (ERP) platforms enables the twin to access design intent, monitor production workflows, and link to supply chain logistics. This enterprise connectivity transforms the digital twin into a strategic asset that spans design, production, and business decision layers [261,262].
Figure 8 illustrates the complete ecosystem of a PBF digital twin.

5.2. Data Sources and Sensor Integration

Sensor technologies and data integration methods form the foundation of PBF digital twin development. Recent advances in thermography, pyrometry, acoustic monitoring, and melt pool imaging enable multi-sensor fusion for defect detection and process control. Challenges persist in spatial–temporal data registration, signal interpretation, and the implementation of FAIR data principles, with gaps remaining in standardization and industrial scalability.
Table 12 summarizes sensor technologies and data fusion approaches for PBF digital twins, evaluating specifications, validation, computational performance, transferability, and industrial readiness.
Figure 9 summarizes defect regimes, detection performance, and sensor fusion gains across the major PBF defect types. Distinct process windows in Ti-6Al-4V delineate lack of fusion at low energy densities, keyhole porosity at high energies, and an optimal zone minimizing both. Detection benchmarking shows X-ray CT as the validation reference, with optical and acoustic sensors excelling at surface and transient defects, respectively, while thermal imaging captures energy coupling variations. Multi-modal fusion combining these signals improves overall accuracy from 83% to 95% and boosts defect-specific detection—especially for gas porosity and lack of fusion—by exploiting complementary sensing physics. These results confirm that integrated sensor fusion surpasses the 90% accuracy threshold required for reliable, real-time defect classification in digital twin-enabled PBF systems.

5.3. AI and Machine Learning Methods

Artificial intelligence (AI) and machine learning (ML) have emerged as transformative paradigms for modeling and controlling powder bed fusion (PBF) processes, addressing the inherent complexity of multiphysics phenomena that govern part quality and process stability [280]. PBF involves rapid thermal cycling, complex melt pool dynamics, and stochastic defect formation mechanisms that are difficult to capture through physics-based models alone [112]. Data-driven ML methods complement traditional computational approaches by learning complex nonlinear mappings from high-dimensional sensor data to process outcomes, enabling predictions of melt pool geometry, thermal history, and defect formation [5]. The integration of AI into PBF process chains has been proposed to facilitate real-time decision-making, adaptive parameter control, and quality assurance, bridging the gap between laboratory-scale understanding and industrial deployment [287].

5.3.1. Supervised Learning for Defect Detection

Supervised learning forms the foundation of most ML applications in PBF, where labeled datasets relating process parameters to outcomes enable predictive modeling. Convolutional neural networks (CNNs) have demonstrated very high accuracy in defect detection from optical and thermal imagery, classifying porosity, lack of fusion, and surface anomalies in real-time layer-wise monitoring [281]. Ensemble methods such as random forest and XGBoost enable online process monitoring [281]. Transfer learning extends model applicability across materials, with cross-material quality monitoring achieving >92% accuracy after fine-tuning with limited target domain data [288].

5.3.2. Self-Supervised and Semi-Supervised Learning

Self-supervised learning addresses the challenge of unlabeled data in PBF through clustering and anomaly detection. Self-supervised learning frameworks reduce labeling requirements by leveraging temporal correlations in high-speed video data, achieving multi-label melt pool anomaly classification without extensive manual annotation [289]. Semi-supervised approaches combine limited labeled data with abundant unlabeled observations to improve model generalization, which is particularly valuable for rare defect types where labeled examples are scarce [290].

5.3.3. Reinforcement Learning and Model Predictive Control

Reinforcement learning (RL) enables autonomous process control by learning optimal policies through interaction with the PBF environment. RL-based scan path optimization has demonstrated improved thermal uniformity and reduced residual stress through dynamic laser parameter adjustment, with geometry-agnostic controllers validated on Ti-6Al-4V and AISI 316L alloys [291]. Model predictive control (MPC) frameworks integrate RL with physics-based surrogate models, achieving precise melt pool temperature tracking that outperforms traditional PID control while maintaining real-time computational feasibility [292].

5.3.4. Physics-Informed Neural Networks

Physics-informed learning represents a paradigm shift toward hybrid modeling, where governing physical laws are embedded directly into neural network architectures. Physics-informed neural networks (PINNs) encode partial differential equations (heat equation, Navier–Stokes) as physics-informed loss functions, achieving temperature field predictions with less than 7% deviation from experimental measurements while requiring significantly reduced computational time compared to finite element simulations [220]. PINNs demonstrate strong generalization across unseen laser scanning strategies and geometries [222].

5.3.5. Integration with Digital Twin Frameworks

The integration of AI with DT frameworks enables predictive and adaptive control capabilities that transcend traditional process monitoring. ML models serve as fast surrogate representations within the DT architecture, replacing computationally expensive physics simulations for real-time inference while maintaining physical consistency through physics-informed training [204]. Bidirectional feedback loops facilitate closed-loop control, where DT predictions inform laser power adjustments, scan path modifications, or layer-wise parameter tuning to prevent defect formation, while in situ sensor data continuously refine model predictions through online calibration [293]. Hybrid physics–ML approaches combine the interpretability and extrapolation capability of physics-based models with the flexibility and speed of data-driven methods [230].
Domain adaptation techniques enhance DT reusability across different machines and sensor configurations, improving anomaly detection accuracy when transferring models between platforms, reducing the retraining burden for industrial deployment [244]. Uncertainty quantification through Bayesian frameworks provides confidence bounds on DT predictions, enabling risk-aware decision-making and robust process optimization under variable conditions [223].

5.3.6. Validation and Scalability Considerations

Validation strategies for AI-enabled DTs include ex situ characterization (X-ray computed tomography, metallography) and in situ monitoring correlation, with model accuracy benchmarked against experimental thermal profiles, melt pool dimensions, and microstructural features [202]. Scalability challenges arise from the computational demands of high-fidelity physics models, addressed through surrogate modeling, reduced-order representations, and GPU acceleration that enable multi-layer simulations with practical latencies [294].
Table 13 provides a systematic comparison of AI and machine learning methods for PBF digital twin integration, evaluating application domains, physics integration strategies, sensor modalities, validation approaches, computational performance, transferability, and key limitations.
Figure 10 presents a performance overview of AI-enhanced digital twins, integrating accuracy, efficiency, and sensor fusion. Physics-based models achieve moderate accuracy, while physics-informed neural networks and hybrid surrogates markedly reduce errors and boost prediction fidelity. Computational trade-offs show hybrid and surrogate models, delivering up to 3900× faster performance than FEM/CFD with ≥97% accuracy, enabling real-time operation. Multi-sensor fusion further enhances defect detection to 98.5% accuracy and 95% POD by combining thermal, acoustic, and optical data. Together, these results highlight hybrid physics–ML digital twins with sensor fusion as the most effective framework for accurate, real-time control in industrial PBF systems.

5.4. Case Studies of PBF Digital Twin Implementation

Recent advances in digital twins (DTs) for industrial PBF highlight the effective integration of physics-based models and machine learning, enabling real-time process monitoring and control. Strengths include the development of hybrid digital twin architectures that combine data-driven and physics-based approaches, enhancing predictive accuracy and process optimization. However, challenges persist in scalability, data heterogeneity, and the integration of multi-source sensor data across diverse industrial contexts. Additionally, while many studies demonstrate promising capabilities in defect detection and process control, the generalizability and industrial deployment of these digital twins remain limited by computational demands and standardization issues. The synthesis highlights the need for modular validation frameworks and adaptable architectures to address evolving manufacturing complexities and ensure sustained digital twin fidelity.
An overview of digital twin developments for PBF is presented in Table 14.

6. Optimization Strategies in PBF Manufacturing

Reliable industrial deployment of powder bed fusion (PBF) increasingly depends on the integration of optimization frameworks with predictive quality assurance. Recent advances have demonstrated that embedding process physics into topology optimization enables designs that account for residual stresses, thermal distortion, and support minimization, although balancing computational fidelity with efficiency remains a challenge. At the same time, machine learning is transforming parameter selection, offering more accurate control of laser power, scan speed, and layer settings to enhance density, surface quality, and mechanical performance. Complementing these developments, predictive models that combine physics-based simulations with multi-sensor monitoring are emerging as powerful tools for defect detection and adaptive control. Collectively, these approaches mark a transition from empirical, trial-and-error strategies to integrated, data-driven frameworks capable of achieving first-time-right manufacturing in PBF.
Figure 11 illustrates the iterative and continuous nature of optimization in PBF through a circular layout. At the center lies the optimization core, symbolizing the driving mechanism of constant improvement. Surrounding it are six interconnected components arranged in a logical sequence, each contributing to the process, a better understanding, and control. Feedback arrows radiating from the central core emphasize the integrated, bidirectional flow of information across all stages. Key annotations highlight the benefits of this framework, including adaptability, scalability, and enhanced process robustness, underscoring its role as a closed-loop pathway to reliable high-quality PBF manufacturing.

6.1. Topology Optimization with PBF Process Constraints

The studies on topology optimization in powder bed fusion additive manufacturing reveal significant advancements in integrating manufacturing constraints and process physics into design frameworks. Notably, the incorporation of thermal distortion, overhang control, and support structure optimization has been addressed with increasing sophistication. However, challenges remain in balancing computational efficiency with model fidelity, especially when simulating complex process-induced phenomena such as residual stresses and material heterogeneity. Furthermore, while multi-material and graded material considerations are emerging, their practical integration into topology optimization frameworks is still limited. The literature also highlights the growing role of artificial intelligence and machine learning in automating design for additive manufacturing, though these approaches require further validation. Overall, the field is progressing towards more holistic and manufacturability-aware optimization methods, yet gaps persist in experimental validation and scalability to industrial applications.
Table 15 highlights strategies and challenges in PBF-oriented topology optimization.

6.2. Machine Learning-Driven Process Parameter Optimization for PBF

Machine learning algorithms described in Section 5.3, including Bayesian optimization, deep neural networks, and reinforcement learning, are increasingly applied to optimize process parameters in powder bed fusion (PBF) additive manufacturing. Studies show significant improvements in build quality metrics such as surface roughness, porosity, and mechanical properties through intelligent parameter selection and real-time process monitoring.
Table 16 summarizes the application of ML and Bayesian optimization methods for PBF parameter optimization. A comparison of ML approaches for key PBF quality indicators is provided in Table 17.

6.3. Quality Assurance in PBF via Predictive Models

Predictive modeling for quality assurance applies the machine learning and sensor fusion methods detailed in Section 5.2 and Section 5.3 to defect detection and process control in PBF. While multi-modal sensor data integration and deep learning architectures enhance prediction accuracy and real-time monitoring capabilities, challenges persist in data acquisition, labeling, model generalizability across materials and process parameters, and computational efficiency for real-time applications. Robust, adaptable models that balance interpretability and performance while addressing practical industrial constraints remain a critical need.

7. Software Platforms and Tools for PBF Modeling

The computational ecosystem for powder bed fusion (PBF) modeling encompasses commercial finite element packages, open-source research codes, machine learning frameworks, and digital twin solutions. Tool selection depends on modeling objectives (process optimization, defect prediction, distortion analysis), required physics fidelity, computational resources, and licensing constraints. Commercial software offers mature implementations and technical support at a substantial cost, while open-source alternatives provide flexibility and transparency with steeper learning curves.
The multi-scale, multiphysics nature of PBF means no single platform comprehensively addresses powder spreading (DEM), laser–powder interaction (ray-tracing), melt pool dynamics (CFD), solidification microstructure (phase-field/cellular automata), and thermomechanical response (FEM) within a unified framework. Users typically employ multiple specialized tools coupled through custom scripts or APIs, introducing workflow complexity. This section reviews the principal software platforms that enable PBF modeling and digital twin development.

7.1. Commercial (e.g., ANSYS, Simufact) and Open-Source Tools

Simulation tools play a critical role in optimizing PBF processes by predicting distortions, residual stresses, and material properties. This section compares commercial and open-source tools, focusing on their technical capabilities, simulation accuracy, and computational efficiency, with specific examples of their applications in PBF.
ANSYS Additive is a commercial tool widely used for simulating PBF processes. It offers advanced capabilities such as predicting distortion, residual stress, and thermal deformation, as well as automatic support generation and scanning pattern optimization. The tool has been successfully applied to simulate the tension specimen of 17-4PH steel, demonstrating its ability to optimize process parameters and improve part quality [384,385]. ANSYS Additive uses finite element analysis to simulate the thermal history and mechanical behavior of parts during PBF. The tool can predict microstructural properties such as porosity and material anisotropy, which are critical for understanding the mechanical properties of the final part [385]. ANSYS Additive has been coupled with machine learning models to predict melt pool geometry and microstructure, significantly reducing simulation times compared to conventional numerical methods [385]. In a study on 17-4PH steel, ANSYS Additive was used to optimize process parameters, resulting in improved part quality and reduced production costs [384].
Simufact Additive is another commercial tool designed for simulating PBF processes. It is particularly useful for optimizing part orientation and predicting the influence of process parameters on part quality. The tool has been used to analyze the effect of weight factors on part orientation, demonstrating its ability to minimize support material volume and building risk [386]. Simufact Additive allows users to predict the optimal orientation of parts based on selected criteria, such as support material volume and building risk. The tool can simulate the residual stress and distortion caused by thermal gradients during the PBF process. Simufact Additive enables the assignment of weight factors to parameters, allowing users to prioritize specific criteria during optimization [386]. In a study on part orientation, Simufact Additive was used to compare three possible orientations of a part, demonstrating that orientation No. 1 achieved better mechanical properties compared to orientation No. 3 [386].
Other commercial tools, such as those from Autodesk and MSC, have also been benchmarked for PBF simulations. These tools were assessed for their ability to predict distortions and residual stresses in test geometries, with results validated using 3D scanning [387]. These tools can simulate the distortion caused by residual stresses during the PBF process. The accuracy of the simulations was validated by comparing the compensated parts with the original CAD design [387]. A case study involving Autodesk and MSC software demonstrated their effectiveness in predicting and compensating for distortions in PBF parts, highlighting their suitability for industrial applications [387].
OpenFOAM is a widely used open-source computational fluid dynamics (CFD) tool that has been adapted for PBF simulations. It has been modified to include the Marangoni effect, which is critical for accurately modeling heat transfer in the melt pool. The tool has been validated against experimental data, demonstrating its ability to simulate melt pool dynamics with high accuracy [388]. OpenFOAM has been enhanced to account for surface tension gradients, which are essential for modeling heat transfer in the melt pool. The tool can simulate multiple physical phenomena, including melting, solidification, and evaporation, making it suitable for complex PBF simulations. OpenFOAM’s parallel performance and scalability make it a viable option for large-scale PBF simulations [389]. OpenFOAM was used to simulate the influence of surface tension modeling on the mushy region in PBF, demonstrating improvements in the accuracy of melt pool dimensions and shape compared to experimental data [388].
AscentAM is an open-source simulation tool specifically developed for PBF processes. It uses a sequentially coupled thermomechanical approach to predict residual stresses and distortions. The tool has been validated for two benchmark geometries, demonstrating high accuracy in predicting distortion at different manufacturing states [390]. AscentAM simulates the thermal and mechanical behavior of parts during PBF, considering the physical relations of the process. The tool includes algorithms for part pre-deformation, enabling the optimization of part geometry before manufacturing. AscentAM’s predictions have been validated against experimental data, showing high accuracy in distortion prediction. AscentAM was used to simulate the distortion of a benchmark geometry, demonstrating significant reductions in form deviation through the application of optimization sub-modules [390].
Other open-source tools, such as those based on the deal.II library and particle-based methods have also been applied to PBF simulations. These tools focus on improving computational efficiency while maintaining high accuracy. For example, a highly parallelized and adaptive finite element method based on deal.II was used to analyze the accuracy and efficiency of thermomechanical simulations for Ti-6Al-4V parts [94]. These tools use adaptive mesh methods to reduce computational effort while maintaining accuracy. The use of parallel computing enables efficient simulation of large-scale PBF processes. Simplified material models and modeling assumptions are used to reduce computational costs without compromising accuracy [94]. A study using deal.II demonstrated that efficient simulations could achieve a maximum deviation of 8% in displacements and 3.5% in residual stresses compared to detailed simulations, while significantly reducing computational time [94].
Both commercial and open-source tools have demonstrated their effectiveness in simulating PBF processes, with each offering unique strengths. Commercial tools such as ANSYS Additive and Simufact Additive provide robust features for part optimization and residual stress prediction, while open-source tools like OpenFOAM and AscentAM offer cost-effective alternatives with high accuracy and scalability. The choice of tool depends on specific requirements, such as simulation accuracy, computational efficiency, and integration with experimental data.

7.2. Machine Learning Frameworks for PBF

Machine learning (ML) frameworks enable the development of data-driven surrogate models, process monitoring algorithms, and physics-informed neural networks (PINNs) for PBF applications [203,391,392]. General-purpose ML libraries provide building blocks (neural network layers, optimization algorithms, automatic differentiation) while PBF-specific implementations typically involve custom Python code, integrating these libraries with domain-specific data processing and physics constraints [367,393].
TensorFlow (Google) and PyTorch (Meta/Facebook) are widely used deep learning frameworks offering comprehensive neural network architectures (feedforward, convolutional, recurrent) and GPU acceleration capabilities. Researchers employ machine learning frameworks for developing defect detection algorithms from thermal images [270], LSTM-based temperature prediction [394], and surrogate models predicting melt pool dimensions or mechanical properties from process parameters [203,367].
Scikit-learn provides classical machine learning algorithms (support vector machines, random forests, Gaussian processes) and pre-processing utilities (feature scaling, dimensionality reduction) commonly used for PBF process parameter optimization, anomaly detection, and feature engineering [367,393,394]. Gaussian process regression (GPR) implementations enable Bayesian optimization approaches for experimental design and process parameter identification [394].
Physics-informed ML libraries embed governing equations into neural network training, improving generalization and data efficiency [391,392]. DeepXDE (Brown University) provides Python implementations of PINNs for solving PDEs, inverse problems, and data assimilation tasks [395]. These physics-informed approaches have been applied to predicting temperature distributions and thermal behavior in PBF processes [203,391].
Digital twin platforms provide infrastructure for real-time data ingestion, model orchestration, and visualization [263,396]. Cloud-based and open-source platforms are being explored for building digital representations of PBF systems, though domain-specific implementations require custom development for integrating these platforms with physics-based models and sensor interfaces [263].

7.3. Data Management and Visualization Tools

Effective data management and visualization tools are essential for handling the large volumes of sensor data, simulation results, and quality measurements generated during PBF processes. Multi-modal in situ monitoring datasets can reach hundreds of gigabytes per build. Managing and organizing these large data spaces require systematic frameworks following FAIR (Findable, Accessible, Interoperable, Reusable) principles to enable data integration and reuse [397].
In situ monitoring software integrates with machine sensors, providing real-time process visualization and anomaly detection [244,398]. Commercial solutions include Sigma Labs PrintRite3D for thermal monitoring [399], EOS EOSTATE for melt pool monitoring and exposure optimization on EOS machines [399], and various proprietary platforms from equipment manufacturers. Industry monitoring systems face challenges in data accessibility and interoperability, with data management and digital platform design being critical for qualification workflows [397,400]. Research applications often develop custom processing pipelines for sensor data acquisition, processing, and analysis, enabling flexibility for experimental investigations [244,398].
Data analysis in PBF research employs computational platforms for processing datasets spanning process parameters, sensor signals, and quality measurements. Python with scientific computing libraries (pandas for tabular data, numpy for numerical operations, scipy for scientific algorithms, matplotlib/seaborn for visualization) is widely used in research implementations due to open-source accessibility and extensive ecosystem support [244,398,401]. MATLAB remains common in academic research for signal processing and statistical analysis applications.
The visualization of three-dimensional simulation results and experimental data requires specialized software tools. Open-source visualization platforms such as ParaView and VisIt provide capabilities for large FEM/CFD datasets, though specific applications to PBF visualization workflows require custom development. Commercial FEM packages (ANSYS, Abaqus) include integrated post-processors with domain-specific visualization features for structural analysis results.

7.4. Integrated Platforms and Digital Twin Solutions

Emerging commercial solutions aim to provide end-to-end workflows spanning design, simulation, manufacturing, and quality assurance within unified platforms. Integration of the complete digital chain from design through manufacturing has been demonstrated through unified data structures and digital thread implementations [402,403,404].
Siemens NX has been employed for integrated CAD and additive manufacturing workflows including topology optimization and build preparation [402,405]. The platform enables design-for-additive-manufacturing (DfAM) workflows where simulation informs iterative design optimization [402,405]. However, for high-fidelity physics modeling, researchers often complement CAD-platform simulations with specialized finite element analysis codes [406].
Commercial CAD platforms offer additive manufacturing extensions with build orientation optimization, support generation, and distortion compensation capabilities, targeting accessibility for design engineers [407,408]. Physics fidelity and customization capabilities may be limited compared to dedicated FEM packages, leading researchers to employ specialized simulation tools for detailed analysis [406].
Research platforms typically involve custom integration of specialized tools through automated workflows and scripting interfaces [408,409]. Custom CAD-FEA integration approaches have been developed to enable DfAM-aware topology optimization and automated design workflows [409]. This approach provides flexibility and access to advanced methods but demands significant software engineering effort and development of custom data integration schemes [403,404].

8. Current Challenges and Future Directions in PBF Modeling

Significant progress in PBF modeling—from physics-based simulations to AI-augmented digital twins—has advanced understanding of process–structure–property relationships and enabled predictive process control in research settings. However, persistent challenges limit industrial deployment and certification for safety-critical applications. Table 18 systematically identifies these limitations across modeling approaches, while the following priorities outline pathways toward scalable, certifiable PBF manufacturing.
Table 18 synthesizes the principal challenges identified across modeling approaches reviewed in Section 3, Section 4, Section 5, Section 6, Section 7 and Section 8, organized by limitation category and supported by representative references.
The overarching challenge lies in achieving robust multi-scale and multiphysics coupling that seamlessly integrates thermal, mechanical, and microstructural phenomena while maintaining computational efficiency. Hybrid modeling frameworks blending physics-based simulations with machine learning surrogates show promise for balancing accuracy and speed. Figure 12 presents an overview of research directions addressing these challenges.
Achieving industrial viability requires immediate focus on validation infrastructure and computational scalability. Community-wide validation benchmarks with open-access experimental datasets—including thermal histories, defect distributions, microstructure maps, and mechanical properties—enable rigorous model verification across materials and machines. Open-source coupled simulation frameworks with GPU acceleration reduce computational barriers, while physics-informed machine learning surrogates achieve 100–1000× speed-ups, bridging high-fidelity predictions and real-time control. GPU-accelerated DEM codes enable part-scale powder spreading predictions, and integrated defect prediction algorithms combining multi-modal sensor fusion with validated ML models provide quantified detection accuracy suitable for quality assurance workflows.
Transitioning to certified manufacturing demands integrated digital twin platforms with closed-loop adaptive control validated on industrial-scale components. Real-time bidirectional synchronization enables layer-wise parameter adjustment and defect mitigation without post-process intervention. Multi-material and functionally graded modeling with validated interface predictions unlocks tailored property gradients and compositionally optimized structures. Uncertainty quantification frameworks providing statistically rigorous confidence bounds enable qualification-by-analysis pathways acceptable to regulatory agencies. Interoperable digital thread platforms with standardized data formats (FAIR compliance) facilitate seamless integration across design, manufacturing, and inspection systems. Explainable AI frameworks combining physics-informed architectures with interpretable feature attribution accelerate material qualification while maintaining regulatory transparency for safety-critical applications.
Realizing these priorities requires coordinated efforts across stakeholders. Academic institutions should prioritize open-source development and the publication of validation dataset. National laboratories should provide advanced characterization facilities, high-performance computing resources, and neutral platforms for pre-competitive consortia. The industry should contribute production process data, define acceptance criteria, and participate in standardized data format development through ASTM and ISO. Regulatory agencies should develop certification frameworks incorporating modeling evidence and establish guidelines for uncertainty quantification, supporting qualification-by-analysis pathways.
By systematically addressing these priorities, the PBF community can transform additive manufacturing into a robust, certifiable technology. The convergence of multiphysics modeling, artificial intelligence, in situ monitoring, and digital twins promises to unlock the full potential of powder bed fusion, enabling design freedoms and performance combinations unattainable through conventional manufacturing.

9. Summary

This review presents a comprehensive synthesis of modeling strategies for powder bed fusion (PBF), tracing their evolution from classical physics-based simulations to AI-augmented digital twins. By systematically evaluating advances across empirical models, numerical frameworks, multi-scale multiphysics approaches, hybrid physics–ML methods, and sensor-fusion-enabled digital twins, this work establishes how modeling has transformed understanding of process–structure–property relationships. Key contributions include the synthesis of scale-specific modeling approaches, critical assessment of coupled physics simulations, evaluation of microstructure prediction methods, analysis of in situ monitoring integration, examination of powder spreading dynamics, and review of digital twin architectures for real-time process control.
The integration of multiphysics modeling, artificial intelligence, in situ monitoring, and digital twins will enable powder bed fusion to evolve from empirical trial-and-error into a predictive, certifiable manufacturing technology, delivering design freedoms and performance characteristics beyond the reach of conventional fabrication methods.

Author Contributions

Conceptualization, Ł.Ł.; methodology, Ł.Ł.; validation, Ł.Ł. and D.S.; formal analysis, Ł.Ł.; investigation, Ł.Ł. and D.S.; resources, Ł.Ł.; data curation, Ł.Ł.; writing—original draft preparation, Ł.Ł.; writing—review and editing, D.S.; visualization, Ł.Ł. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Ministry of Science and Higher Education, Poland, Grant AGH University of Krakow no. 16.16.110.663.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

The following abbreviations are used in this manuscript:
3DThree-Dimensional
AIArtificial Intelligence
AMAdditive Manufacturing
ANNArtificial Neural Network
APIApplication Programming Interface
ARAugmented Reality
ASTMAmerican Society for Testing and Materials
CACellular Automaton
CADComputer-Aided Design
CETColumnar-to-Equiaxed Transition
CFDComputational Fluid Dynamics
CNNConvolutional Neural Network
CPFECrystal Plastivity Finite Element
CTComputed Tomography
CWContinuous Wave
DEMDiscrete Element Method
DfAMDesign for Additive Manufacturing
DMLSDirect Metal Laser Sintering
DOEDesign of Experiments
DTDigital Twin
EBSDElectron Backscatter Diffraction
EBMElectron Beam Melting
ERPEnterprise Resource Planning
FAIRFindable, Accessible, Interoperable, Reusable
FCCFace-Centered Cubic
FDMFused Deposition Modeling
FEMFinite Element Method
FGMFunctionally Graded Material
FVMFinite Volume Method
GANGenerative Adversarial Network
GNNGraph Neural Network
GPUGraphics Processing Unit
HCPHexagonal Close-Packed
HEAHigh-Entropy Alloy
ICMEIntegrated Computational Materials Engineering
IoTInternet of Things
IPFInverse Pole Figure
IRInfrared
ISOInternational Organization for Standardization
LBMLattice Boltzmann Method
LPBFLaser–Powder Bed Fusion
LSTMLong Short-Term Memory
LWIRLong-Wave Infrared
MAPEMean Absolute Percentage Error
MESManufacturing Execution System
MJFMulti Jet Fusion
MLMachine Learning
MPCModel Predictive Control
MQTTMessage Queuing Telemetry Transport
NDENon-Destructive Evaluation
NIRNear-Infrared
OPC-UAOpen Platform Communications - Unified Architecture
PBFPowder Bed Fusion
PDEPartial Differential Equation
PIDProportional–Integral–Derivative
PINNPhysics-Informed Neural Network
PLMProduct Lifecycle Management
PODProper Orthogonal Decomposition/Probability of Detection
PRISMAPreferred Reporting Items for Systematic Reviews and Meta-Analyses
PSPProcess–Structure–Property
RBFRadial Basis Function
RLReinforcement Learning
RNNRecurrent Neural Network
RSMResponse Surface Methodology
SEMScanning Electron Microscopy
SLAStereolithography
SLMSelective Laser Melting
SLSSelective Laser Sintering
SPHSmoothed Particle Hydrodynamics
SVMSupport Vector Machine
SWIRShort-Wave Infrared
TESTemperature–Emissivity Separation
TOTopology Optimization
TRLTechnology Readiness Level
XAIExplainable Artificial Intelligence
XCTX-ray Computed Tomography

References

  1. Soundararajan, B.; Sofia, D.; Barletta, D.; Poletto, M. Review on modeling techniques for powder bed fusion processes based on physical principles. Addit. Manuf. 2021, 47, 102336. [Google Scholar] [CrossRef]
  2. Sarkar, D.; Kapil, A.; Sharma, A. Advances in computational modeling for laser powder bed fusion additive manufacturing: A comprehensive review of finite element techniques and strategies. Addit. Manuf. 2024, 85, 104157. [Google Scholar] [CrossRef]
  3. Attariani, H.; Petitjean, S.R.; Dousti, M. A digital twin of synchronized circular laser array for powder bed fusion additive manufacturing. Int. J. Adv. Manuf. Technol. 2022, 123, 1433–1440. [Google Scholar] [CrossRef]
  4. Mostafaei, A.; Zhao, C.; He, Y.; Reza Ghiaasiaan, S.; Shi, B.; Shao, S.; Shamsaei, N.; Wu, Z.; Kouraytem, N.; Sun, T.; et al. Defects and anomalies in powder bed fusion metal additive manufacturing. Curr. Opin. Solid State Mater. Sci. 2022, 26, 100974. [Google Scholar] [CrossRef]
  5. Li, Y.; Mojumder, S.; Lu, Y.; Al Amin, A.; Guo, J.; Xie, X.; Chen, W.; Wagner, G.J.; Cao, J.; Liu, W.K. Statistical parameterized physics-based machine learning digital shadow models for laser powder bed fusion process. Addit. Manuf. 2024, 87, 104214. [Google Scholar] [CrossRef]
  6. Yeung, H.; Kim, F.; Donmez, A. Application of Digital Twins to Laser Powder Bed Fusion Additive Manufacturing Process Control. In Proceedings of the ASME 2023 18th International Manufacturing Science and Engineering Conference, MSEC 2023, New Brunswick, NJ, USA, 12–16 June 2023; Volume 1. [Google Scholar]
  7. Malik, A.W.; Mahmood, M.A.; Liou, F. Digital twin–driven optimization of laser powder bed fusion processes: A focus on lack-of-fusion defects. Rapid Prototyp. J. 2024, 30, 1977–1988. [Google Scholar] [CrossRef]
  8. Bayat, M.; Dong, W.; Thorborg, J.; To, A.C.; Hattel, J.H. A review of multi-scale and multi-physics simulations of metal additive manufacturing processes with focus on modeling strategies. Addit. Manuf. 2021, 47, 102278. [Google Scholar] [CrossRef]
  9. Zinovieva, O.; Romanova, V.; Dymnich, E.; Zinoviev, A.; Balokhonov, R. A Review of Computational Approaches to the Microstructure-Informed Mechanical Modelling of Metals Produced by Powder Bed Fusion Additive Manufacturing. Materials 2023, 16, 6459. [Google Scholar] [CrossRef]
  10. Chowdhury, S.; Yadaiah, N.; Prakash, C.; Ramakrishna, S.; Dixit, S.; Gupta, L.R.; Buddhi, D. Laser powder bed fusion: A state-of-the-art review of the technology, materials, properties & defects, and numerical modelling. J. Mater. Res. Technol. 2022, 20, 2109–2172. [Google Scholar] [CrossRef]
  11. Ahsan, M.M.; Liu, Y.; Raman, S.; Siddique, Z. Digital Twins in Additive Manufacturing: A Systematic Review. Internet Things 2024, 33, 101692. [Google Scholar] [CrossRef]
  12. Ben Amor, S.; Elloumi, N.; Eltaief, A.; Louhichi, B.; Alrasheedi, N.H.; Seibi, A. Digital Twin Implementation in Additive Manufacturing: A Comprehensive Review. Processes 2024, 12, 1062. [Google Scholar] [CrossRef]
  13. Crump, S.S. Apparatus and Method for Creating Three-Dimensional Objects. U.S. Patent 5,121,329, 30 October 1989. [Google Scholar]
  14. Gong, X.; Anderson, T.; Chou, K. Review on powder-based electron beam additive manufacturing technology. Manuf. Rev. 2014, 1, 2. [Google Scholar] [CrossRef]
  15. Kruth, J.P.; Levy, G.; Klocke, F.; Childs, T.H.C. Consolidation phenomena in laser and powder-bed based layered manufacturing. CIRP Ann. 2007, 56, 730–759. [Google Scholar] [CrossRef]
  16. Kruth, J.P.; Mercelis, P.; Van Vaerenbergh, J.; Froyen, L.; Rombouts, M. Binding mechanisms in selective laser sintering and selective laser melting. Rapid Prototyp. J. 2005, 11, 26–36. [Google Scholar] [CrossRef]
  17. Yap, C.Y.; Chua, C.K.; Dong, Z.L.; Liu, Z.H.; Zhang, D.Q.; Loh, L.E.; Sing, S.L. Review of selective laser melting: Materials and applications. Appl. Phys. Rev. 2015, 2, 041101. [Google Scholar] [CrossRef]
  18. Cao, S.; Zou, Y.; Lim, C.V.S.; Wu, X. Review of laser powder bed fusion (LPBF) fabricated Ti-6Al-4V: Process, post-process treatment, microstructure, and property. Light Adv. Manuf. 2021, 2, 20. [Google Scholar] [CrossRef]
  19. Martucci, A.; Aversa, A.; Lombardi, M. Ongoing Challenges of Laser-Based Powder Bed Fusion Processing of Al Alloys and Potential Solutions from the Literature—A Review. Materials 2023, 16, 1084. [Google Scholar] [CrossRef] [PubMed]
  20. Olakanmi, E.O.; Cochrane, R.F.; Dalgarno, K.W. A review on selective laser sintering/melting (SLS/SLM) of aluminium alloy powders: Processing, microstructure, and properties. Prog. Mater. Sci. 2015, 74, 401–477. [Google Scholar] [CrossRef]
  21. Aboulkhair, N.T.; Simonelli, M.; Parry, L.; Ashcroft, I.; Tuck, C.; Hague, R. 3D printing of Aluminium alloys: Additive Manufacturing of Aluminium alloys using selective laser melting. Prog. Mater. Sci. 2019, 106, 100578. [Google Scholar] [CrossRef]
  22. Zhou, Y.; Kong, D.; Li, R.; He, X.; Dong, C. Corrosion of Duplex Stainless Steel Manufactured by Laser Powder Bed Fusion: A Critical Review. Acta Metall. Sin. Engl. Lett. 2024, 37, 587–606. [Google Scholar] [CrossRef]
  23. Mansoura, A.; Omidi, N.; Barka, N.; Karganroudi, S.S.; Dehghan, S. Selective Laser Melting of Stainless Steels: A review of Process, Microstructure and Properties. Met. Mater. Int. 2024, 30, 2343–2371. [Google Scholar] [CrossRef]
  24. Narasimharaju, S.R.; Zeng, W.; See, T.L.; Zhu, Z.; Scott, P.; Jiang, X.; Lou, S. A comprehensive review on laser powder bed fusion of steels: Processing, microstructure, defects and control methods, mechanical properties, current challenges and future trends. J. Manuf. Process. 2022, 75, 375–414. [Google Scholar] [CrossRef]
  25. Haghdadi, N.; Laleh, M.; Moyle, M.; Primig, S. Additive manufacturing of steels: A review of achievements and challenges. J. Mater. Sci. 2021, 56, 64–107. [Google Scholar] [CrossRef]
  26. Ni, C.; Zhu, J.; Zhang, B.; An, K.; Wang, Y.; Liu, D.; Lu, W.; Zhu, L.; Liu, C. Recent advance in laser powder bed fusion of Ti–6Al–4V alloys: Microstructure, mechanical properties and machinability. Virtual Phys. Prototyp. 2025, 20, e2446952. [Google Scholar] [CrossRef]
  27. Dmytro, S.; Szymon, B.; Michal, K. An extended laser beam heating model for a numerical platform to simulate multi-material selective laser melting. Int. J. Adv. Manuf. Technol. 2023, 128, 3451–3470. [Google Scholar] [CrossRef]
  28. Aufa, A.N.; Hassan, M.Z.; Ismail, Z. Recent advances in Ti-6Al-4V additively manufactured by selective laser melting for biomedical implants: Prospect development. J. Alloys Compd. 2022, 896, 163072. [Google Scholar] [CrossRef]
  29. Raza, S.A.; Canyurt, O.E.; Sezer, H.K. A systematic review of Inconel 939 alloy parts development via additive manufacturing process. Heliyon 2024, 10, e25506. [Google Scholar] [CrossRef]
  30. Mostafaei, A.; Ghiaasiaan, R.; Ho, I.T.; Strayer, S.; Chang, K.C.; Shamsaei, N.; Shao, S.; Paul, S.; Yeh, A.C.; Tin, S.; et al. Additive manufacturing of nickel-based superalloys: A state-of-the-art review on process-structure-defect-property relationship. Prog. Mater. Sci. 2023, 136, 101108. [Google Scholar] [CrossRef]
  31. Sanchez, S.; Smith, P.; Xu, Z.; Gaspard, G.; Hyde, C.J.; Wits, W.W.; Ashcroft, I.A.; Chen, H.; Clare, A.T. Powder Bed Fusion of nickel-based superalloys: A review. Int. J. Mach. Tools Manuf. 2021, 165, 103729. [Google Scholar] [CrossRef]
  32. Tian, Z.; Zhang, C.; Wang, D.; Liu, W.; Fang, X.; Wellmann, D.; Zhao, Y.; Tian, Y. A Review on Laser Powder Bed Fusion of Inconel 625 Nickel-Based Alloy. Appl. Sci. 2019, 10, 81. [Google Scholar] [CrossRef]
  33. Svyetlichnyy, D. Model of the Selective Laser Melting Process-Powder Deposition Models in Multistage Multi-Material Simulations. Appl. Sci. 2023, 13, 6196. [Google Scholar] [CrossRef]
  34. Mehrpouya, M.; Tuma, D.; Vaneker, T.; Afrasiabi, M.; Bambach, M.; Gibson, I. Multimaterial powder bed fusion techniques. Rapid Prototyp. J. 2022, 28, 1–19. [Google Scholar] [CrossRef]
  35. Schneck, M.; Horn, M.; Schmitt, M.; Seidel, C.; Schlick, G.; Reinhart, G. Review on additive hybrid- and multi-material-manufacturing of metals by powder bed fusion: State of technology and development potential. Prog. Addit. Manuf. 2021, 6, 881–894. [Google Scholar] [CrossRef]
  36. Yadav, S.; Liu, S.; Singh, R.K.; Sharma, A.K.; Rawat, P. A state-of-art review on functionally graded materials (FGMs) manufactured by 3D printing techniques: Advantages, existing challenges, and future scope. J. Manuf. Process. 2024, 131, 2051–2072. [Google Scholar] [CrossRef]
  37. Li, Y.; Feng, Z.; Hao, L.; Huang, L.; Xin, C.; Wang, Y.; Bilotti, E.; Essa, K.; Zhang, H.; Li, Z.; et al. A Review on Functionally Graded Materials and Structures via Additive Manufacturing: From Multi-Scale Design to Versatile Functional Properties. Adv. Mater. Technol. 2020, 5, 1900981. [Google Scholar] [CrossRef]
  38. Svyetlichnyy, D. Lattice Boltzmann Modeling of Additive Manufacturing of Functionally Graded Materials. Entropy 2024, 27, 20. [Google Scholar] [CrossRef]
  39. Chen, B. Progress in Additive Manufacturing of High-Entropy Alloys. Materials 2024, 17, 5917. [Google Scholar] [CrossRef] [PubMed]
  40. Shi, Y.; Zheng, Y.; Huang, C.; Zhao, S.; Li, X.; Lu, Y.; Wu, Y.; Li, P.; Wang, L.; Li, Y. Influence of Processing Parameters on Additively Manufactured Architected Cellular Metals: Emphasis on Biomedical Applications. J. Funct. Biomater. 2025, 16, 53. [Google Scholar] [CrossRef]
  41. Svyetlichnyy, D. Simulation of the selective laser sintering/melting process of bioactive glass 45S5. Simul. Model. Pract. Theory 2024, 136, 103009. [Google Scholar] [CrossRef]
  42. Rahmani, R.; Lopes, S.I.; Prashanth, K.G. Selective Laser Melting and Spark Plasma Sintering: A Perspective on Functional Biomaterials. J. Funct. Biomater. 2023, 14, 521. [Google Scholar] [CrossRef]
  43. Al-Maharma, A.Y.; Patil, S.P.; Markert, B. Effects of porosity on the mechanical properties of additively manufactured components: A critical review. Mater. Res. Express 2020, 7, 122001. [Google Scholar] [CrossRef]
  44. du Plessis, A.; Yadroitsava, I.; Yadroitsev, I. Effects of defects on mechanical properties in metal additive manufacturing: A review focusing on X-ray tomography insights. Mater. Des. 2020, 187, 108385. [Google Scholar] [CrossRef]
  45. Singh, S.N.; Deoghare, A.B. Laser Shock Peening of Laser Based Directed Energy Deposition and Powder Bed Fusion Additively Manufactured Parts: A Review. Met. Mater. Int. 2022, 29, 1563–1585. [Google Scholar] [CrossRef]
  46. Bartlett, J.L.; Li, X. An overview of residual stresses in metal powder bed fusion. Addit. Manuf. 2019, 27, 131–149. [Google Scholar] [CrossRef]
  47. Malekipour, E.; El-Mounayri, H. Common defects and contributing parameters in powder bed fusion AM process and their classification for online monitoring and control: A review. Int. J. Adv. Manuf. Technol. 2017, 951, 527–550. [Google Scholar] [CrossRef]
  48. Ross, A.J.; Bitharas, I.; Perkins, K.G.; Moore, A.J. Volumetric heat source calibration for laser powder bed fusion. Addit. Manuf. 2022, 60, 103267. [Google Scholar] [CrossRef]
  49. Nastac, L. A 3D Multiscale Model for Prediction of the Microstructure Evolution in Ti6Al4V Components Produced by Laser Powder Bed Fusion. IOP Conf. Ser. Mater. Sci. Eng. 2023, 1281, 012002. [Google Scholar] [CrossRef]
  50. Malakizadi, A.; Mallipeddi, D.; Dadbakhsh, S.; M’Saoubi, R.; Krajnik, P. Post-processing of additively manufactured metallic alloys—A review. Int. J. Mach. Tools Manuf. 2022, 179, 103908. [Google Scholar] [CrossRef]
  51. Tepylo, N.; Huang, X.; Patnaik, P.C. Laser-Based Additive Manufacturing Technologies for Aerospace Applications. Adv. Eng. Mater. 2019, 21, 1900617. [Google Scholar] [CrossRef]
  52. Dubey, D.; Singh, S.P.; Behera, B.K. A review on recent advancements in additive manufacturing techniques. Proc. Inst. Mech. Eng. Part E J. Process Mech. Eng. 2024. [Google Scholar] [CrossRef]
  53. Chaudhary, S.; Avinashi, S.K.; Rao, J.; Gautam, C. Recent Advances in Additive Manufacturing, Applications and Challenges for Dentistry: A Review. ACS Biomater. Sci. Eng. 2023, 9, 3987–4019. [Google Scholar] [CrossRef]
  54. Salmi, M. Additive Manufacturing Processes in Medical Applications. Materials 2021, 14, 191. [Google Scholar] [CrossRef]
  55. Alinejadian, N.; Kollo, L.; Odnevall, I. Progress in additive manufacturing of MoS2-based structures for energy storage applications—A review. Mater. Sci. Semicond. Process. 2022, 139, 106331. [Google Scholar] [CrossRef]
  56. Srinivasan, D.; Ananth, K. Recent Advances in Alloy Development for Metal Additive Manufacturing in Gas Turbine/Aerospace Applications: A Review. J. Indian Inst. Sci. 2022, 102, 311–349. [Google Scholar] [CrossRef]
  57. Dorussen, B.J.A.; Geers, M.G.D.; Remmers, J.J.C. A discrete element framework for the numerical analysis of particle bed-based additive manufacturing processes. Eng. Comput. 2022, 38, 4753–4768. [Google Scholar] [CrossRef]
  58. Girnth, S.; Heitkamp, T.; Wacker, C.; Waldt, N.; Klawitter, G.; Dröder, K. Dimensionless quantities in discrete element method: Powder model parameterization for additive manufacturing. Prog. Addit. Manuf. 2024, 9, 1967–1983. [Google Scholar] [CrossRef]
  59. Chen, H.; Sun, Y.; Yuan, W.; Pang, S.; Yan, W.; Shi, Y. A Review on Discrete Element Method Simulation in Laser Powder Bed Fusion Additive Manufacturing. Chinese J. Mech. Eng. Addit. Manuf. Front. 2022, 1, 100017. [Google Scholar] [CrossRef]
  60. Cummins, S.; Cleary, P.W.; Delaney, G.; Phua, A.; Sinnott, M.; Gunasegaram, D.; Davies, C. A Coupled DEM/SPH Computational Model to Simulate Microstructure Evolution in Ti-6Al-4V Laser Powder Bed Fusion Processes. Metals 2021, 11, 858. [Google Scholar] [CrossRef]
  61. Aarab, M.; Dorussen, B.J.A.; Poelsma, S.S.; Remmers, J.J.C. Development of optimal L-PBF process parameters using an accelerated discrete element simulation framework. Granul. Matter 2024, 26, 69. [Google Scholar] [CrossRef]
  62. Penny, R.W.; Oropeza, D.; Praegla, P.M.; Weissbach, R.; Meier, C.; Wall, W.A.; John Hart, A. Quantitative analysis of thin metal powder layers via transmission X-ray imaging and discrete element simulation: Blade-based spreading approaches. Powder Technol. 2024, 432, 119106. [Google Scholar] [CrossRef]
  63. Fouda, Y.M.; Bayly, A.E. A DEM study of powder spreading in additive layer manufacturing. Granul. Matter 2020, 22, 10. [Google Scholar] [CrossRef]
  64. Yao, D.; An, X.; Fu, H.; Zhang, H.; Yang, X.; Zou, Q.; Dong, K. Dynamic investigation on the powder spreading during selective laser melting additive manufacturing. Addit. Manuf. 2021, 37, 101707. [Google Scholar] [CrossRef]
  65. Salim, M.A.; Tullis, S.; Elbestawi, M. Accelerating Laser Powder Bed Fusion: The Influence of Roller-Spreading Speed on Powder Spreading Performance. Metals 2024, 14, 1137. [Google Scholar] [CrossRef]
  66. Weissbach, R.; Praegla, P.M.; Wall, W.A.; Hart, A.J.; Meier, C. Exploration of improved, roller-based spreading strategies for cohesive powders in additive manufacturing via coupled DEM-FEM simulations. Powder Technol. 2024, 443, 119956. [Google Scholar] [CrossRef]
  67. Zhang, J.; Tan, Y.; Bao, T.; Xu, Y.; Xiao, X.; Jiang, S. Discrete Element Simulation of the Effect of Roller-Spreading Parameters on Powder-Bed Density in Additive Manufacturing. Materials 2020, 13, 2285. [Google Scholar] [CrossRef]
  68. Wu, Y.; Namilae, S. Parametric Analysis and Designing Maps for Powder Spreading in Metal Additive Manufacturing. Comput. Model. Eng. Sci. 2025, 142, 2067–2090. [Google Scholar] [CrossRef]
  69. Neveu, A.; Francqui, F.; Lumay, G.; Jenkins, B.; Valadian, R. Spreadability of Metal Powders: Combining Powder Characterization And DEM Simulations. In Proceedings of the Euro Powder Metallurgy 2025 Congress & Exhibition, Glasgow, UK, 14–17 September 2025. [Google Scholar]
  70. Zinatlou Ajabshir, S.; Hare, C.; Sofia, D.; Barletta, D.; Poletto, M. Investigating the effect of temperature on powder spreading behaviour in powder bed fusion additive manufacturing process by Discrete Element Method. Powder Technol. 2024, 436, 119468. [Google Scholar] [CrossRef]
  71. Roy, S.; Shaheen, M.Y.; Pöschel, T. Effect of cohesion on structure of powder layers in additive manufacturing. Granul. Matter 2023, 25, 68. [Google Scholar] [CrossRef]
  72. Zhang, T.; Chen, D.; Yang, H.; Zhao, W.; Wang, Y.; Zhou, H.; Zhang, T.; Chen, D.; Yang, H.; Zhao, W.; et al. Spreading Behavior of Non-Spherical Particles with Reconstructed Shapes Using Discrete Element Method in Additive Manufacturing. Polymers 2024, 16, 1179. [Google Scholar] [CrossRef]
  73. He, Y.; Hassanpour, A.; Bayly, A.E. Linking particle properties to layer characteristics: Discrete element modelling of cohesive fine powder spreading in additive manufacturing. Addit. Manuf. 2020, 36, 101685. [Google Scholar] [CrossRef]
  74. Yim, S.; Bian, H.; Aoyagi, K.; Yamanaka, K.; Chiba, A. Spreading behavior of Ti48Al2Cr2Nb powders in powder bed fusion additive manufacturing process: Experimental and discrete element method study. Addit. Manuf. 2022, 49, 102489. [Google Scholar] [CrossRef]
  75. Roy, S.; Pöschel, T. Shape effects in binary mixtures of PA12 powder in additive manufacturing. Powder Technol. 2024, 448, 120326. [Google Scholar] [CrossRef]
  76. Tang, C.M.; Zhao, J.B.; Zhao, Y.H.; Wang, Z.G. Discrete element analysis for effects of recoating blade and surface morphology of substrate on powder spreading quality. J. Phys. Conf. Ser. 2023, 2454, 012016. [Google Scholar] [CrossRef]
  77. Habiba, U.; Hebert, R.J. Powder Spreading Mechanism in Laser Powder Bed Fusion Additive Manufacturing: Experiments and Computational Approach Using Discrete Element Method. Materials 2023, 16, 2824. [Google Scholar] [CrossRef] [PubMed]
  78. Penny, R.W.; Praegla, P.M.; Ochsenius, M.; Oropeza, D.; Weissbach, R.; Meier, C.; Wall, W.A.; Hart, A.J. Spatial mapping of powder layer density for metal additive manufacturing via transmission X-ray imaging. Addit. Manuf. 2021, 46, 102197. [Google Scholar] [CrossRef]
  79. Chen, H.; Chen, Y.; Liu, Y.; Wei, Q.; Shi, Y.; Yan, W. Packing quality of powder layer during counter-rolling-type powder spreading process in additive manufacturing. Int. J. Mach. Tools Manuf. 2020, 153, 103553. [Google Scholar] [CrossRef]
  80. Avrampos, P.; Vosniakos, G.C. A Study on Powder Spreading Quality in Powder Bed Fusion Processes Using Discrete Element Method Simulation. J. Manuf. Mater. Process. 2024, 8, 101. [Google Scholar] [CrossRef]
  81. Cook, P.S.; Phua, A.; Davies, C.H.J.; Delaney, G.W. Modelling the influences of powder layer depth and particle morphology on powder bed fusion using a coupled DEM-CFD approach. Powder Technol. 2023, 429, 118927. [Google Scholar] [CrossRef]
  82. Xu, R.; Nan, W. Analysis of the metrics and mechanism of powder spreadability in powder-based additive manufacturing. Addit. Manuf. 2023, 71, 103596. [Google Scholar] [CrossRef]
  83. Stephan, M.; Roux, G.; Burr, A.; Ablitzer, C.; Garandet, J.P. Identification of the influential DEM contact law parameters on powder bed quality and flow in additive manufacturing configurations. Powder Technol. 2023, 429, 118937. [Google Scholar] [CrossRef]
  84. Bierwisch, C.; Dietemann, B.; Najuch, T. Particle-based modelling of laser powder bed fusion of metals with emphasis on the melting mode transition. Granul. Matter 2024, 26, 71. [Google Scholar] [CrossRef]
  85. Wang, R.; Garcia, D.; Kamath, R.R.; Dou, C.; Ma, X.; Shen, B.; Choo, H.; Fezzaa, K.; Yu, H.Z.; Kong, Z. In situ melt pool measurements for laser powder bed fusion using multi sensing and correlation analysis. Sci. Rep. 2022, 12, 13716. [Google Scholar] [CrossRef]
  86. Ren, Z.; Gao, L.; Clark, S.J.; Fezzaa, K.; Shevchenko, P.; Choi, A.; Everhart, W.; Rollett, A.D.; Chen, L.; Sun, T. Machine learning-aided real-time detection of keyhole pore generation in laser powder bed fusion. Science 2023, 379, 89–94. [Google Scholar] [CrossRef]
  87. Sanchez, S.; Rengasamy, D.; Hyde, C.J.; Figueredo, G.P.; Rothwell, B. Machine learning to determine the main factors affecting creep rates in laser powder bed fusion. J. Intell. Manuf. 2021, 32, 2353–2373. [Google Scholar] [CrossRef]
  88. Shen, T.; Li, B. An integrated computation framework for predicting mechanical performance of single-phase alloys manufactured using laser powder bed fusion: A case study of CoCrFeMnNi high-entropy alloy. Mater. Today Commun. 2024, 39, 109180. [Google Scholar] [CrossRef]
  89. Shen, T.; Li, B.; Zhang, J.; Xuan, F. Integrated computational materials engineering (ICME) for predicting tensile properties of additively manufactured defect-free single-phase high-entropy alloy. Virtual Phys. Prototyp. 2025, 20, e2441947. [Google Scholar] [CrossRef]
  90. Weirather, J.; Rozov, V.; Wille, M.; Schuler, P.; Seidel, C.; Adams, N.A.; Zaeh, M.F. A Smoothed Particle Hydrodynamics Model for Laser Beam Melting of Ni-based Alloy 718. Comput. Math. Appl. 2019, 78, 2377–2394. [Google Scholar] [CrossRef]
  91. Fuchs, S.L.; Praegla, P.M.; Cyron, C.J.; Wall, W.A.; Meier, C. A versatile SPH modeling framework for coupled microfluid-powder dynamics in additive manufacturing: Binder jetting, material jetting, directed energy deposition and powder bed fusion. Eng. Comput. 2022, 38, 4853–4877. [Google Scholar] [CrossRef]
  92. Baehr, S.; Fritz, F.; Adami, S.; Ammann, T.; Adams, N.A.; Zaeh, M.F. Investigations on the Heat Balance of the Melt Pool during PBF-LB/M under Various Process Gases. Metals 2024, 14, 1058. [Google Scholar] [CrossRef]
  93. Gu, D. Tailoring thermodynamics, fluid flow, and surface quality of laser powder bed fusion metallic powder through mass and momentum transfer modeling and monitoring. In Laser Additive Manufacturing of Metallic Materials and Components; Elsevier: Amsterdam, The Netherlands, 2023; pp. 429–482. ISBN 978-0-12-823783-0. [Google Scholar]
  94. Burkhardt, C.; Steinmann, P.; Mergheim, J. Thermo-mechanical simulations of powder bed fusion processes: Accuracy and efficiency. Adv. Model. Simul. Eng. Sci. 2022, 9, 18. [Google Scholar] [CrossRef]
  95. Vallabh, C.K.P.; Zhao, X. Single-camera Two-Wavelength Imaging Pyrometry for Melt Pool Temperature Measurement and Monitoring in Laser Powder Bed Fusion based Additive Manufacturing. J. Manuf. Process. 2021, 79, 486–500. [Google Scholar] [CrossRef]
  96. Flint, T.F.; Robson, J.D.; Esmati, P.; Grilli, N.; Parivendhan, G.; Cardiff, P. Version 2.0—LaserbeamFoam: Laser ray-tracing and thermally induced state transition simulation toolkit. SoftwareX 2024, 25, 101612. [Google Scholar] [CrossRef]
  97. Gusarov, A.V. Radiative transfer, absorption, and reflection by metal powder beds in laser powder-bed processing. J. Quant. Spectrosc. Radiat. Transf. 2020, 257, 107366. [Google Scholar] [CrossRef]
  98. Gu, D. Laser energy absorption and distribution behaviors of randomly packed powder bed using ray tracing method during laser powder bed fusion of metallic materials. In Laser Additive Manufacturing of Metallic Materials and Components; Elsevier: Amsterdam, The Netherlands, 2023; pp. 307–345. [Google Scholar]
  99. Liu, L.; Huang, M.; Ma, Y.H.; Qin, M.L.; Liu, T.T. Simulation of Powder Packing and Thermo-Fluid Dynamic of 316L Stainless Steel by Selective Laser Melting. J. Mater. Eng. Perform. 2020, 29, 7369–7381. [Google Scholar] [CrossRef]
  100. Katagiri, J.; Nomoto, S.; Kusano, M.; Watanabe, M. Particle Size Effect on Powder Packing Properties and Molten Pool Dimensions in Laser Powder Bed Fusion Simulation. J. Manuf. Mater. Process. 2024, 8, 71. [Google Scholar] [CrossRef]
  101. Yuan, W.; Chen, H.; Peng, C.; Lupoi, R.; Yin, S. Revealing melt-vapor-powder interaction towards laser powder bed fusion process via DEM-CFD coupled model. Surf. Sci. Technol. 2023, 1, 14. [Google Scholar] [CrossRef]
  102. Srivastava, S.; Garg, R.K.; Sharma, V.S.; Alba-Baena, N.G.; Sachdeva, A.; Chand, R.; Singh, S. Multi-physics continuum modelling approaches for metal powder additive manufacturing: A review. Rapid Prototyp. J. 2020, 26, 737–764. [Google Scholar] [CrossRef]
  103. Jia, Y.; Naceur, H.; Saadlaoui, Y.; Dubar, L.; Bergheau, J.M. A comprehensive comparison of modeling strategies and simulation techniques applied in powder-based metallic additive manufacturing processes. J. Manuf. Process. 2024, 110, 1–29. [Google Scholar] [CrossRef]
  104. Mayi, Y.A.; Dal, M.; Peyre, P.; Bellet, M.; Metton, C.; Moriconi, C.; Fabbro, R. Transient dynamics and stability of keyhole at threshold in laser powder bed fusion regime investigated by finite element modeling. J. Laser Appl. 2021, 33, 012024. [Google Scholar] [CrossRef]
  105. Tan, W.; Spear, A. Multiphysics Modeling Framework to Predict Process-Microstructure-Property Relationship in Fusion-Based Metal Additive Manufacturing. Acc. Mater. Res. 2024, 5, 10–21. [Google Scholar] [CrossRef]
  106. Afrasiabi, M.; Bambach, M. Modelling and simulation of metal additive manufacturing processes with particle methods: A review. Virtual Phys. Prototyp. 2023, 18, e2274494. [Google Scholar] [CrossRef]
  107. Korneev, B.; Zakirov, A.; Bogdanova, M.; Belousov, S.; Perepelkina, A.; Iskandarova, I.; Potapkin, B. A numerical study of powder wetting influence on the morphology of laser powder bed fusion manufactured thin walls. Addit. Manuf. 2023, 74, 103705. [Google Scholar] [CrossRef]
  108. Zöller, C.; Adams, N.A.; Adami, S. Beam-shaping in laser-based powder bed fusion of metals: A computational analysis of point-ring intensity profiles. Addit. Manuf. 2024, 93, 104402. [Google Scholar] [CrossRef]
  109. Gu, H.; Fu, Y.; Wei, C.; Li, L.; Qian, L.; Zhou, W.; Li, Y.; Ren, X. The impact of thermocapillary on equiaxed/columnar microstructure evolution in laser powder bed fusion: A high-fidelity ray-tracing based finite volume and cellular automaton study. J. Mater. Process. Technol. 2024, 326, 118335. [Google Scholar] [CrossRef]
  110. Lu, H.; Pan, J.; Gu, Y.; Hu, X.; Duan, Z.; Li, H. Effect of molten pool flow and recoil pressure on grain growth structures during laser powder bed fusion by an integrated model. Int. J. Heat Mass Transf. 2024, 223, 125219. [Google Scholar] [CrossRef]
  111. Laskowski, R.; Ahluwalia, R.; Hock, G.T.W.; Ying, C.S.; Sun, C.N.; Wang, P.; Cheh, D.T.C.; Sharon, N.M.L.; Vastola, G.; Zhang, Y.W. Concurrent modeling of porosity and microstructure in multilayer three-dimensional simulations of powder-bed fusion additive manufacturing of INCONEL 718. Addit. Manuf. 2022, 60, 103266. [Google Scholar] [CrossRef]
  112. Sharma, S.; Joshi, S.S.; Pantawane, M.V.; Radhakrishnan, M.; Mazumder, S.; Dahotre, N.B. Multiphysics multi-scale computational framework for linking process–structure–property relationships in metal additive manufacturing: A critical review. Int. Mater. Rev. 2023, 68, 943–1009. [Google Scholar] [CrossRef]
  113. Yan, J.; Zhu, Q.; Zhao, Z. Computational thermal multi-phase flow for metal additive manufacturing. In Modeling and Simulation in Science, Engineering and Technology; Birkhäuser: Cham, Switzerland, 2022; pp. 533–580. [Google Scholar] [CrossRef]
  114. Meier, C.; Penny, R.W.; Zou, Y.; Gibbs, J.S.; Hart, A.J. Thermophysical phenomena in metal additive manufacturing by selective laser melting: Fundamentals, modeling, simulation, and experimentation. Annu. Rev. Heat Transf. 2017, 20, 241–316. [Google Scholar] [CrossRef]
  115. Wang, J.; Zhu, R.; Liu, Y.; Zhang, L. Understanding melt pool characteristics in laser powder bed fusion: An overview of single- and multi-track melt pools for process optimization. Adv. Powder Mater. 2023, 2, 100137. [Google Scholar] [CrossRef]
  116. Shan, X.; Pan, Z.; Gao, M.; Han, L.; Choi, J.-P.; Zhang, H. Multi-Physics Modeling of Melting-Solidification Characteristics in Laser Powder Bed Fusion Process of 316L Stainless Steel. Materials 2024, 17, 946. [Google Scholar] [CrossRef] [PubMed]
  117. Dai, K.; He, X.; Kong, D.; Dong, C. Multi-physical field simulation to yield defect-free IN718 alloy fabricated by laser powder bed fusion. Mater. Lett. 2024, 355, 135437. [Google Scholar] [CrossRef]
  118. Promoppatum, P.; Yao, S.-C. Analytical evaluation of defect generation for selective laser melting of metals. Int. J. Adv. Manuf. Technol. 2019, 103, 1185–1198. [Google Scholar] [CrossRef]
  119. Agrawal, A.K.; Rankouhi, B.; Thoma, D.J. Predictive process mapping for laser powder bed fusion: A review of existing analytical solutions. Curr. Opin. Solid State Mater. Sci. 2022, 26, 101024. [Google Scholar] [CrossRef]
  120. Ökten, K.; Örnek, B.N.; Biyikoğlu, A. Development of an analytical model to investigate temperature distribution and melt geometry in the laser powder bed fusion process under different operating parameters. J. Laser Appl. 2025, 37, 012001. [Google Scholar] [CrossRef]
  121. Estupinan Donoso, A.A.; Peters, B. Exploring a Multiphysics Resolution Approach for Additive Manufacturing. JOM 2018, 70, 1604–1610. [Google Scholar] [CrossRef]
  122. Wang, L.; Guo, Q.; Chen, L.; Yan, W. In-situ experimental and high-fidelity modeling tools to advance understanding of metal additive manufacturing. Int. J. Mach. Tools Manuf. 2023, 193, 104077. [Google Scholar] [CrossRef]
  123. Krauss, H.; Zeugner, T.; Zaeh, M.F. Thermographic process monitoring in powderbed based additive manufacturing. AIP Conf. Proc. 2015, 1650, 177–183. [Google Scholar] [CrossRef]
  124. Moshiri, M.; Pedersen, D.B.; Tosello, G.; Nadimpalli, V.K. Performance evaluation of in-situ near-infrared melt pool monitoring during laser powder bed fusion. Virtual Phys. Prototyp. 2023, 18, e2205387. [Google Scholar] [CrossRef]
  125. Varma, A.; Ganesh, K.V.; Mahapatra, D.R. Multi-Scale Modeling of Selective Laser Melting Process; SAE Technical Paper; SAE International: Warrendale, PA, USA, 2024. [Google Scholar] [CrossRef]
  126. Caprio, L.; Demir, A.G.; Previtali, B. Understanding the effects of temporal waveform modulation of the laser emission power in laser powder bed fusion: Part I—Analytical modelling. J. Phys. D Appl. Phys. 2022, 55, 495101. [Google Scholar] [CrossRef]
  127. Parsazadeh, M.; Ebrahimi, H.; Sichani, M.S.; Dahotre, N. A Novel Physics-Based Model for Predicting Melt Pool Dimensions in Laser Powder Bed Fusion Process. J. Manuf. Sci. Eng. 2024, 146, 081001. [Google Scholar] [CrossRef]
  128. Wang, W.; Ning, J.; Garmestani, H.; Liang, S.Y. Analytical Prediction of Molten Pool Dimensions in Powder Bed Fusion Considering Process Conditions-Dependent Laser Absorptivity. Appl. Sci. 2022, 12, 11926. [Google Scholar] [CrossRef]
  129. Queva, A.; Guillemot, G.; Moriconi, C.; Bergeron, R.; Bellet, M. Mesoscale multilayer multitrack modeling of melt pool physics in laser powder bed fusion of lattice metal features. Addit. Manuf. 2024, 93, 104365. [Google Scholar] [CrossRef]
  130. Gu, D. Mesoscale understanding of particle melting behavior, thermodynamic mechanism, and porosity evolution of randomly packed powder-bed using finite volume method. In Laser Additive Manufacturing of Metallic Materials and Components; Elsevier: Amsterdam, The Netherlands, 2023; pp. 347–395. [Google Scholar]
  131. Queva, A.; Guillemot, G.; Moriconi, C.; Metton, C.; Bellet, M. Numerical study of the impact of vaporisation on melt pool dynamics in Laser Powder Bed Fusion—Application to IN718 and Ti–6Al–4V. Addit. Manuf. 2020, 35, 101249. [Google Scholar] [CrossRef]
  132. Trejos-Taborda, J.; Reyes-Osorio, L.; Garza, C.; del Carmen Zambrano-Robledo, P.; Lopez-Botello, O. Finite element modeling of melt pool dynamics in laser powder bed fusion of 316L stainless steel. Int. J. Adv. Manuf. Technol. 2022, 120, 3947–3961. [Google Scholar] [CrossRef]
  133. Abdelmoula, M.; Musinski, W. Predicting and correlating melt pool characteristics with processing parameters in IN625 powder bed fusion additive manufacturing. J. Manuf. Process. 2024, 121, 427–445. [Google Scholar] [CrossRef]
  134. Peng, J.; Jiang, W.G.; Qin, Q.H.; Xu, G.G.; Liu, F.C. High-fidelity numerical model of melt pool dynamics in selective multi-laser melting of Inconel 718 under different scanning strategies. Mater. Today Commun. 2023, 36, 106913. [Google Scholar] [CrossRef]
  135. Khorasani, M.; Ghasemi, A.H.; Leary, M.; Downing, D.; Gibson, I.; Sharabian, E.G.; Veetil, J.K.; Brandt, M.; Bateman, S.; Rolfe, B. Benchmark models for conduction and keyhole modes in laser-based powder bed fusion of Inconel 718. Opt. Laser Technol. 2023, 164, 109509. [Google Scholar] [CrossRef]
  136. Zöller, C.; Adams, N.A.; Adami, S. Numerical investigation of balling defects in laser-based powder bed fusion of metals with Inconel 718. Addit. Manuf. 2023, 73, 103658. [Google Scholar] [CrossRef]
  137. Krzyzanowski, M.; Svyetlichnyy, D. A multiphysics simulation approach to selective laser melting modelling based on cellular automata and lattice Boltzmann methods. Comput. Part. Mech. 2022, 9, 117–133. [Google Scholar] [CrossRef]
  138. Wei, M.; Ding, W.J.; Vastola, G.; Zhang, Y.W. Quantitative study on the dynamics of melt pool and keyhole and their controlling factors in metal laser melting. Addit. Manuf. 2022, 54, 102779. [Google Scholar] [CrossRef]
  139. Chouhan, A.; Mädler, L.; Ellendt, N. Modeling of rapid solidification in Laser Powder Bed Fusion processes. Comput. Mater. Sci. 2024, 238, 112918. [Google Scholar] [CrossRef]
  140. Manav, M.; Perraudin, N.; Lin, Y.; Afrasiabi, M.; Perez-Cruz, F.; Bambach, M.; De Lorenzis, L. MeltpoolINR: Predicting temperature field, melt pool geometry, and their rate of change in laser powder bed fusion. J. Intell. Manuf. 2025. [Google Scholar] [CrossRef]
  141. Hemmasian, A.; Ogoke, F.; Akbari, P.; Malen, J.; Beuth, J.; Farimani, A.B. Surrogate Modeling of Melt Pool Thermal Field using Deep Learning. Addit. Manuf. Lett. 2022, 5, 100123. [Google Scholar] [CrossRef]
  142. Lopez, E.D.; Fortney, J.J.; Kalourazi, S.F.; Wang, F.; Zhang, H.; Li, Z.; Greenwood, M.; Phillion, A. A phase field methodology for simulating the microstructure evolution during laser powder bed fusion in-situ alloying process. IOP Conf. Ser. Mater. Sci. Eng. 2023, 1281, 012017. [Google Scholar] [CrossRef]
  143. Yang, M.; Wang, L.; Yan, W. Phase-field modeling of grain evolutions in additive manufacturing from nucleation, growth, to coarsening. Npj Comput. Mater. 2021, 7, 56. [Google Scholar] [CrossRef]
  144. Elahi, S.M.; Tavakoli, R.; Romero, I.; Tourret, D. Grain growth competition during melt pool solidification—Comparing phase-field and cellular automaton models. Comput. Mater. Sci. 2022, 216, 56. [Google Scholar] [CrossRef]
  145. Yang, Y.; Kühn, P.; Yi, M.; Egger, H.; Xu, B.X. Non-isothermal Phase-Field Modeling of Heat–Melt–Microstructure-Coupled Processes During Powder Bed Fusion. JOM 2020, 72, 1719–1733. [Google Scholar] [CrossRef]
  146. Choi, J.Y.; Xue, T.; Liao, S.; Cao, J. Accelerating phase-field simulation of three-dimensional microstructure evolution in laser powder bed fusion with composable machine learning predictions. Addit. Manuf. 2024, 79, 103938. [Google Scholar] [CrossRef]
  147. Peivaste, I.; Siboni, N.H.; Alahyarizadeh, G.; Ghaderi, R.; Svendsen, B.; Raabe, D.; Mianroodi, J.R. Accelerating phase-field-based simulation via machine learning. arXiv 2022, arXiv:2205.02121. [Google Scholar]
  148. Xie, Q.; Yan, X.; Yin, S.; Li, X.; Liu, M.; Wang, Y.; Zhuang, L. An extended version of cellular automata model for powder bed fusion to unravel the dependence of microstructure on printing areas for Inconel 625. Addit. Manuf. 2023, 73, 103676. [Google Scholar] [CrossRef]
  149. Andersson, C.; Lundbäck, A. Modeling the Evolution of Grain Texture during Solidification of Laser-Based Powder Bed Fusion Manufactured Alloy 625 Using a Cellular Automata Finite Element Model. Metals 2023, 13, 1846. [Google Scholar] [CrossRef]
  150. Zheng, M.; Wang, G.; Zhou, W.; Wei, L.; Lin, X.; Huang, W. Understanding grain evolution in laser powder bed fusion process through a real-time coupled Lattice Boltzmann model-Cellular Automaton simulation. J. Mater. Process. Technol. 2023, 321, 118126. [Google Scholar] [CrossRef]
  151. Nastac, L. 3D Modeling of the Solidification Structure Evolution and of the Inter Layer/Track Voids Formation in Metallic Alloys Processed by Powder Bed Fusion Additive Manufacturing. Materials 2022, 15, 8885. [Google Scholar] [CrossRef] [PubMed]
  152. Nastac, L. 3D Modeling of the Solidification Structure Evolution of Superalloys in Powder Bed Fusion Additive Manufacturing Processes. Metals 2021, 11, 1995. [Google Scholar] [CrossRef]
  153. Yang, Z.Y.; Kuesters, Y.K.; Logvinov, R.L.; Markl, M.M.; Körner, C.K. SAMPLE3D: A Versatile Numerical Tool for Investigating Texture and Grain Structure of Materials Processed by PBF Processes. In Proceedings of the IVth International Conference on Simulation for Additive Manufacturing (Sim-AM 2023), Munich, Germany, 26–28 July 2023. [Google Scholar]
  154. Maji, S.; Amirthalingam, M. Multi-component and multi-phase-field modelling of solidification microstructural evolution in Inconel 625 alloy during laser powder bed fusion additive manufacturing. Weld. World 2025, 69, 1023–1043. [Google Scholar] [CrossRef]
  155. Okugawa, M.; Saito, K.; Yoshima, H.; Sawaizumi, K.; Nomoto, S.; Watanabe, M.; Nakano, T.; Koizumi, Y. Solute segregation in a rapidly solidified Hastelloy-X Ni-based superalloy during laser powder bed fusion investigated by phase-field and computational thermal-fluid dynamics simulations. Addit. Manuf. 2024, 84, 104079. [Google Scholar] [CrossRef]
  156. Mohammadkamal, H.; Caiazzo, F. Numerical Study to Analyze the Influence of Process Parameters on Temperature and Stress Field in Powder Bed Fusion of Ti-6Al-4V. Materials 2025, 18, 368. [Google Scholar] [CrossRef] [PubMed]
  157. Denlinger, E.R.; Gouge, M.; Irwin, J.; Michaleris, P. Thermomechanical model development and in situ experimental validation of the Laser Powder-Bed Fusion process. Addit. Manuf. 2017, 16, 73–80. [Google Scholar] [CrossRef]
  158. Cao, Y.; Lin, X.; Kang, N.; Ma, L.; Wei, L.; Zheng, M.; Yu, J.; Peng, D.; Huang, W. A novel high-efficient finite element analysis method of powder bed fusion additive manufacturing. Addit. Manuf. 2021, 46, 102187. [Google Scholar] [CrossRef]
  159. Olleak, A.; Xi, Z. A study of modeling assumptions and adaptive remeshing for thermomechanical finite element modeling of the LPBF process. Int. J. Adv. Manuf. Technol. 2021, 115, 3599–3615. [Google Scholar] [CrossRef]
  160. Chen, Q.; Liang, X.; Hayduke, D.; Liu, J.; Cheng, L.; Oskin, J.; Whitmore, R.; To, A.C. An inherent strain based multiscale modeling framework for simulating part-scale residual deformation for direct metal laser sintering. Addit. Manuf. 2019, 28, 406–418. [Google Scholar] [CrossRef]
  161. Bayraktar, C.; Demir, E. A thermomechanical finite element model and its comparison to inherent strain method for powder-bed fusion process. Addit. Manuf. 2022, 54, 102708. [Google Scholar] [CrossRef]
  162. Zhang, M.; Ji, C.; Hou, Y.; Jin, P.; He, J.; Wu, J.; Li, K. Modified inherent strain method coupled with shear strain and dynamic mechanical properties for predicting residual deformation of Inconel 738LC part fabricated by laser powder bed fusion. Int. J. Therm. Sci. 2024, 203, 109163. [Google Scholar] [CrossRef]
  163. Liang, X.; Dong, W.; Chen, Q.; To, A.C. On incorporating scanning strategy effects into the modified inherent strain modeling framework for laser powder bed fusion. Addit. Manuf. 2021, 37, 101648. [Google Scholar] [CrossRef]
  164. Dong, W.; Liang, X.; Chen, Q.; Hinnebusch, S.; Zhou, Z.; To, A.C. A new procedure for implementing the modified inherent strain method with improved accuracy in predicting both residual stress and deformation for laser powder bed fusion. Addit. Manuf. 2021, 47, 102345. [Google Scholar] [CrossRef]
  165. Bhat, A.; Elsen, R.S.; Abishai, D.; Narayan, M.J.; Sakthivel, A.R.; Chadha, U.; Hirpha, B.B. Prediction and Experimental Verification of Distortion due to Residual Stresses in a Ti-6Al-4V Control Arm Plate. Adv. Mater. Sci. Eng. 2022, 2022, 5211623. [Google Scholar] [CrossRef]
  166. Zhou, J.; Barrett, R.A.; Leen, S.B. Finite element modelling for mitigation of residual stress and distortion in macro-scale powder bed fusion components. Proc. Inst. Mech. Eng. Part L J. Mater. Des. Appl. 2023, 237, 1458–1474. [Google Scholar] [CrossRef]
  167. Hassila, C.J.; Malmelöv, A.; Andersson, C.; Hektor, J.; Fisk, M.; Lundbäck, A.; Wiklund, U. Influence of Scanning Strategy on Residual Stresses in Laser-Based Powder Bed Fusion Manufactured Alloy 718: Modeling and Experiments. Materials 2024, 17, 6265. [Google Scholar] [CrossRef] [PubMed]
  168. Palmeri, D.; Pollara, G.; Licari, R.; Micari, F. Finite Element Method in L-PBF of Ti-6Al-4V: Influence of Laser Power and Scan Speed on Residual Stress and Part Distortion. Metals 2023, 13, 1907. [Google Scholar] [CrossRef]
  169. Anand, N.; Chen, Y.-B.; Lee, M.-T. The Development of an Effective and Comprehensive Modeling Technique for Thermomechanical Analysis of Selective Laser Melting Process. In Proceedings of the Conference: International Conference of Asian Society for Precision Engineering and Nanotechnology, Hong Kong, China, 21–24 November 2023; pp. 123–126. [Google Scholar]
  170. Kumar, B.; Dandekar, T.R.; Birosca, S. Probing the temperature field and residual stress transformation in multi-track, multi-layered system. Mater. Today Commun. 2024, 38, 108056. [Google Scholar] [CrossRef]
  171. Tangestani, R.; Sabiston, T.; Chakraborty, A.; Muhammad, W.; Yuan, L.; Martin, É. An Efficient Track-Scale Model for Laser Powder Bed Fusion Additive Manufacturing: Part 1-Thermal Model. Front. Mater. 2021, 8, 753040. [Google Scholar] [CrossRef]
  172. Conti, P.; Cianetti, F.; Pilerci, P. Parametric Finite Elements Model of SLM Additive Manufacturing process. Procedia Struct. Integr. 2018, 8, 410–421. [Google Scholar] [CrossRef]
  173. Jonaet, A.M.; Park, H.S.; Myung, L.C. Prediction of residual stress and deformation based on the temperature distribution in 3D-printed parts. Int. J. Adv. Manuf. Technol. 2021, 113, 2227–2242. [Google Scholar] [CrossRef]
  174. Santos, M.O.; Maiolini, A.S.F.R.; Miranda, F.; Farias, A.; Seriacopi, V.; Bordinassi, E.C.; Batalha, G.F. Numerical and Experimental Analysis of the Influence of Manufacturing Parameters in Additive Manufacturing SLM-PBF on Residual Stress and Thermal Distortion in Parts of Titanium Alloy Ti6Al4V. In Proceedings of the Lecture Notes in Mechanical Engineering; Springer: Cham, Switzerland, 2024; pp. 195–207. [Google Scholar]
  175. Cheng, B.; Shrestha, S.; Chou, K. Stress and deformation evaluations of scanning strategy effect in selective laser melting. Addit. Manuf. 2016, 12, 240–251. [Google Scholar] [CrossRef]
  176. Jia, Y.; Zeng, C.; Xue, J. Scanning strategy optimization for the selective laser melting additive manufacturing of Ti6Al4V. Eng. Res. Express 2023, 5, 015041. [Google Scholar] [CrossRef]
  177. Chen, Q.; Taylor, H.; Takezawa, A.; Liang, X.; Jimenez, X.; Wicker, R.; To, A.C. Island scanning pattern optimization for residual deformation mitigation in laser powder bed fusion via sequential inherent strain method and sensitivity analysis. Addit. Manuf. 2021, 46, 102116. [Google Scholar] [CrossRef]
  178. Phan, M.A.L.; Dew, O.; Todd, I. Predictive process diagram for parameters selection in laser powder bed fusion to achieve high-density and low-cracking built parts. Addit. Manuf. 2024, 85, 104145. [Google Scholar] [CrossRef]
  179. Prasanth Kumar, R.K.; Boggarapu, N.R.; Narayana Murty, S.V.S. Multi-objective optimization to specify optimal selective laser melting process parameters for SS316 L powder. Multidiscip. Model. Mater. Struct. 2024, 20, 59–80. [Google Scholar] [CrossRef]
  180. Mani, M.; Feng, S.; Brandon, L.; Donmez, A.; Moylan, S.; Fesperman, R. Measurement science needs for real-time control of additive manufacturing powder-bed fusion processes. In Additive Manufacturing Handbook; CRC Press: Boca Raton, FL, USA, 2017; pp. 629–652. ISBN 9781315119106. [Google Scholar]
  181. Moges, T.; Ameta, G.; Witherell, P. A Review of Model Inaccuracy and Parameter Uncertainty in Laser Powder Bed Fusion Models and Simulations. J. Manuf. Sci. Eng. Trans. ASME 2019, 141, 040801. [Google Scholar] [CrossRef]
  182. Mwania, F.M.; Maringa, M.; van der Walt, J. Analytical and numerical modelling of laser powder bed fusion (L-PBF) of polymers—A review of the key areas of focus of the process. MATEC Web Conf. 2022, 370, 06001. [Google Scholar] [CrossRef]
  183. Puso, M.A.; Hodge, N.E. An assessment of the utility of multirate time integration for the modeling of laser powder bed fusion. Addit. Manuf. 2023, 73, 103657. [Google Scholar] [CrossRef]
  184. Tong, M. Review of Particle-Based Computational Methods and Their Application in the Computational Modelling of Welding, Casting and Additive Manufacturing. Metals 2023, 13, 1392. [Google Scholar] [CrossRef]
  185. Afrasiabi, M.; Lüthi, C.; Bambach, M.; Wegener, K. Multi-resolution SPH simulation of a laser powder bed fusion additive manufacturing process. Appl. Sci. 2021, 11, 2962. [Google Scholar] [CrossRef]
  186. Liu, W.; Wu, C.Y. Modelling Complex Particle–Fluid Flow with a Discrete Element Method Coupled with Lattice Boltzmann Methods (DEM-LBM). ChemEngineering 2020, 4, 55. [Google Scholar] [CrossRef]
  187. Le, T.N.; Lo, Y.L.; Hung, W. Coupled Computational Fluid Dynamics-Discrete Element Method Model for Investigation of Powder Effects in Nonconventional Laser Powder Bed Fusion Process. 3D Print. Addit. Manuf. 2024, 11, e1656–e1669. [Google Scholar] [CrossRef] [PubMed]
  188. Svyetlichnyy, D.; Krzyzanowski, M.; Straka, R.; Lach, L.; Rainforth, W.M. Application of cellular automata and Lattice Boltzmann methods for modelling of additive layer manufacturing. Int. J. Numer. Methods Heat Fluid Flow 2018, 28, 31–46. [Google Scholar] [CrossRef]
  189. Zhang, B.; Seede, R.; Whitt, A.; Shoukr, D.; Huang, X.; Karaman, I.; Arroyave, R.; Elwany, A. A printability assessment framework for fabricating low variability nickel-niobium parts using laser powder bed fusion additive manufacturing. Rapid Prototyp. J. 2021, 27, 1737–1748. [Google Scholar] [CrossRef]
  190. Holla, V.; Lane, B.; Redford, J.; Kopp, P.; Kollmannsberger, S.; Article, R. Quantifying Thermal Model Accuracy in PBF-LB/M using Statistical Similarity Tests Against Thermographic Measurements. Integr. Mater. Manuf. Innov. 2025. [Google Scholar] [CrossRef]
  191. Yan, W.; Lu, Y.; Jones, K.; Yang, Z.; Fox, J.; Witherell, P.; Wagner, G.; Liu, W.K. Data-driven characterization of thermal models for powder-bed-fusion additive manufacturing. Addit. Manuf. 2020, 36, 101503. [Google Scholar] [CrossRef]
  192. Imani Shahabad, S.; Karimi, G.; Toyserkani, E. An Extended Rosenthal’s Model for Laser Powder-Bed Fusion Additive Manufacturing: Energy Auditing of Thermal Boundary Conditions. Lasers Manuf. Mater. Process. 2021, 8, 288–311. [Google Scholar] [CrossRef]
  193. Zagade, P.R.; Gautham, B.P.; De, A.; DebRoy, T. Analytical modelling of scanning strategy effect on temperature field and melt track dimensions in laser powder bed fusion. Addit. Manuf. 2024, 82, 104046. [Google Scholar] [CrossRef]
  194. Cooke, S.R.; Sinclair, C.W.; Maijer, D.M. Incorporating non-linear effects in fast semi-analytical thermal modelling of powder bed fusion. arXiv 2024, arXiv:2404.03018. [Google Scholar] [CrossRef]
  195. Mahmood, M.A.; Ur Rehman, A.; Khan, T.; Seers, T.D.; Liou, F.; Khraisheh, M. Defects quantification of additively manufactured AISI 316L stainless steel parts via non-destructive analyses: Experiments and semi-FEM-analytical-based modeling. Opt. Laser Technol. 2024, 174, 110684. [Google Scholar] [CrossRef]
  196. Li, Z.J.; Dai, H.L.; Yao, Y.; Liu, J.L. A semi-analytical model for rapid prediction of residual stress and deformation in laser powder bed fusion. Appl. Math. Model. 2024, 125, 672–686. [Google Scholar] [CrossRef]
  197. Olleak, A.; Xi, Z. Efficient LPBF process simulation using finite element modeling with adaptive remeshing for distortions and residual stresses prediction. Manuf. Lett. 2020, 24, 140–144. [Google Scholar] [CrossRef]
  198. Ilani, M.A.; Banad, Y.M. Modeling Melt Pool Geometry in Metal Additive Manufacturing Using Goldak’s Semi-Ellipsoidal Heat Source: A Data-driven Computational Approach. arXiv 2024, arXiv:2404.08834. [Google Scholar]
  199. Wang, J.; Wang, Y.; Shi, J. On efficiency and effectiveness of finite volume method for thermal analysis of selective laser melting. Eng. Comput. 2020, 37, 2155–2175. [Google Scholar] [CrossRef]
  200. Leclerc, W.; Chams, A. Discrete Element Method framework to simulate metallic Laser Powder-Bed Fusion additive manufacturing process. In Proceedings of the Fourteenth International Conference on Computational Structures Technology; Civil-Comp Press: Edinburgh, UK, 2023; Volume 3, pp. 1–7. [Google Scholar] [CrossRef]
  201. Kouraytem, N.; Li, X.; Tan, W.; Kappes, B.; Spear, A.D. Modeling process–structure–property relationships in metal additive manufacturing: A review on physics-driven versus data-driven approaches. J. Phys. Mater. 2021, 4, 032002. [Google Scholar] [CrossRef]
  202. Sharma, R.; Guo, Y.B. Computational modeling and physics-informed machine learning of metal additive manufacturing: State-of-the-art and future perspective. Annu. Rev. Heat Transf. 2021, 24, 303–337. [Google Scholar] [CrossRef]
  203. Riensche, A.; Bevans, B.; King, G.; Krishnan, A.; Cole, K.; Rao, P. Predicting Microstructure Evolution in Laser Powder Bed Fusion Additive Manufacturing Using Physics-Based Machine Learning. In Proceedings of the ASME 2024 19th International Manufacturing Science and Engineering Conference, MSEC 2024, Knoxville, TN, USA, 17–21 June 2024; Volume 1. [Google Scholar]
  204. Moges, T.; Yang, Z.; Jones, K.; Feng, S.; Witherell, P.; Lu, Y. Hybrid Modeling Approach for Melt-Pool Prediction in Laser Powder Bed Fusion Additive Manufacturing. J. Comput. Inf. Sci. Eng. 2021, 21, 050902. [Google Scholar] [CrossRef]
  205. Liao, S.; Xue, T.; Jeong, J.; Webster, S.; Ehmann, K.; Cao, J. Hybrid thermal modeling of additive manufacturing processes using physics-informed neural networks for temperature prediction and parameter identification. Comput. Mech. 2023, 72, 499–512. [Google Scholar] [CrossRef]
  206. La Fé-Perdomo, I.; Ramos-Grez, J.A. An insight into adiabatic efficiency hybrid modeling in Laser-Based Powder Bed Fusion of 316L stainless steel. Opt. Laser Technol. 2023, 161, 109203. [Google Scholar] [CrossRef]
  207. ZainElabdeen, I.H.; Ismail, L.; Mohamed, O.F.; Khan, K.A.; Schiffer, A. Recent advancements in hybrid additive manufacturing of similar and dissimilar metals via laser powder bed fusion. Mater. Sci. Eng. A 2024, 909, 146833. [Google Scholar] [CrossRef]
  208. Gong, J.; Wei, K.; Liu, M.; Song, W.; Li, X.; Zeng, X. Microstructure and mechanical properties of AlSi10Mg alloy built by laser powder bed fusion/direct energy deposition hybrid laser additive manufacturing. Addit. Manuf. 2022, 59, 103160. [Google Scholar] [CrossRef]
  209. Rauch, M.; Hascoet, J.Y. Opening New Opportunities For Aeronautic, Naval And Train Large Components Realization With Hybrid And Twin Manufacturing. J. Mach. Eng. 2022, 22, 5–20. [Google Scholar] [CrossRef]
  210. Loyda, A.; Arizmendi, M.; Ruiz de Galarreta, S.; Rodriguez-Florez, N.; Jimenez, A. Meeting high precision requirements of additively manufactured components through hybrid manufacturing. CIRP J. Manuf. Sci. Technol. 2023, 40, 199–212. [Google Scholar] [CrossRef]
  211. Rabalo, M.A.; García-Domínguez, A.; Rubio, E.M.; Marín, M.M.; Agustina, B. Last Trends in the Application of the Hybrid Additive and Subtractive Manufacturing in the Aeronautic Industry. Adv. Sci. Technol. 2023, 132, 353–362. [Google Scholar] [CrossRef]
  212. Hafenecker, J.; Bartels, D.; Kuball, C.M.; Kreß, M.; Rothfelder, R.; Schmidt, M.; Merklein, M. Hybrid process chains combining metal additive manufacturing and forming—A review. CIRP J. Manuf. Sci. Technol. 2023, 46, 98–115. [Google Scholar] [CrossRef]
  213. Bidare, P.; Jiménez, A.; Abdelkhalek, S.M.; Khaled, I.; Hassan, A.; Abdallah, R.; Essa, K. Hybrid additive manufacturing of AISI H13 through interlayer machining and laser re-melting. Mater. Today Commun. 2023, 37, 107042. [Google Scholar] [CrossRef]
  214. Sommer, D.; Hornung, S.; Esen, C.; Hellmann, R. Optimization of mechanical properties of additive manufactured IN 718 parts combining LPBF and in-situ high-speed milling. In Proceedings of the Laser 3D Manufacturing XI, San Francisco, CA, USA, 27 January–1 February 2024; Volume 12876, pp. 117–125. [Google Scholar]
  215. Sommer, D.; Esen, C.; Hellmann, R. Static and Dynamic Mechanical Behaviour of Hybrid-PBF-LB/M-Built and Hot Isostatic Pressed Lattice Structures. Materials 2023, 16, 3556. [Google Scholar] [CrossRef]
  216. Letenneur, M.; Kreitcberg, A.; Brailovski, V. Optimization of Laser Powder Bed Fusion Processing Using a Combination of Melt Pool Modeling and Design of Experiment Approaches: Density Control. J. Manuf. Mater. Process. 2019, 3, 21. [Google Scholar] [CrossRef]
  217. Tabassum, R.; Uz Zaman, U.K.; Ahmed Baqai, A.; Rehman Mazhar, A.; Ullah Butt, S. Hybrid Additive Manufacturing: A Review from a Process Planning Perspective. J. Adv. Manuf. Syst. 2023, 23, 227–262. [Google Scholar] [CrossRef]
  218. Sajadi, P.; Rahmani Dehaghani, M.; Tang, Y.; Wang, G.G. Physics-Informed Online Learning for Temperature Prediction in Metal AM. Materials 2024, 17, 3306. [Google Scholar] [CrossRef]
  219. Sharma, R.; Guo, Y.B.; Raissi, M.; Guo, W.G. Physics-Informed Machine Learning of Argon Gas-Driven Melt Pool Dynamics. J. Manuf. Sci. Eng. 2024, 146, 081008. [Google Scholar] [CrossRef]
  220. Ghungrad, S.; Faegh, M.; Gould, B.; Wolff, S.J.; Haghighi, A. Architecture-Driven Physics-Informed Deep Learning for Temperature Prediction in Laser Powder Bed Fusion Additive Manufacturing with Limited Data. J. Manuf. Sci. Eng. 2023, 145, 081007. [Google Scholar] [CrossRef]
  221. Jiang, F.; Xia, M.; Hu, Y. Physics-Informed Machine Learning for Accurate Prediction of Temperature and Melt Pool Dimension in Metal Additive Manufacturing. 3D Print. Addit. Manuf. 2024, 11, e1679–e1689. [Google Scholar] [CrossRef] [PubMed]
  222. Zhu, Q.; Liu, Z.; Yan, J. Machine learning for metal additive manufacturing: Predicting temperature and melt pool fluid dynamics using physics-informed neural networks. Comput. Mech. 2021, 67, 619–635. [Google Scholar] [CrossRef]
  223. Nath, P.; Mahadevan, S. Probabilistic predictive control of porosity in laser powder bed fusion. J. Intell. Manuf. 2023, 34, 1085–1103. [Google Scholar] [CrossRef]
  224. Liu, N.; Li, X.; Rajanna, M.R.; Reutzel, E.W.; Sawyer, B.; Rao, P.; Lua, J.; Phan, N.; Yu, Y. Deep Neural Operator Enabled Digital Twin Modeling for Additive Manufacturing. Adv. Comput. Sci. Eng. 2024, 2, 174–201. [Google Scholar] [CrossRef]
  225. Mahmood, M.A.; Khraisheh, M.; Popescu, A.C.; Liou, F. Processing windows for Al-357 by LPBF process: A novel framework integrating FEM simulation and machine learning with empirical testing. Rapid Prototyp. J. 2024, 30, 1846–1858. [Google Scholar] [CrossRef]
  226. Gaikwad, A.; Giera, B.; Guss, G.M.; Forien, J.B.; Matthews, M.J.; Rao, P. Heterogeneous sensing and scientific machine learning for quality assurance in laser powder bed fusion—A single-track study. Addit. Manuf. 2020, 36, 101659. [Google Scholar] [CrossRef]
  227. Liu, S.; Li, R.; Zhou, J.; Dai, C.; Yu, J.; Zhang, Q. Physics-Based Data Augmentation Enables Accurate Machine Learning Prediction of Melt Pool Geometry. Appl. Sci. 2025, 15, 8587. [Google Scholar] [CrossRef]
  228. Guo, L.; Liu, H.; Wang, H.; Cristino, V.A.M.; Kwok, C.T.; Wei, Q.; Tang, Z.; Wu, Y.; Wang, H. Deepening the scientific understanding of different phenomenology in laser powder bed fusion by an integrated framework. Int. J. Heat Mass Transf. 2023, 216, 124596. [Google Scholar] [CrossRef]
  229. Lim, R.E.; Mukherjee, T.; Chuang, C.; Phan, T.Q.; Debroy, T.; Pagan, D.C.; Keckes, J. Combining synchrotron X-ray diffraction, mechanistic modeling and machine learning for in situ subsurface temperature quantification during laser melting. J. Appl. Crystallogr. 2023, 56, 1131–1143. [Google Scholar] [CrossRef] [PubMed]
  230. Kim, J.; Yang, Z.; Lu, Y. A physics-guided deep generative model for predicting melt pool behavior in laser powder bed fusion additive manufacturing. J. Intell. Manuf. 2024, 36, 5715–5735. [Google Scholar] [CrossRef]
  231. Yang, W.; Qiu, Y.; Liu, W.; Qiu, X.; Bai, Q. Defect prediction in laser powder bed fusion with the combination of simulated melt pool images and thermal images. J. Manuf. Process. 2023, 106, 214–222. [Google Scholar] [CrossRef]
  232. Twumasi, A.; Roy, P.C.; Choi, J.Y.; Gan, Z. Brief Paper: Physics-Guided Scan Paths Optimization for Controlled Microstructure in Laser Powder Bed Fusion. In Proceedings of the ASME 2024 19th International Manufacturing Science and Engineering Conference, MSEC 2024, Knoxville, TN, USA, 17–21 June 2024; Volume 1. [Google Scholar]
  233. Riensche, A.R.; Bevans, B.D.; King, G.; Krishnan, A.; Cole, K.D.; Rao, P. Predicting meltpool depth and primary dendritic arm spacing in laser powder bed fusion additive manufacturing using physics-based machine learning. Mater. Des. 2024, 237, 112540. [Google Scholar] [CrossRef]
  234. Penny, R.W.; John Hart, A. Radiometric Temperature Measurement for Metal Additive Manufacturing via Temperature Emissivity Separation. Addit. Manuf. 2025, 110, 104904. [Google Scholar] [CrossRef]
  235. Surana, A.; Lynch, M.E.; Nassar, A.R.; Ojard, G.C.; Fisher, B.A.; Corbin, D.; Overdorff, R. Flaw Detection in Multi-Laser Powder Bed Fusion Using in Situ Coaxial Multi-Spectral Sensing and Deep Learning. J. Manuf. Sci. Eng. 2023, 145, 051005. [Google Scholar] [CrossRef]
  236. Croset, G.; Martin, G.; Josserond, C.; Lhuissier, P.; Blandin, J.J.; Dendievel, R. In-situ layerwise monitoring of electron beam powder bed fusion using near-infrared imaging. Addit. Manuf. 2021, 38, 101767. [Google Scholar] [CrossRef]
  237. Oster, S.; Fritsch, T.; Ulbricht, A.; Mohr, G.; Bruno, G.; Maierhofer, C.; Altenburg, S.J. On the Registration of Thermographic In Situ Monitoring Data and Computed Tomography Reference Data in the Scope of Defect Prediction in Laser Powder Bed Fusion. Metals 2022, 12, 947. [Google Scholar] [CrossRef]
  238. Terrazas-Nájera, C.A.; Fernández, A.; Felice, R.; Wicker, R. On the thermal emissive behavior of four common alloys processed via powder bed fusion additive manufacturing. Addit. Manuf. 2024, 82, 104023. [Google Scholar] [CrossRef]
  239. Diehl, B.; Castro, A.; Jaquemetton, L.; Beckett, D. Thermal calibration of ratiometric, on-axis melt pool monitoring photodetector system using tungsten strip lamp. Mater. Eval. 2022, 80, 64–73. [Google Scholar] [CrossRef]
  240. Lane, B.; Jacquemetton, L.; Piltch, M.; Beckett, D. Thermal Calibration of Commercial Melt Pool Monitoring Sensors on a Laser Powder Bed Fusion System; US Department of Commerce, National Institute of Standards and Technology: Gaithersburg, MD, USA, 2020. [Google Scholar]
  241. Mohr, G.; Altenburg, S.J.; Ulbricht, A.; Heinrich, P.; Baum, D.; Maierhofer, C.; Hilgenberg, K. In-Situ Defect Detection in Laser Powder Bed Fusion by Using Thermography and Optical Tomography—Comparison to Computed Tomography. Metals 2020, 10, 103. [Google Scholar] [CrossRef]
  242. Petrich, J.; Snow, Z.; Corbin, D.; Reutzel, E.W. Multi-modal sensor fusion with machine learning for data-driven process monitoring for additive manufacturing. Addit. Manuf. 2021, 48, 102364. [Google Scholar] [CrossRef]
  243. Snow, Z.; Amir, L.S.; Ziabari, K.; Fisher, B.; Paquit, V. Machine Learning Enabled Sensor Fusion for In Situ Defect Detection in Laser Powder Bed Fusion CRADA Final Report; Oak Ridge National Laboratory (ORNL): Oak Ridge, TN, USA; RTX Technology Research Center: East Hartford, CT, USA, 2024. [Google Scholar]
  244. Snow, Z.; Scime, L.; Ziabari, A.; Fisher, B.; Paquit, V. Scalable in situ non-destructive evaluation of additively manufactured components using process monitoring, sensor fusion, and machine learning. Addit. Manuf. 2023, 78, 103817. [Google Scholar] [CrossRef]
  245. Ero, O.; Taherkhani, K.; Hemmati, Y.; Toyserkani, E. An integrated fuzzy logic and machine learning platform for porosity detection using optical tomography imaging during laser powder bed fusion. Int. J. Extrem. Manuf. 2024, 6, 065601. [Google Scholar] [CrossRef]
  246. Xie, G.; Walker, J.; Middendorf, J.; Spears, T.; Maass, D.; Dunne, B. Deep-learning-based multisensor fusion for in situ defect prediction in additive manufacturing. In Proceedings of the Applications of Machine Learning 2025, Kuala Lumpur, Malaysia, 14–16 February 2025; Volume 13606, pp. 105–119. [Google Scholar]
  247. Bonato, N.; Zanini, F.; Carmignato, S. Deformations modelling of metal additively manufactured parts and improved comparison of in-process monitoring and post-process X-ray computed tomography. Addit. Manuf. 2023, 75, 103736. [Google Scholar] [CrossRef]
  248. Fernandez, A.; Felice, R.; Terrazas-Nájera, C.A.; Wicker, R. Implications for accurate surface temperature monitoring in powder bed fusion: Using multi-wavelength pyrometry to characterize spectral emissivity during processing. Addit. Manuf. 2021, 46, 102138. [Google Scholar] [CrossRef]
  249. Höfflin, D.; Sauer, C.; Schiffler, A.; Manara, J.; Hartmann, J. Pixelwise high-temperature calibration for in-situ temperature measuring in powder bed fusion of metal with laser beam. Heliyon 2024, 10, e28989. [Google Scholar] [CrossRef] [PubMed]
  250. Wang, R.; Standfield, B.; Dou, C.; Law, A.C.; Kong, Z.J. Real-time process monitoring and closed-loop control on laser power via a customized laser powder bed fusion platform. Addit. Manuf. 2023, 66, 103449. [Google Scholar] [CrossRef]
  251. Tao, F.; Zhang, M.; Nee, A.Y.C. Digital Twin Driven Smart Manufacturing; Elsevier: Amsterdam, The Netherlands, 2019; ISBN 9780128176306. [Google Scholar]
  252. Zhu, Z.; Li, W.; Nguyen, Q.B.; An, X.; Lu, W.; Li, Z.; Ng, F.L.; Ling Nai, S.M.; Wei, J. Enhanced strength–ductility synergy and transformation-induced plasticity of the selective laser melting fabricated 304L stainless steel. Addit. Manuf. 2020, 35, 101300. [Google Scholar] [CrossRef]
  253. Colosimo, B.M.; Grasso, M. In-situ monitoring in L-PBF: Opportunities and challenges. Procedia CIRP 2020, 94, 388–391. [Google Scholar] [CrossRef]
  254. McCann, R.; Obeidi, M.A.; Hughes, C.; McCarthy, É.; Egan, D.S.; Vijayaraghavan, R.K.; Joshi, A.M.; Garzon, V.A.; Dowling, D.P.; McNally, P.J.; et al. In-situ sensing, process monitoring and machine control in Laser Powder Bed Fusion: A review. Addit. Manuf. 2021, 45, 102058. [Google Scholar] [CrossRef]
  255. Lu, Y.; Liu, C.; Wang, K.I.K.; Huang, H.; Xu, X. Digital Twin-driven smart manufacturing: Connotation, reference model, applications and research issues. Robot. Comput. Integr. Manuf. 2020, 61, 101837. [Google Scholar] [CrossRef]
  256. Nath, P.; Mahadevan, S. Probabilistic Digital Twin for Additive Manufacturing Process Design and Control. J. Mech. Des. 2022, 144, 091704. [Google Scholar] [CrossRef]
  257. Tapia, G.; Elwany, A. A Review on Process Monitoring and Control in Metal-Based Additive Manufacturing. J. Manuf. Sci. Eng. Trans. ASME 2014, 136, 060801. [Google Scholar] [CrossRef]
  258. Everton, S.K.; Hirsch, M.; Stavroulakis, P.I.; Leach, R.K.; Clare, A.T. Review of in-situ process monitoring and in-situ metrology for metal additive manufacturing. Mater. Des. 2016, 95, 431–445. [Google Scholar] [CrossRef]
  259. Asad, U.; Khan, M.; Khalid, A.; Lughmani, W.A. Human-Centric Digital Twins in Industry: A Comprehensive Review of Enabling Technologies and Implementation Strategies. Sensors 2023, 23, 3938. [Google Scholar] [CrossRef]
  260. Zhu, Z.; Liu, C.; Xu, X. Visualisation of the Digital Twin data in manufacturing by using Augmented Reality. Procedia CIRP 2019, 81, 898–903. [Google Scholar] [CrossRef]
  261. Kritzinger, W.; Karner, M.; Traar, G.; Henjes, J.; Sihn, W. Digital Twin in manufacturing: A categorical literature review and classification. IFAC-PapersOnLine 2018, 51, 1016–1022. [Google Scholar] [CrossRef]
  262. Tao, F.; Zhang, H.; Liu, A.; Nee, A.Y.C. Digital Twin in Industry: State-of-the-Art. IEEE Trans. Ind. Inform. 2019, 15, 2405–2415. [Google Scholar] [CrossRef]
  263. Gaikwad, A.; Yavari, R.; Montazeri, M.; Cole, K.; Bian, L.; Rao, P. Toward the digital twin of additive manufacturing: Integrating thermal simulations, sensing, and analytics to detect process faults. IISE Trans. 2020, 52, 1204–1217. [Google Scholar] [CrossRef]
  264. Li, J.; Zhou, Q.; Cao, L.; Wang, Y.; Hu, J. A convolutional neural network-based multi-sensor fusion approach for in-situ quality monitoring of selective laser melting. J. Manuf. Syst. 2022, 64, 429–442. [Google Scholar] [CrossRef]
  265. Zhang, H.; Vallabh, C.K.P.; Zhao, X. Registration and fusion of large-scale melt pool temperature and morphology monitoring data demonstrated for surface topography prediction in LPBF. Addit. Manuf. 2022, 58, 103075. [Google Scholar] [CrossRef]
  266. Hernandez, K.J.; Ciardi, T.G.; Yamamoto, R.; Lu, M.; Nihar, A.; Jimenez, J.C.; Tripathi, P.K.; Giera, B.; Forien, J.B.; Lewandowski, J.J.; et al. L-PBF High-Throughput Data Pipeline Approach for Multi-modal Integration. Integr. Mater. Manuf. Innov. 2024, 13, 758–772. [Google Scholar] [CrossRef]
  267. Chen, L.; Bi, G.; Yao, X.; Su, J.; Tan, C.; Feng, W.; Benakis, M.; Chew, Y.; Moon, S.K. In-situ process monitoring and adaptive quality enhancement in laser additive manufacturing: A critical review. J. Manuf. Syst. 2024, 74, 527–574. [Google Scholar] [CrossRef]
  268. Gorgannejad, S.; Martin, A.A.; Nicolino, J.W.; Strantza, M.; Guss, G.M.; Khairallah, S.; Forien, J.B.; Thampy, V.; Liu, S.; Quan, P.; et al. Localized keyhole pore prediction during laser powder bed fusion via multimodal process monitoring and X-ray radiography. Addit. Manuf. 2023, 78, 103810. [Google Scholar] [CrossRef]
  269. Liu, H.; Gobert, C.; Ferguson, K.; Abranovic, B.; Chen, H.; Beuth, J.L.; Rollett, A.D.; Kara, L.B. Inference of highly time-resolved melt pool visual characteristics and spatially-dependent lack-of-fusion defects in laser powder bed fusion using acoustic and thermal emission data. Addit. Manuf. 2023, 83, 104057. [Google Scholar] [CrossRef]
  270. Vandecasteele, M.; Searle, S.; Iuso, D.; Nourazar, M.; Ahar, A.; Hamidi Nasab, M.; Vrancken, B.; Booth, B. A High-speed Multimodal In-situ Monitoring System For Detecting Porosity And Deformation Defects In The Powder Bed Fusion Of Metallic Parts. In Proceedings of the Euro Powder Metallurgy 2025 Congress & Exhibition, Glasgow, UK, 14–17 September 2025. [Google Scholar]
  271. Wu, Q.; Yang, F.; Lv, C.; Liu, C.; Tang, W.; Yang, J. In-Situ Quality Intelligent Classification of Additively Manufactured Parts Using a Multi-Sensor Fusion Based Melt Pool Monitoring System. Addit. Manuf. Front. 2024, 3, 200153. [Google Scholar] [CrossRef]
  272. Scime, L.; Joslin, C.; Collins, D.A.; Sprayberry, M.; Singh, A.; Halsey, W.; Duncan, R.; Snow, Z.; Dehoff, R.; Paquit, V. A Data-Driven Framework for Direct Local Tensile Property Prediction of Laser Powder Bed Fusion Parts. Materials 2023, 16, 7293. [Google Scholar] [CrossRef]
  273. Chen, L.; Yao, X.; Liu, K.; Tan, C.; Moon, S.K. Multisensor fusion-based digital twin in additive manufacturing for in-situ quality monitoring and defect correction. Proc. Des. Soc. 2023, 3, 2755–2764. [Google Scholar] [CrossRef]
  274. Hong, G.; Kim, J.; Yang, Z.; Lu, Y.; Lane, B.; Yeung, H. Multi-Scale Model Predictive Control for Laser Powder Bed Fusion Additive Manufacturing. In Proceedings of the ASME 2024 International Design Engineering Technical Conferences and Computers and Information in Engineering Conference, Washington, DC, USA, 25–28 August 2024; Volume 2B-2024. [Google Scholar] [CrossRef]
  275. Estalaki, S.M.; Lough, C.S.; Landers, R.G.; Kinzel, E.C.; Luo, T. Predicting defects in laser powder bed fusion using in-situ thermal imaging data and machine learning. Addit. Manuf. 2022, 58, 103008. [Google Scholar] [CrossRef]
  276. Dubrov, A.V.; Mirzade, F.K.; Dubrov, V.D. On modeling of heat transfer and molten pool behavior in multi-layer and multi-track laser additive manufacturing process. In Proceedings of the SPIE 11057, Modeling Aspects in Optical Metrology VII; SPIE: Bellingham, WA, USA, 2019; Volume 11057, pp. 374–382. [Google Scholar] [CrossRef]
  277. Santi, A.; Bayat, M.; Hattel, J. Multiphysics modeling of metal based additive manufacturing processes with focus on thermomechanical conditions. J. Therm. Stress. 2023, 46, 445–463. [Google Scholar] [CrossRef]
  278. Bevans, B.D.; Riensche, A.; Carrington, A.; Deshmukh, K.; Darji, M.; Plotnikov, Y.; Sions, J.; Snyder, K.; Hass, D.; Rao, P. In-Process Monitoring of Part Quality in Laser Powder Bed Fusion Additive Manufacturing Process Using Acoustic Emission Sensors. J. Manuf. Sci. Eng. 2025, 147, 061010. [Google Scholar] [CrossRef]
  279. Yuan, B.; Guss, G.M.; Wilson, A.C.; Hau-Riege, S.P.; DePond, P.J.; McMains, S.; Matthews, M.J.; Giera, B. Machine-Learning-Based Monitoring of Laser Powder Bed Fusion. Adv. Mater. Technol. 2018, 3, 1800136. [Google Scholar] [CrossRef]
  280. Jin, L.; Zhai, X.; Wang, K.; Zhang, K.; Wu, D.; Nazir, A.; Jiang, J.; Liao, W.H. Big data, machine learning, and digital twin assisted additive manufacturing: A review. Mater. Des. 2024, 244, 113086. [Google Scholar] [CrossRef]
  281. Oster, S.; Breese, P.P.; Ulbricht, A.; Mohr, G.; Altenburg, S.J. A deep learning framework for defect prediction based on thermographic in-situ monitoring in laser powder bed fusion. J. Intell. Manuf. 2024, 35, 1687–1706. [Google Scholar] [CrossRef]
  282. Liu, R.; Liu, S.; Zhang, X. A physics-informed machine learning model for porosity analysis in laser powder bed fusion additive manufacturing. Int. J. Adv. Manuf. Technol. 2021, 113, 1943–1958. [Google Scholar] [CrossRef]
  283. Park, Y.; Love, A.; Behseresht, S.; Pastrana, O.V.; Sakai, J. Digital Twin Development for Additive Manufacturing. In Proceedings of the ASME 2024 Pressure Vessels & Piping Conference, Bellevue, WA, USA, 28 July–2 August 2024; Volume 1. [Google Scholar] [CrossRef]
  284. Horr, A.M. Real-time Modelling and ML Data Training for Digital Twinning of Additive Manufacturing Processesa? BHM Berg Hüttenmännische Monatshefte 2023, 169, 48–56. [Google Scholar] [CrossRef]
  285. Shehab, E.; Jumassultan, A.; Khoyashov, N.; Juziyeva, S.; Jyeniskhan, N.; Ali, H. Data-Driven Digital Twin Requirements for Additive Layer Manufacturing. MATEC Web Conf. 2024, 401, 02012. [Google Scholar] [CrossRef]
  286. Xu, R.; Huang, S.; Song, Z.; Gao, Y.; Wu, J. A deep mixed-effects modeling approach for real-time monitoring of metal additive manufacturing process. IISE Trans. 2024, 56, 945–959. [Google Scholar] [CrossRef]
  287. Papatheodorou, A.; Papadimitriou, N.; Stathatos, E.; Benardos, P.; Vosniakos, G.C. Recent Advances in Sensor Fusion Monitoring and Control Strategies in Laser Powder Bed Fusion: A Review. Machines 2025, 13, 820. [Google Scholar] [CrossRef]
  288. Li, J.; Hu, J.; Zhou, Q.; Zhang, Y. Transfer learning-based quality monitoring of laser powder bed fusion across materials. Expert Syst. Appl. 2024, 252, 124150. [Google Scholar] [CrossRef]
  289. Kim, J.; Yang, Z.; Lu, Y.; Hong, G.S. Self-Supervised Multilabel Melt Pool Anomaly Classification in Powder Bed Fusion Additive Manufacturing. J. Comput. Inf. Sci. Eng. 2025, 25, 071006. [Google Scholar] [CrossRef]
  290. Larsen, S.; Hooper, P.A. Deep semi-supervised learning of dynamics for anomaly detection in laser powder bed fusion. J. Intell. Manuf. 2022, 33, 457–471. [Google Scholar] [CrossRef]
  291. Dou, C.; Chung, J.; Gnanasambandam, R.; Wu, Y.; Li, J.; Kong, Z. Reinforced Scan: A Reinforcement Learning Enabled Optimal Laser Scan Path Planning in Powder Bed Fusion Additive Manufacturing. Res. Sq. 2025. [Google Scholar] [CrossRef]
  292. Xi, Z. Model Predictive Control of Melt Pool Size for the Laser Powder Bed Fusion Process Under Process Uncertainty. ASCE-ASME J. Risk Uncertain. Eng. Syst. Part B Mech. Eng. 2022, 8, 011103. [Google Scholar] [CrossRef]
  293. Tariq, U.; Joy, R.; Wu, S.H.; Mahmood, M.A.; Malik, A.W.; Liou, F. A state-of-the-art digital factory integrating digital twin for laser additive and subtractive manufacturing processes. Rapid Prototyp. J. 2023, 29, 2061–2097. [Google Scholar] [CrossRef]
  294. Tekoğlu, E.; O’Brien, A.D.; Liu, J.; Wang, B.; Kavak, S.; Zhang, Y.; Kim, S.Y.; Wang, S.; Ağaoğulları, D.; Chen, W.; et al. Strengthening additively manufactured Inconel 718 through in-situ formation of nanocarbides and silicides. Addit. Manuf. 2023, 67, 103478. [Google Scholar] [CrossRef]
  295. Zhang, J.; Yin, C.; Xu, Y.; Sing, S.L. Machine learning applications for quality improvement in laser powder bed fusion: A state-of-the-art review. Int. J. AI Mater. Des. 2024, 1, 26–43. [Google Scholar] [CrossRef]
  296. Khan, I.A.; Birkhofer, H.; Kunz, D.; Lukas, D.; Ploshikhin, V.; Wu, Q.; Khan, I.A.; Birkhofer, H.; Kunz, D.; Lukas, D.; et al. A Random Forest Classifier for Anomaly Detection in Laser-Powder Bed Fusion Using Optical Monitoring. Materials 2023, 16, 6470. [Google Scholar] [CrossRef]
  297. Vaghefi, E.; Hosseini, S.; Prorok, B.; Mirkoohi, E. Geometrically-informed predictive modeling of melt pool depth in laser powder bed fusion using deep MLP-CNN and metadata integration. J. Manuf. Process. 2024, 119, 952–963. [Google Scholar] [CrossRef]
  298. Kunkel, N.; Thölken, D.; Behler, K. Deep learning-based automated defect classification for powder bed fusion—Laser beam. Eur. J. Mater. 2024, 4, 2427401. [Google Scholar] [CrossRef]
  299. Bendaouia, A.; El Faqar, A.C.; Ramoni, M.; Abdelwahed, E.H.; Li, J. Deep Learning and Big Data Framework for Real-Time Porosity Monitoring in Laser Powder Bed Fusion. In Proceedings of the IEEE International Conference on Automation Science and Engineering, Auckland, New Zealand, 26–30 August 2025; pp. 1113–1119. [Google Scholar]
  300. Regal, E.; Gawade, V.; Guo, W. Physics-Informed Loss Functions With Explainable AI to Predict Emission in Powder Bed Fusion. In Proceedings of the ASME 2024 19th International Manufacturing Science and Engineering Conference, Knoxville, TN, USA, 17–21 June 2024; Volume 2. [Google Scholar] [CrossRef]
  301. Raut, R.; Ball, A.K.; Basak, A. Temperature Distribution Prediction in Laser Powder Bed Fusion using Transferable and Scalable Graph Neural Networks. arXiv 2024, arXiv:2407.13838. [Google Scholar] [CrossRef]
  302. Ghungrad, S.; Gould, B.; Wolff, S.; Haghighi, A. Physics-Informed Artificial Intelligence for Temperature Prediction in Metal Additive Manufacturing: A Comparative Study. In Proceedings of the ASME 2022 17th International Manufacturing Science and Engineering Conference, MSEC 2022, West Lafayette, IN, USA, 27 June–1 July 2022; Volume 1. [Google Scholar] [CrossRef]
  303. Rahman, M.S.; Sattar, N.S.; Ahmed, R.U.; Ciaccio, J.; Chakravarty, U.K. A Machine Learning Framework for Melt-Pool Geometry Prediction and Process Parameter Optimization in the Laser Powder-Bed Fusion Process. J. Eng. Mater. Technol. 2024, 146, 041006. [Google Scholar] [CrossRef]
  304. Pandita, P.; Ghosh, S.; Gupta, V.K.; Meshkov, A.; Wang, L. Application of Deep Transfer Learning and Uncertainty Quantification for Process Identification in Powder Bed Fusion. ASCE-ASME J. Risk Uncertain. Eng. Syst. Part B Mech. Eng. 2022, 8, 011106. [Google Scholar] [CrossRef]
  305. Mohd Rosli, M.; Rachini, A.; Karim Mahmood, M.; Alam, S. Data-Driven Melt Pool Monitoring and Defect Prediction in LPBF of Ti-6Al-4V Alloy. Babylon. J. Mech. Eng. 2025, 2025, 51–67. [Google Scholar] [CrossRef]
  306. Gor, M.; Dobriyal, A.; Wankhede, V.; Sahlot, P.; Grzelak, K.; Kluczyński, J.; Łuszczek, J. Density Prediction in Powder Bed Fusion Additive Manufacturing: Machine Learning-Based Techniques. Appl. Sci. 2022, 12, 7271. [Google Scholar] [CrossRef]
  307. Bai, F.; Attarzadeh, F.; Xu, J. An Alternative Machine Learning-Based Approach to Establish Stronger Correlations Between Laser Powder Bed Fusion (LPBF) Processing Parameters and Part Properties. In Proceedings of the ASME International Mechanical Engineering Congress and Exposition, Proceedings (IMECE), Hyderabad, India, 10–13 September 2025; Volume 2. [Google Scholar] [CrossRef]
  308. Zou, Z.; Zhang, K.; Liu, T.; Li, J.; Zhu, Z.; Wei, H.; Lu, Y.; Liao, W. In situ monitoring with melt pool data based on multi-signal fusion method in laser powder bed fusion. Measurement 2024, 234, 114877. [Google Scholar] [CrossRef]
  309. Pandiyan, V.; Baganis, A.; Axel Richter, R.; Wróbel, R.; Leinenbach, C. Qualify-as-you-go: Sensor fusion of optical and acoustic signatures with contrastive deep learning for multi-material composition monitoring in laser powder bed fusion process. Virtual Phys. Prototyp. 2024, 19, e2356080. [Google Scholar] [CrossRef]
  310. Debreceni, A.; Bodzás, S. Integration of Machine Learning in Additive Manufacturing for Material Extrusion and Powder Bed Fusion, a Brief Literature Review. Key Eng. Mater. 2025, 1018, 61–66. [Google Scholar] [CrossRef]
  311. Hu, C.C.; Yang, H.C.; Lu, Y.; Yang, C.W.; Kao, Y.C.; Cheng, F.T. A Distributed AM Architecture with Digital Twin for L-PBF Cluster. In Proceedings of the IEEE International Conference on Automation Science and Engineering, Bari, Italy, 28 August–1 September 2024; pp. 698–704. [Google Scholar]
  312. Chen, Y.-P.; Karkaria, V.; Tsai, Y.-K.; Rolark, F.; Quispe, D.; Gao, R.X.; Cao, J.; Chen, W. Real-Time Decision-Making for Digital Twin in Additive Manufacturing with Model Predictive Control using Time-Series Deep Neural Networks. J. Manuf. Syst. 2025, 80, 412–424. [Google Scholar] [CrossRef]
  313. Park, B.; Mishra, S. Geometry-agnostic Melt-pool Homogenization of Laser Powder Bed Fusion through Reinforcement Learning. In Proceedings of the IEEE/ASME International Conference on Advanced Intelligent Mechatronics, AIM, Seattle, DC, USA, 27 June–1 July 2023; Volume 2023, pp. 1014–1019. [Google Scholar]
  314. Riensche, A.; Bevans, B.; Sions, J.; Snyder, K.; Plotnikov, Y.; Hass, D.; Rao, P. Toward Rapid Process Qualification of Laser Powder Bed Fusion: Model Predictive Control of Part Thermal History. In Proceedings of the ASME 2024 19th International Manufacturing Science and Engineering Conference, MSEC 2024, Knoxville, TN, USA, 17–21 June 2024; Volume 1. [Google Scholar]
  315. Huang, R.Y.; Lu, J.Q.; Cheng, C.W.; Tsai, M.C.; Lee, A.C. Multi-data-driven model-based control to improve the accuracy of overhang structures in laser powder bed fusion. Opt. Laser Technol. 2024, 171, 110398. [Google Scholar] [CrossRef]
  316. Horr, A.M.; Drexler, H. Real-Time Models for Manufacturing Processes: How to Build Predictive Reduced Models. Processes 2025, 13, 252. [Google Scholar] [CrossRef]
  317. Horr, A.M. Real-Time Modeling for Design and Control of Material Additive Manufacturing Processes. Metals 2024, 14, 1273. [Google Scholar] [CrossRef]
  318. Chen, L.; Yao, X.; Chew, Y.; Weng, F.; Moon, S.K.; Bi, G. Data-Driven Adaptive Control for Laser-Based Additive Manufacturing with Automatic Controller Tuning. Appl. Sci. 2020, 10, 7967. [Google Scholar] [CrossRef]
  319. Akhavan, J.; Lyu, J.; Manoochehri, S. A deep learning solution for real-time quality assessment and control in additive manufacturing using point cloud data. J. Intell. Manuf. 2024, 35, 1389–1406. [Google Scholar] [CrossRef]
  320. Yeh, H.P.; Bayat, M.; Arzani, A.; Hattel, J.H. Accelerated process parameter selection of polymer-based selective laser sintering via hybrid physics-informed neural network and finite element surrogate modelling. Appl. Math. Model. 2024, 130, 693–712. [Google Scholar] [CrossRef]
  321. Champaney, V.; Chinesta, F.; Cueto, E. Engineering empowered by physics-based and data-driven hybrid models: A methodological overview. Int. J. Mater. Form. 2022, 15, 31. [Google Scholar] [CrossRef]
  322. San, O.; Rasheed, A.; Kvamsdal, T. Hybrid analysis and modeling, eclecticism, and multifidelity computing toward digital twin revolution. GAMM-Mitteilungen 2021, 44, e202100007. [Google Scholar] [CrossRef]
  323. Gunasegaram, D.R.; Murphy, A.B.; Barnard, A.; DebRoy, T.; Matthews, M.J.; Ladani, L.; Gu, D. Towards developing multiscale-multiphysics models and their surrogates for digital twins of metal additive manufacturing. Addit. Manuf. 2021, 46, 102089. [Google Scholar] [CrossRef]
  324. Mondal, S.; Gwynn, D.; Ray, A.; Basak, A. Investigation of Melt Pool Geometry Control in Additive Manufacturing Using Hybrid Modeling. Metals 2020, 10, 683. [Google Scholar] [CrossRef]
  325. Papacharalampopoulos, A.; Christopoulos, D.; Stavropoulos, P. Towards a surrogate spatiotemporal model of additive manufacturing for digital twin-based process control. Procedia CIRP 2024, 121, 73–78. [Google Scholar] [CrossRef]
  326. Marks, M.; Aswani, K.; Weaver, G.; Dababneh, F.; Taheri, H. Applying Digital Twin Methods for Process-Structure-Property Correlation Assessment in Metal Additive Manufacturing with Limited Experimental Data. Res. Nondestruct. Eval. 2025, 36, 74–90. [Google Scholar] [CrossRef]
  327. Lestandi, L.; Wong, J.C.; Dong, G.Y.; Kuehsamy, S.J.; Mikula, J.; Vastola, G.; Kizhakkinan, U.; Ford, C.S.; Rosen, D.W.; Dao, M.H.; et al. Data-driven surrogate modelling of residual stresses in Laser Powder-Bed Fusion. Int. J. Comput. Integr. Manuf. 2024, 37, 685–707. [Google Scholar] [CrossRef]
  328. Afazov, S.; Roberts, A.; Wright, L.; Jadhav, P.; Holloway, A.; Basoalto, H.; Milne, K.; Brierley, N. Metal powder bed fusion process chains: An overview of modelling techniques. Prog. Addit. Manuf. 2021, 72, 289–314. [Google Scholar] [CrossRef]
  329. Kasilingam, S.; Yang, R.; Singh, S.K.; Farahani, M.A.; Rai, R.; Wuest, T. Physics-based and data-driven hybrid modeling in manufacturing: A review. Prod. Manuf. Res. 2024, 12, 2305358. [Google Scholar] [CrossRef]
  330. Catti, P.; Nikolakis, N.; Sipsas, K.; Picco, N.; Alexopoulos, K. A hybrid digital twin approach for proactive quality control in manufacturing. Procedia Comput. Sci. 2024, 232, 3083–3091. [Google Scholar] [CrossRef]
  331. Xie, J.; Yang, Z.; Hu, C.-C.; Yang, H.-C.; Lu, Y.; Zhao, Y.F. Investigation on domain adaptation of additive manufacturing monitoring systems to enhance digital twin reusability. arXiv 2024, arXiv:2409.12785. [Google Scholar]
  332. Megahed, M. Integrated Modeling Tools: Overview. Encycl. Mater. Met. Alloy. 2022, 3, 95–106. [Google Scholar] [CrossRef]
  333. Kumar, P.; Fernández, E. Topology optimization for additive manufacturing with length scale, overhang, and building orientation constraints. arXiv 2022, arXiv:2204.07333. [Google Scholar] [CrossRef]
  334. van de Ven, E.; Maas, R.; Ayas, C.; Langelaar, M.; van Keulen, F. Overhang control in topology optimization: A comparison of continuous front propagation-based and discrete layer-by-layer overhang control. Struct. Multidiscip. Optim. 2021, 64, 761–778. [Google Scholar] [CrossRef]
  335. Thore, C.J.; Grundström, H.A.; Torstenfelt, B.; Klarbring, A. Penalty regulation of overhang in topology optimization for additive manufacturing. Struct. Multidiscip. Optim. 2019, 60, 59–67. [Google Scholar] [CrossRef]
  336. Wang, Y.; Xia, J.; Luo, Z.; Yan, H.; Sun, J.; Lü, E. Self-supporting topology optimization method for selective laser melting. Addit. Manuf. 2020, 36, 101506. [Google Scholar] [CrossRef]
  337. Miki, T. Reducing computational effort in topology optimization considering the deformation in additive manufacturing. arXiv 2024, arXiv:2403.02711. [Google Scholar] [CrossRef]
  338. Miki, T. Self-support topology optimization considering distortion for metal additive manufacturing. Comput. Methods Appl. Mech. Eng. 2022, 404, 115821. [Google Scholar] [CrossRef]
  339. Cheng, L.; Liang, X.; Bai, J.; Chen, Q.; Lemon, J.; To, A. On utilizing topology optimization to design support structure to prevent residual stress induced build failure in laser powder bed metal additive manufacturing. Addit. Manuf. 2019, 27, 290–304. [Google Scholar] [CrossRef]
  340. Zhang, Z.D.; Ibhadode, O.; Ali, U.; Dibia, C.F.; Rahnama, P.; Bonakdar, A.; Toyserkani, E. Topology optimization parallel-computing framework based on the inherent strain method for support structure design in laser powder-bed fusion additive manufacturing. Int. J. Mech. Mater. Des. 2020, 16, 897–923. [Google Scholar] [CrossRef]
  341. Haveroth, G.A.; Thore, C.J.; Ausas, R.F.; Jakobsson, S.; Cuminato, J.A.; Correa, M.R. A thermal model for topology optimization in additive manufacturing: Design of support structures and geometry orientation. Comput. Struct. 2024, 301, 107453. [Google Scholar] [CrossRef]
  342. Lee, K.H.; Yun, G.J. Design optimization of thermally conductive support structure for laser powder-bed fusion process with part-scale thermal history. Addit. Manuf. 2022, 51, 102627. [Google Scholar] [CrossRef]
  343. Lawton, B.; Gordon, K.D.; Qie, L.; Xing, J. Achieving local overheating and length scale control in density-based topology optimization. J. Phys. Conf. Ser. 2023, 2581, 012005. [Google Scholar] [CrossRef]
  344. Ranjan, R.; Ayas, C.; Langelaar, M.; van Keulen, F. Controlling local overheating in topology optimization for additive manufacturing. Struct. Multidiscip. Optim. 2022, 65, 162. [Google Scholar] [CrossRef]
  345. Lan, W.; Panetta, J. Efficient Layer-by-Layer Simulation for Topology Optimization. In Proceedings of the Proceedings—SCF 2022—7th Annual ACM Symposium on Computational Fabrication, Seattle, WA, USA, 26–28 October 2022; Volume 22. [Google Scholar]
  346. Zou, J.; Mou, H. Topology Optimization of Self-Supporting Structures for Additive Manufacturing with Adaptive Explicit Continuous Constraint. Comput. Model. Eng. Sci. 2022, 132, 451–469. [Google Scholar] [CrossRef]
  347. Tajima, M.; Yamada, T. Topology optimization with geometric constraints for additive manufacturing based on coupled fictitious physical model. Comput. Methods Appl. Mech. Eng. 2023, 417, 116415. [Google Scholar] [CrossRef]
  348. Pellens, J.; Lombaert, G.; Lazarov, B.; Schevenels, M. Combined length scale and overhang angle control in minimum compliance topology optimization for additive manufacturing. Struct. Multidiscip. Optim. 2019, 59, 2005–2022. [Google Scholar] [CrossRef]
  349. Xu, S.; Liu, J.; Sun, Y.; Li, X.; Ma, Y. Support structure topology optimization considering the residual distortion for laser powder bed fusion metal additive manufacturing. Struct. Multidiscip. Optim. 2024, 67, 182. [Google Scholar] [CrossRef]
  350. Langelaar, M. Combined optimization of part topology, support structure layout and build orientation for additive manufacturing. Struct. Multidiscip. Optim. 2018, 57, 1985–2004. [Google Scholar] [CrossRef]
  351. Wang, C.; Qian, X. Simultaneous optimization of build orientation and topology for additive manufacturing. Addit. Manuf. 2020, 34, 101246. [Google Scholar] [CrossRef]
  352. Amir, E.; Amir, O. Concurrent high-resolution topology optimization of structures and their supports for additive manufacturing. Struct. Multidiscip. Optim. 2021, 63, 2589–2612. [Google Scholar] [CrossRef]
  353. Liu, Y.; Huang, X.; Xiong, Y.; Zhou, M.; Lin, Z. Concurrent topology optimization of parts and their supports for additive manufacturing with thermal deformation constraint. Comput. Struct. 2023, 289, 107146. [Google Scholar] [CrossRef]
  354. Wu, C.; Luo, J.; Zhong, J.; Xu, Y.; Wan, B.; Huang, W.; Fang, J.; Steven, G.P.; Sun, G.; Li, Q. Topology optimisation for design and additive manufacturing of functionally graded lattice structures using derivative-aware machine learning algorithms. Addit. Manuf. 2023, 78, 103833. [Google Scholar] [CrossRef]
  355. Zhang, B.; Venugopal, V.; Tripathi, M.; Chandran, N.; Sanaei, N.; Ware, L.; Khare, P.; Anand, S. Design for Additive Manufacturing Constrained Topology Optimization of Liquid-to-Air Heat Exchangers With ML-Based Turbulent Flow Considerations. In Proceedings of the ASME 2024 19th International Manufacturing Science and Engineering Conference, MSEC 2024; American Society of Mechanical Engineers Digital Collection: Knoxville, TN, USA, 2024; Volume 1. [Google Scholar] [CrossRef]
  356. Ibhadode, O.; Zhang, Z.; Sixt, J.; Nsiempba, K.M.; Orakwe, J.; Martinez-Marchese, A.; Ero, O.; Shahabad, S.I.; Bonakdar, A.; Toyserkani, E. Topology optimization for metal additive manufacturing: Current trends, challenges, and future outlook. Virtual Phys. Prototyp. 2023, 18, e2181192. [Google Scholar] [CrossRef]
  357. Liu, H.; Wang, C.; Zhang, Y.; Liang, Y. Multi-material structural discrete variable topology optimization with minimum length scale control under mass constraint. Comput. Methods Appl. Mech. Eng. 2024, 420, 116701. [Google Scholar] [CrossRef]
  358. Czink, S.; Holoch, J.; Renz, R.; Schulze, V.; Albers, A.; Dietrich, S. Process-Specific Topology Optimization Method Based on Laser-Based Additive Manufacturing of AlSi10Mg Components: Material Characterization and Evaluation. Processes 2023, 11, 648. [Google Scholar] [CrossRef]
  359. Holoch, J.; Lenhardt, S.; Revfi, S.; Albers, A. Design of Selective Laser Melting (SLM) Structures: Consideration of Different Material Properties in Multiple Surface Layers Resulting from the Manufacturing in a Topology Optimization. Algorithms 2022, 15, 99. [Google Scholar] [CrossRef]
  360. Cao, S.; Wang, H.; Lu, X.; Tong, J.; Sheng, Z. Topology Optimization Considering Porosity Defects in Metal Additive Manufacturing. Appl. Sci. 2021, 11, 5578. [Google Scholar] [CrossRef]
  361. Boos, E.; Ihlenfeldt, S.; Milaev, N.; Thielsch, J.; Drossel, W.-G.; Bruns, M.; Elsner, B.A.M. Topology Optimized Unit Cells for Laser Powder Bed Fusion. BHM Berg-Hüttenmännische Monatshefte 2022, 167, 291–299. [Google Scholar] [CrossRef]
  362. Zhu, J.; Su, Z.; Wang, Q.; Hao, R.; Lan, Z.; Chan, F.S.; Li, J.; Wong, S.W. Process parameter effects estimation and surface quality prediction for selective laser melting empowered by Bayes optimized soft attention mechanism-enhanced transfer learning. Comput. Ind. 2024, 156, 104066. [Google Scholar] [CrossRef]
  363. Le-Hong, T.; Lin, P.C.; Chen, J.Z.; Pham, T.D.Q.; Van Tran, X. Data-driven models for predictions of geometric characteristics of bead fabricated by selective laser melting. J. Intell. Manuf. 2023, 34, 1241–1257. [Google Scholar] [CrossRef]
  364. Hassanin, H.; Zweiri, Y.; Finet, L.; Essa, K.; Qiu, C.; Attallah, M. Laser Powder Bed Fusion of Ti-6Al-2Sn-4Zr-6Mo Alloy and Properties Prediction Using Deep Learning Approaches. Materials 2021, 14, 2056. [Google Scholar] [CrossRef] [PubMed]
  365. Wang, J.; Jeong, S.G.; Kim, E.S.; Kim, H.S.; Lee, B.J. Material-agnostic machine learning approach enables high relative density in powder bed fusion products. Nat. Commun. 2023, 14, 6557. [Google Scholar] [CrossRef]
  366. Tiwari, J.; Cozzolino, E.; Devadula, S.; Astarita, A.; Krishnaswamy, H. Determination of process parameters for selective laser melting of inconel 718 alloy through evolutionary multi-objective optimization. Mater. Manuf. Process. 2024, 39, 1019–1028. [Google Scholar] [CrossRef]
  367. Rahman, M.S.; Ciaccio, J.; Chakravarty, U.K. A Machine Learning Approach for Predicting Melt-Pool Dynamics of Ti-6Al-4V Alloy in the Laser Powder-Bed Fusion Process. In Proceedings of the ASME International Mechanical Engineering Congress and Exposition, Proceedings (IMECE), Columbus, OH, USA, 30 October–3 November 2022; Volume 4. [Google Scholar]
  368. Maitra, V.; Shi, J.; Lu, C. Robust prediction and validation of as-built density of Ti-6Al-4V parts manufactured via selective laser melting using a machine learning approach. J. Manuf. Process. 2022, 78, 183–201. [Google Scholar] [CrossRef]
  369. Özel, T.; Altay, A.; Kaftanoglu, B.; Leach, R.; Senin, N.; Donmez, A. Focus variation measurement and prediction of surface texture parameters using machine learning in laser powder bed fusion. J. Manuf. Sci. Eng. Trans. ASME 2020, 142, 011008. [Google Scholar] [CrossRef]
  370. Funcke, F.; Forster, T.; Mayr, P. Advantages of Hybrid Neural Network Architectures to Enhance Prediction of Tensile Properties in Laser Powder Bed Fusion. Key Eng. Mater. 2023, 964, 65–71. [Google Scholar] [CrossRef]
  371. Wei, B.; Liu, J.; Li, J.; Zhao, Z.; Liu, Y.; Yang, G.; Liu, L.; Chang, H. Research on an Online Monitoring Device for the Powder Laying Process of Laser Powder Bed Fusion. Micromachines 2024, 15, 97. [Google Scholar] [CrossRef] [PubMed]
  372. Marrey, M.; Malekipour, E.; El-Mounayri, H.; Faierson, E.J. A Framework for Optimizing Process Parameters in Powder Bed Fusion (PBF) Process Using Artificial Neural Network (ANN). Procedia Manuf. 2019, 34, 505–515. [Google Scholar] [CrossRef]
  373. Wilkinson, T.; Casata, M.; Barba, D. An alloy-agnostic machine learning framework for process mapping in laser powder bed fusion. Rapid Prototyp. J. 2024, 30, 303–324. [Google Scholar] [CrossRef]
  374. Marrey, M.; Malekipour, E.; El-Mounayri, H.; Faierson, E.J.; Agarwal, M. A Novel Framework of Developing a Predictive Model for Powder Bed Fusion Process. 3D Print. Addit. Manuf. 2024, 11, 179–196. [Google Scholar] [CrossRef] [PubMed]
  375. Gui, Y.; Aoyagi, K.; Bian, H.; Chiba, A. Detection, classification and prediction of internal defects from surface morphology data of metal parts fabricated by powder bed fusion type additive manufacturing using an electron beam. Addit. Manuf. 2022, 54, 102736. [Google Scholar] [CrossRef]
  376. Li, J.; Hu, J.; Cao, L.; Wang, S.; Liu, H.; Zhou, Q. Multi-objective process parameters optimization of SLM using the ensemble of metamodels. J. Manuf. Process. 2021, 68, 198–209. [Google Scholar] [CrossRef]
  377. Fotovvati, B.; Rauniyar, S.; Arnold, J.A.; Chou, K. Experimental, computational, and data-driven study of the effects of selective laser melting (SLM) process parameters on single-layer surface characteristics. Int. J. Adv. Manuf. Technol. 2022, 123, 119–144. [Google Scholar] [CrossRef]
  378. Dong, H.; Liu, F.; Ye, L.; Ouyang, X.; Wang, Q.; Wang, L.; Huang, L.; Tan, L.; Jin, X.; Liu, Y. Process optimization and mechanical property investigation of Inconel 718 manufactured by selective electron beam melting. Mater. Sci. Addit. Manuf. 2022, 1, 23. [Google Scholar] [CrossRef]
  379. Tran, H.C.; Lo, Y.L. Systematic approach for determining optimal processing parameters to produce parts with high density in selective laser melting process. Int. J. Adv. Manuf. Technol. 2019, 105, 4443–4460. [Google Scholar] [CrossRef]
  380. Zhang, Y.; Yang, S.; Dong, G.; Zhao, Y.F. Predictive manufacturability assessment system for laser powder bed fusion based on a hybrid machine learning model. Addit. Manuf. 2021, 41, 101946. [Google Scholar] [CrossRef]
  381. Taherkhani, K.; Eischer, C.; Toyserkani, E. An unsupervised machine learning algorithm for in-situ defect-detection in laser powder-bed fusion. J. Manuf. Process. 2022, 81, 476–489. [Google Scholar] [CrossRef]
  382. Wang, H.; Zhang, S.; Li, B. Machine learning-assisted acoustic emission monitoring for track formability prediction of laser powder bed fusion. Mater. Today Commun. 2024, 38, 108522. [Google Scholar] [CrossRef]
  383. Pandiyan, V.; Wróbel, R.; Richter, R.A.; Leparoux, M.; Leinenbach, C.; Shevchik, S. Monitoring of Laser Powder Bed Fusion process by bridging dissimilar process maps using deep learning-based domain adaptation on acoustic emissions. Addit. Manuf. 2024, 80, 103974. [Google Scholar] [CrossRef]
  384. Vargas Pérez, M.J.; León, R.B.; Gegen, L.K.H. Optimization of simulation variables of powder bed fusion for 17-Ph steel. ARPN J. Eng. Appl. Sci. 2023, 18, 1135–1142. [Google Scholar] [CrossRef]
  385. Qadeer, A.; Akhtar, S.S.; Abubakar, A.A.; Arif, A.F.M.; Mekid, S.; Al-Athel, K.S. Accelerating Microstructure and Melt Pool Prediction in Laser Powder Bed Fusion Processes Using Data Driven Neural Network. In Proceedings of the ASME International Mechanical Engineering Congress and Exposition, Proceedings (IMECE), Hyderabad, India, 10–13 September 2025; Volume 2. [Google Scholar]
  386. Kaščák, Ľ.; Varga, J.; Bidulská, J.; Bidulský, R.; Manfredi, D. Weight Factor as a Parameter for Optimal Part Orientation in the L-PBF Printing Process Using Numerical Simulation. Materials 2024, 17, 3604. [Google Scholar] [CrossRef]
  387. Peter, N.; Pitts, Z.; Thompson, S.; Saharan, A. Benchmarking build simulation software for laser powder bed fusion of metals. Addit. Manuf. 2020, 36, 101531. [Google Scholar] [CrossRef]
  388. Wirth, F.; Tonn, T.; Schöberl, M.; Hermann, S.; Birkhofer, H.; Ploshikhin, V. Implementation of the Marangoni effect in an open-source software environment and the influence of surface tension modeling in the mushy region in laser powder bed fusion (LPBF). Model. Simul. Mater. Sci. Eng. 2022, 30, 034001. [Google Scholar] [CrossRef]
  389. Zirwes, T.; Sontheimer, M.; Zhang, F.; Abdelsamie, A.; Pérez, F.E.H.; Stein, O.T.; Im, H.G.; Kronenburg, A.; Bockhorn, H. Assessment of Numerical Accuracy and Parallel Performance of OpenFOAM and its Reacting Flow Extension EBIdnsFoam. Flow Turbul. Combust. 2023, 111, 567–602. [Google Scholar] [CrossRef]
  390. Goetz, D.; Panzer, H.; Wolf, D.; Bayerlein, F.; Spachtholz, J.; Zaeh, M.F. AscentAM: A Software Tool for the Thermo-Mechanical Process Simulation of Form Deviations and Residual Stresses in Powder Bed Fusion of Metals Using a Laser Beam. Modelling 2024, 5, 841–860. [Google Scholar] [CrossRef]
  391. Safari, H.; Wessels, H. Physics-Informed Surrogates for Temperature Prediction of Multi-Tracks in Laser Powder Bed Fusion. arXiv 2025, arXiv:2502.01820. [Google Scholar]
  392. Depaoli, F.; Felicioni, S.; Ponzio, F.; Aliberti, A.; MacIi, E.; Bondioli, F.; Padovano, E.; Di Cataldo, S. Physics-Informed Neural Networks: A Step Towards Data-Driven Optimization of Additive Manufacturing. In Proceedings of the IEEE International Conference on Emerging Technologies and Factory Automation, ETFA, Padova, Italy, 10–13 September 2024. [Google Scholar]
  393. Kuehne, M.; Bartsch, K.; Bossen, B.; Emmelmann, C. Predicting melt track geometry and part density in laser powder bed fusion of metals using machine learning. Prog. Addit. Manuf. 2023, 8, 47–54. [Google Scholar] [CrossRef]
  394. Bai, F.; Arikatla, S.S.P.R.; Zhang, N.; Gebre, F.L.; Xu, J. Comparison of Machine Learning Models and Analytical Scaling Law for Predicting Melt-Pool Depth in Laser Powder Bed Fusion (LPBF) Additive Manufacturing. In Proceedings of the ASME International Mechanical Engineering Congress and Exposition, Proceedings (IMECE), Portland, OR, USA, 17–21 November 2024; Volume 3. [Google Scholar]
  395. Lu, L.; Meng, X.; Mao, Z.; Karniadakis, G.E. DeepXDE: A Deep Learning Library for Solving Differential Equations. SIAM Rev. 2021, 63, 208–228. [Google Scholar] [CrossRef]
  396. Zhang, L.; Chen, X.; Zhou, W.; Cheng, T.; Chen, L.; Guo, Z.; Han, B.; Lu, L. Digital Twins for Additive Manufacturing: A State-of-the-Art Review. Appl. Sci. 2020, 10, 8350. [Google Scholar] [CrossRef]
  397. Hernandez, K.J.; Barcelos, E.I.; Jimenez, J.C.; Nihar, A.; Tripathi, P.K.; Giera, B.; French, R.H.; Bruckman, L.S. A data integration framework of additive manufacturing based on FAIR principles. MRS Adv. 2024, 9, 844–851. [Google Scholar] [CrossRef]
  398. Scime, L.; Siddel, D.; Baird, S.; Paquit, V. Layer-wise anomaly detection and classification for powder bed additive manufacturing processes: A machine-agnostic algorithm for real-time pixel-wise semantic segmentation. Addit. Manuf. 2020, 36, 101453. [Google Scholar] [CrossRef]
  399. Westman, S.; Poutiainen, I.; Hirvimäki, M. Process Monitoring in Metal Additive Manufacturing; LUT University: Lappeenranta, Finland, 2024. [Google Scholar]
  400. Scime, L.; Halsey, W.; Haley, J.; Singh, A.; Sprayberry, M.; Ziabari, A.; Paquit, V. Development of Monitoring Techniques for Laser Powder Bed Additive Manufacturing of Metal Structures (Progress Report); Oak Ridge National Laboratory (ORNL): Oak Ridge, TN, USA, 2020. [Google Scholar]
  401. Zhao, X.; Zhang, H.; Alam, M.J. Multi-Modality In-Situ Monitoring Big Data Mining for Enhanced Insight into the Laser Powder Bed Fusion Process, Structure, and Properties. In Proceedings of the The International Conference on Computational & Experimental Engineering and Sciences, Hanoi, Vietnam, 1–4 July 2024; Volume 31, pp. 1–2. [Google Scholar]
  402. Duboeuf, F.; Lemaire, E.; Remouchamps, A.; van Eekelen, T.; Chary, C.; François, M.; Vargalui, A.; Rodrigues, G. Integration of AM process in design cycle of metallic parts: Application to space components. In Proceedings of the 24th International Conference on Material Forming, Liège, Belgium, 14–16 April 2021. [Google Scholar]
  403. Poka, K.; Merz, B.; Epperlein, M.; Hilgenberg, K. Integration of the Whole Digital Chain in a Unique File for PBF-LB/M: Practical Implementation Within a Digital Thread and Its Advantages; Springer Tracts in Additive Manufacturing; Springer: Cham, Switzerland, 2024; Volume Part F3256, pp. 91–114. [Google Scholar] [CrossRef]
  404. Dalpadulo, E.; Pini, F.; Leali, F. Powder bed fusion integrated product and process design for additive manufacturing: A systematic approach driven by simulation. Int. J. Adv. Manuf. Technol. 2024, 130, 5425–5440. [Google Scholar] [CrossRef]
  405. Sedlacek, F.; Kalina, T.; Stepanek, M. Optimization of Components with Topology Optimization for Direct Additive Manufacturing by DLMS. Materials 2023, 16, 5422. [Google Scholar] [CrossRef]
  406. Nakai-Chapman, J.; Fietek, C.; Love, A.; Sakai, J.; Park, Y.H. Metal additive manufacturing simulation using sequentially coupled thermomechanical analysis with Johnson–Cook constitutive material model. Multiscale Multidiscip. Model. Exp. Des. 2024, 7, 801–810. [Google Scholar] [CrossRef]
  407. Dalpadulo, E.; Pini, F.; Leali, F. Concurrent Product and Process Design of an Additively Manufactured Engine Piston. In Design Tools and Methods in Industrial Engineering III. ADM 2023. Lecture Notes in Mechanical Engineering; Springer: Cham, Switzerland, 2024; pp. 93–100. [Google Scholar] [CrossRef]
  408. Savane, V.; Fu, R.; Gagorik, P.; Chin, C. Distortion prediction of ti6al4v parts in selective laser melting: An industrial case study. In Proceedings of the AIAA Scitech 2020 Forum, Orlando, FL, USA, 6–10 January 2020; p. 1475. [Google Scholar] [CrossRef]
  409. Celik, H.K.; Elham, A.; Cinar, R.; Erbil, M.A.; Entwistle, R.; Rennie, A.E.; Akinci, I. Development of a CAD–FEA Integrated Automation Add-In for DfAM-Aware Topology Optimization: A Case Study on an Additively Manufactured Pusher Duct Support Bracket for a Novel UAV Prototype. Appl. Sci. 2025, 15, 12341. [Google Scholar] [CrossRef]
  410. Scheel, P.; Hosseini, E. Pulse approach: A physics-guided machine learning model for thermal analysis in laser-based powder bed fusion of metals. Prog. Addit. Manuf. 2025, 10, 1405–1420. [Google Scholar] [CrossRef]
  411. Ogoke, F.; Farimani, A.B. Thermal control of laser powder bed fusion using deep reinforcement learning. Addit. Manuf. 2021, 46, 102033. [Google Scholar] [CrossRef]
  412. Mohammadkamal, H.; Caiazzo, F. Thermomechanical Simulation of Laser Powder Bed Fusion for Ti-6Al-4V: A Study on Residual Stress and Impact of Laser Process Parameters Using Volumetric Heat Source. In Proceedings of the 2024 IEEE International Workshop on Metrology for AeroSpace, MetroAeroSpace 2024—Proceeding, Lublin, Poland, 3–5 June 2024; pp. 260–265. [Google Scholar]
  413. Huang, S.; Huang, D. Deep Learning-Based Cooperative Modelling of Thermal Flow and Stress Fields in Laser Powder Bed Fusion. Int. J. Heat Technol. 2023, 41, 1217–1225. [Google Scholar] [CrossRef]
  414. Chen, J.; Pierce, J.; Williams, G.; Simpson, T.W.; Meisel, N.; Prabha Narra, S.; McComb, C. Artificial intelligence, deep learning, UUV, hull design, energy consumption, computer aided design, data-driven engineering, geometric reasoning, machine learning for engineering applications. J. Comput. Inf. Sci. Eng. 2024, 24, 011004. [Google Scholar] [CrossRef]
  415. Zhang, Y.; Guillemot, G.; Camus, T.; Senninger, O.; Bellet, M.; Gandin, C.A. Part-Scale Thermomechanical and Grain Structure Modeling for Additive Manufacturing: Status and Perspectives. Metals 2024, 14, 1173. [Google Scholar] [CrossRef]
  416. Papadakis, L. Modeling and simulation of additive manufacturing processes with metallic powders—Potentials and limitations demonstrated on application examples. In Additive Manufacturing; Elsevier: Amsterdam, The Netherlands, 2021; pp. 685–721. ISBN 9780128184110. [Google Scholar]
  417. Huang, W.; Wang, W.; Ning, J.; Garmestani, H.; Liang, S.Y. Analytical Model of Quantitative Texture Prediction Considering Heat Transfer Based on Single-Phase Material in Laser Powder Bed Fusion. J. Manuf. Mater. Process. 2024, 8, 70. [Google Scholar] [CrossRef]
  418. Promoppatum, P.; Rollett, A.D. Influence of material constitutive models on thermomechanical behaviors in the laser powder bed fusion of Ti-6Al-4V. Addit. Manuf. 2021, 37, 101680. [Google Scholar] [CrossRef]
  419. Hasanabadi, M.; Imani Shahabad, S.; Keshavarzkermani, A.; Eybel, R.; Gerlich, A.; Toyserkani, E. A numerical modelling for laser Powder-bed fusion of Ti-alloy with a hybrid heat Source: An investigation on solidification and microstructure formation. Opt. Laser Technol. 2024, 174, 110647. [Google Scholar] [CrossRef]
  420. Pérez-Ruiz, J.D.; Galbusera, F.; Caprio, L.; Previtali, B.; de Lacalle, L.N.L.; Lamikiz, A.; Demir, A.G. Laser beam shaping facilitates tailoring the mechanical properties of IN718 during powder bed fusion. J. Mater. Process. Technol. 2024, 328, 118393. [Google Scholar] [CrossRef]
  421. Qadeer, A.; Akhtar, S.S.; Abubakar, A.A.; Arif, A.F.M.; Mekid, S.; Al-Athel, K.S. Multi-Physics Modeling of Continuous and Pulsed Lasers: Impact on Thermal History and Surface Properties in Powder Bed Fusion Additive Manufacturing. In Proceedings of the ASME International Mechanical Engineering Congress and Exposition, Proceedings (IMECE), Hyderabad, India, 10–13 September 2025; Volume 9. [Google Scholar]
  422. Behrens, C.; Ostermann, N.; Sehrt, J.T.; Ploshikhin, V. Temperature control by simulated adaptive layer times in powder bed fusion processes. Prog. Addit. Manuf. 2024, 9, 705–713. [Google Scholar] [CrossRef]
  423. Papazoglou, E.L.; Karkalos, N.E.; Karmiris-Obratański, P.; Markopoulos, A.P. On the Modeling and Simulation of SLM and SLS for Metal and Polymer Powders: A Review. Arch. Comput. Methods Eng. 2021, 292, 941–973. [Google Scholar] [CrossRef]
  424. Tan, P.; Zhou, M.; Tang, C.; Su, Y.; Qi, H.J.; Zhou, K. Multiphysics modelling of powder bed fusion for polymers. Virtual Phys. Prototyp. 2023, 18, e2257191. [Google Scholar] [CrossRef]
  425. Told, R.; Kardos, K.; Paari-Molnar, E.; Szabo, G.; Ujfalusi, Z.; Sahai, N.; Szabo, P.; Maroti, P. Comparative Analysis of Mechanical and Thermal Characteristics of 3D-Printed Polyamide using Material Extrusion and Powder Bed Fusion Process with Industrial and Desktop Printers. Macromol. Mater. Eng. 2025, 310, 2400293. [Google Scholar] [CrossRef]
Figure 1. PRISMA flow diagram of study selection. Original figure created by the authors.
Figure 1. PRISMA flow diagram of study selection. Original figure created by the authors.
Materials 19 00426 g001
Figure 2. Scheme of the PBF process. Figure conceptually synthesized by the authors from widely available literature and domain knowledge.
Figure 2. Scheme of the PBF process. Figure conceptually synthesized by the authors from widely available literature and domain knowledge.
Materials 19 00426 g002
Figure 3. DEM simulation of powder spreading and packing behavior for Ti-6Al-4V powder: (a) particle velocity field, (b) packing density distribution, (c) influence of particle morphology on flowability and density, (d) effect of recoater speed on surface roughness and packing quality. Figure prepared using data and information reported in the cited publications.
Figure 3. DEM simulation of powder spreading and packing behavior for Ti-6Al-4V powder: (a) particle velocity field, (b) packing density distribution, (c) influence of particle morphology on flowability and density, (d) effect of recoater speed on surface roughness and packing quality. Figure prepared using data and information reported in the cited publications.
Materials 19 00426 g003
Figure 4. Comparison of melt pool predictions for Ti-6Al-4V using three modeling approaches: (a) Rosenthal analytical solution; (b) Goldak FEM model; (c) high-fidelity CFD simulation. Figure prepared using data and information reported in the cited publications.
Figure 4. Comparison of melt pool predictions for Ti-6Al-4V using three modeling approaches: (a) Rosenthal analytical solution; (b) Goldak FEM model; (c) high-fidelity CFD simulation. Figure prepared using data and information reported in the cited publications.
Materials 19 00426 g004
Figure 5. Integration of phase-field and cellular automaton methods for microstructure evolution modeling, validated by EBSD data: (a) phase-field simulation captures dendritic arm spacing under LPBF conditions; (b) CA model predicts grain structure and columnar-to-equiaxed transition; (c) EBSD map confirms ⟨001⟩ texture and transition zone. Figure prepared using data and information reported in the cited publications.
Figure 5. Integration of phase-field and cellular automaton methods for microstructure evolution modeling, validated by EBSD data: (a) phase-field simulation captures dendritic arm spacing under LPBF conditions; (b) CA model predicts grain structure and columnar-to-equiaxed transition; (c) EBSD map confirms ⟨001⟩ texture and transition zone. Figure prepared using data and information reported in the cited publications.
Materials 19 00426 g005
Figure 6. Validation of PBF modeling accuracy and efficiency across the full modeling chain: (a) DEM powder packing accuracy [57,58,72,73], (b) melt pool cost–accuracy trade-offs across analytical, FEM, and CFD models [2,121,122], (c) thermomechanical residual stress validation [145,146,154,155,156], and (d) microstructure prediction fidelity of phase-field and cellular automaton methods [128,131,132,137,138,140,141,142], illustrating the balance between physics fidelity and computational cost. Figure prepared using published data.
Figure 6. Validation of PBF modeling accuracy and efficiency across the full modeling chain: (a) DEM powder packing accuracy [57,58,72,73], (b) melt pool cost–accuracy trade-offs across analytical, FEM, and CFD models [2,121,122], (c) thermomechanical residual stress validation [145,146,154,155,156], and (d) microstructure prediction fidelity of phase-field and cellular automaton methods [128,131,132,137,138,140,141,142], illustrating the balance between physics fidelity and computational cost. Figure prepared using published data.
Materials 19 00426 g006
Figure 7. Timeline of PBF modeling evolution. Figure conceptually synthesized by the authors from the widely available literature and domain knowledge.
Figure 7. Timeline of PBF modeling evolution. Figure conceptually synthesized by the authors from the widely available literature and domain knowledge.
Materials 19 00426 g007
Figure 8. Ecosystem of a PBF digital twin. Figure conceptually synthesized by the authors from widely the available literature and domain knowledge.
Figure 8. Ecosystem of a PBF digital twin. Figure conceptually synthesized by the authors from widely the available literature and domain knowledge.
Materials 19 00426 g008
Figure 9. Defect formation and detection in PBF: (a) process maps showing defect regimes for Ti-6Al-4V [4,51,86,91,92,113,116,274,275,276,277], (b) detection accuracy of thermal, acoustic, optical, and XCT methods across defect types [278,279,280,281,282,283,284,285,286], and (c) multi-sensor fusion gains improving overall accuracy to 95% and meeting industrial reliability thresholds [278,279,280,281,285]. Figure prepared using published data.
Figure 9. Defect formation and detection in PBF: (a) process maps showing defect regimes for Ti-6Al-4V [4,51,86,91,92,113,116,274,275,276,277], (b) detection accuracy of thermal, acoustic, optical, and XCT methods across defect types [278,279,280,281,282,283,284,285,286], and (c) multi-sensor fusion gains improving overall accuracy to 95% and meeting industrial reliability thresholds [278,279,280,281,285]. Figure prepared using published data.
Materials 19 00426 g009
Figure 10. Performance of AI-enhanced digital twins: (a) accuracy of physics-based, PINN, data-driven, and hybrid models [210,212,267,276,277,5,194,213]; (b) accuracy–efficiency trade-offs showing 10–3900× speed-ups with ≥97% accuracy [5,194,210,213,214]; (c) multi-sensor fusion boosting defect detection to 98.5% accuracy and 95% POD, demonstrating real-time, closed-loop PBF capability [232,253,254,258,242,243]. Figure prepared using published data.
Figure 10. Performance of AI-enhanced digital twins: (a) accuracy of physics-based, PINN, data-driven, and hybrid models [210,212,267,276,277,5,194,213]; (b) accuracy–efficiency trade-offs showing 10–3900× speed-ups with ≥97% accuracy [5,194,210,213,214]; (c) multi-sensor fusion boosting defect detection to 98.5% accuracy and 95% POD, demonstrating real-time, closed-loop PBF capability [232,253,254,258,242,243]. Figure prepared using published data.
Materials 19 00426 g010
Figure 11. PBF process optimization. Figure conceptually synthesized by the authors from widely available literature and domain knowledge.
Figure 11. PBF process optimization. Figure conceptually synthesized by the authors from widely available literature and domain knowledge.
Materials 19 00426 g011
Figure 12. Overview of future research directions and priorities. Figure prepared using data and information reported in the cited publications.
Figure 12. Overview of future research directions and priorities. Figure prepared using data and information reported in the cited publications.
Materials 19 00426 g012
Table 1. Corpus composition.
Table 1. Corpus composition.
AttributeDistribution
Total Publications408
Year DistributionPre-2020: 33 (8%)|2020: 31 (8%)|2021: 43 (11%)|2022: 62 (15%)|2023: 79 (19%)|2024: 111 (27%)|2025: 49 (12%)
TopicsPhysics-based modeling (35%), ML/AI (28%), Digital twins (18%), Hybrid frameworks (12%), Monitoring/sensing (7%)
PBF VariantsLPBF: 87%|EB-PBF: 9%|Multi-laser/hybrid: 4%
MaterialsTi alloys: 32%|Ni superalloys: 24%|Stainless steels: 21%|Al alloys: 12%|Others: 11%
MethodsFEM: 38%|CFD: 18%|DEM: 14%|ML/NN: 22%|CA/Phase-field: 8%
ValidationExperimental: 68%|Benchmark: 22%|Sensitivity only: 10%
Table 2. Overview of the major PBF variants * [10,14,15,16,17,18].
Table 2. Overview of the major PBF variants * [10,14,15,16,17,18].
AspectSelective Laser Sintering (SLS)Selective Laser Melting (SLM)Electron Beam Melting (EBM)Multi Jet Fusion (MJF)
Energy SourceCO2 or fiber laserHigh-power fiber laserElectron beamThermal inkjet + infrared heating
Processing EnvironmentNitrogen atmosphereInert gas (Ar/N2)High vacuum (~10−4 mbar)Nitrogen atmosphere
Working PrincipleSelective sintering below the melting pointComplete melting of powder particlesElectron beam melting under vacuumChemical agent application + thermal fusion
Temperature Range150–200 °C (powder bed)Room temperature start700–850 °C (preheated bed)150–200 °C (powder bed)
Primary MaterialsPolymers (PA12, PA11, TPU)Metals (Ti, SS, Al, Co-Cr)Metals (Ti alloys, Co-Cr, Inconel)Polymers (PA12, PA11)
Typical Layer Thickness50–200 μm20–100 μm50–200 μm80–120 μm
Build SpeedModerate (10–20 mm/h)Slow (5–15 mm/h)Fast (20–80 mm/h)Fast (15–25 mm/h)
Part Density85–95%>99%>99%90–98%
Surface FinishRa 6–12 μmRa 5–15 μmRa 15–35 μmRa 3–8 μm
Dimensional Accuracy±0.1–0.3 mm±0.05–0.2 mm±0.2–0.5 mm±0.1–0.2 mm
Support StructuresMinimal (self-supporting)Required for overhangsMinimal (powder support)Minimal (self-supporting)
Post-ProcessingPowder removal, surface finishingSupport removal, heat treatmentPowder removal, machiningPowder removal, surface finishing
* (values represent typical ranges reported in the literature for optimized processing conditions; actual performance depends on specific machine configuration, material properties, part geometry, and processing parameters).
Table 3. Material properties relevant to PBF processing [4,18,19,20,21,23,24,25,26,27,28,30,31,32,44,47,48,49].
Table 3. Material properties relevant to PBF processing [4,18,19,20,21,23,24,25,26,27,28,30,31,32,44,47,48,49].
Alloy SystemThermal Conductivity (W/m·K)AbsorptivitySolidification Range (°C)Notable Microstructural TraitsCommon Defects
Ti-6Al-4V6.7–22 (temperature-dependent)0.3–0.4 at laser wavelengths1604–1660Martensitic α’ phase, columnar β grains, epitaxial growth patternsPorosity, lack of fusion, cracking, oxidation
AlSi10MgLow, anisotropic (5–15 typical)0.1–0.3 (enhanced with nanoparticles)577–660 (Al-Si eutectic)Cellular dendritic structure, columnar-to-equiaxed transitionLack of fusion, hot cracking, porosity
316L Stainless Steel15–25 (varies with compaction)0.35–0.451400–1450Cellular austenitic structure, nanoscale oxide particlesPorosity, lack of fusion, residual stress
Inconel 62510–15 (estimated)0.3–0.41290–1350Dendritic structure with Nb segregation, γ matrixSolidification cracking, porosity, microsegregation
CoCrMo14–170.4–0.51350–1450HCP ε-martensite, FCC γ austenitePorosity, phase transformation, and cracking
Maraging Steel (18Ni-300)17–200.3–0.41413–1460Martensitic laths, reverted austeniteLack of fusion, thermal cracking
AlSi128–12 (anisotropic)0.1–0.25575–660Fine cellular structure, Si precipitationHot cracking, porosity
Hastelloy X9–120.35–0.41260–1355Dendritic γ matrix with carbidesSolidification cracking, porosity
Table 4. Comparison of spreader types and their influence on powder layer quality in PBF.
Table 4. Comparison of spreader types and their influence on powder layer quality in PBF.
Spreader TypePacking DensitySurface RoughnessSpeed CapabilityDamage RiskReferences
Rigid blade0.58–0.624–7 μmHigh (>300 mm/s)Low[59,64]
Flexible blade0.60–0.643–5 μmMedium (150–250)Very low[62]
Roller0.60–0.633–6 μmMedium (100–200)Medium[65,66]
Table 5. Influence of powder and process parameters on packing density and uniformity in PBF.
Table 5. Influence of powder and process parameters on packing density and uniformity in PBF.
ParameterConditionDensityCV (%)References
MorphologySpherical (ψ > 0.95)0.62–0.645–8[73,74]
Irregular (ψ < 0.85)0.50–0.5812–18[72]
SpeedOptimal (100–200 mm/s)0.60–0.636–10[64]
High (>300 mm/s)0.52–0.5815–20[79]
RecoaterFlexible blade0.60–0.645–8[62]
Roller (counter-rot.)0.60–0.636–10[66]
TemperatureAmbient (20 °C)0.60–0.636–10[70]
Elevated (200 °C)0.56–0.6010–15[70]
GravityEarth (1 g)0.60–0.636–10[68]
Lunar (0.17 g)0.52–0.5812–18[68]
CohesionFree-flow (H < 1.2)0.62–0.645–8[82]
Cohesive (H > 1.3)0.54–0.6012–20[71]
Table 6. Laser–powder interaction and energy absorption modeling approaches in powder bed fusion.
Table 6. Laser–powder interaction and energy absorption modeling approaches in powder bed fusion.
Model Type/FrameworkPhysics CapturedResolution/Representation of Powder GeometryCore CapabilitiesValidation StrategyComputational Cost/EfficiencyKey LimitationsRepresentative References
Analytical and Ray-Tracing ModelsMultiple reflections, beam scattering, absorption, refraction, volume absorption within powder bed; laser energy distribution and attenuation; radiative transfer in particulate mediaAccounts for particle size distribution, powder packing density, and layer thickness effects; can model random powder packing; some variants use idealized regular packing structuresHigh accuracy in predicting absorptivity variations due to powder bed structure; captures energy deposition patterns influenced by morphology; supports process parameter optimization for defect mitigationExperimental absorptance data, profilometry, single-track experiments, thermal imaging, comparison with melt pool morphologyMedium: Monte Carlo and photon-particle tracking methods vary in efficiency; suitable for powder-scale simulationsOften assumes idealized powder bed structures, neglecting realistic morphology and surface roughness; limited representation of powder heterogeneity; requires coupling with thermal solvers for complete process modeling[96,97,98]
Discrete Element Method (DEM) Coupled Optical–Thermal ModelsPowder packing dynamics, particle-level morphology, surface roughness effects on absorptivity; laser–powder interaction, including multiple reflections; heat transfer through powder bed; powder degradation and recycled powder effectsExplicitly models realistic powder morphology, particle size distribution, shape effects (spherical, ellipsoids, irregular), packing density, and surface roughness; captures powder bed heterogeneity at the particle scaleRobust representation of absorptivity variations due to packing structure and morphology; validated against experimental powder bed characteristics and melt pool dimensions; strong capability for predicting powder bed quality effects on melt pool dynamicsProfilometry, microscopy, X-ray computed tomography (XCT) for powder bed validation, experimental melt pool metrics, single-track experiments, thermal profilesMedium-to-high: GPU acceleration significantly reduces simulation time, enabling multi-layer and bulk-scale simulations; DEM frameworks computationally intensive for large domainsComputational expense limits scalability to full part simulations; multi-layer simulations remain challenging without GPU acceleration; requires extensive calibration for different materials; powder degradation effects underexplored[61,81,99,100,101]
Continuum Models with Effective Medium Approximations (FEM/FVM)Thermomechanical phenomena, heat conduction, phase change, residual stress, distortion; simplified representation of powder bed as homogeneous medium with effective properties; keyhole dynamics at thresholdHomogenizes powder bed characteristics using effective thermal conductivity and absorptivity; limited granularity on particle-level morphology; treats powder bed as a continuumModerate accuracy; effective medium approximations simplify powder-scale heterogeneity, potentially misrepresenting absorption and heat conduction; suitable for macro-scale thermal predictions; widely used for part-scale process control, distortion mitigation, and residual stress predictionExperimental distortion data, residual stress measurements, melt pool metrics, thermal imaging, microstructure validationLow to medium: highly efficient for part-scale simulations; enables multi-layer and full-part thermal–mechanical analysisOversimplifies powder bed characteristics; lacks resolution for local powder heterogeneities and melt pool instabilities; may overlook defect formation mechanisms dependent on powder morphology; limited representation of particle-scale physics[102,103,104,105]
Particle-Based Methods (SPH, LBM) for Radiation–Heat CouplingDetailed melt pool fluid dynamics, phase change, melting mode transitions, recoil pressure, Marangoni convection, evaporation, heat transfer and solidification, powder–melt interactions, powder wetting influence on morphologyModels powder thermal conductivity and particle-scale interactions; captures fluid flow around powder particles; SPH and LBM handle complex free surface dynamics and powder wettingHigh fidelity in capturing melt pool dynamics and phase transitions; accurate representation of melting behavior and keyhole formation; validated against experimental melt pool geometry; effective for detailed process understanding of melt pool instabilitiesX-ray analyses, high-speed imaging, melt pool geometry validation, experimental track morphology, thin wall morphology, in situ monitoringHigh: computationally intensive, limiting scalability to particle or mesoscale domains; restricted to single-track or small domain simulationsComputationally prohibitive for multi-layer or full-part simulations; limited industrial deployment due to computational demands; often excludes full vapor dynamics; scalability challenges for industrial applications[84,106,107,108]
Hybrid Multiphysics Models (Ray-Tracing + FEM/CFD + Microstructure)Comprehensive multiphysics: optical absorption via ray-tracing combined with thermal–fluid dynamics (phase change, recoil pressure, Marangoni effects), solidification, microstructure evolution (cellular automata, phase-field); bubble generation and migrationIntegrates detailed powder packing (via DEM or explicit morphology) with continuum thermal analysis; captures powder bed heterogeneity and its influence on energy absorption and heat transfer; spatial beam shaping effectsHigh accuracy combining detailed optical physics with thermal–fluid predictions; validated against melt pool dimensions, temperature fields, and microstructure; supports digital twin development; enables prediction of defects (porosity, keyholing) and microstructureXCT, thermal imaging, pyrometry, in situ X-ray monitoring, melt pool metrics, experimental microstructure data, single-track and multi-track validationMedium: balances fidelity and efficiency through modular coupling; GPU acceleration improves feasibility for multi-layer simulationsCoupling complexity increases model development effort; trade-offs between resolution and computational feasibility; requires extensive experimental validation; integration of vapor dynamics and spatter remains incomplete; standardized validation protocols lacking[109,110,111,112]
Table 7. Melt pool modeling approaches in powder bed fusion.
Table 7. Melt pool modeling approaches in powder bed fusion.
Model Type/CategoryPhysics CapturedCore CapabilitiesValidation MethodsComputational Cost/EfficiencyLimitations/ChallengesRepresentative References
Analytical and Semi-Analytical ModelsSimplified heat conduction; process-dependent laser absorptivity; heat source modeling (Rosenthal, Goldak); temperature-dependent properties in extended versions; conduction-dominant heat transferClosed-form solutions for thermal field prediction; rapid melt pool geometry estimation (width, depth); process parameter mapping; printability window identification; suitable for rapid process parameter screeningEx situ validation with experimental melt pool dimensions; thermal profile measurements; validated with ~7–8% discrepancy for melt pool dimensionsVery high efficiency: computationally inexpensive (<200 s in some cases); orders of magnitude faster than numerical simulations; suitable for real-time applicationsNeglects fluid flow, Marangoni convection, recoil pressure, and keyhole dynamics; assumes conduction-dominant heat transfer; limited accuracy for transient phenomena and keyhole regimes; requires empirical calibration for quantitative precision; lacks detailed defect prediction capability[126,127,128]
Continuum Numerical Models (FEM/FVM for Thermal–Fluid Coupling)Comprehensive heat transfer (conduction, convection, radiation); fluid flow with Marangoni convection and recoil pressure; phase change and solidification; surface tension; vaporization effects; temperature-dependent material propertiesDetailed thermal history prediction; melt pool morphology and dimensions; thermal gradients and cooling rates; solidification behavior; supports defect prediction including porosity and lack of fusion; multi-layer and multi-track simulationsEx situ validation with experimental temperature profiles, melt pool geometry, thermal history; high-speed imaging correlation; validated with good agreement on temperature gradients and melt pool morphologyModerate to high cost: computationally intensive for multi-layer or multi-track simulations; adaptive meshing and parallelization improve efficiency; ~4.4× faster than traditional FEM in optimized FVM implementationsMesh-based methods face challenges with violent interface dynamics and complex free surface flows; require remeshing or interface tracking algorithms; high computational demands limit scalability to large-scale simulations; simplified powder bed morphology in some implementations[104,129,130,131,132]
High-Fidelity CFD with Multiphysics CouplingComprehensive thermal–fluid dynamics; Marangoni convection; recoil pressure; surface tension; evaporation and vaporization; phase change and solidification; powder–laser interaction; keyhole formation, vapor depression, and oscillations; sulfur-induced flow transitionsHigh-accuracy prediction of melt pool size, shape, temperature distribution, and fluid flow patterns; captures conduction and keyhole modes; defect formation mechanisms including porosity (lack of fusion, keyhole pores), balling, spatter, inter-track voids, and bubble dynamics; fluid instabilities and melting-solidification characteristicsEx situ validation with experimental melt pool dimensions, temperature measurements, high-speed imaging; in situ monitoring data correlation; X-ray imaging for keyhole dynamics; validated against synchrotron dataModerate-to-high demand: multiphysics coupling increases complexity and computational cost; computationally intensive, limiting scalability for multi-layer and complex geometriesComputationally intensive; limited scalability for industrial-scale multi-layer simulations; sensitivity to numerical parameters; requires accurate boundary conditions and material properties; complex powder bed morphology often simplified; integration challenges across scales[116,117,133,134,135]
Mesh-Free and Particle-Based Methods (SPH, LBM, DEM)SPH: fluid flow with complex free surface dynamics, Marangoni convection, recoil pressure, surface tension, evaporation, thermo-capillary flow, phase transitions, melting mode transitions; LBM: fluid flow, heat transfer, phase change, keyhole oscillations, vapor capillary evolution; DEM: powder packing, spreading, granular dynamics, particle-level interactionsSPH: detailed melt pool dynamics, captures powder behavior and melting, fluid flow with complex interfaces, superior handling of violent interface dynamics, balling defects, keyhole formation; LBM: melt pool fluid dynamics, keyhole oscillation dynamics; DEM: powder bed characterization, powder spreading simulation, powder wetting effects on thin wall morphologyEx situ validation with experimental melt pool shapes, surface temperatures, defect mechanisms (balling), thin wall morphology; high-speed imaging and synchrotron data correlation; validated qualitatively and quantitatively with experiments on melting mode transition, vapor depression geometryModerate-to-high cost: SPH moderate with incompressible schemes and optimized solvers; computationally expensive for large-scale simulations; DEM moderate for powder packing; LBM high for detailed fluid dynamics; both limit scalabilitySPH: careful calibration of smoothing lengths and kernel functions required; less mature integration with solid mechanics; particle discretization simplifications; LBM: complexity in multi-phase flows; DEM: simplifications in particle interactions; limited integration with macro-scale models; scalability constraints for industrial applications[84,91,107,136,137,138]
Multi-Scale Coupled Models (Phase-Field, CA, Physics-Informed ML)Phase-Field/CA: coupled thermal–fluid dynamics with rapid solidification kinetics, grain growth mechanisms, solid/liquid phase transitions, Marangoni convection, recoil pressure, microstructure evolution; ML/PINN: implicitly models temperature fields, melt pool boundaries, thermal gradients, physics-guided architectures encode process parameter dependenciesPhase-Field/CA: links melt pool dynamics to microstructure outcomes (grain size, morphology), predicts porosity and grain structure, supports process–structure–property relationships; ML/PINN: rapid prediction of 3D thermal fields and melt pool geometry (2–3% error), strong generalization across process parameters, near-instantaneous inference after trainingPhase-Field/CA: ex situ validation with microstructure characterization (grain size, morphology), porosity measurements, melt pool dimensions; ML/PINN: trained on high-fidelity CFD/FEM simulation data and experimental datasets; validated against experimental melt pool dimensions and thermal measurementsPhase-Field/CA: high cost due to multiphysics and multi-scale coupling; computationally intensive, limiting domain size; ML/PINN: very high efficiency for inference after training; orders of magnitude faster than physics-based simulationsPhase-Field/CA: computationally intensive; limited to small domains or simplified geometries; complex coupling strategies; requires extensive calibration; scalability challenges for industrial-scale simulations; ML/PINN: requires extensive high-quality training data; limited physical interpretability; extrapolation beyond trained parameter ranges unreliable; generalizability depends on training data diversity[110,111,139,140,141]
Table 8. Influence of key parameters on residual stress and distortion in PBF.
Table 8. Influence of key parameters on residual stress and distortion in PBF.
ParameterConditionEffect on Peak Residual StressEffect on DistortionMechanismReferences
Preheating TemperatureAmbient (20 °C)Baseline (500–700 MPa for Ti-6Al-4V)Baseline (0.3–0.5 mm for 50 mm parts)High thermal gradients, limited stress relaxation[156,157]
Elevated (200–300 °C)Reduced by 40–60%Reduced by 50–70%Lower thermal gradients, enhanced plastic relaxation[156,166,168]
Laser AbsorptivityLow (A = 0.30–0.35)Reduced by 15–25%Reduced by 10–20%, risk of lack of fusionReduced energy coupling, smaller melt pool, lower thermal gradients[171,172]
High (A = 0.55–0.70)Increased by 15–25%Increased by 10–20%Enhanced energy deposition, larger melt pool, higher thermal gradients[156,173]
Laser PowerLow (150–180 W)Moderate (300–450 MPa)Moderate (0.2–0.3 mm), fusion quality riskReduced melt pool size, lower peak temperatures[156,168]
High (230–280 W)Elevated (550–750 MPa)Elevated (0.4–0.6 mm)Enlarged melt pool, increased thermal gradients and plastic strain[156,174]
Scan SpeedLow (600–800 mm/s)Elevated (500–700 MPa)Elevated (0.4–0.5 mm)High energy density, extended melt pool, cumulative heating[168,174]
High (1000–1400 mm/s)Reduced (350–500 MPa)Reduced (0.2–0.3 mm), fusion quality riskLow energy density, reduced thermal interaction[156,174]
Hatch SpacingSmall (0.08–0.10 mm)Moderate to elevated (450–600 MPa)Moderate (0.3–0.4 mm)High track overlap, re-melting, cumulative heating[168,172]
Large (0.12–0.15 mm)Reduced (350–500 MPa), lack-of-fusion riskReduced (0.2–0.3 mm)Low track overlap, reduced thermal interaction[172,174]
Laser Operation ModeContinuous Wave (CW)Baseline (500–700 MPa)Baseline (0.3–0.5 mm)Steady-state thermal gradients, continuous melting[156,171]
Pulsed (optimized)Reduced by 20–40% (300–500 MPa)Reduced by 25–45% (0.15–0.3 mm)Intermittent heating, stress relaxation during pulse-off, reduced peak gradients[169,171]
Scanning StrategyUnidirectionalAnisotropic stress (300–600 MPa range)Anisotropic (0.2–0.5 mm directional)Directional thermal gradients, accumulated strain along scan[167,175]
Island/CheckerboardReduced, more isotropic (350–500 MPa)Reduced by 30–50% (0.15–0.3 mm)Localized heating, reduced global thermal gradients, stress compartmentalization[176,177]
Table 9. Comparative evaluation of modeling approaches in PBF.
Table 9. Comparative evaluation of modeling approaches in PBF.
Modeling ApproachPhysics CapturedCore CapabilitiesValidation StrategiesComputational EfficiencyScalabilityTransferabilityIndustrial ApplicabilityKey LimitationsReferences
Empirical and Semi-Empirical Models (Regression-Based Correlations)Effective thermal behavior captured through statistical correlations; limited explicit physics representationRapid prediction of melt pool dimensions, temperature histories, and thermal characteristics; process parameter mapping; printability window identificationCalibrated with experimental data; validated against melt pool geometry and thermal measurementsVery high computational efficiency; low computational cost, enabling rapid predictions suitable for process mapping and controlLimited scalability to complex geometries; primarily suited for part-scale and process-scale simulationsLimited transferability; heavily dependent on experimental calibration; constrained to tested materials and process conditionsSuitable for rapid process optimization, preliminary design space exploration, and printability assessments; enables quick process parameter studiesLack mechanistic insight; oversimplify complex thermal phenomena; neglect fluid flow and phase changes; require extensive calibration; limited generalizability beyond specific experimental conditions[127,189,190,191]
Analytical Models (Rosenthal-Type Heat Conduction, Moving Point Heat Source)Dominant heat conduction with simplified boundary conditions; temperature-dependent properties in extended versions; conduction, convection, radiation, and melting losses in advanced formulationsClosed-form or semi-closed-form solutions for thermal field prediction; melt pool geometry estimation; rapid thermal history predictions; residual stress estimations; process window identificationEx situ validation against experimental thermal profiles and melt pool dimensions; validated with thermographic measurements; sensitivity analyses on assumptionsHighly efficient; orders of magnitude faster than full numerical simulations; suitable for rapid predictions (<200 s in some cases)Part-scale thermal modeling; applicable across various scanning strategies; limited to simplified geometries and boundary conditionsModerate transferability; requires empirical calibration for quantitative precision; assumptions limit applicability to new materials without adjustmentFacilitates rapid process parameter mapping; supports printability assessments and process control strategies; useful for initial design and optimizationRely on simplifying assumptions (steady-state, linear heat conduction, idealized heat sources); neglect fluid flow, keyhole effects, and complex melt pool dynamics; limited accuracy for transient and multiphysics phenomena; require empirical corrections for nonlinear effects[119,128,192,193]
Semi-Analytical Models with Nonlinear CorrectionsHeat conduction with temperature-dependent material properties; incorporates nonlinear effects and phase change; includes convection and radiation boundary conditionsEnhanced thermal field prediction with improved accuracy over linear analytical models; residual stress and deformation prediction; rapid part-scale thermal modelingEx situ validation with experimental data; ~90% accuracy validated in some implementations; sensitivity analyses on material propertiesVery efficient compared to full numerical models; minimal computational overhead vs. linear analytical models; enables rapid iterative designPart-scale thermal and thermomechanical modeling; applicable to various scanning patterns and geometriesImproved transferability over pure analytical models; still requires empirical calibration; better generalization with nonlinear correctionsSupports efficient process optimization and defect prediction; balances accuracy and speed for industrial applications; useful for process controlLinear base model assumptions with empirical calibration; still neglects detailed fluid flow and complex melt pool hydrodynamics; limited to conduction-dominant scenarios[194,195,196]
Finite Element Method (FEM) for Thermomechanical ModelingComprehensive heat transfer (conduction, convection, radiation); thermomechanical coupling; residual stress and distortion; phase transformations; temperature-dependent material propertiesDetailed thermal history prediction; residual stress and deformation analysis; part-scale distortion prediction; multi-layer simulations; microstructure evolution couplingEx situ validation with experimental measurements; validated against distortion, residual stress measurements; ~8% displacement and ~3.5% stress error in optimized implementationsModerate-to-high computational cost; computationally intensive for multi-layer or part-scale simulations; adaptive remeshing and parallelization improve efficiencyPowder to part scale; scalable with adaptive strategies and model simplifications; suitable for multi-layer and complex geometriesGood transferability across materials with proper calibration; material property databases support cross-material applicationIndustry standard for detailed thermal and mechanical analysis; supports design optimization, distortion mitigation, and qualification; enables digital twin integrationHigh computational demands limit real-time applicability; requires extensive material property data; mesh generation complexity; simplified powder behavior in some implementations; linearized material models reduce accuracy[2,94,197,198]
Finite Volume Method (FVM) for Thermal AnalysisHeat conduction dominant; includes convection and radiation; temperature-dependent properties; phase change effectsThermal field prediction; temperature history calculation; melt pool thermal analysis; efficient heat transfer simulationsEx situ validation; good accuracy vs. FEM benchmarks; validated against experimental thermal profilesHigher computational efficiency than FEM (~4.4× faster reported); moderate computational cost with improved solver efficiencyPart-scale thermal analysis; applicable to large domains with efficient discretizationGood transferability with proper boundary condition specification; similar to FEM in cross-material applicabilitySuitable for rapid thermal analysis in industrial settings; supports process optimization with reduced computational burden vs. FEMSimplified fluid flow effects; less mature than FEM for multiphysics coupling; limited representation of complex melt pool dynamics; requires proper boundary condition specification[103,199]
Smoothed Particle Hydrodynamics (SPH) and Particle-Based MethodsComprehensive fluid flow and heat transfer; free surface tracking; phase change and solidification; powder–laser interaction; granular flow dynamics; melt pool hydrodynamicsDetailed melt pool dynamics modeling; captures powder behavior and melting; fluid flow with complex interfaces; powder spreading simulation; multi-phase interactionsEx situ validation with experimental melt pool morphology; validated against synchrotron imaging; benchmarked with experiments on welding and AMComputationally expensive; high computational cost limits large-scale applications; moderate-to-high cost depending on particle resolutionPowder-to-melt pool scale; mesoscale to macro-scale applicability; limited scalability to full part-scale due to computational demandsModerate transferability; flexible for different materials and process conditions; assumptions in particle interactions affect transferabilityEffective for detailed melt pool and powder dynamics research; supports fundamental process understanding; useful for phenomena not captured by continuum methodsComputationally intensive; complex boundary and interface assumptions; less mature than mesh-based methods; particle discretization simplifications; limited to small-scale or mesoscale simulations in practice[91,106,184]
Lattice Boltzmann Method (LBM) and Discrete Element Method (DEM)LBM: fluid flow, heat transfer, phase change at mesoscale; DEM: powder packing, spreading, granular dynamics, particle-level interactionsLBM: melt pool fluid dynamics and thermal modeling; DEM: powder bed characterization, powder spreading simulation, powder–laser interaction modelingEx situ validation; DEM validated against powder bed characteristics; LBM validated for fluid flow and thermal fieldsComputationally expensive; DEM for powder packing moderate cost; LBM for fluid dynamics, high cost; both limit scalabilityPowder scale to mesoscale; DEM effective for powder bed modeling; LBM for melt pool scale; limited to small domainsTransferable across powder materials for DEM; LBM requires calibration for different fluids and materialsDEM useful for powder bed quality assessment and spreading optimization; LBM for detailed melt pool physics research; supports process understandingHigh computational cost limits practical application; DEM simplifications in particle interactions; LBM complexity in multi-phase flows; limited integration with macro-scale models; scalability constraints[103,200]
Multi-Scale and Multiphysics Integrated FrameworksCoupled thermal, fluid, mechanical, and microstructural physics; heat transfer, mass transfer, phase transformations; powder dynamics to part-scale phenomenaComprehensive process simulation across scales; links powder behavior to melt pool dynamics to part-scale properties; microstructure evolution prediction; defect formation analysisMulti-scale validation with experimental data at different scales; validated against thermal, mechanical, and microstructural measurementsComputationally intensive; high computational cost due to multiphysics coupling; efficiency improvements via adaptive strategies and scale separationMulti-scale from powder to part scale; framework supports integration across scales but increases complexityEnhanced transferability through physics-based foundations; multi-scale coupling improves generalization but requires extensive calibrationSupports comprehensive process understanding and optimization; enables process–structure–property linkage; useful for digital twin development and advanced process controlComplexity in coupling strategies; high computational demands limit practical deployment; simplifications in coupling physics; data requirements for validation; integration challenges across scales; model calibration complexity[8,112,201,202,203]
Table 10. Hybrid physics–data-driven modeling approaches.
Table 10. Hybrid physics–data-driven modeling approaches.
Model CategoryApplication DomainPhysics Integration StrategyValidation MethodComputational EfficiencyTransferability/ScalabilityKey LimitationsReferences
Physics-Informed Neural Networks (PINNs)Temperature field and melt pool geometry prediction; thermal history modeling; parameter identification; real-time process monitoring and adaptive controlStrong coupling: governing PDEs (heat equation, Navier–Stokes) embedded as physics-informed loss functions; conservation laws integrated into network training; dynamic weight updates with physics-informed constraints; ontology-integrated frameworksEx situ validation with FE simulations and experimental data; in situ validation with real-time anomaly prognosis; validated across scanning speeds, process parameters, and 3D benchmark problems; errors below 3% for 2D temperature fieldsHigh efficiency: computational time significantly reduced vs. pure FEM/CFD (MAPE 2.8%, R2 0.936); efficient with limited and sparse data; enables real-time monitoring and adaptationGeneralizes to unseen laser scanning strategies and geometries; transferable to different builds without retraining; demonstrated on NIST benchmark parts; adapts to new materials and parametersRequires partial experimental data; accuracy depends on physics model fidelity; most models focus on 2D or single-track simulations; integration of multi-sensor data streams and handling noisy data remain challenges; requires valid digital twin[140,218,219,220,221,222]
Deep Neural Operators and Physics-Based Surrogate Models for Digital TwinsHigh-fidelity melt pool state prediction; closed-loop feedback control; temperature and defect prediction with online calibration; digital twin integration for adaptive manufacturingHybrid: physics-based surrogate models combined with deep neural operators (Fourier Neural Operators); offline fine-tuning with physics simulations; incorporates uncertainty quantification and Bayesian calibrationClosed-loop feedback control with online updates; ex situ validation on experimental datasets; supports adaptive calibration; validated with multi-fidelity FE and ML calibrationHigh efficiency: orders of magnitude faster than traditional FE simulations; supports near-real-time inference; fast surrogate enables efficient closed-loop controlAdaptable via offline fine-tuning and calibration; incorporates uncertainty quantification for evolving conditions; supports layer-by-layer parameter adjustmentHigh computational cost for initial high-fidelity simulations; complexity of model integration; dependence on synthetic or simulated data may not capture all experimental variability; calibration against limited experimental datasets remains bottleneck[5,7,223,224]
Hybrid CFD/FEM–ML Surrogate ModelsMelt pool width and geometry prediction; defect formation prediction; process parameter optimization; printability mapping and qualification accelerationModular coupling: CFD/FEM simulations provide training data for ML surrogates (SVM, U-Net, ensemble methods); integrates experimental and simulation data; data-driven methods augment physics-based predictions; scientific ML with physically intuitive featuresEx situ validation integrating experimental and simulation data; validated on Ti-6Al-4V, AlSi10Mg, and multiple alloys; combines simulation and empirical data; R2 > 0.98 for melt pool geometryHigh efficiency: faster than high-fidelity CFD/FEM; accelerates qualification of LPBF parameters; reduces error norms by up to 75%; relative mean absolute error ~6.77%Effective under sparse data with multiple chemistries; considers powder bed thickness and preheating effects; integrates simulation and experimental data; outperforms black-box modelsNot explicitly real-time but supports process insight; limited to specific parameter sets or small geometries; dependence on simulation quality for training data; transferability across machines and materials requires further validation[204,225,226,227]
Physics-Guided Generative Models and Hybrid Neural NetworksMelt pool behavior prediction; defect identification (lack of fusion, porosity, keyhole); built quality and defect type prediction; anomaly detectionEmbedded physics: generative adversarial networks (GANs) guided by physics; hybrid neural networks fusing thermal images and simulated melt pool images; mechanistic model integration with synchrotron X-ray dataEx situ validation with experimental data; 97.25% defect identification accuracy; multi-classification framework for quality prediction; experimentally validated across multiple regimesHigh efficiency: accelerated prediction vs. traditional methods; reduced CFD computational cost; efficient multi-classification framework; temperature metrics with 5–15% uncertaintyValidated experimentally; combines physical and data features for robustness; explores parameter influence on quality; robust under sparse data; tested across regimesImage prediction quality prioritized over exact quantitative agreement in some cases; limited validation on complex geometries; reliance on synthetic data for training; accuracy depends on quality of physics simulations[228,229,230,231]
Physics-Based Thermal Models with ML for Microstructure EvolutionMicrostructure evolution prediction (primary dendritic arm spacing, grain growth, melt pool depth); scan path optimization for microstructure control; process–structure–property linkagePhysics-based thermal model inputs to ML (SVM, U-Net surrogates); combines phase-field modeling with ML; mechanistic modeling integrated with ML; incorporates recoil pressure and fluid flow effectsEx situ validation; validated on multi-layer and multi-track cases; experimentally validated microstructure predictions; RMSE ~0.012 mm for melt pool depth, 110 nm PDASHigh efficiency: rapid part-level thermal model with ML prediction; reduces computational time by ~1000× vs. phase-field simulations; enables intelligent process optimizationTransferable to different builds without retraining; demonstrated on NIST benchmark parts; geometry-agnostic capabilities; applied to Inconel 625 and molybdenum materialsScalability to full parts or complex geometries limited; high computational cost when applied to full parts; most models focus on single-track or few-layer scenarios; limited validation on complex multi-layer builds[203,232,233]
Table 11. In situ monitoring and data fusion approaches for improved understanding of the PBF process.
Table 11. In situ monitoring and data fusion approaches for improved understanding of the PBF process.
Monitoring ApproachSensor Types and Data CharacteristicsValidation LevelComputational PerformanceTransferabilityIndustrial ReadinessBest Use Cases and Rationale
Thermal/IR Imaging for Melt Pool and Layer MonitoringHigh-speed NIR/LWIR cameras, imaging spectrometers, mid-IR collectors; spatial resolution down to 50 µm (ADM optic); layer-wise thermal maps [234,235,236].Validated via registration to XCT and optical microscopy for pore morphology/distribution across multiple geometries [235,236,237].High frame rates (specific rates vary); temperature–emissivity separation (TES) achieves ±28 K accuracy over 1000 K range [234]; many systems require offline processing for mapping [4,11].Demonstrated on Ti, 316L, Ni superalloys, diverse geometries; cross-machine generalization limited to reported testbeds [235,236,238].TRL 5–6; ADM achieved micro-CT correlation to 4.3 µm pores on testbed; not yet turnkey industrial product [235].Best for layer-wise anomaly localization, thermal history tracking, correlating surface thermal signatures to XCT porosity maps, as spatial resolution enables precise defect localization when coupled with rigorous registration [234,235,236].
Photodiode-Based MonitoringOn-axis photodetectors, ratiometric bichromatic sensors measuring integrated melt-pool emission; calibrated against tungsten lamp and blackbody [239,240].Calibration studies and repeatability checks on testbeds; signals correlated to density patterns and edge effects [239,240].High temporal bandwidth; sampling rates/inference latencies not comprehensively reported [239].Used across different machine optics with normalization strategies; broad cross-machine transfer claims limited [239,240].TRL 7–8; widely integrated on commercial systems; NIST calibration guidance exists [240].Best for fast melt-pool event detection, global trend monitoring, feedstock/optics health checks, as single-point high-bandwidth signals enable rapid anomaly detection without spatial imaging overhead [239,240].
Multi-Sensor Fusion (Visible/NIR/SWIR/LWIR/Acoustic)Fusion of visible, NIR, SWIR, LWIR, optical tomography, acoustic emission, back-reflection, scan metadata in voxelized footprints [241,242,243,244].Strong XCT validation; voxel-by-voxel binary classification: 98.5% accuracy; POD for 200–1000 µm flaws with a 90/95 metrics [242,243,244].Neural networks and variational autoencoders; low-latency designs targeted but detailed metrics sparse; human-in-the-loop annotation improves POD [243,244].Demonstrated on testbeds and industrial components; multi-laser cases explored; generalization often requires re-annotation/retraining [235,243,244].TRL 5–6; INDE framework shows production-scale engineering with POD reporting consistent with NDE practice [243,244].Best for detecting multiple defect types (lack-of-fusion, keyhole, subsurface porosity) and producing probabilistic POD curves for qualification, as multi-modal data captures complementary physics; 98.5% classification accuracy and POD metrics support certification workflows [242,243,244].
Machine Learning for Defect Detection and ClassificationDeep CNNs, U-Net variants, CNN + LSTM hybrids, variational autoencoders on single/fused sensor data [242,245,246].Ground truth: XCT; reported accuracies 97.86–98.5% for defect/regime classification; voxel classification via cross-validation [242,245,246].Rasterized layer images or voxelized footprints; 0.5–4 ms regime detection targeted; many pipelines trained offline [246].Demonstrated across geometries/materials within study scope; applied to industrial geometries with retraining [243,245].TRL 4–6; high-performance lab/testbed demonstrations; some integrated into analytics platforms; full factory closed-loop limited [243,244].Best for supervised in situ detection where extensive XCT ground truth is available; producing POD/PDFA metrics for qualification, as high classification metrics (≥97.86%) when trained on rigorous XCT-labeled data [242,245,246].
Data Registration Techniques (In Situ to XCT)Image-to-volume registration, adaptive volume adjustment, fiducials, deformation modeling to align thermography/optical tomography to XCT [237,247].Necessary and effective for correlating in situ signals to XCT; adaptive methods improve alignment accuracy and reduce false positives/negatives [237,247].Typically offline and computationally intensive; real-time registration latencies not reported for full volumes [237,247].Applied across multiple specimen shapes (cylinders, complex geometries); fiducials and deformation models improve cross-part mappings [237,247].TRL 3–5; critical enabling step for ML training and POD estimation; research-grade, not embedded in real-time factory systems [237,247].Best for producing trustworthy ground truth alignment between in situ signals and XCT for ML training and POD estimation, as registration accuracy directly impacts ML model reliability and defect localization precision [237,247].
Temperature Calibration and Emissivity MeasurementMulti-wavelength pyrometry, temperature–emissivity separation (TES), pixelwise camera calibration, tungsten lamp reference [234,238,239,248,249].TES: ±28 K retrieval accuracy over 1000 K range; pixelwise calibration: 500–1500 K; multi-wavelength studies show 20–300% emissivity variation across alloys/phases [234,238,248,249].Requires spectrally resolved sensors and careful optics; computational cost of TES/pixelwise fitting reported; real-time absolute temperature mapping challenging [234,249].Emissivity varies by alloy and process stage; multi-wavelength methods demonstrate need for in situ measurement vs. fixed models [238,248].TRL 5–6; demonstrated in testbeds and EB-PBF platforms; adoption in production LPBF increasing but requires per-machine calibration [234,238,248,249].Best for obtaining quantitative process temperatures, improving physics-based models, enabling calibrated closed-loop control, as absolute temperature accuracy (±28 K) enables physics-model validation and thermal-based control [234,248,249].
Real-Time Closed-Loop ControlMelt-pool thermal emission feedback to modulate laser power; customized LPBF platforms for on-the-fly control [250].Lab demonstrations of closed-loop power control based on melt-pool emission on custom platforms [250].Real-time control loops implemented in testbeds; explicit latency and industrial cycle-time scalability not comprehensively reported [250].Demonstrations limited to custom/test platforms; industrial scale-up evidence limited in reviewed corpus [250].TRL 4–5; prototype demonstrations with promising regulator designs; widespread factory integration remains active R&D [250].Best for dynamic compensation of melt pool instabilities (over melt, balling) in research or pilot production contexts on custom testbeds with specific materials, as closed-loop feedback demonstrated feasibility for real-time process adjustment in controlled laboratory settings [250].
Table 12. Sensor technologies and data integration in PBF digital twins.
Table 12. Sensor technologies and data integration in PBF digital twins.
Technology/ApproachSpecifications and Data CharacteristicsValidation LevelComputational PerformanceTransferabilityIndustrial ReadinessBest Use Cases and Rationale
Thermography and PyrometryTwo-wavelength coaxial imaging pyrometry and off-axis thermal imaging for melt-pool temperature mapping; photodiode pyrometry sampling >100 kHz; datasets reach hundreds of thousands of frames and hundreds of GB [263,264,265].Correlated to μCT/XCT and operando X-ray radiography for pore/keyhole validation and surface topography prediction [263,264,265].High sampling (>100 kHz) enables sub-ms feature windows for ML; data volumes reach hundreds of GB per multi-build campaign [264,265,266].Demonstrated across lab printers and multiple builds; domain shift issues require domain alignment/augmentation [256,266].TRL 6–7; mature for monitoring; integration with data pipelines shown but FAIR standards incomplete [266,267].Best for temperature control, melt-pool energy tracking, surface topography inference, and as primary features for defect classification models. Why: High temporal resolution (>100 kHz) enables precise thermal event capture; direct correlation to temperature physics supports model validation [263,264,265].
Acoustic Emission MonitoringStructure-borne and airborne AE capture rapid mechanical/pressure transients from keyhole collapse and spatter; signals resolved at sub-ms time windows; contribute strongly to pore prediction [265,268,269].Validated with operando synchrotron X-ray for temporal registration of pore events and XCT comparison for predicted porosity [265,268].Low data volume relative to imaging; enables ML detection within 0.5–4 ms windows [265,268].Shown to transfer across experiments but sensitivity to machine acoustic environment and sensor placement reported [266,267].TRL 5–6; high promise for real-time detection; fewer commercial turnkey AE solutions for PBF than optical tools [267].Best for fast event detection (keyhole pore formation, spatter), early warning signals for closed-loop corrections. Why: Low data overhead with high temporal sensitivity to transient events; strong correlation to pore formation validated by operando X-ray [265,268,269].
Melt Pool Imaging (Optical, Coaxial, Off-Axis)High-speed visible/SWIR cameras and coaxial sensors capture morphology, intensity, area; dynamic ROI cameras operate up to 20 kHz [263,270,271].Co-registered to XCT/μCT and operando X-ray; spatial registration methods map melt pool signatures to part coordinates for topography and defect correlation [263,264,265].Highest data volumes (105–106 frames per build); GPU pipelines required for real-time processing; continuous capture demands ROI or compression [266,270,271].Imaging models show reduced performance under different instruments; domain adaptation improves cross-setting accuracy [256,266].TRL 6–7; widely used in research; scaling to full production requires high-throughput pipelines and ROI strategies [266,271].Best for morphology-based defect detection, surface topography prediction, training labels for other sensors. Why: Spatial resolution enables precise defect localization; registered imaging provides ground truth for ML training [263,264,265].
Multi-Sensor Fusion FrameworksCommon fusions: thermal + optical + acoustic + photodiode; fusion at data/feature/decision levels using CNNs, LSTMs, ensemble classifiers [264,265,268,269].Validated against operando X-ray and XCT; multi-modal ML achieves pore F1 up to 0.95, recall 1.0, classification accuracies ≈ 98% [264,265,268].Fusion improves predictive power but raises synchronization and computational load; ML inference over 0.5–4 ms windows [265,268].Fusion improves robustness but requires per-machine calibration; domain adaptation pipelines increase reusability (+31% detection without labels) [256,266].TRL 5–6; research prototypes show closed-loop correction in L-DED and DT architectures; production adoption partial due to integration complexity [256,267,269].Best for comprehensive defect detection and localization where multiple physical signatures are available. Why: Complementary sensors capture different physics; fusion achieves highest reported defect detection metrics (F1 = 0.95, recall = 1.0) [264,265,268].
Machine Learning for Multi-Sensor ProcessingCNN + LSTM hybrids and feature-level fusion; training datasets range from thousands to millions of frames and hundreds of GB; reported 61% UTS prediction error reduction using fused in situ data [256,268,272].Ground truth: XCT/μCT/metallography/tensile testing; reported classification accuracies ≈ 98% (regime), F1 up to 0.95, defect size classification 98.8% (large pores) [264,265,268,272].Inference times suitable for ms-scale decision windows; full-build scaling needs GPU acceleration and optimized pipelines [266,268,271].Models trained on one setting degrade on others; domain adaptation and augmentation recover performance [256,266,272].TRL 5–6; ML essential for DT updates; publicly shared datasets and HDF5 pipelines improve reproducibility but broader FAIR adoption limited [266,267].Best for automated defect classification, local mechanical property prediction, inputs to closed-loop controllers. Why: ML extracts complex patterns from multi-modal data; validated 61% error reduction in tensile property prediction [268,272].
Spatial–Temporal Data RegistrationMethods combine galvanometer coordinates, laser ON/OFF, camera alignment, ML-based image registration to map sensor records to part coordinates; two-camera (coaxial + off-axis) registration recovers spatial melt pool maps [263,264,265].Registration validated by predicting layer surface topography and correlating melt pool signatures with XCT porosity maps [263,264,265].Registration incurs moderate compute but essential for co-registered datasets; synchronization precision down to 50 µs with synchrotron timing [263,265].Necessary for multi-sensor generalization; registration enables cross-sensor mapping across geometries when scanner coordinates available [256,264,266].TRL 4–5; critical enabling step for DTs and in situ qualification; implementations exist in research toolchains [263,266].Best for spatially resolved defect localization, part-level mapping for mechanical property inference. Why: Registration accuracy directly impacts ML model reliability; enables precise correlation between in situ signals and ex situ validation [263,264,265].
FAIR Principles and Data ManagementPublic release examples: 230 GB HDF5 dataset linking in situ sensor footprints to tensile tests; pipelines process 700 k+ frames for feature extraction [266,272].Ground truth co-registration and labeling practices exist but no universal metadata/schema standard widely adopted [266,267].FAIR pipelines reduce retraining time and enable transfer but add overhead for metadata capture and storage [266].Domain adaptation and open datasets improve reuse and DT reusability across institutions [256,266,272].TRL 3–5; partial adoption: exemplar datasets released but community standards and wide FAIR compliance remain limited [266,267].Best for dataset sharing for ML model training, benchmarking DT models, regulatory qualification. Why: Standardized data formats enable reproducibility and cross-institution model validation; 230 GB public dataset demonstrates feasibility [266,272].
Real-Time Processing and Closed-Loop ControlReal-time ML inference on 0.5–4 ms windows; system demonstrations of automated laser power control or defect correction in lab platforms [268,273].Closed-loop demonstrations in custom platforms and L-DED for defect correction; real-time decision support in digital twins [269,273].Achieved with GPU-accelerated pipelines and ROI imaging; full-build closed-loop across large parts challenging due to data throughput [266,271,273].Lab demonstrations transferable to production with engineering investment; standardization and safety qualification outstanding [267,273].TRL 4–5; early stage for full industrial deployment; specific closed-loop functions (power modulation, local repair) demonstrated [273].Best for on-the-fly power modulation, defect mitigation actions, probabilistic porosity control in DTs. Why: Demonstrated feasibility of ms-scale closed-loop control; probabilistic DT framework shown for online Bayesian calibration [273].
Table 13. AI integration in PBF digital twins.
Table 13. AI integration in PBF digital twins.
AI Method/FrameworkApplication DomainPhysics Integration LevelSensor ModalitiesValidation MethodComputational EfficiencyTransferabilityKey LimitationsReferences
Convolutional Neural Networks (CNNs) and Deep Learning for Defect DetectionDefect detection and classification (porosity, lack of fusion, surface defects, melt pool anomalies); quality assessment; in situ process monitoring; multi-label anomaly detectionPurely data-driven deep learning; transfer learning and self-supervised learning implementations reduce labeling requirementsOptical imagery (powder bed images, melt pool images); multi-modal fusion (visible and infrared cameras); layer-wise optical tomographyEx situ and in situ validation; 82–99.79% accuracy in defect detection and classification; validated against XCT ground truth dataHigh: real-time processing with practical latency (under 18 s per layer); enables real-time multi-label anomaly detection and intra-layer closed-loop controlHigh: transfer learning enables quality monitoring across materials with >92% accuracy; self-supervised learning handles imbalanced datasets without extensive labelingRequires extensive labeled datasets for supervised methods; limited interpretability in complex architectures; limited to single or specific sensor modalities in some implementations; data scarcity for rare defect types[6,272,287,289,295,296,297,298,299]
Physics-Informed Neural Networks (PINNs) and Deep Neural Operators for Thermal PredictionTemperature prediction; full-field thermal modeling; melt pool geometry prediction; parameter identification; high-fidelity melt pool state prediction; closed-loop feedback control; scalable temperature distributionHybrid physics-informed deep learning: custom loss functions enforcing physical behavior (PDEs embedded); Fourier neural operators combine physics-based simulations with data-driven learning; graph neural networks (GNNs) learn physics from FEA simulationsInfrared camera data; thermal imaging; physics-informed variables extracted from simulations; GNNs trained on FEA simulation dataEx situ validation with less than 7% deviation; R2 > 0.98 for melt pool geometry; closed-loop feedback control with online calibration; GNN validation with 3.77% MAPEHigh: computational time significantly reduced vs. pure FEM/CFD (up to 3900× speed-up); efficient with limited and sparse data; architecture-driven approaches effective with minimal dataHigh: transferable to different builds without retraining; demonstrated on NIST benchmark parts; geometry-agnostic capabilities; GNNs are multi-laser capable and transferable across geometriesRequires partial experimental data; accuracy depends on physics model fidelity; validation on limited geometries and materials; computational demands for high-fidelity simulations; surrogate accuracy depends on training data quality[5,220,224,227,280,300,301,302]
Ensemble Methods and Hybrid ML-Physics Models for Process OptimizationMelt pool geometry prediction; defect classification (porosity, melt pool stability); density and hardness prediction; process parameter optimization; robust design optimizationPurely data-driven ML (random forest, XGBoost, extra trees) with physics-augmented data generation and CFD-informed training; Bayesian calibration with stochastic modeling and probabilistic frameworksTrained on process parameters and simulation/experimental data; melt pool monitoring data; infrared thermal imaging with dimensionality reductionEx situ validation; 99.79% accuracy in melt pool stability classification; R2 up to 0.95 for density prediction; in situ layer-wise parameter adjustment; probabilistic approaches support robust optimizationHigh: fast inference enabling online monitoring (0.4 ms in some cases); runtime speed-up of 3900× compared to physics models; surrogate models reduce computation, enabling fast updatesHigh: transferable to different builds without retraining; demonstrated on Ti-6Al-4V, IN 625, SS316L, AlSi10Mg; model tailored for individual parts with model updatingRequires calibration with experimental data; potential for overfitting with complex models; requires high-quality continuous sensor data; computational demands for frequent updating; uncertainty quantification can increase computational burden[5,256,296,303,304,305,306,307]
Multi-Modal Sensor Fusion with Deep LearningFlaw detection; defect identification (subsurface porosity, cracking, keyhole pores); quality assessment; multi-material composition monitoring; comprehensive in situ process monitoringHybrid ML-assisted and physics-based sensor fusion; deep learning with multi-modal integration; contrastive loss functions and attention mechanismsMulti-modal: optical, acoustic emission, thermal imaging (visible and infrared cameras), spectral data, X-ray radiography with spatial–temporal registrationIn situ validation; 95–98.5% accuracy in flaw/defect detection; 0.95 F1-score for keyhole pore prediction; validated against XCT ground truth; high temporal resolution monitoringMedium to High: feature-level fusion balances accuracy and computational cost; high-speed systems enable intra-layer control; computational overhead from sensor synchronizationModerate: contrastive learning improves multi-material composition monitoring; limited cross-material validation in most studiesHigh complexity in sensor synchronization and data heterogeneity; challenges in data volume, standardization, and registration errors; limited validation on complex geometries; multi-modal datasets scarce; limited interpretability[243,268,270,287,295,308,309]
Reinforcement Learning and Model Predictive Control for Adaptive Digital TwinsLaser scan path optimization; thermal uniformity; residual stress reduction; geometry-agnostic melt pool control; real-time parameter adjustment; multi-step temperature tracking; adaptive parameter control; block quality predictionHybrid: reinforcement learning with reduced-order simulation and physics-based models; ML surrogate models with MPC frameworks; hybrid RNN/LSTM + reinforcement learning; ML and Bayesian optimization integratedReal-time sensor feedback for control; thermal sensors for layer-wise control; time-series deep neural networks for MPC; real-time sensor data for temperature predictionIn situ validation; demonstrated error reduction experimentally; MPC achieves precise melt pool temperature tracking and outperforms PID control; real-time parameter tuning via RLHigh: framework supports dynamic process optimization; multi-step predictive models facilitate timely regulation; real-time temperature prediction and process optimizationModerate: geometry-agnostic controller demonstrated; validated on Ti-6Al-4V and AISI 316L; some transferability limitations to different process types (e.g., DED to LPBF); limited cross-material validationRequires extensive training data for RL; validation on specific processes may limit transferability; physics model accuracy critical for MPC; computational demands for frequent updating; model complexity requires extensive training data; limited interpretability[7,291,292,310,311,312,313]
Table 14. Digital twin implementation in PBF.
Table 14. Digital twin implementation in PBF.
Digital Twin Type/ArchitectureCore FunctionsModel CompositionSensor and Data Integration StrategyValidation ApproachComputational Efficiency/ScalabilityIndustrial ReadinessKey LimitationsReferences
Hybrid Physics–ML Digital Twins with Adaptive Control and Bayesian UpdatingPorosity prediction and control; process parameter optimization; layer-wise model updating; real-time parameter tuning; thermal history control; closed-loop feedback; melt pool consistency and geometric accuracyHybrid: physics-based models with Bayesian calibration and probabilistic frameworks; ML surrogates integrated with MPC and reinforcement learning; data-driven adaptive control with G-code manipulationInfrared thermal imaging with dimensionality reduction; real-time sensor feedback (photodiode, melt pool size sensors, thermal sensors); layer-wise control inputsEx situ and in situ validation; layer-wise parameter adjustment; demonstrated process control with reduced defects and improved accuracy; MPC achieves precise temperature trackingHigh: surrogate models enable fast updates (runtime speed-up of 3900×); real-time processing with practical latency (under 18 s per layer); efficient framework balancing physics and MLHigh: layer-wise laser power adjustment, real-time parameter tuning, G-code manipulation, and closed-loop feedback control demonstrated; supports dynamic process optimizationRequires high-quality continuous sensor data; computational demands for frequent updating and MPC; physics model accuracy critical; limited to single sensor modalities in some implementations; validation on specific processes may limit transferability[7,256,314,315,316,317,318,319]
Physics-Informed Neural Network (PINN) and Deep Neural Operator Digital TwinsTemperature prediction; full-field thermal modeling; melt pool geometry prediction; parameter identification with reduced data requirements; real-time anomaly prognosis and diagnosis; accelerated process parameter selectionHybrid physics-informed deep learning: custom loss functions enforcing physical behavior (PDE embedded); physics-based data augmentation; ontology-driven interpretable learning; Fourier neural operators; multi-scale–multiphysics models with ML surrogatesInfrared camera data and thermal imaging; physics-informed variables extracted from simulations; emphasis on interpretable learning; no extensive sensor fusion focusEx situ validation with less than 7% deviation; R2 > 0.98 for melt pool geometry; in situ validation for real-time anomaly prognosis; modular validation frameworksHigh: computational time significantly reduced vs. pure FEM/CFD (up to 3900× speed-up); efficient with limited and sparse data; surrogates enable rapid decision-making across scalesModerate to High: enables faster simulation and potential for real-time process control; real-time monitoring with interpretable physics integration; surrogates enable rapid decision-making for digital twin applicationsRequires partial experimental data; accuracy depends on physics model fidelity; synthetic data may not capture all real-world variability; validation of limited geometries; ontology development requires domain expertise; surrogate accuracy depends on training data quality[205,220,227,320,321,322,323,324]
Multi-Modal Sensor Fusion and Hybrid CFD/FEM-ML Digital TwinsFlaw and defect detection; quality assessment; comprehensive process monitoring; melt pool width prediction; microstructure evolution; process–structure–property correlation; residual stress predictionHybrid: ML-assisted and physics-based sensor fusion with deep learning; combines physics-based simulations (CFD/FEM) with data-driven ML components; ROM with ML integration; combines simulated melt pool images with thermal imagesMulti-modal: optical imagery, acoustic emission, spectral data, thermal imaging with spatial–temporal registration; combines simulation (CFD/FEM) and experimental data for non-fusion approachesIn situ monitoring validation with 97–98.5% accuracy in defect detection; ex situ and modular validation; accurate prediction of meltpool depth and dendritic spacing; validated against XCT ground truthMedium to High: computational cost balanced by sensor fusion and ML; physical supervision network reduces CFD cost; hybrid models faster than pure CFD/FEM with 3900× speed-up for surrogatesModerate to High: validated for in situ monitoring with real-time capability; enables rapid microstructure prediction; model transferable to different builds without retraining in some implementationsHigh complexity in sensor synchronization and data heterogeneity; computational overhead; not real-time control for CFD/FEM-ML variants; limited validation on complex geometries; surrogate accuracy depends on training data and FEM model fidelity[5,203,204,231,242,287,325,326,327]
Comprehensive Industrial Digital Twin Frameworks with Knowledge Transfer and Multi-scale IntegrationComprehensive process monitoring, control, optimization, and cyber–physical system integration; knowledge transfer and domain adaptation for reusability; process chain modeling; predictive maintenance; rapid decision-makingModular hybrid frameworks integrating physics-based and data-driven models; multi-scale–multiphysics models with ML surrogates; domain adaptation techniques for cross-machine/sensor transferability; methodological frameworks for hybrid model constructionSensor integration emphasized for real-time feedback; multi-modal data in comprehensive frameworks; sensor data domain alignment across different machines and sensor configurationsModular validation; framework designed for practical industrial deployment; improved anomaly detection accuracy by 31% via domain adaptation; review frameworks discuss implementation challenges and trustworthinessFramework designed for practical industrial deployment with modular architecture enabling scalability and adaptability; surrogates enable rapid decision-making across scales; enhanced reusability via domain adaptationHigh: real-time process monitoring, control components, data analysis, and predictive maintenance; surrogates enable rapid decision-making; enhanced deployment efficiency via domain adaptation; supports scalability and industrial deploymentLimited details on sensor modalities and specific validation in some frameworks; industrial validation needed; integration complexity requires comprehensive infrastructure development; domain shift challenges; methodological frameworks requiring implementation and validation; standardization challenges[6,293,328,329,330,331,332]
Table 15. Overview of topology optimization methods.
Table 15. Overview of topology optimization methods.
Method CategoryPhysics and ValidationGeometry ComplexityComputational CostTransferabilityReal-Time ReadinessLimitationsReferences
SIMP-Based Methods with Overhang ConstraintsGeometric overhang angle constraints; penalty formulations; validated on 2D/simplified 3D examples2D and low-resolution 3D; limited to simplified geometriesComputationally inexpensive; simple penalty parametersNot reportedNot suitable for real-time controlTrade-offs with structural performance; stress singularities; convergence issues; lacks of thermal/mechanical coupling[333,334,335,336]
Inherent Strain Method with Residual Stress ModelingResidual stress and distortion via the inherent strain method; ex situ validation in selected studies; simplified thermal–mechanical coupling2D and moderate 3D complexity; support structure and topology co-optimizationFast simulation; reduced adjoint sensitivity cost; parallel computing frameworks accelerate optimizationLimited experimental validation across machines/materialsNot suitable for real-time control due to its iterative natureRelies on simplified surrogate models; limited accuracy for complex transient effects; high computational cost for high-fidelity coupling[337,338,339,340,341]
Layer-by-Layer Thermal Process ModelsLocal layer-wise thermal models; identifies heat concentration zones; thermal overheating constraints; limited in situ validationHigh-resolution 2D and 3D; voxel-level simulations; complex geometries supportedHigh efficiency via parallelization; custom solvers enable fast layer-wise simulationNot ReportedPotential for near-real-time feedback with further developmentLocal/layer assumptions may limit accuracy for transient global effects; experimental validation sparse[342,343,344,345]
Coupled Fictitious Physics and Multi-Constraint ModelsCouples geometric constraints (overhang, cavity) with fictitious physics; improved convergence; limited experimental validationAddresses complex geometric constraints; 2D and 3D examplesImproved convergence; computationally manageableNot reportedNot suitable for real-time controlUser-defined parameters reduce generality; simultaneous multi-constraint control remains challenging[346,347,348]
Concurrent Topology, Support, and Build Orientation OptimizationIntegrates thermal deformation, residual stress, and geometric constraints; surrogate models and homogenization; ex situ validation in few studies2D and simplified 3D cases; limited applicability to complex industrial partsHigh computational cost; surrogate models improve tractability; parallel computing usedNot reportedNot suitable for real-time controlScalability limited; interaction between support and orientation not fully captured; high runtime[341,349,350,351,352,353]
Machine Learning-Integrated MethodsML for turbulent flow modeling and multi-scale lattice optimization; derivative-aware algorithms; limited validationApplied to niche problems (heat exchangers, lattice scaffolds); not comprehensiveAccelerates specific subproblems; reduces computational burden in targeted applicationsProblem-specific; lacks general frameworks and cross-material/machine validationEarly stage; not ready for real-time controlNascent stage; data requirements; model generalization challenges; coupling with physics-based simulations underdeveloped[354,355,356]
Multi-Material and Microstructure-Aware MethodsIncorporates material heterogeneity, porosity, graded properties; iterative property updates; limited experimental validationSimplified material models; small-scale examplesNot reportedNot reportedNot suitable for real-time controlEmerging; complexity of capturing microstructural evolution; sparse experimental validation and process simulation integration[357,358,359,360,361]
Table 16. Process parameter optimization by ML algorithm type.
Table 16. Process parameter optimization by ML algorithm type.
Process ParametersBayesian OptimizationDeep Neural NetworksTraditional MLMost Effective Algorithm
Laser Power- Surface quality optimization [362]
- Hyperparameter tuning [363]
- Melt pool prediction [230]
- Property prediction [364]
- Density optimization [365]
- Multi-objective optimization [366]
Bayesian optimization—superior for continuous optimization
Scan Speed- Process parameter effects [362]
- Bead geometry prediction [363]
- Thermal field modeling [141]
- Melt pool dynamics [367]
- Relative density prediction [368]
- Surface roughness control [369]
Deep neural networks—better for complex relationships
Layer Thickness- Multi-parameter optimization [362]- Build quality prediction [370]
- Defect detection [371]
- Mechanical properties [372]
- Energy consumption [366]
Traditional ML—adequate for linear relationships
Hatch Spacing- Integrated optimization [362]- Track morphology [373]
- Process monitoring [308]
- Density control [306]
- Quality optimization [374]
Bayesian optimization—effective for spacing optimization
Energy Density- Comprehensive optimization [362]- Material property prediction [364]- High-density achievement [365]
- Process window determination [375]
Traditional ML—good for energy-based metrics
Multi-Parameter Sets- BOAT framework [362]
- Simultaneous optimization
- Hybrid approaches [370]
- Multi-modal fusion [308]
- Ensemble methods [376]
- Multi-objective optimization [366]
Bayesian optimization—best for complex multi-parameter spaces
Table 17. Build quality metrics and ML outcomes.
Table 17. Build quality metrics and ML outcomes.
Quality MetricML Approaches UsedAchieved ImprovementsPerformance MetricsKey Findings
Surface Roughness- Bayesian optimization with transfer learning [362]
- Machine learning prediction models [369]
- Multi-objective optimization [366]
- Optimized surface quality prediction
- Reduced trial-and-error approaches
- Real-time quality control
- R2 = 98.78% for surface roughness prediction [362]
- Significant parameter contribution analysis [369]
Laser power most significant (48.3% contribution) for surface quality [377]
Porosity/Density- Material-agnostic XGBoost [365]
- Gaussian process regression [378,368]
- Multiple ML algorithms [306]
- >98% relative density achievement [365]
- >99.5% density in EBM [378]
- 99.97% maximum density [379]
- Cross-material validation successful [365]
- R2 = 0.95 (ANN), 0.923 (SVM) [306]
- Mean error: 3% for porosity prediction [364]
Material-agnostic approaches show high transferability across alloy systems
Mechanical Properties- Hybrid neural networks [370]
- Deep learning prediction [364]
- Evolutionary optimization [366]
- Enhanced tensile property prediction
- Improved strength-to-weight ratios
- Optimized mechanical performance
- Good accuracy with small training data [370]
- Mean error: 0.2% for hardness [364]
- 85% correlation coefficient [366]
Microstructure inclusion significantly improves prediction accuracy
Dimensional Accuracy- Process parameter optimization [372]
- Multi-objective approaches [374]
- Predictive modeling frameworks [380]
- Reduced dimensional deviations
- Improved geometric fidelity
- Enhanced manufacturability
- 10% error in predictive models [374]
- High accuracy in manufacturability prediction [380]
Systematic parameter optimization reduces dimensional errors significantly
Defect Detection- Convolutional neural networks [371]
- Unsupervised learning [381]
- Multi-signal fusion [308]
- 98.63% defect detection accuracy [371]
- Real-time monitoring capability
- Automated quality assurance
- Minimum defect size: 0.54 mm [371]
- 74–99% accuracy for different defect sizes [308]
Feature-level fusion outperforms individual signal-based models
Process Stability- Real-time monitoring [382]
- Acoustic emission analysis [382]
- Domain adaptation [383]
- Enhanced process reliability
- Reduced build failures
- Improved consistency
- 74.39% detection accuracy with LSTM [382]
- Successful domain transfer [383]
Multi-modal sensing improves process stability monitoring
Note: ML methods listed here are detailed in Section 5.3. This table focuses on quality metric outcomes and performance indicators.
Table 18. Major limitations and open challenges in PBF modeling approaches.
Table 18. Major limitations and open challenges in PBF modeling approaches.
Area of LimitationLimitationPapers
Computational Cost and EfficiencyMany thermal and mechanical models for PBF are computationally intensive, limiting their applicability for real-time control or large-scale simulations. This constraint reduces external validity for industrial-scale applications and hinders iterative process optimization.[2,8,10,314,410,411,412,413,414]
Limited Multi-scale and Multiphysics IntegrationExisting models often focus on isolated scales or physics, lacking comprehensive coupling of thermal, mechanical, and microstructural phenomena. This methodological constraint weakens the predictive accuracy of process–structure–property relationships.[8,9,103,105,415]
Insufficient Experimental ValidationSeveral modeling approaches rely heavily on simulations with limited experimental calibration or validation, undermining the reliability and generalizability of those findings across different materials and process conditions.[9,111,122,412,416]
Simplifying Assumptions in Thermal ModelingMany thermal models assume linearity, flat molten surfaces, or neglect temperature-dependent material properties, which restricts their quantitative accuracy and external validity in capturing real PBF thermal dynamics.[194,417]
Narrow Focus on Specific Materials or GeometriesResearch often concentrates on particular alloys (e.g., Ti-6Al-4V, IN718) or simple geometries, limiting the applicability of models to diverse materials and complex part designs, thus affecting external validity.[412,417,418,419,420]
Data Limitations for Machine Learning ModelsData-driven and physics-informed neural network models depend on the availability and quality of training data, which is often scarce or expensive to generate, constraining model generalizability and robustness.[141,301,410,413,414]
Process Parameter Variability and Control ChallengesVariability in process parameters and lack of adaptive control strategies in models limit their ability to predict and mitigate defects dynamically, reducing practical applicability for quality assurance.[314,411,421,422]
Limited Comparative Studies Across AM TechniquesFew studies comprehensively compare PBF modeling approaches with those of other AM techniques, such as SLA and FDM, restricting broader contextual understanding and cross-technology insights.[103,423,424,425]
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

Łach, Ł.; Svyetlichnyy, D. Advanced Numerical Modeling of Powder Bed Fusion: From Physics-Based Simulations to AI-Augmented Digital Twins. Materials 2026, 19, 426. https://doi.org/10.3390/ma19020426

AMA Style

Łach Ł, Svyetlichnyy D. Advanced Numerical Modeling of Powder Bed Fusion: From Physics-Based Simulations to AI-Augmented Digital Twins. Materials. 2026; 19(2):426. https://doi.org/10.3390/ma19020426

Chicago/Turabian Style

Łach, Łukasz, and Dmytro Svyetlichnyy. 2026. "Advanced Numerical Modeling of Powder Bed Fusion: From Physics-Based Simulations to AI-Augmented Digital Twins" Materials 19, no. 2: 426. https://doi.org/10.3390/ma19020426

APA Style

Łach, Ł., & Svyetlichnyy, D. (2026). Advanced Numerical Modeling of Powder Bed Fusion: From Physics-Based Simulations to AI-Augmented Digital Twins. Materials, 19(2), 426. https://doi.org/10.3390/ma19020426

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Article metric data becomes available approximately 24 hours after publication online.
Back to TopTop