www.fgks.org   »   [go: up one dir, main page]

Jump to content

High-entropy alloy

From Wikipedia, the free encyclopedia

This is an old revision of this page, as edited by Mr.Z-man (talk | contribs) at 16:54, 29 May 2015 (add more properties and applications). The present address (URL) is a permanent link to this revision, which may differ significantly from the current revision.

High-entropy alloys (HEAs) are a class of multi-component alloys composed of 5 or more principal constituent elements each with a concentration between 5 and 35 atomic %.[1] The defining feature of HEAs over other complex alloys is that they consist entirely or primarily of a simple solid solution phase, thus having a high entropy of mixing.

Atomic structure model of fcc CoCrFeMnNi[2]

HEAs have been described in the literature that have better strength-to-weight ratios, fracture toughness, tensile strength, high-temperature strength, or corrosion resistance than similar conventional alloys.[1][3]

Early development

Although HEAs were described as early as 1996,[4] significant research interest did not develop until after independent 2004 papers by Jien-Wei Yeh and Brian Cantor, with Yeh's first paper on the topic published 2 months sooner. Yeh also coined the term "high-entropy alloy" when he attributed the high configurational entropy as the mechanism stabilizing the solid solution phase.[5] Cantor, not knowing of Yeh's work, did not describe his alloy as a "high entropy" alloy, but the base alloy he developed, equiatomic FeCrMnNiCo, has been the subject of considerable work in the field.[6]

Definition

There is no universally agreed-upon definition of a HEA. Yeh originally defined HEAs as alloys containing at least 5 elements with concentrations between 5 and 35 atomic percent.[5] Later research however, suggested that this definition may be inadequate. Otto et al. suggested that only alloys that form a solid solution with no intermetallic phases should be considered true high-entropy alloys, as the formation of ordered phases decreases the entropy of the system.[7] Some authors have described 4-component alloys as high-entropy alloys[8][9] while others have suggested that alloys meeting the other requirements of HEAs, but with only 2–4 elements[10] or a mixing entropy between R and 1.5R[11] should be considered "medium-entropy" alloys.[12]

Alloy design

In conventional alloy design, one primary element such as iron, copper, or aluminum is chosen for its properties. Then, small amounts of additional elements are added to improve or add properties. Even among binary alloy systems, there are few common cases of both elements being used in nearly-equal proportions such as Pb-Sn solders. Therefore, much is known from experimental results about phases near the edges of binary phase diagrams and the corners of ternary phase diagrams and much less is known about phases near the centers. In higher-order (4+ components) systems that cannot be easily represented on a 2-dimensional phase diagram, virtually nothing is known.[6]

Phase formation

Gibbs' phase rule, , can be used to determine an upper bound on the number of phases that will form in an equilibrium system. In his 2004 paper, Cantor created a 20-component alloy containing 5 at% of Mn, Cr, Fe, Co, Ni, Cu, Ag, W, Mo, Nb, Al, Cd, Sn, Pb, Bi, Zn, Ge, Si, Sb, and Mg. At constant pressure, the phase rule would allow for up to 21 phases at equilibrium, but far fewer actually formed. The predominant phase was a face-centered cubic solid solution phase, containing mainly Fe, Ni, Cr, Co, and Mn. From that result, the FeCrMnNiCo alloy, which forms only a solid solution phase, was developed.[6]

The Hume-Rothery rules have historically been applied to determine whether a mixture will form a solid solution. Research into high-entropy alloys has found that in multi-component systems, these rules tend to be relaxed slightly. In particular, the rule that solvent and solute elements must have the same crystal structure does not seem to apply, as Fe, Ni, Cr, Co, and Mn have 4 different crystal structures as pure elements (and when the elements are present in equal concentrations, there can be no meaningful distinction between "solvent" and "solute" elements).[7]

Thermodynamic mechanisms

The multi-component alloys Yeh developed also consisted mostly or entirely of solid solution phases, contrary to what had been expected from earlier work in multi-component systems, primarily in the field of metallic glasses.[5] [13] Yeh attributed this result to the high configurational, or mixing, entropy of a random solid solution containing numerous elements. Because , and the phase with the lowest Gibbs free energy of formation (ΔG) will be the phase formed at equilibrium, increasing ΔS (entropy) will increase the likelihood of a phase being stable. The mixing entropy for a random ideal solid solution can be calculated by:

where R is the ideal gas constant, N is the number of components, and ci is the atomic fraction of component i. From this it can be seen that alloys in which the components are present in equal proportions will have the highest entropy, and adding additional elements will increase the entropy. A 5 component, equiatomic alloy will have a mixing entropy of 1.61R.[5][14]

Parameter Design guideline
∆Smix Maximized
∆Hmix > -10 and < 5 kJ/mol
Ω ≥ 1.1
δ ≤ 6.6%
VEC ≥ 8 for fcc, <6.87 for bcc
Empirical parameters and design guidelines for forming solid solution HEAs

However entropy alone is not sufficient to stabilize the solid solution phase in every system. The enthalpy of mixing (ΔH), must also be taken into account. This can be calculated using:

where is the binary enthalpy of mixing for A and B[15]. Zhang et al. found, empirically, that in order to form a complete solid solution, ΔHmix should be between -10 and 5 kJ/mol.[14] In addition, Otto et al. found that if the alloy contains any pair of elements that tend to form ordered compounds in their binary system, a multi-component alloy containing them is also likely to form ordered compounds.[7]

Both of the thermodynamic parameters can be combined into a single, unitless parameter Ω:

where Tm is the average melting point of the elements in the alloy. Ω should be greater than or equal to 1.1 to promote solid solution development.[3]

Kinetic mechanisms

The atomic radii of the components must also be similar in order to form a solid solution. Zhang et al. proposed a parameter δ representing the difference in atomic radii:

where ri is the atomic radius of element i and . Formation of a solid solution phase requires a δ≤6.6%, but some alloys with 4%<δ≤6.6% do form intermetallics.[14][3]

Other properties

For those alloys that do form solid solutions, an additional empirical parameter has been proposed to predict the crystal structure that will form. If the average valence electron concentration (VEC) of the alloy is ≥8, the alloy will form a face-centered cubic (fcc) lattice. If the average VEC is <6.87, it will form a body-centered cubic (bcc) lattice. For values in between, it will form a mixture of fcc and bcc.[16] VEC has also been used to predict the formation of σ-phase intermetallics (which are generally brittle and undesirable) in chromium and vanadium-containing HEAs.[17]

Synthesis

High entropy alloys are mostly produced using distinct methods that depend on the initial phase - starting either from a liquid, solid, or gas state.

Other HEAs have been produced by thermal spray, laser cladding, and electrodeposition.[3][20]

Properties and potential uses

Mechanical

The crystal structure of HEAs has been found to be the dominant factor in determining the mechanical properties. bcc HEAs typically have high yield strength and low ductility and vice versa for fcc HEAs. Some alloys have been particularly noted for their exceptional mechanical properties. A refractory alloy, VNbMoTaW maintains a high yield strength (>600 MPa (87 ksi)) even at a temperature of 1,400 °C (2,550 °F), significantly outperforming conventional superalloys such as Inconel 718. However, room temperature ductility is poor, less is known about other important high temperature properties such as creep resistance, and the density of the alloy is higher than conventional nickel-based superalloys.[3]

CoCrFeMnNi has been found to have exceptional low-temperature mechanical properties and high fracture toughness, with both ductility and yield strength increasing as the test temperature was reduced from room temperature to 77 K (−321.1 °F). This was attributed to the onset of nanoscale twin boundary formation, an additional deformation mechanism that was not in effect at higher temperatures. As such, it may have applications as a structural material in low-temperature applications or, because of its high toughness, as an energy-absorbing material.[21] However, later research showed that lower-entropy alloys with fewer elements or non-equiatomic compositions may have higher strength[22] or higher toughness.[23] No ductile to brittle transition was observed in the bcc AlCoCrFeNi alloy in tests as low as 77 K.[3]

Al0.5CoCrCuFeNi was found to have a high fatigue life and endurance limit, possibly exceeding some conventional steel and titanium alloys. But there was significant variability in the results, suggesting the material is very sensitve to defects introduced during manufacturing such as aluminum oxide particles and microcracks.[24]

A single-phase nanocrystalline Al20Li20Mg10Sc20Ti30 alloy was developed with a density of 2.67 g cm−3 and microhardness of 4.9 – 5.8 GPa, which would give it an estimated strength-to-weight ratio comparable to ceramic materials such as silicon carbide,[18] though the high cost of scandium limits the possible uses.[25]

Electrical and magnetic

CoCrCuFeNi is an fcc alloy that was found to be paramagnetic. But upon adding titanium, it forms a complex microstructure consisting of fcc solid solution, amorphous regions and nanoparticles of Laves phase, resulting in superparamagnetic behavior.[26] High magnetic coercivity has been measured in a BiFeCoNiMn alloy[20]

Other

The high concentrations of multiple elements leads to slow diffusion. The activation energy for diffusion was found to be higher for several elements in CoCrFeMnNi than in pure metals and stainless steels, leading to lower diffusion coefficients.[27]

References

  1. ^ a b Tsai, Ming-Hung; Yeh, Jien-Wei (30 April 2014). "High-Entropy Alloys: A Critical Review". Materials Research Letters. 2 (3): 107–123. doi:10.1080/21663831.2014.912690.
  2. ^ Wang, Shaoqing (13 December 2013). "Atomic Structure Modeling of Multi-Principal-Element Alloys by the Principle of Maximum Entropy". Entropy. 15 (12): 5536–5548. doi:10.3390/e15125536.{{cite journal}}: CS1 maint: unflagged free DOI (link)
  3. ^ a b c d e f g h i Zhang, Yong; Zuo, Ting Ting; Tang, Zhi; Gao, Michael C.; Dahmen, Karin A.; Liaw, Peter K.; Lu, Zhao Ping (April 2014). "Microstructures and properties of high-entropy alloys". Progress in Materials Science. 61: 1–93. doi:10.1016/j.pmatsci.2013.10.001.
  4. ^ Huang KH, Yeh JW. A study on multicomponent alloy systems containing equal-mole elements [M.S. thesis]. Hsinchu: National Tsing Hua University; 1996.
  5. ^ a b c d Yeh, J.-W. (May 2004). "Nanostructured High-Entropy Alloys with Multiple Principal Elements: Novel Alloy Design Concepts and Outcomes". Advanced Engineering Materials. 6 (5): 299–303. doi:10.1002/adem.200300567. {{cite journal}}: Unknown parameter |coauthors= ignored (|author= suggested) (help)
  6. ^ a b c Cantor, B. (July 2004). "Microstructural development in equiatomic multicomponent alloys". Materials Science and Engineering: A. 375–377: 213–218. doi:10.1016/j.msea.2003.10.257. {{cite journal}}: Unknown parameter |coauthors= ignored (|author= suggested) (help)
  7. ^ a b c Otto, F. (April 2013). "Relative effects of enthalpy and entropy on the phase stability of equiatomic high-entropy alloys". Acta Materialia. 61 (7): 2628–2638. doi:10.1016/j.actamat.2013.01.042. {{cite journal}}: Unknown parameter |coauthors= ignored (|author= suggested) (help)
  8. ^ Zou, Yu (February 2014). "Size-dependent plasticity in an Nb25Mo25Ta25W25 refractory high-entropy alloy". Acta Materialia. 65: 85–97. doi:10.1016/j.actamat.2013.11.049. {{cite journal}}: Unknown parameter |coauthors= ignored (|author= suggested) (help)
  9. ^ Middleburgh, S.C. (June 2014). "Segregation and migration of species in the CrCoFeNi high entropy alloy". Journal of Alloys and Compounds. 599: 179–182. doi:10.1016/j.jallcom.2014.01.135. {{cite journal}}: Unknown parameter |coauthors= ignored (|author= suggested) (help)
  10. ^ Gali, A. (August 2013). "Tensile properties of high- and medium-entropy alloys". Intermetallics. 39: 74–78. doi:10.1016/j.intermet.2013.03.018. {{cite journal}}: Unknown parameter |coauthors= ignored (|author= suggested) (help)
  11. ^ Miracle, Daniel (10 January 2014). "Exploration and Development of High Entropy Alloys for Structural Applications". Entropy. 16 (1): 494–525. doi:10.3390/e16010494. {{cite journal}}: Unknown parameter |coauthors= ignored (|author= suggested) (help)CS1 maint: unflagged free DOI (link)
  12. ^ Gali, A.; George, E.P. (August 2013). "Tensile properties of high- and medium-entropy alloys". Intermetallics. 39: 74–78. doi:10.1016/j.intermet.2013.03.018.
  13. ^ Greer, A. Lindsay (25 December 1993). "Confusion by design". Nature. 366 (6453): 303–304. doi:10.1038/366303a0.
  14. ^ a b c Zhang, Y. (June 2008). "Solid-Solution Phase Formation Rules for Multi-component Alloys". Advanced Engineering Materials. 10 (6): 534–538. doi:10.1002/adem.200700240. {{cite journal}}: Unknown parameter |coauthors= ignored (|author= suggested) (help)
  15. ^ Takeuchi, Akira (2005). "Classification of Bulk Metallic Glasses by Atomic Size Difference, Heat of Mixing and Period of Constituent Elements and Its Application to Characterization of the Main Alloying Element". Materials Transactions. 46 (12): 2817–2829. doi:10.2320/matertrans.46.2817. {{cite journal}}: Unknown parameter |coauthors= ignored (|author= suggested) (help)
  16. ^ Guo, Sheng (2011). "Effect of valence electron concentration on stability of fcc or bcc phase in high entropy alloys". Journal of Applied Physics. 109 (10): 103505. doi:10.1063/1.3587228. {{cite journal}}: Unknown parameter |coauthors= ignored (|author= suggested) (help)
  17. ^ Tsai, Ming-Hung; Tsai, Kun-Yo; Tsai, Che-Wei; Lee, Chi; Juan, Chien-Chang; Yeh, Jien-Wei (20 August 2013). "Criterion for Sigma Phase Formation in Cr- and V-Containing High-Entropy Alloys". Materials Research Letters. 1 (4): 207–212. doi:10.1080/21663831.2013.831382.
  18. ^ a b Youssef, Khaled M.; Zaddach, Alexander J.; Niu, Changning; Irving, Douglas L.; Koch, Carl C. (9 December 2014). "A Novel Low-Density, High-Hardness, High-entropy Alloy with Close-packed Single-phase Nanocrystalline Structures". Materials Research Letters. 3 (2): 95–99. doi:10.1080/21663831.2014.985855.
  19. ^ Ji, Wei; Wang, Weimin; Wang, Hao; Zhang, Jinyong; Wang, Yucheng; Zhang, Fan; Fu, Zhengyi (January 2015). "Alloying behavior and novel properties of CoCrFeNiMn high-entropy alloy fabricated by mechanical alloying and spark plasma sintering". Intermetallics. 56: 24–27. doi:10.1016/j.intermet.2014.08.008.
  20. ^ a b Yao, Chen-Zhong; Zhang, Peng; Liu, Meng; Li, Gao-Ren; Ye, Jian-Qing; Liu, Peng; Tong, Ye-Xiang (November 2008). "Electrochemical preparation and magnetic study of Bi–Fe–Co–Ni–Mn high entropy alloy". Electrochimica Acta. 53 (28): 8359–8365. doi:10.1016/j.electacta.2008.06.036.
  21. ^ Otto, F. (September 2013). "The influences of temperature and microstructure on the tensile properties of a CoCrFeMnNi high-entropy alloy". Acta Materialia. 61 (15): 5743–5755. doi:10.1016/j.actamat.2013.06.018. {{cite journal}}: Unknown parameter |coauthors= ignored (|author= suggested) (help)
  22. ^ Wu, Z.; Bei, H.; Otto, F.; Pharr, G.M.; George, E.P. (March 2014). "Recovery, recrystallization, grain growth and phase stability of a family of FCC-structured multi-component equiatomic solid solution alloys". Intermetallics. 46: 131–140. doi:10.1016/j.intermet.2013.10.024.
  23. ^ Zaddach, A.J.; Scattergood, R.O.; Koch, C.C. (June 2015). "Tensile properties of low-stacking fault energy high-entropy alloys". Materials Science and Engineering: A. 636: 373–378. doi:10.1016/j.msea.2015.03.109.
  24. ^ Hemphill, M.A. (September 2012). "Fatigue behavior of Al0.5CoCrCuFeNi high entropy alloys". Acta Materialia. 60 (16): 5723–5734. doi:10.1016/j.actamat.2012.06.046. {{cite journal}}: Unknown parameter |coauthors= ignored (|author= suggested) (help)
  25. ^ Shipman, Matt. "New 'high-entropy' alloy is as light as aluminum, as strong as titanium alloys". Phys.org. Retrieved 29 May 2015.
  26. ^ Wang, X.F. (March 2007). "Novel microstructure and properties of multicomponent CoCrCuFeNiTix alloys". Intermetallics. 15 (3): 357–362. doi:10.1016/j.intermet.2006.08.005. {{cite journal}}: Unknown parameter |coauthors= ignored (|author= suggested) (help)
  27. ^ Tsai, K.-Y. (August 2013). "Sluggish diffusion in Co–Cr–Fe–Mn–Ni high-entropy alloys". Acta Materialia. 61 (13): 4887–4897. doi:10.1016/j.actamat.2013.04.058. {{cite journal}}: Unknown parameter |coauthors= ignored (|author= suggested) (help)

See also