Abstract
Electrochemical energy storage (EES) systems demand electrode materials with high power density, energy density, and long cycle life. Metal-organic frameworks (MOFs) are promising electrode materials, while new MOFs with high conductivity, high stability, and abundant redox-reactive sites are demanded to meet the growing needs of EES. Density Functional Theory (DFT) could calculate these properties of MOFs and provide atomic-level insights into the mechanisms, based on which machine learning (ML) can screen MOFs for EES efficiently. In this review, we first review the exploration of mechanisms based on DFT calculations. We focus on the conductivity, stability, and reactivity of MOFs in EES systems. Then, we review the steps to apply ML in screening MOFs. Establishing datasets of MOFs, extracting features from MOF structure, and applying ML in screening MOFs are discussed. Finally, the review proposes the future avenue of DFT and ML to make up the gaps in the knowledge of MOFs.

Similar content being viewed by others
Introduction
Advancements in electrochemical energy storage (EES) systems, such as supercapacitors and batteries, are necessary to meet the demands of rapidly growing electric vehicles and large-scale energy storage1,2,3,4,5. Supercapacitors are distinguished by their high power density, which enables rapid charging and discharging within seconds, as well as their exceptionally long cycle life, exceeding 106 charge-discharge cycles6,7. In contrast, lithium-ion batteries (LIBs) and other metal-ion batteries (sodium, potassium, and zinc, battery systems) are characterized by their high energy densities and relatively limited cycle life (a few thousand cycles)7,8,9,10. Both batteries and supercapacitors encounter challenges in achieving higher energy density, power density, and long cycle life simultaneously7,11. The performance of EES is limited by the electrode materials: (1) low electrical conductivity and the high solubility of organic electrode materials lead to low power density and cycle life12,13, and (2) inorganic electrode materials exhibit low energy density13,14. Consequently, the development of electrode materials with excellent electrical and ionic conductivity for high power density, robust stability for extended cycle life, high accessible specific surface area, and numerous active sites for achieving high energy density has become a critical focus in advancing high-performance EES systems15.
Metal-organic frameworks (MOFs) are a class of crystalline porous materials constructed from metal ions or clusters coordinated with organic ligands, which are distinguished by their high specific surface areas, significant porosity, and tunable structure enabling different functionalities16,17,18. These properties make MOFs highly suitable as electrode materials for EES systems, and several MOFs have already been applied in such systems19,20. For instance, Ni3(2,3,6,7,10,11-hexaiminotriphenylene)2 (Ni3(HITP)2) which exhibits a high surface area of 630 m2 g−1 and a conductivity of 40 S cm−1 has been utilized as an electrode material for supercapacitors without the need for binders or conductive additives20. However, the conductivity of MOFs remains relatively low compared to materials like graphene and graphite, which significantly affects the rate performance21,22,23. Furthermore, the structural integrity of MOFs can be compromised during the charge and discharge cycles, leading to poor cycle stability22. To meet the needs of the growing energy sector, more MOFs with high electrical conductivity, stability, and abundant active sites are demanded11,24,25. One of the most intriguing aspects of MOFs is their tunable structure and composition, which can be tailored by assembling various building blocks to achieve distinct properties suitable for diverse applications26,27,28. Moreover, the disordered nature of porous carbon complicates modeling efforts, making it challenging to investigate structure–performance relationships21. In contrast, the highly organized and well-defined structures of MOFs facilitate modeling, offering clearer insights into the correlation between structures and performance21. Therefore, to design MOFs optimized for EES systems, it is necessitating to investigate the mechanisms of electrical conductivity, stability, and redox reactions.
Using density functional theory (DFT), the Schrodinger equation can be approximately solved and various physical and chemical properties of MOFs, such as band structure, reaction potential, and cohesive energy, can be determined29. These calculations provide insights into the conductivity, stability, and redox activity of MOFs, offering a deeper understanding of the underlying mechanisms30,31. This atomic-level understanding gained through DFT calculations is instrumental in the rational design and screening of MOFs for EES systems. For instance, linker aromaticity of 2D conductive MOFs, as determined by a combination of DFT calculations and NICS-xy scans, was found to reduce charge mobility, suggesting that modifying non-coordinating functional groups can adjust conductivity by influencing linker aromaticity32,33. With the advancement of computational capabilities, high-throughput calculations of MOFs become feasible and tons of data are generated which is beyond the limitation that a human expert can process34,35,36. By identifying patterns within data generated by DFT, machine learning (ML) becomes a powerful tool to rapidly establish complex relationships between MOF structures and their properties, allowing ML to quickly predict MOF properties without solving the complicated Schrodinger equation37. ML enables us to efficiently explore the vast chemical space of MOFs which would be inaccessible using traditional methods because the number of possible structures of MOFs is too large38,39,40,41. With the development of ML, models have become more accurate but increasingly difficult to interpret, limiting our understanding of the relationships between structures and properties37. Enhancing the interpretability of ML models is essential for investigating mechanisms and designing MOFs34.
In this review, we focus on the use of DFT and ML for screening and designing MOFs as electrode materials in EES systems. First, we discuss strategies for enhancing conductivity, chemical and mechanical stability, and redox-active sites based on DFT calculations. We then provide an overview of ML approaches for predicting MOF properties, covering database construction, feature engineering, and various ML applications. Finally, we discuss the remaining challenges and outline future directions to guide the design of MOFs.
DFT insights into the mechanisms of MOF properties
Conductivity of MOFs
Power density is largely influenced by electrode conductivity, as conductivity determines how quickly electrons can be transported within electrode materials11. Most MOFs are composed of carboxylate linkers, which form robust ionic bonds with metal ions, contributing to structural stability but resulting in poor conductivity42,43. However, due to their tunable structures and functionalities, conductive MOFs have been successfully developed44,45. Gaining deeper insights into the relationship between structure and conductivity is crucial for designing more conductive MOFs. Charge transport in electrodes can be described through one of two primary mechanisms: hopping transport and band transport24,30. In the case where sites are spatially separated and the interactions between them are weak, charge carriers transport between sites via hopping, typically found in the disordered materials24. Conversely, when sites are sufficiently close to allow strong interactions, as commonly found in crystalline inorganic materials, continuous energy bands form that enable charge transport, and if the energy levels of these sites are also well-aligned, the electronic structure can be further tuned to enhance charge delocalization24,46.
In the following section, we will first analyze the electronic structure of MOFs calculated using DFT within the framework of band theory to explore their conductivity. For the hopping mechanism, several models have been developed to describe the electronic structure and electrical properties47,48,49. For detailed introductions, please refer to these reviews30,50. We will then discuss the mechanisms behind electrically conductive MOFs and strategies to improve their conductivity.
Electronic structure
Band structure plays a key role in enabling physicists, chemists, and materials scientists to analyze and understand the properties of materials29,51. In band theory, the electronic band structure describes the range of energy levels that an electron in a material can occupy, providing insights into the material’s electrical properties29. In particular, the band structure is plotted along the wave vector (k) in the first Brillouin zone, then the band gap can be obtained from the energy difference between the valance band and the conduction band51. Intrinsic conductivity is largely determined by the band gap, which defines the intrinsic differences between a conductor, semiconductor, and insulator51: A conductor has a band crossing the Fermi level (EF); A semiconductor usually has a band gap (Eg) smaller than 2 eV, allowing charge carriers to cross the band gap to the conduction band through thermal or optical excitation; in contrast, an insulator has a larger band gap52,53,54,55. The band gap is a direct descriptor for identifying conductive MOFs. The accuracy of DFT calculations is influenced by the exchange-correlation (XC) functional. Band gap is usually underestimated by the Perdew–Burke–Ernzerhof (PBE) functional due to self-interaction error56. Although hybrid functionals tend to offer more accurate band gap predictions, their significantly higher computational complexity of \(O({N}_{e}^{3}\log {N}_{e})\) than PBE of \(O({N}_{e}^{2}\log {N}_{e})\) limits their widespread application29,57. Another approach to addressing the PBE underestimation without incurring substantial additional computational costs is incorporating a Hubbard U correction. However, the accuracy of DFT + U calculations depends on the selection of the effective Hubbard parameter, Ueff, which can vary with different systems or conditions58. The GW method, adopting Green’s function and screened Coulomb interaction to represent the self-energy, greatly reduces self-interaction error and yields reliable band gap values59,60. For instance, Nisar et al. 61 evaluated the band gap of pristine kaolinite using PBE, HSE (Heyd-Scuseria-Ernzerhof), and G0W0 (an approximation of the GW method), reporting values of 4.9, 6.2, and 8.2 eV, respectively, and inferring that the true band gap likely lies between 6.2 and 8.2 eV. In a separate comparison, Yang et al. 62 conducted a comparative study of hybrid functionals (HSE and PBE0) and the quasiparticle self-consistent GW method with vertex corrections for several extended systems, finding that the GW approach provided the most accurate band gap predictions when compared with experimental data. Besides the intrinsic conductivity, the conductive direction of MOFs can also be obtained through the band structure. As illustrated in Fig. 1a, the band structure reveals that Ni3(HIB)2 (HIB = hexaiminobenzene) exhibits conductivity in the in-plane direction, as evidenced by bands crossing the Fermi level45. In contrast, LaHHTP (H₆HHTP = 2,3,6,7,10,11-hexahydroxytriphenylene) is conductive in the direction perpendicular to the 2D sheet (Fig. 1b), with bands crossing the Fermi level along the out-of-plane direction63. These anisotropic conductivities exhibited in MOFs contribute to the distinct conductivity mechanisms, which will be discussed in section, “Mechanisms and improvements of conductivity”.
a Calculated band structure of bulk Ni3(HIB)2 and Cu3(HIB)2 and the corresponding first Brillouin zone and high-symmetry K-points, indicating conductive in the in-plane direction. Reproduced with permission from ref. 45 Copyright 2017 American Chemical Society (ACS). b Calculated band structure and density of states (DOS) of LaHHTP and the corresponding first Brillouin zone and high-symmetry K-points, indicating conductive in the direction perpendicular to the 2D sheet. Reproduced with permission from ref. 63 Copyright 2019 Springer Nature.
As for conductive MOFs, their conductivity can be further evaluated. Electrical conductivity (σ) quantifies how efficiently a material can transport electrical charge when subjected to an electric field and is related to charge carrier concentrations (n) and mobility (μ)51:
where q is the elementary charge and the subscripts e and h refer to electrons and holes30. The concentration of charge carriers (n) is primarily governed by the ratio \({E}_{g}/{K}_{B}T\), representing the relationship between the band gap and temperature. When this ratio is large, the intrinsic carrier concentration becomes relatively low51. The charge mobility (μ) is inversely proportional to the effective mass (m*) of the electrons or holes and is influenced by the scattering time (τ) within the material51:
The effective mass m* can be calculated from the band structure, which is denoted as51:
where ℏ is the reduced Plank constant and \(\frac{{\partial }^{2}E}{\partial {k}^{2}}\) is the band curvature at k point. If the band curvature is negative, the charge carrier is considered to be a hole, whereas a positive m* corresponds to an electron. A band with greater curvature means a smaller m*, leading to higher μ and σ. Except for m*, τ is also needed to calculate μ. While m* can be easily obtained from the band structure, accurately calculating n and τ based on DFT is challenging, making it difficult to directly determine conductivity. Most of the studies compare the charge mobility or conductivity by evaluating m* under specific conditions. Clough et al. 64 calculated the effective mass in different directions of a 2D MOF. The smallest effective mass in the in-plane direction is 1.27 me which is much bigger than the minimum effective mass of 0.29 me along the c-axis. This means easier transport in the c-axis and indicates that the main conductive pathway is through the c-axis. Band curvature represents a specific characteristic of energy band dispersion, quantifying the degree of curvature of the energy band. Therefore, we can use the band dispersion instead of calculating band curvature to compare m*. Wang et al. 65 reported that the conductivity is dominated by the interlayer direction by comparing the band dispersion.
DFT calculations are typically based on ideal crystal structures, whereas experimentally synthesized materials inherently contain defects, leading to discrepancies between experimental results and DFT predictions66. Ni3(HITP)2 was first synthesized in 2014, and its conductivity was found to increase with temperature from 77 to 450 K, indicating semiconducting behavior44. However, Ni3(HITP)2 was predicted to be metallic, as DFT calculations revealed a zero band gap67,68,69. To address this discrepancy, Foster et al. 70 constructed MOF structures containing two kinds of defects, grain boundaries, and stacking faults (Perpendicular grain boundary, Strike-slip fault between grains, and layer–layer displacement defect) and calculated the band structure of these structures. The calculation results revealed that all defects would lead to dispersionless or open a band gap. They concluded that the conductivity of Ni3(HITP)2 was dominated by the charge hopping between the microstructure70, and the same inference was made for Ni3(HIB)2 and Cu3(HIB)245. To determine the actual structure of Ni3(HITP)2 at 293 K, Zhang et al. 71 performed ab initio molecular dynamics (AIMD), which revealed that the layers slip in-plane and expand or contract out-of-plane, thereby introducing stacking faults into the structure. These defects lead to an energy difference of 50–200 meV in the in-plane direction and produce a relatively flat band out-of-plane. Debela et al. 66 investigated the impact of hydrogenic defects on the conductivity of Ni3(HITP)2. Hydrogenic vacancies and interstitials were found to significantly affect the electronic structure, often converting the predicted metallic behavior into semiconducting properties.
Ogle et al. 72 synthesized Cu-BHT (BHT = benzenehexathiol) using two different methods: solid–vapor (S-V) and vapor–vapor (V-V) chemical vapor deposition (CVD) approaches. They found that the S-V method produced structures with lower Cu vacancies and larger single-crystalline domain sizes compared to the V-V method72. Additionally, the S-V synthesized Cu-BHT tended to form AB stacking, whereas the V-V synthesized structures exhibited AA stacking72. These structural differences were attributed to the observed electrical properties, with the S-V approach resulting in metallic behavior and the V-V approach yielding semiconducting behavior72. The effect of copper vacancies on conductivity was further investigated by Luo et al.73 and they found that copper vacancies did not alter the metallic behavior. This finding was supported by a zero band gap calculated using density functional theory (DFT) and the reciprocal relationship between temperature and conductivity73. However, Cu-BHT with a higher density of copper vacancies exhibited lower conductivity, as the vacancies disrupted charge transport73. Skorupski et al. 63 investigated linker vacancies in MOFs composed of lanthanide ions and H₆HHTP. They discovered that the absence of a linker disrupts π-π stacking between the linkers, leading to the localization of the linker electronic wavefunctions and reduced band dispersion in the out-of-plane direction63. Additionally, the authors attributed the semiconducting behavior to metal vacancies identified via X-ray diffraction, as well as to the presence of grain boundaries and anisotropic conductivity63.
On the other side, experimental efforts have focused on synthesizing single-crystal MOFs to eliminate the influence of defects on conductivity. Day et al. 74 successfully synthesized a single crystal of Ni3(HITP)2 which exhibited an intrinsic metallic nature and the conductivity was up to 150 S cm−1. The result is consistent with the DFT calculations and provides insights into the charge transport from an experimental point of view. Skorupskii et al. 75 synthesized new porous MOFs, Ln1.5HOTP (Ln = La, Nd; HOTP = 2,3,6,7,10,11-hexaoxytriphenylene), of which single crystals are metallic and the conductivity reached 1000 S cm−1. These reports showed that the conductivities of polycrystalline MOFs are influenced by the crystallinity74,75.
Mechanisms and improvements of conductivity
The ionic nature of the bonds between metal nodes and organic linkers tends to localize electrons, making efficient electrical conduction challenging76. Additionally, the high porosity of MOFs, while beneficial for applications of EES systems, often reduces the overlap between electronic states24. This reduced overlap impedes the free movement of charge carriers, leading to lower conductivity. Therefore, porosity seems to inherently conflict with conductivity25. Through-bond and through-space are two effective methods for enhancing conductivity24. In the through-bond approach, coordination bonds between organic ligands and metal nodes form networks for charge transport24. Pathak et al. 77 reported a MOF, {[Cu2(6-Hmna)(6-mn)]·NH4}n (1, 6-Hmna = 6-mercaptonicotinic acid, 6 mn = 6-mercaptonicotinate). Calculations revealed greater band dispersion along the direction of the (-Cu-S-) plane, with Cu-d and S-p states contributing to the valence band maximum (VBM)77. These results indicate that the continuous transport routes along the (-Cu-S-) plane make the MOF electrically conductive77. In graphene-like 2D MOFs, extended π-d conjugation enables efficient electron delocalization between coordination bonds45. Electrons delocalize not only over the organic ligands but also across the metal centers, leading to a continuous electronic network within the MOF. M3(HIB)2 (M = Ni, Cu) features trigonal organic ligands and square-planar mononuclear metal nodes, which assemble into 2D sheets45. These structural features promote effective charge delocalization via extended π-d conjugation between the metal nodes and ligands45. In the through-space approach, organic ligands interact with each other to form through-space charge transport pathways63,64,78. This interaction, known as π-π stacking, is commonly observed in 2D MOFs63,64,78,79. Lanthanide ions (Ln3+) possess valence electrons in deeply buried 4 f orbitals, leading to predominantly ionic bonding63. However, LnHHTP demonstrated electrical conductivity, attributed to π-π stacking interactions between the ligands, which facilitate effective charge delocalization63. These different approaches can be recognized through DFT calculations. For extended π-d conjugation and through-space mechanisms, MOFs typically exhibit conductivity in specific directions: in-plane direction for extended π-d conjugation approach45 and out-of-plane direction for through-space approach63,64,78. Furthermore, the density of states (DOS) near the Fermi level is jointly contributed by both ligands and metals for the extended π-d conjugation approach45, whereas it is primarily contributed by the ligand in the through-space approach63,64,78. Conductive MOFs based on the through-bond mechanism can conduct electrons either in a specific direction or in all directions. Partial charge density plots of the VBM and conduction band minimum (CBM) can aid in identifying conductive pathways and, thereby, determining the conductive mechanism77.
Based on the understanding of charge transport mechanisms, several methods have been proposed to improve conductivity. Softer coordination atoms, such as nitrogen or sulfur, can enhance the covalency of coordination bonds, thereby improving conductivity42,77. Incorporating bimetallic nodes is also an effective strategy for enhancing conductivity, as the introduction of an additional metal node increases the concentration of charge carriers80,81,82. Demuth et al. 32 suggested that the aromaticity of the linker influences charge mobility, as aromatic linkers tend to stabilize π-electrons. In the through-space approach, a smaller stacking distance results in stronger π-π stacking interactions, leading to higher conductivity63,83. The highest occupied molecular orbital (HOMO) and the lowest unoccupied molecular orbital (LUMO) of ligand core and metal-organic linkage should be well aligned to form dispersed bands crossing the Fermi level, thereby enhancing charge mobility78,79,84.
Additionally, charge transport in MOFs can occur via hopping mechanisms24. Conductivity can be improved by incorporating redox-active metals or ligands into MOFs as sites for charge hopping, which facilitates the formation of continuous charge transport pathways24. NDI (naphthalene diimide) is a redox-active linker that forms a series of conductive MOFs with various metals85,86,87,88. Introducing mixed valency into MOFs has also been shown to be an effective strategy for enhancing conductivity46,89,90. Upon exposure to the atmosphere, Fe2(BDT)3 (H2BDT = 5,5’-(1,4-phenylene)bis(1H-tetrazole)) underwent partial oxidation, resulting in the emergence of mixed valency states of iron, which led to an increase in conductivity by more than 5 orders of magnitude90. Furthermore, incorporating electroactive guest molecules into the pores of MOFs can modulate the conductivity, although this often severely compromises porosity, thereby affecting their applicability in EES24,91. Talin et al. 91 introduced the molecule 7,7,8,8-tetracyanoquinododimethane (TCNQ) into the MOF Cu3(BTC)2 (BTC = benzene-1,3,5-tricarboxylic acid) resulting in six orders of magnitude increase in conductivity compared.
Based on the understanding of the conductive mechanisms, a lot of porous conductive MOFs have been designed and synthesized23,90,92. Nd1.5HOTP exhibits a remarkable conductivity of 1080 S cm−1 at room temperature, demonstrating the highest conductivity reached in porous MOFs or other intrinsically porous materials75.
Stability of MOFs
The stability of electrode materials determines the cycle life of EES systems, but the instability of MOFs significantly limits their application in these systems. Therefore, understanding the factors that influence stability and developing strategies to enhance it are of critical importance to MOF-based EES. The cycle life is influenced by both chemical and mechanical stability, as the framework structures of MOFs are prone to collapse under harsh chemical conditions or mechanical stress, significantly affecting their performance during extended cycling93. Here, we focus on the efforts to understand the mechanisms of chemical stability and mechanical stability and design new stable MOFs based on this understanding.
Chemical Stability
Chemical stability refers to the ability of MOFs to retain their structural integrity under aqueous acidic or alkaline conditions94. The coordination bonds between metal ions and organic linkers play a crucial role in determining the overall chemical stability of the framework3,82,94. To prevent coordination bonds from breaking and subsequent structural, two methods have been developed to protect coordination bonds: constructing robust coordination bonds80,95,96,97,98, and preventing coordination bonds from being attacked80,99,100.
According to Pearson’s hard and soft acid-base (HSAB) theory, strong and stable bonds are preferentially formed when Lewis soft acids coordinate with Lewis soft bases, or when Lewis hard acids coordinate with Lewis hard bases101. Based on the HSAB theory, a series of stable MOFs are designed and synthesized80,95,96,97. Among these MOFs, iron(III)-tetraamino-benzoquinone (Fe-TABQ) demonstrated high chemical stability, retaining 95% of its capacity after 200 cycles as a LIB electrode, which DFT calculations attribute to a strong d-π orbital interaction between Fe and TABQ97. The Paddlewheel framework (Fig. 2a) has been proven to be a stable structure because of multiple bridging-chelating coordination and robust metal-metal bond80,98. Based on the HSAB theory, Chen et al. 102 designed and synthesized a series of {[Ln4(μ4-O)(μ3-OH)3(INA)3(GA)3](CF3SO3)(H2O)6}n (denoted as Ln4-MOFs, Ln = Gd, Tm, and Lu, INA = isonicotinic acid, GA = glycolic acid). These MOFs demonstrated remarkable stability across challenging conditions, including high moisture, acidic and basic pH environments, and extreme temperatures, while maintaining excellent conductivity, a characteristic rarely observed in highly stable MOFs102. Protecting coordination bonds away from the attack of guest molecules can also effectively maintain structural integrity. To enhance water stability, hydrophobic groups, such as methyl groups, are introduced into the ligands to prevent the degradation of coordination bonds by water molecules80,100. Huang et al. 99 designed two identical MOFs, one with cationic and another one with neutral frameworks, namely PFC-8 and PFC-9 (Fig. 2b). They found that PFC-8 exhibited higher chemical stability than PFC-9 under acidic, oxidative, reductive, and high ionic strength conditions which proved that a cationic framework can improve chemical stability. To investigate the protective mechanisms of the cationic framework on coordination bonds, they performed DFT calculations. As evidenced by the DFT calculation, the Ni center in PFC-9 adopts a square planar geometry, while in PFC-8, due to the coordination of Cl- ions, the Ni center forms an octahedral geometry of which the large steric hindrance protects the Ni-N bond from the reactive species. Besides, the Bader charge analysis revealed that the Ni ions in PFC-8 have a higher positive charge density. This increased positive charge enhances the repulsive interaction between the cationic framework and positively charged guest species, effectively protecting the fragile Ni–N bonds. The calculations of the binding energies showed that protons preferentially bind to the nitrogen atoms in PFC-9, leading to the cleavage of Ni–N bonds. In contrast, the presence of Cl- ions in PFC-8 reduces the corrosive effect on the Ni–N bonds, significantly improving its stability.
a Structure of Co-bpy and paddlewheel SBU. Reproduced with permission from Ref. 80 Copyright 2021 Wiley. b Schematic representation of the structures of cationic PFC-8 and neutral PFC-9. Reproduced with permission from Ref. 99 Copyright 2020 Wiley. c The upper line shows the mechanism of water absorption on dehydrated STAM-17-OEt. When water adsorbs onto the open metal site, energy is released and utilized to break the weak interactions between Cu and O. The bottom line illustrates the mechanism of coordination bonds breakdown in HKUST-1. During the Hydration, energy is released and used to break metal–carboxylate oxygen bonds in the paddlewheel. Reproduced with permission from ref. 103 Copyright 2018 Springer Nature.
A novel mechanism is reported to protect the coordination bonds103. STAM-17-OEt exhibits superior hydrolytic stability compared to HKUST-1, which has a similar chemical composition and structure. In STAM-17-OEt, sacrificial Cu-O bonds form during dehydration and break upon rehydration, allowing the structure to revert to its hydrated form as shown in Fig. 2c. In contrast, without these sacrificial bonds, the coordination bonds in HKUST-1 are broken, leading to structural degradation (Fig. 2c). Two idealized models are generated based on the experimental structures, one with weak Cu–O interactions and one without. The DFT results showed that the experimentally observed dehydrated model (with weak Cu–O interactions) is 0.228 eV lower in energy than the hypothetical dehydrated model, which lacks these weak interactions. The calculations demonstrated that the formation of weak Cu–O bonds not only contributes to the thermodynamic stability of the framework but also prevents the release of sufficient energy to break the more critical framework bonds.
Mechanical stability
Mechanical stability refers to the ability to maintain the structure integrity under various mechanical loads. During the activation process of MOFs, where solvents are removed from the pores, capillary forces often induce structural degradation94. In batteries and pseudo-capacitors, redox reactions lead to electrode expansion and contraction due to the insertion and extraction of ions93. These mechanical stresses can result in instability, leading to phase transitions, partial pore collapse, or even amorphization of the material. Therefore, it is crucial to enhance the mechanical stability of MOFs. Similar to chemical stability, mechanical stability benefits from robust coordination bonds and high connectivity. Cavka et al. 43 designed a zirconium-based MOF, UiO-66, which exhibits exceptional stability and they attributed the stability to the robust Zr–O coordination bonds. Redfern et al. 104 validate the conclusion that robust coordination bonds contribute to mechanical stability by exploring the influence of different metal nodes. They reported that the bulk modulus of Ce-UiO-66 is lower than that of Zr-UiO-66 and Hf-UiO-66. The DFT calculations revealed that the coordination bond (M-O bond) shortens more rapidly under pressure in Ce-UiO-66 than in Zr-UiO-66. This suggests that the Ce-O bond is weaker and more flexible, leading to lower bulk modules. Wu et al. 105 further identified high metal-organic coordination as another key factor behind UiO-66’s extremely high bulk modulus. Hobday et al. 106 explored the mechanical stability of two zirconium-based metal-organic frameworks (MOFs) with UiO topology, specifically UiO-67 and UiO-abdc (abdc: 4,4’-azobenzene dicarboxylate). The study combines both computational and experimental approaches to understand how disorder in the linkers affects their behavior under high pressure. The research reveals that dynamic disorder in the ligands has a significant impact on the mechanical properties of the framework. UiO-abdc, which features a more flexible and disordered azobenzene linker, shows a decrease in the elastic modulus and increase in resistance to pressure compared to UiO-67, which has a more rigid biphenyl dicarboxylate linker.
Redox activity of MOFs
MOF-based supercapacitors
Currently, MOFs are frequently used as electrode materials for supercapacitors and batteries. Supercapacitors are generally categorized into two types: electrical double-layer capacitors (EDLCs) and pseudo-capacitors (PCs)6. EDLCs store energy through the reversible adsorption of electrolyte ions onto the nanopores of the electrodes and form electrical double layers (EDLs)6. Conductive MOFs, due to their high surface area and conductivity, are well-suited as electrode materials for EDLCs. For example, Ni₃(HITP)₂ has demonstrated a high conductivity of 40 S cm−1, resulting in an excellent specific gravimetric capacitance of 111 F g−1 (at the discharge rate of 0.05 A g−1)20. The energy storage process of EDLCs is purely physical, involving no chemical reactions, and can be described using classical molecular dynamics simulations. Recent works on EDLCs have been reviewed in recent studies88,107.
Compared to EDLCs, PCs generally exhibit higher capacitance and energy density since they store energy through faradaic reactions occurring on the surface of electrode materials108,109. For instance, the typical redox-active MOF, Ni-HAB (HAB = hexaaminobenzene), exhibits a high capacitance of 420 F g−1, attributed to its pseudocapacitive charge-storage mechanism19,110. Therefore, significant efforts by DFT calculations have been devoted to understanding this pseudocapacitive mechanism in MOFs to further enhance capacitance and energy density111. Ni2[CuPcS8], Ni2[CuPc(NH)8], and Ni2[CuPcO8] exhibited different storage mechanisms in various electrolytes112,113. In a 1 M TEABF4/acetonitrile electrolyte, Ni2[CuPcS8] demonstrated an outstanding pseudo-capacitance of 312 F g−1. DFT calculations based on cluster models suggest that this enhanced capacitance is due to the localized LUMO of the NiS₄ linker (Fig. 3a), which facilitates efficient delocalization of the injected electrons113. In the aqueous electrolyte of 1 M Na2SO4, Ni2[CuPc(NH)8] demonstrated a higher capacitance of 400 F g−1. The calculated molecular electrostatic potential (MESP) for truncated MOF molecules indicates the presence of nucleophilic and electrophilic regions, which enable the adsorption of SO42- and Na+ ions (Fig. 3b). As shown in Fig. 3c, continuous Faradaic reactions occur through the adsorption/desorption of Na+ on the NiN4 linkers and the desorption/adsorption of SO42- on the CuPc building blocks112. Co-CAT and Ni-CAT (CAT = catecholate) also feature dual redox sites in a Na2SO3 electrolyte114. DFT calculations based on cluster models indicate that the binding energy for SO32− adsorption on Co-CAT is lower than that on Ni-CAT, leading to a more effective attraction and reaction of SO32-. Therefore, it is crucial to find a suitable match of MOFs and electrolytes to achieve higher capacitance. MOFs can be modeled either as periodic solids or approximated to clusters. The cluster approximation requires careful validation to ensure the truncated molecular model is sufficiently large to minimize errors115. It is important to note that the cluster approach may yield inaccurate results for properties that rely on the extended symmetry of the material116. On the other hand, for local properties, such as chemical bond analysis, the cluster approximation allows the application of high-level theoretical methods that are often infeasible for solid-state materials115. The comparison of these two methods in terms of accuracy, computational efficiency, and suitability for different scenarios has been discussed in detail in this review115.
a Spatial distribution of LUMOs in the Ni2[CuPcS8], Ni2[CuPcN8], and Ni2[CuPcO8]. Reproduced with permission from ref. 113 Copyright 2023 ACS. b MESP distribution of Ni2[CuPc(NH)8] Reproduced with permission from ref. 112 Copyright 2021 ACS. c Electronic states of the repeating unit of Ni2[CuPc(NH)8] during the charge/discharge process. Reproduced with permission from ref. 112 Copyright 2021 ACS.
MOF-based batteries
Conductive MOFs can also serve as electrodes in batteries, and a bunch of highly conductive have been synthesized and applied in metal-ion batteries117,118,119,120,121,122,123. However, the limited number of redox sites in MOFs results in low energy density. Therefore, designing MOFs with abundant redox-active sites has the potential to significantly enhance energy density, but remains a challenging task.
Redox reactions commonly occur on ligand cores and metal-organic linkages. In conductive 2D MOFs, the metal-organic linkages, CuO4 units, have been proven to effectively enhance the energy density of MOF-based batteries119,124,125,126,127,128,129,130. Dioxolene ligands and the metal Cu form the CuO4 units where the dioxolene moiety undergoes a series of oxidation steps between catechol, semiquinone, and quinone, transferring two electrons127. Additionally, the reduction of Cu2+ to Cu+ contributes a one-electron transfer, further enhancing the density of redox-active (Fig. 4a)124. In addition to metal-organic linkages, efforts have been made to incorporate more redox-active sites into ligands127,131. Incorporating quinone compounds into ligands is a potential approach to enhance the capacity of MOFs, as they contain numerous redox-active sites. Based on this approach, Yan et al. 127 incorporated tricycloquinazoline (TQ) as the ligands into the MOF, Cu-HHTQ, which achieved a high capacity of 657.6 mAh g−1 at 600 mA g−1 and outstanding cycle stability. Through the electrostatic potentials and binding energies of the lithiation of TQ, they confirmed their hypothesis that TQ will undergo a nine-electron reaction for Li+ storage. Similarly, quinone groups in Cu-TBPQ were demonstrated to be attractive to Zn2+ by the calculation of MESP distribution and experimental spectroscopic analysis131.
a Schematic diagram of the insertion and removal processes of Na+ in the (CuO4) units. Reproduced with permission from ref. 130 Copyright 2024 Wiley. b The binding sites of Li+ on Cu-HATN and the distribution electrostatic potential surfaces of Cu-HATN, Cu-HATN+12Li, Cu-HATN+36Li, and Cu-HATN+48Li. Reproduced with permission from ref. 129 Copyright 2023 Wiley. c DFT calculations on two possible distributions of sql-Cu-TBA-MOF with 4 K+ during redox reaction and the corresponding binding energies. Reproduced with permission from ref. 132 Copyright 2024 Wiley. d DFT calculations on three possible distributions of kgm-Cu-TBA-MOF with 6 K+ during redox reaction and the corresponding binding energies. Reproduced with permission from ref. 132 Copyright 2024 Wiley.
It is crucial not only to incorporate redox-active sites but also to maximize their exposure to ions to facilitate effective reactions. Two MOFs, Cu-HATN (HATN = hexaazatriphenylene) and Cu-TAC (TAC = triazacoronene) achieved high capacities of 763 mAh g−1 and 772.4 mAh g−1 at 300 mA g−1, which features redox-active sites exposed within the pores, making them more accessible to Li⁺ ions (Fig. 4b)128,129. Furthermore, MOFs with the same composition but different topologies can result in different numbers of accessible redox-active sites. For instance, the rhombus structure (sql-Cu-TBA-MOF) of Cu-TBA (TBA = octahydroxyltetrabenzoanthracene) exhibited higher capacity in potassium-ion batteries (PIBs) than the kagome structure (kgm-Cu-TBA-MOF) due to its higher surface area, uniform pore size (Fig. 4c), and larger layer space132. The calculated binding energies also revealed that sql-Cu-TBA-MOF is more attractive to K+ as shown in Fig. 4c, d132.
Except for the intercalation of the cations in the metal-ion batteries discussed above, studies have shown a novel storage mechanism in which both anions and cations are stored. Chen et al. 133 compared different storage mechanisms of Zn-HHTP and Cu-HHTP in SIBs. As the experiment observed, Cu will undergo a redox reaction of Cu2+/Cu+ during the charging process while Zn remains redox inactive133. Although Zn2+ is not involved in the redox reaction, the ligand of Zn-HHTP can successively store anions (PF6−) and cations (Na+) while the ligand of Cu-HHTP can only store cations133. The calculated intercalation/de-intercalation potential of PF6- in Zn-HHTP consists well with the experimental redox potential, which further proved the storage of PF6- in Zn-HHTP133. Both Zn-HHTP and Cu-HHTP went through a three-electron redox reaction and exhibited similar capacity, but Zn-HHTP exhibited better cyclability because of redox reactions of only ligands133. Similarly, Ni3(HATQ)2 (HATQ = 2,3,7,8,12,13-hexaaminotricycloquinazoline) achieved a co-storage mechanism of anions (PF6-) and cations (Na+)134. The electrostatic potentials and charge density distribution demonstrate that Na⁺ is attracted to the N atom of the NiN4 unit, while PF6- is drawn to the N atom of the TQ cores, with binding energy results further confirming this tendency134. Cheng et al. 132 reported a MOF with rhombus structure, sql-Cu-TBA which exhibited 146.6 mAh g−1 at 0.1 mA g−1 in PIBs. To gain insight into the charge storage mechanism, they performed cyclic voltammetry (CV) analysis, XPS spectra, and FT-IR spectra and found that both K+ and PF6− could be stored132. They further performed DFT calculations and the calculated reaction voltages accord with the experiment results which demonstrated that 3 K+ and 1 PF6− are combined with each CuO4 unit132. They also calculated the binding energies of different stages during the interaction of K+ and PF6- with sql-Cu-TBA132. The negative values at each stage indicate that the structure is stable throughout the process and the process is feasible132.
ML Workflow for Predicting MOF properties
ML can be divided into three categories: supervised learning, unsupervised learning, and reinforcement learning38,41,135. In supervised learning, both features and labels of materials are available, allowing the establishment of relationships between features and labels41. In unsupervised learning, only features are accessible, enabling the discovery of patterns within the data41. Reinforcement learning focuses on finding optimal behaviors that maximize rewards through an agent’s interaction with its environment136. Supervised learning is the most widely used form of ML today, and it will be the focus of our discussion. Algorithms in machine learning can be broadly divided into two categories: shallowing learning algorithms and deep learning algorithms137. Shallowing learning algorithms, such as support vector machines (SVM)138,139, k-nearest neighbors (KNN)140, decision trees (DT)141, and artificial neural network (ANN)142, rely heavily on expert knowledge to design and generate meaningful descriptors to represent materials features. These algorithms work well with small datasets137. Additionally, they are highly interpretable, making it easier to understand and explain how predictions are made137. With the emergence of big datasets composed of hundreds of thousands of samples, deep learning algorithms, such as convolutional neural networks (CNNs), transformers143, and recurrent neural networks (RNNs), become achievable. They can automatically extract features directly from raw data, eliminating the need for manual feature engineering and facilitating reliable mapping relations between features and targets137. Nevertheless, it is challenging to interpret deep learning models, due to the elusive physical meaning of descriptors and convoluted mapping relations34. Therefore, some methods have been developed to interpret the model34, such as SHAP (SHapley Additive exPlanations)144.
To apply ML in predicting MOF properties, we first need to collect or generate data on these properties. Next, we extract the relevant features of MOFs based on the specific problem and algorithm. Finally, these features and properties are used as inputs to the algorithm to establish the relationships between them. The model generally needs to be validated before being used for making predictions. Typically, the dataset is divided into a training set and a test set, where the training set is used for model training, and the test set is used to evaluate its performance38. High-quality and abundant data, feature engineering, and ML algorithms are essential for ML models to accurately learn the relationships and predict the properties. Furthermore, experts in material science not only need MOFs with high performance but also the influence factors of MOF properties to design new MOFs.
Constructing databases of MOFs
The challenge of rapidly collecting material data has driven the development of several platforms that compile and provide access to extensive material datasets. These platforms, such as Materials Project (MP)145, AFLOW146, OQMD147, Materials Cloud148, and NOMAD149, aim to make material data readily accessible and reusable for users. The quantity and quality of data are crucial for prediction accuracy, which presents a significant challenge in the field of MOFs. Data on MOF structures and properties collected from high-throughput calculations or experiments can often be inconsistent, incomplete, and sourced from different origins38,41. To address this, databases of MOF properties have been developed to facilitate the application of ML in MOF research. The Cambridge Crystallographic Data Centre (CCDC) has characterized various MOFs and generated the Cambridge Structural Database (CSD) subset, which is continuously updated with newly experimental synthesized and reported MOFs150,151,152. The experimental structures of MOFs normally contain solvent molecules and partially occupied or disordered atoms, making it difficult to calculate. In 2014, the Computation-Ready Experimental (CoRE) MOF database was established, comprising approximately 4,700 MOFs from the CSD153. By removing solvent molecules and correcting partially occupied or disordered atoms, MOFs in the CoRE database are well-suited for high-throughput computational screening. Recently, the CoRE MOF database was updated to include over 14,000 MOF structures154. With the structures of MOFs, high-throughput DFT calculations can be performed and various properties of MOFs can be obtained. Rosen et al. 155 introduced the QMOF database based on PBE functional, which includes quantum-chemical properties for over 14,000 MOFs gathered from the CSD and CoRE database, such as optimized geometries, band gaps, density of states, charge densities, partial atomic charges, and net magnetic moments.
Several methods have been developed for generating hypothetical MOFs156,157. Notable examples include the ToBaCCo158, ToBasCCo159, and PORMAKE methods160, which have been employed to generate hundreds of thousands of potential MOFs. Moosavi et al. 161 generated 200 hypothetical MOFs based on the ToBasCCo method and calculated mechanical properties using molecular dynamics simulations. Majumdar et al. 162 produced approximately 20,000 hypothetical MOFs using the ToBaCCo method and evaluated their adsorption performance for CO2 and H2, as well as their effectiveness in separating CO2 from flue gas using Grand-canonical Monte Carlo (GCMC) simulations. Nandy et al. 163 employed ML models to identify stable MOFs, which they deconstructed into organic nodes, organic edges, and inorganic nodes. Using PORMAKE, they recombined these building blocks to generate 50,000 hypothetical MOFs163. Elastic modulus calculations for nearly 10,000 MOFs demonstrated their excellent mechanical stability163. Zhang et al. 164 recombined the organic linkers and metal nodes to generate 1057 bulk and monolayer structures of 2D conductive MOFs, termed EC-MOF database. This database provides various properties of electrically conductive MOFs calculated based on DFT, such as band gap, d-band center, formation energy, largest cavity diameter, gravimetric and volumetric surface area, and void fraction. Burner et al. 165 compiled the ARC-MOF database with approximately 280,000 MOFs, including experimental and computationally generated structures. It features DFT-derived partial atomic charges using the REPEAT method and provides ML-ready descriptors165.
There are also some databases that gather experimentally characterized properties of MOFs from the literature. Nandy et al. 166,167 used natural language processing (NLP) to identify manuscripts reporting solvent removal stability or thermogravimetric analysis data, assembling a dataset of 2179 MOFs with solvent removal stability and 3,132 with thermal decomposition temperatures. Iacomi and Llewellyn168 processed adsorption isotherms from the NIST Isotherm Database (ISODB)169, a repository of experimental and simulated adsorption isotherms collected from the scientific literature. To prepare the data for meaningful analysis and potential machine learning applications, they implemented stringent filtering criteria to ensure relevance and comparability among the isotherms168. Luo et al. 170 used NLP to extract MOF synthesis conditions from the literature, establishing the SynMOF database which contains 983 MOF structures and their synthesis details, including the metal source, linkers, temperature, synthesis time, solvents, and additives).
Feature engineering
In materials science, effective representation of materials is important, as it contains essential information that allows ML models to establish robust relationships between input features and target properties171,172. The process of feature extraction varies depending on the problem and the algorithm used41,171,172. Classical machine learning models rely heavily on domain experts to engineer features that are highly relevant to the problem at hand, allowing for solid performance with fewer data than deep learning typically requires41. In contrast, deep learning integrates feature extraction directly into the model’s learning process, automatically discovering useful representations41,135. As a result, deep learning models reduce the dependence on expert-driven feature engineering but demand significantly more data172. Deep learning can potentially uncover patterns or regularities that are not immediately apparent to human experts137. For MOFs, which are highly ordered crystalline solids, an ML model can recognize the similarities and capture the differences among the structures in the training set and leverage that insight to predict the properties of new structures without calculation or experimental measurements173. However, representing this complex structural information in a form suitable for ML models presents significant challenges due to the need to capture both local atomic environments and long-range periodicity inherent to crystalline materials.
MOFs are composed of secondary building units (SBUs) and organic linkers, which can be either characterized separately or treated as extended periodic crystals115. Janet et al. 174 introduced revised autocorrelation functions (RACs) to capture molecular characteristics based on molecular graphs. Moosavi et al. 175 subsequently revised the RACs to enable their application to MOFs by partitioning the frameworks into SBUs, organic linkers, and functional groups. By calculating RACs for each of these components separately, they generated a unique feature for each MOF. Besides, MOFs are characterized as highly ordered, periodic extended MOFs. Because MOFs are highly ordered and inherently periodic, preserving their translational, rotational, and permutation symmetries when describing their structures remains a critical challenge41. Efforts have been made to describe the structures of crystals. In 2014, Schütt et al. 176 reported a representation that can describe crystal structures better than Coulomb matrices. They proposed partial radial distribution functions (PRDF), which capture the distribution of pairwise atomic distances within a crystal (Fig. 5a) and are invariant to translation, rotation, and the choice of the unit cell176. Faber et al. 177 reported three ways to represent periodic systems: the Ewald sum matrix, extended coulomb-like matrix, and sine matrix. The Ewald sum matrix is an extension of the coulomb matrix used in molecular systems. Each element of the matrix is the Ewald sum, which separates the electrostatic interaction into a short-range component and a long-range component. In this way, the Ewald sum matrix not only represents the infinite repetition of atoms within a crystal system but also ensures convergence in the calculation of these long-range electrostatic interactions. An extended coulomb-like matrix modifies the traditional coulomb matrix by including interactions with atoms in neighboring unit cells to account for periodic systems. The key difference from the Ewald sum matrix is that this method does not aim to compute the interaction over the entire infinite lattice but rather considers only a finite number of neighboring unit cells. Therefore, this method is more computationally efficient than the Ewald sum matrix but less accurate in long-range interactions. The sine matrix is based on using a sine function to model the periodic nature of crystal structures. It is designed to be simpler and computationally cheaper while still capturing the essential periodicity of the atomic arrangement. Schütt and Faber developed methods to modify the Coulomb matrix for representing periodic systems, but these approaches did not account for bond angles. Seko et al. 178 combined elemental and structural representations to describe the chemical composition and atomic arrangement of materials. Elemental properties include atomic number, atomic mass, first ionization energy, electron affinity, covalent radius, and other atomic characteristics. Structure properties include PRDF, generalized radial distribution functions (GRDF), bond-orientational parameters (BOP), and the angular Fourier series (AFS). BOP provides information about local symmetry and the relative orientations of neighboring atoms, while AFS integrates both radial and angular information to describe the full spatial distribution of atoms. Isayev et al. 179 proposed new descriptors called Property-Labelled Materials Fragments (PLMF), which converted the crystal structure into a graph. They used Voronoi Tessellation to partition crystal structures into atom-centered polyhedrals. If atoms share the Voronoi face and are within a certain distance (based on the sum of covalent radii), atoms are considered to be bonded (Fig. 5b). Based on this method, the crystal structure is transformed into a graph containing connectivity within the material. The full graph is divided into two kinds of fragments, path fragments and circular fragments (Fig. 5b). Each fragment is assigned elemental properties based on the atoms it contains. In addition, global structural properties of the entire crystal are also included, such as lattice parameters, angles, space group, and point group.
a Alternative crystal representations. Top: a crystal unit cell indicating the Bravais vectors (blue) and base (pink). Bottom: illustration of one shell of the discrete partial radial distribution function with α and β representing different atom types. Reproduced with permission from ref. 176 Copyright 2014 American Physical Society. b Schematic representing the construction of the Property-Labelled Materials Fragments (PLMF). Reproduced with permission from ref. 179 Copyright 2017 Authors, licensed under a CC BY license. c Schematic construction of the crystal graph with node vectors representing atoms and edge vectors representing bonds. Reproduced with permission from ref. 180 Copyright 2018 American Physical Society.
The descriptors mentioned above, based on classical ML models such as kernel ridge regression, have performed well in predicting various properties of materials. In contrast to classical ML models, Deep learning can automatically learn representations from raw data, reducing the reliance on expert-designed features41. Xie and Grossman180 proposed crystal graph convolutional neural networks (CGCNN) to represent and predict the properties of periodic systems. They employed Isayev’s method to transform the crystal structure into a crystal graph, where the nodes represent atoms and the edges represent the connections between them (Fig. 5c)180. Both the nodes and edges are characterized by feature vectors that encode the information of atoms and bonds, respectively180. Different from the traditional ML model, CGCNN is capable of automatically identifying and learning the most critical characteristics of a material for predicting target properties. Initially, each node starts with basic information about a specific atom180. When the model performs its first “convolution”, the node gathers more details, incorporating not only information about the atom itself but also about the bonds and neighboring atoms within a defined radius180. As this process is repeated through several convolutions, each node ends up with rich information about its surrounding chemical environment180. The number of convolutions determines how far this information reaches, whether it’s just local details or a broader structure180. After this process, the model combines all this information to create a new feature that describes the material180. Through training, this feature is refined to represent the material in a way that best predicts the target properties180. Similarly, Chen et al. 181 proposed MatErials Graph Network (MEGNet) models that utilize graph networks to represent both molecules and crystals. Different from CGCNN, MEGNet incorporates global state variables as additional descriptors to enhance the prediction of properties that depend on external conditions and establish relationships between structure, state, and properties181. MEGNet incorporates temperature, pressure, and entropy as global states and developed a unified molecule free energy model that can predict internal energy at 0k and room temperature, enthalpy, and Gibbs free energy181.
ML models and applications in screening MOFs
ML model plays a crucial role in accurately predicting the properties of MOFs. For a detailed introduction to algorithms, please refer to the following articles38,41,135. We will focus on introducing ML models developed for MOFs, as well as the application of ML techniques for identifying MOFs suitable for EES systems. The transformer architecture is useful in processing long sequences efficiently143. Based on transformer architecture, MOFormer and MOFTransformer are proposed to predict the properties of MOFs182,183. During the self-supervised pretraining phase, MOFormer and CGCNN are jointly employed to capture comprehensive information about the MOFs182. After pretraining, the learned weights are transferred to MOFormer for fine-tuning on downstream tasks182. MOFormer using text string representation (MOIFid) outperforms the structure-agnostic descriptors in predicting band gaps after self-supervised pretraining182. As for MOFTransformer, it incorporates atom-based graph embeddings to capture local features and energy-grid embeddings to capture global features during pretraining183. Atom-based graph embeddings are generated using CGCNN without pooling layers and energy-grid embeddings are calculated using energy grids derived from the interaction between MOFs and methane molecules183. Subsequently, MOFTransformer was integrated into ChatMOF as a predictor184.
Recently, many efforts have been devoted to exploring the properties of MOFs using ML techniques166,185,186. He et al. 185 first applied ML to the exploration of conductive MOFs in 2018. They trained four ML models, logistic regression (LR), support vector classification (SVC), NN, and random forest (RF), using a dataset of 52300 inorganic materials. These models were subsequently applied to screen 2937 MOFs, predicting their band gaps through a multivoting approach that combined the outputs of the four models. Nine MOFs were predicted to be metallic and 6 of them were confirmed to be metallic by the calculations based on semilocal DFT. Dou et al. 186 established a database containing experimentally measured conductivity of 224 MOFs and trained ML models based on the database. These models were subsequently applied to predict the conductivity of the MOFs in the QMOF database, identifying CuTTPD as a candidate with a predicted conductivity of 10−3.3 S/cm. They synthesized CuTTPD and measured the conductivity (10−5.77 S/cm) which is close to the prediction186. Furthermore, they employed SHAP values to identify the influence factors of conductivity, which revealed that the number of hydrogen bond donors (HBD) and the acid dissociation constant (pKa) were the most significant descriptors influencing conductivity186. Notably, these descriptors have been rarely considered in the study of conductive MOFs186.
As for stability, Moghadam et al. 187 conducted high-throughput molecular simulations on a dataset of 3385 MOFs to calculate their mechanical properties and trained ANN using this dataset, enabling rapid and accurate prediction of the MOF bulk modulus. Based on the dataset, certain topologies, high coordination numbers, shorter linkers, and smaller pore sizes are found to contribute to greater mechanical stability187. Batra et al. 188 developed a dataset containing water stability data for 207 MOFs and trained classification models. Their analysis revealed that the number of cyclic divalent nodes (MQNs30) and six-membered rings (MQNs36) were key descriptors influencing water stability188. In particular, MOFs exhibited high water stability when MQN36 was less than 0.04 and MQN30 exceeded 0.01. Subsequently, Zhang et al. 189 constructed a more extensive dataset comprising 1133 MOFs with water stability information. Their study identified critical factors affecting water stability, such as high surface area and narrow pore size, metal with high atomic mass and radius, linker with high connectivity degree, and electronegativity variation. The trained model was subsequently applied to the ARC-MOF database, identifying 129,661 MOFs with potential water stability165,189.
Conclusions and Perspectives
Herein, we give an overview of applying DFT and ML in predicting the properties of MOFs to screen and design high-performance MOFs for EES systems. Our discussion focuses on the conductivity, stability, and redox activity of MOFs which are essential for the performance of the MOF-based electrode materials. Based on DFT calculations, we discuss the mechanisms of charge transport, strategies for improving conductivity, chemical and mechanical stability, and the processes and mechanisms of redox reactions. Additionally, we give an overview of ML workflow to efficiently screen MOFs for EES systems. Specifically, we discussed the database, descriptors, and models developed for the applications of ML in screening MOFs.
It is worth noting that current calculated structures of MOFs are typically idealized as perfect crystals, whereas experimentally synthesized MOFs are often polycrystalline. This discrepancy leads to inconsistencies between calculated and experimental results. Therefore, future research should focus on adjusting calculated MOF structures to better reflect their actual forms, while experimental efforts should aim to synthesize single-crystal MOFs to investigate their intrinsic conductive mechanisms.
3D MOFs generally exhibit high specific surface areas and more accessible redox-active sites. However, 3D conductive MOFs are relatively rare, as the most effective methods for enhancing conductivity—extended conjugation and through-space approaches—are challenging to incorporate into 3D structures190. Identifying more conductive 3D MOFs through high-throughput calculations and ML is essential to explore the factors influencing conductivity and to elucidate the conductive mechanisms.
The customizable structures of MOFs enable them to achieve higher energy densities, making them promising for energy storage applications. For instance, Cu-THQ (THQ = benzoquinoid), when employed as a cathode material for Li-ion batteries, demonstrated a high reversible capacity of 387 mAh g−1, surpassing the theoretical capacities of many transition metal compounds124. Similarly, 2D Ni-MOF nanofilaments, when used as anode materials for Li-ion batteries, achieved a remarkable specific capacity of 719 mAh g−1 at 1 A g−1, exceeding the theoretical energy density of graphite191. While these advancements underscore the potential of MOFs for higher energy densities compared to standard materials, their lower cycling stability poses a significant limitation for practical applications22. Furthermore, achieving high stability seems to conflict with the requirement for high conductivity because the strong coordination bonds, typically designed based on the HSAB theory, tend to localize electrons, resulting in low conductivity or even insulating behavior102. Thus, simultaneously incorporating high conductivity, high stability, and high redox activity in MOF-based materials remains a key challenge.
Rather than screening for MOFs, designing MOF structures tailored for specific functionalities is more desirable. Achieving this goal requires a deep understanding of the relationship between MOF structure and functionality. While advancements in ML have improved predictive accuracy, they have often come at the expense of interpretability, turning the structure-function relationship into a black box34. This challenge presents a promising avenue for future research: developing explainable ML methods to better elucidate these relationships. Integrating domain knowledge of MOFs into ML models offers an effective approach to enhancing both model performance and interpretability192,193,194.
Data availability
Data are available from the referenced original sources.
References
Shao, Y. et al. Design and mechanisms of asymmetric supercapacitors. Chem. Rev. 118, 9233–9280 (2018).
Wang, F. et al. Latest advances in supercapacitors: from new electrode materials to novel device designs. Chem. Soc. Rev. 46, 6816–6854 (2017).
Kong, L., Liu, M., Huang, H., Xu, Y. & Bu, X.-H. Metal/covalent-organic framework based cathodes for metal-ion batteries. Adv. Energy Mater. 12, 2100172 (2022).
Wang, L. et al. Metal–organic frameworks for energy storage: batteries and supercapacitors. Coord. Chem. Rev. 307, 361–381 (2016).
Aricò, A. S., Bruce, P., Scrosati, B., Tarascon, J.-M. & van Schalkwijk, W. Nanostructured materials for advanced energy conversion and storage devices. Nat. Mater. 4, 366–377 (2005).
Simon, P. & Gogotsi, Y. Materials for electrochemical capacitors. Nat. Mater. 7, 845–854 (2008).
Simon, P. & Gogotsi, Y. Perspectives for electrochemical capacitors and related devices. Nat. Mater. 19, 1151–1163 (2020).
Kim, S.-W., Seo, D.-H., Ma, X., Ceder, G. & Kang, K. Electrode materials for rechargeable sodium-ion batteries: potential alternatives to current lithium-ion batteries. Adv. Energy Mater. 2, 710–721 (2012).
Tarascon, J. M. & Armand, M. Issues and challenges facing rechargeable lithium batteries. Nature 414, 359–367 (2001).
Goodenough, J. B. & Kim, Y. Challenges for rechargeable Li batteries. Chem. Mater. 22, 587–603 (2010).
Hong, C. N., Crom, A. B., Feldblyum, J. I. & Lukatskaya, M. R. Metal-organic frameworks for fast electrochemical energy storage: mechanisms and opportunities. Chem. 9, 798–822 (2023).
Lu, Y. & Chen, J. Prospects of organic electrode materials for practical lithium batteries. Nat. Rev. Chem. 4, 127–142 (2020).
Lu, Y., Zhang, Q., Li, L., Niu, Z. & Chen, J. Design strategies toward enhancing the performance of organic electrode materials in metal-ion batteries. Chem. 4, 2786–2813 (2018).
Zeng, X., Zhan, C., Lu, J. & Amine, K. Stabilization of a high-capacity and high-power nickel-based cathode for Li-ion batteries. Chem. 4, 690–704 (2018).
He, B. et al. Freestanding metal-organic frameworks and their derivatives: an emerging platform for electrochemical energy storage and conversion. Chem. Rev. 122, 10087–10125 (2022).
Cai, G., Yan, P., Zhang, L., Zhou, H.-C. & Jiang, H.-L. Metal–organic framework-based hierarchically porous materials: synthesis and applications. Chem. Rev. 121, 12278–12326 (2021).
Du, W. et al. Advanced metal-organic frameworks (MOFs) and their derived electrode materials for supercapacitors. J. Power Sources 402, 281–295 (2018).
Jiao, L., Seow, J. Y. R., Skinner, W. S., Wang, Z. U. & Jiang, H.-L. Metal–organic frameworks: structures and functional applications. Mater. Today 27, 43–68 (2019).
Feng, D. et al. Robust and conductive two-dimensional metal-organic frameworks with exceptionally high volumetric and areal capacitance. Nat. Energy 3, 30–36 (2018).
Sheberla, D. et al. Conductive MOF electrodes for stable supercapacitors with high areal capacitance. Nat. Mater. 16, 220–224 (2017).
Shin, S.-J., Gittins, J. W., Balhatchet, C. J., Walsh, A. & Forse, A. C. Metal–organic framework supercapacitors: challenges and opportunities. Adv. Funct. Mater. 34, 2308497 (2024).
Wu, C., Geng, P., Zhang, G., Li, X. & Pang, H. Synthesis of conductive MOFs and their electrochemical application. Small 20, 2308264 (2024).
Wang, J. et al. Superior charge transport in Ni-diamine conductive MOFs. J. Am. Chem. Soc. 146, 20500–20507 (2024).
Xie, L. S., Skorupskii, G. & Dinca, M. Electrically conductive metal-organic frameworks. Chem. Rev. 120, 8536–8580 (2020).
Sun, L., Campbell, M. G. & Dincă, M. Electrically conductive porous metal–organic frameworks. Angew. Chem. Int. Ed. 55, 3566–3579 (2016).
Yang, D. & Gates, B. C. Catalysis by metal organic frameworks: perspective and suggestions for future research. ACS Catal. 9, 1779–1798 (2019).
Sumida, K. et al. Carbon dioxide capture in metal-organic frameworks. Chem. Rev. 112, 724–781 (2012).
Bi, S. et al. Molecular understanding of charge storage and charging dynamics in supercapacitors with MOF electrodes and ionic liquid electrolytes. Nat. Mater. 19, 552–558 (2020).
He, Q., Yu, B., Li, Z. & Zhao, Y. Density functional theory for battery materials. Energy Environ. Mater. 2, 264–279 (2019).
Nath, A., Asha, K. S. & Mandal, S. Conductive metal-organic frameworks: Electronic structure and electrochemical applications. Chem. – A Eur. J. 27, 11482–11538 (2021).
Kansara, S., Kang, H., Ryu, S., Sun, H. H. & Hwang, J.-Y. Basic guidelines of first-principles calculations for suitable selection of electrochemical Li storage materials: a review. J. Mater. Chem. A 11, 24482–24518 (2023).
Demuth, M. C. & Hendon, C. H. Linker aromaticity reduces band dispersion in 2D conductive metal–organic frameworks. ACS Mater. Lett. 5, 1476–1480 (2023).
Park, G., Demuth, M. C., Hendon, C. H. & Park, S. S. Acid-dependent charge transport in a solution-processed 2D conductive metal-organic framework. J. Am. Chem. Soc. 146, 11493–11499 (2024).
Zhong, X. et al. Explainable machine learning in materials science. npj Computational Mater. 8, 204 (2022).
Xu, Y. et al. High-throughput calculations of magnetic topological materials. Nature 586, 702–707 (2020).
Ludwig, A. Discovery of new materials using combinatorial synthesis and high-throughput characterization of thin-film materials libraries combined with computational methods. npj Computational Mater. 5, 70 (2019).
Choudhary, K. et al. Recent advances and applications of deep learning methods in materials science. npj Computational Mater. 8, 59 (2022).
Chen, A., Zhang, X. & Zhou, Z. Machine learning: accelerating materials development for energy storage and conversion. InfoMat 2, 553–576 (2020).
Rosen, A. S., Notestein, J. M. & Snurr, R. Q. Realizing the data-driven, computational discovery of metal-organic framework catalysts. Curr. Opin. Chem. Eng. 35, 100760 (2022).
Demir, H., Daglar, H., Gulbalkan, H. C., Aksu, G. O. & Keskin, S. Recent advances in computational modeling of MOFs: from molecular simulations to machine learning. Coord. Chem. Rev. 484, 215112 (2023).
Jablonka, K. M., Ongari, D., Moosavi, S. M. & Smit, B. Big-data science in porous materials: materials genomics and machine learning. Chem. Rev. 120, 8066–8129 (2020).
Sun, L., Hendon, C. H., Minier, M. A., Walsh, A. & Dinca, M. Million-fold electrical conductivity enhancement in Fe2(DEBDC) versus Mn2(DEBDC) (E = S, O). J. Am. Chem. Soc. 137, 6164–6167 (2015).
Cavka, J. H. et al. A new zirconium inorganic building brick forming metal organic frameworks with exceptional stability. J. Am. Chem. Soc. 130, 13850–13851 (2008).
Sheberla, D. et al. High electrical conductivity in Ni3(2,3,6,7,10,11-hexaiminotriphenylene)2, a semiconducting metal–organic graphene analogue. J. Am. Chem. Soc. 136, 8859–8862 (2014).
Dou, J.-H. et al. ignature of metallic behavior in the metal–organic frameworks M3(hexaiminobenzene)2 (M = Ni, Cu). J. Am. Chem. Soc. 139, 13608–13611 (2017).
Darago, L. E., Aubrey, M. L., Yu, C. J., Gonzalez, M. I. & Long, J. R. Electronic conductivity, ferrimagnetic ordering, and reductive insertion mediated by organic mixed-valence in a ferric semiquinoid metal-organic framework. J. Am. Chem. Soc. 137, 15703–15711 (2015).
Last, B. J. & Thouless, D. J. Percolation theory and electrical conductivity. Phys. Rev. Lett. 27, 1719–1721 (1971).
Mott, N. F. Electrons in disordered structures. Adv. Phys. 16, 49–144 (1967).
Shklovskiĭ, B. & Éfros, A. L. V. Percolation theory and conductivity of strongly inhomogeneous media. Sov. Phys. Uspekhi 18, 845 (1975).
Stallinga, P. Electronic transport in organic materials: comparison of band theory with percolation/(variable range) hopping theory. Adv. Mater. 23, 3356–3362 (2011).
Kittel, C. & McEuen, P. Introduction to Solid State Physics, 8 (Wiley New York, 1996).
Dong, R. et al. A coronene-based semiconducting two-dimensional metal-organic framework with ferromagnetic behavior. Nat. Commun. 9, 2637 (2018).
Le, K. N., Mancuso, J. L. & Hendon, C. H. Electronic challenges of retrofitting 2D electrically conductive MOFs to form 3D conductive lattices. ACS Appl. Electron. Mater. 3, 2017–2023 (2021).
Saha, S. et al. Ag nanoparticles-induced metallic conductivity in thin films of 2D metal-organic framework Cu3(HHTP)2. Nano Lett. 23, 9326–9332 (2023).
Meng, Z., Aykanat, A. & Mirica, K. A. Welding metallophthalocyanines into bimetallic molecular meshes for ultrasensitive, low-power chemiresistive detection of gases. J. Am. Chem. Soc. 141, 2046–2053 (2019).
Li, W., Walther, C. F. J., Kuc, A. & Heine, T. Density functional theory and beyond for band-gap screening: performance for transition-metal oxides and dichalcogenides. J. Chem. Theory Comput. 9, 2950–2958 (2013).
Hutchinson, M. & Widom, M. VASP on a GPU: application to exact-exchange calculations of the stability of elemental boron. Computer Phys. Commun. 183, 1422–1426 (2012).
Yu, M., Yang, S., Wu, C. & Marom, N. Machine learning the Hubbard U parameter in DFT+U using bayesian optimization. npj Computational Mater. 6, 180 (2020).
Aulbur, W. G., Städele, M. & Görling, A. Exact-exchange-based quasiparticle calculations. Phys. Rev. B 62, 7121–7132 (2000).
Hedin, L. New method for calculating the one-particle Green’s function with application to the electron-gas problem. Phys. Rev. 139, A796–A823 (1965).
Nisar, J., Århammar, C., Jämstorp, E. & Ahuja, R. Optical gap and native point defects in kaolinite studied by the GGA-PBE, HSE functional, and GW approaches. Phys. Rev. B 84, 075120 (2011).
Yang, J., Falletta, S. & Pasquarello, A. One-shot approach for enforcing piecewise linearity on hybrid functionals: application to band gap predictions. J. Phys. Chem. Lett. 13, 3066–3071 (2022).
Skorupskii, G. et al. Efficient and tunable one-dimensional charge transport in layered lanthanide metal-organic frameworks. Nat. Chem. 12, 131–136 (2020).
Clough, A. J. et al. Metallic conductivity in a two-dimensional cobalt dithiolene metal-organic framework. J. Am. Chem. Soc. 139, 10863–10867 (2017).
Wang, Z. et al. Interfacial synthesis of layer-oriented 2D conjugated metal–organic framework films toward directional charge transport. J. Am. Chem. Soc. 143, 13624–13632 (2021).
Debela, T. T., Yang, M. C. & Hendon, C. H. Ligand-mediated hydrogenic defects in two-dimensional electrically conductive metal-organic frameworks. J. Am. Chem. Soc. 145, 11387–11391 (2023).
Chen, S., Dai, J. & Zeng, X. C. Metal-organic kagome lattices M3(2,3,6,7,10,11-hexaiminotriphenylene)2 (M = Ni and Cu): from semiconducting to metallic by metal substitution. Phys. Chem. Chem. Phys. 17, 5954–5958 (2015).
Foster, M. E., Sohlberg, K., Spataru, C. D. & Allendorf, M. D. Proposed modification of the graphene analogue Ni3(HITP)2 to yield a semiconducting material. J. Phys. Chem. C. 120, 15001–15008 (2016).
Zhao, B., Zhang, J., Feng, W., Yao, Y. & Yang, Z. Quantum spin hall and Z2 metallic states in an organic material. Phys. Rev. B 90, 201403 (2014).
Foster, M. E., Sohlberg, K., Allendorf, M. D. & Talin, A. A. Unraveling the semiconducting/metallic discrepancy in Ni3(HITP)2. J. Phys. Chem. Lett. 9, 481–486 (2018).
Zhang, Z., Dell’Angelo, D., Momeni, M. R., Shi, Y. & Shakib, F. A. Metal-to-semiconductor transition in two-dimensional metal–organic frameworks: an ab initio dynamics perspective. ACS Appl. Mater. Interfaces 13, 25270–25279 (2021).
Ogle, J. et al. Semiconducting to metallic electronic landscapes in defects-controlled 2D π-d conjugated coordination polymer thin films. Adv. Funct. Mater. 31, 2006920 (2021).
Luo, Y. et al. Defect engineering to tailor metal vacancies in 2D conductive metal–organic frameworks: an example in electrochemical sensing. ACS Nano 16, 20820–20830 (2022).
Day, R. W. et al. Single crystals of electrically conductive two-dimensional metal–organic frameworks: structural and electrical transport properties. ACS Cent. Sci. 5, 1959–1964 (2019).
Skorupskii, G. et al. Porous lanthanide metal–organic frameworks with metallic conductivity. Proc. Natl Acad. Sci. 119, e2205127119 (2022).
Kharod, R. A., Andrews, J. L. & Dincă, M. Teaching metal-organic frameworks to conduct: ion and electron transport in metal-organic frameworks. Annu. Rev. Mater. Res. 52, 103–128 (2022).
Pathak, A. et al. Integration of a (–Cu–S–)n plane in a metal–organic framework affords high electrical conductivity. Nat. Commun. 10, 1721 (2019).
Yang, M. et al. Two-dimensional conjugated metal–organic frameworks with a ring-in-ring topology and high electrical conductance. Angew. Chem. Int. Ed. 63, e202405333 (2024).
Zhang, J. et al. Wavy two-dimensional conjugated metal–organic framework with metallic charge transport. J. Am. Chem. Soc. 145, 23630–23638 (2023).
Xia, Z. et al. Tailoring electronic structure and size of ultrastable metalated metal–organic frameworks with enhanced electroconductivity for high-performance supercapacitors. Angew. Chem. Int. Ed. 60, 10228–10238 (2021).
Jiao, Y. et al. Mixed-metallic MOF based electrode materials for high performance hybrid supercapacitors. J. Mater. Chem. A 5, 1094–1102 (2017).
Pan, Y. et al. Bimetallic electronic effects of Mn-doped Ni-MOF shuttle-like nanosheets remarkably enhance the supercapacitive performance. Inorg. Chem. Front. 9, 5982–5993 (2022).
Dou, J. H. et al. Atomically precise single-crystal structures of electrically conducting 2D metal-organic frameworks. Nat. Mater. 20, 222–228 (2021).
Hiszpanski, A. M. et al. Halogenation of a nonplanar molecular semiconductor to tune energy levels and bandgaps for electron transport. Chem. Mater. 27, 1892–1900 (2015).
Johnson, B. A., Bhunia, A., Fei, H., Cohen, S. M. & Ott, S. Development of a UiO-type thin film electrocatalysis platform with redox-active linkers. J. Am. Chem. Soc. 140, 2985–2994 (2018).
Li, J., Kumar, A., Johnson, B. A. & Ott, S. Experimental manifestation of redox-conductivity in metal-organic frameworks and its implication for semiconductor/insulator switching. Nat. Commun. 14, 4388 (2023).
Li, J., Kumar, A. & Ott, S. Diffusional electron transport coupled to thermodynamically driven electron transfers in redox-conductive multivariate metal-organic frameworks. J. Am. Chem. Soc. 146, 12000–12010 (2024).
Castner, A. T. et al. Microscopic insights into cation-coupled electron hopping transport in a metal-organic framework. J. Am. Chem. Soc. 144, 5910–5920 (2022).
Sun, L. et al. Is iron unique in promoting electrical conductivity in MOFs? Chem. Sci. 8, 4450–4457 (2017).
Xie, L. S. et al. Tunable mixed-valence doping toward record electrical conductivity in a three-dimensional metal-organic framework. J. Am. Chem. Soc. 140, 7411–7414 (2018).
Talin, A. A. et al. Tunable electrical conductivity in metal-organic framework thin-film devices. Science 343, 66–69 (2014).
Yadav, A., Zhang, S., Benavides, P. A., Zhou, W. & Saha, S. Electrically conductive π-intercalated graphitic metal-organic framework containing alternate π-donor/acceptor stacks. Angew. Chem. Int. Ed. 62, e202303819 (2023).
Cao, Y. et al. Metal-organic frameworks as highly efficient electrodes for long cycling stability supercapacitors. Int. J. Hydrog. Energy 46, 18179–18206 (2021).
Howarth, A. J. et al. Chemical, thermal and mechanical stabilities of metal–organic frameworks. Nat. Rev. Mater. 1, 15018 (2016).
Lv, X. L. et al. A base-resistant metalloporphyrin metal-organic framework for C-H bond halogenation. J. Am. Chem. Soc. 139, 211–217 (2017).
Wang, K. et al. Pyrazolate-based porphyrinic metal-organic framework with extraordinary base-resistance. J. Am. Chem. Soc. 138, 914–919 (2016).
Geng, J. et al. Reversible metal and ligand redox chemistry in two-dimensional iron-organic framework for sustainable lithium-ion batteries. J. Am. Chem. Soc. 145, 1564–1571 (2023).
Song, D. et al. Coordinative reduction of metal nodes enhances the hydrolytic stability of a paddlewheel metal–organic framework. J. Am. Chem. Soc. 141, 7853–7864 (2019).
Huang, G. et al. A comparison of two isoreticular metal–organic frameworks with cationic and neutral skeletons: stability, mechanism, and catalytic activity. Angew. Chem. Int. Ed. 59, 4385–4390 (2020).
Park, K. S. et al. Exceptional chemical and thermal stability of zeolitic imidazolate frameworks. Proc. Natl Acad. Sci. 103, 10186–10191 (2006).
Yuan, S. et al. Stable metal–organic frameworks: design, synthesis, and applications. Adv. Mater. 30, 1704303 (2018).
Chen, C.-L. et al. Conductive lanthanide metal–organic frameworks with exceptionally high stability. J. Am. Chem. Soc. 145, 16983–16987 (2023).
McHugh, L. N. et al. Hydrolytic stability in hemilabile metal-organic frameworks. Nat. Chem. 10, 1096–1102 (2018).
Redfern, L. R. et al. Isolating the role of the node-linker bond in the compression of UiO-66 metal–organic frameworks. Chem. Mater. 32, 5864–5871 (2020).
Wu, H., Yildirim, T. & Zhou, W. Exceptional mechanical stability of highly porous zirconium metal–organic framework UiO-66 and its important implications. J. Phys. Chem. Lett. 4, 925–930 (2013).
Hobday, C. L. et al. A computational and experimental approach linking disorder, high-pressure behavior, and mechanical properties in UiO frameworks. Angew. Chem. Int. Ed. 55, 2401–2405 (2016).
Zeng, L. et al. Molecular dynamics simulations of electrochemical interfaces. J. Chem. Phys. 159, 091001 (2023).
Kumar, A. et al. Theories and models of supercapacitors with recent advancements: impact and interpretations. Nano Express 2, 022004 (2021).
Yan, J., Wang, Q., Wei, T. & Fan, Z. Recent advances in design and fabrication of electrochemical supercapacitors with high energy densities. Adv. Energy Mater. 4, 1300816 (2014).
Lukatskaya, M. R. et al. Understanding the mechanism of high capacitance in nickel hexaaminobenzene-based conductive metal-organic frameworks in aqueous electrolytes. ACS Nano 14, 15919–15925 (2020).
Liu, S., Wei, L. & Wang, H. Review on reliability of supercapacitors in energy storage applications. Appl. Energy 278, 115436 (2020).
Zhang, P. et al. Dual-redox-sites enable two-dimensional conjugated metal-organic frameworks with large pseudocapacitance and wide potential window. J. Am. Chem. Soc. 143, 10168–10176 (2021).
Zhang, P. et al. Largely pseudocapacitive two-dimensional conjugated metal-organic framework anodes with lowest unoccupied molecular orbital localized in nickel-bis(dithiolene) linkages. J. Am. Chem. Soc. 145, 6247–6256 (2023).
Cheng, S. et al. Selective center charge density enables conductive 2D metal-organic frameworks with exceptionally high pseudocapacitance and energy density for energy storage devices. Adv. Mater. 34, 2109870 (2022).
Mancuso, J. L., Mroz, A. M., Le, K. N. & Hendon, C. H. Electronic structure modeling of metal–organic frameworks. Chem. Rev. 120, 8641–8715 (2020).
Nasalevich, M. A., van der Veen, M., Kapteijn, F. & Gascon, J. Metal–organic frameworks as heterogeneous photocatalysts: advantages and challenges. CrystEngComm 16, 4919–4926 (2014).
Park, J. et al. Stabilization of hexaaminobenzene in a 2D conductive metal-organic framework for high power sodium storage. J. Am. Chem. Soc. 140, 10315–10323 (2018).
Wu, Z. et al. Highly conductive two-dimensional metal-organic frameworks for resilient lithium storage with superb rate capability. ACS Nano 14, 12016–12026 (2020).
Nam, K. W. et al. Conductive 2D metal-organic framework for high-performance cathodes in aqueous rechargeable zinc batteries. Nat. Commun. 10, 4948 (2019).
Kon, K. et al. Electron-conductive metal-organic framework, Fe(dhbq)(dhbq = 2,5-dihydroxy-1,4-benzoquinone): coexistence of microporosity and solid-state redox activity. ACS Appl Mater. Interfaces 13, 38188–38193 (2021).
Cai, T., Hu, Z., Gao, Y., Li, G. & Song, Z. A rationally designed iron–dihydroxybenzoquinone metal–organic framework as practical cathode material for rechargeable batteries. Energy Storage Mater. 50, 426–434 (2022).
Kim, T. et al. Radially oriented Ni3(HITP)2 microspheres as high-performance anode materials for Li-ion capacitors with exceptional energy density and cycling stability. J. Power Sources 603, 234449 (2024).
Shuang, W. et al. Engineering the modulation of the active sites and pores of pristine metal–organic frameworks for high-performance sodium-ion storage. Inorg. Chem. Front. 10, 396–405 (2023).
Jiang, Q. et al. A redox-active 2D metal–organic framework for efficient lithium storage with extraordinary high capacity. Angew. Chem. Int. Ed. 59, 5273–5277 (2020).
Guo, L. et al. Bottom-up fabrication of 1D Cu-based conductive metal–organic framework nanowires as a high-rate anode towards efficient lithium storage. ChemSusChem 12, 5051–5058 (2019).
Wang, Z. et al. Ultrathin two-dimensional conjugated metal-organic framework single-crystalline nanosheets enabled by surfactant-assisted synthesis. Chem. Sci. 11, 7665–7671 (2020).
Yan, J. et al. Immobilizing redox-active tricycloquinazoline into a 2D conductive metal–organic framework for lithium storage. Angew. Chem. Int. Ed. 60, 24467–24472 (2021).
Yin, J.-C. et al. Triazacoronene-based 2D conductive metal–organic framework for high-capacity lithium storage. Adv. Funct. Mater. 34, 2403656 (2024).
Yin, J. et al. Stabilizing redox-active hexaazatriphenylene in a 2D conductive metal–organic framework for improved lithium storage performance. Adv. Funct. Mater. 33, 2211950 (2023).
Qi, M. et al. A rhombic 2D conjugated metal–organic framework as cathode for high-performance sodium-ion battery. Adv. Mater. 36, 2401878 (2024).
Liu, J. et al. 2D conductive metal–organic framework with anthraquinone built-in active sites as cathode for aqueous zinc ion battery. Adv. Funct. Mater. 34, 2312636 (2024).
Cheng, L. et al. Conjugation and topology engineering of 2D π-d conjugated metal-organic frameworks for robust potassium organic batteries. Angew. Chem. Int. Ed. 63, e202405239 (2024).
Chen, Y. et al. Successive storage of cations and anions by ligands of π–d-conjugated coordination polymers enabling robust sodium-ion batteries. Angew. Chem. Int. Ed. 60, 18769–18776 (2021).
Chen, D. et al. A tricycloquinazoline based 2D conjugated metal–organic framework for robust sodium-ion batteries with co-storage of both cations and anions. Chem. Sci. 15, 11564–11571 (2024).
Schmidt, J., Marques, M. R. G., Botti, S. & Marques, M. A. L. Recent advances and applications of machine learning in solid-state materials science. npj Computational Mater. 5, 83 (2019).
Sutton, R. S. & Barto, A. G. Reinforcement Learning (The MIT Press, Cambridge, MA, 2018).
Wei, J. et al. Machine learning in materials science. InfoMat 1, 338–358 (2019).
Cortes, C. & Vapnik, V. Support-vector networks. Mach. Learn. 20, 273–297 (1995).
Burges, C. J. C. A. tutorial on support vector machines for pattern recognition. Data Min. Knowl. Discov. 2, 121–167 (1998).
Cover, T. & Hart, P. Nearest neighbor pattern classification. IEEE Trans. Inf. Theory 13, 21–27 (1967).
Quinlan, J. R. Induction of decision trees. Mach. Learn. 1, 81–106 (1986).
Jain, A. K., Jianchang, M. & Mohiuddin, K. M. Artificial neural networks: a tutorial. Computer 29, 31–44 (1996).
Vaswani, A. et al. Attention is all you need. In Advances in Neural Information Processing Systems (NIPS), vol. 30 (2017).
Lundberg, S. M. & Lee, S.-I. A unified approach to interpreting model predictions. In Advances in Neural Information Processing Systems (NIPS), vol. 30 (2017).
Jain, A. et al. Commentary: the materials project: a materials genome approach to accelerating materials innovation. APL Mater. 1, 011002 (2013).
Calderon, C. E. et al. The AFLOW standard for high-throughput materials science calculations. Computational Mater. Sci. 108, 233–238 (2015).
Kirklin, S. et al. The open quantum materials database (OQMD): assessing the accuracy of DFT formation energies. npj Computational Mater. 1, 15010 (2015).
Talirz, L. et al. Materials cloud, a platform for open computational science. Sci. Data 7, 299 (2020).
Scheidgen, M. et al. NOMAD: a distributed web-based platform for managing materials science research data. J. Open Source Softw. 8, 5388 (2023).
Moghadam, P. Z. et al. Development of a Cambridge structural database subset: a collection of metal–organic frameworks for past, present, and future. Chem. Mater. 29, 2618–2625 (2017).
Groom, C. R., Bruno, I. J., Lightfoot, M. P. & Ward, S. C. The Cambridge structural database. Struct. Sci. 72, 171–179 (2016).
Li, A. et al. The launch of a freely accessible MOF CIF collection from the CSD. Matter 4, 1105–1106 (2021).
Chung, Y. G. et al. Computation-ready, experimental metal–organic frameworks: a tool to enable high-throughput screening of nanoporous crystals. Chem. Mater. 26, 6185–6192 (2014).
Chung, Y. G. et al. Advances, updates, and analytics for the computation-ready, experimental metal–organic framework database: CoRE MOF. J. Chem. Eng. Data 2019 64, 5985–5998 (2019).
Rosen, A. S. et al. Machine learning the quantum-chemical properties of metal–organic frameworks for accelerated materials discovery. Matter 4, 1578–1597 (2021).
Boyd, P. G. et al. Data-driven design of metal–organic frameworks for wet flue gas CO2 capture. Nature 576, 253–256 (2019).
Gómez-Gualdrón, D. A. et al. Evaluating topologically diverse metal–organic frameworks for cryo-adsorbed hydrogen storage. Energy Environ. Sci. 9, 3279–3289 (2016).
Colón, Y. J., Gómez-Gualdrón, D. A. & Snurr, R. Q. Topologically guided, automated construction of metal–organic frameworks and their evaluation for energy-related applications. Cryst. Growth Des. 17, 5801–5810 (2017).
Boyd, P. G. & Woo, T. K. A generalized method for constructing hypothetical nanoporous materials of any net topology from graph theory. CrystEngComm 18, 3777–3792 (2016).
Lee, S. et al. Computational screening of trillions of metal–organic frameworks for high-performance methane storage. ACS Appl. Mater. Interfaces 13, 23647–23654 (2021).
Moosavi, S. M., Boyd, P. G., Sarkisov, L. & Smit, B. Improving the mechanical stability of metal–organic frameworks using chemical caryatids. ACS Cent. Sci. 4, 832–839 (2018).
Majumdar, S., Moosavi, S. M., Jablonka, K. M., Ongari, D. & Smit, B. Diversifying databases of metal organic frameworks for high-throughput computational screening. ACS Appl. Mater. Interfaces 13, 61004–61014 (2021).
Nandy, A. et al. A database of ultrastable MOFs reassembled from stable fragments with machine learning models. Matter 6, 1585–1603 (2023).
Zhang, Z. et al. In silico high-throughput design and prediction of structural and electronic properties of low-dimensional metal–organic frameworks. ACS Appl. Mater. Interfaces 15, 9494–9507 (2023).
Burner, J. et al. ARC–MOF: a diverse database of metal-organic frameworks with DFT-derived partial atomic charges and descriptors for machine learning. Chem. Mater. 35, 900–916 (2023).
Nandy, A., Duan, C. & Kulik, H. J. Using machine learning and data mining to leverage community knowledge for the engineering of stable metal–organic frameworks. J. Am. Chem. Soc. 143, 17535–17547 (2021).
Nandy, A. et al. MOFsimplify, machine learning models with extracted stability data of three thousand metal-organic frameworks. Sci. Data 9, 74 (2022).
Iacomi, P. & Llewellyn, P. L. Data mining for binary separation materials in published adsorption isotherms. Chem. Mater. 32, 982–991 (2020).
Siderius, D., Shen, V., Johnson III, R., van Zee, R., (eds.). NIST/ARPA-E database of novel and emerging adsorbent materials. (National Institute of Standard and Technology, Gaithersburg, 2016).
Luo, Y. et al. MOF synthesis prediction enabled by automatic data mining and machine learning. Angew. Chem. Int. Ed. 61, e202200242 (2022).
Ramakrishna, S. et al. Materials informatics. J. Intell. Manuf. 30, 2307–2326 (2019).
Liu, Y., Zhao, T., Ju, W. & Shi, S. Materials discovery and design using machine learning. J. Materiomics 3, 159–177 (2017).
Moosavi, S. M. et al. A data-science approach to predict the heat capacity of nanoporous materials. Nat. Mater. 21, 1419–1425 (2022).
Janet, J. P. & Kulik, H. J. Resolving transition metal chemical space: feature selection for machine learning and structure–property relationships. J. Phys. Chem. A 121, 8939–8954 (2017).
Moosavi, S. M. et al. Understanding the diversity of the metal-organic framework ecosystem. Nat. Commun. 11, 4068 (2020).
Schütt, K. T. et al. How to represent crystal structures for machine learning: towards fast prediction of electronic properties. Phys. Rev. B 89, 205118 (2014).
Faber, F., Lindmaa, A., von Lilienfeld, O. A. & Armiento, R. Crystal structure representations for machine learning models of formation energies. Int. J. Quantum Chem. 115, 1094–1101 (2015).
Seko, A., Hayashi, H., Nakayama, K., Takahashi, A. & Tanaka, I. Representation of compounds for machine-learning prediction of physical properties. Phys. Rev. B 95, 144110 (2017).
Isayev, O. et al. Universal fragment descriptors for predicting properties of inorganic crystals. Nat. Commun. 8, 15679 (2017).
Xie, T. & Grossman, J. C. Crystal graph convolutional neural networks for an accurate and interpretable prediction of material properties. Phys. Rev. Lett. 120, 145301 (2018).
Chen, C., Ye, W., Zuo, Y., Zheng, C. & Ong, S. P. Graph networks as a universal machine learning framework for molecules and crystals. Chem. Mater. 31, 3564–3572 (2019).
Cao, Z., Magar, R., Wang, Y. & Barati Farimani, A. MOFormer: self-supervised transformer model for metal-organic framework property prediction. J. Am. Chem. Soc. 145, 2958–2967 (2023).
Kang, Y., Park, H., Smit, B. & Kim, J. A multi-modal pre-training transformer for universal transfer learning in metal–organic frameworks. Nat. Mach. Intell. 5, 309–318 (2023).
Kang, Y. & Kim, J. ChatMOF: an artificial intelligence system for predicting and generating metal-organic frameworks using large language models. Nat. Commun. 15, 4705 (2024).
He, Y., Cubuk, E. D., Allendorf, M. D. & Reed, E. J. Metallic metal-organic frameworks predicted by the combination of machine learning methods and ab initio calculations. J. Phys. Chem. Lett. 9, 4562–4569 (2018).
Lin, J. et al. Machine learning-driven discovery and structure–activity relationship analysis of conductive metal–organic frameworks. Chem. Mater. 36, 5436–5445 (2024).
Moghadam, P. Z. et al. Structure-mechanical stability relations of metal-organic frameworks via machine learning. Matter 1, 219–234 (2019).
Batra, R., Chen, C., Evans, T. G., Walton, K. S. & Ramprasad, R. Prediction of water stability of metal–organic frameworks using machine learning. Nat. Mach. Intell. 2, 704–710 (2020).
Zhang, Z. et al. Accelerating discovery of water stable metal-organic frameworks by machine learning. Small 20, e2405087 (2024).
Fan, K. et al. Single crystals of a highly conductive three-dimensional conjugated coordination polymer. J. Am. Chem. Soc. 145, 12682–12690 (2023).
Nazir, A., Le, H. T. T., Nguyen, A.-G. & Park, C.-J. Graphene analogue metal organic framework with superior capacity and rate capability as an anode for lithium ion batteries. Electrochim. Acta 389, 138750 (2021).
Karniadakis, G. E. et al. Physics-informed machine learning. Nat. Rev. Phys. 3, 422–440 (2021).
von Rueden, L. et al. Informed machine learning - a taxonomy and survey of integrating prior knowledge into learning systems. IEEE Trans. Knowl. Data Eng. 35, 614–633 (2021).
Karpatne, A. et al. Theory-guided data science: a new paradigm for scientific discovery from data. IEEE Trans. Knowl. Data Eng. 29, 2318–2331 (2017).
Acknowledgements
Acknowledge the funding support from the National Natural Science Foundation of China (T2325012, 92472109), the Program for HUST Academic Frontier Youth Team. Liang Zeng is supported by the Postdoctoral Fellowship Program of CPSF under Grant Number GZC20240532.
Author information
Authors and Affiliations
Contributions
G.F. led the collaboration. T.S. collected publications and completed the framework of the manuscript. Z.-X.W., L.Z. and G.F. thoroughly revised the paper. All authors reviewed and edited the manuscript.
Corresponding author
Ethics declarations
Competing interests
The authors declare no competing interests.
Additional information
Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Rights and permissions
Open Access This article is licensed under a Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License, which permits any non-commercial use, sharing, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if you modified the licensed material. You do not have permission under this licence to share adapted material derived from this article or parts of it. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by-nc-nd/4.0/.
About this article
Cite this article
Sun, T., Wang, Z., Zeng, L. et al. Identifying MOFs for electrochemical energy storage via density functional theory and machine learning. npj Comput Mater 11, 90 (2025). https://doi.org/10.1038/s41524-025-01590-w
Received:
Accepted:
Published:
DOI: https://doi.org/10.1038/s41524-025-01590-w