Next Article in Journal
Effect of Electrodeposition Conditions on Adsorption and Photocatalytic Properties of ZnO
Previous Article in Journal
Oxygen Nonstoichiometry, Electrical Conductivity, Chemical Expansion and Electrode Properties of Perovskite-Type SrFe0.9V0.1O3−δ
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Effects of Hypergravity on Phase Evolution, Synthesis, Structures, and Properties of Materials: A Review

1
Institute of Microstructure and Property of Advanced Materials, Beijing University of Technology, Beijing 100124, China
2
Institute of Quantum Materials and Physics, Henan Academy of Sciences, Zhengzhou 450046, China
*
Author to whom correspondence should be addressed.
Materials 2025, 18(3), 496; https://doi.org/10.3390/ma18030496
Submission received: 13 December 2024 / Revised: 14 January 2025 / Accepted: 16 January 2025 / Published: 22 January 2025

Abstract

In a hypergravity environment, the complex stress conditions and the change in gravity field intensity will significantly affect the interaction force inside solid- and liquid-phase materials. In particular, the driving force for the relative motion of the phase material, the interphase contact interaction, and the stress gradient are enhanced, which creates a nonlinear effect on the movement mode of the phase material, resulting in a change in the material’s behavior. These changes include increased stress and contact interactions; accelerated phase separation; changes in stress distribution; shear force and phase interface renewal; enhanced interphase mass transfer and molecular mixing; and increased volume mass transfer and heat transfer coefficients. These phenomena have significant effects on the synthesis, structural evolution, and properties of materials in different phases. In this paper, the basic concepts of hypergravity and the general rules of the effects of hypergravity on the synthesis, microstructure evolution, and properties of materials are reviewed. Based on the development of hypergravity equipment and characterization methods, this review is expected to broaden the theoretical framework of material synthesis and mechanical property control under hypergravity. It provides theoretical reference for the development of high-performance materials under extreme conditions, as well as new insights and methods for research and application in related fields.

1. Introduction

When a material is on the Earth’s surface, it is subjected to the force of gravity, resulting in a normal-gravity environment with Earth’s average gravitational acceleration, g, of 9.81 m/s2. However, when the acceleration exceeds the temporal gravitational acceleration, the force acting on the material is known as hypergravity, resulting in a hypergravity environment. Hypergravity conditions are widespread in the environments in which people live, such as in matter under conditions of centrifugal motion, where the centrifugal acceleration can be much greater than the gravitational acceleration. Hypergravity environments can be achieved through centrifugal force generated by rotation [1], which includes equipment such as centrifuges, rotating packed beds, and turbines. In centrifugal motion, materials experience centrifugal force, interphase interaction forces, and the Coriolis force. Compared to normal gravity, hypergravity enhances the self-weight stress of multiphase media and increases its gradient. It also increases the driving force for relative motion between different phases by enhancing interphase interaction forces. Therefore, the fluid flow properties and interphase interaction in multiphase systems under hypergravity conditions play a crucial role in the phase evolution of materials. The phase evolution in multiphase materials under hypergravity conditions will be different from that in materials under conventional gravity conditions. For example, under hypergravity conditions, the solidification of alloy melts leads to grain refinement and the weakening of dendrites [2,3], changes in boundary layer structure and flow characteristics in fluids [4,5], strengthening reactions and the synthesis of nanomaterials [6,7], and the separation of two-phase materials in metallurgical processes [8,9].
Hypergravity is also widely present in components with a single solid phase, such as turbines used in gas turbines and engines in industries. The study of the structural evolution of solid materials under hypergravity conditions is of crucial importance for industrial safety. However, in the study of solid materials under hypergravity conditions, the effects of hypergravity are often overlooked. The effects of a hypergravity environment on components such as turbines is often accompanied by high temperatures and stresses [10]. The extreme conditions of high temperature and complex stress have often led previous studied to simplify hypergravity to conventional stress, focusing more on the effects of high-temperature and high-stress coupling on solid materials while neglecting the effects of stress gradients and interphase interaction forces caused by hypergravity. Taking superalloys under high-temperature hypergravity conditions as an example, previous studies have used conventional stress applied under normal gravity to simulate the effects of hypergravity on the structural evolution, mechanical behavior, and properties of materials. This simplification has led to discrepancies in creep between simulated rafting structures and those in service alloys, as well as lower simulated creep life compared to actual service life. This phenomenon may be related to the directional diffusion of Ni/Al atoms caused by centrifugal stress gradients under hypergravity, which determines the dependence of precipitate morphology evolution in Ni-based alloys [11]. Therefore, the enhanced mass transfer under hypergravity and its impact on other characteristics significantly affect the microstructure and behavior of materials, indicating that the influence of hypergravity on the diffusion, microstructural evolution, and mechanical behavior of solid alloy materials cannot be simplified or ignored.
Element diffusion is the basis of structural evolution in solid materials, which in turn affects the microstructural evolution and mechanical behavior and properties of materials. Therefore, studying the diffusion laws and mechanisms of elements under hypergravity conditions and elucidating the physical mechanisms of hypergravity’s effect on solid-state diffusion and phase evolution are fundamental for studying the mechanical behavior and properties of solid materials under hypergravity. This will lead to a deeper understanding of the mechanical behavior of solid materials under hypergravity environments. The aforementioned research provides possibilities for revealing the diffusion laws and mechanisms, mechanical behavior, and other aspects of solid materials under hypergravity conditions. It also provides a theoretical foundation and experimental basis for the development and failure analysis of high-performance alloy materials under high-temperature hypergravity conditions. Furthermore, it has important practical and scientific implications for the application of solid materials in complex hypergravity environments.
The architecture of the paper is organized as follows: Section 1 introduces the background of hypergravity research; Section 2 elaborates on the principles and techniques of hypergravity; Section 3 discusses the liquid-phase synthesis and separation of materials under hypergravity conditions; Section 4 presents the metallurgy of alloys under hypergravity conditions; Section 5 analyzes the structure and properties of alloy materials under hypergravity; and Section 6 concludes with outlooks and future prospects.
In this review, we adopted a systematic approach to collect and analyze the corresponding literature on hypergravity-related research fields, such as the fundamental concept of hypergravity, hypergravity technologies, phase evolution, materials preparation, phase separation, centrifugal casting, and properties. We conducted a comprehensive review of the literature to identify studies directly related to the effects of hypergravity on materials-related multidisciplinarity. It is designed to provide researchers, engineers, and students of disciplines in the field of materials science related to hypergravity with a comprehensive perspective to inform and inspire their research and practice.

2. Principles and Techniques of Hypergravity

2.1. Hypergravity Condition

High-speed rotation of a centrifuge is an effective method to create a stable hypergravity field. For centrifugal hypergravity environments, the origin of the stationary coordinate system and the moving coordinate system are respectively located at the axis of the centrifuge and the center of the moving object (such as the experimental chamber). The governing equation for a moving object of uniform mass, m, in the rotating coordinate system, a hypergravity condition with a constant rotational speed, can be described with Equation (1) [12].
m d r 2 r d t 2 = F m ω × ( ω × R ) m ω × ( m ω × r ) m ω × ν
where F is the external force, ω is the rotational angular velocity vector, R is the rotation radius vector of the stationary coordinate system, r is the position vector of the moving coordinate system of particle (sample in the experimental chamber), and v′ is the velocity vector of the sample relative to the moving coordinate system. The second term on the right-hand side of the equation, m ω × ( ω × R ) , represents the linear hypergravity (centrifugal) force; the third term, m ω × ( ω × r ) , represents the nonlinear centrifugal force; and the last term represents the Coriolis force.
The hypergravity coefficient, G, which describes the magnitude of the hypergravity field, is defined as the ratio of centrifugal acceleration to normal gravitational acceleration. It is defined according to Equation (2):
G = ω 2 R g = N r 2 π 2 R 900 g ,
where Nr is the rotation speed, and the hypergravity intensity could be denoted as Ng, where N is an integer greater than 1. The unit of hypergravity coincides with the unit of gravitational acceleration, i.e., m/s2.
According to Equation (1), the hypergravity acting on samples can be categorized into two main components: the centrifugal force induced by rotation, Fg, and the Coriolis force resulting from variations in movement relative to the coordinate system during rotation, Fc. However, for materials that are not of a single composition, there will be a buoyancy force, Fb, due to variations in the species density, which is one of the interphase forces. Therefore, the above three forces and their effects will be mainly discussed in this review.
The centrifugal force, which is the most significant force acting on the sample, plays a crucial role in the diffusion and structural evolution processes of materials. It can be described as follows:
d F g = S ρ ω 2 R d R ,
where S is the cross-sectional area, and ρ is the density of the sample. Therefore, according to Equation (3), the stress, σ, induced by the centrifugal force can be described as:
σ = R 1 R 2 ρ ω 2 R d R = 1 2 ρ ω 2 R 2 2 R 1 2 ,
where R1 and R2 are the distances of the sample edges from the center of rotation.
Fb can be described as:
F b = m ω 2 R ,
where ∆m is the mass difference of the species, which originates from the density differences in phases, grains, or atoms based on analytical scales (∆m is relevant to ∆ρV, where V is the volume). Obviously, Fb increases with the increasing hypergravity according to Equation (5).
The Coriolis force is described in the following:
F c = 2 m ω × ν ,
It should be noted that both ω and v in this context are vectors, so the magnitude of Fc is related to the cross product of ω and v′.
Obviously, both Fg and Fb are significant for the multiphase system, in accordance with Equations (3)–(5). For the liquid-phase system, the role of Fc cannot be neglected, but for the solid-phase system, Fc can be neglected because v′ is relatively small, according to Equation (6).

2.2. Enhanced Interphase Interaction and Mass Transfer Under Hypergravity Conditions

Equations (1)–(7) describe the stress condition of the sample in hypergravity; most notably, there is an increase in the coefficient of hypergravity and a significant increase in the force on the matter [1], as shown in Figure 1a. Moreover, the buoyancy factor increases between two phases of matter where there is a difference in density, and the interaction force is enhanced, thus enhancing the relative motion between the two phases, as shown in Figure 1b, which shows two solid-phase media in contact with densities of ρ1 and ρ2. Under the action of gravity, the relative motion is the driving force (ρ1ρ2)g at the interface of the media. Hypergravity increases the gravitational acceleration, thereby increasing the driving force of the relative motion to (ρ1ρ2)Ng. Likewise, as the buoyancy factor rises under hypergravity, the driving force of the relative motion of the solid–liquid two-phase medium interface is also enhanced.
On the other hand, the enhanced interphase interaction and mass transfer under hypergravity are also reflected in the enhancement of fluid motion. A higher coefficient of hypergravity also means that the linear velocity of the fluid/matter is faster, so the contact between the two phases is more rapid. The hypergravity environment enhances natural convection due to the concentration, thermal gradients, buoyancy, and hydrostatic pressure, which also promotes interphase interaction and mass transfer. Additionally, the domain in the flat solid–liquid interface is also modulated by the hypergravity condition, since the increasing acceleration/gravity decreases the domain height and increases the interfacial free energy [13]. This will significantly affect the solid–liquid interface structure and promote the interaction between the two phases. It should be noted that the phase components exhibit a completely different motion pattern under hypergravity than normal gravity, resulting in a nonlinear transformation of the material motion pattern, also known as nonlinear effects. This means that under the action of external conditions (such as electric fields, magnetic fields, and a series of external physical fields), the response of the material does not follow a simple linear relationship but shows more-complex nonlinear characteristics [1].
The diffusion mass transfer rate between substances is related to the diffusion coefficient, D, and the diffusion layer thickness, δ, expressed as D/δ. In multiphase media, hypergravity enhances phase contact and mixing and increases the diffusion coefficient [14]. Previous studies have shown that the diffusion layer thickness is inversely proportional to the gravity coefficient in a logarithmic manner [15,16], which reduces the boundary layer thickness, thus significantly promoting mass transfer processes.
The enhanced interphase interaction and mass transfer present in hypergravity environments are interrelated and closely associated with the significant acceleration of each phase under hypergravity. Additionally, the buoyancy factor between phases increases with the augmentation of hypergravity. Moreover, flow characteristics such as turbulence and enhanced shear forces in a hypergravity environment also contribute significantly to interphase interaction and mass transfer.
It is important to note that hypergravity does not affect the chemical equilibrium of the system, which is why conventional thermodynamics do not consider the gravitational field. While hypergravity increases the microscopic chemical potential of substances, since factors such as buoyancy and stress gradient in hypergravity change the distribution of compositions, and the microscopic chemical equilibrium remains unchanged, the distribution of solutes in a solution theoretically changes. However, due to the much stronger intermolecular and interionic forces compared to the effects of hypergravity, the changes in solute distribution are minimal, only becoming significant when the hypergravity coefficient and the height difference of hypergravity direction are substantial. Therefore, solely relying on hypergravity to affect the chemical potential of substances for the enrichment or separation of solutes in a solution or components in a melt is difficult to achieve.
Nevertheless, when there is a concentration gradient of solutes in solution and material melt, hypergravity can accelerate solute transfer by enhancing natural convection. Additionally, when there is a density difference between multiphase media, the effect of hypergravity buoyancy becomes significant. The buoyancy factor, ΔρNg, is directly proportional to the gravity coefficient, as shown in Equation (6). Therefore, hypergravity can accelerate phase separation, achieving the separation of phases with different densities, even when the density difference between the two phases is small, such as in solid–liquid phase separation.
Furthermore, since hypergravity is directional, it is an effective method for the deep separation of fine, dispersed phases in a melt. The directionality of hypergravity is reflected in the directionality of the hypergravity buoyancy force, which affects the relative motion of phases/elements with different masses (relatively light phases/elements/defects move in the direction of the small hypergravity force), thereby affecting the flow properties of the fluid material, as well as the diffusion and elemental distributions of the solid-phase material, resulting in structural and property differences. A more in-depth discussion of the effects of hypergravity on fluids will be further described in the subsequent sections.

2.3. Hypergravity Techniques

High-speed rotational motion is an effective method to create a stable hypergravity field. Technologies based on centrifugal hypergravity environments can be divided into two categories, i.e., equipment applied to liquid and solid phases, respectively. The former is represented by the rotating packed bed (RPB) and the latter by centrifuge-based equipment. Here, we briefly summarize these two types of techniques; relevant details can be found in some of the previous review articles of RPB [17,18,19] and centrifuge-based techniques.

2.3.1. Rotating Packed Bed Techniques

The RPB technique was proposed by Dr. Ramshaw in 1979 [20], in which the hypergravity technology is based on a rotating packed bed (RPB) as the core device, as shown in Figure 2a. The rotating packed bed generates a centrifugal force field through the high-speed rotation of internal rotating components, typically achieving gravity accelerations of hundreds to thousands of times; the technique is also known as hypergravity (Higee) [21,22,23].
The RPB is a tool consisting of a rotating part (i.e., an annular rotor with a bed of packing material) and a static part (i.e., the housing). The rotor and the static part, i.e., the housing, are connected by bearings and seals. The rotor is driven by a motor, and the hollow part of the rotor is referred to as the “eye”, which can be identified by an eye ring and contains the liquid distribution. The rotational speed can be adjusted according to process requirements, and the high-speed rotational motion of the packing bed generates a hypergravity environment. The liquid flows into the eye of the rotor, impacts the rotating packing, flows through the packing in the direction of the centrifugal force applied, then flows outward, along the centrifugal force direction towards the outer edge, and is then sprayed into the housing, exiting the entire device through baffle plates. In a classical rotating packed bed, the porous packing material is generally placed inside the rotating components. Under the strong centrifugal force field, the liquid phase collides with the rapidly rotating packing material, resulting in shearing to form liquid films, liquid lines, and liquid droplets, thus increasing the gas–liquid contact area, enhancing the phase interface renewal rate, and intensifying mass transfer between phases. Compared to traditional tower equipment, the gas–liquid mass transfer coefficient in a rotating packed bed can be increased by one to three orders of magnitude [24]. Hypergravity technology greatly enhances gas–liquid mass transfer processes and liquid–liquid mixing processes; significantly reduces equipment size; lowers investment costs; improves intrinsic safety; and promotes quality improvement, efficiency enhancement, energy conservation, and carbon reduction in the chemical industry. Therefore, it is widely applied in processes such as acid gas removal [25,26,27,28], chemical manufacturing [29], alkali liquor oxidation regeneration [30], continuous distillation [31], hydrogenation [32], the preparation of nanoparticles [33], and so on.
In the classical rotating packed bed, the gas phase serves as the continuous phase, while the liquid phase acts as the dispersed phase. The dispersion and size of the liquid droplets significantly affect the performance of the rotating packed bed. Therefore, extensive experimental [34,35] and computational fluid dynamics studies [36,37] have been conducted on the interaction processes between the liquid phase and surface micro-/nano-structured packing materials.
Figure 2. Centrifugal hypergravity equipment. (a) Diagram of the rotating packed bed [38]; (b) the beam centrifuge [1]; (c) the centrifugal hypergravity apparatus [39].
Figure 2. Centrifugal hypergravity equipment. (a) Diagram of the rotating packed bed [38]; (b) the beam centrifuge [1]; (c) the centrifugal hypergravity apparatus [39].
Materials 18 00496 g002

2.3.2. Centrifuge Techniques

Centrifuge techniques involve the sample being placed in the sample compartment of the centrifuge, as shown in Figure 2b,c, rather than the liquid sample filling the entire rotating part, as in the case of the RPB techniques. By installing different onboard equipment on the centrifuge, an experimental platform is provided to elucidate the material transport mechanisms and phase interactions of multiphase media. Interdisciplinary centrifugal hypergravity experiments have been conducted in the fields of materials science and life sciences. Large centrifuges are commonly used in geotechnical engineering related hypergravity environments, with a capacity of 400 g·t [40], as shown in Figure 2b. Large-capacity centrifuges can be used for simulating the dynamic response of rock and soil infrastructure under seismic or explosive events, as well as for studying the long-term migration of pollutants, simulating dam behavior under earthquakes [41], the accumulation-compaction process of mid-ocean ridge minerals, calculating magma migration rates in the region [42], and researching the effects of gravity in life sciences [43].
While small-capacity centrifuges are used in solid-phase materials science for hypergravity research, as shown in Figure 2c, this is the main tool used in current research studies on materials science performed in hypergravity environments, which we will discuss in detail in this paper. The centrifugal motion of the loading chamber in the centrifuge creates a hypergravity environment for the samples contained within it, allowing for targeted scientific experiments by applying different conditions to the sample chamber. For example, common practices include resistive heating of the sample chamber to conduct high-temperature hypergravity experiments. The structure and materials of the reactor can be designed accordingly based on the experimental objectives. For example, solid samples can be used to study solid-phase diffusion and structural evolution experiments; crucibles can be used for sedimentation, solidification, separation, and enrichment experiments of different phases; and crucibles with microporous filtration plates can be used for thermal filtration and separation experiments of solid and liquid phases.

3. Liquid-Phase Synthesis and Separation of Materials Under Hypergravity

3.1. The Fluid Dynamics Under Hypergravity

In normal gravity, fluid convection is an important phenomenon that describes the motion of fluids under external forces. Turbulence is an irregular motion state in fluids, characterized by the irregular distribution of velocity and pressure [44]. Vorticity is a rotational motion in fluids, commonly observed in atmospheric and oceanic circulations. Additionally, intermolecular forces in liquids are important phenomena in fluids, including surface tension and viscosity. These phenomena interact with each other; for example, turbulence may lead to the formation of vorticity, and intermolecular forces can affect the motion state of fluids. Finally, the influence of external force fields on various phenomena in fluids is also a key focus of research. External force fields can alter the flow patterns and motion states of fluids, thereby affecting the behavior of fluid dynamics. The various phenomena and interactions in fluids and fluid dynamics constitute a complex and intriguing research field that holds significant importance for understanding fluid motion and phenomena in the natural world.
When a fluid experiences hypergravity conditions, the Fc (Coriolis force) and interaction force between phases are unavoidable, so fluid dynamics can be modulated, though the properties are unchanged [45,46]. Natural convection relies on the temperature difference or density difference of the fluid to drive the fluid flow from microscale melts and crystal growth [47,48,49,50] to macroscale meteorological convective activities [51,52], while fluid flow and interactions are enhanced under external driving forces. In hypergravity, the convection promoted by hypergravity accelerates mass transfer and relaxes the solute concentration gradient, and natural convection is usually neglected [53,54]. The Fc (Coriolis force), Fb (centrifugal force), and interaction enhance the forced convection, and the direction of forced convection is inconsistent with those of mass transfer and bubble separation. Therefore, researchers have conducted a study on the flow behavior of the fluid in the rotating bed. Burns et al. used high-frequency flash photography to capture the rotating packed bed and found that three liquid flow patterns were dominant: pore flow, liquid flow, and membrane flow, as shown in Figure 3a. Subsequently, it was also found that the form of the liquid was related to the rotational speed of the rotating packed bed: at low rotational speed (400 rpm), the liquid mostly passed through the packing zone in a radial channel flow mode; at high rotational speed (900 rpm), the liquid passed through the packing zone in a droplet state, though the flow form was not discovered, as shown in Figure 3b [55].
In a study of the microflow behavior of liquid in a rotating packed bed, Wu et al. [56] 3D printed the segment-inlet filler, combined it with CFD simulation and high-speed camera shooting, and studied the liquid form of liquid flow. As such, they found that in addition to droplets and film, a new type of liquid bridge was observed in the packing ring at lower rotational speeds or in the inner packing ring, which was consistent with the liquid content distribution of different packing rings simulated by CFD. Subsequently, with an increase in rotational speed, the ligament will be shorter, and the probability of the liquid form transforming into the liquid drop in the packing ring will be greater, which indirectly indicates that the higher the level of hypergravity, the more easily the liquid will transform into liquid drop form, as shown in Figure 4.
In addition, the liquid flow pattern in the cavity region of the rotationally packed bed (RPB) is also of great concern. Sang et al. observed two typical liquid flow patterns in the RPB cavity—ligament flow and droplet flow, as shown in Figure 5a—and observed a shift in the liquid flow pattern, i.e., the line to droplet flow transition, as shown in Figure 5b,c. This transition is caused by an increase in rotational speed or a decrease in liquid viscosity, which weakens the effect of viscous damping on the liquid, and by the liquid transitions from the ligament to the droplet [57]. This indirectly suggests that the greater the hypergravity, the greater the number of droplets transformed.

3.2. The Liquid-Phase Synthesis of Materials Under Hypergravity

Mass transfer and reaction between multiphase fluids are fundamental in industrial processes for liquid-phase material synthesis. The application of hypergravity technology can improve gas–liquid mass transfer in multiphase fluid systems by significantly increasing the rate of interphase transfer and the degree of microscopic mixing between liquid phases [55]. Packing in a hypergravity field induces significant interfacial shear that breaks down liquids into liquid films, filaments, and droplets of micron to nano unit size, continuously refreshing the phase interfaces and thus improving the micro-mass transfer and separation processes between the solution and the precipitant [58]. For example, in a rotary packed bed (RPB), the centrifugal force generated by high-speed rotation can enhance the renewal of the gas–liquid or liquid–liquid interface, thereby improving the efficiency of processes such as absorption, desulfurization, distillation, etc. [59]. Various organic and inorganic nanomaterials, such as SrCO3 [60], Mg(OH)2 [61], and ZnS [62], etc., have been successfully industrially synthesized. Qi et al. [63] used a hypergravity reactor to develop a new means of rapid synthesis of SiO2-alumina zeolites. The presence of a hypergravity force field significantly increased the gel–liquid interface area in the precursor, alleviated the diffusion resistance of zeolites in the initial stage of nucleation, and thus greatly increased the nucleation rate of zeolites and accelerated their growth. In addition, the study also found that hypergravity promoted the transition of hydrated sodium ions and OH in the liquid phase to the synthetic gel phase, further improving the nucleation efficiency of zeolites. As shown in Figure 6, DFT calculation shows that sodium hydrate can induce Si-OH dehydrogenation to form oxygen free radicals, thus reducing the energy barrier of the synthesis reaction and accelerating the nucleation process.
Similarly, Figure 7 shows the results of a study on the preparation of barium sulfate nanodispersions by precipitation in a RPB, which found that the overall particle size was about 10 to 17 nm according to the transmission electron microscope (TEM) analysis, and the product prepared using RPB had a smaller particle size and a narrower particle size distribution than the barium sulfate prepared in a traditional stirred tank reactor (STR) [64].

3.3. The Separation of Materials Under Hypergravity

Starting with the commonalities among the separation and purification of materials, the basic theories of the equilibrium separation and distillation of substances in a hypergravity environment have been established [65,66]. Combined with the process intensification and coupled separation behaviors, a complete set of process technologies, such as gas purification, gas–solid separation, liquid–liquid separation, and controlling and separating pollutants in water, can be developed using hypergravity technology [67]. In the process of smelting slag at a high temperature, the metal droplets brought about by the violent reaction will be mixed with the slag, thus discharging the slag and causing waste. Valuable metals in the slag in the form of droplets gather with a difficult and diffuse distribution, which is mainly due to the droplets having high surface tension and the buoyant force being too small.
Under hypergravity conditions, the distinct buoyancy forces result in differing distribution trends for the two phases, thereby achieving phase separation, which is known as centrifugal separation. The increased acceleration and enhanced convective effects accelerate the migration speed between phases of different densities, leading to an enhanced separation rate. In contrast, under normal-gravity conditions, phase separation is driven by diffusion and buoyancy effects, with a relatively slower speed. Therefore, hypergravity technology can effectively realize the separation of the liquid phase.
In the preparation process of high-end metal materials, the presence of non-metallic inclusions will destroy the uniform distribution of the substrate, thus affecting the comprehensive properties of the material. Traditional refining processes, such as electromagnetic stirring, gas stirring, etc. [68,69], have limited purification capacity in liquid metal, and it is difficult to completely remove inclusions. Ideally, as the gravity coefficient increases, the inclusions will rise to the top of the sample, as shown in Figure 8, where the SiO2 inclusions in 304 stainless steel are evenly distributed under normal gravity, while the volume fraction and quantity density of SiO2 are distributed in gradient after the introduction of hypergravity, causing them to migrate to the top of the sample along the reverse direction of hypergravity [70].
Similarly, studies on the removal of inclusions by hypergravity in aluminum alloys and copper alloys show that the removal of inclusions by hypergravity is very effective [71,72]. As shown in Figure 9, the inclusions in molten-state Inconel 718 superalloys were concentrated on the top of the sample under the treatment of different gravity conditions, and there was almost no bottom of the sample. The size distribution of the inclusions under normal-gravity conditions (G = 1) was relatively uniform, concentrated in 2~4 μm, while in the hypergravity condition (G = 70~210), the average size of the inclusions on the top of the sample was 7~9 μm, which was much larger than that of the middle and lower part of the sample [73]. It can be seen that the main task in melting recycling alloy materials is to remove inclusions, and the separation processes of hypergravity settling/uplifting or filtration are efficient means of doing so.
Hypergravity technology plays an important role in the enrichment and separation of valuable elements. In the selective separation, enrichment, and recovery of valuable metals in non-ferrous metal ores, it is important to clarify the reaction and separation behavior of multi-metal components to achieve efficient separation. By utilizing the heterogeneous mass transfer and strong phase migration of the hypergravity field, electronic waste is placed in a graphite crucible and a filter, as shown in Figure 10a, for heating and applying hypergravity. The phases with different melting points will be sequentially separated to the bottom, as shown in Figure 10b [39,74].
For example, in waste printed circuit boards (WPCBs), there are valuable metals such as gold and silver, and failing to recover them will lead to serious soil pollution problems. Traditional methods can utilize the difference in vapor pressure of various metals to purify the processed granular metal by combining evaporation and condensation technology (such as Pb and Zn) [75]. Low-temperature alkaline smelting technology can also be used to extract valuable metals from mixed-state metals, which has a good purification effect on amphoteric metals such as Sn, Pb, Al, and Zn [76]. However, these two methods are not suitable for the purification of other metals in WPCBs. When a mixed-metal circuit board is heated to a molten state under normal gravity, the surface tension between different metals hinders their separation. Under hypergravity conditions, however, various metals can be separated according to their different melting points.
Chen et al. [75] used the pyrolysis method to manually dissociate the mixed metal and glass debris of WPCBs and then heated the mixed metal to 1400 °C under 1000× g hypergravity conditions. The ingot along the hypergravity direction had Fe-rich, Cu-rich, Cu-rich, and Pb-rich layers, and the separation rate between the mixed metals could reach more than 94%, as shown in Figure 11, This suggests that hypergravity is conducive to the separation of different metals.
The advantages of hypergravity in accelerating the separation of substances with different densities have been extensively utilized in isotope separation, significantly contributing to the advancement of science and technology [77]. For instance, the gas-phase centrifuge separation of nickel isotopes 62Ni and 64Ni, which form gaseous molecules at near-ambient temperatures, has expedited the development of medical technology [78]. However, isotopes that cannot be vaporized at ambient temperatures, such as group I (alkali metals) and II (alkaline earth metals) elements, are not suitable for gas-phase separation techniques. Notably, the liquid centrifugal method can be employed to separate and extract isotopes of almost all the elements. This method operates under conditions close to ambient temperature and pressure, avoiding the use of harmful gases. The heavier isotopes concentrate at the outer edge of the centrifuge, while the lighter isotopes accumulate at the inner edge [79]; such a phenomenon is also observed in the solid phase [80]. Nevertheless, there are challenges associated with this process; centrifugal separation equipment must withstand the immense centrifugal force generated by high-speed rotation, requiring materials with exceptional strength and corrosion resistance.

4. Metallurgy of Alloys Under Hypergravity

4.1. Convection and Segregation Under Hypergravity

The solidification of the alloy melt causes differences in the composition of the solute, meaning the concentration gradient will cause the natural convection of the melt. Solidification and melt flow are strongly coupled, and this is reflected in the solute removal during the crystallization process of the primary crystal, which causes solute pair flow at the solid–liquid interface, and the melt convection in turn affects the growth rate and direction of the crystal [81,82]. In addition, the morphology of the grains during crystal growth will be affected by the melt flow, which is mainly caused by the solute pair flow [83,84]. The influence of alloy melt convection on the solidification process under a normal-gravity field has been extensively studied, as well as in microgravity or hypergravity field experiments [85,86,87,88,89]. The alloy melt will show different flow states under the action of different gravity fields. The convective instability of the liquid phase can be measured according to a physical quantity, namely Rayleigh number Ra, which is expressed as outlined in Equation (7) [5]:
R a = β g L 3 ν k
where g is the acceleration of gravity, β is the isobaric thermal expansion coefficient, ν is the kinematic viscosity, κ is the thermal diffusivity, Δ is the temperature gradient between the hot and cold plates, and L is the thickness of the fluid layer between the above plates. The melt does not flow when the Rayleigh number is small. When the Rayleigh number exceeds a critical value, the buoyancy overcomes the viscous force, and natural convection occurs. Under the conditions of natural convection, when the Raleigh number is small, it is a laminar flow. With an increase in Rayleigh number, the change from laminar flow to turbulent flow will occur [90]. In the hypergravity field, the Rayleigh number further increases, and the convection state returns from turbulence to laminar flow, also known as “re-laminarization”. However, under the conditions of centrifugal force and buoyancy introduced in the hypergravity force field, the Coriolis force generated by rotation around the centrifugal axis will change the flow state of the melt. In particular, when the Coriolis force is coupled with a stable axial temperature gradient, it mainly redistributes the convection in the melt [91,92], while in the case of an unstable temperature gradient, the Coriolis force promotes the basic rearrangement of the flow [93]. At present, scientists use the Rossby dimensionless number Ro to evaluate the influence of the Coriolis force on convection and describe the influence of the Coriolis force in the centrifugal rotating system. Under a hypergravity field, convection is strengthened, and Rossby, Ro, is expressed as outlined in Equation (8) [94]:
R O = U Ω L = ν G r 1 2 Ω L 2 ,
where ν is the kinematic viscosity of the liquid, U is the characteristic velocity, Ω is the rotational angular velocity, L is the characteristic length, and Gr is the Grashof number. In general, the Rossby number is greater than 1, indicating that the influence of the Coriolis force can be ignored.
When the Rossby number is much less than 1, the Coriolis force completely determines the flow pattern of the fluid. Cisternas Fernández et al. [95] showed the deflection effect of the Coriolis force on alloy melts in experiments where all Rayleigh numbers were much lower than 1. In this case, the Coriolis force pushes the lighter liquid toward the sample’s flight speed, resulting in a large circular backflow throughout the entire liquid region. However, in Battaglioli’s dimensionless analysis [96], Coriolis forces increase proportionally to the ratio of buoyancy brought about by hypergravity sites, and their magnitude is negligible relative to the buoyancy caused by centrifugal acceleration.
The metallurgical process that materials undergo is a complex liquid–solid phase transition reaction, accompanied by a series of physical and chemical reactions such as heat loss, melt convection, and solute redistribution [97]. The solidification interface is affected by heat or solute redistribution, resulting in dendritic growth during solidification [98]. The distribution of the solute may lead to the formation of harmful casting defects such as “freckles” in the microstructure of metal alloys [99]. These “freckles” have been observed in many materials, such as directionally solidified turbine blades, cast ingots, etc. The convection driven by the buoyancy of the hot melt of an alloy liquid will lead to the uneven distribution of solute between solid and liquid phases in a specific thermodynamic state, which is one of the reasons for the segregation. The segregation behavior of the solidification process is often accompanied by a gradual change in composition. Numerous studies have shown that segregation and melt convection have a direct effect on the formation of casting defects [100,101,102].
When the gravitational acceleration is much less than the Earth’s gravitational acceleration (g ≈ 9.81 m/s2), the environmental state is called a microgravity field. At this point, the gravitational effect on the object is negligible, putting the object in a near-weightless state. The buoyancy effect of gravity causes solid particles to settle or float during solidification [103]. The magnitude of gravity makes it difficult to understand the equiaxed crystal transformation that occurs during solidification, while natural convection and sedimentation effects are greatly weakened or disappear in microgravity [104,105]. During solidification, solute diffusion and crystal growth are mainly controlled by the diffusion process rather than convection [106,107]. This is conducive to improving the nucleation and growth conditions of equiaxed crystals. For example, in the dendrite growth of Ni-50%Cu supercooled melt under microgravity and with no container, the crystal growth morphology changes from coarse dendrites to uniform equiaxed crystals with an increase in the supercooling degree [108].
The buoyancy drive of the centrifugal casting process will be affected by the acceleration of the non-inertial frame, which is due to the combined effect of centrifugal acceleration and the Coriolis force. In the hypergravity field, there is a certain centrifugal force gradient in the alloy melt in different regions of the rotating packed bed, which may aggravate the flow behavior, make the fluid state transition from laminar flow to turbulent flow mode, and finally cause it to reverse back to laminar flow mode [109]. When centrifugal casting turbine blades of high-Nb TiAl alloy [110], it was found that dendrite γ segregation occurred at the edge of the casting surface due to rapid cooling. This is because hypergravity increases the equivalent interfacial heat transfer coefficient between the alloy melt and the die, increasing the cooling rate of the surface [111].
High temperatures affect the diffusion rates, chemical reaction rates, and thermodynamics within materials, while hypergravity influences interphase interaction forces and stresses. Therefore, for certain material systems, there may exist specific temperature ranges where these effects are most pronounced. Under a hypergravity force field, convection and segregation will result in the inhibition of nucleation and precipitation of the alloy. As shown in Figure 12, a microstructural analysis conducted using scanning electron microscopy (SEM) reveals the morphology and variations in area fraction of the precipitated phase δ at the grain boundary of GH4169 superalloy after a centrifugation experiment. It was discovered that [112] under normal gravity of 1× g, as the temperature increases (720–800–850 °C), the amount of δ phase precipitation of GH4169 alloy gradually rises, and the morphology changes from rods to large-angle needles. At 50,000× g, the distribution and area fraction of the δ phase decrease. The precipitation of the δ phase is inhibited by the hypergravity level, and the inhibitory effect intensifies with the increase in aging temperature. The mechanism of inhibition is that hypergravity reduces the possibility of lattice distortion and internal accumulation fault during the precipitation of the γ phase from the matrix, thus reducing the nucleation site of the δ phase.

4.2. Grain Refinement of Solidification

As we all know, the grain size of metal materials significantly affects their mechanical and processing properties: the finer the grain, the higher the strength of the material. The grain refinement process of cast parts has been widely reported in previous studies. Hypergravity has a non-negligible role in the solidification process. As shown in Figure 13a, the influence of the hypergravity coefficient on the grain size during the solidification process of pure aluminum, under the same temperature conditions, shows that the larger the hypergravity coefficient, the more significant the effect of grain refinement [113]. Yang [114] studied the solidification process of Cu-Sn alloy under hypergravity conditions. The results show that the higher the gravity coefficient, the more significant the refinement effect on the solidification structure. Hypergravity causes the solidification structure to change from coarse columnar crystals and developed dendrites to fine, equiaxed crystals and spherical crystals, as shown in Figure 13b. It is proposed that hypergravity affects grain refinement not during the melt or grain growth stages but during the nucleation stage. Zimmermann et al. [115] used transparent model alloy C5H12O2-C10H16O to observe the increase in nucleation rate and the precipitation of dendritic crystals during solidification under a hypergravity field.
During metal solidification under normal-gravity conditions, the gravity difference between the crystalline droplets and the molten liquid phase is minimal, resulting in very slow movement between the two phases. The presence of the hypergravity field makes the high-density metal particles overcome the resistance caused by the liquid phase. A large gravity difference between the two phases causes the crystalline particles to continue moving in the direction of hypergravity, resulting in the phenomenon of “crystal rain” [116] in the molten metal, which promotes the proliferation of crystal nuclei and further refines the grains. As shown in Figure 14, hypergravity-field-induced grain refinement occurs in the early stages of solidification, with the fine-crystal region gradually moving in the opposite direction to hypergravity in the later stages of solidification.
In addition, dendrites generated during the casting process form new nucleated nuclei, which are reinforced by the intervention of external fields, such as electromagnetic fields and ultrasonic waves [117,118]. The strong convection phenomenon caused by the hypergravity field will also locally alter the subcooling degree of the melt (dendrites), resulting in changes to the nucleation mode and growth rate of the liquid phase during crystallization [119]. Phase-field simulations of mono-crystalline and polycrystalline materials also demonstrate that hypergravity can promote dendrite growth direction and fragmentation [120]. As shown in Figure 15, an X-ray system was used to observe the changes in the solidification process of Al-20 wt% Cu. It was found that the dendrites broke along the solid–liquid interface, and the fragments floated from the bottom to the top due to buoyancy. The dendrite fragments melted rapidly in the hot region of the specimen until they disappeared completely. When hypergravity reaches 1.8 g, more equiaxed grain nucleation appears at the front of the columnar crystal [121]. The increase in heterogeneous nucleation sites caused by dendrite fragmentation plays a key role in grain refinement.
You et al. [122] found that in centrifugally cast Al7050 aluminum alloy, the hypergravity force field enhanced the weak parts of dendrites (Figure 16a, blue arrow), resulting in dendrite fragments of different sizes, which were used as the sites of heterogeneous nucleation, thus improving nucleation efficiency. Hypergravity causes the initial crystal and dendrite debris to settle at the bottom (Figure 16b, green arrow). The broken dendrites are recognized as the new nucleation sites in the mechanism of grain refinement under the strong convection conditions brought about by the hypergravity.
However, the specific effects of a hypergravity field on dendrite morphology and size during nucleation, including but not limited to the initial formation of dendrite, growth direction, branching characteristics, etc., still need to be further studied.

4.3. The Dendrite Structure of Alloy Under Hypergravity Conditions

In the actual process of crystal growth, the intervention of an external field can control the growth form of the crystal, its evolution can lead to the formation of a cellular interface due to the instability of the crystal surface, and further instability will cause dendrite growth. Moreover, dendrite growth in alloy materials can produce micro-segregation and other internal defects in castings and ingots, thus affecting the performance of castings [123]. For example, the primary dendrite arm spacing affects the hardness of the casting [124]. The hypergravity field simulated in the centrifugal casting process exacerbates the buoyancy-driven flow behavior and thus alters the alloy melt mass and heat transfer processes. When the solidification of molten metal in a hypergravity field occurs at certain nucleation points, the growth of dendrites and the morphology of the solidified structure are related to heat and mass transfer. The pressure on the liquid in a hypergravity field is anisotropic, which may cause the anisotropy of heat and mass transfer, which in turn affects the growth of dendrites and the morphology of solidified tissues; e.g., when the direction of hypergravity is opposite to the upward-growing dendritic dendrites, the primary dendritic arm spacing (PDAS) decreases with an increase in the hypergravity level [125]. As shown in Figure 17, phase-field simulation in Al-Cu alloys reveals that the dendrite growth morphology is related to the magnitude and direction of gravity. Under positive gravity, the primary dendrites grow longer with an increase in hypergravity, the growth of other secondary dendrites becomes shorter and shorter, and the growth of bottom dendrites is completely inhibited at +50 g. Under negative gravity, the increase in hypergravity makes the length of primary dendrites shorter and shorter, while the number of branches of secondary dendrites and other dendrites increases [126]. The change in the direction of positive gravity is related to the melt flow pattern caused by hypergravity. In the enhanced convection mode, the flow caused by the density difference is downward, and the latent heat of the melt is released to the bottom, making the concentration and temperature gradients larger, and thus promoting the growth of the dendrite tip [127]. It is worth noting that in the centrifugal casting of Ti-Al alloys, the effect of hypergravity on the initial dendrites obtained from numerical simulations is different [128]. Therefore, the processes related to dendrite growth during solidification under a hypergravity field need more reliable experiments to validate the simulation results.
In the experiment, You et al. [122] studied the centrifugal casting process of Al7050 aluminum alloy and found that obvious microstructure changes occurred along the direction of hypergravity, resulting in different types of dendritic regions (I-II-III-IV regions, as shown in Figure 18a; IIT and IIB are, respectively, the top and bottom II structures). It was found that type-II columnar dendrites appear under hypergravity and grow gradually with an increase in the hypergravity coefficient. As shown in Figure 18b,c, from the area and elemental distribution of the second phase between dendrites, it can be observed that the strong convection caused by hypergravity will push elements, such that the concentration gradient at the tip of the dendrites becomes too large, accelerating the dendrite growth process. Once the solute concentration reaches dynamic equilibrium, it inhibits the further growth of the columnar crystals, resulting in a type-IV structure with a higher secondary phase area fraction.

4.4. The Gradient Structure Under Hypergravity Conditions

For fast sample preparation and rapid optimization of performance, conventional metal materials can be specially prepared as gradient functional materials with gradient changes in composition or properties. Gradient materials with continuous variations in composition, structure, and properties in specific directions require one of the methods of high-throughput preparation [129,130]. Laser fusion deposition [131], self-propagating high-temperature synthesis [74], chemical vapor deposition [132], and many other methods have been successful in preparing gradient functional materials, and hypergravity technology has been equally effective [133]. As shown in Figure 19, in a hypergravity field of 500 g, the microstructure of the alloy undergoes a large change in phase morphology along the direction of hypergravity [134], and the apparent gradient structure is due to the effect of the gradient distribution of trace alloy oxides on nucleation during centrifugation. As the gravity field reaches 60,000× g, an obvious crystalline phase delamination phenomenon occurs in the Mg56Al30Li7Cu7 metallic glass organization, and the high-gravity region is replaced by a new copper-rich binary eutectic organization, which reveals the solidification sequence in multicomponent alloys under a hypergravity field and provides a guideline for the development of new bulk metallic glass materials [135].

5. Structural and Properties of Alloy Materials Under Hypergravity

5.1. Solid-State Diffusion Under Hypergravity Conditions

Elemental diffusion is a phenomenon of atomic migration in solid materials. This process is based on Gibbs free energy or chemical potential [136]. Under normal-gravity conditions, the most common diffusion mechanism in solid-phase alloy systems is vacancy diffusion, i.e., when atoms migrate and diffuse through vacancies in the crystal lattice, a mechanism that has been confirmed by numerous experiments, including the famous Kirkendall void phenomenon [137]. In a normal-gravity field, the diffusion coefficient is related to the product of the vacancy concentration and its jump probability [138]. Under the applied stress field, diffusion parameters such as the lattice constant, hopping mechanism, and vacancy activation energy will change, thus changing the diffusion behavior [139,140]. When applying conventional stress conditions under normal gravity, Cowern [141] and Kringhøj [142] proposed a stress-based pair. The vacancy diffusion mechanism model outlines the influence of atomic diffusion activation energy. The basic concept of this model is that when the material is subjected to tensile stress, the activation energy for atomic diffusion increases, and the atomic diffusion is inhibited. In the case of compressive stress, the change in activation energy is opposite due to the decrease in diffusion activation enthalpy, which promotes atomic diffusion.
The hypergravity environment differs significantly from the conventional stress environment experienced under normal-gravity conditions. Early research involved fundamental studies on the effects of high-temperature hypergravity environments on solid-phase materials, conducted by simulating hypergravity using devices such as centrifuges with integrated heating. These studies revealed that hypergravity fields influence the diffusion of elements within solid-phase alloys. For instance, investigations of gold–potassium (Au-K) alloys at different hypergravity levels (achieved by changing the speed of the centrifuge) have shown that gold atoms with larger atomic masses exhibit faster diffusion rates [143].
This phenomenon is attributed to the centrifugal force induced by hypergravity, which increases the chemical potential of the solution composition and promotes the deposition of heavier gold atoms. Atomic deposition with a higher atomic mass has also been confirmed in selenium–tellurium (Se-Te) alloys [144], such as gradient structures with different lattice constants formed along the direction of hypergravity, as shown in Figure 20. Therefore, hypergravity enhances the atomic diffusion of heavier species.
Studies of the interface between copper (Cu) and brass (CuZn) alloys [145,146] show that hypergravity affects the distribution of vacancy concentrations in the material and thus the diffusion of elements in the alloy, as shown in Figure 21. It can be seen that the diffusion of heavier atoms can be enhanced under a hypergravity field, which does not occur under normal-gravity conditions. It is worth noting that hypergravity is a combination of centrifugal force, buoyancy force, and the Coriolis force, so the influence of hypergravity on diffusion cannot be simplified to traditional tensile stress or compressive stress, and the influence of centrifugal force and buoyancy on the diffusion system of binary alloys must be further studied.
Some studies have evaluated the reliability of this simplification. By studying CuZn diffusion pairs treated under hypergravity, they found that the mutual diffusion behavior of an intermetallic compound (IMC) layer had directional effects: the hypergravity of G+ samples effectively inhibited the formation of diffusion void defects, but this effect was weakened at 4700 g, as shown in Figure 22 [147]. It was thus revealed that the centrifugal force and buoyancy caused by hypergravity significantly affect the diffusion characteristics of the CuZn binary system. Centrifugal force enhances interfacial contact and promotes interfacial diffusion, while buoyancy promotes the diffusion of vacancies and atoms, resulting in the directional effects and structural behavior observed under hypergravity.

5.2. Structural Evolution of Alloy Materials Under Hypergravity Conditions

The evolution of material phases is closely linked to solid-phase diffusion, and hypergravity also influences the phase transformation of alloy materials, such as the formation of intermediate metal compounds. Bartek Wierzba investigated hypergravity’s impact on a material’s pressure and density. He proposed a self-consistent coupled diffusion equation to illustrate the influence of hypergravity on the two-phase growth rate of Cu6Sn5 and Cu3Sn in Cu-Sn diffusion couples, both theoretically and experimentally [148]. Qiao et al. [149] studied the diffusion behavior of a CuSn alloy under high temperature and hypergravity simulated by high-speed rotation and found that the growth of Cu3Sn in CuSn diffusion couples under hypergravity was inhibited, as shown in Figure 23. At 210 °C, hypergravity inhibits the diffusion between CuSn couples by inhibiting the formation of the intermetallic compound Cu3Sn and reducing the number of Kirkendall holes in Cu3Sn layer.
Simultaneously, it is essential to investigate the behavior of defects under hypergravity to complement the exploration of element diffusion mechanisms and phase evolution in solid materials. Defects present in solid crystalline materials, such as vacancies, dislocations, layer faults, twins, and grain boundaries, not only influence element diffusion and microstructure evolution but also significantly impact the mechanical properties of materials. Under normal-gravity conditions, differential diffusion rates lead to the formation of Kirkendall pores, which impede diffusion and contribute to material failure. Dislocation movement, stratification, and twin crystal formation, as well as grain boundary deformation, are critical factors influencing the atomic migration/diffusion and mechanical properties of materials. Compared to their behavior under traditional stress, the thermal diffusion and microstructure evolution of Cu-Zn diffusion couples show significant differences when subjected to hypergravity conditions. Xie et al. [147] conducted an atomic-scale mechanical study on the evolution of defect behavior under the action of hypergravity and found that the β phase, when generated under different gravity conditions, such as dislocation and other defects, decreases with an increase in hypergravity, as shown in Figure 24, especially under 4700 g hypergravity. The decrease in defect density represents a decrease in vacancy concentration, and the vacancy diffusion mechanism is inhibited.
The above studies fully reflect that the diffusion/movement behavior of atoms and defects (vacancy, dislocation, grain boundary) in the material will be different from the behavior under normal-gravity conditions, and the solid phase diffusion behavior and physical mechanism of the material extending from the binary alloy system to the multicomponent alloy system under hypergravity need to be further studied. It will be helpful to understand the evolution of microstructure, mechanical behavior, and properties of materials under extreme conditions.

5.3. The Properties of Alloy Materials Under Hypergravity Conditions

An alloy material under long-term high-speed rotating conditions is affected by the hypergravity field, which easily causes a change in the alloy structure, and the change in the alloy structure under extremely high-temperature conditions is particularly obvious after a change in the rotating device. For example, over-burnt aero-engine high-pressure turbine blades during high-speed rotation will greatly change the diffusion mode and path, change the microstructure, and even recrystallize, resulting in blade failure and engine failure [150,151,152].
For high-speed rotating systems, the traditional study of mechanical properties often uses a unified constant external force (such as creep) or different frequencies of periodic constant force (such as fatigue) loading conditions to study the effect of external forces on metals. However, due to the force transfer effect of the rigid body, the centrifugal hypergravity field has a gradient stress from the core to the outer end; that is, the magnitude of the hypergravity at different positions is different, and the structure under the coupling of temperature and stress will also be different. For typical rotating parts such as high-pressure turbine blades made of superalloys, different temperatures and external forces lead to different structural changes. Figure 25 shows that single-temperature action can only make the microstructure of superalloys coarse, and under coupled thermo-mechanical conditions, the structural rafting structure changes and ultimately changes the mechanical properties [153].
A material’s properties, such as hardness and strength, are strongly associated with the microstructure of that material. Therefore, the effects of hypergravity on material properties should be quantitatively assessed based on physical property parameters and the corresponding microstructure morphology of the materials through mechanical property measurements and corresponding microstructural analyses (e.g., SEM, TEM, EBSD). And these results can then be compared with those obtained under normal-gravity conditions. Among these analyses, the examination of the material’s microstructure and the corresponding distribution of defects such as dislocations has proven to be the most effective, as conventional material strengthening mechanisms are closely related to the interaction between microstructure and defects such as dislocations.
In the previous study mentioned, there was a lack of high-temperature and hypergravity environment equipment to dynamically simulate the extreme environment under which the service conditions of alloy materials could be assessed in situ. Moreover, most characterization studies have been conducted only at the micron scale, and the analytical problems have also been limited by the diffusion of elements and the growth of corresponding intermediate metal compounds. The lack of research on basic problems related to hypergravity leads to an insufficient understanding of the influence of hypergravity. As a result, in the field of applied research, the influence mechanism of hypergravity on the microstructure evolution, deformation and failure mechanism, and creep properties of nickel-based single-crystal superalloy materials in service remains unclear, and the correlation between microstructure evolution and properties and material damage evaluation methods under hypergravity environment is also ignored.
Based on this, some researchers have used dog bone samples to study the failure behavior of gradient hypergravity on 1060 pure aluminum. By comparing the constant force and the hypergravity force, as shown in Figure 26 [154], it has been observed that crack morphology changes under hypergravity conditions. The impact of hypergravity on crack growth is more significant than that of a constant force. At different gravity levels, the crack tip exhibits plastic deformation, and the size of the pit diminishes as gravity increases. As hypergravity intensifies, both the crack area and the number of pits increase. At the same temperature, the stress required for the sample to fracture under normal-gravity conditions and constant stress is 5–10 times that in a hypergravity field, and with an increase in the level of hypergravity, the fracture temperature of 1060 pure aluminum also decreases.
Similarly, the breaking strength of 7075 series aluminum alloy under normal-gravity conditions is about 10–18 times greater than that under hypergravity at the same temperature, and the differential mechanism of the cracking of alloy materials under different gravity levels/stress conditions was discussed [155], as shown in Figure 27. Under hypergravity conditions, the specimen is affected by the coupling of normal hypergravity and tangential torsional force. The torsional force splits the extended grain along the vertical direction, weakens the grain boundary, and accelerates the grain thinning at the same time. According to a geometrically necessary dislocation (GND) analysis using the electron backscattered diffraction (EBSD) method, this causes a large number of dislocations to cluster at the upper end, resulting in the vertical crack shape changing into a zigzag crack.
The above results demonstrate that hypergravity affects the evolution of a material’s structure and its properties. Specifically, for solid alloy materials subjected to a hypergravity environment, dislocations within the microstructure tend to migrate towards regions of lower hypergravity, thereby accelerating the material’s deformation and failure processes. However, the mechanism of its effect still needs to be combined with in situ mechanical transmission electron microanalysis technology to further study the influence mechanism of hypergravity on the movement of defects in materials, the crystal deformation mechanism, the initiation and expansion of microstructure defects, and material instability. In particular, the atomic-scale mechanism of the structure and property evolution of alloy materials under the coupled environment of high temperature and hypergravity has been expounded, and the similarities and differences between the effects of hypergravity and normal gravity on alloy materials have been revealed.
It is worth noting that there are some limitations in the analysis of mechanical behavior in the processes involved in hypergravity services by finite element simulation. For example, the mechanical behavior of materials in hypergravity environments involves the coupling of several physical processes (such as temperature–stress coupling), and the complexity of these processes makes it difficult to establish and solve the model accurately.

6. Conclusions and Perspective

This paper explores the fundamental principles and technologies associated with hypergravity conditions, liquid-phase reaction synthesis and separation, alloy metallurgy processes, and the structural and property characteristics of alloy materials. Specifically, it encompasses a comprehensive examination of hypergravity conditions, including their definitions, characteristics, and the interphase interactions and mass transfer phenomena that occur under these conditions. The research delves into the application of hypergravity techniques, such as rotating packed beds and centrifuge technologies, to enhance the efficiency of material synthesis and separation processes in both liquid and solid phases. And we also elaborate on the dynamic behavior of fluids in hypergravity environments; the mechanisms underlying liquid-phase reaction synthesis; material separation techniques; and the impact of hypergravity on various aspects of alloy metallurgy, including convection, segregation, grain refinement, dendrite texture, and the formation of gradient structures. Additionally, we also review the effects of hypergravity on solid-state diffusion, the evolution of microstructures, and the analysis of performance and failure behavior in solid-phase alloy materials.
While significant progress has been made in materials science research under hypergravity conditions, numerous scientific and engineering challenges remain to be addressed due to the complex interactions of multiphase changes in these environments.
For instance, accurately describing interphase interactions and mass transfer in hypergravity requires a comprehensive understanding and quantification of the relationships between different phases and the mass transfer processes occurring under such conditions. Further research is necessary to explore how hypergravity technology enhances the efficiency of material synthesis and separation, including its effects on reaction rates, separation efficiency, and product quality, as well as the fundamental physical and chemical mechanisms involved. Moreover, precise modeling and forecasting of hydrodynamic properties in hypergravity environments is essential. Simulation methods such as finite element analysis should develop more accurate three-dimensional models in order to explore the effects of hypergravity on the microstructure of polycrystalline materials. In metallurgy, it is vital to further elucidate the specific mechanisms by which hypergravity affects alloy metallurgy. Studies on the influence of hypergravity on solid alloy materials should concentrate on the atomic-scale mechanisms and principles that govern changes in the properties of alloy materials under hypergravity, including the modulation of material structure and properties under hypergravity conditions, which has not been explored. Ultra-high-gravity equipment (such as >100,000× g) is limited by high construction and maintenance costs, and relatively little research has been conducted on the evolution of structural and mechanical properties in complex material systems. These topics will be addressed in future research endeavors.

Author Contributions

Conceptualization, L.X. and X.H.; methodology, Y.Z. and L.X.; software, Y.Z.; validation, Y.Z., L.X., Y.C. and X.H.; formal analysis, Y.Z. and L.X.; investigation, Y.Z. and L.X.; resources, Y.Z. and L.X.; data curation, Y.Z. and L.X; writing—original draft preparation, Y.Z. and L.X.; writing—review and editing, Y.Z., L.X., Y.C. and X.H.; visualization, Y.Z. and L.X.; supervision, L.X., Y.C. and X.H.; project administration, L.X., Y.C. and X.H.; funding acquisition, L.X. and X.H. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by High-level Talent Research Start-up Project Funding of Henan Academy of Sciences (Project No. 241827232) and the Basic Science Center Program for Multiphase Media Evolution in Hypergravity of the National Natural Science Foundation of China (No. 51988101).

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

The following abbreviations are used in this manuscript:
CFDComputational fluid dynamics
DFTDensity functional theory
EDSEnergy dispersive spectrometer
EBSDElectron back scatter diffraction
GNDGeometrically necessary dislocation
HRTEMHigh-resolution transmission electron microscope
IMCIntermetallic compound
IPFInverse pole figure
PDASPrimary dendritic arm spacing
RPBRotating packed bed
SEMScanning electron microscope
STRStirred tank reactor
TEMTransmission electron microscope
WPCBWaste printed circuit boards

References

  1. Chen, Y.; Tang, Y.; Ling, D.; Wang, Y. Hypergravity experiments on multiphase media evolution. Sci. China Technol. Sci. 2022, 65, 2791–2808. [Google Scholar] [CrossRef]
  2. Yang, Y.; Song, B.; Song, G.; Cai, Z. Effect of Super Gravity on the Solidification Structure and C Segregation of High-Carbon Steel. In Proceedings of the 8th International Symposium on High-Temperature; Springer: Cham, Switzerland, 2017. [Google Scholar]
  3. Takaki, T.; Kashima, H. Numerical investigations of stress in dendrites caused by gravity. J. Cryst.Growth. 2011, 337, 97–101. [Google Scholar] [CrossRef]
  4. Jiang, H.; Zhu, X.; Mathai, V.; Verzicco, R.; Lohse, D.; Sun, C. Controlling heat transport and flow structures in thermal tur-bulence using ratchet surfaces. Phys. Rev. Lett. 2018, 120, 044501. [Google Scholar] [CrossRef] [PubMed]
  5. Jiang, H.; Zhu, X.; Wang, D.; Huisman, S.G.; Sun, C. Supergravitational turbulent thermal convection. Sci. Adv. 2020, 6, eabb8676. [Google Scholar] [CrossRef]
  6. van der Laan, G.P.; Beenackers, A.A.C.M.; Krishna, R. Multicomponent Reaction Engineering Model for Fe-Catalyzed Fischer–Tropsch Synthesis in Commercial Scale Slurry Bubble Column Reactors. Chem. Eng. Sci. 1999, 54, 5013–5019. [Google Scholar] [CrossRef]
  7. Li, Y.-B.; Wen, Z.-N.; Xu, H.-Z.; Chu, G.-W.; Zhang, L.-L.; Chen, J.-F. Mass transfer modeling for viscous fluids in a disk-distributor rotating packed bed. Chem. Eng. J. 2023, 453, 139604. [Google Scholar] [CrossRef]
  8. Gao, J.-T.; Guo, L.; Zhong, Y.-W.; Ren, H.-R.; Guo, Z.-C. Removal of phosphorus-rich phase from high-phosphorous iron ore by melt separation at 1573 K in a super-gravity field. Int. J. Miner. Met. Mater. 2016, 23, 743–750. [Google Scholar] [CrossRef]
  9. Lan, X.; Gao, J.; Li, Y.; Guo, Z. A green method of respectively recovering rare earths (Ce, La, Pr, Nd) from rare-earth tailings under super-gravity. J. Hazard. Mater. 2019, 367, 473–481. [Google Scholar] [CrossRef]
  10. Aminzadeh, A.; Dimitrova, M.; Meiabadi, M.S.; Karganroudi, S.S.; Taheri, H.; Ibrahim, H.; Wen, Y. Non-Contact Inspection Methods for Wind Turbine Blade Maintenance: Techno–Economic Review of Techniques for Integration with Industry 4.0. J. Nondestruct. Evaluation 2023, 42, 54. [Google Scholar] [CrossRef]
  11. Xie, Y.; Zhao, J.; Fu, H.; Wei, H. Dependence of microstructural evolution on the geometric structure for serviced DZ125 turbine blades. Front. Mater. 2023, 10, 1165971. [Google Scholar] [CrossRef]
  12. Ling, D.; Shi, C.; Zheng, J.; Zhao, Y.; Chen, Y. Non-inertial effects on matter motion in centrifugal model tests. Chin. J. Geotech. Eng. 2021, 43, 226. [Google Scholar]
  13. Patuelli, C.; Tognato, R. Effect of High Gravity on the Solid-Liquid Interfacial Free Energy. In Processing by Centrifugation; Regel, L.L., Wilcox, W.R., Eds.; Springer: Boston, MA, USA, 2001; pp. 361–366. [Google Scholar]
  14. Wang, M.; Wang, Z.; Gong, X.; Guo, Z. The intensification technologies to water electrolysis for hydrogen production—A review. Renew. Sustain. Energy Rev. 2014, 29, 573–588. [Google Scholar] [CrossRef]
  15. Liu, T.; Guo, Z.-C.; Wang, Z.; Wang, M.-Y. Effects of gravity on the electrodeposition and characterization of nickel foils. Int. J. Miner. Met. Mater. 2011, 18, 59–65. [Google Scholar] [CrossRef]
  16. Wang, M.; Wang, Z.; Guo, Z. Deposit structure and kinetic behavior of metal electrodeposition under enhanced gravity-induced convection. J. Electroanal. Chem. 2015, 744, 25. [Google Scholar] [CrossRef]
  17. Wang, Z.; Yang, T.; Liu, Z.; Wang, S.; Gao, Y.; Wu, M. Mass Transfer in a Rotating Packed Bed: A Critical Review. Chem. Eng. Process. 2019, 139, 78. [Google Scholar] [CrossRef]
  18. Wenzel, D.; Górak, A. Review and analysis of micromixing in rotating packed beds. Chem. Eng. J. 2018, 345, 492–506. [Google Scholar] [CrossRef]
  19. Neumann, K.; Gladyszewski, K.; Groß, K.; Qammar, H.; Wenzel, D.; Górak, A.; Skiborowski, M. A guide on the industrial application of rotating packed beds. Chem. Eng. Res. Des. 2018, 134, 443–462. [Google Scholar] [CrossRef]
  20. Ramshaw, C.; Mallinson, R.H. Mass Transfer Process. U.S. Patent 4283255A, 11 August 1981. [Google Scholar]
  21. Tung, H.-H.; Mah, R.S. Modeling Liquid Mass Transfer In Higee Separation Process. Chem. Eng. Commun. 1985, 39, 147–153. [Google Scholar] [CrossRef]
  22. Wang, G.; Xu, Z.; Ji, J. Progress on Higee distillation—Introduction to a new device and its industrial applications. Chem. Eng. Res. Des. 2011, 89, 1434–1442. [Google Scholar] [CrossRef]
  23. Chen, W.-H.; Syu, Y.-J. Hydrogen production from water gas shift reaction in a high gravity (Higee) environment using a rotating packed bed. Int. J. Hydrogen Energy 2010, 35, 10179–10189. [Google Scholar] [CrossRef]
  24. Jiang, L.; Chu, G.-W.; Liu, Y.-Z.; Liu, W.; Wen, L.-X.; Luo, Y. Preparation of cordierite monolithic catalyst for α-methylstyrene hydrogenation in a rotating packed bed reactor. Chem. Eng. Process. Process. Intensif. 2020, 150, 107882. [Google Scholar] [CrossRef]
  25. Jiao, W.; Yang, P.; Qi, G.; Liu, Y. Selective absorption of H2S with High CO2 concentration in mixture in a rotating packed bed. Chem. Eng. Process. Process. Intensif. 2018, 129, 142–147. [Google Scholar] [CrossRef]
  26. Zhang, L.; Wu, S.; Gao, Y.; Sun, B.; Luo, Y.; Zou, H.; Chu, G.; Chen, J. Absorption of SO2 with calcium-based solution in a rotating packed bed. Sep. Purif. Technol. 2019, 214, 148–155. [Google Scholar] [CrossRef]
  27. Chu, G.-W.; Fei, J.; Cai, Y.; Liu, Y.-Z.; Gao, Y.; Luo, Y.; Chen, J.-F. Removal of SO2 with Sodium Sulfite Solution in a Rotating Packed Bed. Ind. Eng. Chem. Res. 2018, 57, 2329–2335. [Google Scholar] [CrossRef]
  28. Freguia, S.; Rochelle, G.T. Modeling of CO2 capture by aqueous monoethanolamine. AIChE J. 2003, 49, 1676–1686. [Google Scholar] [CrossRef]
  29. Wang, B.-J.; Tang, Z.-Y.; Wang, D.; Luo, Y.; Liu, Z.-H.; Zou, H.-K.; Chu, G.-W. Process intensification of reactive extraction for hydrogen peroxide production in a rotating packed bed reactor. Chem. Eng. J. 2022, 428, 132066. [Google Scholar] [CrossRef]
  30. Zhan, Y.-Y.; Wan, Y.-F.; Su, M.-J.; Luo, Y.; Chu, G.-W.; Zhang, L.-L.; Chen, J.-F. Spent Caustic Regeneration in a Rotating Packed Bed: Reaction and Separation Process Intensification. Ind. Eng. Chem. Res. 2019, 58, 14588–14594. [Google Scholar] [CrossRef]
  31. Luo, Y.; Chu, G.-W.; Zou, H.-K.; Xiang, Y.; Shao, L.; Chen, J.-F. Characteristics of a two-stage counter-current rotating packed bed for continuous distillation. Chem. Eng. Process. Process. Intensif. 2012, 52, 55–62. [Google Scholar] [CrossRef]
  32. Wang, D.; Liu, Y.-Z.; Wang, B.-J.; Chu, G.-W.; Sun, B.-C.; Luo, Y. Process Intensification of Quasi-Homogeneous Catalytic Hydrogenation in a Rotating Packed Bed Reactor. Ind. Eng. Chem. Res. 2020, 59, 1383–1392. [Google Scholar] [CrossRef]
  33. Wu, K.; Wu, H.; Dai, T.; Liu, X.; Chen, J.-F.; Le, Y. Controlling Nucleation and Fabricating Nanoparticulate Formulation of Sorafenib Using a High-Gravity Rotating Packed Bed. Ind. Eng. Chem. Res. 2018, 57, 1903–1911. [Google Scholar] [CrossRef]
  34. Tang, Y.-Y.; Su, M.-J.; Chu, G.-W.; Luo, Y.; Wang, Y.-Y.; Zhang, L.-L.; Chen, J.-F. Impact phenomena of liquid droplet passing through stainless steel wire mesh units. Chem. Eng. Sci. 2019, 198, 144–154. [Google Scholar] [CrossRef]
  35. Wen, Z.-N.; Li, Y.-B.; Liu, W.; Luo, Y.; Zhang, L.-L.; Chu, G.-W. Flow behavior in a rotating packed bed reactor with single-layer mesh: Effect of fiber cross-sectional shape. Chem. Eng. Sci. 2022, 248, 117147. [Google Scholar] [CrossRef]
  36. Liao, H.-L.; Ouyang, Y.; Zhang, J.-P.; Zou, H.-K.; Chu, G.-W.; Luo, Y. Numerical Studies of a Liquid Droplet Impacting on Single-Layer Hydrophilic and Hydrophobic Wire Meshes. Ind. Eng. Chem. Res. 2022, 61, 7154–7162. [Google Scholar] [CrossRef]
  37. Wang, Y.; Li, Y.-B.; Su, M.-J.; Chu, G.-W.; Sun, B.-C.; Luo, Y. Liquid droplet dispersion in a rotating packed bed: Experimental and numerical studies. Chem. Eng. Sci. 2021, 240, 116675. [Google Scholar] [CrossRef]
  38. Chen, Y.-S.; Liu, H.-S. Absorption of VOCs in a Rotating Packed Bed. Ind. Eng. Chem. Res. 2002, 41, 1583–1588. [Google Scholar] [CrossRef]
  39. Meng, L.; Wang, Z.; Zhong, Y.; Guo, L.; Gao, J.; Chen, K.; Cheng, H.; Guo, Z. Supergravity separation for recovering metals from waste printed circuit boards. Chem. Eng. J. 2017, 326, 540–550. [Google Scholar] [CrossRef]
  40. Chen, Y.; Han, C.; Ling, D.; Kong, L.; Zhou, Y. Development of geotechnical centrifuge ZJU400 and performance assessment of its shaking table system. Chin. J. Geotech. 2011, 33, 1887–1894. [Google Scholar]
  41. Ng, C.W.; Li, X.; A Van Laak, P.; Hou, D.Y. Centrifuge modeling of loose fill embankment subjected to uni-axial and bi-axial earthquakes. Soil. Dyn. Earthq. Eng. 2004, 24, 305–318. [Google Scholar] [CrossRef]
  42. Connolly, J.A.D.; Schmidt, M.W.; Solferino, G.; Bagdassarov, N. Permeability of asthenospheric mantle and melt extraction rates at mid-ocean ridges. Nature 2009, 462, 209–212. [Google Scholar] [CrossRef]
  43. Van Loon, J.J. Centrifuges for Microgravity Simulation. The Reduced Gravity Paradigm. Front. Astron. Space Sci. 2016, 3, 21. [Google Scholar] [CrossRef]
  44. Crowley, C.J.; Pughe-Sanford, J.L.; Toler, W.; Krygier, M.C.; Grigoriev, R.O.; Schatz, M.F. Turbulence tracks recurrent solutions. Proc. Natl. Acad. Sci. USA 2022, 119, 2120665119. [Google Scholar] [CrossRef] [PubMed]
  45. Yeh, H. Geophysical fluid dynamics in the hypergravity field. Acta Mech. Sin. 2024, 40, 723296. [Google Scholar] [CrossRef]
  46. Nagel, T. Centrifugal hypergravitational scaling experiments of fluid convection with evaluation and correction of Coriolis effect. Phys. Fluids 2023, 35, 0149123. [Google Scholar] [CrossRef]
  47. Carruthers, J.R.; Nassau, K. Nonmixing Cells due to Crucible Rotation during Czochralski Crystal Growth. J. Appl. Phys. 1968, 39, 5205–5214. [Google Scholar] [CrossRef]
  48. Müller, G.; Schmidt, E.; Kyr, P. Investigation of convection in melts and crystal growth under large inertial accelerations. J. Cryst. Growth 1980, 49, 387–395. [Google Scholar] [CrossRef]
  49. Hurle, D. Interface stability during the solidification of a stirred binary-alloy melt. J. Cryst. Growth 1969, 5, 162–166. [Google Scholar] [CrossRef]
  50. Feigelson, R.S. Crystal growth History: Theory and melt growth processes. J. Cryst. Growth 2022, 594, 126800. [Google Scholar] [CrossRef]
  51. Zhang, G. Warming-induced contraction of tropical convection delays and reduces tropical cyclone formation. Nat. Commun. 2023, 14, 6274. [Google Scholar] [CrossRef]
  52. Hartmann, D.L.; Moy, L.A.; Fu, Q. Tropical Convection and the Energy Balance at the Top of the Atmosphere. J. Climate. 2001, 14, 4495. [Google Scholar] [CrossRef]
  53. Wang, M.; Wang, Z.; Gong, X.; Guo, Z. Progress toward Electrochemistry Intensified by using Supergravity Fields. Chem. Electro. Chem. 2015, 2, 1879–1887. [Google Scholar] [CrossRef]
  54. Gao, Y.; Qiao, F.; You, J.; Ren, Z.; Li, N.; Zhang, K.; Shen, C.; Jin, T.; Xie, K. Effect of the supergravity on the formation and cycle life of non-aqueous lithium metal batteries. Nat. Commun. 2022, 13, 1–12. [Google Scholar] [CrossRef] [PubMed]
  55. Burns, J.; Ramshaw, C. Process intensification: Visual study of liquid maldistribution in rotating packed beds. Chem. Eng. Sci. 1996, 51, 1347–1352. [Google Scholar] [CrossRef]
  56. Wu, W.; Luo, Y.; Chu, G.-W.; Su, M.-J.; Cai, Y.; Zou, H.-K.; Chen, J.-F. Liquid flow behavior in a multiliquid-inlet rotating packed bed reactor with three-dimensional printed packing. Chem. Eng. J. 2020, 386, 121537. [Google Scholar] [CrossRef]
  57. Sang, L.; Luo, Y.; Chu, G.-W.; Zhang, J.-P.; Xiang, Y.; Chen, J.-F. Liquid flow pattern transition, droplet diameter and size dis-tribution in the cavity zone of a rotating packed bed: A visual study. Chem. Eng. Sci. 2017, 158, 429. [Google Scholar] [CrossRef]
  58. Anlar, O.; Kisli, M.; Tombul, T.; Ozbek, H. A new impinging stream–rotating packed bed reactor for improvement of mi-cromixing iodide and iodate. J. Neurosci. 2003, 113, 483. [Google Scholar]
  59. Zhao, H.; Shao, L.; Chen, J.F. High-gravity process intensification technology and application. Chem. Eng. J. 2010, 156, 588–593. [Google Scholar] [CrossRef]
  60. Chen, J.-F.; Wang, Y.-H.; Guo, F.; Wang, X.-M.; Zheng, C. Synthesis of Nanoparticles with Novel Technology: High-Gravity Reactive Precipitation. Ind. Eng. Chem. Res. 2000, 39, 948–954. [Google Scholar] [CrossRef]
  61. Sun, Q.; Chen, B.; Wu, X.; Wang, M.; Zhang, C.; Zeng, X.-F.; Wang, J.-X.; Chen, J.-F. Preparation of Transparent Suspension of Lamellar Magnesium Hydroxide Nanocrystals Using a High-Gravity Reactive Precipitation Combined with Surface Modifica-tion. Ind. Eng. Chem. Res. 2015, 54, 666. [Google Scholar] [CrossRef]
  62. Chen, J.; Li, Y.; Wang, Y.; Yun, J.; Cao, D. Preparation and characterization of zinc sulfide nanoparticles under high-gravity environment. Mater. Res. Bull. 2004, 39, 185–194. [Google Scholar] [CrossRef]
  63. Qi, T.; Shi, J.; Li, Y.; Chu, G.-W.; Zhang, L.-L.; Sun, B.-C. Interfacial nanoarchitectonics for accelerated growth of zeolite via enlarging gel-liquid interfacial area using rotating packed bed. Microporous Mesoporous Mater. 2022, 342, 112112. [Google Scholar] [CrossRef]
  64. Fang, L.; Sun, Q.; Duan, Y.-H.; Zhai, J.; Wang, D.; Wang, J.-X. Preparation of transparent BaSO4 nanodispersions by high-gravity reactive precipitation combined with surface modification for transparent X-ray shielding nanocomposite films. Front. Chem. Sci. Eng. 2021, 15, 902–912. [Google Scholar] [CrossRef]
  65. Song, Z.-M.; Zhang, H. Present Situation of Producing and Quality for Bearing Steel in China. J. Iron Steel Res. 2000, 12, 59–63. [Google Scholar] [CrossRef]
  66. Reddy, A.; Zitoun, E. Tensile properties and fracture behavior of 6061/Al2O3 metal matrix composites fabricated by low pressure die casting process. J. Mater. Sci. 2011, 6, 147. [Google Scholar]
  67. Guo, J.; Jiao, W.; Qi, G.; Yuan, Z.; Liu, Y. Applications of high-gravity technologies in gas purifications: A review. Chin. J. Chem. Eng. 2019, 27, 1361–1373. [Google Scholar] [CrossRef]
  68. Cremer, P.; Driole, J. Effects of the electromagnetic stirring on the removal of inclusions of oxide from liquid steel. Met. Trans. B 1982, 13, 45–52. [Google Scholar] [CrossRef]
  69. Lou, W.; Zhu, M. Numerical Simulations of Inclusion Behavior in Gas-Stirred Ladles. Met. Mater. Trans. B 2013, 44, 762–782. [Google Scholar] [CrossRef]
  70. Guo, L.; Gao, J.; Li, C.; Guo, Z. Removal of Fine SiO2 Composite Inclusions from 304 Stainless Steel Using Super-gravity. ISIJ Int. 2020, 60, 238–246. [Google Scholar] [CrossRef]
  71. Wang, L.; Lan, X.; Wang, Z.; Guo, Z. Removal of Low-Content Impurities from Pure Al by Supergravity Combined with Semi-Solid Method. Metals 2023, 13, 1945. [Google Scholar] [CrossRef]
  72. Wen, X.-C.; Guo, L.; Bao, Q.-P.; Guo, Z.-C. Rapid removal of copper impurity from bismuth-copper alloy melts via super-gravity separation. Int. J. Miner. Met. Mater. 2021, 28, 1929–1939. [Google Scholar] [CrossRef]
  73. Shi, A.; Wang, Z.; Shi, C.; Guo, L.; Guo, C.; Guo, Z. Supergravity-Induced Separation of Oxide and Nitride Inclusions from Inconel 718 Superalloy Melt. ISIJ Int. 2020, 60, 205–211. [Google Scholar] [CrossRef]
  74. Wen, X.; Guo, L.; Bao, Q.; Gao, J.; Guo, Z. Efficient separation of lead and antimony metals from the Pb-Sb alloy by su-per-gravity technology. J. Alloys Compd. 2019, 806, 1012. [Google Scholar] [CrossRef]
  75. Chen, B.; He, J.; Xi, Y.; Zeng, X.; Kaban, I.; Zhao, J.; Hao, H. Liquid-liquid hierarchical separation and metal recycling of waste printed circuit boards. J. Hazard. Mater. 2019, 364, 388–395. [Google Scholar] [CrossRef] [PubMed]
  76. Guo, X.-Y.; Liu, J.-X. Optimization of low-temperature alkaline smelting process of crushed metal enrichment originated from waste printed circuit boards. J. Central South. Univ. 2015, 22, 1643–1650. [Google Scholar] [CrossRef]
  77. Beams, J.W.; Haynes, F.B. The Separation of Isotopes by Centrifuging. Phys. Rev. B 1936, 50, 491–492. [Google Scholar] [CrossRef]
  78. Cheltsov, A.N.; Sosnin, L.Y.; Khamylov, V.K. Centrifugal enrichment of nickel isotopes and their application to the de-velopment of new technologies. J. Radioanal. Nucl. Chem. 2014, 299, 981. [Google Scholar] [CrossRef]
  79. Wild, J.F.; Chen, H.; Liang, K.; Liu, J.; Cox, S.E.; Halliday, A.N.; Yang, Y. Liquid solution centrifugation for safe, scalable, and efficient isotope separation. Sci. Adv. 2023, 9, eadg8993. [Google Scholar] [CrossRef]
  80. Osawa, T.; Ono, M.; Esaka, F.; Okayasu, S.; Iguchi, Y.; Hao, T.; Magara, M.; Mashimo, T. Mass-dependent isotopic fractionation of a solid tin under a strong gravitational field. EPL Europhysics Lett. 2009, 85, 64001. [Google Scholar] [CrossRef]
  81. Murakami, K.; Fujiyama, T.; Koike, A.; Okamoto, T. Influence of melt flow on the growth directions of columnar grains and columnar dendrites. Acta Metall. 1983, 31, 1425–1432. [Google Scholar] [CrossRef]
  82. Boden, S.; Eckert, S.; Gerbeth, G. Visualization of freckle formation induced by forced melt convection in solidifying GaIn alloys. Mater. Lett. 2010, 64, 1340–1343. [Google Scholar] [CrossRef]
  83. Sample, A.K.; Hellawell, A. The mechanisms of formation and prevention of channel segregation during alloy solidification. Met. Trans. A 1984, 15, 2163–2173. [Google Scholar] [CrossRef]
  84. Yuan, L.; Lee, P.D. A new mechanism for freckle initiation based on microstructural level simulation. Acta Mater. 2012, 60, 4917–4926. [Google Scholar] [CrossRef]
  85. Mamiya, M.; Nagai, H.; Castillo, M.; Okutani, T. The analysis of CdTe solidification in absence of thermal convection via short-duration microgravity. J. Cryst. Growth 2006, 295, 209–216. [Google Scholar] [CrossRef]
  86. Polovko, Y.; Sazonov, V.; Yuferev, V. Effect of spacecraft rotation on convection and impurity segregation during the growth of crystals from a melt under microgravity. J. Cryst. Growth 1999, 198–199, 182–187. [Google Scholar] [CrossRef]
  87. Ananth, R.; Gill, W.N. Dendritic growth in microgravity and forced convection. J. Cryst. Growth 1997, 179, 263–276. [Google Scholar] [CrossRef]
  88. Li, X.; Fang, X.; Fang, Y.; He, Z.; Shang, H. Experimental study on subcooled pool boiling heat transfer under hypergravity. Int. J. Heat. Mass. Transf. 2023, 217, 124697. [Google Scholar] [CrossRef]
  89. Xu, Y.; Wang, J.; Yan, Z. Experimental investigation on melting heat transfer characteristics of a phase change material under hypergravity. Int. J. Heat. Mass. Transf. 2021, 181, 122004. [Google Scholar] [CrossRef]
  90. Weber, W.; Neumann, G.; Müller, G. Stabilizing influence of the coriolis force during melt growth on a centrifuge. J. Cryst. Growth 1990, 100, 145–158. [Google Scholar] [CrossRef]
  91. Fikri, M.; Labrosse, G.; Betrouni, M. The melt phase hydrodynamics for the “stabilized” Bridgman procedure applied under centrifugation; preliminary analysis and numerical results. J. Cryst. Growth 1992, 119, 41–60. [Google Scholar] [CrossRef]
  92. Arnold, W.A.; Wilcox, W.R.; Carlson, F.; Chait, A.; Regel, L.L. Transport modes during crystal growth in a centrifuge. J. Cryst. Growth 1992, 119, 24–40. [Google Scholar] [CrossRef]
  93. Müller, G.; Neumann, G.; Weber, W. Natural convection in vertical Bridgman configurations. J. Cryst. Growth 1984, 70, 78–93. [Google Scholar] [CrossRef]
  94. Yuferev, V.S. Coriolis Force on Melt Convection During Growth of Crystals. In A Centrifuge and Under Weightlessness, in Growth of Crystals; Givargizov, E.I., Melnikova, A.M., Eds.; Springer: Boston, MA, USA, 1996; pp. 117–127. [Google Scholar]
  95. Fernández, M.C.; Založnik, M.; Combeau, H.; Hecht, U. Thermosolutal convection and macrosegregation during directional solidification of TiAl alloys in centrifugal casting. Int. J. Heat. Mass. Transf. 2020, 154, 119698. [Google Scholar] [CrossRef]
  96. Battaglioli, S.; Robinson, A.; McFadden, S. Influence of natural and forced gravity conditions during directional columnar solidification. Int. J. Heat. Mass. Transf. 2018, 126, 66–80. [Google Scholar] [CrossRef]
  97. Jiang, P.; Gao, S.; Geng, S.; Han, C.; Mi, G. Multi-physics multi-scale simulation of the solidification process in the molten pool during laser welding of aluminum alloys. Int. J. Heat Mass. Transf. 2020, 161, 120316. [Google Scholar] [CrossRef]
  98. Zhao, Y.; Zhang, B.; Hou, H.; Chen, W.; Wang, M. Phase-field simulation for the evolution of solid/liquid interface front in directional solidification process. J. Mater. Sci. Technol. 2019, 35, 1044–1052. [Google Scholar] [CrossRef]
  99. Han, D.; Jiang, W.; Xiao, J.; Li, K.; Lu, Y.; Lou, L. Investigating the evolution of freckles into sliver defects in Ni-based sin-gle-crystal superalloy castings. Mater. Today Commun. 2021, 27, 102350. [Google Scholar] [CrossRef]
  100. Sun, Q.Y.; Ren, Y.; Liu, D.-R. Numerical investigations of freckles in directionally solidified nickel-based superalloy casting with abrupt contraction in cross section. Results Phys. 2019, 12, 1547–1558. [Google Scholar] [CrossRef]
  101. Chen, C.; Sun, J.; Diao, A.; Yang, Y.; Li, J.; Zhou, Y. On the dendrite deformation and evolution mechanism of Ni-based sup-eralloy during directional solidification. J. Alloys Compd. 2022, 891, 161949. [Google Scholar] [CrossRef]
  102. Liu, Y.; Wang, F.; Ma, D.; Yang, Q.; Xu, W.; Zhao, Y.; Dong, H.; Liu, Y.; Li, D. Freckle prediction model incorporating geometrical effects for Ni-based single-crystal superalloy components. Acta Mater. 2024, 266, 119702. [Google Scholar] [CrossRef]
  103. Nguyen-Thi, H.; Bogno, A.; Reinhart, G.; Billia, B.; Mathiesen, R.H.; Zimmermann, G.; Houltz, Y.; Löth, K.; Voss, D.; Verga, A.; et al. Investigation of gravity effects on solidification of binary alloys with in situ X-ray radiography on earth and in microgravity environment. J. Phys. Conf. Ser. 2011, 327, 012012. [Google Scholar] [CrossRef]
  104. Wu, Y.; Wang, W.; Xia, Z.; Wei, B. Phase separation and microstructure evolution of ternary Fe–Sn–Ge immiscible alloy under microgravity condition. Comput. Mater. Sci. 2015, 103, 179–188. [Google Scholar] [CrossRef]
  105. Zhang, Q.; Mo, D.; Moon, S.; Janowitz, J.; Ringle, D.; Mays, D.; Diddle, A.; Rexroat, J.; Lee, E.; Luo, T. Bubble nucleation and growth on microstructured surfaces under microgravity. npj Microgravity 2024, 10, 13. [Google Scholar] [CrossRef]
  106. Thi, H.N.; Dabo, Y.; Drevet, B.; Dupouy, M.D.; Camel, D.; Billia, B.; Hunt, J.D.; Chilton, A. Directional solidification of Al–1.5 wt% Ni alloys under diffusion transport in space and fluid-flow localisation on earth. J. Cryst. Growth 2005, 281, 654. [Google Scholar]
  107. Akamatsu, S.; Nguyen-Thi, H. In situ observation of solidification patterns in diffusive conditions. Acta Mater. 2016, 108, 325–346. [Google Scholar] [CrossRef]
  108. Wen-Jing, Y.; Chun, Y.; Xiu-Jun, H.; Min, C.; Bing-Bo, W.; Zeng-Yuan, G. Rapid dendritic growth in an undercooled Ni-Cu alloy under the microgravity condition. Acta Phys. Sin. 2003, 52, 448–453. [Google Scholar] [CrossRef]
  109. Kühnen, J.; Scarselli, D.; Schaner, M.; Hof, B. Relaminarization by Steady Modification of the Streamwise Velocity Profile in a Pipe. Flow Turbul. Combust. 2018, 100, 919–943. [Google Scholar] [CrossRef]
  110. Yang, L.; Chai, L.; Liang, Y.; Zhang, Y.; Bao, C.; Liu, S.; Lin, J. Numerical simulation and experimental verification of gravity and centrifugal investment casting low pressure turbine blades for high Nb–TiAl alloy. Intermetallics 2015, 66, 149–155. [Google Scholar] [CrossRef]
  111. Kobryn, P.A.; Semiatin, S.L. Determination of interface heat-transfer coefficients for permanent-mold casting of Ti-6Al-4V. Met. Mater. Trans. B 2001, 32, 685–695. [Google Scholar] [CrossRef]
  112. Wang, J.; Zhao, X.; You, F.; Yue, Q.; Xia, W.; Gu, Y.; Zhang, Z. Effects of high gravity on the nucleation and precipitation of δ phase of GH4169 alloy. Scr. Mater. 2024, 255, 116409. [Google Scholar] [CrossRef]
  113. Zhao, L.; Guo, Z.; Wang, Z.; Wang, M. Influences of Super-Gravity Field on Aluminum Grain Refining. Met. Mater. Trans. A 2010, 41, 670–675. [Google Scholar] [CrossRef]
  114. Yang, Y.; Song, B.; Cheng, J.; Song, G.; Yang, Z.; Cai, Z. Effect of Super-gravity Field on Grain Refinement and Tensile Properties of Cu–Sn Alloys. ISIJ Int. 2018, 58, 98–106. [Google Scholar] [CrossRef]
  115. Zimmermann, G.; Hamacher, M.; Sturz, L. Effect of zero, normal and hyper-gravity on columnar dendritic solidification and the columnar-to-equiaxed transition in Neopentylglycol-(D)Camphor alloy. J. Cryst. Growth 2019, 512, 47–60. [Google Scholar] [CrossRef]
  116. Yang, Y.; Song, B.; Yang, Z.; Song, G.; Cai, Z.; Guo, Z. The Refining Mechanism of Super Gravity on the Solidification Structure of Al-Cu Alloys. Materials. 2016, 9, 1001. [Google Scholar] [CrossRef] [PubMed]
  117. Räbiger, D.; Zhang, Y.; Galindo, V.; Franke, S.; Willers, B.; Eckert, S. The relevance of melt convection to grain refinement in Al–Si alloys solidified under the impact of electric currents. Acta Mater. 2014, 79, 327–338. [Google Scholar] [CrossRef]
  118. Wang, X.; Zhai, W.; Wang, J.; Wei, B. Strength and ductility enhancement of high-entropy FeCoNi2Al0.9 alloy by ultrasonically refining eutectic structures. Scr. Mater. 2023, 225, 115154. [Google Scholar] [CrossRef]
  119. Huang, C.; Hecht, U.; Bührig-Polaczek, A. Numerical Modeling of Melting and Columnar Solidification with Convection in a Gradient Zone Furnace in a Centrifuge. Met. Mater. Trans. B 2020, 51, 2252–2267. [Google Scholar] [CrossRef]
  120. Zhang, Z.; Hou, X.; Zhang, Y.; Wei, H.; Wang, J. Phase field simulation of solidification under supergravity. Acta Mech. Sin. 2022, 38, 122031. [Google Scholar] [CrossRef]
  121. Abou-Khalil, L.; Salloum-Abou-Jaoude, G.; Reinhart, G.; Pickmann, C.; Zimmermann, G.; Nguyen-Thi, H. Influence of gravity level on Columnar-to-Equiaxed Transition during directional solidification of Al-20 wt.% Cu alloys. Acta Mater. 2016, 110, 44–52. [Google Scholar] [CrossRef]
  122. You, F.; Zhao, X.; Yue, Q.; Wang, J.; Gu, Y.; Zhang, Z. Evolution of solidification structure and mechanical properties of Al7050 alloy under hypergravity. J. Alloy. Compd. 2024, 985, 174014. [Google Scholar] [CrossRef]
  123. Glicksman, M.E. Fundamentals of Dendritic Growth. In Crystal Growth in Science and Technology; Arend, H., Hulliger, J., Eds.; Springer: Boston, MA, USA, 1989; Volume 210, pp. 167–183. [Google Scholar]
  124. Cui, C.; Liu, W.; Deng, L.; Wang, Y.; Liu, Y.; Wang, S.; Tian, L.; Su, H. Primary dendrite arm spacing and preferential orientations of the Ni–Si hypereutectic composites at different solidification rates. Appl. Phys. A 2020, 126, 898. [Google Scholar] [CrossRef]
  125. Grugel, R.N.; Hmelo, A.B.; Battaile, C.C.; Wang, T.G. Microstructural Development in Pb-Sn Alloys Subjected to High-Gravity during Controlled Directional Solidification. In Materials Processing in High Gravity; Regel, L.L., Wilcox, W.R., Eds.; Springer: Boston, MA, USA, 1994; pp. 101–110. [Google Scholar]
  126. Zhang, Y.; Dou, R.; Wang, J.; Liu, X.; Wen, Z. Numerical simulation of the effect of hypergravity on the dendritic growth characteristics of aluminum alloys. Heliyon 2024, 10, e27008. [Google Scholar] [CrossRef]
  127. Ren, N.; Panwisawas, C.; Li, J.; Xia, M.; Dong, H.; Li, J. Solute enrichment induced dendritic fragmentation in directional so-lidification of nickel-based superalloys. Acta Mater. 2021, 215, 117043. [Google Scholar] [CrossRef]
  128. Viardin, A.; Souhar, Y.; Fernández, M.C.; Apel, M.; Založnik, M. Mesoscopic modeling of equiaxed and columnar solidification microstructures under forced flow and buoyancy-driven flow in hypergravity: Envelope versus phase-field model. Acta Mater. 2020, 199, 680–694. [Google Scholar] [CrossRef]
  129. Naebe, M.; Shirvanimoghaddam, K. Functionally graded materials: A review of fabrication and properties. Appl. Mater. Today 2016, 5, 2352–9407. [Google Scholar] [CrossRef]
  130. Bui, T.Q.; Do, T.V.; Ton, L.H.T.; Doan, D.H.; Tanaka, S.; Pham, D.T.; Nguyen-Van, T.-A.; Yu, T.; Hirose, S. On the high tem-perature mechanical behaviors analysis of heated functionally graded plates using FEM and a new third-order shear deformation plate theory. Compos. B Eng. 2016, 92, 218. [Google Scholar] [CrossRef]
  131. Ren, H.; Liu, D.; Tang, H.; Tian, X.; Zhu, Y.; Wang, H. Microstructure and mechanical properties of a graded structural material. Mater. Sci. Eng. A 2014, 611, 362–369. [Google Scholar] [CrossRef]
  132. Sasaki, M.; Hirai, T. Thermal fatigue resistance of CVD SiC/C functionally gradient material. J. Eur. Ceram. Soc. 1994, 14, 257–260. [Google Scholar] [CrossRef]
  133. Löffler, J.F.; Johnson, W.L. Crystallization of Mg–Al and Al-Based Metallic Liquids under Ultra-High Gravity. Intermetallics. 2002, 10, 1167–1175. [Google Scholar] [CrossRef]
  134. Li, R.; Wang, Z.; Guo, Z.; Liaw, P.K.; Zhang, T.; Li, L.; Zhang, Y. Graded microstructures of Al-Li-Mg-Zn-Cu entropic alloys under supergravity. Sci. China Mater. 2019, 62, 736–744. [Google Scholar] [CrossRef]
  135. Löffler, J.F.; Peker, A.; Bossuyt, S.; Johnson, W.L. Processing of metallic glass-forming liquids under ultra-high gravity. Mater. Sci. Eng. A 2004, 375–377, 341–345. [Google Scholar] [CrossRef]
  136. Wijmans, J.; Baker, R. The solution-diffusion model: A review. J. Membr. Sci. 1995, 107, 1–21. [Google Scholar] [CrossRef]
  137. Sequeira, C.; Amaral, L. Role of Kirkendall effect in diffusion processes in solids. Trans. Nonferrous Met. Soc. China 2014, 24, 1–11. [Google Scholar] [CrossRef]
  138. Tahir-Kheli, R.A. Correlation factors for atomic diffusion in nondilute multicomponent alloys with arbitrary vacancy concen-tration. Phys. Rev. B 1983, 28, 3049. [Google Scholar] [CrossRef]
  139. Mundy, J.N. Effect of Pressure on the Isotope Effect in Sodium Self-Diffusion. Phys. Rev. B 1971, 3, 2431–2445. [Google Scholar] [CrossRef]
  140. Dederichs, P.H.; Schroeder, K. Anisotropic diffusion in stress fields. Phys. Rev. B 1978, 17, 2524–2536. [Google Scholar] [CrossRef]
  141. Cowern, N.E.B.; Zalm, P.C.; Van Der Sluis, P.; Gravesteijn, D.J.; De Boer, W.B. Diffusion in Strained Si(Ge). Phys. Rev. Lett. 1994, 72, 2585–2588. [Google Scholar] [CrossRef]
  142. Kringhøj, P.; Larsen, A.N.; Shirayev, S.Y. Diffusion of Sb in Strained and Relaxed Si and SiGe. Phys. Rev. Lett. 1996, 76, 3372–3375. [Google Scholar] [CrossRef]
  143. Barr, L.W.; Smith, F.A. Observations on the equilibrium distribution of gold diffusing in solid potassium in a centrifugal field. Philos. Mag. 1969, 20, 1293. [Google Scholar] [CrossRef]
  144. Huang, X.; Ono, M.; Ueno, H.; Iguchi, Y.; Tomita, T.; Okayasu, S.; Mashimo, T. Formation of atomic-scale graded structure in Se-Te semiconductor under strong gravitational field. J. Appl. Phys. 2007, 101, 113502. [Google Scholar] [CrossRef]
  145. Mashimo, T. Self-consistent approach to the diffusion induced by a centrifugal field in condensed matter: Sedimentation. Phys. Rev. A 1988, 38, 4149–4154. [Google Scholar] [CrossRef]
  146. Wierzba, B. The vacancies formation and agglomeration under centrifugal force. Phys. A Stat. Mech. Its Appl. 2017, 484, 482–487. [Google Scholar] [CrossRef]
  147. Xie, L.; Zheng, Y.; Lu, H.; Jiao, Y.; Qu, Y.; Cai, J.; Zhai, Y.; Chen, Y.; Mao, S.; Han, X. Thermal diffusion and microstructural evolution of Cu-Zn binary system under hypergravity. Acta Mater. 2024, 268, 119790. [Google Scholar] [CrossRef]
  148. Wierzba, B.; Mashimo, T.; Danielewski, M. Competition between Chemical and Gravity Forces in Binary Alloys. High. Temp. Mater. Process. 2017, 37, 285–288. [Google Scholar] [CrossRef]
  149. Qiao, S.; Chen, Y.; An, Z.; Jiao, Y.; Li, A.; Zhai, Y.; Han, X. Hypergravity suppressed thermal diffusion at the Cu-Sn couple interface. J. Alloy. Compd. 2022, 928, 167231. [Google Scholar] [CrossRef]
  150. Guo, X.; Zheng, W.; Xiao, C.; Li, L.; Antonov, S.; Zheng, Y.; Feng, Q. Evaluation of microstructural degradation in a failed gas turbine blade due to overheating. Eng. Fail. Anal. 2019, 103, 308–318. [Google Scholar] [CrossRef]
  151. Alozie, O.; Li, Y.; Diakostefanis, M.; Wu, X.; Shong, X.; Ren, W. Assessment of degradation equivalent operating time for aircraft gas turbine engines. Aeronaut. J. 2020, 124, 549–580. [Google Scholar] [CrossRef]
  152. Fu, C.; Chen, Y.; Li, L.; Antonov, S.; Feng, Q. Evaluation of service conditions of high pressure turbine blades made of DS Ni-base superalloy by artificial neural networks. Mater. Today Commun. 2020, 22, 100838. [Google Scholar] [CrossRef]
  153. Lu, H.; Zhang, W.; Chen, Y.; Zhai, Y.; Wang, W.; Long, H.; Li, A.; Han, X. Structural degradation and elemental variations in an ex-service first-stage gas turbine blade. Mater. Charact. 2023, 196, 112596. [Google Scholar] [CrossRef]
  154. Yang, F.; Jiao, Y.; Xie, L.; Qiao, S.; Qu, Y.; Zhai, Y.; Li, A.; Chen, Y. Hypergravity prompt thermal crack in 1060 aluminium slat. Mater. Sci. Eng. A 2024, 899, 146437. [Google Scholar] [CrossRef]
  155. Jiao, Y.; Yang, F.; Niu, H.; Xie, L.; Zhai, Y.; Li, A.; Chen, Y. Hypergravity-exacerbated cracking in high-speed rotating 7075 aluminum blades. J. Mater. Res. Technol. 2024, 30, 542–551. [Google Scholar] [CrossRef]
Figure 1. Effects of hypergravity on the interaction forces of matters [1]. (a) Increasing the force and gradient; (b) strengthening the relative motion between phases.
Figure 1. Effects of hypergravity on the interaction forces of matters [1]. (a) Increasing the force and gradient; (b) strengthening the relative motion between phases.
Materials 18 00496 g001
Figure 3. Fluid morphology change in a rotating packed bed [55]. (a) The three typical liquid forms; (b) morphological differences at different rotational speeds.
Figure 3. Fluid morphology change in a rotating packed bed [55]. (a) The three typical liquid forms; (b) morphological differences at different rotational speeds.
Materials 18 00496 g003
Figure 4. The liquid flow in the MLI-RPB reactor with 3D-printed wire mesh packing was compared with CFD simulation and high-speed camera data (N = 600 rpm) [56].
Figure 4. The liquid flow in the MLI-RPB reactor with 3D-printed wire mesh packing was compared with CFD simulation and high-speed camera data (N = 600 rpm) [56].
Materials 18 00496 g004
Figure 5. Research on liquid flow behavior in the cavity region of a rotating packed bed [57]. (a) Droplet morphological changes in the cavity region at different rotational speeds; (b) liquid distribution; (c) images of liquid morphological changes. From left to right: The first and the second, the liquid membrane is gradually decomposed into droplets, the third; the droplets are separated from the ligament, and the fourth: the mother droplets are separated to form consecutive droplets.
Figure 5. Research on liquid flow behavior in the cavity region of a rotating packed bed [57]. (a) Droplet morphological changes in the cavity region at different rotational speeds; (b) liquid distribution; (c) images of liquid morphological changes. From left to right: The first and the second, the liquid membrane is gradually decomposed into droplets, the third; the droplets are separated from the ligament, and the fourth: the mother droplets are separated to form consecutive droplets.
Materials 18 00496 g005
Figure 6. Synthesis of alumina zeolite by rotating packed bed in hypergravity. The red arrow shows the direction of sodium-induced nucleation [63].
Figure 6. Synthesis of alumina zeolite by rotating packed bed in hypergravity. The red arrow shows the direction of sodium-induced nucleation [63].
Materials 18 00496 g006
Figure 7. TEM image and particle size distribution of BaSO4 nanoparticles prepared at different rotational speeds. (a) The stirred tank reactor uses a rotational speed of 500 r/min (b) RPB uses a rotational speed of 500 r/min (c) RPBuses a rotational speed of 1500 r/min (d) RPB uses a rotational speed of 2500 r/min (e) Particle size distribution [64].
Figure 7. TEM image and particle size distribution of BaSO4 nanoparticles prepared at different rotational speeds. (a) The stirred tank reactor uses a rotational speed of 500 r/min (b) RPB uses a rotational speed of 500 r/min (c) RPBuses a rotational speed of 1500 r/min (d) RPB uses a rotational speed of 2500 r/min (e) Particle size distribution [64].
Materials 18 00496 g007
Figure 8. The growth of SiO2 inclusions in 304 stainless steel after treatment for 1 min under different gravity conditions [70]. (a) Volume fraction; (b) density distribution.
Figure 8. The growth of SiO2 inclusions in 304 stainless steel after treatment for 1 min under different gravity conditions [70]. (a) Volume fraction; (b) density distribution.
Materials 18 00496 g008
Figure 9. Hypergravity-enhanced separation for the removal of inclusions from Inconel 718 superalloys [73]. (a) Macroscopic and microscopic imaging of samples before and after treatment under hypergravity; (b) density and particle size distribution of inclusions after treatment of Inconel 718 superalloys under different hypergravity conditions (A: The central area of the ingot under normal gravity; B–F: Different areas of sample cross section obtained by hypergravity facilitation separation).
Figure 9. Hypergravity-enhanced separation for the removal of inclusions from Inconel 718 superalloys [73]. (a) Macroscopic and microscopic imaging of samples before and after treatment under hypergravity; (b) density and particle size distribution of inclusions after treatment of Inconel 718 superalloys under different hypergravity conditions (A: The central area of the ingot under normal gravity; B–F: Different areas of sample cross section obtained by hypergravity facilitation separation).
Materials 18 00496 g009
Figure 10. Hypergravity separation of valuable components in mixed metals [39,74]. (a) Schematic diagram of filtration mechanism of hypergravity technology (a1): The initial phase of random distribution of antimony phase,(a2): Hypergravity is applied for separation, (a3): The phase with small particle size is separated to the bottom, black arrow: direction of phase migration); (b) Hypergravity technology used for fractional separation and recovery of Pb, Sn, Zn, and Cu.
Figure 10. Hypergravity separation of valuable components in mixed metals [39,74]. (a) Schematic diagram of filtration mechanism of hypergravity technology (a1): The initial phase of random distribution of antimony phase,(a2): Hypergravity is applied for separation, (a3): The phase with small particle size is separated to the bottom, black arrow: direction of phase migration); (b) Hypergravity technology used for fractional separation and recovery of Pb, Sn, Zn, and Cu.
Materials 18 00496 g010
Figure 11. WPCB metal recovery process combining the pyrolysis process and hypergravity technology separation [75].
Figure 11. WPCB metal recovery process combining the pyrolysis process and hypergravity technology separation [75].
Materials 18 00496 g011
Figure 12. Distribution and SEM of the precipitated phase in GH4169 alloy under diverse gravity conditions and temperature treatments [112]. (a) Statistics of the δ phase area fraction; (b) SEM images of the rod-shaped δ at 720 °C, 24 h, 1 G; (c) SEM images showing no δ at 720 °C, 24 h, 50,000 G; (d) SEM images of the needle-like δ at 800 °C, 24 h, 1 G; (e) SEM images of the needle-like δ at 800 °C, 24 h, 50,000 G; (f) SEM images of the needle-like δ at 850 °C, 24 h, 1 G; (g) SEM images of the needle-like δ at 850 °C, 24 h, 50,000 G.
Figure 12. Distribution and SEM of the precipitated phase in GH4169 alloy under diverse gravity conditions and temperature treatments [112]. (a) Statistics of the δ phase area fraction; (b) SEM images of the rod-shaped δ at 720 °C, 24 h, 1 G; (c) SEM images showing no δ at 720 °C, 24 h, 50,000 G; (d) SEM images of the needle-like δ at 800 °C, 24 h, 1 G; (e) SEM images of the needle-like δ at 800 °C, 24 h, 50,000 G; (f) SEM images of the needle-like δ at 850 °C, 24 h, 1 G; (g) SEM images of the needle-like δ at 850 °C, 24 h, 50,000 G.
Materials 18 00496 g012
Figure 13. Effect of hypergravity coefficient on solidified grains of different metals [113,114]. (a) Pure aluminum; (b) Cu-11%Sn alloy.
Figure 13. Effect of hypergravity coefficient on solidified grains of different metals [113,114]. (a) Pure aluminum; (b) Cu-11%Sn alloy.
Materials 18 00496 g013
Figure 14. Microstructure of Al-8 wt% Cu alloy after hypergravity treatment with G = 300 at different solidification stages [116]. (a) Starting from high-temperature liquid and ending with a solid phase fraction of 20%; (b) Starting with a solid phase fraction of 20% and ending with a solid phase fraction of 50%; (c) Starting with a solid phase fraction of 50% and ending with a solid phase fraction of 80%; (d) Starting with a solid phase fraction of 80% and ending with a solid phase fraction of 100%; (e) The complete solidification stage, where the red wireframe represents the fine crystal zone (Red dotted line: the fine grain zone of each sample).
Figure 14. Microstructure of Al-8 wt% Cu alloy after hypergravity treatment with G = 300 at different solidification stages [116]. (a) Starting from high-temperature liquid and ending with a solid phase fraction of 20%; (b) Starting with a solid phase fraction of 20% and ending with a solid phase fraction of 50%; (c) Starting with a solid phase fraction of 50% and ending with a solid phase fraction of 80%; (d) Starting with a solid phase fraction of 80% and ending with a solid phase fraction of 100%; (e) The complete solidification stage, where the red wireframe represents the fine crystal zone (Red dotted line: the fine grain zone of each sample).
Materials 18 00496 g014
Figure 15. Evolution of dendrite structure at solid–liquid interface under different gravity conditions. Red dashed circle: broken dendrites [121].
Figure 15. Evolution of dendrite structure at solid–liquid interface under different gravity conditions. Red dashed circle: broken dendrites [121].
Materials 18 00496 g015
Figure 16. Mechanism of grain refinement of 7050 aluminum alloy under hypergravity (a) Dendrite fragmentation. (b) Dense primary crystals settle to the bottom (blue arrow: hypergravity causes the dendrite root to break; dashed arrow: broken dendrites begin to form primary crystals; green arrow: direction of settling of primary crystal) [122].
Figure 16. Mechanism of grain refinement of 7050 aluminum alloy under hypergravity (a) Dendrite fragmentation. (b) Dense primary crystals settle to the bottom (blue arrow: hypergravity causes the dendrite root to break; dashed arrow: broken dendrites begin to form primary crystals; green arrow: direction of settling of primary crystal) [122].
Materials 18 00496 g016
Figure 17. The growth process of columnar crystals under different gravity conditions. A indicates the primary dendrites; the direction of gravity vector is the same as that of main dendrite growth under positive gravity and the opposite direction under negative gravity; blue is the primary dendritic phase, and green or red is the copper-rich liquid; the scale is 100 μm [126].
Figure 17. The growth process of columnar crystals under different gravity conditions. A indicates the primary dendrites; the direction of gravity vector is the same as that of main dendrite growth under positive gravity and the opposite direction under negative gravity; blue is the primary dendritic phase, and green or red is the copper-rich liquid; the scale is 100 μm [126].
Materials 18 00496 g017
Figure 18. Effect of hypergravity on dendrite growth [122]. (a) Dendrite structure at different locations; (b) the typical element distribution and microstructure of the second phase of Al7050 alloy at 1023 K (b1): 1 g—region I, (b2): 3000 g—region II, (b3): 3000 g—region IV. I–IV represent different regions). (c) area fractions of the second phases of different structures, B and T represent the bottom and top regions, respectively.
Figure 18. Effect of hypergravity on dendrite growth [122]. (a) Dendrite structure at different locations; (b) the typical element distribution and microstructure of the second phase of Al7050 alloy at 1023 K (b1): 1 g—region I, (b2): 3000 g—region II, (b3): 3000 g—region IV. I–IV represent different regions). (c) area fractions of the second phases of different structures, B and T represent the bottom and top regions, respectively.
Materials 18 00496 g018
Figure 19. The gradient structure was prepared under hypergravity [134]. (a) Samples after hypergravity treatment; (b) low-magnification and (c) high-magnification SEM and EDS analysis for the AlZn0.4Li0.2Mg0.2Cu0.2.
Figure 19. The gradient structure was prepared under hypergravity [134]. (a) Samples after hypergravity treatment; (b) low-magnification and (c) high-magnification SEM and EDS analysis for the AlZn0.4Li0.2Mg0.2Cu0.2.
Materials 18 00496 g019
Figure 20. Gradient structure in Se-Te alloy after hypergravity treatment. a–b solid red line indicates the direction of component detection [144].
Figure 20. Gradient structure in Se-Te alloy after hypergravity treatment. a–b solid red line indicates the direction of component detection [144].
Materials 18 00496 g020
Figure 21. Effect of high-temperature hypergravity on copper–brass diffusion. (a) Centrifuge-modified high-temperature hypergravity environment equipment [145]; (b) copper–brass diffusion couple treated with high-temperature hypergravity (400 °C, 400,000× g); (c) approximate vacancy concentration distribution in copper–brass diffusion couple after high-temperature and hypergravity treatment. Points A and B represent the top and bottom of the sample [146].
Figure 21. Effect of high-temperature hypergravity on copper–brass diffusion. (a) Centrifuge-modified high-temperature hypergravity environment equipment [145]; (b) copper–brass diffusion couple treated with high-temperature hypergravity (400 °C, 400,000× g); (c) approximate vacancy concentration distribution in copper–brass diffusion couple after high-temperature and hypergravity treatment. Points A and B represent the top and bottom of the sample [146].
Materials 18 00496 g021
Figure 22. Backscattering SEM images of the sample interface at a hypergravity level of 870, 2260, 3500, and 4700× g, with diffusion at 350 °C for 6 h [147]. (ad) G+ sample, the hypergravity direction is Cu to Zn; (eh) G- sample, the hypergravity direction is Zn to Cu. The yellow arrows indicates the defects (voids), the red arrows indicate the ε phase diffusion layer, and the black arrows indicate the β phase (not obvious).
Figure 22. Backscattering SEM images of the sample interface at a hypergravity level of 870, 2260, 3500, and 4700× g, with diffusion at 350 °C for 6 h [147]. (ad) G+ sample, the hypergravity direction is Cu to Zn; (eh) G- sample, the hypergravity direction is Zn to Cu. The yellow arrows indicates the defects (voids), the red arrows indicate the ε phase diffusion layer, and the black arrows indicate the β phase (not obvious).
Materials 18 00496 g022
Figure 23. SEM and EDS maps and IMC layer thicknesses of CuSn diffusion couples treated at 210 °C for 100 h under normal/hypergravity. G represents the direction of hypergravity [149].
Figure 23. SEM and EDS maps and IMC layer thicknesses of CuSn diffusion couples treated at 210 °C for 100 h under normal/hypergravity. G represents the direction of hypergravity [149].
Materials 18 00496 g023
Figure 24. Atomic structure of β phase (CuZn) under different gravity conditions, HRTEM images, and corresponding superimposed images after filtering. (a,b) HRTEM filtered images and corresponding overlapping images of the corresponding variable distribution of β under 870 g hypergravity. (c,d) HRTEM filtered images and corresponding overlapping images of the corresponding variable distribution of β under 3500 g hypergravity. (e,f) HRTEM filtered images and corresponding overlapping images of the corresponding variable distribution of β under 4700 g hypergravity. The yellow marks indicate the presence of dislocations, and the black arrows indicate the presence of strain but no dislocation. The atomic strain distribution was obtained by applying the geometric phase analysis method for TEM images [147].
Figure 24. Atomic structure of β phase (CuZn) under different gravity conditions, HRTEM images, and corresponding superimposed images after filtering. (a,b) HRTEM filtered images and corresponding overlapping images of the corresponding variable distribution of β under 870 g hypergravity. (c,d) HRTEM filtered images and corresponding overlapping images of the corresponding variable distribution of β under 3500 g hypergravity. (e,f) HRTEM filtered images and corresponding overlapping images of the corresponding variable distribution of β under 4700 g hypergravity. The yellow marks indicate the presence of dislocations, and the black arrows indicate the presence of strain but no dislocation. The atomic strain distribution was obtained by applying the geometric phase analysis method for TEM images [147].
Materials 18 00496 g024
Figure 25. Structural evolution of turbine blades under high temperature and a hypergravity field [153]. (a) Schematic; (b) different position points. 1–13 represent a severely degraded portion from tip to stalk near the trailing edge of the blade.
Figure 25. Structural evolution of turbine blades under high temperature and a hypergravity field [153]. (a) Schematic; (b) different position points. 1–13 represent a severely degraded portion from tip to stalk near the trailing edge of the blade.
Materials 18 00496 g025
Figure 26. Fracture failure behavior of 1060 aluminum alloy under different hypergravity conditions [154]. (a) The stress of the specimen; (b) difference in crack limit tension under constant stress and hypergravity gradient stress; (c) SEM fracture morphology under different hypergravity magnitude and temperature conditions.
Figure 26. Fracture failure behavior of 1060 aluminum alloy under different hypergravity conditions [154]. (a) The stress of the specimen; (b) difference in crack limit tension under constant stress and hypergravity gradient stress; (c) SEM fracture morphology under different hypergravity magnitude and temperature conditions.
Materials 18 00496 g026
Figure 27. Fracture failure behavior of 7075 aluminum alloy under different hypergravity conditions [155]. (a) Comparison of crack tips at different gravity levels; (b) EBSD analysis of different crack propagation locations (IPF diagram, GND diagram and dislocation frequency distribution of the location in the dotted box, red dotted box is the upper edge of the sample, blue dotted box is the lower edge of the sample); (c) Schematic diagram of crack patterns under hypergravity/conventional uniaxial stress (the dotted lines are in the direction of torsion).
Figure 27. Fracture failure behavior of 7075 aluminum alloy under different hypergravity conditions [155]. (a) Comparison of crack tips at different gravity levels; (b) EBSD analysis of different crack propagation locations (IPF diagram, GND diagram and dislocation frequency distribution of the location in the dotted box, red dotted box is the upper edge of the sample, blue dotted box is the lower edge of the sample); (c) Schematic diagram of crack patterns under hypergravity/conventional uniaxial stress (the dotted lines are in the direction of torsion).
Materials 18 00496 g027
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

Zheng, Y.; Xie, L.; Chen, Y.; Han, X. Effects of Hypergravity on Phase Evolution, Synthesis, Structures, and Properties of Materials: A Review. Materials 2025, 18, 496. https://doi.org/10.3390/ma18030496

AMA Style

Zheng Y, Xie L, Chen Y, Han X. Effects of Hypergravity on Phase Evolution, Synthesis, Structures, and Properties of Materials: A Review. Materials. 2025; 18(3):496. https://doi.org/10.3390/ma18030496

Chicago/Turabian Style

Zheng, Yisheng, Lilin Xie, Yanhui Chen, and Xiaodong Han. 2025. "Effects of Hypergravity on Phase Evolution, Synthesis, Structures, and Properties of Materials: A Review" Materials 18, no. 3: 496. https://doi.org/10.3390/ma18030496

APA Style

Zheng, Y., Xie, L., Chen, Y., & Han, X. (2025). Effects of Hypergravity on Phase Evolution, Synthesis, Structures, and Properties of Materials: A Review. Materials, 18(3), 496. https://doi.org/10.3390/ma18030496

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