Conductive Metal-Organic Frameworks for Electrocatalysis ...

14
Acta Phys. -Chim. Sin. 2021, 37 (7), 2010025 (1 of 14) Received: October 13, 2020; Revised: November 25, 2020; Accepted: November 25, 2020; Published online: November 30, 2020. * Corresponding authors. Emails: [email protected] (Z.Z.); [email protected] (W.H.); Tel: +86-22-83613363 (Z.Z.). These authors contributed equally to this work. This work was supported by the National Key R&D Program (2017YFA0204503), and the National Natural Science Foundation of China (22071172, 91833306, 21875158, 51633006, 51733004). 国家重点研发计划(2017YFA0204503)和国家自然科学基金(22071172, 91833306, 21875158, 51633006, 51733004)资助 © Editorial office of Acta Physico-Chimica Sinica [Review] doi: 10.3866/PKU.WHXB202010025 www.whxb.pku.edu.cn Conductive Metal-Organic Frameworks for ElectrocatalysisAchievements, Challenges, and Opportunities Zengqiang Gao 1,, Congyong Wang 2,3,† , Junjun Li 1 , Yating Zhu 1 , Zhicheng Zhang 1,* , Wenping Hu 1,2,* 1 Tianjin Key Laboratory of Molecular Optoelectronic Sciences, Department of Chemistry, School of Science, Tianjin University & Collaborative Innovation Center of Chemical Science and Engineering, Tianjin 300072, China. 2 Joint School of National University of Singapore and Tianjin University, International Campus of Tianjin University, Binhai New City, Fuzhou 350207, China. 3 Department of Chemistry, Faculty of Science, National University of Singapore, Singapore 117543, Singapore. Abstract: To fulfill the demands of green and sustainable energy, the production of novel catalysts for different energy conversion processes is critical. Owing to the intriguing advantages of the intrinsic active species, tunable crystal structure, remarkable chemical and physical properties, and good stability, metal-organic frameworks (MOFs) have been extensively investigated in various electrochemical energy conversions, such as the CO2 reduction reaction, N2 reduction reaction, oxygen evolution reaction, hydrogen evolution reaction, and oxygen reduction reaction. More importantly, it is feasible to change the chemical environments, pore sizes, and porosity of MOFs, which will theoretically facilitate the diffusion of reactants across the open porous networks, thereby improving the electrocatalytic performance. However, owing to the high energy barriers of charge transfer and limited free charge carriers, most MOFs show poor electrical conductivity, thus limiting their diverse applications. As reported previously, MOFs were used as a porous substrate to confine the growth of nanoparticles or co-doped electrocatalysts after annealing. The conductive MOFs can combine the advantages of conventional MOFs with electronic conductivity, which significantly enhance the electrocatalytic performance. In addition, conductive MOFs can achieve conductivity via electronic or ionic routes without post-annealing treatment, thereby extending their potential applications. Different synthesis strategies have recently been developed to endow MOFs with electrical conductivity, such as post-synthesis modification, guest molecule introduction, and composite formatting. The performance of conductive MOFs can even outperform those of commercial RuO2 catalysts or Pt-group catalysts. However, it is difficult to endow most MOFs with high conductivity. This review summarizes the mechanisms of constructing conductive MOFs, such as redox hopping, through-bond pathways, through-space pathways, extended conjugation, and guest-promoted transport. Synthetic methods, including hydro/solvothermal synthesis and interface-assisted synthesis, are introduced. Recent advances in the use of conductive MOFs as heterogeneous catalysts in electrocatalysis have been comprehensively elucidated. It has been reported that conductive MOFs can demonstrate considerable catalytic activity, selectivity, and stability in different electrochemical reactions, revealing the immense potential for future displacement of Pt-group catalysts. Finally, the challenges and opportunities of conductive MOFs in electrocatalysis are discussed. Based on systematic synthesis strategies, more conductive MOFs can be constructed for electrocatalytic reactions. In addition, the morphology and structure of conductive MOFs, which can change the electrochemical accessibility between substrates and MOFs, are also crucial for catalysis, and thus, they should be extensively studied in the future. It is believed that a breakthrough for high-performance conductive MOF-based electrocatalysts could be achieved.

Transcript of Conductive Metal-Organic Frameworks for Electrocatalysis ...

Page 1: Conductive Metal-Organic Frameworks for Electrocatalysis ...

物 理 化 学 学 报

Acta Phys. -Chim. Sin. 2021, 37 (7), 2010025 (1 of 14)

Received: October 13, 2020; Revised: November 25, 2020; Accepted: November 25, 2020; Published online: November 30, 2020. *Corresponding authors. Emails: [email protected] (Z.Z.); [email protected] (W.H.); Tel: +86-22-83613363 (Z.Z.). †These authors contributed equally to this work.

This work was supported by the National Key R&D Program (2017YFA0204503), and the National Natural Science Foundation of China (22071172,

91833306, 21875158, 51633006, 51733004).

国家重点研发计划(2017YFA0204503)和国家自然科学基金(22071172, 91833306, 21875158, 51633006, 51733004)资助

© Editorial office of Acta Physico-Chimica Sinica

[Review] doi: 10.3866/PKU.WHXB202010025 www.whxb.pku.edu.cn

Conductive Metal-Organic Frameworks for Electrocatalysis:Achievements, Challenges, and Opportunities

Zengqiang Gao 1,†, Congyong Wang 2,3,†, Junjun Li 1, Yating Zhu 1, Zhicheng Zhang 1,*, Wenping Hu 1,2,* 1 Tianjin Key Laboratory of Molecular Optoelectronic Sciences, Department of Chemistry, School of Science, Tianjin University &

Collaborative Innovation Center of Chemical Science and Engineering, Tianjin 300072, China. 2 Joint School of National University of Singapore and Tianjin University, International Campus of Tianjin University, Binhai New

City, Fuzhou 350207, China. 3 Department of Chemistry, Faculty of Science, National University of Singapore, Singapore 117543, Singapore.

Abstract: To fulfill the demands of green and sustainable energy, the production

of novel catalysts for different energy conversion processes is critical. Owing to the

intriguing advantages of the intrinsic active species, tunable crystal structure,

remarkable chemical and physical properties, and good stability, metal-organic

frameworks (MOFs) have been extensively investigated in various electrochemical

energy conversions, such as the CO2 reduction reaction, N2 reduction reaction,

oxygen evolution reaction, hydrogen evolution reaction, and oxygen reduction

reaction. More importantly, it is feasible to change the chemical environments, pore

sizes, and porosity of MOFs, which will theoretically facilitate the diffusion of

reactants across the open porous networks, thereby improving the electrocatalytic

performance. However, owing to the high energy barriers of charge transfer and

limited free charge carriers, most MOFs show poor electrical conductivity, thus

limiting their diverse applications. As reported previously, MOFs were used as a porous substrate to confine the growth of

nanoparticles or co-doped electrocatalysts after annealing. The conductive MOFs can combine the advantages of

conventional MOFs with electronic conductivity, which significantly enhance the electrocatalytic performance. In addition,

conductive MOFs can achieve conductivity via electronic or ionic routes without post-annealing treatment, thereby

extending their potential applications. Different synthesis strategies have recently been developed to endow MOFs with

electrical conductivity, such as post-synthesis modification, guest molecule introduction, and composite formatting. The

performance of conductive MOFs can even outperform those of commercial RuO2 catalysts or Pt-group catalysts. However,

it is difficult to endow most MOFs with high conductivity. This review summarizes the mechanisms of constructing

conductive MOFs, such as redox hopping, through-bond pathways, through-space pathways, extended conjugation, and

guest-promoted transport. Synthetic methods, including hydro/solvothermal synthesis and interface-assisted synthesis,

are introduced. Recent advances in the use of conductive MOFs as heterogeneous catalysts in electrocatalysis have been

comprehensively elucidated. It has been reported that conductive MOFs can demonstrate considerable catalytic activity,

selectivity, and stability in different electrochemical reactions, revealing the immense potential for future displacement of

Pt-group catalysts. Finally, the challenges and opportunities of conductive MOFs in electrocatalysis are discussed. Based

on systematic synthesis strategies, more conductive MOFs can be constructed for electrocatalytic reactions. In addition,

the morphology and structure of conductive MOFs, which can change the electrochemical accessibility between substrates

and MOFs, are also crucial for catalysis, and thus, they should be extensively studied in the future. It is believed that a

breakthrough for high-performance conductive MOF-based electrocatalysts could be achieved.

Page 2: Conductive Metal-Organic Frameworks for Electrocatalysis ...

物理化学学报 Acta Phys. -Chim. Sin. 2021, 37 (7), 2010025 (2 of 14)

Key Words: Conductive metal-organic frameworks; Electrocatalysis; CO2 reduction reaction; N2 reduction

reaction; Oxygen evolution reaction; Hydrogen evolution reaction; Oxygen reduction reaction

导电金属有机框架材料在电催化中的成就,挑战和机遇

高增强 1,†,王聪勇 2,3,†,李俊俊 1,朱亚廷 1,张志成 1,*,胡文平 1,2,* 1 天津市分子光电科学重点实验室,天津大学理学院化学系,天津化学化工协同创新中心,天津 300072 2 天津大学-新加坡国立大学福州联合学院,天津大学福州国际校区,滨海新城,福州 350207 3 化学系,理学院,新加坡国立大学,新加坡 117543

摘要:开发用于各种能量转化过程的新型催化剂对于满足绿色和可持续能源的需求至关重要。由于其具有可调节的晶体

结构,显著的化学和物理性质以及稳定性,金属有机骨架(MOFs)已经广泛应用于电化学能量转换领域,比如CO2还原反

应、N2还原反应、析氧反应、析氢反应和氧还原反应。更重要的是,MOFs具有可调节的化学环境、孔径和孔隙率,这

些性质将促进反应物在多孔网络中的扩散,从而改善其电催化性能。但是,由于高的电荷转移能垒和受限的自由载流子,大

多数MOFs展示了差的导电性,阻碍了其多样化应用。在先前的报道中,MOFs常被用作多孔基质来限制纳米颗粒生长或

经退火处理作为共掺杂电催化剂。而导电MOFs不仅结合了传统MOFs的优点,还具有电子导电性和高电催化活性,使其

无需退火处理就可以通过电子或离子途径实现导电,从而极大提高了电催化性能,这有助于拓宽其在电化学能源领域或

其他方面的潜在应用。在一些催化反应中,导电MOFs的催化活性甚至超过了商业化的RuO2催化剂或Pt基催化剂。本文

主要总结了构建导电MOFs的机制,并概述了其合成方法,如水/溶剂热合成和界面辅助合成。此外,本文阐述了导电MOFs

在电催化应用中的最新研究进展。值得一提的是,导电MOFs的形态和结构可改变底物与MOFs之间的界面接触,从而影

响其催化性能,需要进一步深入研究。基于系统的合成策略,在未来可以根据各种电催化反应的需求设计合成更多的导

电MOFs。高性能的导电MOF基催化剂将有望获得突破。

关键词:导电金属有机框架;电催化;二氧化碳还原反应;氮还原反应;析氧反应;析氢反应;氧还原反应

中图分类号:O643

1 Introduction Metal-organic frameworks (MOFs), as hybrid crystalline

reticular materials composed of organic linkers and metal-

containing nodes, have drawn tremendous attention from

researchers because of their captivating and unique advantages

like outstanding surface areas, tunable porosity, and framework

topologies 1–5. The combination of these compelling features

with well-defined ordered structures makes MOFs promising

candidates for a wide variety of applications, such as molecular

separations, energy storage and conversion, sensors, magnets,

catalysis, biomedicine, and so on 6–17. However, the electronic

properties and applications of MOFs have attracted relatively

less interest as the majority of MOFs are insulators with

inherently weak conductivity (< 10−10 S·cm−1) 18,19. With the

emergence of novel designing and synthetic strategies,

considerable research interests have been focused on fabricating

conductive MOFs and broadening their electrical applications 20–24.

Until recent years, numerous two dimensional (2D) and three

dimensional (3D) MOFs have been reported to be electrically

conductive 25, showing great potential in diverse electrical-

Zhicheng Zhang is a Professor in the Department of

Chemistry, School of Science, Tianjin University. He

received his Ph.D. degree from the College of

Chemical Engineering, China University of

Petroleum (Beijing) in 2012. Then, he worked as a

Postdoctoral Researcher in the Department of

Chemistry, Tsinghua University. Since 2014, he

worked as a Research Fellow in the School of Materials Science and

Engineering, Nanyang Technological University, Singapore. In 2019, he

joined Tianjin University as a Full Professor. His research interests include the

design and synthesis of functional nanomaterials and their applications in

energy conversion, catalysis, and organic optoelectronics.

Wenping Hu is a Professor at Tianjin University and

a Cheung Kong Professor of the Ministry of

Education, China. He received his Ph.D. from

ICCAS in 1999. Then, he joined Osaka University

and Stuttgart University as a Research Fellow of the

Japan Society for the Promotion of Sciences and an

Alexander von Humboldt fellow, respectively. In

2003, he worked with Nippon Telephone and Telegraph and then joined

ICCAS as a Full Professor. He worked for Tianjin University in 2013. His

research focuses on organic optoelectronics.

Page 3: Conductive Metal-Organic Frameworks for Electrocatalysis ...

物理化学学报 Acta Phys. -Chim. Sin. 2021, 37 (7), 2010025 (3 of 14)

related applications, such as electrocatalysis 26–30, batteries 31,32,

supercapacitors 33–35, chemiresistive sensors 36–39, field-effect

transistors (FET) 40–43, organic spin valves 44, and other

electronics or spintronics 45–47. Among these applications, the

utilization of conductive MOFs as electrocatalysts for

electrochemical energy conversion has received special attention

due to their favorable characteristics: (i) good electrical

conductivity is beneficial for rapid charge carrier transfer; (ii)

high specific surface areas and greatly exposed active sites are

capable of enhancing the catalytic efficiency; (iii) well-defined

structure and highly ordered porosity can facilitate the

incorporation of metal sites and additional catalytic sites acting

as hosts; (iv) tunable size, morphology, and structure with

accessibly abundant active sites. These compelling features

make the conductive MOFs extremely attractive for

electrocatalytic reactions, including CO2 reduction reaction

(CO2RR), oxygen evolution reaction (OER), hydrogen evolution

reaction (HER), N2 reduction reaction (NRR), and O2 reduction

reaction (ORR).

Electrocatalysis has been perceived as one of the cleanest and

the most renewable strategies to mitigate environmental issues

and energy shortage problems 48–50. More recently, conductive

MOFs have been widely investigated as catalysts for

electrocatalytic application 10,51. However, a focused review of

the synthesis and electrocatalytic application of conductive

MOFs is still lacking. Herein, we aim to summarize the recent

progress of conductive MOFs with a particular emphasis on their

synthetic strategies and electrocatalytic applications. Firstly, two

kinds of synthetic approaches, including hydro/solvothermal

synthesis and interface-assisted synthesis, are summarized to

fabricate conductive MOFs. Secondly, the applications of

conductive MOFs as electrocatalysts toward catalyzing energy-

conversion reactions (CO2RR, HER, OER, NRR, and ORR) are

elaborated. Finally, the existing problems and perspectives are

highlighted, promoting their further development in the field of

electrocatalysis. It is envisioned that this review can offer

guidance and inspiration to design and develop highly efficient

conductive MOFs for electrocatalytic applications.

Conductivity can quantify the efficacy of charge transfer,

involving carrier mobility and the carrier concentration. And

band-like and hopping transport are the general mechanisms, the

former relies on strong interactions and can form continuous

energy bands. For the latter, due to the discretely localized

charge carriers, there are high barriers between the networks.

Considering that the complex factors in the transfer mechanisms,

five broad approaches are proposed to achieve the conductivity

for MOFs 52. (i) Continuous coordination bonds within MOFs

can form a system of delocalization, contributing to high charge

mobilities and small band gaps (Fig. 1a). (ii) The organic and

inorganic components endow a part of MOFs with metal-organic

analogs of graphene, which can transport charge by extended

conjugation (Fig. 1b). (iii) MOFs with π–π interactions of

organic linkers can transfer charges through space (Fig. 1c). (iv)

MOFs with obfuscated grain boundary resistances and unknown

defect concentrations are hard to distinguish the transfer

mechanism. Thus, the small spatial separation and the redox-

active components can improve charge transfer (Fig. 1d). (v) The

presence of the porosity can post-synthetically introduce

electroactive guests into MOFs to promote the conductivity (Fig.

1e).

2 Synthetic strategies of conductive MOFs

2.1 Hydro/solvothermal synthesis

The hydro/solvothermal synthesis is a conventional method

that has been extensively employed to fabricate 2D or 3D MOFs

with different morphologies (powders, single crystals, or films),

especially for 2D conjugated conductive MOFs 46. In this

method, the reactants (monomers, solvents, and surfactant

additives) are added into a sealed autoclave, where some factors,

including the concentration of monomers, the types of solvents,

reaction temperature, and time, have a great influence on the final

products. Thanks to its simple operation, the hydro/solvothermal

synthesis method is capable of preparing conductive MOFs in a

high-yield, large scale, and low-cost way 20.

For example, Dincă et al. applied 2,3,6,7,10,11-

hexaaminotriphenylene (HITP) to react with Ni2+ for forming

2D conductive MOFs of Ni3(HITP)2 under solvothermal

conditions 53. The resultant Ni3(HITP)2 bulk powders displayed

a conductivity of only 2 S·cm−1, which could be attributed to the

electrode contact resistances and the presence of grain

boundaries. By utilizing different metal centers, the same group

prepared another electrically conductive MOF of Cu3(HITP)2

with a conductivity of 0.2 S·cm−1 using a hydrothermal

approach 36. Powder X-ray diffraction (PXRD) measurement

indicated that the as-obtained Cu3(HITP)2 exhibited a 2D lattice

with slipped-parallel stackings and hexagonal unit cells (a = b =

2.23 nm and c = 0.66 nm). In addition to conductive MOFs based

on the organic linkers of HITP, the tetrathiafulvalene (TTF)-

based MOFs have also been widely prepared by the

hydrothermal method 54–56. For instance, three different TTF-

based conductive MOFs with various topologies, including

La4(HTTFTB)4, La(HTTFTB), and La4(TTFTB)3, have been

constructed from the same TTFTB ligand and La3+ under distinct

solvothermal conditions 56. It was found that their electrical

Fig. 1 (a) The through-bond pathway; (b) The extended conjugation

pathway; (c) The through-space pathway; (d) A redox hopping

mechanism; (e) A guest-promoted pathway.

(a–e) Adapted from ACS Publications publisher 52.

Page 4: Conductive Metal-Organic Frameworks for Electrocatalysis ...

物理化学学报 Acta Phys. -Chim. Sin. 2021, 37 (7), 2010025 (4 of 14)

conductivity was related to the stacking motifs, and the longer

S…S contact distances would lead to the reduction of electrical

conductivity. Additionally, the hydrothermal synthesis method

has also been developed to synthesize conductive MOFs single

crystals with exceptional electrical conductivities 57,58. Recently,

the 2D conductive MOFs (Ni3- and Cu3(HHTP)2) were obtained

as single crystals via the solvothermal method 58. Remarkably,

single rods displayed a conductivity of 150 S·cm−1 without

distinct degradation over prolonged exposure to an ambient

environment. These conductive MOF's single crystals afforded

remarkable promise as the ideal platforms for electrical-related

applications.

2.2 Interface-assisted synthesis

Although great achievements have been made regarding the

synthesis of conductive MOFs, the majority of products are

obtained as powders or single crystals which are difficult to be

integrated into different devices. In contrast, conductive MOFs

thin films provide more substantial opportunities for practical

electronic applications and attract more attention from

researchers owing to their good solution-based processability.

Until now, a large number of conductive MOFs thin films have

been prepared by interface-assisted synthesis methods including

Langmuir-Blodgett (LB) methodology, air-liquid interfacial

synthesis, and liquid-liquid interfacial synthesis.

The LB approach is proven to be effective for fabricating

monolayer film on-air/liquid surfaces, where the surface

pressure is used to promote the reaction of monomers in LB

trough and the formed monolayer film can be readily transferred

onto the suitable substrates. The utilization of LB methodology

for the preparation of the MOFs monolayer began in 2010,

Kitagawa group successfully constructed 2D MOFs of CoTCPP-

py-Cu consisting of CoTCPP and pyridine at the surface of H2O 59.

Since then, the method of LB is extensively employed for

fabricating various MOF films.

Feng et al. fabricated large-area, free-standing, and

single-layer 2D MOFs films from the π-conjugated building

block of 1,2,5,6,9,10-triphenylenehexathiol (THT) at an air-

water interface through LB methodology (Fig. 2a) 60. In detail,

the THT monomers were firstly compressed into a densely single

layer on the surface in a LB groove. With the injection of nickel

salts solution into the aqueous solution, the coordination

polymerization between the nickel and dithiolene unites

occurred, leading to the formation of single-layer 2D MOFs with

a thickness of ~0.7 nm and large lateral dimensions up to square

millimeters (Fig. 2b,c).

In 2017, Xu and co-workers adopted an air-liquid interfacial

approach to synthesize a conductive MOFs film of Ni3(HITP)2,

exhibiting a smooth surface (average surface roughness of ~1.43

nm) (Fig. 3a–c) 40. Such high-quality and flat MOF film was

formed by the constant assembly of HITP and nickel ions on the

water surface. The thickness of Ni3(HITP)2 film could be well

controlled by reaction time and a MOF film with a thickness of

approximately 100 nm was afforded upon a 3-min reaction. The

as-synthesized Ni3(HITP)2 film was further used to construct a

FET device as the active channel material, with a device

geometry of bottom-gate top-contact. The fabricated FET

displayed exceptional charge mobility of 48.6 cm2·V−1·s−1,

which was comparable with other reported organic or inorganic-

based FET. By varying metal centers, Lahiri et al. prepared a

series of conjugated thin 2D MOF films by coordinating

hexaiminobenzene (HAB) ligands with different metal ions

(Cu2+, Ni2+, and Co2+) utilizing the air-liquid interfacial synthesis

method 42. However, poor crystallinity was observed in these

MOF films, resulting in their poor charge transporting ability and

insulating property.

The liquid-liquid interfacial reaction is regarded as another

promising strategy to fabricate conductive MOFs films.

In 2013, Nishihara et al. proposed H2O/CH2Cl2 interfacial

reactions of fabricating conductive MOFs of Ni-BHT (BHT =

benzenehexathiol) composed of nickel bis(dithiolene) units 61. It

was found that the resultant Ni-BHT films exhibited good

crystallinity, a large lateral size of up to100 μm, and a thickness

of 1–2 μm. Subsequently, some other conductive MOFs,

including PdDt 62, NiAT 63, and PtDT 64, have also been prepared

by the Nishihara group using the liquid-liquid interfacial

Fig. 3 (a) Schematic illustration for fabricating Ni3(HITP)2-based

FETs; (b) SEM and (c) AFM images of Ni3(HITP)2 film.

(a–c) Adapted from ACS Publications publisher 40.

Fig. 2 (a) Schematic representation of the preparation of

2D single-layer MOFs using LB method; (b) AFM image of MOFs

single layers showing thickness of ~0.7 nm; (c) TEM image of the

single-layer MOFs on Cu grids.

(a–c) Adapted from Wiley publisher 60.

Page 5: Conductive Metal-Organic Frameworks for Electrocatalysis ...

物理化学学报 Acta Phys. -Chim. Sin. 2021, 37 (7), 2010025 (5 of 14)

approach, indicating the generality of this method in fabricating

electrically conducting MOFs films.

Additionally, Zhu and co-workers utilized the

dichloromethane (CH2Cl2)/water interface to fabricate highly

crystalline MOFs films composed of copper bis(dithiolene)

complex of Cu-BHT with an adjustable thickness (20–140 nm)

(Fig. 4a–e) 41. According to synchrotron radiation grazing

incident X-ray diffraction (GIXRD) characterizations, the as-

formed Cu-BHT films were piled up by 2D plate-like sheets of

[Cu3(C6S6)]n, which possessed a hexagonal lattice of a = b =

0.876 nm. By replacing metal centers of Cu with Ag, Zhu et al.

also prepared highly crystalline Ag-BHT film through a

coordination reaction between BHT and silver (I) nitrate at the

toluene/H2O interface 65. It was worth noting that the Ag-BHT

film displayed remarkable conductivity (250 S·cm−1), showing

great potential for electronic applications.

3 Conductive MOFs for electrocatalysis Electrocatalysis is one of the particularly promising strategies

to produce carbon-neutral fuels or industrial chemicals derived

from petroleum, due to its mild soft reaction condition,

controllable selectivity, and scalable application 66. The strategy

can efficiently decrease the emission of environmental pollutants

by coupling with a renewable energy source, such as wind or

solar energy 67. Due to the components of MOF precisely

modified at the molecular level, MOF has an intriguing potential

in electrocatalysis 68.

3.1 CO2 reduction reaction (CO2RR)

The electrochemical reduction of CO2 to chemical products is

an important route for meeting the energy demand, reducing the

dependence on fossil fuel 69–74. For electrochemical

thermodynamics such as the formal potentials, the proton-

coupled pathway, such as 4e−, 6e−, and 8e− processes, is more

favorable 75.

Inspired by electrocatalytic degradation of carbon

tetrachloride by Co-porphyrinic MOF 76, Kubiak’s group applied

electrophoretic deposition to construct a heterogeneous system

by integrating redox-conductive Fe-porphyrins into thin MOF-

525 films (Fig. 5a), which electrochemically covered catalytic

sites (~1015 sites·cm−2) in conductive surface 77. The active

Fe(0)-porphyrin species in Fe-based MOF films can produce

15.3 μmol·cm−2 of CO and 14.9 μmol·cm−2 of H2, whose current

densities can reach 2.3 mA·cm−2 for 0.5 h, and turnover number

(TON) for CO in controlled potential electrolysis (CPE) was

272. With acidic proton donor (trifluoroethanol, TFE), the Fe(0)-

porphyrin system of TON in this system attained 1520 (Fig. 5b)

and it exhibited a current density of 5.9 mA·cm−2 and catalyst

stability for 3.2 h, better than that without TFE. Whereafter,

Albo et al. further investigated the electrocatalytic reduction of

CO2 by using MOFs (HKUST-1, CuAdeAce) and metal-organic

aerogels (CuDTA, CuZnDTA) (Fig. 5c) 78. By cyclic

voltammograms (CV), HKUST-1 showed the best current

densities after five scans among four materials, revealing its

excellent performance in transferring electrons. At the best

condition, the HKUST-1 exhibited yield rates of 5.62 × 10−6 and

5.28 × 10−6 mol·m−2·s−1 for CH3OH and C2H5OH, respectively

(Fig. 5d).

The charge can be transported within MOF films by either

linker or electron/hole redox hopping. 3D porphyrin-based

MOF-525 catalyst synthesized by dip-coating method admirably

converted CO2 to CO with low overpotential, achieving a

maximum FECO of 91% and a TOF of 0.336 site−1·s−1 after 10-

hour electrolysis 79. Due to the porosity and conformance of the

inherited catalytic active sites after carbonization, MOF was

considered a promising candidate in the field of non-noble metal

catalysts. And metal-metal interface played an important role in

improving the activity and selectivity of the catalysts. Three

heterometallic HKUST-1 dopped ions (Zn, Ru, Pd) reported by

Albo et al. were also applied to investigate the catalytic

performance 80. Ru(III)-electrocatalyst showed the highest yield

of alcohol with the FE of 47.2%. Li group applied Cu3(BTC)2

MOF based on GDE to capture CO2 for synthesizing

hydrocarbons, in which capture capacity reaches up to 1.8

mmol·g−1 under 1.0 × 105 Pa 81. Under relatively negative

potentials, the FE of CH4 for Cu3(BTC)2 at a weight ratio of

7.5%–10% can double that without the addition of Cu3(BTC)2,

and for the competitive HER, the FE was declined to 30%.

Moreover, MOF with the function of support for catalytic

molecular-type catalysts for electrochemical conversion of CO2

to CO is an intriguing strategy, which is capable of tailoring

active molecular in the porous network to maximize active sites.

And charge transport could be controlled by the morphology of

MOF. Chang et al. incorporate the active molecule (TCPP-Co)

into an Al-based MOF 82. All of the active sites in homogeneous

catalysts (porphyrin molecular) can contact the electrolyte,

electrons and their transfer can be changed by adjusting the

morphology and thickness of MOF. The current selectivity for

CO was over 76% at −0.7 V versus RHE, running for over 7 h.

Fig. 4 (a) The digital photo of preparing Cu-BHT film at

CH2Cl2/H2O; (b) Upside-up face (right) and upside-down (left)

face of the as-fabricated film; (c, d) SEM images of the downside and

upside surfaces (scale bar of 200 nm in (c), scale bar of 400 nm in (d),

scale bar of 100 nm in the insets of (c) and (d)). (e) SEM image

showing the cross-section of Cu-BHT film (scale bar: 400 nm).

(a–e) Adapted from ACS Publications publisher 41.

Page 6: Conductive Metal-Organic Frameworks for Electrocatalysis ...

物理化学学报 Acta Phys. -Chim. Sin. 2021, 37 (7), 2010025 (6 of 14)

Further, a hybrid thin film (Ag@Al-PMOF) was synthesized to

improve the reduction of CO2 and suppress HER by changing

the interface between the nanoparticles and the MOFs 83. Wu et

al. investigated 2D Cu2(CuTCPP) nanosheets for selective

formate and acetate production with a FE of 68.4% and 85.2%

(Fig. 5e,f), superior to other Cu catalysts (Cu, CuTCPP, Cu2O,

and CuO) 84. More recently, Cao et al. applied in situ

electrochemical transformations to obtain atomically thin

bismuthene (Bi-ene) that can produce formate with a FE of

~100 % in a wider potential range 85. Based on the study of in

situ ATR-IR spectra, the authors hypothesized that the

absorption of HCO−3 in the surface led to the formation of

OCHO* intermediate, which gives a new insight into the CO2

electroreduction process.

3.2 Oxygen reduction reaction (ORR)

For biological and energy systems, the transformation of

Fig. 6 (a) The 2D layered structure of Ni3(HITP)2; (b) The polarization curves of different electrodes (glassy carbon electrode, GCE).

(a, b) Adapted from Springer publisher 27.

Fig. 5 (a) The structure of Fe-porphyrin MOF; (b) TON versus time for FTO without or with Fe-porphyrin MOF; (c) TON versus time for Fe-

porphyrin MOF without or with TFE; (d) Rates and the FE of HKUST-1; (e) Crystal structure and (f) the reaction system of Cu2(CuTCPP).

(a–c) Adapted from ACS Publications publisher 77. (d) Adapted from Wiley publisher 78. (e, f) Adapted from Royal Society of Chemistry publisher 84.

Page 7: Conductive Metal-Organic Frameworks for Electrocatalysis ...

物理化学学报 Acta Phys. -Chim. Sin. 2021, 37 (7), 2010025 (7 of 14)

oxygen (O2) to water (H2O) is a critical reaction, which is a

multi-electron/multi-proton pathway (4e–/4H+) 86–91.

Ni3(HITP)2 film with electrical conductivity of 40 S·cm−1

grown in the electrode surface exhibited competitive ORR

activity in alkaline solution (Fig. 6a) 27. The film with a specific

surface area of 629.9 m2·g−1 exhibited the current density of

50 μA·cm−2 at 0.82 V (Fig. 6b), which could keep 88% of the

initial current density over 8-h measurements. Liu et al. reported

that the metal-catecholate with well-defined M-O6 octahedral

coordination (M = Ni or Co) could display good electrocatalytic

performance towards ORR 92. Afterwards, bimetallic conductive

2D CoxNiy-CAT was developed via a ball-mill reactor for ORR

to replace Pt-based catalysts 93. The activity of bimetallic

CoxNiy-CATs for ORR can be controlled by adjusting the ratio of

two metal ions. Compared to the counterparts, the hybrids

exhibited better ORR activity, by combining the high diffusion-

limiting current density of Co-CAT with the high onset potential

of Ni-CAT. Chen et al. prepared a series of porphyrinic MOFs

((Co)PCN222) with size from 200 to 1000 nm via a coordination

modulation synthesis 94. The constructed MOFs with a size of

200 nm exhibited excellent stability with no obvious changes in

particle size or composition after keeping in 0.1 mol·L−1 HClO4

for 3 days. Due to the large surface area, the MOFs with a size

of 200 nm showed good ORR performance, indicating that

particle size and morphology of MOFs are important for

catalytic performance.

3.3 Hydrogen evolution reaction (HER)

For the new green energy economy, numerous efforts have

been made to fabricate efficient electrocatalysts for HER 95–98.

Zhao and co-workers applied ultrathin NiFe-MOF nanosheets on

nickel foam for electrochemical hydrogen evolution (Fig. 7a) 98.

When the current density reached 10 mA·cm−2, NiFe-MOF

nanosheets possessed the overpotential of 134 mV, smaller than

that of analogs (Ni- and bulk NiFe-MOF). Electrocatalytic

performances of ultrathin nanosheets can achieve no detectable

activity decay under robust operation for 2000 s (Fig. 7b,c).

Zheng et al. applied electrically conductive MOF-74-M (Co, Ni,

Zn) as templates to load tiny Pd nanoclusters (Pd@MOF-74),

which displayed outstanding catalytic performance for HER 99.

Owing to the high dispersion and the tiny size of Pd clusters, the

Pd@MOF-74-Co-3 exhibited the excellent HER activity with

the Tafel slope of 57 mV·dec−1. Huang et al. used

hexaiminohexaazatrinaphthalene (HAHATN) as organic ligands

to construct bimetallic 2D conductive MOFs 100.

Ni3(Ni3·HAHATN)2 nanosheets with a small bandgap of 0.19 eV

displayed a high conductivity and the overpotential of

nanosheets was just 115 mV at 10 mA·cm−2. Owing to the extra

unsaturated M-N2 moiety and the expanded in-plane porous

structure of 2.7 nm, the fabricated sheets showed a low Tafel

slope of 45.6 mV·dec−1 and excellent electrocatalytic stability.

3.4 Nitrogen reduction reaction (NRR)

Nitrogen conversion is a crucial reaction for many products

such as fertilizers, drugs as well as chemicals 101–104. Thanks to

the difficulty of activation of the inertness of the triple bonds,

nitrogen conversion is hard to achieve under ambient conditions.

Through the Haber-Bosch process, the production of NH3 is

usually energy-intensive and demands extreme reaction

conditions such as high temperature and pressure 62,105. The

Fig. 7 (a) The synthesis process of NiFe-MOF; (b) LSV curves of different materials;

(c) Chronoamperometric tests of NiFe-MOF before and after 2000 s.

(a–c) Adapted from Springer publisher 98.

Page 8: Conductive Metal-Organic Frameworks for Electrocatalysis ...

物理化学学报 Acta Phys. -Chim. Sin. 2021, 37 (7), 2010025 (8 of 14)

highly ordered MOFs with microporous structure was crucial in

gas-phase catalysis 106. Sun and coworkers fabricated 2D

conductive MOF (Mo3(HAB)2) by coordination of transition

metal ions (TM) with Hexa-aminobenzene, exhibiting high

catalytic activity and selectivity for NRR 107. Through DFT

calculations, the first hydrogenation step of *NNH formation

with positive ΔG values (+0.34 eV) limited the reaction rate in

the distal and alternating pathways. And the Mo-based MOF can

reduce overpotential for the NRR, compared to traditional pure

metals and some transition metal nitrides, or even some of the

single metal-doped materials (Fig. 8). Recently, Xiong et al.

employed Co3(HHTP)2 MOF (HHTP =

hexahydroxytriphenylene) as efficient catalysts for

electrocatalytic N2-NH3 conversion 108, which achieved a high

yield of 22.14 μg·h–1·mg–1 for NH3, and there is no obvious

change after a 24-hour operation.

3.5 Oxygen evolution reaction (OER)

OER is a significant process in water splitting electrolyzers

and metal-air batteries 109. The overpotential needed to be

applied for promoting OER performance, due to the sluggish

kinetics. Although many benchmark electrocatalysts, such as

IrO2, and RuO2, have been widely investigated, the efficient and

abundant MOF-based materials as alternative electrocatalysts

can control the cost and improve the performance 110,111.

Shen and co-workers applied a host (Fe3(μ3-O)(BDC)3 (Fe3)

to fit and anchor Co2(RCOO)4(H2O)2 clusters which were

unstable in aqueous solution, reaching extraordinary OER

activity at pH = 13 with a low overpotential of 225 mV 112. Wang

group reported a Ni/Fe bimetallic MOF (Ni/Fe-BTC) for oxygen

evolution with a low overpotential of 270 mV at 10 mA·cm–2 and

a high turnover frequency of 468 h–1 (Fig. 9a) 113. Subsequently,

ultrathin bimetal-MOF nanosheets (NiCo-UMOFNs) with three

coordination structural layers have been synthesized by the same

group, which exhibited a high electrochemical OER

performance in 1 mol·L−1 KOH 114. As shown in Fig. 9b, the

OER process on UMOFNs involves four steps ((i) adsorption,

(ii, iii) dissociation, and (iv) desorption). Lu and coworkers

employed post-synthetic ion-exchange to prepare MAF-X27-

OH (Co2(μ-OH)2(bbta)) which drastically improved the

performance for OER (Fig. 9c) 115. MAF-X27-OH(Cu)

displayed the overpotential of 292 mV at 10.0 mA·cm−2 at pH =

14 (Fig. 9d), better than that of MAF-X27-OH, Co3O4, and

Co(OH)2, which attributed to the lower energy barrier by the

direct participation of the μ-OH− ligands through the

intraframework.

Cao et al. employed the ultrasmall Fe(Ni)-MOF cluster with

a size of 2–5 nm decorated on Ni(Fe)-MOF nanosheets on NiFe

alloy foam (NFF) for the OER, which achieved a current density

of 10 mA·cm−2 within 227 mV and showed a small charge-

transfer resistance of 1.54 Ω (Fig. 10a,b) 116. The electron transfer

between Ni2+ and Fe3+ in NiFe-NFF modulated the electronic

environment of the Ni centers, thus enhancing the coupling

effect between Ni and Fe active sites. The features of the unique

nanostructure, the coupling effect, and higher intrinsic activity

make the OER performance of catalysts better than that of the

benchmark of RuO2. Besides, the better interface contact

between substrate and electrocatalyst contributes to the

improvement of the OER performance. Subsequently, the same

group vertically inlaid ultrathin NiFe MOF nanosheets (1.5 nm)

into a 3D ordered macroporous structure of NiFe hydroxide

(OM-NFH) by in situ growth method for enhancing the OER

performance (Fig. 10c) 117. Attributing to the efficient mass

conduction between ordered macroporous hydroxide and

nanosheets, effective electron transfer across hybrid material,

and the highly exposed active centers, the electrocatalyst with

high conductivity (Fig. 10d) exhibited an excellent catalytic

activity with a current density of 10 mA·cm−2 at 270 mV and

long-term stability up to 10 h, much better than that of RuO2. In

the same year, Zheng et al. synthesized several Fe/Ni bimetallic

MOFs by solvothermal approach for improving OER

performance, which can achieve a current density of 50

Fig. 8 The mechanisms of the NRR (distal, right; alternating, left).

Adapted from Royal Society of Chemistry publisher 107.

Fig. 9 (a) The scheme of Ni/Fe bimetallic MOF; (b) DFT calculation

(NiCo-UMOFNs), (Ni/Co, green; Ni/Co, purple; O, red; C, grey;

H, white); (c) 3D Structure of MAF-X27-OH; (d) LSV curves of

different simples.

(a) Adapted from ACS Publications publisher 113. (b) Adapted from Springer

publisher 114. (c, d) Adapted from ACS Publications publisher 115.

Page 9: Conductive Metal-Organic Frameworks for Electrocatalysis ...

物理化学学报 Acta Phys. -Chim. Sin. 2021, 37 (7), 2010025 (9 of 14)

mA·cm–2 at 270 mV 118.

4 The challenges and opportunities for

conductive MOFs To date, various MOF-based catalysts have been fabricated

and widely applied in catalysis as shown in Table 1. Particularly,

the conductive MOF-based materials have exhibited

significantly improved performance in different energy

conversion reactions. Therefore, the development of novel

MOFs with high conductivity is highly desired, yet remains a

great challenge.

(1) Conductivity

Due to the inherent property of organic ligands, the

conductivity of pristine MOFs is limited, which may become the

dominant factor in restricting the advance of MOFs with high

electroactivity. Choosing a proper strategy, such as fabricating

continuous metal-ligand chains by coordination bonds,

improving the π–π stacking interactions of organic components,

or integrating guest molecular in the inherent porosity for

Table 1 The performance of conductive MOFs for different energy conversion reactions.

Type Catalyst Electrolyte Potential FE Ref.

CO2RR Fe_MOF-525 films 1.0 mol·L−1 TBAPF6 + 1 mol·L−1 TFE −1.3 V vs. NHE ∼100% (CO + H2) 77

HKUST-1 (Cu,Ru) 0.5 mol·L−1 KHCO3 −1.0 V vs. Ag/AgCl 47.2% (alcohol) 80

Al2(OH)2TCPP-Co 0.5 mol·L−1 potassium carbonate −0.7 V vs. RHE 76% (CO) 82

Ag@Al-PMOFs 0.1 mol·L−1 KHCO3 −1.1 V vs. RHE 55.8% (CO) 83

Cu2(CuTCPP) 0.5 mol·L−1 EMIMBF4 −1.55 V vs. Ag/Ag+ 68.4% (formate); 16.8% (acetate) 84

bismuthine (Bi-ene) 1 mol·L−1 KOH -0.57V vs. RHE 99.8% (formate) 85

ORR Ni3(HITP)2 0.1 mol·L−1 KOH ∼0.75 V vs. RHE 63% (H2O2) 27

(Co)PCN222 0.1 mol·L−1 HClO4 0.43V vs. RHE 94

NRR Mo3(HAB)2 0.18 V 107

Co3(HHTP)2 0.5 mol·L−1 LiClO4 –0.40 V vs. RHE 3.34% 108

OER Fe3(μ3-O)(bdc)34Co2(na)4(LT)23 water at pH = 13 225 mV 112

Fe/Ni-BTC 0.1 mol·L−1 KOH 270 mV 95% 113

NiCo-UMOFNs 1.0 mol·L−1 KOH ∼189 mV 99.3% 114

MAF-X27-OH(Cu) 1.0 mol·L−1 KOH 292 mV 100% 115

NiFe-NFF 1.0 mol·L−1 KOH 227 mV ~100% 116

NiFe MOF/OM-NFH 1.0 mol·L−1 KOH 270 mV 117

HER NiFe-MOF 0.1 mol·L−1 KOH 240 mV 98

Pd@MOF-74 0.5 mol·L−1 H2SO4 −0.106 V vs. RHE 99

Ni3(Ni3·HAHATN)2 0.1 mol·L−1 KOH 115 mV 100

The potential for OER and the HER is the overpotential or potential at 10.0 mA·cm−2.

Fig. 10 (a) Scheme of synthesis of the NiFe-NFF electrode; (b) electrochemical impedance curves (EIS) conducted in O2-saturated 1.0 mol·L−1 KOH

solution at 270 mV; (c) The fabrication of NiFe MOF/OM-NFH; (d) EIS test in O2-saturated 1.0 mol·L−1 KOH solution.

(a, b) Adapted from Wiley publisher 116. (c, d) Adapted from Wiley publisher 117.

Page 10: Conductive Metal-Organic Frameworks for Electrocatalysis ...

物理化学学报 Acta Phys. -Chim. Sin. 2021, 37 (7), 2010025 (10 of 14)

improving the ability of the transfer of electrons and charge for

MOF is crucial. Besides, anchoring the MOF films on the

surface of the conductive substrates is demonstrated to be an

effective approach as well.

(2) Catalytic activity

The existing catalytic reaction system faces a series of

problems, such as high overpotential, low conversion efficiency,

and unsatisfactory product selectivity. Adjusting the adsorption

of intermediates on the active sites via tuning metal nodes or

organic ligands to decrease the free energy of adsorption and

desorption and maximizing the utilization of active sites via

introducing co-catalysts with highly dispersed single-metal

atoms, can effectively improve the electrocatalytic performance.

Appropriate design of morphology and size can increase the

number of exposed active sites, and thus amplify the catalytic

performance, such as higher FE and selectivity. Furthermore,

interface properties between different components,

electrocatalysts, and electrolytes are all relevant to catalytic

activity. (3) The selectivity

The selectivity is a crucial factor for the electrocatalytic

reactions, such as ORR, CO2RR, and NRR. A high-performance

electrocatalyst should not only exhibit prominent catalytic

activity but also own an outstanding selectivity. For example, the

conversion of CO2 to multi-carbon products on conductive

MOFs still suffers from a big challenge in selectivity. Tandem

catalysis in CO2RR could favor the formation of multi-carbon

products, such as ethanol and ethylene. Besides, the inter-

distance which is often ignored parameter is a key factor to

control the intermediates’ diffusion and adsorption, contributing

to improving the selectivity.

(4) The stability

Stability is an important factor in long-term operation. The

changes in composition, morphology, and structure could

elucidate stability trends 119,120 More importantly, similar to

catalytic activity, stability can be enhanced by adjusting the

metal centers or organic ligands. Besides, employing metal

centers with high oxidation states may also achieve this aim.

(5) The reaction mechanism

The reaction mechanisms should also be investigated to assess

the effect of various factors on kinetics to design and prepare

high-performance catalysts. Inspiringly, by combining the

theory of computer science and quantum chemistry with

experiments, lots of opportunities would be provided to promote

the evolution of high-performance MOFs, and it can also

decrease the workload of researchers. To further investigate the

reaction mechanism, more advanced in situ characterization

technologies, such as in situ X-ray photoelectron spectroscopy,

operando Raman spectra, inductively coupled plasma mass

spectrometry, X-ray absorption near-edge structure

spectroscopy, should be applied to detect electrochemical

reaction or monitor the change of surface state and the

transformation of atomic structure. For example, the detection of

the active sites on catalysts is beneficial to find the mechanisms

and optimize the structure of electrocatalysts. Moreover,

developing a large-scale synthesis strategy is of significance to

put conductive MOF into practice 121.

5 Conclusions and perspectives In summary, this review summarizes the recent progress of

conductive MOFs materials for a variety of energy conversion

reactions, such as CO2RR, NRR, HER, ORR, and OER.

Benefiting from their unique and intriguing physical and

chemical properties of crystallinity, large surface area, high

porosity, ordered porous structure, the abundant and ultrahigh

density of catalytic active sites, and the combination of the

merits of both heterogeneous and homogeneous electrocatalysts,

conductive MOFs (powders, single crystals, or thin films) have

been investigated for numerous applications, especially for

electrocatalytic applications. Besides, various synthetic

strategies (hydro/solvothermal synthesis and interface-assisted

synthesis) have been demonstrated to prepare conductive MOFs

with good charge-carrier mobility, controllable morphology, and

tailorable structures.

Various strategies have been developed to fabricate

conductive MOFs to improve pathways for charge transport and

increase of mobile charge carriers, and the morphology control,

component manipulation, and structural engineering, some

potential directions are listed as follows: (i) Synthesizing large

size of conductive MOFs single crystals is highly desirable. The

single crystals of high-quality MOFs can not only avoid the loss

of conductivity and improve intrinsic charge ability but also

favor the studies of structure-property relationships. (ii)

Preparing highly crystalline conductive MOFs thin films with

few defects is essential for enhancing charge transport and

broadening their electrical-related applications. (iii) it is

imperative to develop a novel synthetic methodology for

increasing the number of catalytic sites within the frameworks,

thus improving their catalytic performances. (iv) The fabrication

of heterostructured MOFs-covalent organic frameworks (COFs)

or MOFs-MOFs may provide more opportunities for designing

highly efficient MOFs-based electrocatalysts for different

energy conversion reactions. For example, the composite

containing intrinsic mixed valence of metals can promote the

conductivity, compared to single species. Compared with

amorphous materials, MOF-based materials are more

structurally characterizable due to their crystallinity. The

tolerability and tunability of molecular structure resulting from

tuning metal node and organic ligand, make it possible to design

and acquire desired materials with specific structures and

corresponding functions. Therefore, more advanced and

powerful in situ characterization techniques should be adapted

to enclose the catalytic mechanism of active sites, thereby

affording guidance to obtain MOFs-based electrocatalysts with

improved catalytic performances (activity and selectivity) and

enhanced stabilities. It is believed that all of these could endow

Page 11: Conductive Metal-Organic Frameworks for Electrocatalysis ...

物理化学学报 Acta Phys. -Chim. Sin. 2021, 37 (7), 2010025 (11 of 14)

MOFs materials with promising potential in the electrocatalytic

applications.

References

(1) Li, H.; Eddaoudi, M.; O’Keeffe, M.; Yaghi, O. M. Nature 1999, 402,

276. doi: 10.1038/46248

(2) Tranchemontagne, D. J.; Mendoza-Cortés, J. L.; O’Keeffe, M.; Yaghi,

O. M. Chem. Soc. Rev. 2009, 38, 1257. doi: 10.1039/B817735J

(3) Zhang, Z.; Chen, Y.; Xu, X.; Zhang, J.; Xiang, G.; He, W.; Wang, X.

Angew. Chem. Int. Ed. 2014, 53, 429. doi: 10.1002/anie.201308589

(4) Zhou, H.-C.; Long, J. R.; Yaghi, O. M. Chem. Rev. 2012, 112, 673.

doi: 10.1021/cr300014x

(5) Furukawa, H.; Cordova, K. E.; O’Keeffe, M.; Yaghi, O. M. Science

2013, 341, 1230444. doi: 10.1126/science.1230444

(6) Chen, X.-H.; Wei, Q.; Hong, J.-D.; Xu, R.; Zhou, T.-H. Rare Met.

2019, 38, 413. doi: 10.1007/s12598-019-01259-6

(7) Bavykina, A.; Kolobov, N.; Khan, I. S.; Bau, J. A.; Ramirez, A.;

Gascon, J. Chem. Rev. 2020, 120, 8468.

doi: 10.1021/acs.chemrev.9b00685

(8) Wang, Y.; Li, Q.; Shi, W.; Cheng, P. Chin. Chem. Lett. 2020, 31,

1768. doi: 10.1016/j.cclet.2020.01.010

(9) Zhang, K.; Liang, Z.; Zou, R. Sci. China Mater. 2020, 63, 7.

doi: 10.1007/s11426-019-9613-1

(10) Song, Z.; Zhang, L.; Doyle-Davis, K.; Fu, X.; Luo, J.-L.; Sun, X. Adv.

Energy Mater. 2020, 10, 2001561. doi: 10.1002/aenm.202001561

(11) Zhao, M.; Huang, Y.; Peng, Y.; Huang, Z.; Ma, Q.; Zhang, H. Chem.

Soc. Rev. 2018, 47, 6267. doi: 10.1039/C8CS00268A

(12) Liu, J.; Wöll, C. Chem. Soc. Rev. 2017, 46, 5730.

doi: 10.1039/C7CS00315C

(13) Li, B.; Wen, H.-M.; Cui, Y.; Zhou, W.; Qian, G.; Chen, B. Adv. Mater.

2016, 28, 8819. doi: 10.1002/adma.201601133

(14) Ding, M.; Flaig, R. W.; Jiang, H.-L.; Yaghi, O. M. Chem. Soc. Rev.

2019, 48, 2783. doi: 10.1039/C8CS00829A

(15) Li, J.-R.; Sculley, J.; Zhou, H.-C. Chem. Rev. 2012, 112, 869.

doi: 10.1021/cr200190s

(16) Zhang, Z.; Chen, Y.; He, S.; Zhang, J.; Xu, X.; Yang, Y.; Nosheen, F.;

Saleem, F.; He, W.; Wang, X. Angew. Chem. Int. Ed. 2014, 53, 12517.

doi: 10.1002/anie.201406484

(17) Dhakshinamoorthy, A.; Asiri, A. M.; Garcia, H. Adv. Mater. 2019, 31,

1900617. doi: 10.1002/adma.201900617

(18) Sun, L.; Campbell, M. G.; Dincă, M. Angew. Chem. Int. Ed. 2016, 55,

3566. doi: 10.1002/anie.201506219

(19) Talin, A. A.; Centrone, A.; Ford, A. C.; Foster, M. E.; Stavila, V.;

Haney, P.; Kinney, R. A.; Szalai, V.; El Gabaly, F.; Yoon, H. P.; et al.

Science 2014, 343, 66. doi: 10.1126/science.1246738

(20) Li, W.-H.; Deng, W.-H.; Wang, G.-E.; Xu, G. Energy Chem. 2020, 2,

100029. doi: 10.1016/j.enchem.2020.100029

(21) Li, W. H.; Ding, K.; Tian, H. R.; Yao, M.-S.; Nath, B.; Deng, W.-H.;

Wang, Y.; Xu, G. Adv. Funct. Mater. 2017, 27, 1702067.

doi: 10.1002/adfm.201702067

(22) Ko, M.; Mendecki, L.; Mirica, K. A. Chem. Commun. 2018, 54, 7873.

doi: 10.1039/C8CC02871K

(23) Li, P.; Wang, B. Isr. J. Chem. 2018, 58, 1010.

doi: 10.1002/ijch.201800078

(24) Stavila, V.; Talin, A. A.; Allendorf, M. D. Chem. Soc. Rev. 2014, 43,

5994. doi: 10.1039/C4CS00096J

(25) Bhardwaj, S. K.; Bhardwaj, N.; Kaur, R.; Mehta, J.; Sharma, A. L.;

Kim, K.-H.; Deep, A. J. Mater. Chem. A 2018, 6, 14992.

doi: 10.1039/C8TA04220A

(26) Clough, A. J.; Yoo, J. W.; Mecklenburg, M. H.; Marinescu, S. C.

J. Am. Chem. Soc. 2015, 137, 118. doi: 10.1021/ja5116937

(27) Miner, E. M.; Fukushima, T.; Sheberla, D.; Sun, L.; Surendranath, Y.;

Dincă, M. Nat. Commun. 2016, 7, 10942. doi: 10.1038/ncomms10942

(28) Miner, E. M.; Wang, L.; Dincă, M. Chem. Sci. 2018, 9, 6286.

doi: 10.1039/C8SC02049C

(29) Cheng, W.-Z.; Liang, J.-L.; Yin, H.-B.; Wang, Y.-J.; Yan, W.-F.;

Zhang, J.-N. Rare Met. 2020, 39, 815.

doi: 10.1007/s12598-020-01440-2

(30) Liu, X.; Yue, T.; Qi, K.; Qiu, Y.; Xia, B. Y.; Guo, X. Chin. Chem. Lett.

2020, 31, 2189. doi: 10.1016/j.cclet.2019.12.009

(31) Zhao, R.; Liang, Z.; Zou, R.; Xu, Q. Joule 2018, 2, 2235.

doi: 10.1016/j.joule.2018.09.019

(32) Shinde, S. S.; Lee, C. H.; Jung, J.-Y.; Wagh, N. K.; Kim, S.-H.; Kim,

D.-H.; Lin, C.; Lee, S. U.; Lee, J.-H. Energy Environ. Sci. 2019, 12,

727. doi: 10.1039/C8EE02679C

(33) Liu, J.; Song, X.; Zhang, T.; Liu, S.; Wen, H.; Chen, L. Angew. Chem.

Int. Ed. 2020, 59, 2. doi: 10.1002/anie.202006102

(34) Sheberla, D.; Bachman, J. C.; Elias, J. S.; Sun, C.-J.; Shao-Horn, Y.;

Dincă, M. Nat. Mater. 2017, 16, 220. doi: 10.1038/nmat4766

(35) Du, W.; Bai, Y.-L.; Yang, Z.; Li, R.; Zhang, D.; Ma, Z.; Yuan, A.; Xu,

J. Chin. Chem. Lett. 2020, 31, 2309. doi: 10.1016/j.cclet.2020.04.017

(36) Campbell, M. G.; Sheberla, D.; Liu, S. F.; Swager, T. M.; Dincă, M.

Angew. Chem. Int. Ed. 2015, 54, 4349. doi: 10.1002/anie.201411854

(37) Campbell, M. G.; Liu, S. F.; Swager, T. M.; Dincă, M. J. Am. Chem.

Soc. 2015, 137, 13780. doi: 10.1021/jacs.5b09600

(38) Aubrey, M. L.; Kapelewski, M. T.; Melville, J. F.; Oktawiec, J.;

Presti, D.; Gagliardi, L.; Long, J. R. J. Am. Chem. Soc. 2019, 141,

5005. doi: 10.1021/jacs.9b00654

(39) Meng, Z.; Aykanat, A.; Mirica, K. A. J. Am. Chem. Soc. 2019, 141,

2046. doi: 10.1021/jacs.8b11257

(40) Wu, G.; Huang, J.; Zang, Y.; He, J.; Xu, G. J. Am. Chem. Soc. 2017,

139, 1360. doi: 10.1021/jacs.6b08511

Page 12: Conductive Metal-Organic Frameworks for Electrocatalysis ...

物理化学学报 Acta Phys. -Chim. Sin. 2021, 37 (7), 2010025 (12 of 14)

(41) Huang, X.; Sheng, P.; Tu, Z.; Zhang, F.; Wang, J.; Geng, H.; Zou, Y.;

Di, C.-A.; Yi, Y.; Sun, Y.; Xu, W.; Zhu, D. Nat. Commun. 2015, 6,

7408. doi: 10.1038/ncomms8408

(42) Lahiri, N.; Lotfizadeh, N.; Tsuchikawa, R.; Deshpande, V. V.; Louie,

J. J. Am. Chem. Soc. 2017, 139, 19. doi: 10.1021/jacs.6b09889

(43) Wang, B.; Luo, Y.; Liu, B.; Duan, G. ACS Appl. Mater. Interfaces

2019, 11, 35935. doi: 10.1021/acsami.9b14319

(44) Song, X.; Wang, X.; Li, Y.; Zheng, C.; Zhang, B.; Di, C.-A.; Li, F.;

Jin, C.; Mi, W.; Chen, L.; Hu, W. Angew. Chem. Int. Ed. 2020, 59,

1118. doi: 10.1002/anie.201911543

(45) Zhao, W.; Peng, J.; Wang, W.; Liu, S.; Zhao, Q.; Huang, W. Coordin.

Chem. Rev. 2018, 377, 44. doi: 10.1016/j.ccr.2018.08.023

(46) Dong, R.; Zhang, Z.; Tranca, D. C.; Zhou, S.; Wang, M.; Adler, P.;

Liao, Z.; Liu, F.; Sun, Y.; Shi, W.; et al. Nat. Commun. 2018, 9, 2637.

doi: 10.1038/s41467-018-05141-4

(47) Yang, C.; Dong, R.; Wang, M.; Petkov, P. S.; Zhang, Z.; Wang, M.;

Han, P.; Ballabio, M.; Bräuninger, S. A.; Liao, Z.; et al. Nat.

Commun. 2019, 10, 3260. doi: 10.1038/s41467-019-11267-w

(48) Qiu, T.; Liang, Z.; Guo, W.; Tabassum, H.; Gao, S.; Zou, R. ACS

Energy Lett. 2020, 5, 520. doi: 10.1021/acsenergylett.9b02625

(49) Chu, S.; Majumdar, A. Nature 2012, 488, 294.

doi: 10.1038/nature11475

(50) Liu, J.; Zhu, D.; Guo, C.; Vasileff, A.; Qiao, S.-Z. Adv. Energy Mater.

2017, 7, 1700518. doi: 10.1002/aenm.201700518

(51) Wang, H.-F.; Chen, L.; Pang, H.; Kaskel, S.; Xu, Q. Chem. Soc. Rev.

2020, 49, 1414. doi: 10.1039/C9CS00906J

(52) Xie, L.; Skorupskii, G.; Dincă, M. Chem. Rev. 2020, 120, 8536.

doi: 10.1021/acs.chemrev.9b00766

(53) Sheberla, D.; Sun, L.; Blood-Forsythe, M. A.; Er, S.; Wade, C. R.;

Brozek, C. K.; Aspuru-Guzik, A.; Dincă, M. J. Am. Chem. Soc. 2014,

136, 8859. doi: 10.1021/ja502765n

(54) Narayan, T. C.; Miyakai, T.; Seki, S.; Dincă, M. J. Am. Chem. Soc.

2012, 134, 12932. doi: 10.1021/ja3059827

(55) Park, S. S.; Hontz, E. R.; Sun, L.; Hendon, C. H.; Walsh, A.; Van

Voorhis, T.; Dincă, M. J. Am. Chem. Soc. 2015, 137, 1774.

doi: 10.1021/ja512437u

(56) Xie, L. S.; Alexandrov, E. V.; Skorupskii, G.; Proserpio, D. M.;

Dincă, M. Chem. Sci. 2019, 10, 8558. doi: 10.1039/C9SC03348C

(57) Pathak, A.; Shen, J.-W.; Usman, M.; Wei, L.-F.; Mendiratta, S.;

Chang, Y.-S.; Sainbileg, B.; Ngue, C.-M.; Chen, R.-S.; Hayashi, M.;

et al. Nat. Commun. 2019, 10, 1721.

doi: 10.1038/s41467-019-09682-0

(58) Xie, L.; Skorupskii, G.; Dincă, M. Chem. Rev. 2020, 120, 8536.

doi: 10.1021/acs.chemrev.9b00766

(59) Makiura, R.; Motoyama, S.; Umemura, Y.; Yamanaka, H.; Sakata, O.;

Kitagawa, H. Nat. Mater. 2010, 9, 565. doi: 10.1038/nmat2769

(60) Dong, R.; Pfeffermann, M.; Liang, H.; Zheng, Z.; Zhu, X.; Zhang, J.;

Feng, X. Angew. Chem. Int. Ed. 2015, 54, 12058.

doi: 10.1002/anie.201506048

(61) Kambe, T.; Sakamoto, R.; Hoshiko, K.; Takada, K.; Miyachi, M.;

Ryu, J.-H.; Sasaki, S.; Kim, J.; Nakazato, K.; Takata, M.; et al. J. Am.

Chem. Soc. 2013, 135, 2462. doi: 10.1021/ja312380b

(62) Pal, T.; Kambe, T.; Kusamoto, T.; Foo, M. L.; Matsuoka, R.;

Sakamoto, R.; Nishihara, H. ChemPlusChem 2015, 80, 1255.

doi: 10.1002/cplu.201500206

(63) Sun, X.; Wu, K.-H.; Sakamoto, R.; Kusamoto, T.; Maeda, H.; Ni, X.;

Jiang, W.; Liu, F.; Sasaki, S.; Masunaga, H.; et al. Chem. Sci. 2017, 8,

8078. doi: 10.1039/C7SC02688A

(64) Pal, T.; Doi, S.; Maeda, H.; Wada, K.; Tan, C. M.; Fukui, N.;

Sakamoto, R.; Tsuneyuki, S.; Sasaki, S.; Nishihara, H. Chem. Sci.

2019, 10, 5218. doi: 10.1039/C9SC01144G

(65) Huang, X.; Li, H.; Tu, Z.; Liu, L.; Wu, X.; Chen, J.; Liang, Y.; Zou,

Y.; Yi, Y.; Sun, J.; et al. J. Am. Chem. Soc. 2018, 140, 15153.

doi: 10.1021/jacs.8b07921

(66) Sheberla, D.; Bachman, J. C.; Elias, J. S.; Sun, C.-J.; Shao-Horn, Y.;

Dincă, M. Nat. Mater. 2017, 16, 220. doi: 10.1038/nmat4766

(67) Du, W.; Bai, Y.-L.; Yang, Z.; Li, R.; Zhang, D.; Ma, Z.; Yuan, A.; Xu,

J. Chin. Chem. Lett. 2020, 31, 2309. doi: 10.1016/j.cclet.2020.04.017

(68) Campbell, M. G.; Sheberla, D.; Liu, S. F.; Swager, T. M.; Dincă, M.

Angew. Chem. Int. Ed. 2015, 54, 4349. doi: 10.1002/anie.201411854

(69) Campbell, M. G.; Liu, S. F.; Swager, T. M.; Dincă, M. J. Am. Chem.

Soc. 2015, 137, 13780. doi: 10.1021/jacs.5b09600

(70) Dunwell, M.; Lu, Q.; Heyes, J. M.; Rosen, J.; Chen, J. G.; Yan, Y.;

Jiao, F.; Xu, B. J. Am. Chem. Soc. 2017, 139, 3774.

doi: 10.1021/jacs.6b13287

(71) Zhao, C.; Dai, X.; Yao, T.; Chen, W.; Wang, X.; Wang, J.; Yang, J.;

Wei, S.; Wu, Y.; Li, Y. J. Am. Chem. Soc. 2017, 139, 8078.

doi: 10.1021/jacs.7b02736

(72) Lu, Y.; Zhang, J.; Wei, W.; Ma, D. D.; Wu, X. T.; Zhu, Q. L. ACS

Appl. Mater. Interfaces 2020, 12, 37986.

doi: 10.1021/acsami.0c06537

(73) Li, X; Zhu, Q. L. EnergyChem 2020, 2, 100033.

doi: 10.1016/j.enchem.2020.100033

(74) Ma, D. D.; Zhu, Q. L. Coord. Chem. Rev. 2020, 422, 213483.

doi: 10.1016/j.ccr.2020.213483

(75) Aubrey, M. L.; Kapelewski, M. T.; Melville, J. F.; Oktawiec, J.;

Presti, D.; Gagliardi, L.; Long, J. R. J. Am. Chem. Soc. 2019, 141,

5005. doi: 10.1021/jacs.9b00654

(76) Meng, Z.; Aykanat, A.; Mirica, K. A. J. Am. Chem. Soc. 2019, 141,

2046. doi: 10.1021/jacs.8b11257

(77) Hod, I.; Sampson, M. D.; Deria, P.; Kubiak, C. P.; Farha, O. K.;

Hupp, J. T. ACS Catal. 2015, 5, 6302. doi: 10.1021/acscatal.5b01767

Page 13: Conductive Metal-Organic Frameworks for Electrocatalysis ...

物理化学学报 Acta Phys. -Chim. Sin. 2021, 37 (7), 2010025 (13 of 14)

(78) Albo, J.; Vallejo, D.; Beobide, G.; Castillo, O.; Castaño, P.; Irabien,

A. ChemSusChem 2017, 10, 1100. doi: 10.1002/cssc.201600693

(79) Dong, B.-X.; Qian, S.-L.; Bu, F.-Y.; Wu, Y.-C.; Feng, L.-G.; Teng,

Y.-L.; Liu, W.-L.; Li, Z.-W. ACS Appl. Energy Mater. 2018, 1, 4662.

doi: 10.1021/acsaem.8b00797

(80) Perfecto-Irigaray, M.; Albo, J.; Beobide, G.; Castillo, O.; Irabien, A.;

Pérez-Yáñez, S. RSC Adv. 2018, 8, 21092.

doi: 10.1039/C8RA02676A

(81) Qiu, Y.-L.; Zhong, H.-X.; Zhang, T.-T.; Xu, W.-B.; Su, P.-P.; Li, X.-F.;

Zhang, H.-M. ACS Appl. Mater. Interfaces 2018, 10, 2480.

doi: 10.1021/acsami.7b15255

(82) Kornienko, N.; Zhao, Y.; Kley, C. S.; Zhu, C.; Kim, D.; Lin, S.;

Chang, C. J.; Yaghi, O. M.; Yang, P. J. Am. Chem. Soc. 2015, 137,

14129. doi: 10.1021/jacs.5b08212

(83) Guntern, Y. T.; Pankhurst, J. R.; Vávra, J.; Mensi, M.; Mantella, V.;

Schouwink, P.; Buonsanti, R. Angew. Chem. Int. Ed. 2019, 58, 12632.

doi: 10.1002/anie.201905172

(84) Wu, J.-X.; Hou, S.-Z.; Zhang, X.-D.; Xu, M.; Yang, H.-F.; Cao, P.-S.;

Gu, Z.-Y. Chem. Sci. 2019, 10, 2199. doi: 10.1039/C8SC04344B

(85) Cao, C.; Ma, D. D.; Gu, J. F.; Xie, X.; Zeng, G.; Li, X.; Han, S.

G.; Zhu, Q. L.; Wu, X. T.; Xu, Q. Angew. Chem. Int. Ed. 2020, 59,

15014. doi: 10.1002/anie.202005577

(86) Brezny, A. C.; Johnson, S. I.; Raugei, S.; Mayer, J. M. J. Am. Chem.

Soc. 2020, 142, 4108. doi: 10.1021/jacs.9b13654

(87) Pegis, M. L.; Wise, C. F.; Martin, D. J.; Mayer, J. M. Chem. Rev.

2018, 118, 2340. doi: 10.1021/acs.chemrev.7b00542

(88) Zhao, S.; Yin, H.; Du, L.; He, L.; Zhao, K.; Chang, L.; Yin, G.; Zhao,

H.; Liu, S.; Tang, Z. ACS Nano 2014, 8, 12660.

doi: 10.1021/nn505582e

(89) Lai, Q.; Zheng, L.; Liang, Y.; He, J.; Zhao, J.; Chen, J. ACS Catal.

2017, 7, 1655. doi: 10.1021/acscatal.6b02966

(90) Guo, J.; Li, Y.; Cheng, Y.; Dai, L.; Xiang, Z. ACS Nano 2017, 11,

8379. doi: 10.1021/acsnano.7b03807

(91) Yin, P.; Yao, T.; Wu, Y.; Zheng, L.; Lin, Y.; Liu, W.; Ju, H.; Zhu, J.;

Hong, X.; Deng, Z.; et al. Angew. Chem. Int. Ed. 2016, 55, 10800.

doi: 10.1002/anie.201604802

(92) Liu, X.-H.; Hu, W.-L.; Jiang, W.-J.; Yang, Y.-W.; Niu, S.; Sun, B.;

Wu, J.; Hu, J.-S. ACS Appl. Mater. Interfaces 2017, 9, 28473.

doi: 10.1021/acsami.7b07410

(93) Yoon, H.; Lee, S.; Oh, S.; Park, H.; Choi, S.; Oh, M. Small 2019, 15,

1805232. doi: 10.1002/smll.201805232

(94) Chen, G.; Stevens, M. B.; Liu, Y.; King, L. A.; Park, J.; Kim, T. R.;

Bao, Z.; Sinclair, R.; Jaramillo, T. F.; Bao, Z. Small Methods 2020, 4,

2000085. doi: 10.1002/smtd.202000085

(95) Roger, I.; Shipman, M. A.; Symes, M. D. Nat. Rev. Chem. 2017, 1, 1.

doi: 10.1038/s41570-016-0003

(96) Chen, W.; Pei, J.; He, C.-T.; Wan, J.; Ren, H.; Wang, Y.; Dong, J.;

Wu, K.; Cheong, W.-C.; Mao, J.; et al. Adv. Mater. 2018, 30,

1800396. doi: 10.1002/adma.201800396

(97) Liu, T.; Li, P.; Yao, N.; Cheng, G.; Chen, S.; Luo, W.; Yin, Y. Angew.

Chem. Int. Ed. 2019, 306, 627. doi: 10.1002/anie.201901409

(98) Duan, J.; Chen, S.; Zhao, C. Nat. Commun. 2017, 8, 15341.

doi: 10.1038/ncomms15341

(99) Zheng, F.; Zheng, C.; Gao, X.; Du, C.; Zhang, Z.; Chen, W.

Electrochim. Acta 2019, 7, 9743.

doi: 10.1016/j.electacta.2019.03.175

(100) Huang, H.; Zhao, Y.; Bai, Y.; Li, F.; Zhang, Y.; Chen, Y. Adv. Sci.

2020, 7, 2000012. doi: 10.1002/advs.202000012

(101) Yang, C.; Zhu, Y.; Liu, J.; Qin, Y.; Wang, H.; Liu, H.; Chen, Y.;

Zhang, Z.; Hu, W. Nano Energy 2020, 77, 105126.

doi: 10.1016/j.nanoen.2020.105126

(102) Geng, Z.; Liu, Y.; Kong, X.; Li, P.; Li, K.; Liu, Z.; Du, J.; Shu, M.;

Si, R.; Zeng, J. Adv. Mater. 2018, 30, 1803498.

doi: 10.1002/adma.201803498

(103) Guo, C.; Ran, J.; Vasileff, A.; Qiao, S.-Z. Energy Environ. Sci. 2018,

11, 45. doi: 10.1039/C7EE02220D

(104) Yuan, L.; Wu, Z.; Jiang, W. Tang, T.; Niu, S.; Hu J.-S. Nano Res.

2020, 13, 1376. doi: 10.1007/s12274-020-2637-8

(105) Abghoui, Y.; Garden, A. L.; Howalt, J. G.; Vegge, T.; Skúlason, E.

ACS Catal. 2016, 6, 635. doi: 10.1021/acscatal.5b01918

(106) Fukushima, T.; Drisdell, W.; Yano, J.; Surendranath, Y. J. Am.

Chem. Soc. 2015, 137, 10926. doi: 10.1021/jacs.5b06737

(107) Cui, Q.; Qin, G.; Wang, W.; K. R, G.; Du, A.; Sun, Q. J. Mater.

Chem. A 2019, 7, 14510. doi: 10.1039/C9TA02926E

(108) Xiong, W.; Cheng, X.; Wang, T.; Luo, Y.; Feng, J.; Lu, S.; Asiri,

A. M.; Li, W.; Jiang, Z.; Sun, X. Nano Res. 2020, 13, 1008.

doi: 10.1007/s12274-020-2733-9

(109) Zhou, J.; Dou, Y.; Zhou, A.; Guo, R.-M.; Zhao, M.-J.; Li, J.-R. Adv.

Energy Mater. 2017, 7, 1602643. doi: 10.1002/aenm.201602643

(110) Li, M.; Xia, Z.; Huang, Y.; Tao, L.; Chao, Y.; Yin, K.; Yang, W.;

Yang, W.; Yu, Y.; Guo, S. Acta Phys. -Chim. Sin. 2020, 36, 1912049.

[李蒙刚, 夏仲泓, 黄雅荣, 陶璐, 晁玉广, 尹坤, 杨文秀, 杨微

微, 于永生, 郭少军. 物理化学学报, 2020, 36, 1912049.]

doi: 10.3866/PKU.WHXB201912049

(111) Zheng, F.; Zhang, Z.; Xiang, D.; Li, P.; Du, C.; Zhuang, Z.; Li,

X.; Chen, W. J. Colloid Interf. Sci. 2019, 555, 541.

doi: 10.1016/j.jcis.2019.08.005

(112) Shen, J.-Q.; Liao, P.-Q.; Zhou, D.-D.; He, C.-T.; Wu, J.-X.; Zhang,

W.-X.; Zhang, J.-P.; Chen, X.-M. J. Am. Chem. Soc. 2017, 139,

1778. doi: 10.1021/jacs.6b12353

(113) Wang, L.; Wu, Y.; Cao, R.; Ren, L.; Chen, M.; Feng, X.; Zhou, J.;

Wang, B. ACS Appl. Mater. Interfaces 2016, 8, 16736.

Page 14: Conductive Metal-Organic Frameworks for Electrocatalysis ...

物理化学学报 Acta Phys. -Chim. Sin. 2021, 37 (7), 2010025 (14 of 14)

doi: 10.1021/acsami.6b05375

(114) Zhao, S.; Wang, Y.; Dong, J.; He, C.-T.; Yin, H.; An, P.; Zhao, K.;

Zhang, X.; Gao, C.; Zhang, L.; et al. Nat. Energy 2016, 1, 16184.

doi: 10.1038/nenergy.2016.184

(115) Lu, X.-F.; Liao, P.-Q.; Wang, J.-W.; Wu, J.-X.; Chen, X.-W.; He,

C.-T.; Zhang, J.-P.; Li, G.-R.; Chen, X.-M. J. Am. Chem. Soc. 2016,

138, 8336. doi: 10.1021/jacs.6b03125

(116) Cao, C.; Ma, D. D.; Xu, Q.; Wu, X. T.; Zhu, Q. L. Adv. Funct.

Mater. 2019, 29, 1807418. doi: 10.1002/adfm.201807418

(117) Li, X.; Ma, D. D.; Cao, C.; Zou, R.; Xu, Q.; Wu, X. T.; Zhu, Q.

L. Small 2019, 15, 1902218. doi: 10.1002/smll.201902218

(118) Zheng, F.; Zhang, C.; Gao, X.; Du, C.; Zhuang, Z.; Chen, W.

Electrochim. Acta 2019, 306, 627.

doi: 10.1016/j.electacta.2019.03.175

(119) Liu, J.; Zhu, D.; Guo, C.; Vasileff, A.; Qiao, S.-Z. Adv. Energy

Mater. 2017, 7, 1700518. doi: 10.1002/aenm.201700518

(120) Zheng, F.; Zhang, Z.; Zhang, C.; Zhang, C.; Chen, W. ACS

Omega 2020, 5, 2495. doi: 10.1021/acsomega.9b03295

(121) Centi, G. SmartMat 2020, e1005. doi: 10.1002/smm2.1005