Next Article in Journal
A Study on Metallurgical Artifacts Excavated from Luojiaba Site H235 in the Eastern Sichuan Region during the Eastern Han Dynasty
Previous Article in Journal
A Kinetic Model for Oxide–Carbonitride Inclusion Heterogeneous Nucleation and Precipitation during Superalloy Solidification
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Laser Cutting of Titanium Alloy Plates: A Review of Processing, Microstructure, and Mechanical Properties

1
Aviation Maintenance NCO Academy, Air Force Engineering University, Xinyang 464001, China
2
School of Mechanics and Safety Engineering, Zhengzhou University, Zhengzhou 450001, China
3
School of Materials Science and Engineering, Guilin University of Electronic Technology, Guilin 541004, China
*
Author to whom correspondence should be addressed.
Metals 2024, 14(10), 1152; https://doi.org/10.3390/met14101152
Submission received: 28 July 2024 / Revised: 25 September 2024 / Accepted: 1 October 2024 / Published: 9 October 2024

Abstract

:
The growing use of titanium alloys has led to the gradual replacement of traditional processing methods by laser cutting technology, making it the preferred method for processing titanium alloy plates due to its high efficiency, precision, and adaptability. In this review, the characteristics of laser cutting technology and its application in titanium alloy plate processing are summarized, outlining several aspects of the cutting process, microstructure, and mechanical properties of the material after cutting, along with simulation predictions. Previous research categorized laser-cutting input parameters into beam parameters and process parameters, with the commonly used parameters being the laser power, cutting speed, and gas pressure. Various parameter combinations can achieve different cutting qualities, and seven indices can be used to evaluate the cutting process, with the surface roughness and slit width serving as the most common indices. Different auxiliary gases have shown a significant impact on the laser cutting quality, with commonly used gases consisting of nitrogen, argon, and air. Argon-assisted cutting generally results in better surface quality. Due to the rapid temperature change, the titanium alloy microstructure will undergo a non-diffusive martensitic phase transformation during laser cutting, producing a heat-affected zone. Experimental studies and simulations of the mechanical properties have shown that the occurrence of a martensitic phase transformation increases the hardness and residual tensile stress of the material, which reduces the fatigue strength and static tensile properties. In addition, studies have found that the more streaks appear on the cut surface, the lower the fatigue strength is, with fatigue cracks arising from the stripes. Hence, the established analytical solution model and three-dimensional finite element model can effectively predict the temperature distribution and residual stress during the cutting process. This can provide a better understanding of the high residual stress characteristics of the cutting edge and the stripe formation mechanism, allowing researchers to better explore the mechanism of laser cutting.

1. Introduction

Titanium and titanium alloys are important metallic materials that were developed in the 20th century. Titanium alloys have been extensively utilized in the automotive, chemical, and aerospace industries due to their high specific strengths, low densities, and high fatigue strengths [1,2]. Research on titanium alloys originated from the development of the aviation industry, with the demand in the aerospace sector constituting approximately 40% of the global demand. These alloys are used in applications including engine components, fasteners, and landing gear [3,4]. In the chemical industry, titanium alloys serve as excellent corrosion-resistant materials and are employed as structural materials in petroleum refining, including reactors. In the automotive sector, the use of titanium alloys contributes to weight reductions in vehicles, thereby reducing fuel consumption and diminishing engine noise and vibration. In the medical field, titanium and its alloys are highly desirable biomaterials due to their non-toxicity, high strength, low weight, and excellent biocompatibility, commonly used for medical implants and surgical instruments. With the advancement of technology, the annual production of titanium alloys has increased year by year, and widespread applications across various industries have been found. However, the low thermal conductivities and the tendency for adhesion of titanium alloys result in relatively poor mechanical cutting performances, leading to high application costs that have significantly impeded the widespread use of titanium alloys [5]. Laser cutting involves using a laser beam to cut materials by directing a high-energy-density laser beam onto the surface of a workpiece. The material absorbs energy, causing a rapid temperature increase at the focal point, resulting in melting and vaporization. High-pressure airflow is then utilized to remove the molten material from the cutting seam, thereby achieving cutting [6,7].
In 2008, Dubey and Yadava reviewed the experimental studies on the effect of process parameters on the machining performance of Nd:YAG lasers and summarized the influence of process parameters on cutting pure titanium and Ti–6Al–4V titanium alloys [8]. In 2014, Pedram et al. reviewed the modeling and simulation of laser beam machining processes, discussing the three stages of material removal during laser beam machining: melting, evaporation, and chemical degradation, and the effect of these processes on the quality of the process [9]. In 2022, Naresh and Khatak reviewed the application of CO2 laser cutting technology in metal plates, thermoplastic polymers, polypropylene (PP), etc., mainly focusing on laser cutting parameters, which can obtain better cutting quality and improve material removal rates [10]. In 2023, Pramanik et al. reviewed the laser beam processing mechanism, surface formation mechanism, heat affected zone, taper formation, and dimensional deviations of titanium alloys, and noted that laser beam machining can initiate non-diffusive phase transitions in titanium alloys, slightly altering the mechanical properties of their surfaces, in addition to optimizing the parameters [11]. In 2024, Elaziem et al. comprehensively reviewed the application of titanium-based biomaterials in the field of orthopedic implants and summarized the application of laser technology in the surface modification of titanium alloys, including laser surface alloying and laser cladding, which have improved the thermal stability, corrosion resistance, and hardness of titanium alloys [12]. In 2024, Alsaadawy et al. reviewed the effects of laser cutting parameters on the surface and kerf quality of metals (titanium alloys, steel alloys, and aluminum alloys), described the advantages of laser cutting compared to other processing methods, summarized the cutting performance parameters in detail through charts, and discussed the effects of parameters on cutting quality [13]. This study not only included the analysis and prediction of process parameters and modeling, but also placed special emphasis on the relationship between the microstructure and mechanical properties of materials during laser cutting, systematically describing the research findings on laser cutting titanium alloys.
This review was conducted from three perspectives: experiments, result analysis, and numerical simulation. As an emerging technology, laser cutting is widely utilized, and its advantages and classifications need to be clarified. When cutting titanium alloys, the selection of suitable process parameters can yield better cutting quality. Additionally, due to the influence of the laser, the microstructure of titanium alloys changes, which simultaneously affects mechanical properties such as strength, hardness, and stress. Therefore, the effects of laser cutting on alloys were summarized from the perspectives of microstructure and mechanical properties. Finally, the review summarized the analytical solution models and finite element models established by researchers, including the temperature distribution and residual stresses during the cutting process, as well as the cutting results. These contents are summarized as a comprehensive roadmap, as shown in Figure 1. This review aims to present readers with a more comprehensive research perspective.

1.1. Advantages and Classification of Laser Cutting

Compared to traditional methods such as machining [14], wire-cut electric discharge machining (WEDM; 5-axis ePULSE-40, Electronica Machine Tool Ltd., Kolkata, India) [15,16], and short electric arc milling (SEAM, DHX33A3000/28FS SEAM, Sun Hung, Taiwan) [17], laser cutting technology offers several advantages, including a small heat-affected zone in the processed material, minimal thermal deformation of the workpiece, easy laser beam guiding, and high processing productivity [18]. Nevertheless, the high costs of machines and their operation, coupled with the significant impact of input factors on the cutting quality, necessitate the optimization of the process parameters and the monitoring of the material surface microstructural changes to achieve high-quality cuts [19]. Additionally, laser cutting offers faster cutting speeds and smoother cutting surfaces [20], eliminating the need for metal hardening before cutting [21]; this is in contrast to non-traditional methods such as laser machining (LM, MFSC-4000W laser, Dechuan Laser Technology Co., Ltd., Dongguan, China) [10], abrasive water jet machining (AWJM, OMAX2626, OMAX Corporation, Kent, WA, USA) [22], ultrasonic machining (USM, 5e-axis DMU 50 evo linear, Deckel Maho Gildemeister, Bielefeld, Germany) [23,24], and abrasive water-jet-assisted laser machining (AWJALM, YLR-300-MM-WC-Y11, IPG, Su Zhou, China) [25].
Laser cutting can be divided into fusion cutting, oxidative cutting, vaporization cutting, and controlled fracture techniques. When the high-power laser is focused on the material through a lens, it can be instantaneously injected into the material, causing the material to change through melting, chemical bond breaking, and evaporation. The high-pressure auxiliary gas will then eject the evaporated and melted material from the cutting zone to complete the precise cutting of the material [26].
(1)
Laser fusion cutting
Fusion cutting refers to the action of a laser beam on a metal material during the laser cutting process, which raises the temperature to the fusion point of the metal and causes it to melt, releasing gases that cannot easily burn with high energy, such as Ar, N2, and He. High-energy compressed air is then pushed to the metal surface as a promoter to achieve efficient cutting. Laser fusion technology can be used to cut hard metals that are difficult to oxidize. For example, the precise cutting of stainless steel, titanium, aluminum, and their alloys only consumes one-tenth of the original laser energy.
(2)
Laser oxidative cutting
Laser oxidative cutting is a special melting cutting method that uses oxygen as the auxiliary gas type. During the cutting process, the high-energy laser beam heats the material to the ignition temperature without melting it. After the ignition point is reached, oxygen will violently react with the material, releasing a significant amount of oxidation reaction heat. Subsequently, the material will begin to melt to form a molten metal accompanied by a large number of oxides. These melts will be expelled from the kerf surface via auxiliary gas flow, completing the cutting process.
(3)
Laser vaporization cutting
When intense laser radiation hits a workpiece, it rapidly heats the material to the boiling state, turning some of the material into a liquid, while the other portion is discharged from the cutting slit with the flow of hot gas, producing 10 times more laser energy than the effect of fusion cutting. Laser vaporization cutting is a specialized cutting method for materials that cannot be fused, such as wood, rubber, or certain metals. The laser emits laser light on the surface of an object, during which one part is refracted and the other part is absorbed. As the temperature of an object increases, its refractive index decreases. When the energy of a laser beam is released, it heats the material until the surface temperature reaches the boiling point, causing steam to be ejected from the workpiece at the speed of sound.
(4)
Laser cutting through controlled fracture
In laser-controlled fracture cutting, cracking results from the material absorbing laser energy upon heating, and the stress near the laser focus will change due to the high temperature. As the laser beam passes through, the relaxation of compressive stress creates localized residual tensile stress [27]. Under the action of stress, the crack will continue to extend in the direction of laser movement, and the material will break and separate in the path of the laser beam. This cutting method is often used for brittle materials such as ceramics [28].

1.2. Classification of Lasers

A laser, as a laser-generating device, is mainly composed of the pumping source, gain media, and optical resonant cavity, as shown in Figure 2. The pumping source supplies energy to the laser. The gain medium absorbs the energy from the pumping source, causing population inversion and providing the medium with laser amplification through stimulated emission. The optical resonant cavity can provide the space required for energy excitation, with light repeatedly reflected, enhancing the stimulated emission and ultimately releasing coherent laser light [29].
Laser cutting can be divided into three categories based on the different types of lasers, namely CO2 laser cutting, pulsed Nd:YAG laser cutting, and fiber laser cutting, as shown in Figure 3. Lasers can be divided into solid-state lasers, semiconductor lasers, gas lasers, and fiber lasers.
Solid-state lasers are based on crystals or glass as the matrix material and doped with rare-earth metal ions, transition metal ions, and other activating ions. These systems offer advantages such as high laser powers, stable energy outputs, and compact sizes. Currently, Nd:YAG lasers are the most commonly used. Semiconductor lasers mainly use semiconductor materials as the working medium, providing high energy efficiency and long service life. Another laser type is gas lasers, such as CO2 lasers and argon lasers. CO2 lasers have the characteristics of large power ranges, high energy conversion rates, and good beam quality. Commonly used argon laser wavelengths include 458 nm and 488 nm for blue light and 514.5 nm for green light. Argon lasers are frequently applied in the medical field, such as using blue light for dental restoration [30] and the treatment of infant jaundice [31], and for ophthalmic procedures utilizing blue-green light [32]. In the industrial manufacturing sector, blue and green lasers are used for processing highly reflective metals, resulting in a smaller heat-affected zone and reduced deformation. However, compared to fiber lasers and CO2 lasers, argon lasers have lower output power, lower conversion efficiency, and higher operating and maintenance costs, making them less common in industrial applications [33]. Fiber lasers are a type of laser based on optical fibers, utilizing fibers infused with various rare-earth metal ions as the working medium. According to their time-domain characteristics, these systems can be divided into continuous wave fiber lasers and pulsed fiber lasers. Due to their advantages of high efficiencies, low thresholds, excellent beam quality, and cost-effectiveness, lasers have gradually become the mainstream choice in fields such as industrial processing, medical treatment, military defense, aerospace, automobile manufacturing, shipbuilding, and scientific research [34].
Figure 3. Laser diagram: (a) CO2 laser cutting [35], (b) fiber laser cutting [36], and (c) pulsed Nd:YAG laser cutting [37].
Figure 3. Laser diagram: (a) CO2 laser cutting [35], (b) fiber laser cutting [36], and (c) pulsed Nd:YAG laser cutting [37].
Metals 14 01152 g003

2. Laser Cutting Parameters

The laser cutting process involves changes in energy, temperature fields, thermochemical reactions, and fluid mechanics, including aerodynamics. The input parameters of laser cutting can be divided into two parts, namely beam parameters and process parameters, as shown in Figure 4 and Table 1. In actual processing, the relevant properties and parameters of the material and equipment are definitive. To improve the cutting quality and efficiency, the effects of different processing parameters on the results should be considered.
The use of different laser parameters results in different cutting qualities. Currently, there is no uniform standard for determining laser cutting quality, with differences in evaluation criteria among different laser equipment manufacturers worldwide. According to the international standards for thermally cut products and various studies, commonly used criteria for cutting quality can be divided into seven types: surface roughness (Ra) [38], kerf width (KW) [39], kerf taper (KT) [36], heat-affected zone (HAZ) [40], kerf deviation (KD) [41], material removal rate (MRR) [42], and hanging slag (S) [39], as shown in Figure 5.
The formulas for three parameters, KT, KD, and MRR, are as follows:
KT = UKW LKW 2 π T 180 ,
KD = Avg . Max . TKW Avg . Min . TKW ,
MRR ( kg / min ) = W 1 W 2 t ,
where UKW is the upper kerf width, LKW is the lower kerf width, T is the thickness of the workpiece, Avg. Max. TKW is the average maximum taper kerf width, Avg. Min. TKW is the average minimum taper kerf width, W1 is the mass of the workpiece before cutting, W2 is the mass of the workpiece after cutting, and t is the cutting time. Based on several previous studies, Ra and KW are the most commonly used parameters [13].
Understanding the effect of machining parameters on the cutting quality helps to achieve accurate cuts. Some researchers have used different test methods for different evaluation quality parameters, including the Taguchi method (TM), full factorial design (FFD), responsive design (RD), the multiple regression method (MRM), single-factor experiments (SFEs), neural network models (NNMs), and a genetic algorithm (GA). The influences of different cutting parameters have been assessed, and optimal parameter combinations have been identified, as shown in Table 2, offering a reference for practical applications of laser cutting.
Alsaadawy et al. [51] studied the laser cutting of a Ti-6Al-4V (TC4) titanium alloy sheet with a 4 mm thickness and investigated the effects of the laser cutting power, cutting speed, and gas pressure on the surface roughness, kerf width, kerf taper, and slag hanging height. In the laser cutting process, commonly used auxiliary gases include air, oxygen (O2), nitrogen (N2), and argon (Ar). O2-assisted cutting will generate hard and brittle oxides on the cutting surface, and even an uncontrollable burning phenomenon. If the cutting plate is thicker and the quality of the cutting surface is not considered, O2 can be used to assist [46]. Air- and N2-assisted laser cutting produces a thin layer of hard, brittle oxides and nitrides on the surface, leading to the formation of micro-cracks and a reduction in surface quality. In comparing nitrogen- and argon-assisted laser cutting of commercially pure titanium and Ti-6Al-4V titanium alloys, the results show that the surface is pale yellow and TiN is formed during nitrogen cutting. Compared with the substrate, the Vickers microhardness of the remelted zone is increased by about two times on the surface of pure titanium and three times on the surface of TC4 titanium alloy. In argon cutting, the surface is usually a bright white metallic color, and the cut edges are more regular and flatter [52,53]. By diluting nitrogen with argon, the properties of the nitride layer in surface roughness and crack can be improved, while the more argon content, the lower the microhardness of the titanium alloys [54]. In contrast, Ar-assisted cutting can produce a better cutting quality, a smaller HAZ, and minimal slag hanging [46], and the use of Ar-assisted cutting can produce a smaller HAZ and low slag hanging [53,55].

3. Microstructural Evolution Induced by Laser Cutting of Titanium Alloys

Titanium has two crystal structures, a closely packed hexagonal αphase below 882 °C and a body-centered cubic β phase at higher temperatures [56]. These structures serve as the basis for the classification of titanium alloys as α, near-α, α-β, and β [57]. Phase transitions occur in nearly all manufacturing processes involving rapid temperature changes, such as welding, forging, machining, and additive manufacturing [58]. These transitions induce significant microstructural changes and high residual stresses, thereby reducing the mechanical properties of the material [59]. Burgers [60] first obtained a certain orientation relation for the transition of hexagonal close-packed α phase to body-centered cubic β phase in zirconium alloys, which was later confirmed in titanium alloys. Therefore, this relationship is also known as the Burgers orientation relationship, defined as {110} β//{0001} α, <111>β//<11-20>α.
Shanjin et al. [46] used an Nd:YAG pulsed laser to cut TC1 titanium alloy plates with a thickness of 1 mm. The effects of three auxiliary gases—compressed air, N2, and Ar—on the thickness of the HAZ, surface topography, and corrosion resistance were studied. The researchers found that a non-diffusion phase transition occurred in the HAZ, and the acicular martensite α′ structure formed. The effect of the pulse energy on the thickness of the HAZ is shown in Figure 6. Pulsed energies of 1.5 and 2.5 J produced HAZs with thicknesses of about 10 μm. Correspondingly, thicker HAZs were observed below and above 1.5 and 2.5 J. Due to the narrow kerf width at lower energies, the cleaning ability of the auxiliary gas was limited. This allowed additional molten material to adhere to the cutting surface, thus transferring more heat to the workpiece. Therefore, pulse energies below 1.5 J could produce a thicker HAZ layer. As the pulse energy increased above 2.5 J, more energy accumulated on the material, transferring more energy to the material and producing a thicker HAZ [46]. When the pulse frequency and cutting speed were maintained constant, at lower pulse widths, the kerf taper (KT) decreased with the increase in assist gas pressure. This was primarily due to the inverse relationship between pulse width and peak power, where an increase in peak power led to an increase in the laser penetration depth, resulting in a larger bottom kerf. Consequently, the difference in the kerf widths between the top and bottom decreased, leading to a reduction in KT. At higher pulse widths, KT initially decreased and then increased with the augmentation of assist gas pressure. This was mainly because the decrease in peak power caused the melt zone to approach the top, and the deposition of molten material on the sides and bottom reduced as the gas pressure increased, diminishing the difference in kerf widths. However, as the pulse width continued to increase, the amount of molten material produced at the top increased, causing the top kerf to widen, which in turn led to an increase in KT [5].
Scintilla et al. [61] used a fiber laser to cut a TC4 titanium alloy sheet with a thickness of 1 mm, employing Ar as the cutting auxiliary gas. They studied the effects of different cutting parameters on the cutting surface quality, and the results showed that different microstructures appeared in the TC4 during laser cutting. The microstructure of the matrix consisted of intergranular β (black) and equiaxed α (white) phases. The microstructure of the HAZ consisted of an acicular α phase, primary α phase, martensitic α′ phase, and β phase. The microstructure of the remelting zone (RL) was dominated by a β-transformation into the martensite α′ phase, as shown in Figure 7. With an increase in the cutting speed, the thicknesses of the HAZ and RL increased, primarily due to reduced heat input and the cooling effect of the auxiliary gas pressure, which led to a decrease in the temperature. However, the increased molten material that adhered to the section acted as an additional heat source, thereby increasing the RL and HAZ widths.
Aoud et al. [62] used a CO2 laser to cut a TC4 titanium alloy sheet with a thickness of 3 mm, and N2 was selected as the auxiliary gas. The microstructure was observed by SEM. The researchers found that the surface of the cut titanium alloy could be divided into three regions, namely the top, middle, and bottom, as shown in Figure 8a. The top area was the initial region of laser cutting where the surface material started to melt at high temperatures. The acicular martensite phase was observed in the middle area. The main reason was that when the laser beam passed through, the surface of the plate rapidly cooled due to the action of the auxiliary gas and heat conduction, resulting in a martensitic transformation, as shown in Figure 8b. These results were consistent with the findings obtained by Shanjin et al. [46] on the pulsed laser cutting of the titanium alloy. As shown in Figure 9, microcracks and voids formed on the TC4 cut surface. Microcracks are mainly due to thermal stresses generated during laser cutting, and increasing the cutting speed reduces the number and size of microcracks [63].
Parmar et al. [64] used argon as an auxiliary gas to cut 3-mm-thick TC4 using a fiber laser and confirmed that the TC4 titanium alloy could be largely converted into oxides after cutting with X-ray photoelectron spectroscopy (XPS, SPECS, Berlin, Germany), X-ray diffraction (XRD, Ultima-IV, Rigaku, Woodlands, TX, USA), and energy-dispersive X-ray spectroscopy (EDX, Swift ED3000, Oxford Instruments Analytical Limited, Wycombe, UK). The XPS analysis shown in Figure 10a indicated that Ti–6Al–4V was transformed into various oxides of its constituent metals, where TiO2 was enhanced and transformed into Ti2O3. The XRD analysis in Figure 10b revealed that Ti, TiN, and TiO2 were transformed into Ti2O3, and the EDX analysis in Figure 10c showed that the O2 content increased by 19%. Kalyanasundaram et al. [65] proposed that the lattice spacing of the oxides was larger than that of the matrix material and hypothesized that the oxides could expand and help the molten metal penetrate and separate the substrate [66,67].
Electron backscatter diffraction (EBSD) orientation difference analysis serves as an important technical method for plasticity characterization, and it is widely used for the failure analysis of polycrystalline alloys under various operating conditions, such as tension, impulse, creep [68,69], creep-fatigue [70,71], and fatigue crack growth [72]. Accumulated plasticity in polycrystalline alloys can induce dislocation multiplication and lattice curvature within the grains, which can be measured by the EBSD local mismatched angle (e.g., kernel average misorientation (KAM) and grain reference orientation deviation (GROD)) [73]. The KAM method is based on the measurement of the core region, which can be used to explain the local strain distributions of crystalline materials. This method is especially suitable for illustrating the strain distribution at the grain and phase boundaries of crystalline materials after deformation, and it can be used to calculate the geometrically necessary dislocation (GND) density. The GROD method is based on grain measurement, with a higher KAM value indicating a higher geometrically necessary dislocation density inside the material.
Sun et al. established a quasi-in situ EBSD observation method to assess the microstructures and damage evolution of titanium alloys under cyclic loading [74]. The fatigue specimen axis of the Ti–6Al–4V ELI alloy was along the rolling direction, and its microstructure is shown in Figure 11, which consisted of equiaxed α-phase and lamellar β-transformation structures. Figure 11d presents the KAM diagram.
In a word, the microstructure of the cut surface after laser cutting consists of three typical regions: remelting zone, heat-affected zone, and matrix. Combined with the scanning electron microscope (SEM) image in Figure 12, it is evident that the matrix of the TC4 sheet consists of primary equiaxed α phase and intergranular β phase. The remelted zone consists of fine acicular martensite, with no β phase detected. The main reason is that the high-energy-density laser beam makes the temperature of the cutting zone rise rapidly, reaching and exceeding the phase transition temperature of the β phase, and then cools rapidly from above the β-phase transition temperature under the effect of auxiliary gases and heat conduction with the matrix and undergoes a non-diffusive martensitic phase transition into the acicular α’ phase [46]. The typical microstructure in the heat-affected zone includes the primary equiaxed α phase, black β phase, and a small amount of acicular α’ phase. The maximum temperature and cooling rate in this area are smaller than those in the remelting zone, and the martensitic transformation is less; the pinch α’ phase is finer.

4. Mechanical Properties of Cut Titanium Alloy

Laser cutting of titanium alloys can lead to changes in the surface properties of the processing area, mainly due to the high energy density of the laser beam acting on the surface of the workpiece. This causes the temperature to increase, the heat to diffuse along the edge of the kerf under the action of heat conduction, and rapid cooling to occur after the laser beam passes. Due to temperature changes, the microstructure of the titanium alloy changes, resulting in the formation of an HAZ. The emergence of this zone changes the hardness, plasticity, and other mechanical properties of the material.
Wu et al. [75] investigated the effects of laser cutting on the fatigue performance of TC4 titanium alloys by setting up three groups of specimens containing an HAZ, after removing the HAZ, and with no HAZ. The results showed that (1) the HAZ constituted a fatigue source zone, exhibiting brittle fracture characteristics. (2) The fatigue lives of the laser-cut specimens were significantly reduced by the presence of the HAZ, which were an order of magnitude lower than those of specimens without the HAZ. Hou et al. [76] set up three sets of specimens in the same manner and found that the presence of an HAZ affected the fatigue life of the material, and the fatigue life of the laser-cutting aircraft skin material was reduced by 20% to 25%. Germain et al. [77] analyzed the effect of laser-assisted processing on the fatigue lives of TC4 titanium alloys from a microstructural perspective. The researchers found that the fatigue life decreased from 580 to 550 MPa due to the formation of martensite in the HAZ. With increasing laser power, the residual stress increased and tended toward tensile stress or normal stress.
Ullah et al. [78] obtained the static tensile and fatigue properties of 0.5- and 1-mm-thick AA2B06–T4 alloy plates cut with different parameters. The results showed that the plate has higher tensile strength values at a cutting speed of 9 m/min, and the 1.5 MPa auxiliary gas pressure produces a better ultimate tensile strength (UTS), as shown in Figure 13a. For both thicknesses, UTS increases with increasing auxiliary gas pressure and decreasing severance energy (SE), as shown in Figure 13b,c. The combination of low SE (6.67–13.34 J/mm2) and high auxiliary gas pressure (1.5 MPa) results in better static mechanical properties and fatigue strength as shown in Figure 13d.
The failure mode in the fatigue test involved the generation, growth, and catastrophic failure of cracks. Surface irregularities such as indents, notches, stripes, or splines often serve as the sources of fatigue cracks. Figure 14 shows the SEM image of the fracture surface, which shows that a crack originated from the fringe, and the fatigue source area produced a depression structure, which is characteristic of ductile fracture. The V-shaped markings in the direction of fatigue extension indicate that catastrophic failure occurred as a result. Zhang et al. [79] investigated the deformation behavior of the Ti-17 titanium alloy under uniaxial stretching via in situ SEM. The results showed that slip was the main deformation mechanism of the primary α phase in the Ti-17 tensile test.
The hardness of the HAZ produced during laser cutting of the material was significantly higher than that of the base material, and in some instances, it was even twice as hard [80,81], where the final performance of the HAZ was mainly related to the martensite near the edge of the slit. Bursi et al. [82] analyzed the microhardness, residual stress, and high-cycle fatigue performance of laser-cutting metallic materials and found that the hardness of the HAZ was significantly improved. XRD revealed the visible peaks of martensite and the γ phase, as shown in Figure 15. The increase in the number of section stripes was identified as a key feature of the fatigue properties. The residual tensile stress in the material increased the external stress in the structural features, resulting in a decrease in the fatigue strength. Slow quenching of the cutting edge after the thermal cutting operation significantly reduced the residual stress.
Digital image correlation (DIC) technology is a non-contact optical measurement method, based on the principle of human binocular stereo imaging, mainly using the scattering pattern on the object and two camera-type image acquisition instruments. The image is compared before and after object deformation to obtain real-time strain field results [83]. DIC has been widely used to examine the formability and local deformation behavior during tensile testing [84]. Specifically, it has been used to investigate the local tensile properties of welds obtained by friction stir welding [85,86,87] and laser welding [88,89] of different metal materials. This technology has also been applied to the laser welding of Ti–6Al–4V [90,91]. Scintilla et al. [92] investigated the mechanical properties of welded joints using DIC technology after cutting and welding 1-mm-thick TC4 plates. Data acquisition was carried out using a LabVIEW-specific virtual instrument, and the DIC system used the ARAMIS three-dimensional (3D) optical deformation analysis system. The experimental setup is shown in Figure 16, where global strain maps were obtained from two cameras. Before the test, a pattern of scattered black spots was applied to a matte white background to prepare the tensile specimen. This pattern was used to correlate the images at different stages in the tensile process and calculate the strain data for each stage. Specifically, the influence of the weld geometry and microstructural changes on the mechanical properties of the joints could be assessed.

5. Numerical Simulations of Laser Cutting

The theoretical modeling of laser cutting can be used to establish a mathematical model of the relevant characteristics of the material cutting process through the use of physical principles and energy balance equations, such as temperature distribution and residual stresses.

5.1. Heat Source Model

In numerical calculations, the laser is typically represented using a relevant heat source expression, which is why the choice of heat source expression can be directly related to the results of the numerical calculations. Commonly used heat source models include the Gaussian heat source, rotary Gauss body heat source, double-ellipsoidal heat source, and composite heat source models [93].

5.1.1. Gaussian Heat Source Model

The Gaussian heat source model [94] is spatially in the form of a normal distribution, with the laser energy concentrated within the laser-spot diameter, which is in line with the distribution of the laser energy density in actual processing. The mathematical expression of the heat flux shown in Figure 17 is as follows:
q = 3 η P π r 2 exp 3 x 2 + y 2 r 2 ,
where q is the heat flux, η is the material absorption coefficient of the laser, P is the laser power, r is the laser spot radius, and x and y are the spatial position coordinates of the laser.

5.1.2. Rotary Gaussian Body Heat Source Model

The rotary Gaussian body heat source model [96] assumes that the laser energy density is concentrated in the Gaussian rotating body during laser action, which improves the energy distribution in the depth direction and improves the calculation accuracy of the puddle. As shown in Figure 18, the mathematical expression is given by the following:
q = 9 P π h R 2 1 e 3 exp 9 x 2 + y 2 R 2 log h z ,
where P is the laser power, R is the radius of the heat source, and h is the height of the heat source.

5.1.3. Double-Ellipsoidal Heat Source Model

In the actual application process, the laser energy distribution does not consist of a 3D rotating Gaussian body. Some researchers have proposed an ellipsoidal heat source model, as shown in Figure 19. However, based on a large number of experimental observations, studies have found that the ellipsoidal model does not match the actual heat source distribution. As a result of this, some researchers proposed a double-ellipsoid heat source model [98] to model the laser energy density of the surface in different regions. The expressions are given by the following:
q f = 12 3 f f a f b c π π Q exp 3 x 2 a f 2 3 y 2 b 2 3 z 2 c 2 ,
q r = 12 3 f r a r b c π π Q exp 3 x 2 a r 2 3 y 2 b 2 3 z 2 c 2 , x < 0 ,
where the subscripts f and r denote the front and back of the semi-ellipsoid, respectively; a, b, and c indicate the long, middle, and short semi-axes of the ellipsoid, respectively; and ff, fr are the percentages of the front and back of ellipsoid energy scores, ff + fr = 2.

5.1.4. Recombination Heat Source Model

The single heat source model cannot accurately represent the actual laser energy. Therefore, a composite heat source model was proposed to establish different heat sources for various regions affected by the laser. This approach can provide a heat distribution that more closely aligns with the actual processing conditions with improved calculation accuracy. However, the selection of a combined heat source and the energy distribution coefficient must be based on empirical results and should be verified and refined in subsequent experiments.
Modest et al. first proposed the finite difference method (FDM) to model the beam machining process in 1988. Subsequent researchers gradually developed various finite element models (FEMs) to describe the characteristics of the temperature distribution, residual stress, and other factors in the material removal process, explaining the formation mechanisms affecting the kerf width and HAZ width. As shown in Figure 20, Yilbas et al. investigated the modeling of laser machining of thicker samples [100]. They developed a numerical model of laser machining using Abaqus software (https://www.3ds.com/products/simulia/abaqus, accessed on 30 September 2024) and analyzed the stresses during laser machining [101,102]. A temperature higher than the melting point indicated the transformation of the solid phase into the fluid phase. The titanium alloy remained in the liquid phase until it reached 3000 °C [103], and the molten metal started to partially vaporize from the fluid phase until it reached the vaporization temperature of 3558 °C. [104] The oxidation temperature of the titanium/titanium alloy was higher than 600 °C [105].

5.2. Laser Cutting Temperature and Stress Field Simulation

The temperature field during laser cutting consists of a nonlinear transient heat transfer problem, accompanied by phase transitions [112,113,114], where the changes in the stress field and the distributions of thermal and residual stress serve as the key parts of the laser cutting process design. Kardas et al. [115] experimented with high-pressure nitrogen-assisted laser cutting and predicted the temperature and thermal stress fields in the cutting zone using Abaqus finite element software. Morphological and elemental changes in the cut sections were analyzed by light microscopy, SEM, and EDX. The study found that the von Mises stress reached high values at the cutting edges, while in the mid-thickness position of the cut section, thermal strain during the cooling cycle was higher, which increased the stress at this location. Gu et al. [116] established a 3D transient FEM of Ni–Ti alloy powders as the raw material, observed the distribution of thermal stresses in the laser-selective melting process, and proposed a method for calculating the distribution of residual stresses to investigate the effect of residual stresses on the workpiece after cooling. Samanta [117] studied the cutting forces and residual stresses in laser-assisted processing and developed a thermo-mechanical coupled heat source model to predict the temperature, cutting forces, and residual stresses. Gaussian heat source models have often been used to predict the temperature field distributions of laser cutting [100,114,118]. Sharmlooei et al. [119] proposed an improved heat source model based on a Gaussian distribution, where the parameters of the model consisted of the laser power, cutting speed, spot size, and specimen thickness. The model was used to simulate the temperature field of the laser cutting in Abaqus, and an infrared thermal imager was used to detect the laser cutting process and the position of trace components during the test. The test results were consistent with the simulated temperature curve, and the distribution of trace components along the depth matched the temperature distribution predicted with the following equation, as shown in Figure 21:
Q = c 1 w i + w u     w i h z + c 2 q h π w i h + w u w r z 2 exp K 2 x 2 + y 2 h 2 w i h + w u w i z 2
where Q is the heat flux, c2 is a constant, q’ is the effective laser power, h is the thickness of the steel element, x and y denote the coordinates of the Gaussian point from the center of the beam, z is the coordinate of the laser along the beam from the inlet face of the element, wi and wu are the beam radii at the inlet and outlet surfaces of the laser beams, respectively, and K2 is the modulator coefficient.
Through simulation of the stress field of the temperature field in the cutting process, the study found that the temperature rose sharply in the region close to the laser heat source and then rapidly decayed as the laser heat source moved away. During the cutting process, the von Mises stress near the cutting edge achieved a higher value. However, due to changes in the temperature gradient along the cutting direction, the thermal strain generated in the cutting area changed, which reduced the von Mises stress to a local minimum. As the cutting progressed, the initially heated part of the cutting edge cooled to a low temperature, and the stress field transitioned to residual stress. The residual tensile stress at the cutting edge remained high [100], as shown in Figure 22.

5.3. Simulation of Laser Cutting Quality Evaluation Parameter

Zhai et al. [120] conducted a simulation and experimental study on the HAZ during the laser cutting of a TC4 titanium alloy. The results showed that the width of the HAZ was mainly affected by the laser power and cutting speed. Ding et al. [121] analyzed and predicted the effect of the phase transformation on the laser cutting of AISI 1045 steel based on metallo-thermomechanical coupling theory. In the study, a metallo-thermomechanically coupled model was established, and Abaqus software was used to simultaneously solve for the evolution of the phase constituent, cutting temperature, section shape, and cutting force. The experimental data matched well with the predicted values of the model. Fu et al. [122] developed a conical volumetric heat flow model to investigate the effects of each pulse-cutting process parameter on the kerf width, temperature, stress, and HAZ width of an Ni–Ti alloy, as shown in Figure 23. The moving heat flux of the pulsed laser was simulated using the Abaqus DFLUX subroutine. The study found that increasing the cutting speed reduced the kerf width, stress magnitude, and HAZ thickness, while the peak power pulses had the opposite effect. The stress was also mainly affected by the peak pulse power, and the slit was mainly affected by the average power. Yang et al. [114] established a 3D transient finite element model for the laser processing of Ti–6Al–4V alloy plates to predict the depth and width of the HAZ. This model explored the effects of the laser power, spot size, speed, and material properties on the depth and width of the HAZ, which provided a basis for optimizing and improving laser-assisted processing technology.
The formation of stripes has been shown to affect the quality of laser cutting. Laser cutting consists of a surface Fres absorption process, where the laser is converted into heat in the workpiece, melting the material. The molten material is then ejected from the area by an auxiliary gas jet. Wee et al. [123] developed a two-dimensional analytical model that focused on the effects of laser cutting parameters such as power, speed, and spot size on the cutting stripes, which predicted the energy absorption and oxidation processes at the cutting front. Pietro et al. [107] proposed a model to predict surface roughness during laser cutting by analyzing the dynamic phenomena at the cutting front, with the frequency of streak formation and the depth of the periodic structure as the output variables. Tani et al. [124] developed an analytical model of the quality, force, and energy balance, which was used to analyze the 3D geometry of the cutting front, the geometry of the hanging slag, and the temperature field. In addition, a method of interpreting the streak was proposed based on the slag evolution and the effect of the auxiliary gas pressure. Kovalev et al. [125] developed a mathematical model of viscous compressible gas supersonic flow and the experimental results of gas jet visualization in laser cutting by numerically solving the 3D Navier–Stokes equation. As shown in Figure 24, the airflow separation phenomenon at the cutting front was numerically simulated and experimentally verified, providing a model airflow structure of the conical supersonic nozzle. A mechanism was also proposed to control the airflow separation, which indicated that the airflow separation directly affected the changes in the structure of the streak, serving as one of the main reasons for the poor surface quality in laser cutting.
Kar et al. [126] developed a lumped parameter model that combined the width of the incision with various process parameters and obtained an empirical expression similar to the one proposed by Atsuta et al. [127] However, the model contained several constants that were difficult to measure. Based on this model, Li et al. [108] modified and re-established a lumped parameter model for the laser cutting depth, laser cutting parameters, and material properties. The model took into account the threshold power of the incident laser beam and modified earlier cutting models to make them suitable for a range of processes from low to high power and slow to fast cutting. Using ordinary steel as an example, the effects of the laser power, spot size, cutting speed, and other process parameters on the cutting depth were investigated. The model assumed that the temperature of the entire region reached at least the melting temperature and that kerf was generated by the auxiliary gas blowing off the melt. Therefore, this model was suitable for large-nozzle and high-pressure jet cutting. Considering various laser–material–water interaction phenomena, different depletion mechanisms, and the shear forces provided by water jets, Mullick et al. established a lumped parameter model for the water-jet-assisted underwater laser cutting of steel plates [128]. The model predicted the maximum laser cutting speed, notch width, and cutting front under different processing conditions, which were in good agreement with the experimental results.

6. Conclusions

With the increased use of titanium alloys in the aerospace field, laser cutting has gradually become widely used in the processing and manufacturing of metal materials as a non-contact advanced processing technology. In practical applications, this technique can effectively improve work efficiency and achieve a better cutting quality compared to traditional processing methods.
(1)
The quality evaluation factors of laser-cut metal sheets include surface roughness, kerf width, kerf taper, HAZ width, kerf deviation, material removal rate, and slag hanging. The parameters affecting the cutting quality can be categorized into beam parameters and process parameters, and commonly studied process parameters include laser power, cutting speed, and gas pressure. The surface roughness and kerf width have often been used to evaluate the cutting quality.
(2)
In the laser cutting process, the auxiliary gases most commonly used consist of nitrogen, argon, and air, with nitrogen being the most prevalent. Air- and N2-assisted laser cutting can produce a thin layer of hard and brittle oxides and nitrides on the surface, leading to microcracks and reduced surface quality. In contrast, argon-assisted cutting results in a better surface quality and a smaller HAZ.
(3)
Martensite is a pure metal or alloy transformed from one solid phase to another solid phase. The composition of the two phases remains unchanged before and after the transformation, and only the crystal structure changes, which is called non-diffusion martensitic transformation. In the process of laser cutting of titanium alloy, the temperature in the heat-affected zone rises rapidly due to the action of a high-energy-density laser beam, reaching and exceeding the phase transition temperature of the β phase. Then rapidly cooled from above the β-phase transition temperature under the action of heat conduction of the auxiliary gases and the substrate and undergoes a non-diffusive martensitic phase transition to the acicular α’ phase. Non-diffusive martensitic phase transformation can occur in the HAZ during the laser cutting of titanium alloys.
(4)
The modeling and simulation of laser-cut metal sheets can better describe the characteristics of the temperature distribution and residual stress during the material removal process. Simulations have revealed that the temperature will sharply increase in regions close to the laser heat source and rapidly decay as the laser heat source moves away. After the cutting is complete, the temperature will gradually decrease, with the stress field changing to residual stress. At this point, the residual tensile stress at the cutting edge will be very high.
(5)
Analytical models of the quality, force, and energy balance, as well as 3D finite element models of laser cutting with heat flow coupling, have been established. These models can explain the stripe formation mechanism, allowing researchers to explore the overall mechanisms of laser cutting.

Author Contributions

Conceptualization, data curation, methodology, and writing—original draft preparation, Y.Z.; writing—review and editing, C.W.; conceptualization, formal analysis, and validation, C.W., W.X., and S.W.; investigation and validation, W.X. and K.R.; data curation and methodology, Q.H. and X.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

The original contributions presented in the study are included in the article; further inquiries can be directed to the corresponding author.

Conflicts of Interest

The authors declare that they have no conflicts of interest.

References

  1. Hourmand, M.; Sarhan, A.A.D.; Sayuti, M.; Hamdi, M. A Comprehensive Review on Machining of Titanium Alloys. Arab. J. Sci. Eng. 2021, 46, 7087–7123. [Google Scholar] [CrossRef]
  2. Li, G.; Chandra, S.; Rahman Rashid, R.A.; Palanisamy, S.; Ding, S. Machinability of additively manufactured titanium alloys: A comprehensive review. J. Manuf. Process. 2022, 75, 72–99. [Google Scholar] [CrossRef]
  3. Huang, Z.; Qu, H.; Deng, C.; Yang, J.J.M.R. Development and application of aerial titanium and its alloys. Mater. Rep. 2011, 25, 102–107. [Google Scholar]
  4. Liu, Q.M.; Zhang, Z.H.; Liu, S.F.; Yang, H.Y. Application and development of titanium alloy in aerospace and military hardware. J. Iron. Steel. Res. Int. 2015, 29, 4–10. [Google Scholar] [CrossRef]
  5. Pandey, A.K.; Dubey, A.K. Simultaneous optimization of multiple quality characteristics in laser cutting of titanium alloy sheet. Opt. Laser. Technol. 2012, 44, 1858–1865. [Google Scholar] [CrossRef]
  6. Dubey, A.K.; Yadava, V. Laser beam machining—A review. Int. J. Mach. Tool. Manu. 2008, 48, 609–628. [Google Scholar] [CrossRef]
  7. Nagimova, A.; Perveen, A. A review on laser machining of hard to cut materials. Mater. Today. Proc. 2019, 18, 2440–2447. [Google Scholar] [CrossRef]
  8. Dubey, A.K.; Yadava, V. Experimental study of Nd:YAG laser beam machining—An overview. J. Mater. Process. Technol. 2008, 195, 15–26. [Google Scholar] [CrossRef]
  9. Parandoush, P.; Hossain, A. A review of modeling and simulation of laser beam machining. Int. J. Mach. Tools Manuf. 2014, 85, 135–145. [Google Scholar] [CrossRef]
  10. Naresh; Khatak, P. Laser cutting technique: A literature review. Mater. Today 2022, 56, 2484–2489. [Google Scholar] [CrossRef]
  11. Pramanik, A.; Basak, A.K. Laser Beam Machining of Titanium Alloy—A Review. Metals 2023, 13, 1536. [Google Scholar] [CrossRef]
  12. Abd-Elaziem, W.; Darwish, M.A.; Hamada, A.; Daoush, W.M. Titanium-Based alloys and composites for orthopedic implants Applications: A comprehensive review. Mater. Des. 2024, 241, 112850. [Google Scholar] [CrossRef]
  13. Alsaadawy, M.; Dewidar, M.; Said, A.; Maher, I.; Shehabeldeen, T.A. A comprehensive review of studying the influence of laser cutting parameters on surface and kerf quality of metals. Int. J. Adv. Manuf. Technol. 2024, 130, 1039–1074. [Google Scholar] [CrossRef]
  14. Yuan, C.G.; Pramanik, A.; Basak, A.; Prakash, C.; Shankar, S. Drilling of titanium alloy (Ti6Al4V)—A review. Mach. Sci. Technol. 2021, 25, 637–702. [Google Scholar] [CrossRef]
  15. Marelli, D.; Singh, S.; Nagari, S.; Subbiah, R.J.M.T.P. Optimisation of machining parameters of wire-cut EDM on super alloy materials–A review. Mater. Today. Proc. 2020, 26, 1021–1027. [Google Scholar] [CrossRef]
  16. Murray, J.; Zdebski, D.; Clare, A.T. Workpiece debris deposition on tool electrodes and secondary discharge phenomena in micro-EDM. J. Mater. Process. Technol. 2012, 212, 1537–1547. [Google Scholar] [CrossRef]
  17. Chen, X.; Zhou, J.; Wang, K.; Xu, Y.; Hu, G. A study on machining characteristics of nickel-based alloy with short electric arc milling. Int. J. Adv. Manuf. Technol. 2019, 105, 2935–2945. [Google Scholar] [CrossRef]
  18. Riveiro, A.; Quintero, F.; Lusquiños, F.; Del Val, J.; Comesaña, R.; Boutinguiza, M.; Pou, J. Experimental study on the CO2 laser cutting of carbon fiber reinforced plastic composite. Compos. Part A Appl. S 2012, 43, 1400–1409. [Google Scholar] [CrossRef]
  19. Thomas, D.J. Optimising laser cut-edge durability for steel structures in high stress applications. J. Constr. Steel. Res. 2016, 121, 40–49. [Google Scholar] [CrossRef]
  20. Hock, K.; Adelmann, B.; Hellmann, R. Comparative Study of Remote Fiber Laser and Water-Jet Guided Laser Cutting of Thin Metal Sheets. Phys. Proc. 2012, 39, 225–231. [Google Scholar] [CrossRef]
  21. Krajcarz, D. Comparison Metal Water Jet Cutting with Laser and Plasma Cutting. Proc. Eng. 2014, 69, 838–843. [Google Scholar] [CrossRef]
  22. Ramakrishnan, S.; Senthilkumar, V.; Singaravelu, D.L. Effect of cutting parameters on surface integrity characteristics of Ti-6Al-4V in abrasive water jet machining process. Mater. Res. Express 2019, 6, 116583. [Google Scholar] [CrossRef]
  23. Feng, Y.; Hsu, F.-C.; Lu, Y.-T.; Lin, Y.-F.; Lin, C.-T.; Lin, C.-F.; Lu, Y.-C.; Lu, X.; Liang, S.Y. Surface roughness prediction in ultrasonic vibration-assisted milling. J. Adv. Mech. Des. Syst. Manuf. 2020, 14, JAMDSM0063. [Google Scholar] [CrossRef]
  24. Jiajia, L.; Xinggang, J.; Ze, G.; Mingliang, Z.; Deyuan, Z. Investigation of the Effect of Vibration Amplitude on the Surface Integrity in High-Speed Rotary Ultrasonic Elliptical Machining for Side Milling of Ti-6Al-4V. J. Mech. Eng. 2019, 55, 215–223. [Google Scholar] [CrossRef]
  25. Weiss, L.; Tazibt, A.; Aillerie, M.; Tidu, A. Surface and microstructure modifications of Ti-6Al-4V titanium alloy cutting by a water jet/high power laser converging coupling. Mater. Res. Express 2018, 5, 016528. [Google Scholar] [CrossRef]
  26. Riveiro, A.; Quintero, F.; Boutinguiza, M.; Del Val, J.; Comesaña, R.; Lusquiños, F.; Pou, J.J.M. Laser cutting: A review on the influence of assist gas. Material 2019, 12, 157. [Google Scholar] [CrossRef] [PubMed]
  27. Nisar, S.; Sheikh, M.A.; Li, L.; Safdar, S. The effect of material thickness, laser power and cutting speed on cut path deviation in high-power diode laser chip-free cutting of glass. Opt. Laser. Technol. 2010, 42, 1022–1031. [Google Scholar] [CrossRef]
  28. Tsai, C.H.; Chen, C.J. Application of iterative path revision technique for laser cutting with controlled fracture. Opt. Lasers Eng. 2004, 41, 189–204. [Google Scholar] [CrossRef]
  29. Yuan, Y.; Gu, C.; Huang, S.; Song, L.; Fang, F. Advances on studying optical forces: Optical manipulation, optical cooling and light induced dynamics. J. Phys. D Appl. Phys. 2020, 53, 283001. [Google Scholar] [CrossRef]
  30. Kwaśna, M.; Cłapińska, P.; Piosik, Z.; Barysz, K.; Dubiec, I.; Bęben, A.; Ordyniec-Kwaśnica, I. Intraoral Applications of Lasers in the Prosthetic Rehabilitation with Fixed Partial Dentures—A Narrative Review. Dent. J. 2024, 12, 164. [Google Scholar] [CrossRef]
  31. Abdulrahman, H.J.; Mohammed, S.B. Using the Visible Blue Light of Argon Laser Technology for Treatment of Jaundice in Infants. J. Nat. Sci. Biol. Med. 2024, 15, 184. [Google Scholar] [CrossRef]
  32. Jiang, Y.; Cai, Y.; Zhang, X.; Chen, L.; Zhou, X.; Chen, Y.J. A Two-Decade Bibliometric Analysis of Laser in Ophthalmology: From Past to Present. Clin. Ophthalmol. 2024, 18, 1313–1328. [Google Scholar] [CrossRef] [PubMed]
  33. Yang, Q.; Zhang, P.; Lu, Q.; Yan, H.; Shi, H.; Yu, Z.; Sun, T.; Li, R.; Wang, Q.; Wu, Y.; et al. Application and development of blue and green laser in industrial manufacturing: A review. Opt. Laser. Technol. 2024, 170, 110202. [Google Scholar] [CrossRef]
  34. Jauregui, C.; Limpert, J.; Tünnermann, A. High-power fibre lasers. Nat. Photonics 2013, 7, 861–867. [Google Scholar] [CrossRef]
  35. Salim, A.A.; Bidin, N.; Islam, S.J.S. Low power CO2 laser modified iron/nickel alloyed pure aluminum surface: Evaluation of structural and mechanical properties. Surf. Coat. Technol. 2017, 315, 24–31. [Google Scholar] [CrossRef]
  36. Pramanik, D.; Kuar, A.; Sarkar, S.; Mitra, S.J.O.; Technology, L. Enhancement of sawing strategy of multiple surface quality characteristics in low power fiber laser micro cutting process on titanium alloy sheet. Opt. Laser Technol. 2020, 122, 105847. [Google Scholar] [CrossRef]
  37. Sharma, A.; Yadava, V.J.O. Experimental analysis of Nd-YAG laser cutting of sheet materials–A review. Opt. Laser Technol. 2018, 98, 264–280. [Google Scholar] [CrossRef]
  38. Pramanik, D.; Roy, N.; Kuar, A.; Sarkar, S.; Mitra, S.J.O. Experimental investigation of sawing approach of low power fiber laser cutting of titanium alloy using particle swarm optimization technique. Opt. Laser Technol. 2022, 147, 107613. [Google Scholar] [CrossRef]
  39. Marimuthu, S.; Dunleavey, J.; Liu, Y.; Antar, M.; Smith, B.J.O. Laser cutting of aluminium-alumina metal matrix composite. Opt. Laser Technol. 2019, 117, 251–259. [Google Scholar] [CrossRef]
  40. Joshi, P.; Sharma, A.J.P.E. Simultaneous optimization of kerf taper and heat affected zone in Nd-YAG laser cutting of Al 6061-T6 sheet using hybrid approach of grey relational analysis and fuzzy logic. Precis. Eng. 2018, 54, 302–313. [Google Scholar] [CrossRef]
  41. Tamilarasan, A.; Rajamani, D.; Esakki, B. Parametric optimisation in Nd-YAG laser cutting of thin Ti-6Al-4V super alloy sheet using evolutionary algorithms. Int. J. Mater. Prod. Technol. 2018, 57, 71–91. [Google Scholar] [CrossRef]
  42. Perveen, A.; Molardi, C.; Fornaini, C.J.M. Applications of laser welding in dentistry: A state-of-the-art review. Micromachines 2018, 9, 209. [Google Scholar] [CrossRef] [PubMed]
  43. El Aoud, B.; Boujelbene, M.; Boudjemline, A.; Bayraktar, E.; Ben Salem, S.; Elbadawi, I. Investigation of cut edge microstructure and surface roughness obtained by laser cutting of titanium alloy Ti-6Al-4V. Mater. Today 2021, 44, 2775–2780. [Google Scholar] [CrossRef]
  44. Shrivastava, P.K.; Pandey, A.K. Multi-objective optimization of cutting parameters during laser cutting of titanium alloy sheet using hybrid approach of genetic algorithm and multiple regression analysis. Mater. Today 2018, 5, 24710–24719. [Google Scholar] [CrossRef]
  45. Boudjemline, A.; Boujelbene, M.; Bayraktar, E. Surface quality of Ti-6Al-4V titanium alloy parts machined by laser cutting. Eng. Technol. Appl. Sci. Res. 2020, 10, 6062–6067. [Google Scholar] [CrossRef]
  46. Shanjin, L.; Yang, W. An investigation of pulsed laser cutting of titanium alloy sheet. Opt. Lasers Eng. 2006, 44, 1067–1077. [Google Scholar] [CrossRef]
  47. Kochergin, S.; Morgunov, Y.A.; Saushkin, B. Particularities of pulse laser cutting of thin plate titanium blanks. Procedia Eng. 2017, 206, 1161–1166. [Google Scholar] [CrossRef]
  48. Boujelbene, M.; El Aoud, B.; Bayraktar, E.; Elbadawi, I.; Chaudhry, I.; Khaliq, A.; Ayyaz, A.; Elleuch, Z. Effect of cutting conditions on surface roughness of machined parts in CO2 laser cutting of pure titanium. Mater. Today 2021, 44, 2080–2086. [Google Scholar] [CrossRef]
  49. El Aoud, B.; Boujelbene, M.; Bayraktar, E.; Ben Salem, S. Optimization of kerf quality during CO2 laser cutting of titanium alloy sheet Ti-6Al-4V and pure titanium Ti. In Mechanics of Composite, Hybrid and Multifunctional Materials, Volume 5: Proceedings of the 2018 Annual Conference on Experimental and Applied Mechanics, Greenville, SC, USA, 4–7 June 2018; Springer: Cham, Switzerland, 2019; pp. 213–219. [Google Scholar]
  50. Pandey, A.K.; Dubey, A.K. Modeling and optimization of kerf taper and surface roughness in laser cutting of titanium alloy sheet. J. Mech. Sci. Technol. 2013, 27, 2115–2124. [Google Scholar] [CrossRef]
  51. Alsaadawy, M.; Dewidar, M.; Said, A.; Maher, I.; Shehabeldeen, T. Investigation of the Effect of Laser Cutting Parameters on Surface and Kerf Quality of Thick Ti–6Al–4V Alloy Sheets. Arab. J. Sci. Eng. 2024, 1–18. [Google Scholar] [CrossRef]
  52. Almeida, I.A.; de Rossi, W.; Lima, M.S.F.; Berretta, J.R.; Nogueira, G.E.C.; Wetter, N.U.; Vieira, N.D. Optimization of titanium cutting by factorial analysis of the pulsed Nd:YAG laser parameters. J. Mater. Process. Techol. 2006, 179, 105–110. [Google Scholar] [CrossRef]
  53. Rao, B.T.; Kaul, R.; Tiwari, P.; Nath, A.K. Inert gas cutting of titanium sheet with pulsed mode CO2 laser. Opt. Lasers Eng. 2005, 43, 1330–1348. [Google Scholar] [CrossRef]
  54. Li, G.; Yao, X.; Wood, R.J.; Guo, J.; Shi, Y. Laser Surface Nitriding of Ti–6Al–4V Alloy in Nitrogen–Argon Atmospheres. Coatings 2020, 10, 1009. [Google Scholar] [CrossRef]
  55. Hou, H.; Wu, C.; Lv, R.; Li, Q. Study on Laser Drilling and Cutting Technology of TC4 Titanium Alloy. Appl. Lasers 2019, 39, 614–620. [Google Scholar]
  56. Salvador, C.A.F.; Maia, E.L.; Costa, F.H.; Escobar, J.D.; Oliveira, J.P. A compilation of experimental data on the mechanical properties and microstructural features of Ti-alloys. Sci. Data 2022, 9, 188. [Google Scholar] [CrossRef] [PubMed]
  57. de Lima, M.S.F. Phase Transformations during Laser Processing of Aerospace Metallic Materials. Adv. Mater. Res. 2016, 1135, 179–201. [Google Scholar] [CrossRef]
  58. Thijs, L.; Verhaeghe, F.; Craeghs, T.; Humbeeck, J.V.; Kruth, J.P. A study of the microstructural evolution during selective laser melting of Ti–6Al–4V. Acta Mater. 2010, 58, 3303–3312. [Google Scholar] [CrossRef]
  59. Pan, Z.; Liang, S.Y.; Garmestani, H. Finite element simulation of residual stress in machining of Ti-6Al-4V with a microstructural consideration. Proc. Inst. Mech. Eng. Part B J. Eng. Manuf. 2019, 233, 1103–1111. [Google Scholar] [CrossRef]
  60. Burgers, W.G. On the process of transition of the cubic-body-centered modification into the hexagonal-close-packed modification of zirconium. Physica 1934, 1, 561–586. [Google Scholar] [CrossRef]
  61. Scintilla, L.D.; Sorgente, D.; Tricarico, L. Experimental Investigation on Fiber Laser Cutting of Ti6Al4V Thin Sheet. Adv. Mater. Res. 2011, 264–265, 1281–1286. [Google Scholar] [CrossRef]
  62. Aoud, B.E.; Boujelbene, M.; Bayraktar, E.; Salem, S.B.; Boudjemline, A. Experimental Comparison of the Microstructure and Surface Roughness in CO2 Laser Cutting of the Titanium Alloy Ti-6Al-4V and the Pure Titanium Ti. In Proceedings of the SEM International Congress & Exposition on Experimental and Applied Mechanics, Reno, NV, USA, 3–6 June 2019. [Google Scholar]
  63. Yang, C.; Fu, Y.; Liu, X.-H.; Yin, X.-H.; Liu, K.; Li, C.-B.; Zheng, X.; Yang, B.-H. Micro-Crack Formation in Laser Structuring of Titanium Alloys for Orthopedic Applications. In Proceedings of the ASME 2022 International Mechanical Engineering Congress and Exposition, Columbus, OH, USA, 29 October–2 November 2022. [Google Scholar]
  64. Parmar, V.; Kumar, A.; Prakash, G.V.; Datta, S.; Kalyanasundaram, D. Investigation, modelling and validation of material separation mechanism during fiber laser machining of medical grade titanium alloy Ti6Al4V and stainless steel SS316L. Mech. Mater. 2019, 137, 103125. [Google Scholar] [CrossRef]
  65. Kalyanasundaram, D.; Shrotriya, P.; Molian, P. Fracture mechanics-based analysis for hybrid laser/waterjet (LWJ) machining of yttria-partially stabilized zirconia (Y-PSZ). Int. J. Mach. Tools Manuf. 2010, 50, 97–105. [Google Scholar] [CrossRef]
  66. Haripriya, M.; Manimekala, T.; Dharmalingam, G.; Minakshi, M.; Sivasubramanian, R.J.C.A.A.J. Asymmetric supercapacitors based on ZnCo2O4 nanohexagons and orange peel derived activated carbon electrodes. Chem. Asian. J. 2024, 19, e202400202. [Google Scholar] [CrossRef] [PubMed]
  67. Vasudevan, S.; Manickam, M.; Sivasubramanian, R.J.D.T. A sol–gel derived LaCoO3 perovskite as an electrocatalyst for Al–air batteries. Dalton Trans. 2024, 53, 3713–3721. [Google Scholar] [CrossRef]
  68. Fujiyama, K.; Mizutani, Y.; Taniguchi, Y.; Kimachi, H. EBSD Analysis of Creep Damage Process in SUS304HTB Steel. J. Soc. Mater. Sci. Jpn. 2013, 62, 305–310. [Google Scholar] [CrossRef]
  69. Kobayashi, D.; Miyabe, M.; Kagiya, Y.; Sugiura, R.; Yokobori, A.T. An Assessment and Estimation of the Damage Progression Behavior of IN738LC under Various Applied Stress Conditions Based on EBSD Analysis. Metall. Mater. Trans. A 2013, 44, 3123–3135. [Google Scholar] [CrossRef]
  70. Fujiyama, K.; Kimachi, H.; Watanabe, Y.; Hijikuro, K.; Tsuboi, T. The concept of “EBSD strain analysis” and its application to creep and creep-fatigue damage assessment of ferritic and austenitic heat resistant steels. Strength 2011, 7, 123–135. [Google Scholar] [CrossRef]
  71. Kobayashi, D.; Miyabe, M.; Kagiya, Y.; Nagumo, Y.; Sugiura, R.; Matsuzaki, T.; Yokobori, A.T. Analysis of damage behaviour based on EBSD method under creep–fatigue conditions for polycrystalline nickel base superalloys. Mater. Sci. Technol. 2014, 30, 24–31. [Google Scholar] [CrossRef]
  72. Kobayashi, D.; Takeuchi, T.; Achiwa, M. J0310104 Evaluation of Fatigue Crack Growth Rate by the EBSD Method. Proc. Mech. Eng. Congr. Jpn. 2015, 2015, J0310104. [Google Scholar] [CrossRef]
  73. Rui, S.S.; Shang, Y.B.; Su, Y.; Qiu, W.; Niu, L.S.; Shi, H.J.; Matsumoto, S.; Chuman, Y. EBSD analysis of cyclic load effect on final misorientation distribution of post-mortem low alloy steel: A new method for fatigue crack tip driving force prediction. Int. J. Fatigue 2018, 113, 264–276. [Google Scholar] [CrossRef]
  74. Sun, C.; Sun, J.; Chi, W.; Wang, J.; Wang, W. A method of quasi in-situ EBSD observation for microstructure and damage evolution in fatigue and dwell fatigue of Ti alloys. Int. J. Fatigue 2023, 176, 107897. [Google Scholar] [CrossRef]
  75. Wu, W.; Wang, Y.; Zhang, X.; Wang, J.; Wei, S. Influence of laser cutting on TC4 titanium alloy sheet. Ordnance Mater. Sci. Eng. 2013, 36, 35–38. [Google Scholar]
  76. Hongling, H.; Yuhai, P.; Yukun, X.; Kai, P.; Hui, Z.J.L.T. Effect of laser cutting on fatigue performance of aircraft skin. Laser Technol. 2016, 40, 417–421. [Google Scholar]
  77. Germain, G.; Morel, F.; Lebrun, J.L.; Morel, A.; Huneau, B. Effect of Laser Assistance Machining on Residual Stress and Fatigue Strength for a Bearing Steel (100Cr6) and a Titanium Alloy (Ti 6Al 4V). Mater. Sci. Forum 2006, 524–525, 569–574. [Google Scholar] [CrossRef]
  78. Ullah, S.; Li, X.; Guo, G.; Riveiro, R.A.; Li, D.; Du, J.; Cui, L.; Wei, L.; Liu, X. Influence of the fiber laser cutting parameters on the mechanical properties and cutedge microfeatures of a AA2B06T4 aluminum alloy. Opt. Lasers Technol. 2022, 156, 108395. [Google Scholar] [CrossRef]
  79. Zhen, A.; Tianqi, L.; Yuanyuan, L.; Xu, D.; Hui, X.; Xiaonan, M.; Pingxiang, Z.; Jinshan, L.; University, X.A. In-Situ SEM Observations of Tensile Deformation Behavior of Ti555211 Alloy with Dual Structure. Rare Met. Mater. Eng. 2019, 48, 1584–1590. [Google Scholar]
  80. Wang, J. An Experimental Analysis and Optimisation of the CO2 Laser Cutting Process for Metallic Coated Sheet Steels. Int. J. Adv. Manuf. Technol. 2000, 16, 334–340. [Google Scholar] [CrossRef]
  81. Ramaswamy, M.T. Heat Affected Zone Studies of Thermally Cut Structural Steels. Master’s Thesis, Oregon Graduate Center, Beaverton, OR, USA, 1989. [Google Scholar]
  82. Bursi, O.S.; D’Incau, M.; Zanon, G.; Raso, S.; Scardi, P. Laser and mechanical cutting effects on the cut-edge properties of steel S355N. J. Constr. Steel Res. 2017, 133, 181–191. [Google Scholar] [CrossRef]
  83. Chen, S.; Zhang, R.; Shi, Z.; Lin, J. An improved DIC method for full-field strain measurement in tensile tests on aluminium alloys under hot stamping conditions. Int. J. Lightweight Mater. Manuf. 2024, 7, 438–449. [Google Scholar] [CrossRef]
  84. Palumbo, G.; Sorgente, D.; Tricarico, L. The design of a formability test in warm conditions for an AZ31 magnesium alloy avoiding friction and strain rate effects. Int. J. Mach. Tools Manuf. 2008, 48, 1535–1545. [Google Scholar] [CrossRef]
  85. Hatamleh, O. Effects of peening on mechanical properties in friction stir welded 2195 aluminum alloy joints. Mater. Sci. Eng. A 2008, 492, 168–176. [Google Scholar] [CrossRef]
  86. Leitao, C.; Galvao, I.; Leal, R.M.; Rodrigues, D.M. Determination of local constitutive properties of aluminium friction stir welds using digital image correlation. Mater. Design 2012, 33, 69–74. [Google Scholar] [CrossRef]
  87. Nielsen, K.L.; Pardoen, T.; Tvergaard, V.; Meester, B.D.; Simar, A. Modelling of plastic flow localisation and damage development in friction stir welded 6005A aluminium alloy using physics based strain hardening law. Int. J. Solids Struct. 2010, 47, 2359–2370. [Google Scholar] [CrossRef]
  88. Boyce, B.L.; Reu, P.L.; Robino, C.V. The constitutive behavior of laser welds in 304L stainless steel determined by digital image correlation. Metall. Mater. Trans. A 2006, 37, 2481–2492. [Google Scholar] [CrossRef]
  89. Scintilla, L.D.; Tricarico, L.; Brandizzi, M.; Satriano, A.A. Nd:YAG laser weldability and mechanical properties of AZ31 magnesium alloy butt joints. J. Mater. Process. Technol. 2010, 210, 2206–2214. [Google Scholar] [CrossRef]
  90. Abu, S.H.K.; Cao, X.J.; Gholipour Baradari, J.; Wanjara, P.; Cuddy, J.; Birur, A.; Medraj, M. Determination of Global and Local Tensile Behaviours of Laser Welded Ti-6Al-4V Alloy. Adv. Mater. Res. 2012, 409, 859–864. [Google Scholar] [CrossRef]
  91. Kabir, A.S.H.; Cao, X.; Gholipour, J.; Wanjara, P.; Cuddy, J.; Birur, A.; Medraj, M. Effect of Postweld Heat Treatment on Microstructure, Hardness, and Tensile Properties of Laser-Welded Ti-6Al-4V. Metall. Mater. Trans. A 2012, 43, 4171–4184. [Google Scholar] [CrossRef]
  92. Scintilla, L.D.; Palumbo, G.; Sorgente, D.; Tricarico, L. Fiber laser cutting of Ti6Al4V sheets for subsequent welding operations: Effect of cutting parameters on butt joints mechanical properties and strain behaviour. Mater. Design 2013, 47, 300–308. [Google Scholar] [CrossRef]
  93. Luo, J.; Zhu, G.; Li, J.J. Numerical simulation on temperature field and flow field of molten pool in laser cladding. Surf. Technol. 2023, 52, 67–84. [Google Scholar]
  94. Pavelic, V.; Tanbakuchi, R.; Uyehara, O.A.; Myers, P.S. Experimental and computed temperature histories in gas tungsten-arc welding of thin plates. Weld. Res. Suppl. 1969, 48, 296–305. [Google Scholar]
  95. Jiang, H.-J.; Xu, W.-L.; Gao, N.-H.; Jia, C.-S.; Wang, X.-G. 3D thermoelastoplastic study of FML structure under the action of a moving Gaussian-distributed laser heat source. Thin-Walled Struct. 2022, 179, 109759. [Google Scholar] [CrossRef]
  96. Wu, S.; Zhao, H.; Wang, Y.; Zhang, X. New Heat Source Model in Numerical Simulation of High Energy Beam Welding. China Weld. 2004, 25, 91–94. [Google Scholar]
  97. Unni, A.K.; Vasudevan, M. Determination of heat source model for simulating full penetration laser welding of 316 LN stainless steel by computational fluid dynamics. Mater. Today 2021, 45, 4465–4471. [Google Scholar] [CrossRef]
  98. Ou, D.; Wang, F.; Zhao, S.; Zheng, J.; Liu, T. Welding Deformations and Stress Simulations Control of Large-scale Complex Structures. Chin. J. Mech. Eng. 2018, 29, 616. [Google Scholar]
  99. Kik, T.; Garašić, I.; Perić, M.; Landek, D.; Jurica, M.; Tonković, Z. Modifications of the heat source model in numerical analyses of the metal-cored arc welding process. Energy 2024, 302, 131811. [Google Scholar] [CrossRef]
  100. Arif, A.; Yilbas, B.; Aleem, B.A. Laser cutting of thick sheet metals: Residual stress analysis. Opt. Lasers Technol. 2009, 41, 224–232. [Google Scholar] [CrossRef]
  101. Yilbas, B.S. Laser cutting of thick sheet metals: Effects of cutting parameters on kerf size variations. J. Mater. Process. Technol. 2008, 201, 285–290. [Google Scholar] [CrossRef]
  102. Yilbas, B.S.; Akhtar, S.S. Laser Cutting of Alloy Steel: Three-Dimensional Modeling of Temperature and Stress Fields. Adv. Manuf. Process. 2011, 26, 104–112. [Google Scholar] [CrossRef]
  103. Wang, R.R.; Chang, C.T. Thermal modeling of laser welding for titanium dental restorations. J. Prosthet. Dent. 1998, 79, 335–341. [Google Scholar] [CrossRef]
  104. Bandyopadhyay, S.; Sundar, J.S.; Sundararajan, G.; Joshi, S. Geometrical features and metallurgical characteristics of Nd: YAG laser drilled holes in thick IN718 and Ti–6Al–4V sheets. J. Mater. Process. Technol. 2002, 127, 83–95. [Google Scholar] [CrossRef]
  105. Kofstad, P. High-temperature oxidation of titanium. J. Less Common Met. 1967, 12, 449–464. [Google Scholar] [CrossRef]
  106. Abakians, H.; Modest, M.F. Evaporative cutting of a semitransparent body with a moving CW laser. J. Heat Transf. 1988, 110, 924–930. [Google Scholar] [CrossRef]
  107. Di Pietro, P.; Yao, Y. A new technique to characterize and predict laser cut striations. Int. J. Mach. Tools Manuf. 1995, 35, 993–1002. [Google Scholar] [CrossRef]
  108. Li, Y.; Latham, W.P.; Kar, A. Lumped parameter model for multimode laser cutting. Opt. Lasers Eng. 2001, 35, 371–386. [Google Scholar] [CrossRef]
  109. Aloke, R.; Girish, V.; Scrutton, R.F.; Molian, P.A. A model for prediction of dimensional tolerances of laser cut holes in mild steel thin plates. Int. J. Mach. Tools Manuf. 1997, 37, 1069–1078. [Google Scholar] [CrossRef]
  110. Poprawe, R.; König, W. Modeling, Monitoring and Control in High Quality Laser Cutting. CIRP Ann.-Manuf. Technol. 2001, 50, 137–140. [Google Scholar] [CrossRef]
  111. Yilba, B.S.; Sahin, A.Z. Oxygen assisted laser cutting mechanism—A laminar boundary layer approach including the combustion process. Opt. Laser Technol. 1995, 27, 175–184. [Google Scholar] [CrossRef]
  112. Kim, J.H.; Kim, E.J.; Lee, C.M. A study on the heat affected zone and machining characteristics of difficult-to-cut materials in laser and induction assisted machining. J. Manuf. Process. 2020, 57, 499–508. [Google Scholar] [CrossRef]
  113. Rozzi, J.C.; Pfefferkorn, F.E.; Incropera, F.P.; Shin, Y.C. Transient, three-dimensional heat transfer model for the laser assisted machining of silicon nitride: I. Comparison of predictions with measured surface temperature histories. Int. J. Heat Mass Transf. 2000, 43, 1409–1424. [Google Scholar] [CrossRef]
  114. Yang, J.; Sun, S.; Brandt, M.; Yan, W. Experimental investigation and 3D finite element prediction of the heat affected zone during laser assisted machining of Ti6Al4V alloy. J. Mater. Process. Technol. 2010, 210, 2215–2222. [Google Scholar] [CrossRef]
  115. Kardas, O.O.; Keles, O.; Akhtar, S.; Yilbas, B.S. Laser cutting of rectangular geometry in 2024 aluminum alloy: Thermal stress analysis. Opt. Laser. Technol. 2014, 64, 247–256. [Google Scholar] [CrossRef]
  116. Gu, D.; He, B. Finite element simulation and experimental investigation of residual stresses in selective laser melted Ti–Ni shape memory alloy. Comput. Mater. Sci. 2016, 117, 221–232. [Google Scholar] [CrossRef]
  117. Samanta, A.; Teli, M.; Singh, R. Experimental characterization and finite element modeling of the residual stresses in laser-assisted mechanical micromachining of Inconel 625. Proc. Inst. Mech. Eng. Part B J. Eng. Manuf. 2017, 231, 1735–1751. [Google Scholar] [CrossRef]
  118. Mittal, V.; Sharma, P.; Patel, V.K.; Pandey, V.; Kumar, V. Numerical Study of Laser Cutting of Titanium Alloy (TI-6Al-4V). Appl. Mech. Mater. 2014, 592, 579–583. [Google Scholar] [CrossRef]
  119. Shamlooei, M.; Zanon, G.; Valli, A.; Bison, P.; Bursi, O.S. Investigation of thermal behaviour of structural steel S235N under laser cutting process: Experimental, analytical, and numerical studies. Eng. Struct. 2022, 269, 114754. [Google Scholar] [CrossRef]
  120. Zhai, C.; Xu, J.; Li, Y.; Hou, Y.; Liu, Q. Study on surface heat-affected zone and surface quality of Ti-6Al-4V alloy by laser-assisted micro-cutting. Int. J. Adv. Manuf. Technol. 2020, 109, 2337–2352. [Google Scholar] [CrossRef]
  121. Ding, H.; Shin, Y.C. A Metallo-Thermomechanically Coupled Analysis of Orthogonal Cutting of AISI 1045 Steel. J. Manuf. Sci. Eng. 2012, 134, 051014. [Google Scholar] [CrossRef]
  122. Fu, C.H.; Sealy, M.P.; Guo, Y.; Wei, X.T. Finite element simulation and experimental validation of pulsed laser cutting of nitinol. J. Manuf. Process. 2015, 19, 81–86. [Google Scholar] [CrossRef]
  123. Wee, L.M.; Li, L. An analytical model for striation formation in laser cutting. Appl. Surf. Sci. 2005, 247, 277–284. [Google Scholar] [CrossRef]
  124. Tani, G.; Tomesani, L.; Campana, G.; Fortunato, A. Quality factors assessed by analytical modelling in laser cutting. Thin Solid Films 2004, 453, 486–491. [Google Scholar] [CrossRef]
  125. Kovalev, O.; Yudin, P.; Zaitsev, A. Modeling of flow separation of assist gas as applied to laser cutting of thick sheet metal. Appl. Math. Model. 2009, 33, 3730–3745. [Google Scholar] [CrossRef]
  126. Kar, A.; Scott, J.; Latham, W. Theoretical and experimental studies of thick-section cutting with a chemical oxygen–iodine laser (COIL). J. Laser Appl. 1996, 8, 125–133. [Google Scholar] [CrossRef]
  127. Atsuta, T.; Yasuda, K.; Matsumoto, T.; Sakurai, T.; Okado, H. COIL and the Material Processing. In Proceedings of the Conference on Lasers and Electro-Optics, Amsterdam, The Netherlands, 28 August–2 September 1994; p. CThI52. [Google Scholar]
  128. Mullick, S.; Madhukar, Y.K.; Roy, S.; Nath, A.K. An investigation of energy loss mechanisms in water-jet assisted underwater laser cutting process using an analytical model. Int. J. Mach. Tools Manuf. 2015, 91, 62–75. [Google Scholar] [CrossRef]
Figure 1. The overview of issues covered in this review.
Figure 1. The overview of issues covered in this review.
Metals 14 01152 g001
Figure 2. Basic structure of a laser.
Figure 2. Basic structure of a laser.
Metals 14 01152 g002
Figure 4. Schematic diagram of laser cutting [13].
Figure 4. Schematic diagram of laser cutting [13].
Metals 14 01152 g004
Figure 5. Schematic diagram of laser cutting quality evaluation parameters [13].
Figure 5. Schematic diagram of laser cutting quality evaluation parameters [13].
Metals 14 01152 g005
Figure 6. Effect of (a) pulse energy, (b) speed, and (c) gas pressure on the heat-affected zone (HAZ) thickness during cutting with an Nd:YAG pulsed laser [46].
Figure 6. Effect of (a) pulse energy, (b) speed, and (c) gas pressure on the heat-affected zone (HAZ) thickness during cutting with an Nd:YAG pulsed laser [46].
Metals 14 01152 g006
Figure 7. HAZ section [61]: (a) optical microscopy photographs of Ti–6Al–4V laser cut kerf cross section (50×) and (b) recast layer of Ti–6Al–4V laser cut edge (1000×).
Figure 7. HAZ section [61]: (a) optical microscopy photographs of Ti–6Al–4V laser cut kerf cross section (50×) and (b) recast layer of Ti–6Al–4V laser cut edge (1000×).
Metals 14 01152 g007
Figure 8. Cutting edge microstructure [62]: (a) overall view. Divided by the blue line into the narrow top zone, intermediate zone and bottom zone. (b) middle region enlargement.
Figure 8. Cutting edge microstructure [62]: (a) overall view. Divided by the blue line into the narrow top zone, intermediate zone and bottom zone. (b) middle region enlargement.
Metals 14 01152 g008
Figure 9. Scanning electron microscopy (SEM) photographs of microcracks and voids at cutting edges [43]: (a) V = 480 mm/min and (b) V = 2400 mm/min.
Figure 9. Scanning electron microscopy (SEM) photographs of microcracks and voids at cutting edges [43]: (a) V = 480 mm/min and (b) V = 2400 mm/min.
Metals 14 01152 g009
Figure 10. X-ray photoelectron spectroscopy (XPS), X-ray diffraction (XRD), and energy-dispersive X-ray spectroscopy (EDX) analysis diagram [64]: (a) different oxides formed on bare/pristine and laser-machined surfaces of Ti–6Al–4V alloy (Peak 1 denotes the metallic Ti, 2p3/2 peaks of TiN, Ti2O3 and TiO2 are shown by dotted line 2, 3 and 4 respectively. The 2p1/2 peaks of TiN, Ti2O3 and TiO2 are indicated by 2′, 3′ and 4′ respectively), (b) Cu-Kα XRD, and (c) bar chart of EDX results for bare and laser-processed surfaces.
Figure 10. X-ray photoelectron spectroscopy (XPS), X-ray diffraction (XRD), and energy-dispersive X-ray spectroscopy (EDX) analysis diagram [64]: (a) different oxides formed on bare/pristine and laser-machined surfaces of Ti–6Al–4V alloy (Peak 1 denotes the metallic Ti, 2p3/2 peaks of TiN, Ti2O3 and TiO2 are shown by dotted line 2, 3 and 4 respectively. The 2p1/2 peaks of TiN, Ti2O3 and TiO2 are indicated by 2′, 3′ and 4′ respectively), (b) Cu-Kα XRD, and (c) bar chart of EDX results for bare and laser-processed surfaces.
Metals 14 01152 g010
Figure 11. Electron backscatter diffraction (EBSD) microstructure results [74]: (a) boundary contrast (BC) map, (b) phase diagram, blue α phase, yellow β phase, (c) inverse pole figure inverse pole figure (IPF), and (d) kernel average misorientation (KAM) diagram.
Figure 11. Electron backscatter diffraction (EBSD) microstructure results [74]: (a) boundary contrast (BC) map, (b) phase diagram, blue α phase, yellow β phase, (c) inverse pole figure inverse pole figure (IPF), and (d) kernel average misorientation (KAM) diagram.
Metals 14 01152 g011
Figure 12. Microstructure evolution of Ti–6Al–4V.
Figure 12. Microstructure evolution of Ti–6Al–4V.
Metals 14 01152 g012
Figure 13. Effect of processing parameters on ultimate tensile strength (UTS) [78]: (a) Main effect plots of laser cutting parameters on the UTS for 0.5 mm and 1 mm; (b) SE-pressure effect on 0.5 mm thick sheet; (c) SE-pressure effect on 1 mm thick sheet; (d) fatigue test for samples processed at a SE = 6.67 and 13.34 J/mm2; gas pressure was 1.5 MPa.
Figure 13. Effect of processing parameters on ultimate tensile strength (UTS) [78]: (a) Main effect plots of laser cutting parameters on the UTS for 0.5 mm and 1 mm; (b) SE-pressure effect on 0.5 mm thick sheet; (c) SE-pressure effect on 1 mm thick sheet; (d) fatigue test for samples processed at a SE = 6.67 and 13.34 J/mm2; gas pressure was 1.5 MPa.
Metals 14 01152 g013aMetals 14 01152 g013b
Figure 14. SEM images of fracture surface in fatigue test [78]: (a) Fracture surface view; (b) The dimple structure at the crack initiation region; (c) The crack initiation region from the view, perpendicular to the fracture surface; (d) The crack propagation or growing region.
Figure 14. SEM images of fracture surface in fatigue test [78]: (a) Fracture surface view; (b) The dimple structure at the crack initiation region; (c) The crack initiation region from the view, perpendicular to the fracture surface; (d) The crack propagation or growing region.
Metals 14 01152 g014
Figure 15. XRD patterns of laser cutting surface [82].
Figure 15. XRD patterns of laser cutting surface [82].
Metals 14 01152 g015
Figure 16. Use scenarios of digital image correlation (DIC) technology tensile test [92]: (a) Aramis three-dimensional DIC system integrated into INSTRON 4485 universal tensile testing machine and (b) main strain fields obtained in welded specimens by the DIC technique during the pre-fracture phase.
Figure 16. Use scenarios of digital image correlation (DIC) technology tensile test [92]: (a) Aramis three-dimensional DIC system integrated into INSTRON 4485 universal tensile testing machine and (b) main strain fields obtained in welded specimens by the DIC technique during the pre-fracture phase.
Metals 14 01152 g016
Figure 17. Planar Gaussian heat source model [95].
Figure 17. Planar Gaussian heat source model [95].
Metals 14 01152 g017
Figure 18. Rotary Gaussian body heat source model [97].
Figure 18. Rotary Gaussian body heat source model [97].
Metals 14 01152 g018
Figure 19. Double-ellipsoidal heat source model [99].
Figure 19. Double-ellipsoidal heat source model [99].
Metals 14 01152 g019
Figure 20. Laser cutting simulation [106,107,108,109,110,111].
Figure 20. Laser cutting simulation [106,107,108,109,110,111].
Metals 14 01152 g020
Figure 21. Experimental and simulated temperature test paths [119].
Figure 21. Experimental and simulated temperature test paths [119].
Metals 14 01152 g021
Figure 22. Laser cutting temperature field and stress field distribution [100].
Figure 22. Laser cutting temperature field and stress field distribution [100].
Metals 14 01152 g022
Figure 23. Schematic of laser cutting, conical volume heat flux (The colors in the figure indicating the temperature distribution of the heat source, with red indicating higher temperatures and the central area of the heat source, blue indicating lower temperatures and areas far away from the heat source, and green and yellow indicating the medium temperature range.), and pulse cutting [122].
Figure 23. Schematic of laser cutting, conical volume heat flux (The colors in the figure indicating the temperature distribution of the heat source, with red indicating higher temperatures and the central area of the heat source, blue indicating lower temperatures and areas far away from the heat source, and green and yellow indicating the medium temperature range.), and pulse cutting [122].
Metals 14 01152 g023
Figure 24. Influence of airflow separation on laser cutting surface quality [125]: (a) Schlieren picture; (b) streamlines of the gas, numerical simulations; and (c) density gradient of the gas, numerical results. Samples of the surface with defects produced by the laser treatment: (d) stainless steel 5 mm thick; (e) stainless steel 16 mm thick; and (f) titanium 30 mm thick.
Figure 24. Influence of airflow separation on laser cutting surface quality [125]: (a) Schlieren picture; (b) streamlines of the gas, numerical simulations; and (c) density gradient of the gas, numerical results. Samples of the surface with defects produced by the laser treatment: (d) stainless steel 5 mm thick; (e) stainless steel 16 mm thick; and (f) titanium 30 mm thick.
Metals 14 01152 g024
Table 1. Classification of laser cutting parameters.
Table 1. Classification of laser cutting parameters.
Beam ParameterProcess Parameter
Laser power (Pu)Auxiliary gas and pressure (p)
Wavelength (λ)Cutting speed (v)
Spot size (r)Cutting distance (D)
Types of beamsNozzle diameter (R)
Pulse width (PW)
Pulse frequency (Pf)
Pulse energy (Q)
Defocusing quantity (f)
Table 2. Research on laser cutting process parameters of titanium alloys.
Table 2. Research on laser cutting process parameters of titanium alloys.
AuthorLaserMaterialMethodProcess ParametersAuxiliary GasCutting Quality Evaluation ParameterOptimal Combination of Parameters
Aoud et al. [43]CO2 laser cutting3 mm Ti-6Al-4VTMPu, v, pN2RaPu = 3 kW, v = 2400 mm/min, and p = 2 bar
Shrivastava et al. [44]Pulsed Nd:YAG laser cutting1.6 mm Ti-6Al-4VGA and MRMPW, Pf, v, pAirKW, KTp = 9.959 bar, PW = 1.447 ms, Pf = 6 Hz, v = 5.06 mm/min
Boudjemline et al. [45]CO2 laser cutting5 mm Ti-6Al-4VFFDPu = 2000 W, v, pN2Rav = 2250–2400 mm/min, p = 12–14 bar
Shanjin et al. [46]pulsed Nd:YAG laser cutting1 mm TC1SFEQ, PW, v, pair, Ar, N2HAZMedium-Q, high-PW, high-v, and high-p, Ar
Kochergin et al. [47]fiber laser cutting2 mm BT1-0SFEv, Pf, p, fArKW, Ra, Sv = 3 mm/min, Pf = 100–200 Hz, p = 1–1.2 MPa, f = 0 mm
Kumar et al. [5]pulsed Nd:YAG laser cutting1.4 mm Ti-6Al-4VTMp, v, PW, PfN2Ra, KTLow-PW, low-Pf, high-v, medium-p
Boujelbene et al. [48]CO2 laser cutting2 mm pure TiTMv, Pu, pN2Rav = 2400 mm/min, Pu = 2 kW, p = 14 bar
El Aoud B et al. [49]CO2 laser cutting3 mm Ti-6Al-4V and TiTMPu, v, pN2KWPu = 2 kW, v = 2400 mm/min, p = 8 bar
Pandey et al. [50]pulsed Nd:YAG laser cutting1.4 mm Ti-6Al-4VNNM and GAPW, Pf, v, p Rap = 50.04 N/2, v = 0.278 mm/min, Pf = 13.9995 Hz, PW = 1.7 ms
Tamilarasan et al. [41]pulsed Nd:YAG laser cutting1 mm Ti-6Al-4VRD and GAPW, Q, v, p KD, MRRPW = 1.789 ms, Q = 4.574 J, v = 10 mm/min, p = 7.674 kg/cm2
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

Zhang, Y.; Wang, C.; Xu, W.; Zhang, X.; Ren, K.; Wang, S.; Hua, Q. Laser Cutting of Titanium Alloy Plates: A Review of Processing, Microstructure, and Mechanical Properties. Metals 2024, 14, 1152. https://doi.org/10.3390/met14101152

AMA Style

Zhang Y, Wang C, Xu W, Zhang X, Ren K, Wang S, Hua Q. Laser Cutting of Titanium Alloy Plates: A Review of Processing, Microstructure, and Mechanical Properties. Metals. 2024; 14(10):1152. https://doi.org/10.3390/met14101152

Chicago/Turabian Style

Zhang, Ya, Chunyu Wang, Wentao Xu, Xianfeng Zhang, Kerong Ren, Shuai Wang, and Qing Hua. 2024. "Laser Cutting of Titanium Alloy Plates: A Review of Processing, Microstructure, and Mechanical Properties" Metals 14, no. 10: 1152. https://doi.org/10.3390/met14101152

APA Style

Zhang, Y., Wang, C., Xu, W., Zhang, X., Ren, K., Wang, S., & Hua, Q. (2024). Laser Cutting of Titanium Alloy Plates: A Review of Processing, Microstructure, and Mechanical Properties. Metals, 14(10), 1152. https://doi.org/10.3390/met14101152

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

Article Metrics

Back to TopTop