Top Banner
VOLUME 5 • ISSUE 5 • SEPTEMBER 2018 Special Topic: Single-Atom Catalysts
72

Special Topic: Single-Atom Catalysts - Western Engineering...Predictive approach of heterogeneous catalysis Metal-organic frame-work assisted synthesis of single-atom catalysts for

Oct 07, 2020

Download

Documents

dariahiddleston
Welcome message from author
This document is posted to help you gain knowledge. Please leave a comment to let me know what you think about it! Share it to your friends and learn new things together.
Transcript
Page 1: Special Topic: Single-Atom Catalysts - Western Engineering...Predictive approach of heterogeneous catalysis Metal-organic frame-work assisted synthesis of single-atom catalysts for

VOLUME 5 • ISSUE 5 • SEPTEMBER 2018

Volume 5

, Issue 5, 6

05

–78

3 S

eptember 2

01

8Special Topic:

Single-Atom Catalysts

nsr_5_5_cover.indd 1nsr_5_5_cover.indd 1 03/10/18 4:10 PM03/10/18 4:10 PM

Page 2: Special Topic: Single-Atom Catalysts - Western Engineering...Predictive approach of heterogeneous catalysis Metal-organic frame-work assisted synthesis of single-atom catalysts for

Special Topic: Single-Atom Catalysts

Guest Editors: Nanfeng Zheng and Tao Zhang

Abhaya Datye (pictured)and Yong Wang

Preface: single-atom catalysts as a new generation of hetero-geneous catalysts

Predictive approach of heterogeneous catalysis

Metal-organic frame-work assisted synthesis of single-atom catalysts for energy applications

Single-atom catalysts by the atomic layer deposition technique

Atom trapping: a novel approach to generate thermally stable and regen-erable single-atom catalysts

Yao Zheng and Shi-Zhang Qiao (pictured)

Lei Zhang, Mohammad Norouzi Banis and Xueliang Sun (pictured)

625 626 628

633

Coordination chemistry of atomically dispersed catalysts

Pengxin Liu and Nanfeng Zheng (pictured)

636

Theoretical understanding of the stability of single-atom catalysts

Jin-Cheng Liu, Yan Tang, Yang-Gang Wang, Tao Zhang and Jun Li (pictured)

638

630

Page 3: Special Topic: Single-Atom Catalysts - Western Engineering...Predictive approach of heterogeneous catalysis Metal-organic frame-work assisted synthesis of single-atom catalysts for

COVER IMAGE

Single-atom catalysts (SACs) with well-defined mononuclear active sites on a solid support are

expected to bridge homogeneous and heterogeneous catalysis. See the section of Special Topic

(pages 625–693).

Image credit: Leilei Zhang, Yujing Ren, Aiqin Wang and Xiaoling Yu.

Single-atom catalysis: a new field that learns from tradition

Jean-Marie Basset (Reporter: Philip Ball)

690

Single-atom heterogeneous catalysts based on distinct carbon nitride scaffolds

642

Single-atom catalyst: a rising star for green synthesis of fine chemicals

Leilei Zhang, Yujing Ren, Wengang Liu, Aiqin Wang (left) and Tao Zhang (right)

653

Recent advances in the precise control of isolated single-site catalysts by chemical methods

Zhijun Li, Dehua Wang , Yuen Wu (left) and Yadong Li (right)

673

RESEARCH ARTICLE

Page 4: Special Topic: Single-Atom Catalysts - Western Engineering...Predictive approach of heterogeneous catalysis Metal-organic frame-work assisted synthesis of single-atom catalysts for

GUEST EDITORIALNational Science Review

5: 625, 2018

doi: 10.1093/nsr/nwy095

Advance access publication 14 September 2018

Special Topic: Single-Atom Catalysts

Preface: single-atom catalysts as a new generation of heterogeneouscatalystsNanfeng Zheng 1,∗,† and Tao Zhang2,∗,†

Heterogeneous catalysis plays a leading role in the chemical in-dustry. Creating heterogeneous catalysts with minimal use ofmetals but well-defined catalytic sites is highly desired for de-veloping cost-effective and green chemical processes. With sin-gle metal atoms dispersed on supports, single-atom catalysts arefeatured in their optimized usage of metal and also the similarchemical environments of metal sites. It has been long believedthat the single-atom catalysts should not be stable enough tobe prepared via conventional catalyst-preparation techniques.However, in 2011, Tao Zhang’s group in Dalian Institute ofChemical Physics, cooperating with Jun Li from Tsinghua Uni-versity and Jingyue Liu from Arizona State University, success-fully demonstrated the feasibility to fabricate stable and efficientsingle-atomPt catalysts using a simple co-precipitationmethod.The research also presented a systematic methodology to in-vestigate single-atom catalysts using advanced characterizationtechniques. During the past several years, the work has greatlystimulated the rapid development of the field.There are nownu-merous studies demonstrating that single-atom catalysts serveas an excellent catalyst system to bridge homogeneous catalystswith well-defined molecular structures and heterogeneous cat-alysts with structural complexity. With more and more contri-butions to this rapidly growing field, single-atom catalysts havebeen emerging as next-generation heterogeneous catalysts forvarious applications.

This special topic aims at highlighting the latest research ad-vances in the field of single-atom catalysts. The research high-light by Shi-Zhang Qiao et al. focuses on how to use metal-organic framework materials to assist the synthesis of single-atom catalysts for energy applications, which has been recentlystudied extensively by many research groups in China. The re-view article by Tao Zhang and Aiqin Wang et al. summarizesthe developments of single-atom catalysts for green synthesis offine chemicals. Yadong Li and Yuen Wu et al. reviewed the re-cent advances in the precise control of single-atom catalysts bychemical methods. The perspective articles in this special topiccover several issues of single-atom catalysts ranging from prepa-ration techniques, catalyticmechanisms and also theoretical un-derstanding. Xueliang Sun et al. discuss the development of the

atomic layer deposition technique as an effectivemethod to pre-pare single-atom catalysts for various applications. In the per-spective article by Abhaya Datye et al., atom trapping is demon-strated as a viable method for the synthesis of oxide-supportedsingle atoms of platinum-group metals. Jean-Marie Basset et al.discuss the research advances to design single-atom catalystsbased on surface organometallic chemistry. From the viewpointof coordination chemistry, Nanfeng Zheng et al. discuss in theirperspective article different roles of supports in single-atom cat-alysts. The stability of single-atom catalysts is a critical issuefor their industrial applications. The perspective article by JunLi et al. focuses on the theoretical understanding of the sta-bility issue of single-atom catalysts. This special topic also of-fers a research article contributed by Javier Perez-Ramırez andSharonMitchell et al.which demonstrates that palladium atomscan be effectively isolated on carbon nitride scaffolds beyondgraphitic heptazine-based polymers for creating stable and effi-cient single-atom Pd catalysts for selective hydrogenation. Aninterview with Jean-Marie Basset is also included in this spe-cial topic to share his personal views on the development andprospects of the field of single-atom catalysts.

As guest editors, we would like to express our sincere appre-ciation to all the authors, reviewers and also the editorial officeofNSR for their efforts tomake this special topic possible.Wehopethat this special topic will gain broad attention from chemistry,physics, materials science and other related fields.

Nanfeng Zheng1,∗ ,† and Tao Zhang2,∗ ,†1State Key Laboratory for Physical Chemistry of Solid Surfaces, Collaborative

Innovation Center of Chemistry for Energy Materials, and Department of

Chemistry, College of Chemistry and Chemical Engineering, Xiamen University,

China2State Key Laboratory of Catalysis, iChEM (Collaborative Innovation, Center of

Chemistry for Energy Materials), Dalian Institute of Chemical Physics, Chinese

Academy of Sciences, China∗Corresponding authors.E-mails: [email protected]; [email protected]†Guest Editor of Special Topic

C©TheAuthor(s) 2018. Published by Oxford University Press on behalf of China Science Publishing &Media Ltd. All rights reserved. For permissions, please e-mail:[email protected]

Page 5: Special Topic: Single-Atom Catalysts - Western Engineering...Predictive approach of heterogeneous catalysis Metal-organic frame-work assisted synthesis of single-atom catalysts for

RESEARCH HIGHLIGHTNational Science Review

5: 626–627, 2018

doi: 10.1093/nsr/nwy010

CHEMISTRY

Special Topic: Single-Atom Catalysts

Metal-organic framework assisted synthesis of single-atom catalystsfor energy applicationsYao Zheng1 and Shi-Zhang Qiao1,2,∗

Metal single-atom catalysts (M-SACs)are emerging as a new research fron-tier because of their low-coordinationmetal atoms and uniform structures. Asa representation of maximum atom uti-lization efficiency, M-SACs demonstrateunique properties in awide variety of het-erogeneous catalysis and electrocatalyticprocesses [1,2]. Metal-organic frame-works (MOFs) are composed of metal-containingnodes andorganic linkers, andare typically characterized as atomicallydispersed metal sites. With the versatilityof metal ions and ligands, the direct py-rolysis of MOFs serves as an ideal routefor preparing variousM-SACs. However,this seemingly simple strategy is stillhighly challenging due to the drastic reac-tive processes that occur at high temper-ature that rapidly convert metal ions intoaggregated nanoparticles. Very recently,Yadong Li et al. developed a general syn-thetic strategy for M-SAC preparationby utilizing particular zeolitic imidazo-late frameworks (ZIFs)—a kind ofMOFthat are topologically isomorphic withzeolites [3–5]. The resultant M-SACs(M = Fe, Co, Ni) demonstrated excel-lent activities for energy-related electro-catalytic reactions.

The most important principle ofthis methodology is the utilization oflow-boiling-point atoms, such as Zn (mp420◦C, bp 907◦C), in ZIF-8 and Zn/Cobimetallic ZIF-67, which evaporatedat high temperature, leaving abundantN-rich defects. As a result, the iso-lated metal ions were closely anchoredthrough N-coordination, avoiding their

aggregation during pyrolysis. For thesynthesis of Fe and Ni single atoms inan N-carbon matrix (Fe-SAs/CN andNi-SAs/CN), Fe(acac)3 and Ni(NO3)2were selected as metal sources, re-spectively, and were confined in themolecular-cages of ZIF-8 (Fig. 1) [3,4].Co-SAs/CN was synthesized from a par-tially Zn-replaced ZIF-67 with a Zn/Co= 1:1 (Zn/Co-ZIF-67) precursor (Fig.1) [5]. After pyrolysis at 900–1000◦C inan Ar atmosphere, the metal ions werethermally reduced and atomically dis-

ZIF-8Ni-SAs/CN

Co-SAs/CN

Zn/Co-ZIF-67

Fe-SAs/CN

O2

CO2

CO

OH-

O2

OH-

Figure 1. Schematic illustration of the precur-sor and structures of Fe, Ni and Co single atoms

supported in N-carbon matrices. Color code:

grey: carbon; royal blue: nitrogen; white: hydro-

gen; red: oxygen; light blue: zinc; brown: iron;

green: nickel; purple: cobalt.

persed in the N-carbon matrix. This wasdirectly observed by high-angle annulardark-field scanning transmission elec-tron microscopy. The precise molecularstructures of these isolated metal atomswere well identified by extended X-rayabsorption fine structure and X-rayabsorption near-edge structure. Asshown in Fig. 1, the ZIF-8-derivedFe-SAs/CN and Ni-SAs/CN, andZn/Co-ZIF-67-derived Co-SAs/CN,show a 4, 3 and 2 metal coordinationnumber with surrounding N atoms,respectively.

WithpreciseN-coordination andhighmetal loadings of 1.5–4wt.%, allM-SACsshowed superior electrocatalytic activi-ties. For example, Fe-SAs/CN demon-strated a half-wave potential of 0.9 V forthe oxygen-reduction reaction in alka-line electrolyte—a key cathodic step forproton exchange membrane fuel cells,even outperforming commercial Pt/Cand most non-precious-metal catalysts[3]. Ni-SAs/CN exhibited an excellentturnover frequency for the electroreduc-tion of CO2 (5273 h–1), with a Faradaicefficiency for CO generation of 71.9%[4].

In summary, Yadong Li et al. de-veloped a general and well-designedmethod for the simple preparation ofM-SACs with the assistance of ZIFs.Due to the wide variety of metal atomsin versatile MOFs, this accurately con-trolled methodology may present someguidelines for the rational design andmodulation of M-SACs for broaderapplications.

C©TheAuthor(s) 2018. Published by Oxford University Press on behalf of China Science Publishing &Media Ltd. All rights reserved. For permissions, please e-mail:[email protected]

Page 6: Special Topic: Single-Atom Catalysts - Western Engineering...Predictive approach of heterogeneous catalysis Metal-organic frame-work assisted synthesis of single-atom catalysts for

RESEARCH HIGHLIGHT Zheng and Qiao 627

Yao Zheng1 and Shi-Zhang Qiao1,2,∗1School of Chemical Engineering, The University of

Adelaide, Australia2School of Materials Science and Engineering,

Tianjin University, China∗Corresponding author.E-mail: [email protected]

REFERENCES1. Yang X, Wang A and Qiao B et al. Acc Chem Res

2013; 46: 1740–8.2. Liu P, Zhao Y and Qin R et al. Science 2016; 352:797–800.

3. Chen Y, Ji S and Wang Y et al. Angew Chem Int Ed

2017; 56: 6937–41.

4. Zhao C, Dai X and Yao T et al. J Am Chem Soc 2017;

139: 8087–91.5. Yin P, Yao T and Wu Y et al. Angew Chem Int Ed

2016; 55: 10800–5.

Page 7: Special Topic: Single-Atom Catalysts - Western Engineering...Predictive approach of heterogeneous catalysis Metal-organic frame-work assisted synthesis of single-atom catalysts for

PERSPECTIVES

MATERIALS SCIENCE

Special Topic: Single-Atom Catalysts

Single-atom catalysts by the atomic layer deposition techniqueLei Zhang, Mohammad Norouzi Banis and Xueliang Sun∗

Noble metal nanocatalysts have beenwidely applied in the petrochemicalindustry, medicine, environmental pro-tection and energy sectors. To decreasethe cost and maximize the utilizationefficiency, single-atom catalysts (SACs),as a new frontier in heterogeneouscatalysis, have attracted considerableattention due to their unique catalyticproperties. Unlike traditional nanocat-alysts, the catalytic performance ofsingle-atom catalysts is highly dependenton their low-coordination environment,quantum size effects and metal–supportinteraction. Recently, several methodshave been developed for the preparationof single-atomcatalysts, including the im-pregnation method [1], co-precipitationmethod [2], photo-reduction method[3] and atomic layer deposition (ALD)method [4–6]. Among these, the ALDprocess is a powerful approach forstudying the relationship between thecatalysts’ structure and their catalyticperformances, as it has a great capabilityto precisely control the deposition ofsingle atoms and nanoclusters. In thisperspective, we will briefly discuss therecent progress in the rational designof single-atom catalysts through atomiclayer deposition. In addition, we willsummarize the key issues for the de-velopment of ALD methods and theoutlook for future research trends.

Generally, one complete ALD cyclecontains two main processes (Fig. 1a)[7]. Taking the deposition of single Ptatoms on graphene as an example, thefirst step is the reaction of the Pt pre-cursor with the adsorbed oxygen on thesurface of the graphene. The second step

is an oxygen pulse to convert the pre-cursor ligands to Pt–O species to forma new adsorbed oxygen layer on the Ptsurface (Fig. 1b). During the ALD pro-cess, Pt catalysts with atomically precisedesign and control can be synthesizedby adjusting the ALD cycles. In 2013,Sun and co-workers fabricated single Ptatoms on graphene nanosheets by ALDfor the first time [3]. The graphene sub-strate exhibited many carbon vacanciesand defects, which were favorable for an-choring Pt precursors. Up to now, single-atomPt catalysts have also been achievedon nitrogen-doped graphene and CeO2supports [4,5]. When single Pt atoms aredepositedonCeO2, the atoms tend tode-posit on the Ce rows of the CeO2(110)and (100) facets, instead of the (111)facet [5]. In addition to single Pt atoms,single Pd atom catalysts have also beensuccessfully deposited on graphene by al-ternately exposing Pd(hfac)2 and forma-lin at 150◦C [6]. Although several typesof single-atom catalysts have been ob-tained through the ALD method, large-scale production of single-atom catalystsusing the ALD process is still challeng-ing, and optimization of the ALDprocessis required. For example, the use of thespatial atomic layer deposition techniquecould greatly improve the industrializa-tion process [8].

Atomic-resolution transmissionelectron microscopy (TEM), scanningtunneling microscopy (STM) andradiation-ray absorption spectroscopyinvestigations are three typical charac-terization methods for detecting SACs.We can directly observe the single atomsthrough atomic-resolution TEM and

STM images while the synchrotron-based X-ray absorption spectrum canprovide information on the overall struc-ture of the catalysts. The formation ofSACs can be concluded by studying theextendedX-ray-absorption fine-structure(EXAFS) spectrum of the correspondingcatalysts. Local atomic structure infor-mation including coordination numberand atomic distances can be derivedfrom the fitted EXAFS spectra in Rspace. The reduction of peak intensity ofmetal–metal peaks in these spectra canbe a good indicator for the formationof SACs. Though we can conclude theformation of single-atomic catalyststhrough these methods, it remains agreat challenge to identify the accuratebinding site between the single atomsand the support. As a result, the devel-opment of advanced characterizationmethods is essential for widespread syn-thesis and application of SACs. In addi-tion, to attainphysical insights into the re-actions, advanced in situ/operandocharacterization techniques would benecessary. A systematic understanding ofthe reaction mechanisms will be crucialfor discovering the growth behavior ofsingle atoms on different substrates andits practical applications.

The SACs obtained through ALDhave extremely high atom-utilization ef-ficiency, which makes them highly ac-tive catalysts for several catalytic re-actions. The SACs have been provento be highly efficient catalysts for sev-eral electrochemical reactions, includingmethanol oxidation, the hydrogen evo-lution reaction and the oxygen reduc-tion reaction (ORR) [3,4,9,10]. Sun and

C©TheAuthor(s) 2018. Published by Oxford University Press on behalf of China Science Publishing &Media Ltd. All rights reserved. For permissions, please e-mail:[email protected]

Page 8: Special Topic: Single-Atom Catalysts - Western Engineering...Predictive approach of heterogeneous catalysis Metal-organic frame-work assisted synthesis of single-atom catalysts for

PERSPECTIVES Zhang et al. 629

co-workers reported that the single-atomPt catalysts on graphene showed 10 timeshigher activity for methanol oxidationand superior CO tolerance compared toconventional Pt/C catalysts [3]. In ad-dition, they found that the mass activityof single-atom Pt catalysts on nitrogen-doped graphene was around 37.4 timesgreater than that of the Pt/C catalyst [4].Thanks to the strong bonding energy be-tween Pt and nitrogen-doped graphene,the activity of the catalysts dropped only4% after 1 000 cycles, indicating a highdurability. For theoxygen reduction reac-tion, twodifferentpathwaysweredemon-stratedon single-atomPt catalysts [9,10].The support plays a significant role in theselective catalytic reaction route and fi-nal product. For instance, single Pt atomson a sulfur-doped zeolite-templated car-bon support could selectively transferthe O2 to H2O2 through a two-electronpathway [9], while single Pt atoms onnitrogen-doped carbon black exhibited

Single atomic catalysts

Step 1 Step 2

Step 1

Step 2

Substrate

Chamber

Substrate Substrate Substrate

Two main half-reactions

(a)

(b)

(c)

Figure 1. (a) Schematic illustrations of the deposition of single-atomic Pt on substrates through the ALD method. (b) The detailed two main half-

reactions during a whole ALD cycle. (c) Four key research areas for the development of the ALD technique for single-atom catalysts.

great performance for highly efficientfour-electron ORR [10]. Up till now, theinitial active sites for the enhancedmech-anismare still unclear.More fundamentalstudies should focus on the influence ofthe support on the performance of single-atomcatalysts,whichmightprovide a sys-tematic understanding of the interactionsbetween the SAC and substrates. In ad-dition to electrochemical catalytic reac-tions, the SACs also exhibited superiorperformance as heterogeneous catalystsin several traditional catalytic reactions.For example, Pt/FeOx and Pt/CeO2catalysts exhibited a remarkable activitytowards both CO oxidation and pref-erential oxidation of CO in H2 [2,5].The single-atom Pd/graphene catalystsobtained through ALD showed 100%butene selectivity in selective hydrogena-tion of 1,3-butadiene [6]. The adsorp-tion mode of 1,3-butadiene and the en-hanced steric effect induced by singlePd atoms played key roles in improv-

ing the butene selectivity. To furtherincrease the activity, the rational de-sign of SACs with multiple compositionsshould be explored in the future, be-cause bimetallic catalysts can exhibit im-proved performance compared to puremetals. The area-selective ALD methodprovides an effective way to fabricate cat-alysts withmulti-compositions. Elam andco-workers synthesized several types ofbimetallic nanoparticles by precisely con-trolling the ALD conditions [11]. Re-cently, Pt2 dimers have been fabricatedthrough selective deposition of a sec-ondary Pt atom onto the preliminary one[12]. Based on the above points, we caninfer that, under certain ALD conditions,the fabrication of bimetallic dimers is fea-sible throughdeposition of the secondarymetal atom on the first single-atomicmetal.With the formation of dimer struc-tures, the electronic structure of SACscould be tuned, which would further in-crease the activity of SACs to a new level.

Page 9: Special Topic: Single-Atom Catalysts - Western Engineering...Predictive approach of heterogeneous catalysis Metal-organic frame-work assisted synthesis of single-atom catalysts for

630 Natl Sci Rev, 2018, Vol. 5, No. 5 PERSPECTIVES

However, the metal loading is usuallyvery low due to the problems associatedwith agglomeration during the prepara-tion process. Further Pt deposition dur-ing the ALD process tends to adsorb onthe existing Pt atoms, which will also re-sult in the formation of clusters. As a re-sult, the development of ALD techniquesto achieve high loading of SACs is stilla challenge for their future commercialapplication. Furthermore, the selectivityof substrates is important for the dis-persion of singe atoms [3,11]. Recently,Zeng et al. found that the mass loadingof single Pt atom catalysts can reach upto 7.5% on MoS2 [13]. In addition tothe low loading issues, another key pa-rameter for the SACs is the durabilityof the catalysts during the catalytic re-actions. Single-atom catalysts are highlymobile and tend to aggregate during thecatalytic reactions due to the high sur-face energy of SACs. The interaction be-tween SAC and substrate plays an impor-tant role in their stability. Several stud-ies have shown that nitrogen-doped car-bon black, graphene and nitrogen-dopedgraphene can form strong coordinationsites with metal, thus observably improv-ing their stability [3,4,10]. However, thesingle atoms still tend to aggregate on tra-ditional substrates. More studies shouldfocus on the development of effective

routes to stabilize the catalysts. The ALDdeposition of metal oxide around Pt cat-alysts has been proven to be an effectiveway to stabilize the Pt clusters. For exam-ple, the ORR stability of Pt clusters wassignificantly increased by selective depo-sition of ZrO2 around Pt clusters to forma cage structure [14].This method of sta-bilization may also be applied in single-atom systems. In addition, the fabricationof porous structures to pin the atoms inthe hole might also increase the stabilityof single-atom catalysts. We believe thatthe ALDmethodmight open up new op-portunities in fabrication and optimiza-tion of SACs for improved activity anddurability in heterogeneous reactions.

FUNDINGThis work was supported by the NaturalSciences and Engineering Research Council ofCanada (NSERC), Canada Research Chair (CRC)Program, Canada Foundation for Innovation(CFI) and the University of Western Ontario.

Lei Zhang, Mohammad Norouzi Banis and

Xueliang Sun∗

Department of Mechanical and Materials

Engineering, The University of Western Ontario,

Canada∗Corresponding author.E-mail: [email protected]

REFERENCES1. Yang S, Kim J and Tak YJ et al. Angew Chem Int

Ed 2016; 55: 2058–62.2. Qiao B, Wang A and Yang X et al. Nat Chem 2011;

3: 634–41.3. Liu P, Zhao Y and Qin R et al. Science 2016; 352:797–800.

4. Cheng N, Stambula S and Wang D et al. Nat

Commun 2016; 7: 13638.5. Wang C, Gu XK and Yan H et al. ACS Catal 2017;

7: 887–91.6. Yan H, Cheng H and Yi H et al. J Am Chem Soc

2015; 137: 10484–7.7. Sun S, Zhang G and Gauquelin N et al. Sci Rep

2013; 3: 1775.8. George SM. Chem Rev 2010; 110: 111–31.9. Choi CH, Kim M and Kwon HC et al. Nat Commun

2016; 7: 10922.10. Liu J, Jiao M and Lu L et al. Nat Commun 2017;

8: 15938.11. Lu J, Low K-B and Lei Y et al. Nat Commun 2014;

5: 3264.12. Yan H, Lin Y and Wu H et al. Nat Commun 2017;

8: 1070.13. Li H, Wang L and Dai Y et al. Nat Nanotechnol

2018; 13: 411–7.14. Cheng N, Banis MN and Liu J et al. Adv Mater

2015; 27: 277–81.

National Science Review

5: 628–630, 2018

doi: 10.1093/nsr/nwy054

Advance access publication 22 May 2018

CHEMISTRY

Special Topic: Single-Atom Catalysts

Atom trapping: a novel approach to generate thermally stable andregenerable single-atom catalystsAbhaya Datye 1,∗ and Yong Wang2,3

An important goal of heterogeneous cat-alyst synthesis is the dispersion of the ac-tive metal uniformly on a catalyst sup-port, ideally achieving atomic dispersion.Isolated single atoms dispersed on oxidesupports provide efficient utilization ofscarce platinum group metals (PGMs).The enhanced reactivity of Pt1/FeOx by

Qiao et al. [1] generated a lot of ex-citement in this field and helped set inmotion the field of single-atom catal-ysis (SAC). Since then, we have seennumerous reports of higher reactivityand better selectivity for SACs in arange of catalytic reactions. Synthesismethods for depositing transitionmetals,

through strong electrostatic adsorption(SEA), ion exchange, co-precipitation,grafting, impregnation or deposition–precipitation, are well developed. Togenerate and maintain single-atomcatalysts using these currently availablemethods, it is necessary to use lowmetal loading and to limit the operating

Page 10: Special Topic: Single-Atom Catalysts - Western Engineering...Predictive approach of heterogeneous catalysis Metal-organic frame-work assisted synthesis of single-atom catalysts for

PERSPECTIVES Datye and Wang 631

Wavenumber (cm-1)

Abs

orba

nce

(a.u

.)

2400 2300 2200 2100 2000 1900

1wt.% Pt/CeO2 oxidized

CO Oxidation CO flow stopped -0 minCO Oxidation CO flow stopped -2 minCO Oxidation CO flow stopped -10 min

(a) (c)(b)

Figure 1. (a) AC-STEM of Pt/CeO2 nanorods, prepared via atom trapping, (b) DRIFTS of adsorbed CO after 30 min of reaction and with flowing O2 while

CO flow was stopped and (c) AC-STEM image of the catalyst after multiple cycles of CO oxidation showing that the single-atom species are stable

under lean conditions. Adapted from [9].

temperatures to prevent agglomerationof single atoms into nanoparticles [2].For industrial applications, it would bedesirable to achieve high metal loadingsand stable performance at high tempera-tures, which can be achieved by themeth-ods of atom trapping that we describehere.

One of the earliest reports on the sta-bilization of isolated metal atoms camefrom the work of Kwak et al. [3] onPt atoms on penta coordinated Al3+ ingamma alumina and a similar mecha-nism was also invoked for amorphousmesoporous alumina [2]. Aberration-corrected STEM images provided evi-dence for single-atom Pt species and thespecific sites on the alumina were iden-tified using NMR spectroscopy [3]. Inlater work, it was reported that the keyparameter helping to prevent sintering ofPt was the formation of a stable coher-ent interface with the support, such asPt(111) with MgAl2O4(111) helping tostabilize Pt against sintering [4]. How-ever, on both of these supports, aluminaand MgAl2O4, heating the Pt catalyst tohigh temperatures (800◦C) in air leadsto growth of large Pt metal particles thatyield sharpXRDpeaks.The reason for therapid growth of Pt particle size is the highvapor pressure of PtO2 when Pt is heatedin air [5]. Since a number of PGMs formvolatile metal oxides [5], it is importantto develop methods to stabilize PGMsunder oxidative conditions.

In the course of our work on diesel ox-idation catalysts, we discovered [6] that

mobile Pt oxides reacted with PdO form-ing metallic Pt–Pd alloy particles. Thereaction of PtO2 with PdO is thermo-dynamically favored even in flowing air,forming metallic alloy Pt–Pd alloys. Sin-tering of Pt is slowed down considerablyin the presence of Pd. But excess PdOis needed to provide the sites for captur-ing mobile Pt species.Themechanism bywhich PdO helps to slow the sintering ofPt was termed ‘Regenerative Trapping’[7], analogous to the self-regeneratingperovskite catalyst reported previously[8]. When applying this approach toother oxides, we discovered that CeO2can effectively trap the PtO2 vapor, yield-ing stable isolated single atomsof Pt2+ onthe surface of ceria [9].

Figure 1a shows aberration-correctedHAADF images of the Pt/CeO2 catalystas prepared by heating the Pt precursorwith ceria at 800◦C in flowing air [9].The sample was used for multiple runsof CO oxidation under lean conditions,with 1% CO, 1.5% O2 and the balanceHe, and temperatures up to 300◦C.DRIFTS spectra of this catalyst duringCO oxidation are shown in Fig. 1b indi-cating a single prominent band at∼2100cm−1 that is characteristic of Pt2+ on theceria. When the CO flow was stoppedand O2 flow continued, the CO banddid not decrease significantly in inten-sity, indicating that this CO is stronglybound to ionic Pt. Figure 1c shows theimage after reaction, showing that theisolated single atoms of Pt are stable andthere is no change in the catalyst after

multiple reaction cycles. These resultsconfirm that the atom-trapping approachis successful in generating atomically dis-persed Pt catalysts that are thermallystable at 800◦C in air. However, thesecatalysts are not very reactive for CO ox-idation due to strong CO adsorption onionic Pt. We therefore explored variousmethods of activation to enhance the re-activity. For example, treatment in steamat 750◦C activated the ceria by creatingoxygen vacancies and hydroxyls in thevicinity of Pt single sites that were stableat high temperatures and allowed the cat-alyst to achieve high reactivity for COox-idation at low temperatures [10].

Atom trapping is not restricted to ce-ria supports. It was previously reportedthat heating a Pd/La-alumina catalyst at700◦C in air also forms isolated Pd2+ onthe alumina support [11]. Similarly, theregeneration of Pt–Sn/Al2O3 catalystswas assisted by the presence of atomi-cally dispersed Snon the alumina support[12]. It is evident then that, in the case ofAl2O3 supports, it is necessary to providetrapping sites, such as La or Sn, to helpcreate atomically dispersed species. Incontrast, the ceria support is unique in itsability to trap Pt species.We further stud-ied Pt–Sn/CeO2 catalysts for propanedehydrogenation, where they show sta-ble performance with high selectivity[13]. Due to the reducing conditions,the Pt atoms become mobile and alloywith the Sn, forming bimetallic nanopar-ticles in the working catalyst (Fig. 2a).However, the catalyst can be readily

Page 11: Special Topic: Single-Atom Catalysts - Western Engineering...Predictive approach of heterogeneous catalysis Metal-organic frame-work assisted synthesis of single-atom catalysts for

632 Natl Sci Rev, 2018, Vol. 5, No. 5 PERSPECTIVES

(b)(a)

PtSn(Sn-rich)cluster

Pt singleatom

Highlydispersed

SnO2SnO2 particle

Steady state Regenerated

Oxidation

Figure 2. (a) AC-STEM image of the working catalyst, after 6 h at 680◦C in C3H8 and H2 and (b)

after oxidative regeneration in flowing air at 580◦C. Adapted from [13].

regenerated under oxidizing conditions,with the Pt forming single-atom specieswhile the Sn remains in the form of SnO2particles (Fig. 2b). These results demon-strate the easewithwhich ceria supportedcatalysts can be regenerated to recreatethe single-atom species.

In summary, high-temperature vaporphase synthesis is demonstrated as a vi-able method for synthesis of atomicallydispersed PGM catalysts.There are otherreports of transformation of noble metalnanoparticles into thermally stable singleatoms using high temperatures [8,14].Oxidizing conditions are always used forcatalyst regeneration [15], since this rep-resents themost commonapproach to re-move carbonaceous deposits (coke) thatlead to catalyst deactivation. PGMs canform mobile oxides in most cases [5].Even when the oxide has a low vaporpressure, as in the case of PdO, it is

possible to emit mobile Pd species underoxidizing conditions as demonstrated bythe regeneration of Pd/La-Al2O3 creat-ing single-atom Pd species [11]. There-fore, understanding the sites for atomtrapping could pave the way for broaderapplication of this approach for synthesissingle-atom catalysts.

FUNDINGThis work was supported by the US Department ofEnergy (DOE) ( DE-FG02–05ER15712).

Conflict of interest statement.None declared.

Abhaya Datye 1,∗ and Yong Wang2,3

1Department of Chemical and Biological

Engineering and Center for Micro-Engineered

Materials, University of New Mexico, USA2The Gene & Linda Voiland School of Chemical

Engineering and Bioengineering, Washington

State University, USA

3Institute for Integrated Catalysis, Pacific

Northwest National Laboratory, USA∗Corresponding author.E-mail: [email protected]

REFERENCES1. Qiao B, Wang A and Yang X et al. Nat Chem 2011;

3: 634–41.2. Zhang Z, Zhu Y and Asakura H et al. Nat Commun

2017; 8: 16100.3. Kwak J, Hu J and Mei D et al. Science 2009; 325:1670–3.

4. Li W-Z, Kovarik L and Mei D et al. Chem Mater

2014; 26: 5475–81.5. McCarty J, Lau K and Hildenbrand D et al. Stud-

ies in Surface Science and Catalysis. Amsterdam:

Elsevier, 1997, 601–7.

6. Carrillo C, Johns T and Xiong H et al. J Phys Chem

Lett 2014; 5: 2089–93.7. Carrillo C, DeLaRiva A and Xiong H et al. Appl

Catal B-Environ 2017; 218: 581–90.8. Nishihata Y, Mizuki J and Akao T et al. Nature

2002; 418: 164–7.9. Jones J, Xiong H and Delariva A et al. Science

2016; 353: 150–4.10. Nie L, Mei D and Xiong H et al. Science 2017; 358:

1419–23.

11. Peterson E, DeLaRiva A and Lin S et al. Nat Com-

mun 2014; 5: 4885.12. Pham H, Sattler J and Weckhuysen B et al. ACS

Catal 2016; 6: 2257–64.13. Xiong H, Lin S and Goetze J et al. Angew Chem

Int Ed 2017; 56: 8986–91.14. Wei S, Li A and Liu J-C et al. Nat Nanotechnol

2018; 13: 856–61.15. Sattler J, Ruiz-Martinez J and Santillan-Jimenez

E et al. Chem Rev 2014; 114: 10613–53.

National Science Review

5: 630–632, 2018

doi: 10.1093/nsr/nwy093

Advance access publication 5 September 2018

Page 12: Special Topic: Single-Atom Catalysts - Western Engineering...Predictive approach of heterogeneous catalysis Metal-organic frame-work assisted synthesis of single-atom catalysts for

PERSPECTIVES Basset and Pelletier 633

CHEMISTRY

Special Topic: Single-Atom Catalysts

Predictive approach of heterogeneous catalysisJean-Marie Basset∗ and Jeremie D.A. Pelletier

‘Predictive catalysis’ or ‘catalysis bydesign’ has recently advanced heteroge-neous catalysis by using the conceptualtool of surface fragments [1] such as ‘sur-face organometallic fragments’ (SOMF)or ‘surface coordination fragments’(SCF) to achieve and understand apresumed catalytic cycle (see Fig. 1).One or several fragments of the moleculeare linked to one metal atom linked tothe surface ([M]-H, [M]-R, [M]=CR2,[M]≡CR, [M]=O, [M]=NR, [M]-O-OH in which [M] is a surface metal atomlinked to an oxide by one, two or severalsigma or pi bonds). Surface fragments arethe logical continuation of the abundantwork published in the field of SurfaceOrganometallic Chemistry (SOMC).In this paradigm based on molecularunderstanding of surface catalytic sites,one ‘single’ metal atom is surrounded byligands and linked to the surface of anoxide via covalent or ionic bonds [2]. It isa continuation of classical heterogeneouscatalysis in the direction of single-atomcatalysis (SAC). In SAC, single metalatoms are also linked to an oxide bycoordination, covalent or ionic bondsbut are at the other extreme of catalysisby metal nanoparticles because SAC isthe result of a conceptual evolution frommetal nanoparticles to a single atom bysize reduction of the nanoparticle. It is acontinuation of homogeneous catalysisbut with a rigid surface as a ligand.

SOMC has allowed the discoveryof new catalytic reactions (e.g. Ziegler-Natta depolymerization [3], alkanemetathesis [1], non-oxidative methanecoupling [1], cyclo-alkane metathesis[4], etc.) and has improved the ac-tivity, the selectivity or the lifetime ofknown ones. The concepts of molecularchemistry (organic, organometallic,coordination chemistry) are the keys toexplaining how bonds can be broken andformed [2]. In this context, the reactivity

of SOMF or SCF and their sequence inthe cycle are pivotal to the overall out-come of catalysis.

SOMC can generate catalytic sitesthat are in principle identical (single-siteor close to single atom) by graftingtransition metal atoms onto highly dehy-droxylated metal oxide support handledunder a controlled atmosphere. Thisstrategy, limited to metal-oxides ormetallic surfaces, presents considerableadvantages over traditional hetero-geneous catalysts in which variouspopulations of potentially active metal-lic sites coexist. All the steps of thepreparation are carefully controlledusing the methods of organometallicand coordination chemistry. Hence,the coordination sphere of the graftedmetal can be accurately determined(well-defined catalytic site) by modernsolid/surface characterization tech-niques (elemental analysis, in situ IR, insitu UV, Solid State Nuclear MagneticResonance spectroscopy (SS NMR),Extended X-ray absorption fine structure(EXAFS) and in operando EXAFS, etc.)[2]. The surface should be consideredas a bulky rigid ligand preventing mostundesired interferences between cat-alytic sites (e.g. leading to bimoleculardeactivation). The relationship betweenstructure and activity become possibleto establish; with the addition of theSOMF tools, it is now a predictablediscipline.

The various steps of the catalytic cy-cle are monitored to understand deacti-vation, to increase activity and/or selec-tivity by changing the support or ligandenvironment of the ‘active site’ (Fig. 2).The existing gap between heterogeneouscatalysis and homogeneous catalysis hasalmost completely disappeared, becausethe elementary steps of molecular chem-istry are applicable to ‘single-atom catal-ysis’. We shall review here some of

the recent catalytic results obtained onoxides.

Metal hydrides are the simplest andmost frequent surface fragments, yet donot belong to ‘organometallic’ classifi-cation stricto sensu. ([M]-H) are mostlygenerated by hydrogenolysis of metal-alkyl ([M]-R) and generally promotelow-temperature C-H bond activationof alkanes (i.e. methane activation) [5].([M]-H) and ([M]-R) can convert intoeach other by β-hydride elimination ofthe metal-alkyl and CH insertion. Thishas been evidenced in alkane depolymer-ization [3] (polyethylene is transformedinto diesel-range gasoline by group 4metal-based catalysts under hydrogen).

([M]-R) SOMF have numerousexamples as polymerization catalysts[6,7]. Another case is the bis-alkyl SOMFwith dual ([M]-R) SOMF for propanehomologation to higher alkanes [8].

Surface carbenes were first evidencedwith Nb [9]([(≡Si-O)2Nb(=CH)],Mo Mo[(≡ Si-O-)Mo(= CHCMe3)(= NH)Np] [10] and Re [(≡Si-O-)Re(Np) (≡C-CMe3) (=CHCMe3)])[10]. They were eventually successfullyemployed as catalysts for olefin metathe-sis. This reactivity is specific to the([M]=CR2) SOMF, consistently withthe metallacyclobutane intermediateproposed by Chauvin [11]. Variation ofactivity and selectivity had been linkedto both the nature of the metal employedand that of the spectator ligands (i.e. oxo,imido, amido, alkyl, etc.) [12,13].

The first multifunctional SOMF frag-ments [M](H) (=CR2)] were identi-fied following the discovery of alkanemetathesis using tantalum, tungsten andthenmolybdenum hydride [14] catalystssupported on alumina or silica in 1997[15]. In this reaction, saturated hydro-carbons, linear and branched, were re-arranged to longer or shorter paraffins.For example, n-propane canbe converted

Page 13: Special Topic: Single-Atom Catalysts - Western Engineering...Predictive approach of heterogeneous catalysis Metal-organic frame-work assisted synthesis of single-atom catalysts for

634 Natl Sci Rev, 2018, Vol. 5, No. 5 PERSPECTIVES

Hydroaminoalkylation Metallaziridine

Ziegler-NattadepolymerizationAlkanehydrogenolysisHydrogenolysis ofwaxes

Ziegler-Nattapolymerization

Metal alkyls

Metallocarbenes

Cyclotrimerization of alkynes

CO2 and epoxidecycloaddition

MetallocarbenesHydrides / alkyl

Metallocarbenes

Metal halide

Metalloxo

Oxidation

Oxidative dehydrogenation of propane

Metalloalkoxo

Olefins epoxidation

Imine metathesis

Metallimine

Metal hydridesNeutral / cationic

(H)n

[M]

SOMF

SCF

(R)n

[M]

[M]

O

[M]

O OR

[M]

OR[M]

NR

[M]

Cl

[M]

CHR

RNCH2[M]

[M] R′(H)

CHR

[M] R′

CR

n-alkane metathesisBranched alkane metathesisCycloalkane isometathesisMethane coupling to ethaneCleavage alkanes by methaneEthylene to propyleneButane to dieselHydrometathesisIsometathesis of n-decane1-butene to propylene2-butene to propylene

OO

O

OM

M′

R

R

y n

X

R = Functionalligand

X = Spectatorligand

Olefin metathesisCycloolefin ROMP

Metalloxoalkoxo

Figure 1. Overview of surface fragments: surface organometallic fragments (SOMF) and surface coordination fragments (SCF). Revised version from

the figure published by Pelletier and Basset, in Acc Chem Res 2016; 49: 664–77.

into (C1, C2, C4, C5 . . .) paraffins un-der mild conditions. It was a chemicalbreakthrough taking into account the in-ertness of the sp3 carbon-hydrogen orC–C bonds.

The multifunctional character of theW(=CH2)(H) was at the origin of thefirst cascade reaction on a single metalatom leading from ethylene to propy-lene by a succession of dimerizationof ethylene to butene-1, isomeriza-tion of butene-1 to butene-2 andcross-metathesis between butene-2 andethylene to give propylene.

Another notable application of[M](H)(=CR2) was cyclooctanemetathesis [16] to produce both ringcontraction and ring dimerization

using [(≡Si-O-)WMe5] as a precursor.Again we have a cascade reaction with[M](H)(=CR2) that allows RingOpen-ing Metathesis (ROM) dimerization ordouble bond migration followed by aring-closing metathesis step.

Another ‘polyfunctional’ fragment([M](H)(≡CR)) explains catalyticterminal alkyne cyclotrimerization.Silica-supported tungsten carbynecomplexes have shown high Turn OverNumbers (TONs) without generatingsignificant alkyne metathesis products[4].

The first example of ([M] ′η−2

(N(R)CH2)) SOMF was obtainedwith the isolation of silica-supportedzirconiaaziridine in 2013 [16]—anactive

catalyst for the hydroamino-alkylation ofolefins [17].

The first example of ([M]=N) SCFwas evidenced by being the first hetero-geneous catalyst for imine metathesisusing [(≡Si–O–)Zr(=NEt)NEt2] [18].Treatment with an imine substrateresulted in imido/imine (=NRi, R: Et,Ph) exchange (metathesis) with theformation of [(≡Si–O–)Zr (= NPh)NEt2]. ([M]=N) effectively catalyseimine/imine cross-metathesis.

([M]-OR) SCF were involved inolefin catalysed by silica-supportedtitanium complexes [2]. Various(≡SiO)nTi(OCap)4-n (OCap =OR,OSiR3, OR; R=hydrocarbyl) supportedon MCM-41 have been evaluated as

Page 14: Special Topic: Single-Atom Catalysts - Western Engineering...Predictive approach of heterogeneous catalysis Metal-organic frame-work assisted synthesis of single-atom catalysts for

PERSPECTIVES Basset and Pelletier 635

Start the reactionby a Fragment or itsclosest precursor

Fragment

Fragment 1

Fragment 4Reversible andirreversible processes(deactivation)

Fragment 3

Fragment 2

Catalysis

by DesignBA

ReactionintermediatesSOMF or SCF

ElementarySteps

Fragment

Figure 2. Typical schematic to build ‘catalysis by design’ mechanism. Version from the figure pub-

lished by Pelletier and Basset, in Acc Chem Res 2016; 49: 664–77.

catalysts for 1-octene epoxidation by tert-butylhydroperoxide.

Although [M]-Cl are not stricto sensureaction intermediates in CO2 reactionwith epoxides to give cyclic carbonates,the work below is a rare example of coop-erating surface bimetallic catalysis. CO2and epoxide are each activated by twoseparate Lewis acid centers (≡Si-O-NbCl4.OEt2) maintained in very close prox-imity by silica [19].

SOMC, alongside surface fragments,allow the prediction of catalysis by de-termination of the sequence of inter-mediates and to control the coordina-tion sphere of the metal to achievetargeted reactions. In SAC, the atomgrafted onto the surface depends on thereagents/substrates to adopt the right co-ordination sphere. The SOMC strategymay be seen as the result of a molecu-lar understanding of the elementary stepsnecessary to achieve a given reaction.One of the questions raised could bethe advantage of using SOMC ratherthan classical heterogeneous catalysis. Itis true that ‘classical catalysis’, mostly

based on a ‘trial and error approach’,could be considered as easier in termsof practical advantages, but the SOMCapproach offers several competitiveadvantages: prediction of new catalyticreactions, never observed in classicalheterogeneous catalysis; a reliable struc-ture activity relationship because weare dealing with well-defined structureswhere the physicochemical tools areused with maximum efficiency; andthe possibility to control the activityand selectivity with a careful choiceof ‘ligands’, ‘spectators ligands’ andsupports. This strategy progressivelyremoves the existing gap betweenheterogeneous and homogeneous catal-ysis mainly because the concepts ofmolecular chemistry (in particular theelementary steps) are easily applied toheterogeneous catalysis.

Jean-Marie Basset∗ and Jeremie D.A. PelletierKing Abdullah University of Science and

Technology, Saudi Arabia∗Corresponding author.E-mail: [email protected]

REFERENCES1. Pelletier JDA and Basset JM. Acc Chem Res

2016; 49: 664–77.2. Basset JM, Psaro R and Roberto D et al.

Modern Surface Organometallic Chemistry 2009,

23–135.

3. Dufaud VR and Basset JM. Angew Chem Int Ed

1998; 37: 806–10.4. Riache N, Dery A and Callens E et al.

Organometallics 2015; 34: 690–5.5. Corker J, Lefebvre F and Lecuyer C et al. Science

1996; 271: 966–9.6. Millot N, Soignier S and Santini CC et al. J Am

Chem Soc 2006; 128: 9361–70.7. Williams LA, Guo N and Motta A et al. Proc Natl

Acad Sci USA 2013; 110: 413–8.8. Thieuleux C, Maraval A and Veyre L et al. Angew

Chem Int Ed 2007; 46: 2288–90.9. Asakura K, Nishimura M and Iwasawa Y. J Mol

Catal 1989; 55: 159–69.10. Lesage A, Emsley L and ChabanasM et al. Angew

Chem Int Ed 2002; 41: 4535–8.11. Herisson JL and Chauvin Y. Makromol Chem

1971; 141: 161–76.12. Mazoyer E, Merle N and de Mallmann A et al.

Chem Commun 2010; 46: 8944–6.13. Basset JM, Coperet C and Soulivong D et al. Acc

Chem Res 2010; 43: 323–34.14. Blanc F, Coperet C and Thivolle-Cazat J et al.

Angew Chem Int Ed 2006; 45: 6201–3.15. Vidal V, Theolier A and ThivolleCazat J et al. Sci-

ence 1997; 276: 99–102.16. Hamzaoui B, El Eter M and Abou-Hamad E et al.

Chem Eur J 2015; 21: 4294–9.17. Hamzaoui B, Pelletier JDA and El EterM et al. Adv

Synth Catal 2015; 357: 3148–54.18. Hamzaoui B, Pelletier JDA and Abou-Hamad E

et al. Chem Commun 2016; 52: 4617–20.19. D’Elia V, Dong HL and Rossini AJ et al. J AmChem

Soc 2015; 137: 7728–39.

National Science Review

5: 633–635, 2018

doi: 10.1093/nsr/nwy069

Advance access publication 12 July 2018

Page 15: Special Topic: Single-Atom Catalysts - Western Engineering...Predictive approach of heterogeneous catalysis Metal-organic frame-work assisted synthesis of single-atom catalysts for

636 Natl Sci Rev, 2018, Vol. 5, No. 5 PERSPECTIVES

CHEMISTRY

Special Topic: Single-Atom Catalysts

Coordination chemistry of atomically dispersed catalystsPengxin Liu and Nanfeng Zheng ∗

The last decade has witnessed the rapiddevelopment of atomically dispersedcatalysts. With every active metal atomanchored on the surface of supports,atomically dispersed catalysts offer themaximum atom efficiency, helping tocreate cost-effective catalysts, particu-larly those based on earth-scarce metals.While it is still in intense debate whethersingle metal atoms can provide bettercatalytic activities than surface atomsof metal nanoparticles, an increasingnumber of studies have been revealingthat the subtle chemical environmentsurrounding the catalytic metal atomsplays a critical role in determining theoverall catalytic performance of anatomically dispersed catalyst. To makesingle-atom catalysts stable during cat-alytic reactions, metal atoms thereon areoften chemically anchored on the surfaceof supports through strong metal–metalbonds or coordination bonds with O,N, S or other atoms on supports. Thisperspective focuses on how the bindingsites from supports are involved in thecatalytic mechanisms of atomicallydispersed catalysts. As shown in Fig. 1a,similarly to the important roles of ligandsin homogenous catalysts, changing thebinding sites in the first coordinationshell surrounding the metal centers ofatomically dispersed catalysts is expectedto alter their electronic structures andthus catalytic properties. What is evenmore interesting is that, beyond thefirst coordination shell, metal cationson supports readily serve as the secondcoordination shell and are involved incatalytic reactions together with theprimary catalytic metal atoms.

With the same binding atoms on sup-ports, lowering the coordination num-ber would alter the electronic and geo-metric structure of single metal centers,thus changing the adsorption behavior ofreactants. For instance, in an atomically

dispersed Fe catalyst system, four Fe-Nx(x = 4–6) coordination environmentsgave different catalytic activities in the ox-idation of alkanes (Fig. 1b). The FeIIIN5structure showed over 1 order of mag-nitude more active than the high-spinand low-spin FeIIIN6 structures and threetimes more active than FeIIN4 in the se-lective oxidation of the C−H bond [1].In the electroreduction of CO2, Co-N-C

(a)

M’M’

M’M’ XX

Catalyticcenter

Support

X X

X

M

(b) TOF

Ph Pho

Fe-N-C

Mostactive site

N NNN

N

FeN NNN

Fe

N NNN

Y

X

Fe

Co-N4 Co-N2

(c)

Co Zn C N

Co-N2

Co-N3

Co-N4

Co NPsBackground

-0.8 -0.6 -0.4 -0.2 0E (V) vs. RHE

0

-15

-30

-45

j tota

l (m

A cm

-2) Co-N2

Co-N3

Co NPs

100

75

50

25

0

F.E

. (%

)

-0.8 -0.7 -0.6 -0.5E (V) vs. RHE

(d)

Pd1/TiO2-EG Ti(III)-O-Pd

Ti O

Pd

O Ti

O

O

O

Ti3+

O

OH2

Ti O PdCalcine

3+

H2

Pd O

Pd O

H Hδ

Pd O

Pd O

CD3OD

D

OO

OO

Pd1/TiO2-EG in hydrogenation

δH Hδδ

Pd1/TiO2-cal in oxidation

Ti O Pd4+O O

330 360 390 420Temperature (K)

Con

vers

ion

(%)

100

80

60

40

20

0

Pd1/TiO2-calPd1/TiO2-EG

Figure 1. Important roles of vicinal coordination environments in determining the catalysis of atom-ically dispersed metal catalysts. (a) Scheme of the first and second coordination shells of atomically

dispersed catalysts that determine the overall catalytic performances. The binding atoms and metal

cations from supports are located in the first and second shells, respectively. (b) Comparedwith other

structures, FeIIIN5 showed the highest activity in the selective oxidation of the C−H bond (adapted

with permission from Ref. [1], ACS). (c) Lower N-coordinated Co showed better performances in elec-

troreduction of CO2 (adapted with permission from Ref. [2], Wiley). (d) Two atomically dispersed Pd

catalysts showed distinct catalytic properties due to the different coordination structures of their

Pd atoms. Pd1/TiO2-EG activated H2 heterolytically with the assistance of EG ligands, showing a

homogeneous-like performance in hydrogenation (adapted with permission from Ref. [4], AAAS).

Pd1/TiO2-cal activated O2 into superoxide ions due to Ti(III)-O-Pd interfaces, boosting oxidation re-

actions of CO and VOCs (adapted with permission from Ref. [7], Elsevier).

catalysts with a lower Co-N coordinationnumber showedmuchhigher activity andselectivity to CO (Fig. 1c) [2]. In bothcases, the fabrication of active centers ofdifferent coordination environments wasachieved by thermal treatment at differ-ent temperatures. The delicate design ofsurface defects on carbon supports wasalso recently demonstrated as an effec-tive method to fabricate Ni single-atom

Page 16: Special Topic: Single-Atom Catalysts - Western Engineering...Predictive approach of heterogeneous catalysis Metal-organic frame-work assisted synthesis of single-atom catalysts for

PERSPECTIVES Liu and Zheng 637

catalysts with excellent electrocatalyticperformance in hydrogen evolution reac-tion under acidic conditions [3].

Besides the anionic components(e.g. O2−, N3−, S2−) of supports, organicligands on the surface of supports havepotentially important roles in shapingthe catalytic performances of atomi-cally dispersed catalysts. For instance,the surface EG (ethylene glycolate)ligands on the photochemically pre-pared Pd1/TiO2-EG catalyst played acrucial role in catalysing hydrogenationreactions [4]. As shown in Fig. 1d, theEG ligands not only served as bindingsites to anchor and thus stabilize Pdatoms onto the TiO2 support, but alsohelped to activate H2 heterolytically into‘H+–H−’ pairs. Such a unique activationpathway makes it possible to preciselycontrol the transfer of activated H atomsduring hydrogenation reactions. Such acontrol results in a perfect homogeneouscatalyst-like catalytic selectivity, whichis hardly achieved by other Pd-basedheterogeneous catalysts.

The second coordination shell ofatomically dispersed metals are metalcations from supports, which can alsomake a significant contribution to theiroverall catalytic performances throughelectronic effect and/or oxygen dona-tion. For example, both MgO and HYzeolite-supported single-site Ir complex[Ir(C2H4)2(acac)] have the identicalfirst-shell coordination environment.Each Ir is coordinated by two O atomsfrom the supports and two C2H4 lig-ands. However, the electron densityof zeolite-supported Ir was lower thanMgO-supported Ir due to the elec-tron withdrawal effect of Al in zeolite.Such an electron effect led to a muchhigher activity of the MgO-supportedIr catalyst than the zeolite-supportedone in the hydrogenation of ethylene[5]. In this case, local electron distri-bution of catalytic atoms depends onthe electron affinity of metal cations insupports.

When the metal cations in supportshave variable oxidation states, their sur-rounding O atoms can participate in re-actions. Thus, the reactivity depends onthe reducibility of the metal cations insupports. For example, a surface science

study on Au1/CuO demonstrated thatthe charge transfer from CuO to Aumakes the Au atoms negatively charged,making themactive forCOoxidation [6].Lattice O2− anions adjacent to the Auatoms can then reactwithCO to generateCO2 and O2− vacancies, leaving the Auatoms neutralized and inactive. O2− va-cancies can then be fixed by reacting withO2 to make up the complete catalytic cy-cle. Another nice example on the directinvolvement of metal cations in supportsis the vicinal effect for promoting oxida-tion catalysis of Pd1/TiO2-cal, whichwasmade from Pd1/TiO2-EG by removingEG through thermal treatment in air [7].Although Pd sites in both Pd1/TiO2-caland Pd1/TiO2-EG catalysts were presentas atomically dispersed species in the ox-idation state +2, they exhibited distinctcatalytic behaviors. While Pd1/TiO2-calexhibited a dramatically decreased activ-ity in hydrogenation catalysis, the cat-alyst displayed a significantly increasedactivity in catalytic CO oxidation, fivetimes higher than that by Pd1/TiO2-EG.As revealed by atomic-resolution elec-tron energy-loss spectroscopy (EELS),a well-defined atomic Ti(III)-Pd-O in-terface was created after the thermal re-moval of EG. As shown in Fig. 1d, theexposed Ti3+ sites facilitated the activa-tion of O2 into superoxide ions (O2

−)and thus suppressed the CO poisoningeffect, resulting in a significant enhance-ment in low-temperature CO oxidation.The direct involvement of metal cationson oxide supports in the catalysis sug-gests that the real active sites of atom-ically dispersed metal catalysts can befar beyond isolated metal atoms them-selves. Supports of atomically dispersedcatalysts play multiple roles than simplyserving as ligands.

The active involvement of a support’smetal cations that are vicinal to the pri-mary catalytic metal sites may answerthe long-debated issue of why atom-ically dispersed catalysts of the samemetal show different activities over dif-ferent supports. While Pt1/FeOx exhib-ited high activity inCOoxidation at 27◦C[8], Pt1/SiO2 and Pt1/HZSM-5 showedno activity below 100◦C [9]. Vapor-phase-synthesized Pt1/CeO2 started toshow activity over 150◦C [10]. More-

over, the preparation and applicationconditions (e.g. annealing temperature,pre-reduction) of atomically dispersedcatalysts are expected tohave a significantinfluence on the local coordination envi-ronments of their catalytic metal centers.

Although there have been a largenumber of reports showing the superiorcatalytic performances of atomically dis-persedmetal catalysts over their nanopar-ticulate counterparts, pictures of the co-ordination structures around catalyticmetal atoms are still blurry inmanypiecesof work. However, the atomic resolu-tion of the chemical structure of thedispersed metal atoms is crucial to de-code the catalytic mechanisms of atom-ically dispersed metal catalysts, whichis currently limited by the lack of ef-fective characterization techniques. To-gether with the development of atomic-resolution spectroscopy techniques (e.g.electron energy-loss spectroscopy) andhigh-resolution scanning tunneling mi-croscopy, creating atomically dispersedcatalysts that are easy to be character-ized is a promising solution. For in-stance, using ultrathin 1D or 2D nano-materials as supports for the synthesisof atomically dispersed metal catalystsnot only increases the loading amountof single atoms and thus promotes spec-troscopic signal–noise ratio, but also de-creases the background intensity in mi-croscopy. Meanwhile, the ultrathin fea-ture of the supports allows simplificationof the creation of structural models fortheoretical calculations to gainmolecularunderstanding on the structure-propertyrelationships that are crucial to the ratio-nal design of practical catalysts.

In summary, from the viewpoint of co-ordination chemistry, the supports notonly serve as ligands to stabilize singleatoms, but also help to tune the electronicand/or geometric structures of the atom-ically dispersed metal species. In somecases, metal cations from the supportin the second coordination shell of sin-gle atoms directly participate in the cat-alytic reactions as well. It is thus impor-tant to understand the catalytic mech-anism of atomically dispersed catalystsat the atomic level. We believe the fun-damental understanding on the coordi-nation chemistry of atomically dispersed

Page 17: Special Topic: Single-Atom Catalysts - Western Engineering...Predictive approach of heterogeneous catalysis Metal-organic frame-work assisted synthesis of single-atom catalysts for

638 Natl Sci Rev, 2018, Vol. 5, No. 5 PERSPECTIVES

catalysts will promote the developmentof the whole field and push them into in-dustrial applications.

FUNDINGThis work was supported by the National KeyR&D Program of China (2017YFA0207302) andthe National Natural Science Foundation of China(21731005, 21420102001). Dr. P. X. Liu thanksthe National Postdoctoral Program for InnovativeTalents (BX201600093) and the China Postdoc-toral ScienceFoundationProject (2017M610392).

Pengxin Liu and Nanfeng Zheng ∗

Collaborative Innovation Center of Chemistry for

Energy Materials, State Key Laboratory for

Physical Chemistry of Solid Surfaces, and

Department of Chemistry, College of Chemistry

and Chemical Engineering, Xiamen University,

China∗Corresponding author.E-mail: [email protected]

REFERENCES1. Liu W, Zhang L and Liu X et al. J Am Chem Soc

2017; 139: 10790–8.2. Wang X, Chen Z and Zhao X et al. Angew Chem

Int Ed 2018; 57: 1944–8.3. Zhang L, Jia Y and Gao G et al. Chem 2018; 4:285–97.

4. Liu PX, Zhao Y and Qin RX et al. Science 2016;

352: 797–800.

5. Lu J, Serna P and Aydin C et al. J Am Chem Soc

2011; 133: 16186–95.6. Zhou X, Shen Q and Yuan K et al. J Am Chem Soc

2018; 140: 554–7.7. Liu PX, Zhao Y and Qin RX et al. Sci Bull 2018; doi:

10.1016/j.scib.2018.03.002.

8. Qiao B, Wang A and Yang X et al. Nat Chem 2011;

3: 634–41.9. Ding K, Gulec A and Johnson AM et al. Science

2015; 350: 189–92.10. Jones J, Xiong H and DeLaRiva AT et al. Science

2016; 353: 150–4.

National Science Review

5: 636–638, 2018

doi: 10.1093/nsr/nwy051

Advance access publication 30 April 2018

CHEMISTRY

Special Topic: Single-Atom Catalysts

Theoretical understanding of the stability of single-atom catalystsJin-Cheng Liu1, Yan Tang1, Yang-Gang Wang1,2, Tao Zhang3 and Jun Li1,∗

As a new frontier, the rapid developmentof single-atom catalysts (SACs) in het-erogeneous catalysis has attracted exten-sive attention since the concept of single-atom catalysis was first coined in 2011[1,2]. Supported metal single atoms usu-ally possess unique chemical and phys-ical properties and have a special lo-cal chemical environment that is dis-tinctly different from conventional sup-ported nanoparticles and metal catalysts.Studies in the past several years haveshown four main advantages of SACs:high selectivity, possibly high stability,high atomic efficiency and tunable highactivity, as shown in Fig. 1. For efficiency,SAC maximizes the utilization of expen-sive metals by exposing each single metalatom to reactants. The local configura-tions of the active centers of an ideal SACcan be highly identical, leading to excel-lent selectivity compared to supportednanoparticles and metal surfaces that of-ten havemultiple yet rather diverse activesites.Unique coordinationof singlemetalatomswith neighboring atoms of support

may lead to high activity for specific reac-tions. All the above merits are on the ba-sis of the stability of catalysts that peopleare most concerned about. To avoid ag-gregation under catalytic reaction condi-tions, singlemetal atoms canbe stabilizedby anchoring at specific sites on the sup-port, including embedding and surface-adsorbing. Considering the complex re-alistic conditions, the stability of SACsmay also depend on the surface condi-tion, support type, reactant species, andfinite temperature and pressure, whichare difficult to ascertain and vary fromone system to another. It remains a grandchallenge to guide the prediction and de-sign of highly stable and reactive SACstoday.

Four types of single atoms on an oxidesurface are shown in Fig. 2b. The singleatomcanbe adsorbedonperfect ordefec-tive surfaces and can also be embedded(doped) into cation or oxygen vacancy.Most fabricated SACs from experimentsare the embedded type, with strong cova-lent metal–support interaction (CMSI),

while the sintering of single atoms anddispersion of nanoparticles should in-volve the diffusion of supported zerova-lent single atoms. Therefore, both thesupported andembedded single-atomac-tive centers are possible under realisticconditions and worth thorough investi-gation.

The intrinsic stability of SAC arisesfrom the support-assisted lower chemicalpotential when compared to nanopar-ticles. When the free-energy changefrom nanoparticles to single atoms isnegative, nanoparticles can be dispersedto single atoms spontaneously, whichleads to thermodynamic stability ofSACs (Fig. 2a, black curve). However,even if the free-energy change is pos-itive, single atoms can also be stablewhen the aggregation barrier is highenough to prevent sintering, which isthe kinetic stability of SACs (Fig. 2a, redcurve). Following the principles of theatomistic theory of Ostwald ripeningdeveloped by W-X Li and cowork-ers, we have achieved a quantitative

Page 18: Special Topic: Single-Atom Catalysts - Western Engineering...Predictive approach of heterogeneous catalysis Metal-organic frame-work assisted synthesis of single-atom catalysts for

PERSPECTIVES Liu et al. 639

Stability Selectivity effic

iency

Act

ivity

Pt1/CeO2 800°C

Au1/FeOx 400°C

SAC

Metal

Nanoparticle

Single atom

O O O

TiIII TiIV TiIV TiIV

PdCl2/TiO2

Pd

Cl Cl

H2C

HC

OH

Figure 1. Examples illustrating the advantages of single-atom catalysts (SACs) include activity, sta-

bility, selectivity and atomic efficiency. (a) High loadings of palladium atoms on TiO2 exhibit high

catalytic activity in hydrogenation of C= C bonds [3]. (b) Pt1/CeO2 and Au1/FeOx are stable at 800◦C

and 400◦C in oxidation conditions, respectively [4,5]. (c) Iron single atom embedded on a SiO2 sur-

face exhibits high selectivity for conversion of methane to ethylene, aromatics [6]. (d) SACs are less

expensive than supported nanoparticles and metal catalysts.

description and comparison of the twomodels by taking the reaction environ-ment, particle size and morphology, sup-port type anddefects, andmetal–reactantinteraction into account [7,8]. We haveproposed a complete theoretical modelto predict the chemical potential of sup-ported single metal atoms and supportednanoparticles, and taken the CO oxida-tion reaction as an example to guide thedesign of stable SACs. As a measure ofthe degree of change of free energy of asystem by adding 1 mole of metal atomsto the system at a given temperatureand pressure, the chemical potentials forSACs and metal particles were discussedin our recent work [8]. Both thermo-dynamic and kinetic criteria are consid-ered to determine the stability of SACs.The thermodynamic part includes (i) the

energetics of supported metal particles,which is based on the Gibbs−Thomson(G–T) relation with considering the ad-sorbed reactants and (ii) the chemicalpotential of monomers (both the metaladatoms and metal–reactant complexes)on supports.And the kinetic part includes(i) the diffusion barrier of monomers onperfect surfaces and defects and (ii) thebarrier of moving one metal atom froma supported metal nanoparticle to a sub-strate surface with corresponding sinter-ing rate equations.

Here, we discuss the support effecton the stability of supported metal atomsfirst using selected examples [9]. Wehave found that the quantum primogeniceffect plays a vital role in determining thevalence states and charge distributionof single-atom gold and the adsorption

mode of CO on various supports, asshown in Fig. 2d, which is consistentwith results from others [10]. Remark-ably, Au1 atoms are positively chargedon reductive supports (e.g. CeO2, TiO2)by charge transfer from metal adatomsto support cations, which leads to stronginteraction between CO’s 5σ orbital and6 s orbital of Au+, whereas Au1 keeps azero oxidation state on the irreduciblemetal dioxides (e.g. ZrO2, HfO2 andThO2) and has a weak interaction withCO via a bent Au–CO coordinationwith partial electron donation to theCO π∗ anti-bonding orbital. For thesame substrate, exposing different sur-faces also affects the relative stability ofsingle atoms. For example, the stabilitysequence of doped Pt single atoms ondifferent surfaces of CeO2 follows (110)> (100)> (111) [11].The high stabilityof Pt single atoms on (110) surfacebenefits from the spontaneous formationof an O2

2− species from two surfaceoxygen atoms that reduces PtIV to PtII. Adifferent anchoring site for a single atomin a given surface is another significantfactor of stability. It has been found thatthe presence of special defects such asvacancies or steps can improve the stabil-ity of an Au single atom on a CeO2(111)surface and the stability sequence iscalculated as cation vacancy > steps >

oxygen vacancy > perfect (defect-free)surface [8].

Besides the support effect, reactantspecies also change the stability ofSACs. It is now well recognized thatsintering is often accelerated by thepresence of reducing gases that converthigh- and medium-valent metal ions intozerovalent atoms, whereas nanoparticlescan also re-disperse to single atoms insome cases. Jones et al. reported that Ptnanoparticles can disperse onto ceria sur-faces under an oxygen atmosphere and800◦C [4]. Li and colleagues recentlyobserved noble metal nanoparticlestransforming to thermally stable singleatoms [12]. Parkinson et al. observedCO-induced coalescence of Pd adatomssupported on the Fe3O4(001) surface atroom temperature [13]. Those phenom-ena can also be attributed to the chemicalpotential of single atoms affected byreactants. As shown in Fig. 2c, reducing

Page 19: Special Topic: Single-Atom Catalysts - Western Engineering...Predictive approach of heterogeneous catalysis Metal-organic frame-work assisted synthesis of single-atom catalysts for

640 Natl Sci Rev, 2018, Vol. 5, No. 5 PERSPECTIVES

(a) (b)

(c)

(d) (e)

Figure 2. (a) Schematic illustration of free-energy diagram of sintering and dispersion processes

between Au NPs and SAs. (b) Four types of single atom on oxide support. (c) Aggregation and re-

dispersion processes of Pt1/FeOx under CO and O2 atmosphere. (d) Schematic representation of the

relationship between Hartree potential and charge transfer on reducible and irreducible substrates,

and the molecular orbital of Au+CO species on CeO2 support [9]. (e) Dynamic formation of single-

atom catalytic active sites on ceria-supported gold nanoparticles (adapted from [15]).

gases such as CO and H2 can react withthe lattice oxygen after adsorption ontosingle metal atoms, which breaks thecovalent bond between metal and oxy-gen, especially for doped atoms at cationvacancy. When all the relevant metal–support covalent bonds are broken byCO, the charge state of Pt changes frompositive to neutral. And, simultaneously,the chemical potential of single Pt atomsincreases to be higher than Pt nanopar-ticles. Such Pt0 species become highlymobile with CO adsorption and thus ag-gregated to particles rapidly by overcom-ing the diffusion barrier. On the contrary,when Pt nanoparticles are exposed tooxygen atmospheres at high temperature,the chemical potential of separated PtO2species can become lower than that onnanoparticles. Should there be enoughsuitable surface sites (such as cation va-cancies or steps on ceria) to trap PtO2, Ptnanoparticleswill transfer to single atomson support.

It is worth noting that single metalatomsmay also be dynamically generated

during the catalytic processes. Based onlarge-scale ab initio molecular dynamic(AIMD) simulations, we have foundthat, on reducible oxide-supported goldnanocatalysts [14,15], the gold cationcan migrate from the gold nanoparticleto support to catalyse CO oxidation andreintegrate back to the nanoparticle aftercompleting the reaction. This dynamicphenomenon of SACs is named as dy-namic single-atom catalysis [15]. Espe-cially, by a combination of ab initio elec-tronic structure and molecular dynamicssimulations, as well as amicrokinetic sim-ulation, we have shown that, on TiO2-supportedAunanocatalysts, formationofdynamic SACs is the dominant reactionpathway for CO oxidation under oxidiz-ing conditions and T < 400 K [14]. Thedynamic formation of single gold atomsunder realistic conditions is ultimately at-tributed to the reducibility of the oxidesupport that strongly couples with thecharge state of gold and makes the singlegold atom active during the catalytic pro-cesses, which is attributed to dynamic sta-

bility. In recent work, we found that dy-namic single atoms under reaction condi-tions account for the notorious size effectin gold nanocatalysts [16].

In summary, intrinsic thermodynamicstability, kinetic stability and dynamicstability are the key factors in deter-mining the reactivity of SACs. Althoughexperimental results present a compli-cated picture from system to system,it all complies with the rules of chem-ical potential, which can be quantita-tively evaluated from thermodynamicand dynamic aspects. Single-atom cataly-sis pushes the traditional catalytic theoryfrom a band-structure-based interpreta-tion to a more local chemical-bondinginterpretation, where the local chemi-cal coordination and local molecular or-bitals dominate the physical and chemi-cal properties, especially the stability andactivity. Exploring these intriguing fac-tors on stability of SACs based on the-oretical models [7,8] may lead to betterand deeper understanding of single-atomcatalysis and provide guidance for the ra-tional design of exactly controllable cat-alytic reactions.

Jin-Cheng Liu1, Yan Tang1, Yang-Gang Wang1,2,

Tao Zhang3 and Jun Li1,∗1Department of Chemistry and Key Laboratory of

Organic Optoelectronics & Molecular Engineering

of Ministry of Education, Tsinghua University,

China2Department of Chemistry, Southern University of

Science and Technology, China3Dalian Institute of Chemical Physics, Chinese

Academy of Sciences, China∗Corresponding author.E-mail: [email protected]

REFERENCES1. Qiao B, Wang A and Yang X et al. Nat Chem 2011;

3: 634–41.2. Yang X-F, Wang A and Qiao B et al. Acc Chem Res

2013; 46: 1740–8.3. Liu P, Zhao Y and Qin R et al. Science 2016; 352:797–800.

4. Jones J, Xiong H and DeLaRiva AT et al. Science

2016; 353: 150–4.5. Qiao B, Liang J-X and Wang A et al. Nano Res

2015; 8: 2913–24.6. Guo X, Fang G and Li G et al. Science 2014; 344:616–9.

Page 20: Special Topic: Single-Atom Catalysts - Western Engineering...Predictive approach of heterogeneous catalysis Metal-organic frame-work assisted synthesis of single-atom catalysts for

PERSPECTIVES Liu et al. 641

7. Ouyang R, Liu JX and LiWX. JAmChemSoc 2013;

135: 1760–71.8. Liu JC, Wang YG and Li J. J Am Chem Soc 2017;

139: 6190–9.9. Tang Y, Zhao S and Long B et al. J Phys Chem C

2016; 120: 17514–26.10. Bruix A, Lykhach Y and Matolinova I et al. Angew

Chem Int Ed 2014; 53: 10525–30.

11. Tang Y, Wang YG and Li J. J Phys Chem C 2017;

121: 11281–9.12. Wei S, Li A and Liu J-C et al. Nat Nanotech 2018;

doi: 10.1038/s41565-018-0197-9.

13. Parkinson GS, Novotny Z and Argentero G et al.

Nat Mater 2013; 12: 724–8.14. Wang Y-G, Cantu DC and Lee M-S et al. J Am

Chem Soc 2016; 138: 10467–76.

15. Wang YG, Mei D and Glezakou VA et al. Nat Com-

mun 2015; 6: 6511.16. He Y, Liu J-C and Luo L et al. Proc Natl Acad Sci

USA 2018; doi: 10.1073/pnas.1800262115.

National Science Review

5: 638–641, 2018

doi: 10.1093/nsr/nwy094

Advance access publication 5 September 2018

Page 21: Special Topic: Single-Atom Catalysts - Western Engineering...Predictive approach of heterogeneous catalysis Metal-organic frame-work assisted synthesis of single-atom catalysts for

RESEARCH ARTICLENational Science Review

5: 642–652, 2018

doi: 10.1093/nsr/nwy048

Advance access publication 17 April 2018

CHEMISTRY

Special Topic: Single-Atom Catalysts

Single-atom heterogeneous catalysts based on distinctcarbon nitride scaffoldsZupeng Chen1, Evgeniya Vorobyeva1, Sharon Mitchell1,∗, Edvin Fako2, Nuria Lopez2,Sean M. Collins3, Rowan K. Leary3, Paul A. Midgley3, Roland Hauert4 and

Javier Perez-Ramırez1,∗

1Institute for Chemical

and Bioengineering,

Department of

Chemistry and Applied

Biosciences, ETH

Zurich, 8093 Zurich,

Switzerland; 2Institute

of Chemical Research

of Catalonia (ICIQ),

The Barcelona

Institute of Science

and Technology,

43007 Tarragona,

Spain; 3Department of

Materials Science and

Metallurgy, University

of Cambridge,

Cambridge, CB3 0FS,

UK and 4Empa, Swiss

Federal Laboratories

for Materials Science

and Technology, 8600

Dubendorf,

Switzerland

∗Correspondingauthors. E-mails:[email protected].

ch; [email protected]

Received 5 February2018; Revised 10April 2018; Accepted13 April 2018

ABSTRACTCarbon nitrides integrating macroheterocycles offer unique potential as hosts for stabilizing metal atomsdue to their rich electronic structure. To date, only graphitic heptazine-based polymers have been studied.Here, we demonstrate that palladium atoms can be effectively isolated on other carbon nitride scaffoldsincluding linear melem oligomers and poly(triazine/heptazine imides). Increased metal uptake was linkedto the larger cavity size and the presence of chloride ions in the polyimide structures. Changing the hoststructure leads to significant variation in the average oxidation state of the metal, which can be tuned byexchange of the ionic species as evidenced by X-ray photoelectron spectroscopy and supported by densityfunctional theory. Evaluation in the semi-hydrogenation of 2-methyl-3-butyn-2-ol reveals an inversecorrelation between the activity and the degree of oxidation of palladium, with oligomers exhibiting thehighest activity.These findings provide newmechanistic insights into the influence of the carbon nitridestructure on metal stabilization.

Keywords: single-atom heterogeneous catalysts, carbon nitride scaffolds, alkyne semi-hydrogenation,density functional theory, metal–host interaction

INTRODUCTIONThe exploration of single-atom heterogeneouscatalysts (SACs) based on noble metals has beenstimulated by the prospect of improving metalutilization and selectivity simultaneously in sus-tainable catalytic processes [1–6]. Unfortunately,atomically dispersed metals on common hosts(e.g. metals, metal oxides and carbons) are oftenthermodynamically unstable and aggregate intoclusters or nanoparticles, especially at elevatedtemperature [7]. In this regard, graphitic carbonnitride (herein denoted as GCN) emerges as aunique host for preparing SACs due to the presenceof nitrogen-rich macroheterocycles in the lattice,which can anchor metal atoms firmly [8,9]. Thedensity of the adsorption pockets also helps tomaintain dispersion by configurational entropyconsiderations. In comparison to SACs supported

on other nitrogen-doped carbons, which typicallyexhibit significant structural heterogenity, the highercontent and uniform type and arrangement of nitro-gen species within GCN materials offer abundantand more precisely defined coordination sites.

Graphitic carbon nitride is regarded as themost stable polymorph upon polymerization ofcommon nitrogen-rich precursors (e.g. cyanamide,dicyanamide and melamine) under ambient condi-tions, and is widely used as a photocatalyst [10–14].In agreement with density functional theory (DFT)predictions of the higher thermodynamic stability,most experimental studies report the formationof heptazine- rather than triazine-based molecularstructures [8,9,15–19]. Since the preparation ofGCN is a stepwise polymerization process, variousintermediate phases including melam, melem andlinear melem oligomers (LMO) can be obtained by

C©TheAuthor(s) 2018. Published by Oxford University Press on behalf of China Science Publishing &Media Ltd. All rights reserved. For permissions, please e-mail:[email protected]

Page 22: Special Topic: Single-Atom Catalysts - Western Engineering...Predictive approach of heterogeneous catalysis Metal-organic frame-work assisted synthesis of single-atom catalysts for

RESEARCH ARTICLE Chen et al. 643

Figure 1. Idealized structure motifs and high-resolution TEM images of LMO, GCN, PTI and PHI. Sites that further polymerize in the extended structure

are indicated with parentheses. Color codes: gray, C; blue, N; white, H; pink, Li; purple, K; orange, Cl. The relaxed structures are shown in Supplementary

Fig. 16.

varying the synthesis conditions (temperature, pres-sure or atmosphere) [20,21]. On the other hand,highly crystalline carbon nitrides comprising or-dered poly(triazine imide) (PTI) or poly(heptazineimide) (PHI) can be assessed either by increasingthe temperature or pressure [22,23] or by ionother-mal synthesis employing eutectic salt mixtures ofLiX/KX (X = Cl or Br) [23–26] or simply singlealkaline metal chlorides [27,28] as the solvent.

Various metals (Pd, Ag, Pt or Ir) have beenstabilized as single atoms on GCN by using bothdirect (e.g. copolymerization, in-situ doping) andpost-synthetic (e.g. wet deposition optionally as-sisted by microwave irradiation and/or combinedwith chemical reduction) approaches [29–36]. Themethod of metal introduction is known to im-pact the distribution of metal centers within thehost, the post-synthetic deposition resulting inhigher surface metal densities and consequentlyincreased turnover frequencies in the three-phasesemi-hydrogenation of alkynes. Furthermore, boththe accessibility and electronic properties of ad-sorbed metal species could also be altered by vary-ing the morphology and porosity of GCN [33].By doping carbon into the lattice of GCN, we

recently reported the controlled variation of theC/N ratio, pointing out the potentially critical roleof the strength of the metal–host interaction [35].However, to date, the preparation of SACs based onother carbon nitrides including LMOand crystallinePTI and PHI phases has not been attempted.

To guide the design of improved SACs andgain insight into the effect of the host structure onmetal stabilization, a series of carbon nitride mate-rials (LMO, GCN, PTI and PHI) have been pre-pared (the idealized structure motifs are illustratedin Fig. 1).The comparative properties of the distinctscaffolds are studied in depth before and after the in-troduction of palladium via microwave-assisted de-position.The single-atom dispersion is confirmed inall cases by aberration-corrected scanning transmis-sion electron microscopy, while analysis by X-rayphotoelectron spectroscopy reveals significant vari-ation in the formal oxidation state of themetal. DFTcalculations are conducted to shed further light onthe interaction of the metal with the different scaf-folds, which is fundamental to understand the per-formance that in our case was interrogated throughthe semi-hydrogenation of 2-methyl-3-butyn-2-ol.The possibility to tune the electronic properties of

Page 23: Special Topic: Single-Atom Catalysts - Western Engineering...Predictive approach of heterogeneous catalysis Metal-organic frame-work assisted synthesis of single-atom catalysts for

644 Natl Sci Rev, 2018, Vol. 5, No. 5 RESEARCH ARTICLEIn

tens

ity (a

.u.)

2Theta (degrees)

(110

)

PHI

PTI

GCN

LMO

(002

)

(100

)

5 15 25 35 45 55

(a)

Inte

nsity

(a.u

.)

Wavenumbers (cm-1)

NHX CN

*

4000 3000 2000 1000

(b)

Nor

mal

ized

inte

nsity

(a.u

.)

Chemical shift (ppm)

*

CN3

CN2(NHX)(c)

200 175 150 125

Inte

nsity

(a.u

.)

(d)

Binding energy (eV)408 403 398 393

Figure 2. (a) XRD patterns, (b) DRIFTS spectra, (c) 13C CP/MAS NMR spectra and (d) N

1s XPS spectra of the investigated carbon nitride scaffolds. The sample color codes in

(a) apply to all panels. The transparent blue boxes indicate the reflections and stretch-

ing assignments, where the asterisks indicate bands characteristic of the PTI structure.

In (d), the black lines show the fitted result of the raw data (open symbols), whereas

the green, red and purple peaks corresponding to the deconvoluted C-N = C, NC3 and

-NHx components.

themetal via the exchange of intercalated ions in thepolyimides is also demonstrated.

RESULTS AND DISCUSSIONHost propertiesThe distinct carbon nitride scaffolds were preparedadapting previously reported protocols [22–25,37].In particular, LMO and GCN were obtained by di-rect polymerization of melamine at different tem-peratures, while PTI and PHI were synthesized by asimilar approach exploiting eutectic salt mixtures ofLiCl/KCl as the solvent. Analysis by X-ray diffrac-tion (XRD) confirms the characteristic crystallinestructures of the resultingmaterials (Fig. 2a). In par-ticular, GCN exhibits an intense reflection at 27.3◦

2θ (002) associated with the graphite-like interlayerstacking and a weak in-plane reflection stemmingfrom heptazine repeating units at 13.1◦ 2θ (100).PTI features a number of well-resolved reflections,which are consistent with the expected hexagonalstructure andP63cm space group [23].The strongestreflection at 26.8◦ 2θ indexed as the (002) planecorresponds to an interlayer distance of 0.33 nm,

whereas the (002) reflection of PHI was found tobe at 26.2◦ 2θ (0.32 nm). Note that the reflectionat 12.2◦ 2θ of PTI corresponding to the (100) in-plane periodicity shifts to a lower angle (8.3◦ 2θ) inPHI. In the case of LMO, the XRD pattern agreeswell with previously reported observations, wherethe (002) reflection at 25.5◦ 2θ features an interlayerdistance of 0.35 nm [37]. The high crystalline orderof the carbon nitride hosts was further evidenced byhigh-resolution transmission electron microscopy(TEM) imaging (Fig. 1). Although not observed inGCN due to the in-plane structural disorder andbeam sensitivity, LMO, PTI and PHI exhibited lat-tice fringes with spacings of 0.35, 0.33 and 0.32 nm,respectively, corresponding to the (002) planes inthese stacked aromatic structures. Additional latticefringes with distances of 0.44 and 0.74 nm were alsoobserved in the case of PTI (Supplementary Fig. 1),which canbe assigned to the (110) and(100)planes,respectively.

The distinct structures were further corrobo-rated by diffuse reflectance infrared Fourier trans-form spectroscopy (DRIFTS), 13C solid-state cross-polarization/magic angle spinning nuclear mag-netic resonance (CP/MAS NMR) spectroscopyand X-ray photoelectron spectroscopy (XPS). TheDRIFTS spectra (Fig. 2b) evidence the existenceof the aromatic heterocycles in all hosts, showingthe stretching at 1100–1650 cm–1, while the broadbands at 3000–3300 cm–1 are assigned to the bend-ing of -NHx terminations [38]. Though bearingthe same building units of heptazine, LMO showsmore intense -NHx breathing modes than GCN, in-dicative of more peripheral -NHx terminations inLMO. The cumulated double bonds (-N=C=N-)at 2184 cm–1 are obvious in PTI and PHI, evidenc-ing the presence of -NH- bridges as in ketene imines[39]. In addition to the deformation vibrations ofthe triazine or heptazine rings at 814 cm–1, a uniqueband at 670 cm–1 in PTI suggests that triazine ringsare the building units [23,40], instead of the hep-tazine in the other cases. LMO and GCN show sim-ilar 13CNMR spectra (Fig. 2c) with twomain peaksat 164 and 155–157 ppm, attributed to CN2(NHx)and CN3 moieties, respectively. These signals arealso present in PTI and PHI, but the ratio betweenthe intensity at 164 and 157 ppm is much higher,demonstrating more carbon species close to periph-ery -NHx. On the other hand, an additional peakat 168 ppm indicates the presence of triazine ringsin PTI, which can be ascribed to the carbons withfew protons in its proximity [41]. Furthermore, themain contribution at 288.3 eV in the C 1sXPS spec-tra (Supplementary Fig. 2) originates from the car-bon species in the triazine or heptazine rings. Com-paratively, deconvolution of N 1s spectra (Fig. 2d)

Page 24: Special Topic: Single-Atom Catalysts - Western Engineering...Predictive approach of heterogeneous catalysis Metal-organic frame-work assisted synthesis of single-atom catalysts for

RESEARCH ARTICLE Chen et al. 645

Table 1. Characterization data of the carbon nitride scaffolds and associated SACs.

Hosta FormulabSBETc

(m2 g−1)Pdb

(wt.%)

Loadingefficiencyd

(%)

Pd surfacedensitye

(μmolPd m−2)

LMO C3N4.86H2.69O0.11 3 (29) 0.66 33 8.6GCN C3N4.64H1.59O0.10 8 (11) 0.58 29 63.8PTI C3N4.52H2.89O0.82Li0.16K0.05Cl0.12 68 (76) 0.56 100 1.7PHI C3N4.27H3.11O1.43Li0.04K0.27Cl0.01 31 (30) 0.47 95 4.7PTI-Mg C3N4.55H2.96O0.84Li0.15K0.01Cl0.08Mg0.02 80 (110) 0.50 100 1.6PHI-Mg C3N4.36H3.38O1.49Li0.00K0.09Cl0.01Mg0.10 31 (40) 0.48 96 4.2

aLMO, linearmelem oligomer; GCN, graphitic carbon nitride; PTI, poly(triazine imide); PHI, poly(heptazine imides); PTI-Mg and PHI-Mg,magnesiumion-exchanged PTI and PHI.bDetermined by elemental analysis (non-metals) or ICP-OES (metals).cBETmethod (in parentheses, the surface area of the SACs).dDetermined by 100× (actual metal content/targeted metal content).eDetermined from the surface Pd concentration (from XPS) and area (from gas sorption) of the SACs.

of LMO, GCN and PHI evidence three main peaksat 398.8, 399.9 and 401.0 eV, which can be as-cribed to the ringnitrogen (C-N=C), tertiary nitro-gen (NC3) and terminal -NHx groups. The ratio ofC-N=C/NC3 was calculated to be around 6, con-firming the presence of heptazine as the buildingunit [25]. Taking account of the relative nitrogencontent from elemental analysis (Table 1), the sur-face NHx concentration for LMO was calculatedto be 6.8 mmol g−1, which is 1.3-fold more thanthat of GCN (5.3 mmol g−1). On the other hand,the absence of the tertiary nitrogen (NC3) and theC-N=C/NC3 ratio of around 2, accompanied bysome shift in the terminal -NHx, suggest again thatthe obtained PTI is built of triazine instead of hep-tazine units.

The chemical composition of the hosts was de-termined by elemental analysis and inductively cou-pled plasma-optical emission spectrometry (ICP-OES) (Table 1). The C/N molar ratio of LMOis 0.62, which is lower than its GCN counterpart(0.65), likely due to the abundant -NHx termina-tions in LMO. Meanwhile, the C/N ratios of PTIand PHI are 0.66 and 0.70, respectively, which arevery close to the theoretical values (0.67 for PTIand 0.71 for PHI). The Li, K, Cl molar contents(in mol.%) in PTI and PHI (PTI/PHI) were cal-culated to be 2.3/0.3, 0.5/2.2 and 1.0/0.1, respec-tively. Therefore, it can be concluded that PTI isintercalated by Li+ and Cl– simultaneously, whilePHI is preferentially intercalated with K+. The pres-ence of exchangeable ions in PTI and PHI offersa further possibility to tune the electronic proper-ties of the hosts. Magnesium was chosen as the ex-changed species, since Mg2+ is known to interactwell with the N species in porphyrins and it has acomparable or smaller ionic radius (rion = 86 pm)than that of the Li+ (90 pm) and K+ (152 pm)ions initially present in the structures. As shown in

Table 1, only 0.41 wt.% Mg was exchanged into theframework of PTI, while the Mg content can be in-troduced into PHI up to 1.94 wt.%, indicating thatK+ in PHI can be efficiently exchanged with Mg2+.The structure of PTI and PHI remains unchangeduponMg2+ exchange, as suggested by the similarityof the XRD patterns (Supplementary Fig. 3).

The distinct morphology of the applied hostswas visualized by TEM and scanning electron mi-croscopy (SEM) (Supplementary Fig. 4). In con-trast to the irregularly shaped LMO and GCN, PTIdisplays fiber-like morphology constituted by cu-bic and hexagonal nanocrystals, while PHI presentsa mixture of rod-like structures and plates. As evi-denced by argon sorption (Supplementary Fig. 2),PTI and PHI exhibit higher surface areas (68 and31 m2 g–1, respectively) than non-porous LMO andGCN (3 and 8m2 g–1, respectively), which is linkedto the nanostructured characteristics of the poly-imide structures.

Metal stabilizationTo study the capacity of the distinct carbon nitridescaffolds as hosts for single atoms, palladium wasintroduced via a microwave-assisted deposition tar-geting a loading of 2 wt.%. Increased metal depo-sition was observed for the polyimide structures,taking up 66% (PTI) and 68% (PHI) of the avail-able metal leading to palladium contents of 1.32and 1.35 wt.%, respectively. Comparatively, loweruptakes 33% (LMO) and 29% (GCN) were ob-served for the other carbon nitrides (incorporating0.66 and 0.58 wt.%, respectively). This is tentativelyattributed to the specific binding of palladium bythe former carriers (vide infra) although the highersurface area of these materials could also enhancethe capacity as a metal host. For an improved cat-alytic evaluation, two additional SACs based on PTI

Page 25: Special Topic: Single-Atom Catalysts - Western Engineering...Predictive approach of heterogeneous catalysis Metal-organic frame-work assisted synthesis of single-atom catalysts for

646 Natl Sci Rev, 2018, Vol. 5, No. 5 RESEARCH ARTICLE

Pd0

Pd2+

Pd4+

Pd-PHI

Pd-PTI

Pd-GCN

Pd-LMONor

mal

ized

inte

nsity

(a.u

.)

Binding energy (eV)348 344 340 336 332

50 nm1 nm 1 nm 1 nm 1 nm

1 nm 1 nm 1 nm

Fresh

Used

Pd-LMO Pd-GCN Pd-PTI

1 nm

Pd-PHI

Figure 3. AC-HAADF-STEM images of the fresh and used Pd-SACs based on different carbon nitride scaffolds and Pd 3d core-

level XPS spectra of the fresh catalysts. Some isolated Pd atoms are identified by yellow circles. Additional low-magnification

HAADF-STEM images of the fresh samples are shown in Supplementary Fig. 17. In the XPS spectra, the black lines show

the fitted result of the raw data (open symbols), whereas the magenta and green peaks correspond to the deconvoluted

components. The dashed lines indicate the positions formally assigned to Pd4+, Pd2+ and Pd0 species.

and PHI were prepared with metal contents close to0.5 wt.% (Table 1).

The examination by aberration-correctedhigh-angle annular dark-field scanning transmis-sion electron microscopy (AC-HAADF-STEM)verified the single-atom dispersion of palladium,where the higher-atomic-number metal atoms arevisible as the sub-nanometer bright spots on asmoothly varying gray background signal from thelower-atomic-number hosts (Fig. 3 and Supple-mentary Fig. 5). To assess the macroscopic metaldistribution, thin cross-sections of the embeddedmaterials were mapped by energy-dispersive X-ray(EDX) spectroscopy, indicating a relatively uniformpresence of palladium throughout LMO, PTI andPHI (Supplementary Fig. 6). In contrast, a surfaceenrichment in the concentration of palladium wasobserved for GCN, suggesting that the metal isunable to penetrate deeply into the material.

The influence of the host structure on the elec-tronic properties of palladium was studied by XPS.The presence of two different oxidation states wasclearly distinguishable from the Pd 3d core-levelspectra (Fig. 3), at around 338.3 and 336.5 eV, re-spectively. Based on formal assignments, these peakscan be attributed to Pd4+ (338.3 eV) and Pd2+

(336.5 eV). Notably, no signal corresponding tothe metallic Pd fingerprint (appearing at 334.9 eV)was detected in any of the catalysts. These observa-tions are consistent with the expected strong inter-action between the isolated atoms and hosts and theabsence of nanoparticles evidenced by microscopy.Notice that the direct assignment of formal chargesis debatable, and the values above are only used forreference purposes indicating the existence of palla-

dium species with differing degrees of oxidation orcoordination to atoms with different electronegativ-ity. Significant variation of the ratio of Pd2+/Pd4+

was observed, ranging from 0.19 (Pd-PHI) and 0.31(Pd-PTI) to 1.04 (Pd-GCN) and 1.32 (Pd-LMO),indicating that the strength of the metal–host inter-action can be manipulated depending on the frame-work structure of carbon nitrides.The relatively highcontribution of Pd4+ in PTI and PHI is rationalizedby the possible interaction of palladium with inter-calated ionic species in these materials (vide infra).Toassess the chemical state of sub-surfacepalladiumatoms in the SACs, a depth-profiling analysis byXPScoupledwithAr+ beametchingwas conducted to re-move the surface layer (Supplementary Fig. 7). In allcases, the Pd 3d core-level spectra are slightly shiftedto higher binding energies, showing an evolution to-wards Pd4+.The charge assignment is done by com-parison with standards and thus they might differfrom the charges obtained in the DFT calculations.

To gain insight into the relative thermal stabil-ity, the distinct Pd-SACs were treated both in airat 673 K and in a flowing 5%H2/He mixture at433 K, the latter conditions representative of thosetypically employed in gas-phase hydrogenation re-actions [32,42]. The Pd atoms over LMO, PTI andPHI exhibit high resistance to sintering with no signof nanoparticle formation (Supplementary Fig. 8),while abundant Pd clusters become visible at thesurface of GCN. The lower stability of single atomsover GCN can be due to several reasons. Compara-tively, Pd atoms bind weakly to GCN, as evidencedby the lowest average oxidation state of themetal ob-served by XPS and supported by DFT calculations(vide infra). In addition, the surface density of metal

Page 26: Special Topic: Single-Atom Catalysts - Western Engineering...Predictive approach of heterogeneous catalysis Metal-organic frame-work assisted synthesis of single-atom catalysts for

RESEARCH ARTICLE Chen et al. 647

Surface Pd Coordinated PdCl2

(-1.73 eV; 355.50 eV) (-1.91 eV; 356.80 eV)

(-1.47 eV; 356.45 eV) (-3.19 eV; 356.58 eV)

(-1.76 eV; 356.17 eV) (-2.00 eV; 357.44 eV)

(-2.43 eV; 357.95 eV)

(-2.78 eV; 356.43 eV)

(-2.75 eV; 356.06 eV)

Sub-surface Pd

PTI-9N

PHI-15N

GCN-6N

Figure 4. Optimized Pd coordination sites within different carbon nitride scaffolds with6 N, 9 N and 15 N pockets. Values in parentheses beneath each image indicate the

corresponding formation energies (left), calculated versus an isolated Pd atom or PdCl2coordination and the relaxed scaffold (the chlorinated system contains Mg2+ counter

cations between planes) and the calculated Pd 3d XPS assignments (right), respectively.

The top view representations are shown in Supplementary Fig. 18. Coordinated PdCl2within GCN was calculated for reference purposes despite the absence of Cl− in GCN.

Color codes: gray, C; blue, N; white, H; green, Pd; orange, Cl.

atoms in the case ofGCNis significantly higher com-pared to the other hosts (Table 1), which couldalso contribute to their lower stability under thermaltreatment. No reflections associated with Pd phasesare observed in the XRD patterns (SupplementaryFig. 9) after calcination, in agreement with the pre-served high dispersion in all cases. Note that, in thecase of LMO, the appearance of a reflection at 6.3◦

2θ and the merging of the reflections at around 12.6and 27.4◦ 2θ after calcination indicate the likely for-mation of a new layered complex, which could havea similar structure to previously reported complexesbetween melamine and cyanuric acid [43]. On theother hand, the disappearance of the in-plane reflec-tion peak at 8.3◦ suggests a lower stability in the caseof PHI.

DFT calculationsDFT has been employed to gain insight into thestabilization of palladium in the distinct carbon ni-tride scaffolds and their speciation in terms of Badercharge andXPS shift (Fig. 4, Supplementary Table 2and Supplementary Discussion). The optimized lat-

tices present cavitieswith different numbers of nitro-gen atoms (denoted as 6 N, 9 N and 15 N pocketsfor GCN, PTI and PHI, respectively).The small 6 Ncavity in GCN can efficiently stabilize Pd in the cen-ter of the pocket (Pd-6 N) or alternatively by coor-dinating with four N atoms (Pd-4 N) between twoneighboring planes. In the case of the larger cavities(9N and 15N), Pd is found to adsorb close to twoNcenters. A consistent range of Pd3d shiftwas noticedduring the simulation of the XPS profiles. In general,the Pd atoms preferentially reside at the subsurfacelayer of carbon nitrides, particularly for PTI. For allscaffolds, the surface Pd appears to be the least oxi-dized (Pd2+) and they are suggested to be more ac-tive than their more oxidized counterparts (Pd4+)that are buried more deeply in the material (espe-cially in the case of PTI and PHI) and therefore lessaccessible, which is consistent with the XPS depth-profiling analysis. In the presence of chlorine, palla-dium can coordinate in the form of PdN2Cl2 that isalso more oxidized than their surface counterparts.The PdCl2 are less likely in the smallest cavities, asthey imply a larger perturbation in the scaffold. Thepresence of Cl− negatively affects the activity, sinceit lowers the d-band position of the Pd levels (Sup-plementary Table 2) and it would be necessary tocleave Pd-Cl bonds (energetically more favorablethan cleavage of Pd-N bonds) to adsorb reactants.Furthermore, the incorporation of Mg2+ in the sub-surface vacancy of PTI and PHI steadily propels thePd species to the surface, which would be expectedto positively influence the activity. Ab-initiomolecu-lar dynamics (MD) simulations were conducted onPd-GCN, highlighting the high stability of this sys-tem.Upon increasing the temperature, the Pd atomsare observed to fluctuate between the surface (lessoxidized) and subsurface (more oxidized) configu-rations (Supplementary Fig. 10).The stability of thePd-SACs was further assessed in the presence of O2or H2, sampling the potential energy surface withrelevant intermediates (Supplementary Fig. 11). Inall cases, the adsorption of oxygen is found to havea stabilizing effect. Two distinct scenarios are ob-served for the activation of hydrogen, which is foundto occur homolytically, resulting in improved sta-bility of the single atoms on PTI and PHI, but oc-curs heterolytically, slightlyweakening the coordina-tion of Pd (by 0.24 eV), on GCN with smaller cav-ity. The adsorption energy of 2-methyl-3-butyn-2-ol was probed, confirming the preferred interactionwith surface sites (Supplementary Table 3).

Alkyne semi-hydrogenation performanceThe catalytic performance was evaluated in thesemi-hydrogenation of 2-methyl-3-butyn-2-ol,

Page 27: Special Topic: Single-Atom Catalysts - Western Engineering...Predictive approach of heterogeneous catalysis Metal-organic frame-work assisted synthesis of single-atom catalysts for

648 Natl Sci Rev, 2018, Vol. 5, No. 5 RESEARCH ARTICLE

Sel

ectiv

ity (%

)

Con

vers

ion

(%)

(a)100

80

60

40

20

0

100

80

60

40

20

0Pd-LMO Pd-PTI Pd-PHIPd-GCN

Pd average oxidation state (-)

(b)

100

0

200

300

2. 6 2.8 3.0 3.2 3.4 3.6 3.8

r (m

olal

keno

l mol

Pd-1

h-1) LMO

GCN

GCN-DCDA

MCNPTI-Mg

PHI

PTIOMCN

PHI-Mg

Pd NPs-MCN

ECN

(c)

Time-on-stream (h)

r/r0 (

-)

0 1 2 3 4 50.0

0.5

1.0

1.5

2.0Pd-LMO

Pd-PTIPd-PHI

Pd-GCN

Figure 5. (a) Conversion and selectivity to 2-methyl-3-buten-2-ol in the hydrogenationof 2-methyl-3-butyn-2-ol of Pd-SACs based on carbon nitride scaffolds of different lat-

tice structure. (b) Correlation of the rate of 2-methyl-3-buten-2-ol formation with the

palladium average oxidation state. This trend was generalized by considering addi-

tional Pd-SACs based on carbon nitrides with different morphology and ion-exchanged

form, and the nanoparticle-based catalyst on a mesoporous carbon nitride host. The

metal loading was ∼0.5 wt.% in all cases. Characterization data of the additional

samples is shown in Supplementary Table 1. (c) The relative rate (r/r0) as a function

of time-on-stream over Pd-SACs based on different carbon nitride scaffolds.

which is an important reaction in fine-chemicalmanufacturing [44]. Despite comparable metalcontents of the examined SACs, Pd-LMO exhibitsa significantly higher conversion (52%) than othercatalysts (Pd-GCN, 33%; Pd-PTI, 11%; Pd-PHI,7%) (Fig. 5a). Under the conditions investigated,the selectivity towards 2-methyl-3-buten-2-olapproaches 100% over all samples, evidencing thehigh chemoselectivity of isolated single atoms. Aspresented in Fig. 5b, the rate of alkenol (2-methyl-3-buten-2-ol) formation is well correlated with theaverage oxidation state of palladium (Pdavg). For in-stance, the rate over Pd-LMO with the lowest Pdavgof 2.86 reaches 311 molalkenol molPd–1 h–1, which ismore than five times higher than that observed overPd-PHI (57 molalkenol molPd–1 h–1) with a Pdavg of3.68. These findings suggest the critical role of tun-ing the electronic properties of the host structure intailoring the strength of metal–host interaction.Thestability of the SACs was further evaluated incontinuous mode in order to exclude the effectsof deactivation due to Pd leaching or aggregation.Importantly, all Pd-SACs display a constant ratetowards 2-methyl-3-buten-2-ol formation for 5 h onstream with no variation in conversion or selectivity(Fig. 5c). As additional references, four SACsbased on GCN were specifically prepared fromdifferent precursor (dicyandiamide (DCDA)) anddifferent morphology (exfoliated, mesoporousand ordered mesoporous carbon nitride (denotedas ECN, MCN and OMCN)) (SupplementaryTable 1). Impressively, the rates towards 2-methyl-3-buten-2-ol also fall in the same correlation,despite presenting distinct morphology andsingle-atom distribution. Although it cannot beassessed by standard techniques, the abundant-NHx terminations over the oligomer were sug-gested to improve the accessibility of the activePd centers. Analysis of the used catalysts con-firms the virtually identical atomic dispersion(Fig. 3), electronic properties (SupplementaryFig. 12 and Supplementary Table 4) and crystallinestructure (Supplementary Fig. 13) compared tothe fresh materials, verifying the stability of theSACs. For reference, the traditional catalyst for theliquid-phase selective hydrogenation of alkynesbased on supported lead-modified palladiumnanoparticles (Lindlar catalyst, 5 wt.% Pd-3 wt.%Pb/CaCO3) was evaluated as a benchmark. Whilethis catalyst yields a slightly higher rate towards 2-methyl-3-buten-2-ol (517 molalkenol molPd−1 h−1),it exhibits significantly reduced selectivity (78% to2-methyl-3-buten-2-ol) due to over-hydrogenation(22% selectivity to 2-methyl-3-butan-2-ol). Thisfurther highlights the superior performance of the

Page 28: Special Topic: Single-Atom Catalysts - Western Engineering...Predictive approach of heterogeneous catalysis Metal-organic frame-work assisted synthesis of single-atom catalysts for

RESEARCH ARTICLE Chen et al. 649

SACs (>95% selectivity to 2-methyl-3-buten-2-ol).To address the impact of magnesium incorporationon the catalytic performance, ∼0.5 wt.% Pd wasintroduced into PTI-Mg and PHI-Mg (Table 1,Supplementary Fig. 3 and Supplementary Fig. 14).The alkenol formation rates over Pd-PTI-Mg andPd-PHI-Mg are 102 and 94 molalkenol molPd–1 h–1,respectively, which is 1.2 and 1.6 times bettercompared to those without Mg2+ incorporation,while the selectivity towards 2-methyl-3-buten-2-olis 100% all samples. Again, the reaction rates alsocorrelate with the palladium average oxidation state(Fig. 5b), which further demonstrates the impactof tuning metal–host interaction by tailoring theelectronic properties of carbon nitride scaffolds. Forcomparative purposes, a nanoparticle-containingcatalyst based on a mesoporous carbon nitridehost (Pd NPs-MCN, with average Pd oxidationstate of 2.7; see STEM image in SupplementaryFig. 15) was prepared. Evaluation of this materialevidenced a lower rate towards 2-methyl-3-buten-2-ol than the expected trend (Fig. 5), which isconsistent with the different metal speciation in thiscatalyst.

CONCLUSIONSThis study demonstrated the obtainment of SACson three previously unreported carbon nitride scaf-folds, namely linear melem oligomers poly(triazineimides) and poly(heptazine imides).The larger cav-ity size and presence of chloride ions in the poly-imide structures facilitate the accommodation ofpalladium and enhance the metal–host interactionand thus higher resistance to sintering. An inversecorrelation is observedbetween the activity for semi-hydrogenation of 2-methyl-3-butyn-2-ol and the de-gree of oxidation of palladium, where the oligomersexhibit the highest activity. This was further gener-alized over additional previously reported materials,highlighting the critical importance of controllingthe oxidation state of isolated metal atoms. The in-tercalated alkaline metals within the network of PTIand PHI were demonstrated to be able to exchangewith other ions such as magnesium, presenting an-other opportunity to tune the catalytic performancein hydrogenation. The least oxidized surface Pd2+

species are suggested to be more active than thePd4+ species that reside deeper in thematerial; how-ever, this highpositive charge can also appear if someligand remains as PdN2Cl2 coordination appears.The findings provide an opportunity to systemicallydesign effective SACs to boost the atom efficiencyat an atomic level by constructing the host latticestructure.

METHODSCarbon nitride synthesisLMO and polymeric GCN were prepared by cal-cining melamine (8 g) at the desired temperature(723 K for LMO and 823 K for GCN with a ramprate of 2.3 K min−1) in a crucible for 4 h under anitrogen flow (15 cm3 min−1). PTI and PHI wereprepared by ball milling eutectic salt mixtures ofLiCl (4.52 g)/KCl (5.48 g) together with a corre-sponding precursor (melamine (1 g) for PTI and3-amino-1,2,4-triazole-5-thiol (2 g) for PHI) for10 min, in a Retsch PM 100 bioMETA planetaryball mill (500 rpm). Afterwards, the mixtures weretransferred into a crucible and calcined at 823 Kfor 4 h (ramp rate, 2.3 K min−1) under a nitrogenflow (15 cm3 min−1). The resulting products werewashed with hot water for 48 h to remove any excesssalts. Finally, the carbon nitride products were col-lected by filtration, washed thoroughly with distilledwater and ethanol, and dried at 338 K overnight.

Metal introduction bymicrowave-assisted depositionDifferent carbon nitride hosts (0.5 g) was first dis-persed in H2O (20 cm3) under sonication for 1 h.Then, an aqueous solution of Pd(NH3)4(NO3)2containing 5 wt.% Pd (0.05 cm3, targeting 0.5 wt.%;0.2 cm3, targeting 2 wt.%) was added and stirredovernight for complete adsorption. The result-ing solution was placed in a microwave reactor(CEM Discover SP), applying a cyclic program (20repetitions)of irradiation (15 s) and cooling (3min)using a power of 100 W. The resulting powder wascollected by filtration, washed with distilled waterand ethanol, and dried at 333 K overnight.

Ion exchangeIn a typical synthesis, PTI or PHI (0.7 g) wasdispersed in an aqueous solution of MgCl2·6H2O(3.4 g, 20 cm3) and then stirred at room temperaturefor 24 h, after which the solids were collected by cen-trifugation. These steps were repeated three timesand the products were thoroughly washed with wa-ter and ethanol, and subsequently dried at 338 Kovernight. The samples are denoted as PTI-Mg andPHI-Mg.

CharacterizationXRD was performed in a PANalytical X’Pert PRO-MPD diffractometer operated in Bragg-Brentanogeometry using Ni-filtered Cu Kα (λ = 0.1541 nm)

Page 29: Special Topic: Single-Atom Catalysts - Western Engineering...Predictive approach of heterogeneous catalysis Metal-organic frame-work assisted synthesis of single-atom catalysts for

650 Natl Sci Rev, 2018, Vol. 5, No. 5 RESEARCH ARTICLE

radiation. Data were recorded in the range of 5–60◦

2θ with an angular step size of 0.05◦ and a countingtime of 2 s per step. DRIFTS was performed usinga Bruker Optics Vertex 70 spectrometer equippedwith a high-temperature DRIFT cell (Harrick) andan MCT detector. The samples were pretreatedat 423 K for 1 h under Ar before analysis. Spectrawere recorded in the range of 4000–400 cm−1

under Ar flow (20 cm3 min–1) and at room tem-perature by co-addition of 64 scans with a nominalresolution of 4 cm−1. The 13C solid-state cross-polarization/magic angle spinning nuclearmagneticresonance (CP/MAS NMR) spectra were recordedon a Bruker AVANCE III HD NMR spectrometerat a magnetic field of 16.4 T corresponding to a1H Larmor frequency of 700.13 MHz. A 4-mmdouble resonance probe head at a spinning speedof 10 kHz was used for all experiments. The 13Cspectra were acquired using a cross-polarizationexperiment with a contact time of 2ms and a recycledelay of 1 s. A total of 64 × 103 scans were addedfor each sample. Between 39 × 103 and 96 × 103

scans were acquired, depending on the sample. The13C experiments used high-power 1H decouplingduring acquisition using a SPINAL-64 sequence.X-ray photoelectron spectroscopy (XPS) wasperformed in a Physical Electronics InstrumentsQuantum 2000 spectrometer using monochromaticAl Kα radiation generated from an electron beamoperated at 15 kV and 32.3 W. The spectra werecollected under ultra-high vacuum conditions(residual pressure = 5 × 10−8 Pa) at a pass energyof 46.95 eV. All spectra were referenced to theC 1s peak of ternary carbon at 288.3 eV. Prior topeak deconvolution, X-ray satellites and inelasticbackground (Shirley type) were subtracted for allthe spectra. The Pd average oxidation state wascalculated based on the relative content determinedby the peak area of different palladium species fromPd3d core-level XPS spectra. Elemental analysis wasdetermined by infrared spectroscopy using a LECOCHN-900 combustion furnace. Inductively coupledplasma-optical emission spectrometry (ICP-OES)was conducted using a Horiba Ultra 2 instrumentequipped with photomultiplier tube detection. Thesamples were dissolved in a piranha solution andleft under sonication until the absence of visiblesolids in the solution. SEM images were acquiredusing a Zeiss ULTRA 55 operated at 5 kV. Argonsorption was measured at 77 K in a Micrometrics3Flex instrument, after evacuation of the samplesat 423 K for 10 h. The specific surface area wasdetermined via the Brunauer-Emmett-Teller (BET)method. Samples for TEM studies were preparedby dusting respective powders onto lacey-carbon-coated copper or nickel grids. High-resolution

TEM, conventional STEM, and energy-dispersiveX-ray spectroscopy (EDX) measurements wereperformed on aTalos F200X instrument operated at200 kV and equipped with an FEI SuperX detector.Thincross-sectionswerepreparedbyembedding thepowder in a suitable resin (Polysciences Inc., HardGrade) followed by cutting the sections (100 nm)with a diamond knife. The sections were mountedon carbon-coated copper grids. AC-HAADF-STEM was performed using an FEI Titan3 80–300(Thermo Fisher Scientific) microscope equippedwith a high-brightness extreme field emission gunand a CEOS (Corrected Electron Optical SystemsGmbH) aberration corrector for the probe-forminglenses, operated at 300 kV. The AC-STEM imageswere acquired with an illumination semi-angle of18 mrad and a detector inner semi-angle greaterthan 35 mrad, chosen to minimize any possibleBragg diffraction contrast and maximize overallimage signal and especially atomic number (Z) con-trast between Pd atoms/clusters and the underlyingsupport. Images were obtained in suitably thin re-gions of the specimen for minimal background fromthe support and for reduced overlap of Pd atoms.Per-pixel dwell times of 5–10μs and probe currentsof 40–60 pA were selected to achieve sufficientsignal-to-noise for single Pd atom visibility whilstminimizing beam-induced changes, providing im-ages representative of the Pd atom species and theirdistribution on the LMO,GCN, PTI and PHI hosts.

Hydrogenation of 2-methyl-3-butyn-2-olThe hydrogenation was carried out in a microwavereactor (CEM Discover SP) with a pressure-controlled vessel under continuous stirring. In atypical reaction, the feed solution containing 0.4 Msubstrate in toluene (1.5 cm3) was microwavedin the presence of the catalyst (15 mg) for 1 h at323 K. The initial hydrogen pressure was 3 bar inall experiments. The resulting reaction mixture wasfiltered (pore size, 0.45 μm) and the products werecollected. The continuous tests were carried outin a flooded-bed micro-reactor (ThalesNano H-Cube ProTM), in which the liquid-feed-containing0.4-M substrate in toluene and gaseous hydrogen(generated in situ by Millipore water electrolysis)flowed concurrently upward through a cylindricalcartridge (3.5-mm internal diameter) containinga fixed bed of catalyst (0.1 g) and silicon carbide(0.2 g) particles, both with a size of 0.2–0.4 mm.The reactions were conducted at T = 323 K, P =3 bar, liquid (1 cm3 min−1) andH2 (36 cm3 min−1)flow rates.The products were collected every 20minafter reaching steady state. All collected samples

Page 30: Special Topic: Single-Atom Catalysts - Western Engineering...Predictive approach of heterogeneous catalysis Metal-organic frame-work assisted synthesis of single-atom catalysts for

RESEARCH ARTICLE Chen et al. 651

were analyzed offline using a gas chromatograph(HP-6890) equipped with a HP-5 capillary columnand a flame ionization detector. The conversion(X) of the substrate was determined as the amountof reacted substrate divided by the amount ofsubstrate at the reactor inlet. The selectivity (S) toeach product was quantified as the amount of theparticular product divided by the amount of reactedsubstrate. The reaction rate (r) was expressed as thenumber of moles of product formed per mole of Pdand unit of time.

DFT simulationsSlab models representing different carbon nitridescaffolds have been described through the DFT asimplemented in the Vienna Ab initio SimulationPackage (VASP) code [45], using the Perdew-Burke-Ernzerhof (PBE) functional together withD3 dispersion terms [46,47]. Core electrons werereplaced by projector augmented wave (PAW)method [48] and the valence electrons were ex-panded in plane waves with a kinetic cut-off energyof 450 eV. The bulk structures were derived fromour previous carbon nitride systemby expanding thebuildingmotifs.The k-point densitieswere 5×5×5for GCN and PTI system and 3 × 3 × 5 for PHI.Slabmodelswere cut along the vanderWaals planes;the slabs contain four layers and are interleaved byat least 12 A of vacuum. The optimized crystal lat-tices are presented in the Supplementary Material.The k-point sampling in these cases was 3 × 3 × 1(GCN and PTI) and 1 × 1 × 1 (PHI). The PdCl2precursor and Pd atoms were anchored at differentpositions in the cavities and within the two upper-most layers, which are preferential sites for the sim-plest carbon nitrides as observed in first-principlesMD. The Heyd-Scuseria-Ernzerhof (HSE03) [49]functional was used to generate the partial densityof states (PDOS) of the relaxed structures, includ-ing 25% of exact Hartree-Fock exchange. Ab-initioMD simulations were conducted on the Perdew-Burke-Ernzerhof (PBE) level, and comprised heat-ing/equilibration cycles in which the system washeated to 500 K with a cycle step of 100 K for a to-tal duration of 10 ps.The structures can be retrievedfrom the ioChem-BD database [50] at the followingdataset: DOI:10.19061/iochem-bd-1-75.

SUPPLEMENTARY DATASupplementary data are available atNSR online.

ACKNOWLEDGEMENTSScopeM at ETH Zurich for access to its facilities. The service ofMicroelemental Analysis at ETH Zurich for CHN analysis. BSC-RES for providing generous computational resources. Dr R. Verelfor NMRmeasurements. Dr A.J. Martın for SEMmeasurements.

FUNDINGThis work was supported by ETH Zurich, the Swiss NationalScience Foundation (200021–169679), the Spanish Ministeriode Economıa y Competitividad (CTQ2015–68770-R) and theEuropean Research Council under the European Union’s Sev-enth Framework Program (291522–3DIMAGE). E.F. thanks theSpanish Ministerio de Economıa y Competitividad (MINECO)La Caixa-Severo Ochoa for a pre-doctoral grant through SeveroOchoa Excellence Accreditation 2014–2018 (SEV-2013–0319).S.M.C. acknowledges support from the Henslow Research Fel-lowship at Girton College, Cambridge.

Conflict of interest statement. None declared.

REFERENCES1. Qiao B, Wang A and Yang X et al. Single-atom catalysis of CO

oxidation using Pt1/FeOx. Nat Chem 2011; 3: 634–41.2. Kamiya K, Kamai R and Hashimoto K et al. Platinum-modified

covalent triazine frameworks hybridized with carbon nanopar-

ticles as methanol-tolerant oxygen reduction electrocatalysts.

Nat Commun 2014; 5: 5040.3. Lucci FR, Liu J andMarcinkowskiMD et al. Selective hydrogena-

tion of 1,3-butadiene on platinum-copper alloys at the single-

atom limit. Nat Commun 2015; 6: 8550.4. Wei H, Liu X andWang A et al. FeOx-supported platinum single-

atom and pseudo-single-atom catalysts for chemoselective hy-

drogenation of functionalized nitroarenes. Nat Commun 2014;

5: 5634.5. Kyriakou G, Boucher MB and Jewell AD et al. Isolated metal

atom geometries as a strategy for selective heterogeneous hy-

drogenations. Science 2012; 335: 1209–12.6. Sun S, Zhang G and Gauquelin N et al. Single-atom catalysis us-

ing Pt/graphene achieved through atomic layer deposition. Sci

Rep 2013; 3: 1775.7. Taylor HS. A theory of the catalytic surface. Proc R Soc London

Ser A 1925; 108: 105–11.8. Wang Y, Wang X and Antonietti M. Polymeric graphitic carbon

nitride as a heterogeneous organocatalyst: from photochem-

istry to multipurpose catalysis to sustainable chemistry. Angew

Chem Int Ed 2012; 51: 68–89.9. Thomas A, Fischer A and Goettmann F et al. Graphitic carbon ni-

tride materials: variation of structure and morphology and their

use as metal-free catalysts. J Mater Chem 2008; 18: 4893–908.10. Liu AY and Cohen ML. Prediction of new low compressibility

solids. Science 1989; 245: 841–2.11. Molina B and Sansores LE. Electronic structure of six phases of

C3N4: a theoretical approach. Mod Phys Lett B 1999; 13: 193–201.

Page 31: Special Topic: Single-Atom Catalysts - Western Engineering...Predictive approach of heterogeneous catalysis Metal-organic frame-work assisted synthesis of single-atom catalysts for

652 Natl Sci Rev, 2018, Vol. 5, No. 5 RESEARCH ARTICLE

12. Lin Z andWang X. Nanostructure engineering and doping of conjugated carbon

nitride semiconductors for hydrogen photosynthesis. Angew Chem Int Ed 2013;

52: 1735–8.13. Wang X, Chen X and Thomas A et al. Metal-containing carbon nitride com-

pounds: a new functional organic–metal hybrid material. Adv Mater 2009; 21:1609–12.

14. Chen X, Zhang J and Fu X et al. Fe-g-C3N4-catalyzed oxidation of benzene to

phenol using hydrogen peroxide and visible light. J Am Chem Soc 2009; 131:11658–9.

15. Kroke E, Schwarz M and Horath-Bordon E et al. Tri-s-triazine derivatives. Part

I. From trichloro-tri-s-triazine to graphitic C3N4 structures. New J Chem 2002;

26: 508–12.16. Liu J, Wang H and Antonietti M. Graphitic carbon nitride ‘reloaded’: emerging

applications beyond (photo)catalysis. Chem Soc Rev 2016; 45: 2308–26.17. Chen ZP, Antonietti M and Dontsova D. Enhancement of the photocatalytic ac-

tivity of carbon nitrides by complex templating. Chem Eur J 2015; 21: 10805–11.18. Fina F, Callear SK and Carins GM et al. Structural investigation of graphitic

carbon nitride via XRD and neutron diffraction. Chem Mater 2015; 27: 2612–8.19. Wang X,Maeda K and Thomas A et al.Ametal-free polymeric photocatalyst for

hydrogen production from water under visible light. Nat Mater 2009; 8: 76–80.20. Sattler A, Pagano S and Zeuner M et al. Melamine-melem adduct phases:

investigating the thermal condensation of melamine. Chem Eur J 2009; 15:13161–70.

21. Kessler FK, Zheng Y and Schwarz D et al. Functional carbon nitride materials-

design strategies for electrochemical devices. Nat Rev Mater 2017; 2: 17030.22. Doblinger M, Lotsch BV and Wack J et al. Structure elucidation of polyhep-

tazine imide by electron diffraction-a templated 2D carbon nitride network.

Chem Commun 2009; 1541–3.

23. Wirnhier E, Doblinger M and Gunzelmann D et al. Poly(triazine imide) with

intercalation of lithium and chloride ions [(C3N3)2(NHxLi1−x)3�LiCl]:

a crystalline 2d carbon nitride network. Chem Eur J 2011; 17:3213–21.

24. Bojdys MJ, Muller J-O and Antonietti M et al. Ionothermal synthesis of crys-

talline, condensed, graphitic carbon nitride. Chem Eur J 2008; 14: 8177–82.25. Dontsova D, Pronkin S and Wehle M et al. Triazoles: a new class of precursors

for the synthesis of negatively charged carbon nitride derivatives. ChemMater

2015; 27: 5170–9.26. Savateev A, Dontsova D and Kurpil B et al. Highly crystalline poly(heptazine

imides) by mechanochemical synthesis for photooxidation of various organic

substrates using an intriguing electron acceptor-elemental sulfur. J Catal 2017;

350: 203–11.27. Liang Q, Huang Z-H and Kang F et al. Facile synthesis of crystalline polymeric

carbon nitrides with an enhanced photocatalytic performance under visible

light. ChemCatChem 2015; 7: 2897–902.28. Chen Z, Savateev A and Pronkin S et al. ‘The easier the better’ preparation

of efficient photocatalysts-metastable poly(heptazine imide) salts. Adv Mater

2017; 29: 1700555.29. Vile G, Albani D and Nachtegaal M et al. A stable single-site palladium catalyst

for hydrogenations. Angew Chem Int Ed 2015; 54: 11265–9.30. Chen Z, Pronkin S and Fellinger T-P et al.Merging single-atom-dispersed silver

and carbon nitride to a joint electronic system via copolymerization with silver

tricyanomethanide. ACS Nano 2016; 10: 3166–75.

31. Li X, Bi W and Zhang L et al. Single-atom Pt as co-catalyst for enhanced pho-

tocatalytic H2 evolution. Adv Mater 2016; 28: 2427–31.32. Huang X, Xia Y and Cao Y et al. Enhancing both selectivity and coking-

resistance of a single-atom Pd1/C3N4 catalyst for acetylene hydrogenation.

Nano Res 2017; 10: 1302–12.33. Chen Z, Mitchell S and Vorobyeva E et al. Stabilization of single metal atoms

on graphitic carbon nitride. Adv Funct Mater 2017; 27: 1605785.34. Liu W, Cao L and Cheng W et al. Single-site active cobalt-based photocata-

lyst with a long carrier lifetime for spontaneous overall water splitting. Angew

Chem Int Ed 2017; 56: 9312–7.35. Vorobyeva E, Chen Z and Mitchell S et al. Tailoring the framework composition

of carbon nitride to improve the catalytic efficiency of the stabilised palladium

atoms. J Mater Chem A 2017; 5: 16393–403.36. Zheng Y, Jiao Y and Zhu Y et al.Molecule-level g-C3N4 coordinated transition

metals as a new class of electrocatalysts for oxygen electrode reactions. J Am

Chem Soc 2017; 139: 3336–9.37. Lau VW-h, Mesch MB and Duppel V et al. Low-molecular-weight carbon ni-

trides for solar hydrogen evolution. J Am Chem Soc 2015; 137: 1064–72.38. Lotsch BV, DoblingerM and Sehnert J et al.Unmaskingmelon by a complemen-

tary approach employing electron diffraction, solid-state NMR spectroscopy,

and theoretical calculations-structural characterization of a carbon nitride poly-

mer. Chem Eur J 2007; 13: 4969–80.39. Zimmerman JL, Williams R and Khabashesku VN et al. Synthesis of spherical

carbon nitride nanostructures. Nano Lett 2001; 1: 731–4.40. Komatsu T. Attempted chemical synthesis of graphite-like carbon nitride. J

Mater Chem 2001; 11: 799–801.41. Chong SY, Jones JTA and Khimyak YZ et al. Tuning of gallery heights in a crys-

talline 2D carbon nitride network. J Mater Chem A 2013; 1: 1102–7.42. Yan H, Cheng H and Yi H et al. Single-atom Pd1/graphene catalyst achieved by

atomic layer deposition: remarkable performance in selective hydrogenation of

1,3-butadiene. J Am Chem Soc 2015; 137: 10484–7.43. Jun Y-S, Lee EZ and Wang X et al. From melamine-cyanuric acid supramolec-

ular aggregates to carbon nitride hollow spheres. Adv Funct Mater 2013; 23:3661–7.

44. Vernuccio S, von Rohr PR and Medlock J. General kinetic modeling of the se-

lective hydrogenation of 2-methyl-3-butyn-2-ol over a commercial palladium-

based catalyst. Ind Eng Chem Res 2015; 54: 11543–51.45. Kresse G and Furthmuller J. Efficient iterative schemes for ab initio total-energy

calculations using a plane-wave basis set. Phys Rev B 1996; 54: 11169–86.46. Perdew JP, Chevary JA and Vosko SH et al. Atoms, molecules, solids, and sur-

faces: applications of the generalized gradient approximation for exchange and

correlation. Phys Rev B 1992; 46: 6671–87.47. Grimme S, Antony J and Ehrlich S et al. A consistent and accurate ab initio

parametrization of density functional dispersion correction (DFT-D) for the 94

elements H-Pu. J Chem Phys 2010; 132: 154104.48. Kresse G and Joubert D. From ultrasoft pseudopotentials to the projector

augmented-wave method. Phys Rev B 1999; 59: 1758–75.49. Heyd J, Scuseria GE and Ernzerhof M. Hybrid functionals based on a screened

coulomb potential. J Chem Phys 2003; 118: 8207–15.50. Alvarez-Moreno M, de Graaf C and Lopez N et al.Managing the computational

chemistry big data problem: the ioChem-BD platform. J Chem Inf Model 2015;

55: 95–103.

Page 32: Special Topic: Single-Atom Catalysts - Western Engineering...Predictive approach of heterogeneous catalysis Metal-organic frame-work assisted synthesis of single-atom catalysts for

REVIEW National Science Review

5: 653–672, 2018

doi: 10.1093/nsr/nwy077

Advance access publication 2 August 2018

CHEMISTRY

Special Topic: Single-Atom Catalysts

Single-atom catalyst: a rising star for green synthesis offine chemicalsLeilei Zhang1,†, Yujing Ren1,2,†, Wengang Liu1,2,†, Aiqin Wang1,∗ and Tao Zhang1,∗

1State Key Laboratory

of Catalysis, Dalian

Institute of Chemical

Physics, Chinese

Academy of Sciences,

Dalian 116023, China

and 2University of

Chinese Academy of

Sciences, Beijing

100049, China

∗Correspondingauthors. E-mails:[email protected];

[email protected]†Equally contributedto this work.

Received 20 April2018; Revised 24June 2018; Accepted31 July 2018

ABSTRACTThe green synthesis of fine chemicals calls for a new generation of efficient and robust catalysts. Single-atomcatalysts (SACs), in which all metal species are atomically dispersed on a solid support, and which oftenconsist of well-defined mononuclear active sites, are expected to bridge homogeneous and heterogeneouscatalysts for liquid-phase organic transformations.This review summarizes major advances in theSAC-catalysed green synthesis of fine chemicals in the past several years, with a focus on the catalyticactivity, selectivity and reusability of SACs in various organic reactions.The relationship between catalyticperformance and the active site structure is discussed in terms of the valence state, coordinationenvironment and anchoring chemistry of single atoms to the support, in an effort to guide the rationaldesign of SACs in this special area, which has traditionally been dominated by homogeneous catalysis.Finally, the challenges remaining in this research area are discussed and possible future research directionsare proposed.

Keywords: single-atom catalysts, fine chemicals, green synthesis, structure–reactivity relationship,heterogeneous catalysis

INTRODUCTIONFine chemicals (<1000 tons/year) are importantand valuable (>$10/kg) ingredients and inter-mediates for the manufacture of pharmaceuticals,agrochemicals and other specialty chemicals such asadhesives, sealants, dyestuffs, pigments, flavors,fragrances, food additives, biocides and corrosioninhibitors [1]. Due to their high purity, specializednature, technology intensiveness and high addedvalue, the manufacture of fine chemicals is alwaysamong the most active and strategic areas of thechemical industries [2].

Traditional synthetic routes for fine chemicalsgenerally involve stoichiometric chemical reactionsor use hazardous catalysts and thus produce seri-ous environmental pollution and waste; the quan-tity of unwanted by-products may even exceed theamount of the desired products produced. Typi-cal examples include the coupling of aromatic com-pounds with diazonium salts derived from stoi-chiometric amounts of nitrite salts to synthesize

aromatic azo compounds [3] and the stoichiomet-ric reduction of nitroarenes by iron metal to pro-duce functional anilines [4]. With the increasingconcern for the environment and safety, there isan urgent demand for the development of new,mild, efficient and straightforward methodologiesand catalysts to achieve the green and sustainablesynthesis of fine chemicals [5]. Such green strate-gies are characterized by several principles: atom-efficient catalytic processes, environmentally benignreagents and solvents, andone-pot tandem syntheticroutes. On these grounds, a great number of effi-cient organometallic compounds and accompany-ing straightforward synthetic processes have beendeveloped, which led to the flourishing of the finechemical industry in the twentieth century, such ashomogeneous Rh catalysts for hydroformylation re-actions of olefins [6,7]. However, although thesetransition metal complexes exhibit high catalytic ac-tivity and selectivity, they are usually sensitive tomoisture and/or air and are difficult to separate from

C©TheAuthor(s) 2018. Published by Oxford University Press on behalf of China Science Publishing &Media Ltd. All rights reserved. For permissions, please e-mail:[email protected]

Page 33: Special Topic: Single-Atom Catalysts - Western Engineering...Predictive approach of heterogeneous catalysis Metal-organic frame-work assisted synthesis of single-atom catalysts for

654 Natl Sci Rev, 2018, Vol. 5, No. 5 REVIEW

the products, which sometimes leads to the contam-ination of the products. With the rapid growth ofnanoscience, heterogeneous catalysts, mainly sup-ported transition metal nanoparticles, have beenexploited to tackle these problems; however, theiroverall catalytic efficiencies are usually inferior totheir homogeneous analogs.Therefore, the develop-ment of a new class of catalysts that integrates themerits of homogeneous and heterogeneous catalystsfor the green synthesis of fine chemicals is highly de-sirable.

Single-atom catalysts (SACs) are defined as cat-alysts in which all of the active metal species exist asisolated single atoms stabilized by the support of orby alloying with another metal [8,9]. Since the sem-inal work of Zhang, Li, Liu and coworkers, who re-ported in 2011 that the Pt1/FeOx SAC was threetimes more active than its nano-Pt counterpart forCO oxidation [8], single-atom catalysis has becomea new frontier in heterogeneous catalysis. SACs aredistinguished from nanoparticle (NP) catalysts inthat they donot containmetal–metal bonds and thatthe singlemetal atoms are usually positively charged.These unique geometric and electronic propertiesbring about significant alterations in the interac-tions with reactants/intermediates/products, lead-ing to enhanced activity and/or selectivity,which areparticularly desired for fine-chemicals production.Therefore, SACs are expected to combine the advan-tages of high catalytic activity, selectivity, stabilityand reusability for the green synthesis of fine chemi-cals.The past seven years have confirmed this expec-tation; a variety of SACs have been developed andhave demonstrated excellent catalytic performancefor various organic transformations, including se-lective hydrogenation, oxidation, hydroformylationand C–C coupling reactions. For example, FeOx-supported Pt single-atom and pseudo-single-atomcatalysts exhibited the best catalytic activity (at least20-fold more active than any previously reportedcatalyst) and selectivity (∼99%) for the rather chal-lenging selective hydrogenation of 3-nitrostyrene[10]; single-atom Rh1/ZnO catalysts showed com-parable activity to the benchmark homogeneous cat-alyst RhCl(PPh3)3 in the hydroformylation reactionof olefins [11]; and non-precious-metal Co(Fe)–N–C catalysts demonstrated ‘platinum-like’ perfor-mance for selective hydrogenation or oxidation re-actions [12–18].

This review summarizes recent advances in SAC-catalysed organic reactions for the green synthe-sis of fine chemicals under liquid-phase reactionconditions—an area in which SACs are expected tofind wide application and bridge homogeneous andheterogeneous catalysis. It shouldbementioned thatmany of the examples discussed in this review are

only simple model reactions, rather than real finechemical substrates of intermediates. As preparationmethods and characterization techniques for SACshave been described in detail in several recent excel-lent reviews [19–24], they arenot emphasized in thisreview.The focusof this review is the catalytic perfor-mance of SACs in organic transformations as well asthe coordination structure and oxidation state of thecentral single metal atoms. The stabilization mecha-nisms of the single atoms against leaching or aggre-gation in the liquid phase are also discussed.The aimof this review is to provide illustrative accounts of therecent progress in this research field and to extractfundamental principles to guide the rational designof SACs for green chemical synthesis.

SYNTHESIS OF FINE CHEMICALSPROMOTED BY SACSSACs, with their well-defined mononuclear struc-tures, have been expected to bridge homogeneousand heterogeneous catalysis since the term ‘SAC’was first introduced in 2011 [8,9]. Over the pastyears, various supported SACs have been testedfor a plethora of organic transformations, includ-ing selective hydrogenation of nitroarenes, ketones,alkynes and alkenes; selective oxidation of alcoholsto aldehydes and ketones, of benzene to phenol andof silanes to silanols; hydroformylation of olefins;C–C coupling reactions; and biomass-related hy-drodeoxygenation reactions. Compared with theirNP counterparts, SACs afford enhanced activity bya factor of several to hundreds per metal atom andunparalleled selectivity in some reactions. The ex-cellent stability of SACs provides additional advan-tages over their homogeneous analogs.While the ac-tive site structures of SACsmay be less uniform thanthose of homogeneous catalysts due to their hetero-geneous support surface, the well-defined structuresarising from the coordination between single atomsand the support provide SACs with promising prop-erties to mimic homogeneous catalysis in organictransformations.

Catalytic hydrogenationCatalytic hydrogenations of unsaturated organiccompounds represent one of the most importantclasses of chemical transformations and are widelyapplied in both the chemical industry and labora-tory organic synthesis [25]. The majority of hydro-genation reactions involve the direct use ofH2 as thehydrogen source and are catalysed by VIII–X groupmetals such as Ni, Pd, Ru and Pt.

Page 34: Special Topic: Single-Atom Catalysts - Western Engineering...Predictive approach of heterogeneous catalysis Metal-organic frame-work assisted synthesis of single-atom catalysts for

REVIEW Zhang et al. 655

Pt/FeOx

Selectivity (%)Whiteline intensity

(a) (b)

100

90

80

70

0 1 2 3 4 5

1.7

1.6

1.5

1.4

TOF (h-1) Selectivity (%)Whitelineintensity

1500

1350

1200

1050

900

750

0 1 2 3 4 5

1.55

1.50

1.45

1.40

1.35

1.30

98

96

94

92

NO2NH2

Figure 1. The dependence of activity/selectivity on the Pt content of the Pt/FeOx catalysts (a) and on the Na content of the

Na-2.16%Pt/FeOx catalysts (b) for the chemoselective hydrogenation of 3-nitrostyrene, as well as the correlation between

activity/selectivity and the oxidation state (white-line intensity) of Pt in the two systems.

Themanufacture of functional anilines is of greatsignificance in the chemical industry due to theirversatility in biologically active natural products,agrochemicals, pharmaceuticals, dyes and ligandsof organometallic complexes [4,26,27]. Comparedwith traditional synthetic methodologies (e.g. re-duction by the transition metals iron and tin), cat-alytic hydrogenation is more efficient and environ-mentally benign. However, when one or more re-ducible organic groups (e.g. C = C) are presenton the aryl ring containing the nitro group, thechemoselective hydrogenation of the nitro groupis very challenging. Modification of the transitionmetal catalysts with proper additives achieved en-hanced selectivity [28]; however, the catalytic ac-tivity was severely compromised. The hydrogena-tion of the nitro group is generally considered to besize-insensitive, whereas the reduction of the C=Cgroup is highly sensitive to the size of the transi-tion metal, with the reaction rate increasing withparticle size [29]. Consequently, downsizing the ac-tive metal to its ultimate dispersion, namely single-atom dispersion, is expected to accomplish excel-lent chemoselectivity by suppressing theC=Cdou-ble bond hydrogenation. Zhang, Wang and cowork-ers designed a FeOx-supported Pt single/pseudo-single-atom catalyst. The term ‘pseudo-single-atomcatalyst’ indicates a special structure composed of afew to tens of atoms that are loosely and randomlyassociated with each other, but do not form strongmetallic bonds, and thus show structure and func-tion similar to that of isolated single atoms.The cat-alyst was able to hydrogenate a broad scope of ni-troareneswith different functional groups to the cor-responding functionalized anilines with a turnoverfrequency (TOF) as high as 1514 h−1 and a chemos-electivity above 95% [10]. The loading of Pt on the

FeOx support had a significant effect on both the ac-tivity and chemoselectivity.As shown inFig. 1a, boththe activity per metal atom (TOF) and the chemos-electivity increased with decreasing Pt content un-til atomic dispersion was reached at 0.08 wt% Pt,at which the catalyst contained exclusively singleatoms of Pt, and therefore achieved the highestactivity and chemoselectivity (98.6% at a conver-sion of 96.5%). Interestingly, the white-line inten-sity of Pt in XANES (X-ray Adsorption Near-EdgeStructure spectroscopy), which reflects the oxida-tion state of Pt, followed the same trend as thecatalytic activity and chemoselectivity—that is, thehigher the oxidation state, the higher the activityand chemoselectivity. This result indicated that theisolated, positively charged Pt on the FeOx sup-port acted as the main active site for the chemose-lective hydrogenation of functionalized nitroarenes.Keeping this point in mind, and considering thatthe alkalimetals could yield positively chargedmetalcenters and also promote the dispersion of metalspecies [30,31], the same authors further tunedthe electronic properties of the Pt center by intro-ducing alkali metals to the high-Pt-loading catalyst(2.16wt%Pt/FeOx) [32]. By increasing the amountof Na to 5.03 wt%, the chemoselectivity increasedremarkably from 66.4 to 97.4% (Fig. 1b); concur-rently, the oxidation state of Pt increased as well.Moreover, detailed EXAFS (extended X-ray absorp-tion fine structure) data analysis reveals that thestructure of the central Pt sites changes with vary-ing the amount of sodium; the Pt–Pt contributiondecreases while the Pt–O contribution increases.Consistently with this trend, HAADF-STEM (highangle annular dark field scanning transmission elec-tron microscopy) images clearly show that the Ptparticles in the Na-containing samples all featured

Page 35: Special Topic: Single-Atom Catalysts - Western Engineering...Predictive approach of heterogeneous catalysis Metal-organic frame-work assisted synthesis of single-atom catalysts for

656 Natl Sci Rev, 2018, Vol. 5, No. 5 REVIEW

abundant dark regions, which are probably due tolighter elements such as Na and Fe. In combina-tion with the 57Fe Mossbauer spectroscopy resultthat theNaFeO2 species forms in theNa-containingsample, the authors proposed that the active siteswere likely Pt–O–Na–O–Fe species. This scenariois reminiscent of that of a homogeneous catalyst, ifthe Na–O–Fe surrounding the single-atom Pt cen-ter is viewed as a robust ligand bridged by oxygen.Due to the higher contribution from Pt–O bond-ing than Pt–Pt bonding, the performance of thisNa-modified Pt/FeOx catalyst is quite similar tothat of the SAC in terms of activity and selectivity.Nevertheless, in comparison to the SACs with rela-tively low metal loading, (e.g. 0.08% Pt/FeOx), thehigh-metal-loading Na-modified Pt/FeOx catalystsafforded amore than 20-fold increase in the produc-tivity (activity per catalyst mass, rather than per Pt),which is of high importance in practical applications.In order to make the process more environmentallybenign, the authors further studied the reaction ina CO2-expanded toluene/THF system. The resultsshowed that both the conversion andchemoselectiv-ity could reach above 95% over Pt/FeOx SAC, whilethe amount of the solvent toluene could be reducedby 90% in the CO2-expanded toluene system [33].

It is noted that, even with the same single-atom dispersion and the same support, the activityand chemoselectivity of an SAC may vary depend-ing on the intrinsic properties of the metal used.For example, while both Pt1/FeOx and Ir1/FeOxSACs are highly chemoselective towards the hy-drogenation of 3-nitrostyrene to 3-aminostyrene,the Pd/FeOx SAC is poorly selective (selectivitylower than 20%) for this reaction due to the intrin-sically high activity of Pd towards the hydrogena-tion of C = C bonds [10]. Similarly, Zhang et al.found that sub-nanometric Pd clusters supported onCeO2 nanorods were highly active (TOF as high as44 059 h−1) for hydrogenation of 4-nitrophenol andchemoselective for a broad scope of nitroarenes con-taining –OH, –X, –C = O and –CN groups, butpoorly selective for substrates possessing a –C = Cgroup [34]. It would be highly interesting to engi-neer themetal–support interaction to develop a newPd SAC that is both highly active and chemoselec-tive for the hydrogenation of nitro-styrenes.

The reaction mechanism of the chemoselectivehydrogenation of nitroarenes over SACs has yet tobe clarified. The results of control experiments in-volving the competitive adsorption of nitrobenzeneand styrene revealed that the unique selectivity ofSACs is not due to their intrinsically low activity forstyrene hydrogenation, but instead due to the fa-vorable adsorption of the nitro group in the pres-ence of –C = C group [10], which is similar to

the case behavior of the nanogold catalyst as well asthe TiO2-decorated nano-Pt catalyst [29,35]. Basedon Fourier transform infrared spectroscopy (FT-IR) and density functional theory (DFT) theoreti-cal studies [36], the nitro group is preferentially ad-sorbed on the support; the oxygen vacancies on thesupport surface (e.g. reducible supports like FeOxand CeO2) or the basicity of the support (e.g. theaddition of Na increases the basicity of the supportin the above-mentioned work) facilitate the prefer-ential adsorption of the nitro group. This interac-tion can be understood in terms of Lewis acid–baseinteractions. The oxygen atoms in the nitro groupare Lewis bases, while the oxygen vacancies of thereducible support act as Lewis acids; this ensuresthe strong and preferential adsorption of the nitrogroups on the support surface, even in the presenceof other reducible groups. Additionally, hydrogencan be easily dissociated at the isolated Pt (or othernoble metal) single atoms, and the hydrogenationreaction probably occurs at the interface betweenthe single metal atoms and the support, as illus-trated in Fig. 2a. Nevertheless, this mechanism can-not completely justify the key role of single atoms ingoverning the high chemoselectivity. Alternatively,the nitro group could be adsorbed at the interfaceof the single metal atom and the support via the in-teraction of its two oxygen atoms with both the sin-gle metal atom and a nearby support cation, as il-lustrated in Fig. 2b. In either of the plausible sce-narios above, the key point to be underscored isthat the isolated andpositively chargedmetal centerscan successfully suppress the adsorption of C = Cbonds, which would occur easily on their nanopar-ticle counterparts. Therefore, the single-atom dis-persion is of paramount importance to achieve thechemoselective hydrogenation of nitroarenes to ani-lines, although cooperationbetween the singlemetalatoms and the specific support is required.Achievingan atomic-scale understanding of the reactionmech-anism will require combined spectroscopic and the-oretical studies.

There are two reaction pathways for the ni-troarene reduction: a direct pathway involvingnitro-nitroso-hydroxylamine-aniline steps and acondensation pathway involving nitro-nitroso-hydroxylamine-azoxy-azo-hydrazo-aniline steps[37]. In the above SAC systems, steric hindranceshould prevent the reaction from proceeding via thecondensation pathway, which requires the adsorp-tion of two nitroarene molecules. However, in somecases, the azo is the desired product rather thanthe aniline, because aromatic azo compounds, withtheir great diversity of structures, represent themostwidely used class of synthetic organic colorantsby far [3]. For this purpose, Wang, Zhang and

Page 36: Special Topic: Single-Atom Catalysts - Western Engineering...Predictive approach of heterogeneous catalysis Metal-organic frame-work assisted synthesis of single-atom catalysts for

REVIEW Zhang et al. 657

(a)

(b)

M

M M M

PtO O O NO2 H2

M

Pt

M M

MO O O

OON

NO2

NH2

NH2

H2H2

MM

M

M

PtO O O

MM

M MMPtO

O O OO

M MM

M MMPtO

O O OO

MPtM

N

M M MMO

O O

OO

OO

PtM

N

M M MMO

O O

OO

O

Figure 2. Two possible scenarios for the hydrogenation mechanism of nitroarenes on Pt-SACs with a reducible support.

coworkers explored atomically dispersed Co–N–Ccatalysts for the direct synthesis of azo compoundsfrom nitroarenes [12]. The synthesis of this typeof SACs involved the impregnation of a Co, N,C-containing precursor (e.g. a Co2+ phenanthro-line complex) on a basic support (e.g. Mg(OH)2)followed by pyrolysis at 600–800◦C and etchingwith acid to remove the basic support as well as anyparticles of Co0 or CoOx. Extensive characterizationwith aberration-corrected HAADF-STEM, XAS(X-ray absorption spectroscopy) and XPS (X-rayphotoelectron spectroscopy) in combinationwith DFT calculations revealed that the resultingmaterial was a truly atomically dispersed catalystin which cobalt existed exclusively as single atomsand was bonded strongly to four pyridinic N atomswithin a deformed graphitic layer and weakly to twooxygen molecules in the axial direction, as shownin Fig. 3a and b. Differently from the rigid planestructure generally proposed in the literature [38],the deformed CoN4C8–2O2 configuration mightimpart structural flexibility to the SACs similarto that of metal–ligand complexes and thus theircatalytic activities may be comparable or evensuperior to that of their homogeneous analogs. Inthe hydrogenation of nitroarenes under mild condi-tions and in the presence of additional base (80◦C,3 MPa H2, 1.5 h, tert-butyl alcohol as the solvent,catalyst loading of 0.7 mol% Co, 0.2 equivalents ofNaOH), the Co–N–C SACs afforded good activity(TOF 35.9 h−1) and excellent selectivity to azocompounds (Fig. 3c). Various functional groupsincluding –CH3, –C = C, –CF3, –Cl, –Br and –Iwere tolerated in the reaction, and the catalyst couldbe reused without decay in the selectivity. The basicadditive is the key to the condensation betweennitrosobenzene and hydroxylamine to producethe target azo product; otherwise, anilines wouldbe produced as the final product, as shown by thepioneering work of Beller and coworkers [39].

While the molecular-scale understanding of thereaction mechanism of these atomically dispersed

Co–N–C catalysts has yet to be clarified, the M–Nxmotif (where M refers to transition metals such asCo, Fe, Ni, Cu, etc.) is probably the active site, evenin the earlier reported CoxOy/N–C or FexOy/N–C systems [39,40]. To determine the real activespecies, Zhang et al. prepared a catalyst composed ofmetallic Co, CoxOy, CoxCy and CoNx species usingcarbon as the support [41]. After acid etching, 78%of the Co-containing species had leached out andthe residual Co species were encapsulated in thickgraphitic nanoshells and thus inaccessible to anymolecules, except for Co–Nx, which is stable againstacid etching. Interestingly, the performance of thecatalyst in the oxidative cross-coupling of primaryand secondary alcohols remained the same, stronglysuggesting that the atomically dispersedM–Nx is theactive site. This work also indicates that some com-plicated heterogeneous systems may need to be re-visited to identify the real active sites by means ofstate-of-the-art characterization techniques.

For thehydrogenation reactionsoccurringon theSACs, one key question is how hydrogen moleculesare activated on the single atoms. Recently, an in-creasing number of studies have suggested thatH2 is activated on SACs via a heterolytic pathway[42–45], which is quite different from the case ofNP systems in which the homolytic dissociation ofH2 is favorable. For example, Zheng and cowork-ers reported a high-loading Pd1/TiO2 SAC that wasobtained via a photochemical route on an ethyleneglycolate (EG)-protected TiO2 nanosheet material[42]. In theirmethod, the EG radicals that were gen-erated by ultraviolet (UV) radiation played a keyrole in producing and stabilizing the Pd single atomsat a relatively high loading (1.5 wt%). The coordi-nation of the single Pd atoms with oxygen atomsfrom the EG molecule allowed the dissociation ofH2 in a heterolytic mode to produce O–Hδ+ andPd–Hδ− (Fig. 4), as proven by both kinetic isotopeeffect (KIE) and DFT calculations. The Pd1/TiO2SAC exhibited high activity and stability for C = Cand C=Ohydrogenations; the TOFwas enhanced

Page 37: Special Topic: Single-Atom Catalysts - Western Engineering...Predictive approach of heterogeneous catalysis Metal-organic frame-work assisted synthesis of single-atom catalysts for

658 Natl Sci Rev, 2018, Vol. 5, No. 5 REVIEW

N N N N

N N

N NN N

N NN N

N N

l l

5

8

1 2

4

7

97%, 1.5 h 94%, 1 h

N N

3

6

9

88%, 2.5 h

94%, 2 h

97%, 1.5 h

99%, 2.3 h

96%, 1.3 h

63%, 7 h

85%, 1 h

CF3

Cl

Br Br

Cl

F3C

NO2NO2

R2R1

Co-N-C SAC (0.7 mol%)

0.2 equiv. NaOH, 2 mL TBA

R2

R1N

N H2O

(a)

(c)

(b)

Energy (eV)7710 7720 7730 7740 7750 7760 7770

TheoryCo-N-C

XANES simulation

Inte

nsity

(a.u

.)

Figure 3. (a) HAADF-STEM image of the Co–N–C catalyst; (b) X-ray absorption spectroscopy (XAS) fitting and density func-

tional theory (DFT) calculations of the Co–N–C catalyst; (c) substrate scope of the hydrogenation reaction on the Co–N–C

catalyst. Adapted with permission from [12].

by a factor of 9 for styrene hydrogenation and a fac-tor of 55 for aldehyde hydrogenation compared tothe commercial Pd/C catalyst.The higher activity ofthePd1/TiO2 for thehydrogenationof polar unsatu-ratedbondsprobablyoriginated fromtheheterolyticdissociation of H2 at the Pd1–O interface. Simi-larly, Yan and coworkers reported that single Pt1atoms anchored on active-carbon-supported phos-phomolybdic acid (PMA) were more active for thehydrogenation of polar –C=Ogroups than of non-polar –C=Cgroups, implying the heterolytic disso-ciation of hydrogen on the Pt1–O4 active sites [43].In our recent work on the single/pseudo-single-atom catalyst Pt/WOx, hydrogen was also found todissociate in a heterolytic fashion at the interface of

Pt and WOx. The H+ formed on WOx and the H−

on the single Pt atom provided Brønsted acid sitesand hydrogenation sites, respectively, for the con-certed dehydration–hydrogenation reaction of glyc-erol to produce 1,3-propanediol [44], as shown inFig. 5. Moreover, the introduction of single goldatoms to this Pt/WOx system further promoted theheterolytic dissociation of H2, possibly by modulat-ing the interaction between Pt and WOx, leading toremarkably enhanced activity and chemoselectivitytowards 1,3-propanediol [45].

The heterolytic dissociation of H2 favored onSACs is reminiscent of that of frustrated Lewis pairs(FLPs) in homogeneous catalysts, where H2 isheterolytically dissociated at the sterically

Page 38: Special Topic: Single-Atom Catalysts - Western Engineering...Predictive approach of heterogeneous catalysis Metal-organic frame-work assisted synthesis of single-atom catalysts for

REVIEW Zhang et al. 659

H2C O

OO

Ti5c

O

HH

PdIX

H2C

IVTi5cIII Ti5c

IV Ti5cIV

Pd1H-OH-0.98

0.69-0.29

0.40

IX

Pd1-H2

Pd1

VII

VIII

TS1

Figure 4. Energies and a model of the intermediates and

transition states in the heterolytic H2 activation process of

Pd1/TiO2. Adapted with permission from [42].

encumbered Lewis acids and bases [46]. Insupported SACs, the single metal atoms are usuallypositively charged and therefore act as a Lewis acid,while the nearby oxygen or nitrogen atoms from thesupport can act as a Lewis base; the combination ofthe two facilitates the dissociation of hydrogen via aheterolytic pathway.

SACs not only show higher activity than theirNPs counterparts in catalysing the hydrogenationof polar unsaturated groups, but also catalyse thehydrogenation of alkynes with high selectivitytowards alkenes. This kind of reaction was initiallyinvestigated on single-atom alloy (SAA) catalysts.The term SAA was first introduced by Sykes andcoworkers [47]. SAA catalysts are special SACs inwhich the active metal atom (e.g. Pd) is isolatedcompletely by the surrounding less active metalatoms (e.g. Cu). This configuration allows forthe easy dissociation of hydrogen on the activesingle metal atoms and simultaneously weakensthe binding of intermediates by spillover to the lessactive metal metals, thus accomplishing both highactivity and high selectivity in the hydrogenationof alkynes to alkenes. Various combinations ofSAA catalysts have been explored, such as Au–Pd[48], Ag–Pd [49], Cu–Pd [50], Zn–Pd [51], In–Pd[52] and Pt–Cu [53]. In addition to these SAAs,single Pd atoms on other supports have also beendemonstrated to be effective [54–56]. For example,Perez-Ramirez and coworkers reported a Pt SACsupported on mesoporous polymeric graphiticcarbon nitride (mpg–C3N4) that was 3 orders ofmagnitudemore active than traditional NP catalysts(e.g. Au, Ag and CeO2) in the hydrogenation of1-hexyne, with 100% 1-hexene selectivity [56].When the reaction was performed at 70◦C and5 bar H2, [Pd]mpg–C3N4 was more active than

Step 6

Step 1

PtW W

O O

O

O O O O

O

O O O

OOW W W

Pt

PtH

2

H2

H 2H 2 C CC

H2

OH

OH

HO

Step 2

PtW W

O O

O

OO

C C

C

OH

OH

HO O O

O

O O O

OOW W W

Pt

Pt

H2

H 2H 2 H2

δ-H

Hδ+

Step 3

PtW W

O O

O

OO

C C

C

OH-H 2

O

OH

HO O O

O

O O O

OOW W W

Pt

Ptδ-H

Hδ+

Step 4

PtW W

O O

O

OO

C C

C OH

HO O O

O

O O O

OOW W W

Pt

Ptδ-

H

+

Step 5

PtW W

O O

O

OO

C CHC OH

HO O O

O

O O O

OOW W W

Pt

Pt

PtW W

O O

O

OO

C CHC

OH

HO

O O O

O

O O O

OOW W W

Pt

Pt

Figure 5. Proposed reaction scheme for the hydrogenolysisof glycerol to 1,3-PD over single-atom and pseudo-single-

atom Pt/WOx catalysts. Adapted with permission from [44].

Pd–Pb/CaCO3 by a factor of 4, more selectivethan Pd/Al2O3 (90 vs 69%) and was as active asthe Pd/TiS modified by ligands. Furthermore,no decrease in activity or selectivity was observedduring 20 h of time-on-stream. Other substrates,such as 2-methyl-3-butyn-2-ol and 3-hexyne, couldalso be smoothly converted with excellent chemo-and stereo-selectivity (cis/trans ratio >20). DFTcalculations revealed that hydrogen underwent het-erolytic dissociation assisted by the support, leavingone H atom bonded to a N atom in the support andthe secondone to aPdatom.Whenalkynemoleculeswere adsorbed, the semi-hydrogenation reactionproceeded smoothly and the alkene product waseasily desorbed without over-hydrogenation or

Page 39: Special Topic: Single-Atom Catalysts - Western Engineering...Predictive approach of heterogeneous catalysis Metal-organic frame-work assisted synthesis of single-atom catalysts for

660 Natl Sci Rev, 2018, Vol. 5, No. 5 REVIEW

oligomerization.The proposed reaction mechanismis quite similar to that of the gas-phase selectivehydrogenation of acetylene to ethylene over Au(Ag,Cu) alloyed Pd SACs and Pd-Zn intermetallic com-pounds reported by Zhang and coworkers [48–51],in which ethylene was demonstrated to interact withthe single Pd atoms via a weak π -bonding pattern,and could be readily desorbed without saturation.

Oxidation reactionSelective oxidation is an important strategy for pro-ducing oxygenates, such as aldehydes, ketones oracid products, to be employed as building blocksin chemical processes that range from kilogram-scale applications in pharmaceuticals to 1000-ton-scale in chemicals. Traditional catalytic routes gen-erally use expensive homogeneous complexes andare performed under harsh operating conditions,which bring about serious environmental issues andinevitable over-oxidation side reactions.

With the successful synthesis of SACs, a numberof SACshavebeen explored in selective oxidation re-actions and have shown promising catalytic perfor-mances. Pioneering work in this area was reportedin 2007 by Lee and coworkers, in which a single-site Pd/Al2O3 catalyst showed 30 times greater cat-alytic activity (TOF: 4096–7080 h−1) and selectiv-ity (>91%) than its NP counterpart in the aerobicoxidation of allylic alcohols including cinnamyl al-cohol, crotyl alcohol and benzyl alcohol [57]. Thissingle-site Pd/Al2O3 catalyst was obtained by us-ing mesoporous alumina as the support and an ex-tremely low loading of Pd (0.03wt%), both of whichensured the atomic dispersion of the Pd species.TheisolatedPdII surface species, whichwas identifiedus-ing atomic-resolution HAADF-STEM, EXAFS andXANES, were proposed to be the active sites forthe reaction.This catalyst also demonstrated impres-sive stability; the catalytic performance remainedunchanged during the course of reaction over pe-riods of days. This exceptional durability could beattributed to the high stability of the isolated PdII

species, as well as that of the mesoporous structureof the support. Very recently, Li et al. reported anAu1/CeO2 SAC system that also showed excellentactivity, selectivity anddurability for the selective ox-idation of alcohols to the corresponding aldehydes[58]. Based on isotopic exchange experiments, theyproposed that the lattice oxygen from theCeO2 sup-port participated in theoxidation reaction, leading tothe high selectivity towards aldehyde. The SAC sys-tem facilitated the removal and replenishment of thelattice oxygen by maximizing the interfacial sites be-tween the single gold atoms and the CeO2 support.

In contrast, the gas-phase oxygen activation on thesurface of gold NPs was less selective towardsthe aldehyde. Another important factor determin-ing the selectivity is the adsorption of the aldehyde,which was found to be much weaker on the singleatoms than on the NPs.

Toshima and coworkers constructed a ‘crown-jewel’-structured Pd-alloyed Au SAC via replacingthe Pd atoms at the top position of Pd147 particleswith an Au atom [59]. The as-prepared catalystsshowed excellent catalytic activity in the oxidationof glucose. The specific activity of the Au SAC was20–30-fold higher than that of thePd andAumotherclusters and 8–10 times those of the Pd/Au alloyswith different Au/Pd ratios, although all had simi-lar average particle sizes. The higher catalytic activ-ity of the Pd-alloyed Au SAC than the Pd–Au al-loys indicated position-dependent catalysis beyondthe synergistic effects of Au and Pd, as the top atomsare assumed to be more active than the edge andface atoms owing to themore unsaturated coordina-tion. To confirm this proposal, the authors preparedthree Pd-alloyed Au SACs with different amountsof Au using the same method, in which Au wouldpreferentially replace the top Pd atoms, then edgeand lastly face Pd atoms. As the Au concentrationwas increased, the specific activity normalized to theAu mass decreased, which unambiguously demon-strated the lower activity of the edge and face Auatoms compared to that of top Au atoms. The Au4f XPS spectra of the Pd-alloyed Au SACs showeda negative shift in binding energy compared withthat ofmonometallicAu, indicating electron transferfrom Pd to Au; thus, Au was negatively charged.TheAuδ– species could effectively donate electrons toO2 togeneratehydroperoxo-like species,whichwereconsidered to play an important role in the oxida-tion of glucose. Additionally, Tsukuda and cowork-ers reported a Pd1Au24 SAC supported on a carbonnanotube (CNT) and investigated its catalytic per-formance in the aerobic oxidation of benzyl alcohol[60].Comparedwith themonometallic Au25/CNT,the single-Pd-atom-doped Pd1Au24/CNTexhibitedsignificantly enhanced conversion of benzyl alcohol(from 22 to 74%).The synergistic effect between Auand Pd was attributed to electron transfer from Pdto Au, which promoted the activation of O2. How-ever, the selectivity of the catalyst was unsatisfac-tory and aldehyde, acid and ester products were pro-duced with similar yields.

It should be noted that the oxidation mechanismon the single Pd atomsmight be quite different fromthat on the single gold atoms or NPs. In the formercase, the active oxygen mostly likely arises from theisolated PdO species and the Pd reduced during thereaction can be quickly re-oxidized by oxygen gas,

Page 40: Special Topic: Single-Atom Catalysts - Western Engineering...Predictive approach of heterogeneous catalysis Metal-organic frame-work assisted synthesis of single-atom catalysts for

REVIEW Zhang et al. 661

thus completing the redox cycle [57], while, in thelatter case, oxygen is first activated on the electron-rich gold atoms to formO2−, which thenoxidizes thealcohol substrate [59].

In addition to noble metal SACs, atomically dis-persedM–N–Ccatalysts have also been explored foraerobic oxidation reactions. In 2016, Davis’s groupreported a series of M–N–C SACs with non-noblemetals, including Cr, Fe, Co, Ni and Cu, confinedin a nitrogen-containing carbon matrix for the aer-obic oxidation of alcohols in aqueous media [14].Among these catalysts, Cu–N–C showed the high-est catalytic activity for benzyl alcohol oxidation(TOF = 3.4 × 10−2 s−1), followed by Co–N–Cand Cr–N–C. The activity of Fe–N–C was lowerthan that of Cu–N–C by a factor of 50, althoughFe–N–C is usually more active than Cu–N–C inthe oxygen reduction reaction (ORR). KIE experi-ments were conducted to investigate the differencein catalytic activity between them. A large KIE ef-fect (4.6) was observed on Fe–N–C, indicating thatthe elimination of β-H in benzyl alcohol was therate-determining step, whereas a weaker KIE effect(2.2) was observed for Cu–N–C, suggesting Cu–N–C could abstract the β-H more easily. However,it should be noted that the as-prepared catalystsalso contained nanoparticles encapsulated by car-bon shells, whose role in the catalysis was not clearlydetermined.

Guan and coworkers also developed atomicallydispersed Co SACs supported on nitrogen-dopedgraphene and investigated their catalytic behaviorin the selective oxidation of alcohols [15]. TheCo SAC exhibited excellent catalytic activity fora broad scope of substrates including aromaticand aliphatic alcohols, and moderate to excellentyields (54–92.4%) of the corresponding aldehy-des were obtained. The base-free conditions andthe use of air/oxygen as an oxidant seem attrac-tive for practical application, although the organicsolvent N,N-dimethylformamide (DMF) was used.Furthermore, some critical points, such as the dop-ing effect of N in graphene, structure of the Co–Nmoiety and effect of calcination temperature, haveyet to be clarified.

The selective oxidation of C–H bonds is a moredemanding class of reactions than alcohol oxida-tion due to the high dissociation energy of C–Hbonds and the over-oxidation side reactions [61].Very recently, Liu et al. studied the selective oxi-dation of C–H bonds in aromatic and aliphatic hy-drocarbons by using an atomically dispersed Fe–N–C catalyst [16]. By employing tert-butyl hydroper-oxide (TBHP) as the oxidant, Fe–N–C SAC af-forded excellent activity and selectivity even at roomtemperature, and a broad spectrum of substrates

could be smoothly converted into the correspond-ing ketones. The performance was highly depen-dent on the pyrolysis temperature at which the SACwas derived from the Fe(phen)x (phen = 1,10-phenanthroline) precursor. Although the Fe–N–Ccatalysts obtained at different pyrolysis tempera-tures (ranging from 700 to 900◦C) included almostexclusively or even exclusively atomically dispersedFe species, their exact structures were rather com-plicated and in fact consisted of a mixture of FeNxspecies (x = 4–6). The employment of the power-ful 57FeMossbauer spectroscopy technique allowedthe identification and quantification of the differentFe centers. As shown in Fig. 6a, the relative con-centration of each species differed with the pyroly-sis temperature, and the most active species was themedium-spin FeN5 site (D4 in Fig. 6a), which af-forded a TOF of 6455 h−1 for the oxidation of ethyl-benzene to acetophenone.This TOFwas more than1 order of magnitude higher than that of the high-spin FeN6 species (D2) and low-spin FeN6 species(D3), and even several times more active than theFe(II)N4 species (D1).Thehigh activity of theFeN5structure was attributed to the unsaturated coordi-nationof theFe single atoms (actuallyFe3+ cations),which provided adsorption sites for TBHP (H–O–O–, Fig. 6b). This behavior resembles the oxygenactivation of the hemoglobin molecule [62]. In thisaspect, the Fe–N–C SACs are expected to mimicthe metalloenzymes in more important but compli-cated organic transformations. However, in the cur-rent Fe–N–C SACs, the most active species are un-fortunately the least abundant (28% or less based onFig. 6a). Evidently, there is plenty of room for fu-ture research in the design and synthesis of atomi-cally dispersed single-site catalysts.

The direct catalytic conversion of benzene tophenol is a demanding and practically importantreaction in the field of C–H activation. Bao andcoworkers prepared a highly dispersed FeN4/GNcatalyst via the high-energy ball milling of ironphthalocyanine (FePc) and graphene nanosheetsunder controllable conditions [17]. The graphene-matrix-confined coordinatively unsaturated ironsites demonstrated excellent catalytic activity inthe oxidation of benzene with H2O2 as the oxidantat room temperature. For example, the TOF valuereached 84.7 h−1 and a phenol yield of 18.7% wasobtained at a conversion of 23.4%. Notably, thereaction even proceeded efficiently at 0◦C, reachinga phenol yield of 8.3% after 24 h of reaction. Further-more, the catalyst could be reused six times withouta decrease in its catalytic activity. In situ XAS andMoossbauer measurements were performed toinvestigate the catalytic mechanism. For the freshcatalyst, both the Fe K-edge XANES spectra and

Page 41: Special Topic: Single-Atom Catalysts - Western Engineering...Predictive approach of heterogeneous catalysis Metal-organic frame-work assisted synthesis of single-atom catalysts for

662 Natl Sci Rev, 2018, Vol. 5, No. 5 REVIEW

Fe-N-C SAC (0.6 mol%)

25 oC,TBHP (2-6 equiv)

H2O (7 mL) GC yield

R1 R2

R1 R2

O TOFD4 = 6455.2 h-1

TOFD1 = 2062.7 h-1

TOFD2 = 587.7 h-1

(a)

D2

NN

NN

Y

X

Fe

D4

N

N

N

NN

Fe

D1D2D3D4 Sext 1S1

Fe-N-C-600 Fe-N-C-700 Fe-N-C-800

70

60

50

40

30

20

10

0

Are

a (%

)

D1

N

N

NN

Fe

OO

H O+ OH

O

Ph

t-BuOH

Ph

OH

PhOH

Ph

OO

Ph

O

t-BuOH

Ph

OOH

Ph

O

GC-MS analysis

H2OPh O

GC-MS analysis

EPR

EPR

(1)

(3)

(4)

(5)

(6)

H

OH

CH3OH

Ph OGC-MS analysis

Fe-N-C-700O

H ONN

NNN

N

N

NN

N

Ph

H

(2)Fe-N-C-700 O/NO/N

Ph

HH

Ph

H*

Ph

H*

Ethylbenzene Adsorption

(b)

Fe

FeFe

-6 -4 -2 0 2 4 6

D1D2D3

Fe-N-C-800

Fe-N-C-700

Fe-N-C-600

Abs

(%)

D4D3D2

Velocity (mm s-1)

D2Sext 1D4S1

Figure 6. (a) 57Fe Mossbauer spectra of the Fe−N−C-600/700/800 catalysts and the relative concentration of each species;

(b) proposed reaction mechanism of ethylbenzene oxidation on the FeN5 site. Adapted with permission from [16].

Page 42: Special Topic: Single-Atom Catalysts - Western Engineering...Predictive approach of heterogeneous catalysis Metal-organic frame-work assisted synthesis of single-atom catalysts for

REVIEW Zhang et al. 663

Fourier transformed EXAFS spectra in r-space werequite similar to that of FePc, indicating that the Fein FeN4/GN adopted the same FeN4 structure,with Fe coordinated to four N atoms within a planarsheet, as that of FePc. Upon treatment with H2O2,the pre-edge peak in the Fe K-edge XANES spectrabroadened and increased in intensity due to theformation of Fe=O.The intensity of the first strongpeak in the FT-EXAFS spectra in r-space was alsoenhanced, suggesting that a sharp increase in thecoordination of Fe resulted from the formation ofFe = O/O = Fe = O. 57Fe Moossbauer measure-ments also revealed that the amount of Fe = Oincreased after H2O2 treatment and then decreasedupon contact with benzene. These results, togetherwith DFT calculations, showed that the single Featomwas oxidized byH2O2 to form anO= Fe=Ointermediate, which then oxidized benzene to formphenol.

Li et al. also constructed Fe–N–C SACs by thepyrolysis of metal hydroxides or oxides coated withpolymers of dopamine, followed by acid leaching[18].TheobtainedFe–N–CSACexhibitedhigh cat-alytic activity for the hydroxylation of benzene tophenol (45% benzene conversion with 94% selec-tivity). Based on electron paramagnetic resonance(EPR)analysis andDFTcalculations, theyproposeda similar reactionmechanism to that of Bao, inwhichthe H2O2 oxidant was activated on the single Featom, and the resulting FeIV = O species was con-sidered to be the active intermediate in the reaction.Notably, althoughFe–N4 moietieswere proposed asthe active species in both works, the selectivity to-wards phenol (94%) in Li’s report was higher thanthat of Bao’s (∼80%), indicating that theremight besome subtle discrepancy in the exact structure of thetwo active sites. Further efforts are expected to iden-tify the active species using XAS and Moossbauercharacterization and/or poisoning experiments. Inaddition, the use of a large excess ofH2O2 as oxidantdecreased the catalytic efficiency to some extent inboth cases.

Silanols are important building blocks in polymerchemistry. The selective oxidation of silanes withwater to synthesize silanols has attracted increasingattention for its safety (nonexplosive, nonflammableand nontoxic) and atomic economy (hydrogen gasas the only by-product) [63]. Chen and coworkersprepared Au SACs supported by mesoporouscarbon graphitic carbon nitride (mpg–C3N4),which showed excellent catalytic performance in theoxidation of silanes by water [64]. Nevertheless, incomparison with the gold NPs supported on silicapreviously reported for the same reaction [65], theAu1/C3N4 showed only comparable activity. Oneadvantage of the Au1/C3N4 SAC is its excellentstability; it could be recovered and reused at least

10 times with its activity and selectivity beingmaintained well. The characteristic coordinationcavity surrounded by N atoms in the mpg–C3N4could anchor and stabilize the Au single atoms;this effect was believed to contribute to the highdurability of the catalyst.

Hydroformylation of olefinsThe hydroformylation of olefins involves the addi-tion of syngas (a mixture of CO and H2) to olefinsfor the production of aldehydes and represents atypical example of an efficient and clean chemicalprocess with 100% atom economy. The aldehydesare valuable final products and important inter-mediates for the synthesis of bulk chemicals suchas alcohols, esters and amines; more than 10 mil-lion tons of aldehydes are manufactured per yearglobally [66]. Cobalt-based catalysts were the firstgeneration of catalysts for these transformations;however, they required harsh reaction conditionsand had low productivity. In 1968, a Rh-basedcatalyst [RhCl(PPh3)3] was reported by Wilkinsonet al. [67], and was much more active and selectivethan the Co-based catalysts under mild reactionconditions, thus opening a new era of rhodium–alkylphosphine/phosphite complexes for thehydroformylation reaction [66]. The alkyl–phosphorous ligands, with their unique steric bulkand electron-donating effects, contribute greatlyto the high activity and selectivity. However, thesehomogeneous complexes are difficult to recover.Thus, great efforts have been devoted to the hetero-genization of Rh-based hydroformylation catalysts,including immobilization of the homogeneouscomplexes on a solid support by ion exchange,adsorption or covalent grafting [68]. However, theinteraction between the post-heterogenization Rhmetallorganics and the support is relatively weak.Degradation of the ligands and/or theRh complexestends to occur during the course of the reaction,which necessitates the periodic supplementationof the active components. Furthermore, only alow concentration of phosphorous ligands can begrafted onto the surface of the support using thesemethods, which usually results in poor catalyticactivity, selectivity and stability because of the lowP/Rh ratio [69,70].

Recently, the groups of Ding and Xiao groupshave made great progress in the heterogenizationof homogeneous Rh catalysts by designing porousorganic ligands (POLs) to support Rh [71–73].They chose the well-known electron-donating lig-and triphenylphosphine (PPh3) as the backboneof the monomer and grafted a vinyl group tothe three benzene rings in PPh3 to constructvinyl-functionalized PPh3 (3V–PPh3), which then

Page 43: Special Topic: Single-Atom Catalysts - Western Engineering...Predictive approach of heterogeneous catalysis Metal-organic frame-work assisted synthesis of single-atom catalysts for

664 Natl Sci Rev, 2018, Vol. 5, No. 5 REVIEW

PRh

P P

O OPO

OP O

O

OPP

orP PPPh3 P: :

Xantphos Biphephos

(a) (b)

PPh3

RhPh3P PPh3

Figure 7. The proposed structures of the (a) Rh/POL–PPh3 and (b) bidentate-ligand-doped Rh/POPs-PPh3 catalysts.

underwent polymerization under solvothermal con-ditions to yield a porous organic polymer (POP)with high surface area and rich porosity (Fig. 7a).One of the most significant advantages of thismethod is that the PPh3 moieties can be anchoredwith a high loading and are homogeneously dis-tributed throughout the support via covalent bond-ing (C–C). This not only stabilizes the active Rhspecies exclusively as single atoms, but also ensuresthe high P/Rh ratio that is required for high catalyticactivity and selectivity. In the hydroformylation of1-dodecene, the conversion reached 86.4% with ahigh linear/branched aldehyde ratio (l/b = 6.6).The high selectivity towards linear aldehydes wasthought to arise from the confinement effect of themicropores (0.7–1.5 nm) in the support. More im-pressively, in contrast to the vulnerability of mostpreviously reported heterogenized molecular cata-lysts [68,74], the Rh SAC was quite stable for morethan 500 h time-on-stream in continuous flow re-action, demonstrating its great potential for practi-cal applications. However, the substrate toleranceof the catalyst was unsatisfactory and, for the hy-droformylation of the shorter-chain substrate 1-octene, the linear/branched ratio was poor (45/55),and a large amount of isomerized alkene was pro-duced. To improve the regio-selectivity, the authorsdoped bidentate ligands such as Xantphos, whichwas reported by Van Leuween [75], and biphephos,which was developed by the Union Carbide Corp.[76,77], onto thePOP–PPh3 (Fig. 7b).Thesebiden-tate ligands usually exhibit large natural bite angles(∼120o) and, when coordinated with an Rh cation,the formed complex preferentially adopted a diequa-torial rather than an equatorial-apical configuration,which led to an increase in the steric congestionaround the Rh center and resulted in more selec-tive formation of the sterically less demanding linear

alkyl rhodium species and subsequently of the linearaldehydes [75]. Using a similar method, Ding andcoworkers functionalized the Xantphos and biphep-hos ligandswith vinyl groups, which then underwentco-polymerization with the 3V–PPh3 to yield thesupport [78]. The Rh species in these catalysts stillexisted as single atoms. Compared with Rh/POP–3V–PPh3, the Xantphos- and biphephos-doped cat-alysts afforded greatly enhanced regio-selectivity, al-though the catalytic activity was decreased to someextent, most likely due to the limited mass trans-fer. For example, the l/b ratio in the hydroformyla-tion of 1-octene over Xantphos-doped Rh/POPs–PPh3 increased from 45/50 to 90/10 and, in thehydroformylation of propene on biphephos-dopedcatalysts, the l/b ratio reached as high as 24.2. Im-pressively, the biphephos-doped catalyst was also ef-fective for the isomerization–hydroformylation ofinternal olefins and the regio-selectivity for linearaldehyde reached up to 92%. The authors pro-posed that the PPh3 ligands contributed mainly tothe overall catalytic activity and stability owing totheir strong electron-donating capacity and multi-ple binding with Rh cations, whereas the biden-tate ligands were mainly responsible for the highregio-selectivity due to their steric hindrance. ThePOLs strategy developed by Ding and Xiao openeda new approach for the heterogenization of met-allorganic molecular catalysts that could maintaintheir high activity and regio-selectivitywhile creatingrobustness. Nevertheless, the tedious procedures re-quired to prepare the polymers and the use of ex-pensive ligands andGrignard reagents represent bar-riers to their future industrial application. Furtherefforts must be devoted to developing facile andinexpensive synthetic routes and exploring otherP-containing materials (e.g. MOFs, COFs) for suchtransformations.

Page 44: Special Topic: Single-Atom Catalysts - Western Engineering...Predictive approach of heterogeneous catalysis Metal-organic frame-work assisted synthesis of single-atom catalysts for

REVIEW Zhang et al. 665

Given the high cost and vulnerability of the phos-phorous ligands, supported Rh NPs catalysts with-out P-ligands were developed for the hydroformyla-tion of olefins; however, their catalytic performanceswere inferior to those of their homogeneous analogs[79]. SACs, inwhich the singlemetal atomsare coor-dinated to heteroatoms on the support, are regardedas a mimic of supported homogeneous catalysts andthus are expected toperformwell in the hydroformy-lation of olefins. Indeed, the Rh SACs met some ofthe criteria for the hydroformylation of olefins. First,in most Rh homogeneous catalysts, Rh(I) speciesare considered to be the active sites; fortunately, Rhsingle atoms are generally positively charged. Sec-ond, Rh single atoms are reported to effectively ac-tivate both H2 and CO, and the isolated Rh cen-ters are believed to interact with olefins weakly with-out hydrogenation. Third, the confinement in someSACs is expected to have steric effects similar tothose of bulky P-ligands. Recently, several excitingexamples have demonstrated the feasibility of RhSACs for the hydroformylation of olefins. Lang et al.fabricated a Rh1/ZnO catalyst for the hydroformy-lation reaction [11]. They chose ZnO nanowires(ZnO-nw) as a support and a Rh SAC with a load-ing of 0.006 wt% was prepared. For comparison,ZnO-nw-supported Rh NPs were also synthesized.In the hydroformylation of styrene, when the Rhloading was decreased from 0.3 to 0.006 wt%, theturn over number (TON) increased from7× 103 to4 × 104. Notably, the TONs of Rh1/ZnO-nw wereeven higher than those of homogeneous RhCl3 andWilkinson’s catalyst RhCl(PPh3)3 (TON = 3324and 1.9× 104, respectively).Moreover, the chemos-electivity of the Rh1/ZnO-nw catalyst reached 99%,which was much higher than those of its homoge-neous counterparts (83–92%). Almost no hydro-genation of olefins occurred on the catalyst, in-dicating that the single Rh atoms interacted withthe olefins through a π -mode interaction, whichwas in good agreement with the behavior of otherSACs for hydrogenation reactions [10,42,56]. Un-fortunately, the Rh1/ZnO-nw SAC did not exhibitregio-selectivity towards the linear aldehydes; thelinear/branched ratio was about 1. Generally, highregio-selectivity can only be obtained when the Rhcenters have significant steric hindrance during theinsertion of CO.The unsatisfactory regio-selectivityof the Rh1/ZnO-nw SAC may imply that the Rhcenter had low steric congestion. XPS and in situXANES characterization results indicated that thesingleRh atomswere in a nearlymetallic state, whichimplied that the single Rh atoms might have beenlocated atop a Zn atom. This position would leavethe Rh single atoms open to CO insertion from ev-ery direction and thus lead to poor regio-selectivity.

However, high regio-selectivity and high activityover Rh SACs was achieved by fine-tuning the va-lence state and location of the single Rh atoms onthe support, as demonstrated by Zeng and cowork-ers [80]. They constructed a Rh1/CoO SAC byconfining the single Rh atoms in the CoO lay-ers via galvanic replacement between Rh(III) andCoO. Inductively coupled plasma atomic emissionspectroscopy (ICP-AES) examination andHAADF-STEM images suggested that the single Rh atomsoccupied the Co positions. In the liquid-phase hy-droformylation of propene, the 0.2% Rh/CoO SACachieved a TOF value of 2065 h−1, which was evenhigher than that of Rh/POPs–PPh3. Furthermore,the catalyst showed impressive chemoselectivity(>99%) and regio-selectivity for 1-butaldehyde(94.4%), higher than those of its nano-counterparts(68.7 and 53.9% for 1.0% Rh/CoO and 4.8%Rh/CoO). This high chemo- and regio-selectivitymay arise from two features. First, XPS measure-ments demonstrated that the single Rh atoms werepositively charged even under the reaction condi-tions, allowing the single Rh atoms to stably fill theCo(II) vacancies; that is, the Rh single atoms werecoordinated to four oxygen atoms within the CoOlayer in the equatorial directionof anoctahedral con-figuration. Second, the single Rh atoms were con-fined within the CoO sheets, although according toDFT calculations, Rh moved 1.3 A out of the planeduring the reaction. This confinement limited thepossible CO insertion pathways, which in turn pref-erentially yielded the linear aldehyde due to sterichindrance. Furthermore, the CoO support mighthave assisted in the activation of H2 and CO, as inthe case of Au/Co3O4 [81], which would increasethe catalytic activity. This work subverted the tra-ditional opinion that high regio-selectivity can onlybe obtained with the assistance of expensive phos-phorous ligands, and demonstrated that the het-eroatoms (e.g. O) in the support can fulfill the roleof the bulky P-ligands in a carefully designed SAC.

C–C coupling reactionThe construction of C–C bonds, which is a vitalmethod for the formation of complex moleculesfrom simple substrates, is one of the central themesin modern synthetic chemistry. Typical powerfulapproaches include the Suzuki cross-coupling re-action [82], Heck reaction [83] and α-alkylationof enolates [84], which are generally catalysed byPd, Ir or Pt complexes. When aryl halides are usedas substrates, the chlorides are cheaper and morereadily available than the bromides and iodides, butare more difficult to activate because of the higher

Page 45: Special Topic: Single-Atom Catalysts - Western Engineering...Predictive approach of heterogeneous catalysis Metal-organic frame-work assisted synthesis of single-atom catalysts for

666 Natl Sci Rev, 2018, Vol. 5, No. 5 REVIEW

O O

O

O

O

OCNNC F F

O O

X=Cl

X=Br

X=I

89% 97% 98%

84% 80% 73%

99% 77% 88%

99% 93% 91%O

O

90%

99%

(a) (b)

(c)

0

300

600

900

1200

TON

0 20 40 60 80 100Pd (mol%)

Au

Pd

Au-alloyed Pd single-atom

XR R RR = Cl, Br, I 14 examples

73-99% yields

Cl Cl

ClCl

Cl Cl

Cl Cl

AuPd

Figure 8. (a) Relationship between activity (turnover number, TON) and Pd concentra-tion in the Au–Pd/resin catalysts for the Ullmann reaction of chlorobenzene; (b) pro-

posed reaction mechanism; (c) substrate tolerance (the percentages in C indicate the

product yield). Adapted with permission from [85].

dissociation energy of the C–Cl bond. Zhang et al.prepared an ion-exchange-resin-supported Au–PdSAA catalyst that exhibited high catalytic activityand selectivity in the Ullmann reaction of aryl chlo-rides [85]. BothCO-DRIFTs (diffuse reflectance in-frared Fourier transform spectroscopy) and EXAFScharacterization confirmed the formation of an SAAstructure at Au/Pd ≥ 6, with the Pd atoms isolatedby the surrounding Au atoms. In the Ullmann re-action of chlorobenzene, an exponential increase ofthe TON (normalized by Pd mass) was observedwith decreasing Pd concentration (Fig. 8a). This in-teresting trend can be rationalized by the assump-tion that, with an increase in the Au/Pd ratio, thePd single atoms are mainly located at the edges andcorner positions of the Au nanoparticles (Fig. 8b),which should bemore active than those on facets be-causeof themoreunsaturated coordination environ-ment, as in the case of the aforementioned ‘crown-

jewel’-structured Au SACs [59]. Furthermore, thecatalyst exhibited promising substrate tolerance anddurability for Ullmann reactions of aryl chlorides,bromides and iodides (Fig. 8c) and could be reusedeight times without decay in their activity. Hot fil-tration experiments demonstrated a heterogeneousreaction mechanism. It was proposed that the Au inthe catalysts not only played a role in separating Pdto form single atoms, but also promoted the dissoci-ation of the C–Cl bond and the coupling of two arylgroups, thus opening a newmethod of fabricating ef-ficient alloyedSACs for other reactions. Futureworkcan be devoted to exploring other inexpensive tran-sition metals to isolate Pd, in light of the high priceof Au.

In addition to the SAA, a TiO2-supported PdSAC was also explored for the Sonogashira cou-pling reactions of aryl bromides and iodides withphenylacetylene [86]. The catalyst afforded com-plete conversion of phenylacetylene after reaction at60◦C for 3 h in the Sonogashira coupling reactionof phenylacetylene and iodobenzene, with the prod-uct selectivity reaching over 90%. The coupling re-action also proceeded readily for substituted pheny-lacetylenes with different functional groups and arylbromides and iodides, and the corresponding prod-ucts were obtained in high yields. However, for themore challenging aryl chloride substrates, the cata-lyst showed rather low activity. XAS and XPS char-acterization, in combination with DFT calculations,revealed a Pd1O4 structure in which the positivelycharged and isolated Pdδ+ species interact stronglywith four surface lattice O atoms of the TiO2 sup-port. Such a structure provides multifunctional Pd,Oad and Ti5c atoms for the activation of the reac-tants, and therefore a lower apparent activation en-ergy of 28.9 kJ/mol is required in comparison withthat of the homogeneous catalyst Pd(PPh3)2Cl2(51.7 kJ/mol), demonstrating the concerted cataly-sis by the TiO2 support and single Pd atoms.

Kim et al. reported a thiolated multiwalled nan-otube (MWNT) supported Pt SAC that acted asa good catalyst for the Suzuki cross-coupling reac-tion of 4-iodoanisole and 4-methylbenzene boronicacid [87]. The Pt-S-MWNT SAC afforded a highproduct yield of 99.5%, which wasmuch higher thanthat on Pd-S-MWNTs, Pd/C or H2PtCl6. How-ever, for bromide and chloride, much lower yields of10.4 and 4.8% were obtained, respectively. The cat-alyst could be recovered by filtration and could bereused 12 times without significant decrease in theyield. XANES spectra showed a white-line intensityslightly higher than that of Pt foil, but much lowerthan that of H2PtCl6, suggesting that the single Ptsingle atoms were positively charged. The positivelycharged single Pt atoms could easily dissociate the

Page 46: Special Topic: Single-Atom Catalysts - Western Engineering...Predictive approach of heterogeneous catalysis Metal-organic frame-work assisted synthesis of single-atom catalysts for

REVIEW Zhang et al. 667

C–I bond and were believed to be responsible forthe high activity. It was also noted that, in the Pt-S-MWNTSACs, the single Pt atomswere coordinatedto the thiol groups, which not only resulted in a posi-tive charge on the single Pt atoms, but also helped tostabilize them against aggregation and leaching.

Recently, Wang et al. constructed a PdII@PDMSsingle-atom catalyst for the oxidative coupling ofbenzene to synthesize biphenyl compounds usingO2 as the terminal oxidant with only H2O as a by-product [88]. In this catalyst, the porous organiccopolymer PDMS (a copolymer of divinylbenzene,maleic anhydride and p-styrene sulfonate) contain-ing abundant carboxylic acid and sulfonate groupswas used as the support for anchoring single Pdatoms. In the oxidative coupling reaction of ben-zene, a 26.1% yield of biphenyl with a high selec-tivity of 98.3% was obtained after 4 h at 120◦C, af-fording a TON value of 363. Under the optimizedreaction conditions, an even higher TON of 487was obtained, which surpassed those achieved onPd(OAc)2 and other reported heterogeneous cata-lysts. Benzene, toluene, o-xylene, m-xylene and p-xylene were also good substrates, and the corre-sponding products were obtained with high selec-tivity (90.8–97.6%) and yields (10.3–30.8%). Hotfiltration experiment results revealed that the re-action did not proceed after the catalyst was re-moved, thus ruling out the possible contribution ofleached Pd to the catalysis. The bidentate –COOHin the PdII@PDMS catalyst contributed greatly tothe stabilization of Pd(II) against reduction as wellas the subsequent catalytic activity; in contrast, forcatalysts with both –SO3H and mono-carboxylicgroups, or those possessing solely –COOH or−SO3H groups, much lower yields (0.2–14.2%)were obtained because of their weaker ability to sta-bilize Pd(II). Evidently, the isolated Pd(II) specieswas the active site for this reaction.

STABILITY OF SACSOne of the greatest concerns regarding the use ofSACs for liquid-phase reactions is their stability.Generally, leaching (the detachment of the singleatoms from the support) and aggregation (the for-mation of clusters and NPs from the single atomsthrough migration) are the two major factors thatseverely deteriorate the stability, and the stability ofthe SACs is strongly dependent on the interactionbetween the single atoms and the support.

Free single atoms are known to have an ex-tremely high surface energy and therefore cannotbe stable. In contrast, upon being fixed on a sup-port, whether on the surface, in the cation position

(lattice substitution), confined in a cavity or alloyedwith another metal, the single atoms tend to inter-act strongly with the support via chemical bondingwith the donor atoms of the support, such as oxy-gen or nitrogen atoms or a second metal atom inthe case of SAA; these atoms are analogous to theligands of organometallic complexes. This bondinglowers the surface energy and sometimes even hascovalent characteristics [9]. The strong ionic or co-valent bonding of the single atoms with the supportsignificantly improves their stability against leachingand/or aggregation. Here, we provide several exam-ples that demonstrate the superior stability of SACsin biomass conversions.

The transformation of biomass into fuelsand value-added chemicals is an important andactive research area. Biomass-related hydrogena-tion/hydrodeoxygenation reactions often requireharsh reaction conditions, such as elevatedhydrogenpressure, moderate to high reaction temperaturesand hot water or even acidic reaction media. Forexample, in the hydrogenation of levulinic acid (LA)to γ -valerolactone (GVL), commercially availableRu/C has proven to be highly active and selective.However, it suffers from severe deactivation underhydrothermal and/or acidic reaction conditions dueto Ru leaching [89]. Wang, Zhang and coworkersreported a single-atom Ru/ZrO2@C (0.85 wt%)catalyst that was highly active and ultra-stable forthe hydrogenation reaction of LA [90]. When thereaction was performed in water, the conversionof LA dropped from 69.2% in the first run to 40%in the third cycle on Ru/C, whereas the conver-sion remained unchanged even after six runs overRu/ZrO2@C. The stability of Ru/ZrO2@C wasfurther demonstrated under harsher conditions(pH=1); theLAconversiondecreased from77.7 to25% on Ru/C after three runs, whereas no apparentdrop of activity was observed over Ru/[email protected] results indicated that 14.8 and 7.4% ofthe Ru leached into acid media and H2O for Ru/C,respectively, which caused the severe deactivationof the catalyst. In contrast, no leaching of Ru wasdetected for Ru/ZrO2@C, which was attributedto the strong interaction between Ru and ZrO2.Similar results were also reported by other groups[91], although the unambiguous identification ofsingle atoms of Ru on ZrO2 remains a challenge dueto the similar atomic numbers of the two elements.

Another class of ultra-stable SACs are the M–N–C catalysts, in which M is exclusively dispersedas single atoms and bonds with the nearby Natoms to form M–Nx active sites. The M–Nx struc-ture is expected to be highly resistant against high-temperature aggregation and acid leaching based onthe strong bonding between the M cation and N

Page 47: Special Topic: Single-Atom Catalysts - Western Engineering...Predictive approach of heterogeneous catalysis Metal-organic frame-work assisted synthesis of single-atom catalysts for

668 Natl Sci Rev, 2018, Vol. 5, No. 5 REVIEW

Ni-N-C-fresh Ni-N-C-used

Ni/AC-usedNi/AC-fresh

(a)(b)

(c)

HxWO4HydrolysisO

OO

OH

OHO

OHOH

HO OH

O

n

OHO

OH

HOOH

OH HO OH ......HO

ONi-N-C

Cellulose Glucose Glycolaldehyde Ethylene glycol

EG1,2-BDO

HAOthers

iso-SorEry

GlyMan

1,2-PG

1st 2nd 1st 2nd3rd 4th 5th 6th 7th

Ni-N-C SAC

Wei

ght y

ield

(%)

Ni/AC

100

90

80

70

60

50

40

30

20

10

0

Figure 9. (a) Stability tests of the Ni–N–C catalyst and Ni/AC catalyst, and HAADF-STEM images of fresh and used (b) Ni–N–C and (c) Ni/AC catalysts.

Adapted with permission from [13].

atoms. As expected, Wang, Li and coworkers re-cently reported an ultra-stable Ni–N–C SAC with aNi loading of 7.5 wt% that exhibited excellent per-formance in the one-pot conversion of cellulose toEG [13], which is an important reaction for thevalorization of biomass to value-added chemicals[92–94]. The Ni–N–C SAC exhibited superior sta-bility compared to its Ni/ACNP counterpart underrelatively harsh reaction conditions (245◦C, 60 barH2, presence of tungstic acid in hot water); it couldbe recycled seven times without any deactivation,while the Ni/AC lost half of its initial activity dur-ing the second run (Fig. 9). Characterizations ofthe used catalysts showed severe agglomeration ofthe Ni nanoparticles whereas the Ni–N–C catalystmaintained atomic dispersion. In another biomass-valorization reaction, an MoS2 monolayer dopedwith single-atom Co was reported to show superioractivity and stability for the hydrodeoxygenation of4-methylphenol to toluene [95], which is a modelreaction for lignin transformations [96]. The singleCo atomswere believed to fill S-atomvacancies to beimmobilized as a part of the basal plane.The durabil-itywas greatly improved—that is, the catalysts couldbe reused at least seven times for a total reaction timeof 56 h without decay in their activity and selectiv-ity;moreover, no sulfur detachmentwas detected by

ICP-AES. This work demonstrated that, when sin-gle atoms—100% dispersion of active metal—meet with single layer sheets—the ultimate exfo-liation of 2D materials—powerful catalysts can becreated.

The above examples, among others, clearlydemonstrate that single atoms that are stronglychemically bonded with the donor atoms of thesupport can behave even more stably than NPs incertain reactions.

CONCLUSION AND PERSPECTIVESACs are of great interest and importance for thedevelopment of a new generation of low-cost, ef-ficient and robust catalysts. Especially in the fieldof the green synthesis of fine chemicals, variousSACs have been fabricated, and their catalytic po-tentials have been exploited in a variety of organictransformations, including, but not limited to, selec-tive hydrogenation, oxidation, hydrogenolysis, hy-drodeoxygenation, hydroformylation and C–C cou-pling reactions. In particular, some reactions thatrepresent very important industrial processes, suchas the hydrogenolysis of glycerol to 1,3-propanedioland the hydroformylation of propene, would be

Page 48: Special Topic: Single-Atom Catalysts - Western Engineering...Predictive approach of heterogeneous catalysis Metal-organic frame-work assisted synthesis of single-atom catalysts for

REVIEW Zhang et al. 669

very profitable if efficient and robust SACs could besuccessfully exploited. Traditional homogeneous orheterogeneousNP catalysts involve compromises interms of activity, selectivity or recyclability. SACs in-tegrate the merits of both these types of catalysts—that is, the unsaturated coordination environmentof the single atoms imparts superior catalytic ac-tivity per metal atom; the uniform structure of theSACs results in unparalleled selectivity for the de-sired products; and strong covalent or electronic in-teractions with the support or another metal pro-vide excellent stability under liquid-phase operatingconditions. Therefore, in the field of green chemicalsynthesis, one can expect further progress in the useof SACs to bridge homogeneous and heterogeneouscatalysis.

In spite of its exciting andencouragingbeginning,the SAC field is still in its infancy. The following is-sues must be addressed to obtain an in-depth un-derstanding of single-atom catalysis and eventuallyachieve the rational design of SACs for specific or-ganic transformations.

(1) The role of the support must be further clar-ified. In SACs, the single atoms of the active metalinteract strongly with the support, and the supportserves as the ligands of the active metal centers,just as in homogeneous organometallic molecules.Therefore, the properties of the support can greatlyaffect the chemical state and coordination structureof the single atoms, and thus their ultimate catalyticperformance. By tuning the properties of the sup-port, for example, by grafting functional groups ontothe support surface or changing the concentrationof defect sites via thermal treatment, the electronicand geometric properties of the central single metalatoms can be changed, providing an effective wayto tune the catalytic performance of the SACs. Thesingle atoms may be located in the cation vacancies,the anion vacancies or atop the cation sites of theoxide/sulfide support, which results in different lo-cal structures around the single central single atoms,and thus different catalytic performance. For exam-ple, in the Co1/MoS2 catalyst developed by Tsanget al. [95], DFT calculations revealed that, when theCo single atom was located in the sulfur vacancieswithin the MoS2 monolayer, the catalyst exhibitedthe highest stability. In addition, the supportmay di-rectly participate in the reaction in concert with thesingle atoms. In such cases, the choice and the mod-ification of the support become key factors in thesuccess of the targeted reaction. Therefore, the elu-cidation of the multifarious roles played by the sup-port will definitely help to develop an understandingthe mechanism of single-atom catalysis and to de-velop efficient and robust SACs for green chemicalsynthesis.

(2) Multifunctional SACs should be developed.The green synthesis of fine chemicals calls forthe combination of multi-step syntheses into one-pot tandem (domino) reactions. Thus, great effortshould be devoted to the development of SACs withmultiple functions; for example, the development ofa dehydration–hydrogenation bi-functional catalystthrough the use of an acidic support might be ex-plored [44,45].

(3) Construction of SACs with a high surfacedensity of the active metal species must be achieved.In order to ensure atomic dispersion, most of theSACs reported to date have had rather low surfacedensities of the active species (e.g.<0.5wt%),whichleads to a low-volume productivity, although theatom utilization is maximized. For practical appli-cations, SACs with a high surface density of the ac-tive metal species are required; this can be accessedby engineering the metal–support interaction. Thefabrication of support materials with a high den-sity of defect sites will be favorable considering thatthe single atoms are usually anchored to the de-fect sites of the support (vacancies, coordination-unsaturated sites, etc.). For example, in the Pt/WOxsystem [44,45,97], the mesoporous WOx preparedby the alcoholysis method is rich in oxygen vacan-cies, which allows the Pt species interact stronglywith the support and thereby be dispersed as sin-gle/pseudo single atoms even at a high loading of∼2.6 wt%. Nano-engineering of the support is an-other useful strategy to create abundant surface de-fects.When the supportmaterial is downsized to thenanoscale, both the surface area and density of de-fect sites will increase greatly, which should be favor-able to the incorporation ofmore single atoms of theactive metal. Typical examples include graphene-supported-Fe [17], MoS2-monolayer-confined Co[95],CoO-nanosheet-supportedRh [80] andTiO2-nanosheet-anchored Pd [42], all of which featurea practically high density of single atoms. Further-more, because these 2D materials have distinctphysical and chemical properties from their bulkcounterparts, they might provide unique catalyticperformance when loaded with single atoms. Insomecases, dopingheteroatoms into the support hasalso been found tobe effective to stabilize a highden-sity of single atoms. For example, the sodium cationsin thePt–Na/FeOx SACswere found to enhance thedispersion and stabilization of Pt single atoms with ahigh loading (2.16 wt%) [32]; doping N atoms intocarbon sheets can be effective to achieve a high load-ing of single Co (3.6 wt%) [12], Fe (1.0 wt%) [16]and Ni (7.5 wt%) [13] atoms.

Lastly, more in situ characterization techniquesshould be developed to monitor the dynamic struc-ture of single atoms under operating conditions, as

Page 49: Special Topic: Single-Atom Catalysts - Western Engineering...Predictive approach of heterogeneous catalysis Metal-organic frame-work assisted synthesis of single-atom catalysts for

670 Natl Sci Rev, 2018, Vol. 5, No. 5 REVIEW

this would be helpful for the understanding of thecatalytic mechanism of SACs and the developmentof more efficient and robust SACs for the green syn-thesis of fine chemicals. Operando characterizationis particularly challenging for liquid-phase reactionsdue to the involvement of solvent molecules, but itis of vital importance to efforts to establish a bridgebetween homogeneous and heterogeneous catalysisby the use of SACs.

FUNDINGThis work was supported by the National Natural ScienceFoundation of China (21373206, 21690080, 21690084,21503219 and 21673228), the Strategic Priority Research Pro-gram of the Chinese Academy of Sciences (XDB17020100) andthe National Key Projects for Fundamental Research and Devel-opment of China (2016YFA0202801).

REFERENCES1. Pollak P. Fine Chemicals: the Industry and the Business. Hobo-

ken: John Wiley & Sons, 2011.

2. Joshi SS and Ranade VV. Industrial Catalytic Processes for Fine

and Specialty Chemicals. Amsterdam: Elsevier, 2016.

3. Merino E. Synthesis of azobenzenes: the coloured pieces of

molecular materials. Chem Soc Rev 2011; 40: 3835–53.4. Blaser HU, Steiner H and Studer M. Selective catalytic hydro-

genation of functionalized nitroarenes: an update. Chem Cat

Chem 2009; 1: 210–21.5. Polshettiwar V and Varma RS. Green chemistry by nano-

catalysis. Green Chem 2010; 12: 743–54.6. Trost BM. Atom economy—a challenge for organic synthesis:

homogeneous catalysis leads the way. Angew Chem Int Ed Engl

1995; 34: 259–81.7. Beller M, Cornils B and Frohning CD et al. Progress in hydro-

formylation and carbonylation. J Mol Catal A 1995; 104: 17–85.8. Qiao B, Wang A and Yang X et al. Single-atom catalysis of CO

oxidation using Pt1/FeOx. Nat Chem 2011; 3: 634–41.9. Yang X, Wang A and Qiao B et al. Single-atom catalysts: a new

frontier in heterogeneous catalysis. Acc Chem Res 2013; 46:1740–8.

10. Wei H, Liu X andWang A et al. FeOx-supported platinum single-

atom and pseudo-single-atom catalysts for chemoselective hy-

drogenation of functionalized nitroarenes. Nat Commun 2014;

5: 5634.11. Lang R, Li T and Matsumura D et al. Hydroformylation of olefins

by a rhodium single-atom catalyst with activity comparable to

RhCl(PPh3)3. Angew Chem Int Ed 2016; 55: 16054–8.12. Liu W, Zhang L and Yan W et al. Single-atom dispersed Co–N–

C catalyst: structure identification and performance for hydro-

genative coupling of nitroarenes. Chem Sci 2016; 7: 5758–64.13. Liu W, Chen Y and Qi H et al. Ultra-durable Ni−N−C single-

atom catalyst for biomass conversion under harsh conditions.

Angew Chem Int Ed 2018; 57: 7071–5.

14. Xie J, Yin K and Serov A et al. Selective aerobic oxidation of al-

cohols over atomically-dispersed non-precious metal catalysts.

Chem Sus Chem 2017; 10: 359–62.15. Li M, Wu S and Yang X et al. Highly efficient single atom cobalt

catalyst for selective oxidation of alcohols. Appl Catal A Gen

2017; 543: 61–6.16. Liu W, Zhang L and Liu X et al. Discriminating catalytically ac-

tive FeNx species of atomically dispersed Fe-N-C catalyst for

selective oxidation of the C-H bond. J Am Chem Soc 2017; 139:10790–8.

17. Deng D, Chen X and Yu L et al. A single iron site confined in a

graphene matrix for the catalytic oxidation of benzene at room

temperature. Sci Adv 2015; 1: e1500462.18. Zhang M, Wang YG and Chen W et al. Metal (hydr) oxides@

polymer core-shell strategy to metal single-atom materials. J

Am Chem Soc 2017; 139: 10976–9.19. WangA, Li J and Zhang T. Heterogeneous single-atom catalysis.

Nat Rev Chem 2018; 2: 65–81.20. Liu J. Catalysis by supported single metal atoms. ACS Catal

2017; 7: 34–59.21. Qin R, Liu P and Fu G et al. Strategies for stabilizing atomically

dispersed metal catalysts. Small Methods 2018; 2: 1700286.22. Zhang H, Liu G and Shi L et al. Single-atom catalysts: emerg-

ing multifunctional materials in heterogeneous catalysis. Adv

Energy Mater 2018; 8: 1701343.23. Asokan C, DeRita L and Christopher P. Using probe molecule

FTIR spectroscopy to identify and characterize Pt-group

metal based single atom catalysts. Chin J Catal 2017; 38:1473–80.

24. Han B, Lang R and Qiao B et al. Highlights of the major progress

in single-atom catalysis in 2015 and 2016. Chin J Catal 2017;

38: 1498–507.25. Nishimura S. Handbook of Heterogeneous Catalytic Hydrogena-

tion for Organic Synthesis. New York: Wiley, 2001.

26. Downing RS, Kunkeler PJ and van Bekkum H. Catalytic synthe-

ses of aromatic amines. Catal Today 1997; 37: 121–36.27. Ren Y,Wei H and Yin G et al.Oxygen surface groups of activated

carbon steer the chemoselective hydrogenation of substituted

nitroarenes over nickel nanoparticles. Chem Commun 2017; 53:1969–72.

28. Siegrist U, Baumeister P and Blaser HU et al. Chemical Indus-

tries. New York: Marcel Dekker, 1998, 207–20.

29. Corma A, Serna P and Concepcioon P et al. Transforming non-

selective into chemoselective metal catalysts for the hydro-

genation of substituted nitroaromatics. J Am Chem Soc 2008;

130: 8748–53.30. Zhai Y, Pierre D and Si R et al. Alkali-stabilized Pt-OHx species

catalyze low-temperature water-gas shift reactions. Science

2010; 329: 1633–6.31. Yang M, Li S and Wang Y et al. Catalytically active Au-O(OH)x-

species stabilized by alkali ions on zeolites and mesoporous ox-

ides. Science 2014; 346: 1498–501.32. Wei H, Ren Y and Wang A et al. Remarkable effect of alkalis on

the chemoselective hydrogenation of functionalized nitroarenes

over high-loading Pt/FeOx catalysts. Chem Sci 2017; 8:5126–31.

Page 50: Special Topic: Single-Atom Catalysts - Western Engineering...Predictive approach of heterogeneous catalysis Metal-organic frame-work assisted synthesis of single-atom catalysts for

REVIEW Zhang et al. 671

33. Xu G, Wei H and Ren Y et al. Chemoselective hydrogenation of 3-nitrostyrene

over a Pt/FeOx pseudo-single-atom-catalyst in CO2-expanded liquids. Green

Chem 2016; 18: 1332–8.34. Zhang S, Chang CR and Huang ZQ et al. High catalytic activity and chemose-

lectivity of sub-nanometric Pd clusters on porous nanorods of CeO2 for hydro-

genation of nitroarenes. J Am Chem Soc 2016; 138: 2629–37.35. Corma A and Serna P. Chemoselective hydrogenation of nitro compounds with

supported gold catalysts. Science 2006; 313: 332–4.36. Wang L, Guan E and Zhang J et al. Single-site catalyst promoters accelerate

metal-catalyzed nitroarene hydrogenation. Nat Commun 2018; 9: 1362.37. Makosch M, Sa J and Kartusch C et al. Hydrogenation of nitrobenzene over

Au/MeOx catalysts-a matter of the support. Chem Cat Chem 2012; 4: 59–63.38. Zitolo A, Goellner V and Armel V et al. Identification of catalytic sites for oxygen

reduction in iron- and nitrogen-doped graphene materials. Nat Mater 2015; 14:937–42.

39. Westerhaus FA, Jagadeesh RV andWienhofer G et al.Heterogenized cobalt ox-

ide catalysts for nitroarene reduction by pyrolysis of molecularly defined com-

plexes. Nat Chem 2013; 5: 537–43.40. Jagadeesh RV, Surkus AE and Junge H et al. Nanoscale Fe2O3-based cata-

lysts for selective hydrogenation of nitroarenes to anilines. Science 2013; 342:1073–6.

41. Zhang L,Wang A andWangW et al. Co-N-C catalyst for C-C coupling reactions:

on the catalytic performance and active sites. ACS Catal 2015; 5: 6563–72.42. Liu P, Zhao Y and Qin R et al. Photochemical route for synthesizing atomically

dispersed palladium catalysts. Science 2016; 352: 797–800.43. Zhang B, Asakura H and Zhang J et al. Stabilizing a platinum1 single-atom cata-

lyst on supported phosphomolybdic acid without compromising hydrogenation

activity. Angew Chem Int Ed 2016; 55: 8319–23.44. Wang J, Zhao X and Lei N et al. Hydrogenolysis of glycerol to 1,3-propanediol

under low hydrogen pressure over WOx-supported single/pseudo-single atom

Pt catalyst. Chem Sus Chem 2016; 9: 784–90.45. Zhao X, Wang J and Yang M et al. Selective hydrogenolysis of glycerol to

1,3-propanediol: manipulating the frustrated lewis pairs by introducing gold

to Pt/WOx. Chem Sus Chem 2017; 10: 819–24.46. Stephan DW. Frustrated Lewis pairs: from concept to catalysis. Acc Chem Res

2015; 48: 306–16.47. Kyriakou G, Boucher MB and Jewell AD et al. Isolated metal atom geometries

as a strategy for selective heterogeneous hydrogenations. Science 2012; 335:1209–12.

48. Pei GX, Liu XY and Wang A et al. Promotional effect of Pd single atoms on Au

nanoparticles supported on silica for the selective hydrogenation of acetylene

in excess ethylene. New J Chem 2014, 38: 2043–51.49. Pei GX, Liu XY and Wang A et al. Ag alloyed Pd single-atom catalysts for effi-

cient selective hydrogenation of acetylene to ethylene in excess ethylene. ACS

Catal 2015; 5: 3717–25.50. Pei GX, Liu XY and Yang X et al. Performance of Cu-alloyed Pd single-atom cat-

alyst for semihydrogenation of acetylene under simulated front-end conditions.

ACS Catal 2017; 7: 1491–500.51. Zhou H, Yang X and Li L et al. PdZn intermetallic nanostructure with Pd-Zn-Pd

ensembles for highly active and chemoselective semi-hydrogenation of acety-

lene. ACS Catal 2016; 6: 1054–61.52. Feng Q, Zhao S and Wang Y et al. Isolated single-atom Pd sites in intermetallic

nanostructures: high catalytic selectivity for semihydrogenation of alkynes. J

Am Chem Soc 2017; 139: 7294–301.53. Lucci FR, Liu J and Marcinkowski MD et al. Selective hydrogenation of 1,3-

butadiene on platinum–copper alloys at the single-atom limit. Nat Commun

2015; 6: 8550.

54. Yan H, Cheng H and Yi H et al. Single-atom Pd1/graphene catalyst achieved by

atomic layer deposition: remarkable performance in selective hydrogenation of

1,3-butadiene. J Am Chem Soc 2015; 137: 10484–7.55. Huang X, Xia Y and Cao Y et al. Enhancing both selectivity and coking-

resistance of a single-atom Pd1/C3N4 catalyst for acetylene hydrogenation.

Nano Res 2017; 10: 1302–12.56. Vile G, Albani D and Nachtegaal M et al. A stable single-site palladium catalyst

for hydrogenations. Angew Chem Int Ed 2015; 54: 11265–9.57. Hackett SFJ, Brydson RM and Gass MH et al. High-activity, single-site meso-

porous Pd/Al2O3 catalysts for selective aerobic oxidation of allylic alcohols.

Angew Chem Int Ed 2007; 46: 8593–6.58. Li T, Liu F and Tang Y et al.Maximizing the number of interfacial sites in single-

atom catalysts for the highly selective, solvent-free oxidation of primary alco-

hols. Angew Chem Int Ed 2018; 57: 7795–9.59. Zhang H, Watanabe T and Okumura M et al. Catalytically highly active top gold

atom on palladium nanocluster. Nat Mater 2012; 11: 49–52.60. Xie S, Tsunoyama H and Kurashige W et al. Enhancement in aerobic alcohol

oxidation catalysis of Au25 clusters by single Pd atom doping. ACS Catal 2012;

2: 1519–23.61. Song G, Wang F and Li X. C–C, C–O and C–N bond formation via rhodium(iii)-

catalyzed oxidative C–H activation. Chem Soc Rev 2012; 41: 3651–78.62. Sahu S and Goldberg DP. Activation of dioxygen by iron and manganese com-

plexes: a heme and nonheme perspective. J Am Chem Soc 2016; 138: 11410–28.

63. Chandrasekhar V, Boomishankar R and Nagendran S. Recent developments

in the synthesis and structure of organosilanols. Chem Rev 2004; 104:5847–910.

64. Chen Z, Zhang Q and Chen W et al. Single-site AuI catalyst for silane oxidation

with water. Adv Mater 2018; 30: 1704720.65. Li W, Wang A and Yang X et al. Au/SiO2 as a highly active catalyst

for the selective oxidation of silanes to silanols. Chem Commun 2012; 48:9183–5.

66. Franke R, Selent D and Boorner A. Applied hydroformylation. Chem Rev 2012;

112: 5675–732.67. Evans D, Osborn JA and Wilkinson G. Hydroformylation of alkenes by use of

rhodium complex catalysts. J Chem Soc, A 1968; 3133–42.

68. Ford ME and Premecz JE. Preparation and evaluation of ion exchange resin-

immobilized rhodium-phosphine hydroformylation catalysts. J Mol Catal 1983;

19: 99–112.69. Li C, Wang W and Yan L et al. A mini review on strategies for heterogenization

of rhodium-based hydroformylation catalysts. Front Chem Sci Eng 2018; 12:113–23.

70. Sun Q, Dai Z and Meng X et al. Enhancement of hydroformylation performance

via increasing the phosphine ligand concentration in porous organic polymer

catalysts. Catal Today 2017; 298: 40–5.71. Sun Q, Dai Z and Meng X et al. Task-specific design of porous polymer het-

erogeneous catalysts beyond homogeneous counterparts. ACS Catal 2015; 5:4556–67.

72. JiangM, Yan L and Ding Y et al. Ultrastable 3V-PPh3 polymers supported single

Rh sites for fixed-bed hydroformylation of olefins. J Mol Catal A: Chem 2015;

404: 211–7.73. Li C, Yan L and Lu L et al. Single atom dispersed Rh-biphephos&PPh3@porous

organic copolymers: highly efficient catalysts for continuous fixed-bed hydro-

formylation of propene. Green Chem 2016; 18: 2995–3005.74. Huang L and Kawi S. An active and stable Wilkinson’s complex-derived SiO2-

tethered catalyst via an amine ligand for cyclohexene hydroformylation. Catal

Lett 2004; 92: 57–62.

Page 51: Special Topic: Single-Atom Catalysts - Western Engineering...Predictive approach of heterogeneous catalysis Metal-organic frame-work assisted synthesis of single-atom catalysts for

672 Natl Sci Rev, 2018, Vol. 5, No. 5 REVIEW

75. van der Veen LA, Keeven PH and Schoemaker GC et al. Origin of the bite angle

effect on rhodium diphosphine catalyzed hydroformylation. Organometallics

2000; 19: 872–83.76. Billig E, Abatjoglou AG and Bryant DR. (Union Carbide Corp.) Bis-phosphite com-

pounds. 1987; Eur. Pat. EP0213639.

77. Billig E, Abatjoglou AG and Bryant DR. (Union Carbide Corp.) Bis-phosphite com-

pounds. 1988; US Pat. 4748261.

78. Li C, Sun K andWangW et al. Xantphos doped Rh/POPs-PPh3 catalyst for highly

selective long-chain olefins hydroformylation: chemical and DFT insights into

Rh location and the roles of Xantphos and PPh3. J Catal 2017; 353: 123–32.79. Liu J, Yan L and Ding Y et al. Promoting effect of Al on tethered ligand-modified

Rh/SiO2 catalysts for ethylene hydroformylation. Appl Catal, A 2015; 492: 127–32.

80. Wang L, Zhang W and Wang S et al. Atomic-level insights in optimizing reac-

tion paths for hydroformylation reaction over Rh/CoO single-atom catalyst.Nat

Commun 2016; 7: 14036.81. Liu X, Hu B and Fujimoto K et al. Hydroformylation of olefins by Au/Co3O4 cat-

alysts. Appl Catal, B 2009; 92: 411–21.82. Miyaura N, Yanagi T and Suzuki A. The palladium-catalyzed cross-coupling re-

action of phenylboronic acid with haloarenes in the presence of bases. Synth

Commun 1981; 11: 513–9.83. Heck RF. Palladium-catalyzed vinylation of organic halides. Org React 1982; 27:

345–90.

84. Stork G, Rosen P and Goldman NL. The α-alkylation of enolates from the

lithium-ammonia reduction ofα,β-unsaturated ketones. J Am Chem Soc 1961;

83: 2965–6.85. Zhang L, Wang A andMiller JT et al. Efficient and durable Au alloyed Pd single-

atom catalyst for the Ullmann reaction of aryl chlorides in water. ACS Catal

2014; 4: 1546–53.86. Zhang X, Sun Z and Wang B et al. C-C coupling on single-atom based hetero-

geneous catalyst. J Am Chem Soc 2018; 140: 954–62.

87. Lee EK, Park SA and Woo H et al. Platinum single atoms dispersed on carbon

nanotubes as reusable catalyst for Suzuki coupling reaction. J Catal 2017; 352:388–93.

88. Liu Y, Zhou Y and Li J et al. Direct aerobic oxidative homocoupling of benzene

to biphenyl over functional porous organic polymer supported atomically dis-

persed palladium catalyst. Appl Catal B 2017; 209: 679–88.89. Yan Z, Lin L and Liu S. Synthesis of γ -valerolactone by hydrogenation of

biomass-derived levulinic acid over Ru/C catalyst. Energy Fuels 2009; 23: 3853–8.

90. Cao W, Luo W and Ge H et al. UiO-66 derived Ru/ZrO2@C as a highly stable

catalyst for hydrogenation of levulinic acid to γ -valerolactone. Green Chem

2017; 19: 2201–11.91. Ftouni J, Munnoz-Murillo A and Goryachev A et al. ZrO2 is preferred over

TiO2 as support for the Ru-catalyzed hydrogenation of levulinic acid to γ -

valerolactone. ACS Catal 2016; 6: 5462–72.92. Ji N, Zhang T and Zheng M et al. Direct catalytic conversion of cellulose into

ethylene glycol using nickel-promoted tungsten carbide catalysts.AngewChem

2008; 120: 8638–41.93. Pang J, Zheng M and Sun R et al. Synthesis of ethylene glycol and terephthalic

acid from biomass for producing PET. Green Chem 2016; 18: 342–59.94. Wang A and Zhang T. One-pot conversion of cellulose to ethylene gly-

col with multifunctional tungsten-based catalysts. Acc Chem Res 2013; 46:1377–86.

95. Liu G, Robertson AW and Li MMJ et al. MoS2 monolayer catalyst doped with

isolated Co atoms for the hydrodeoxygenation reaction. Nat Chem 2017; 9:810–6.

96. Li C, Zhao X and Wang A et al. Catalytic transformation of lignin for the pro-

duction of chemicals and fuels. Chem Rev 2015; 115: 11559–624.97. YangM, Zhao X and Ren Y et al. Pt/Nb-WOx for the chemoselective hydrogenol-

ysis of glycerol to 1,3-propanediol: Nb dopant pacifying the over-reduction of

WOx supports. Chin J Catal 2018; 39: 1027–37.

Page 52: Special Topic: Single-Atom Catalysts - Western Engineering...Predictive approach of heterogeneous catalysis Metal-organic frame-work assisted synthesis of single-atom catalysts for

REVIEW National Science Review

5: 673–689, 2018

doi: 10.1093/nsr/nwy056

Advance access publication 1 June 2018

MATERIALS SCIENCE

Special Topic: Single-Atom Catalysts

Recent advances in the precise control of isolatedsingle-site catalysts by chemical methodsZhijun Li1,†, Dehua Wang3,†, Yuen Wu1,∗ and Yadong Li1,2,∗

1Department of

Chemistry, iChEM

(Collaborative

Innovation Center of

Chemistry for Energy

Materials), University

of Science and

Technology of China,

Hefei 230026, China;2Department of

Chemistry, Tsinghua

University, Beijing

100084, China and3School of

Pharmaceutical and

Chemical Engineering,

Taizhou University,

Taizhou 318000, China

∗Correspondingauthors. E-mails:[email protected];

[email protected]†Equally contributedto this work.

Received 21 March

2018; Revised 10May 2018; Accepted31 May 2018

ABSTRACTThe search for constructing high-performance catalysts is an unfailing topic in chemical fields. Recently, wehave witnessed many breakthroughs in the synthesis of single-atom catalysts (SACs) and their applicationsin catalytic systems.They have shown excellent activity, selectivity, stability, efficient atom utilization andcan serve as an efficient bridge between homogeneous and heterogenous catalysis. Currently, most SACsare synthesized via a bottom-up strategy; however, drawbacks such as the difficulty in accessing high massactivity and controlling homogeneous coordination environments are inevitably encountered, restrictingtheir potential use in the industrial area. In this regard, a novel top-down strategy has been recentlydeveloped to fabricate SACs to address these practical issues.Themetal loading can be increased to 5% andthe coordination environments can also be precisely controlled.This review highlights approaches to thechemical synthesis of SACs towards diverse chemical reactions, especially the recent advances in improvingthe mass activity and well-defined local structures of SACs. Also, challenges and opportunities for the SACswill be discussed in the later part.

Keywords: single-atom catalysts, bottom-up, top-down, catalytic performance

INTRODUCTIONIn the worldwide theme of exploring efficient andlow-cost technologies for energy conversion andchemical transformations, substantial effort hasbeendevoted to the development of general, practical andsimple chemical approaches for catalyst preparationin past decades [1–6]. Studies have shown that ul-trasmall assemblies, compared to their macroscopiccounterparts [7], can exhibit essentially differentphysical and chemical properties. These uniqueproperties would drastically alter their practical ap-plications in a variety of areas, such as cataly-sis, biomedical research, energy and environmentalfields [6]. Therefore, metal nanoparticles representa rich resource for a variety of chemical processes,employed both in industry and in academia [8].The maximized surface area of support, increasednumber of catalytic active sites, minimized catalystloading and strong catalyst-support interaction de-termine the nature of nanocatalysts [2,6]. The sup-ported metal nanoparticles are frequently employed

in heterogeneous catalysis; however, to greatly in-crease the turnover frequency of surface active sitesand to enhance themass activity remain the primarygoals in catalysis [9,10]. In most circumstances, ithas been demonstrated that the surface atoms of thenanomaterials in an unsaturated coordination envi-ronment generally act as the active sites to catalysespecific reactions [11]. Therefore, extensive stud-ies have been devoted to rationally controlling theshapes, structures, crystal phases and compositionsof nanocatalysts [3,8,12–18].

With decreasing the size of nanomaterials, thenumber of surface atoms is increased substantially,exposing more defects and active sites, tuninggeometric and electronic properties involved inchemical reactions [10]. Nanoclusters have shownintriguing properties because of the reduced sizecompared to nanoparticles, exposing more uncoor-dinated active sites and changing molecular orbitalenergy levels [19,20]. Heiz and co-workers founda pronounced size effect for model catalysts of

C©TheAuthor(s) 2018. Published by Oxford University Press on behalf of China Science Publishing &Media Ltd. All rights reserved. For permissions, please e-mail:[email protected]

Page 53: Special Topic: Single-Atom Catalysts - Western Engineering...Predictive approach of heterogeneous catalysis Metal-organic frame-work assisted synthesis of single-atom catalysts for

674 Natl Sci Rev, 2018, Vol. 5, No. 5 REVIEW

size-selected Pdn (n ≤ 30) clusters supported onMgO(100) for the cyclotrimerization of acetyleneto benzene [21]. Anderson et al. studied size-selected palladium clusters and deposited them onthe rutile phase of titanium dioxide (TiO2) for COoxidation [22]. By employing X-ray photoemissionspectroscopy and temperature-programmed reac-tion measurements, they found that the activity ofthese catalysts was associated with Pd3d bindingenergy. Li et al. utilized a double-solvent methodcombined with a photoreduction process to prepareactive Pd nanoclusters encapsulated inside the cageof NH2-Uio-66 [23]. The resultant catalyst showedexceptional performance for a Suzuki coupling re-action under visible-light irradiations. Nevertheless,although the sizes of the nanoclusters have beenreduced, their multiple distributions of active metalsites alongside different geometric and electricstructures might not be ideal for specific catalyticreactions [9,10].The stability of nanoclusters wouldalso be a problem for their application in hetero-geneous catalysis, especially at higher operationaltemperatures [24].

Further downsizing nanoclusters to the atomiclevel, namely single atoms (SAs), maximum atomutilization and superior/distinguishing catalyticperformance are supposed to be obtained [9,10].Of note is the fact that unexpected superior cat-alytic performances of the single-atom catalysts(SACs) are often observed as one of the keyadvances of these novel catalysts versus theirnanoscale counterparts. This can be ascribed tothe unsaturated environments of metal activesites, quantum size effects and metal–supportinteractions [25–28]. Therefore, the research onSACs has rapidly progressed from their funda-mental aspects to pursing practical applicationsin areas of nanotechnology and materials science.A study that has attracted considerable attentionsince its publication in 1995 is that of Thomaset al., who reported that direct grafting oforganometallic complexes onto the walls ofmesoporous silica gives a shape-selective high-performance catalyst with well-separated, homoge-neously dispersed and high surface concentrationsof active sites for the epoxidation of cyclohexenesand their derivatives [29]. Later, in 2003, Flytzani-Stephanopoulos et al. discovered that the water–gasshift reaction was not affected by the catalyticactivity of metallic Au or Pt nanoparticles; instead,nonmetallic Au or Pt species on the ceria surfaceplayed a key role in this reaction [30]. In 2007, Leeet al. successfully synthesized Pd-Al2O3 catalyst andvalidated that the extremely low metal loading leadsto the formation of atomically isolated PdII species,which greatly contribute to the excellent selox ac-

tivity of allylic alcohols [31]. The key discovery wasthat the employment of homogeneously dispersedSACs generally confers a dramatic improvement incatalytic activity, selectivity and stability, or evenconsiderably different catalytic properties thanthe corresponding nanoparticles and nanoclusters[32–35]. This is highly desirable and has attractedextensive scientific attention, as they might poten-tially act as alternatives to circumvent the problemsof scarcity and high cost of the noble-metal catalystsused in large-scale catalysis applications [36].Specifically, the active single-atom sites are welldefined and atomically stabilized on the supports,and the identical geometric structure of each activesite is similar to that of a homogeneous catalyst.Recently, studies have clearly demonstrated thatthe utility and uniqueness of these SACs have greatpotential to bridge the gap between homogeneousand heterogeneous catalysis [37–42]. This wouldsolve the problems of the difficulty in separatingthe homogeneous catalysts from raw materials andproducts, as well as combining the merits of bothhetero- and homogeneous catalysts. The SACsalso provide a good avenue to identify the detailedstructural features for the active sites and an idealmodel to elucidate the structure–activity relation-ship [43–45]. Such catalysts have shown intriguinginterests in the catalysis field [10,32,34,43,46–49].To meet the practical demand, the most importantchallenges for fabricating the SACs are to increasethe density of active sites and to improve theirintrinsic activities [32,36,45]. The first challenge isthe high propensity for aggregation of SAs once thesize of the nanomaterials is greatly reduced [10].The second challenge is the rational control of thecoordination environment of the singlemetal atoms.

Recently, several strategies for constructingatomically dispersedmetal sites on catalyst supportshave been extensively studied [9,10,32,42]. Thesestrategies include enhancing the metal–supportinteractions, engineering vacancy defects and voidson the supports, and modifying surface functionalgroups [9,42]. In most cases, the supports forisolated SACs are chosen on purpose, as they canstabilize the isolated catalytic SAs or activate nearbyreactants to form intermediate species for thecatalytic active sites [50–52]. For example, zeolitescould provide effective voids to anchor individualmetal atoms to maintain the high dispersion ofthe isolated metal atoms and prevent them fromsintering at high temperatures under oxidative orreductive atmospheres during catalysis processes[53]. Nanoparticles and nanoclusters can also serveas supports. Through elegant studies of supportmaterials, Sykes et al. showed that the isolatedPd atoms can be supported on a Cu surface and

Page 54: Special Topic: Single-Atom Catalysts - Western Engineering...Predictive approach of heterogeneous catalysis Metal-organic frame-work assisted synthesis of single-atom catalysts for

REVIEW Li et al. 675

Adsorption

Co-precipitation

Galvanicreplacement

Bottom-up

High-tempmigration

High-temppyrolysis

Top-down

SACs

Figure 1. Schematic representation of the bottom-up and

top-down strategies for the synthesis of SACs.

significantly lower the energy barrier to hydrogenuptake on and subsequent desorption from thenearby Cu atoms [48]. This facile hydrogen disso-ciation at the isolated Pd atoms and weak bindingto the Cu surface together facilitate selective hydro-genation of styrene and acetylene. Toshima et al.described a crown-jewel concept for the construc-tion of catalytically highly active top gold atomson palladium nanoclusters [54]. Interestingly, thegold atoms can be controllably assembled at the topposition on the cluster and exhibit high catalytic ac-tivity because of their high negative-charge densityand unique structure. Recent reports have begunto document that the defects in reducible oxides(e.g. TiO2 and CeO2), graphene or C3N4 also helpto stabilize isolated metal atoms [32,42,55]. Forexample, Du et al. investigated the favorable role ofisolated palladium and platinum atoms supportedon graphitic carbon nitride (g-C3N4) to act asphotocatalysts for CO2 reduction [56]. Overall,an important conclusion derived from these worksis that the further development of this SACs fieldrequires a more fundamental understanding of SAformation at the atomic scale.

This review covers the preparation strategiesfor SACs, which can be categorized according tohow their components are integrated, namely viabottom-up and top-down approaches (Fig. 1).Currently, a large majority of SACs are synthesizedvia a bottom-up strategy by using oxides or carbonsupports to construct N or O defects to enable thedeposition of metal precursors.This is followed by achemical reduction process to generate SACs fromhigh-oxidation-state ions to low oxidation state.However, the following drawbacks are encounteredfrequently: difficulty in accessing highmetal loading

because of their high propensity for aggregation andthe difficulty in constructing homogeneous coor-dination environments for the reactive sites. Thesedrawbacks lead to limited selectivity and stabilityof the SACs, greatly limiting their potential use invarious industrial fields. In this regard, Li and Wuproposed a top-down strategy to construct SACsby the pyrolysis of metal nodes in metal-organicframeworks (MOFs) for the first time [57]. In thiscase, the introduction of Zn atoms intoMOFs is im-portant and can effectively prevent the formation ofCo NPs (nanoparticles) during the high-temperature pyrolysis process. The resultingCo SAC has a high metal loading close to 5% andshowed exceptional chemical and thermal stability.A distinguishing feature of this strategy is not onlythat the metal loading can be substantially increasedfrom 1% to 5%, which is important from practicalperspectives, but it can control the coordinationenvironments to construct high-performanceSACs by exposing real active sites. This top-downstrategy overcome challenges in the fabrication ofSACs with a traditional bottom-up strategy andhas great potential to meet the requirement foruse in practical applications. In addition to thefabrication strategies, the use of these methods indifferent chemical reactions will also be presented.Finally, future challenges and opportunities will bediscussed.

BOTTOM-UP SYNTHETICMETHODOLOGIES FOR THECONSTRUCTION OF SACSThe bottom-up strategy is the most commonmethod to synthesize metal SACs, during whichthe metal precursors are adsorbed, reduced andconfined by the vacancies or defects of the supports[9,10,32,52]. Nevertheless, how to effectivelyincrease the SACs loading with well-defined dis-persion on the supports is still challenging. First,aggregation would occur during a chemical synthe-sis or catalytic process when high loading of SAsis required. Second, the architectural structures ofanchor sites for confining and stabilizing the metalSACs on the support remain elusive; therefore, thecoordination environment for metal SAs might beinhomogeneous and poorly defined [58]. Optimiza-tion of the precursors and supports and controllingof the synthetic procedures play a key role in tuningthe metal–support interaction and guaranteeing thehomogeneous dispersion of SACs.

For the wet-chemistry strategy, the precursor so-lutions of mononuclear metal complexes are firstanchored to the supports by a coordination effect

Page 55: Special Topic: Single-Atom Catalysts - Western Engineering...Predictive approach of heterogeneous catalysis Metal-organic frame-work assisted synthesis of single-atom catalysts for

676 Natl Sci Rev, 2018, Vol. 5, No. 5 REVIEW

Co-precipitation

(c)

(b)

Nor

mal

ized

abs

orpt

ion

(a.u

.)

3.0

2.5

2.0

1.5

1.0

0.5

0.0

11.540 11.560 11.580 11.600 11.620Energy (eV)

PtO2

0.08% Pt/FeOx R2000.08% Pt/FeOx R2500.75% Pt/FeOx R2502.73% Pt/FeOx R2504.30% Pt/FeOx R250Pt foil

(a)

( )

(b)

Pt foilPtO2Sample ASample B

Inte

nsity

(a.u

.)

0 1 2 3 4 5 6R (A)

× 13

× 14

Inte

nsity

(a.u

.)

11.55 11.60 11.65E (keV)

Sample B

Sample A (used)PtO2

Pt foil

Sample A

Figure 2. (a) HAADF-STEM images of Pt1/FeOx. Adapted with permission from [46]. (b) HAADF-STEM image of 0.08%Pt/FeOx-R200. Adapted with

permission from [60]. (c) HAADF-STEM images of Ir1/FeOx. Adapted with permission from [61].

between the metal complexes and the functionalgroups of the support surfaces [32]. Then, the or-ganic ligands of themetal complexes are removed bya post-treatment to expose more active sites to meetthe requirement of catalytic reactions. Particularly,the advantage of wet chemistry for preparing SACsis that this method does not require specializedequipment and can be routinely practiced in anychemistry lab [59].

Co-precipitation approachCo-precipitation is one of the commonly employedapproaches for preparing SACs, during which thesubstances that are normally soluble under theconditions would be precipitated. A significant ad-vantage of this method lies in its extreme simplicity,as no additional complicated steps are involved.For a classical example, Zhang et al. employed thismethod to fabricate single Pt atoms supportedon iron oxide nanocrystallites (Pt1/FeOx) [46].The metal precursor of H2PtCl6·H2O was mixedwith Fe(NO3)3·9H2O in a proper molar ratio andpH. After recovery, the precipitate was dried andcalcined, resulting in the formation of Pt1/FeOx.The aberration-corrected scanning transmissionelectron microscopy (AC-STEM) and extendedX-ray absorption fine structure (EXAFS) spec-tra demonstrated the individual Pt atoms wereuniformly dispersed on FeOx support, with ametal loading level of 0.17 wt% (Fig. 2a). This

SAC showed extremely high atom efficiency, excel-lent stability and superior activity for both CO oxi-dation and preferential oxidation of CO in H2.Theyfound that these merits can be attributed to the par-tially vacant 5d orbitals of the positively chargedhigh-valent Pt atoms, as they can effectively reduceCO-adsorption energy and activation barriers thatare required for CO oxidation. This study demon-strated the feasibility of using the defects of oxidesupports to serve as anchoring sites for metal clus-ters and single metal atoms. Subsequently, the fea-sibility and efficiency of this approach were furtherdemonstrated by the Zhang group showing that thehigh-performance Pt- and Ir-based SACs could alsobe obtained (Fig. 2b and c) for use in organic trans-formation [60] and water–gas shift reactions [61].In these examples, defects in the oxide supports andthe amount of metal loading were found to be criti-cal for accessing high-performance SACs that wouldnormally lead to aggregation.

Adsorption approachThe adsorption method is one of the most funda-mental approaches for constructing isolated metalatoms on the supports [62,63]. It is simple, directand has been widely used in the preparation of sup-portedmetal catalysts.Generally, after themetal pre-cursors are adsorbed on the support, the residual so-lution is removed and then the catalysts are driedand calcined. To ensure the SAs could be stably

Page 56: Special Topic: Single-Atom Catalysts - Western Engineering...Predictive approach of heterogeneous catalysis Metal-organic frame-work assisted synthesis of single-atom catalysts for

REVIEW Li et al. 677

anchored onto the supports with atomic disper-sion, appropriate functional groups on the supportsshould be given consideration.

Oxides are generally employed as an efficientsupport for preparing catalysts. In 2013, Narulaet al. [64] reported single Pt atoms supportedon θ -Al2O3(010) prepared by a wet-impregnationmethod using alumina powder and chloroplatinicacid. In this work, water was gradually evapo-rated before the resulting powder was transferredto an alumina crucible and subjected to a pyrol-ysis process. The resultant catalyst was catalyti-cally active in its ability to oxidize CO to CO2. Inaddition, serials of Pt/θ -Al2O3 catalysts with dif-ferent metal loadings were prepared, and the re-sults reveal that they are highly active towards NOoxidation [65].

In the same year, Tao et al. [66] developedan impregnation-reduction method for preparingsingly dispersed Rh atoms supported on Co3O4nanorods. This method involves the impregnationof Rh3+ on Co3O4 nanorods followed by on-sitereduction of Rh3+ using NaBH4. In situ character-izations reveal evidence of the active sites of iso-latedRh atoms in the formationofRhCon onCo3O4nanorods, whichwere generated through restructur-ing of Rh1/Co3O4 at 220◦C in reactant gases. Theresulting new catalytic phase exhibits a high selectiv-ity to produce N2 in the reduction of NO with H2between 180◦C and 300◦C.

A report by Li et al. demonstrated that single Pt1and Au1 atoms can be stabilized by lattice oxygenon ZnO{1010} surface via an adsorption method[67]. In detail, ZnO-nanowires (nws) were dis-persed in de-ionized water followed by the additionof H2PtCl6·6H2O or HAuCl4 solution. After an ag-ing process, the suspension was filtered, washed anddried to givePt1/ZnOandAu1/ZnOcatalysts. Simi-larly, Zhang et al. fabricated anRhSACsupportedonZnO nws by introducing RhCl3 solution into ZnOnws that were dispersed in de-ionized water [40].After stirring and aging processes, the resulting pre-cipitate was filtered, washed, dried and reduced.As the weight loading of Rh reduced from 0.03%to 0.006%, the isolated Rh SACs can be clearlyobserved. During the synthetic process, the Rhatoms bond with proximal Zn atoms which loseone or more O atoms. Therefore, electrons trans-fer from metallic Zn to Rh atoms to generatenear-metallic Rh species. The results show that theas-obtained Rh1/ZnO-nws SACs exhibited compa-rable efficiency in the hydroformylation of severalolefins to the homogeneous Wilkinson’s catalyst,along with superior catalytic activity to those of themost highly reported heterogeneous nanoparticle-based catalysts.

In a more recent piece of work, Wang and co-workers described a convenient two-step synthesisof an atomically dispersed Pt catalyst supported onceria (CeO2), with 1 wt.%metal loading, by wetnessimpregnation and steam treatment [68]. Chloropla-tinic acid was added drop-wise to the CeO2 supportwhile being ground in a mortar and pestle. The as-obtained powder was then dried, calcined, thermalaged and stream treated to give the catalyst. The au-thors demonstrated that the activation of SACs onCeO2 via high-temperature steam treatment can ac-complish excellent low-temperature CO-oxidationactivity and superior thermal stability. This is be-cause the steam treatment can enable the formationof active surface lattice oxygennear isolatedPt atomsto considerably enhance catalytic performance. Fur-ther investigation of the nature of this active surfacelattice oxygen on Pt/CeO2 was supported by den-sity functional theory (DFT) calculations and re-action kinetic analyses. They found the oxygen va-cancies from the CeO2 bulk can redistribute to theCeO2(111) surface when exposed to water at a hightemperature. During the steam-treatment process,H2Omolecules can fill out the oxygen vacancy overthe atomically dispersed Pt/CeO2 surface, affordingtwo neighboring active Olattice[H] sites around Pt.This provides the significantly improved reactivityand stability.

Yan and co-workers developed a unique adsorp-tion approach to construct Pt SACs, anchored in theinternal surface ofmesoporous Al2O3, by amodifiedsol-gel solvent vaporization self-assembly method[69], as shown in Fig. 3a. Triblock copolymersP123, C9H21AlO3 and H2PtCl6 were first mixedin ethanol. With continued evaporation of the sol-vent, the amphiphilic P123 macromolecules andC9H21AlO3 assembled into a highly ordered hexag-onally arranged mesoporous structure, with Pt pre-cursor encapsulated in the matrix. The as-obtainedgel was then calcined in air to decompose the P123template. Meanwhile, the C9H21AlO3 was trans-formed into a rigid, well-aligned mesoporous Al2O3framework. This was followed by a reducing stepin 5% H2/N2 to give the isolated Pt SAs stabilizedby the unsaturated pentahedral Al3+ centers. Theauthors showed that the catalyst retained its struc-tural integrity and exceptional catalytic performancein several reactions under harsh conditions, such ashydrogenation of 1,3-butadiene after exposure to areductive atmosphere at 200◦C for 24 h, n-hexanehydroreforming at 550◦C for 48 h and CO oxida-tion after 60 cycles between 100◦C and 400◦C over1 month.

Zeolites are crystalline materials with well-defined structures and high surface area, along withmore sites for robust bonding with catalytic species

Page 57: Special Topic: Single-Atom Catalysts - Western Engineering...Predictive approach of heterogeneous catalysis Metal-organic frame-work assisted synthesis of single-atom catalysts for

678 Natl Sci Rev, 2018, Vol. 5, No. 5 REVIEW

Adsorption (a)

(b) (c)

(d)

Pyrolysis HF etching

C N ZrO2Ru

a)PtCl62- Pt C H

O CI AI

P123

Al(OCHCH3CH3)3

Self-assemblyComplexing

Gel

Calcination

Reduction

0.2Pt/m-Al2O3-H2

Single atoms

Nano clusters

0.2-0.3 0.4-0.5 0.6-0.9 ~1.0 nm

Cou

nt p

erce

ntag

e (%

) 80

60

40

20

0

Figure 3. (a) Schematic illustration of the 0.2Pt/m-Al2O3-H2 synthesis process. Adapted with permission from [69]. (b) HAAD–STEM images of the

0.25Au-Na/[Si]MCM41 catalyst. Adapted with permission from [72]. (c) TEM and HAADF-STEM images of 0.35 wt% Pt/TiN. Adapted with permission

from [73]. (d) Scheme of proposed formation mechanisms, TEM and HAADF-STEM images for Ru SAs/N–C. Adapted with permission from [76].

[24,70]. Specifically, zeolites could provide effectivevoids to anchor individual metal atoms to maintainthe high dispersion and prevent them from sinteringat high temperatures under oxidative or reductiveatmospheres during the catalysis processes [53]. In2012, Gates et al. reported that atomically dispersedgold atoms catalyse with a high degree of uniformitysupported on zeolite NaY [71]. The site-isolatedgold complexes retained after CO-oxidation cataly-sis, confirming the robust stabilization effect of thezeolite channels for gold species.

The addition of alkali ions, such as sodiumor potassium, on inert KLTL-zeolite and meso-porous MCM-41 silica materials could structurallystabilize the single gold sites in Au–O(OH)x– en-sembles (Fig. 3b), as demonstrated by Flytzani-Stephanopoulos and co-workers [72]. They haveshown evidence that the active catalyst was com-posed of alkali ions linked to the gold atom through–O ligands, not merely on the support, making thereducible oxide supports no longer an essential re-quirement.The validation tests show that the single-site gold atoms were homogeneously dispersed andhighly active for the industrially important low-temperature water–gas shift reaction.

In addition to metal oxides and zeolites, othersupports such as nitrides and carbides have alsobeen explored and shownpromise for stabilizing SAsfor use in catalysis. Lee et al. described a Pt SACsupported on titanium nitride (TiN) nanoparticleswith the aid of chlorine ligands [73].H2PtCl6·6H2Owas dissolved in anhydrous ethanol and mixedwith acid-treated TiN nanoparticles before the re-sulting sample was dried and reduced. Transmis-sion electron microscopy (TEM) and HAADF-STEM images of the samples are shown in Fig. 3c.

The results show that the 0.35 wt% Pt/TiN sampleaffords a high mass activity and a unique selectivitytowards electrochemical oxygen reduction, formicacid oxidation and methanol oxidation.

Carbon nitride (C3N4) has been proved as an al-ternative support material by virtue of their poros-ity and high surface area [55]. Li et al. used an im-pregnation method to access isolated Au atoms an-chored on polymeric mesoporous graphitic C3N4(mpg-C3N4) [74].The catalytically active AuI atomwas coordinated by three nitrogen or carbon atomsin tri-s-triazine repeating units. This coordinationfeature significantly prevents the Au atoms from ag-gregation and makes the AuI surface highly active.Moreover, they demonstrated this catalyst as highlyactive, selective and stable for silane oxidation withwater.

In 2017, Ma et al. developed a highly efficientcatalyst consisting of isolated Pt atoms uniformlydispersed on an α-molybdenum carbide (α-MoC)support that can enable low-temperature, base-freehydrogen production through aqueous-phase re-forming of methanol [75]. They found that theα-MoC displays stronger interactions with Pt thanother oxide supports or β-Mo2C; therefore, atom-ically dispersed Pt atoms can be formed on anα-MoC support following a high-temperature acti-vation process.This generates an exceptionally high-density electron-deficient surface to stabilize Pt sitesfor the adsorption/activation of methanol. This cat-alyst affords an excellent turnover frequency andthe corresponding hydrogen production greatly ex-ceeds those of previously reported catalysts for low-temperature aqueous-phase reforming of methanol.They deduce that the unique structure of α-MoC,which affects water dissociation, and the synergic

Page 58: Special Topic: Single-Atom Catalysts - Western Engineering...Predictive approach of heterogeneous catalysis Metal-organic frame-work assisted synthesis of single-atom catalysts for

REVIEW Li et al. 679

effects betweenPt andα-MoCtogether affect the ac-tivation of methanol and the subsequent reformingprocess.

In 2017, Wu et al. reported a novel syntheticapproach to construct isolated single Ru atoms onnitrogen-doped porous carbon (Ru SAs/N–C) bya coordination-assisted strategy usingMOFs for thehydrogenation of quinolones [76]. It is noticed thatthe strong coordination effect between the lonepair of nitrogen and d-orbital of Ru atoms is cru-cial for the formation of stable Ru SAs (Fig. 3d).Without the dangling −NH2 groups, the Ru atomstend to aggregate into nanoclusters, even confinedin the pores of MOFs. The results demonstrate theRu SAs serve as an effective semi-homogeneous cat-alyst to the chemoselective catalyse hydrogenationof quinolones. This method has been shown to po-tentially broaden the substrate scope for the synthe-sis of SACs with unique properties for use in variouschemical reactions.

Together, the ease of preparation for SACs us-ing a wet-chemistry strategy envisages a promisingfuture in the field. However, these methods havetheir own disadvantages. For example, some metalatoms might be buried either in the interfacial re-gions of the support agglomerates or within the bulkof the support when co-precipitation methods areapplied [43]. In addition, when high metal loadingis required for the construction of SACs, aggregationwould inevitably occur [9]. This trade-off should beminimized by developing new synthesis methods.

Other methodologies have also been ex-plored to design and synthesize SACs with varieschemical and physical functionalities and futureunderpinned studies in these directions.The photo-chemical method becomes particularly appealing toassist the effective adsorption of SAs on the supportsand has been proven to be effective for the synthesisof nanocrystals, such as gold, silver, platinum,palladium, etc. [77–80]. In this process, regulatingthe nucleation and growth processes of nanocrystalshas been a major topic. Flytzani-Stephanopouloset al. constructed isolated gold atoms supportedon titania with a loading of approximately 1 wt%under ultraviolet (UV) irradiation [81].They foundthat the addition of ethanol can serve as a chargescavenger to facilitate the donation of electrons fromgold atoms to −OH groups on the titania support.The catalytic performance was examined and theresults showed that this catalyst displayed excellentactivity for the low-temperature water–gas shiftreaction, as well as admirable stability in long-termcool-down and startup operations.

An important study by Zheng et al. demon-strated a room-temperature photochemical strategyto construct atomically dispersed palladium atoms

supported on ethylene glycolate (EG)-stabilizedultrathin TiO2 nanosheets (Pd1/TiO2 catalyst)with a Pd loading up to 1.5% [82]. Typically,two-atom-thick TiO2 nanosheets were prepared byreacting TiCl4 with EG and used as the support.H2PtCl6 was then added to the TiO2 dispersion foradsorption of Pd species followed by irradiation byUV to give the Pd1/TiO2 catalyst. TEM, STEM andEXAFS revealed that the isolated Pd atoms wereevenly dispersed over theTiO2 support, without anyobservable evidence of NPs (Fig. 4a). The catalystexhibited excellent catalytic performance in the hy-drogenation of C = C bonds, outperforming thosecommercial Pd catalysts. In addition, there wasno observable decay in the catalytic activity for20 cycles, suggesting the robustness of thePd1/TiO2 catalyst. Importantly, they foundthis catalyst can activate H2 in a heterolytic pathwayto drastically enhance its catalytic activity in thehydrogenation of aldehydes. This mechanismhas been commonly observed for homogeneouscatalysts, such as Au, Pd and Ru complexes;however, there is no report for heterogeneousPd catalysts. This study set a good example usingatomically dispersed metal catalysts for bridgingthe gap between heterogeneous and homogeneouscatalysis.

Very recent work byWu and co-workers showeda novel synthetic approach to accessing atomicallydispersed platinum species on mesoporous carbonvia iced-photochemical reduction of frozen chloro-platinic acid solution (Fig. 4b) [83]. In this report,H2PtCl6 solution was first frozen by liquid nitro-gen followed by irradiation using a UV lamp. TheH2PtCl6 ice was kept overnight in dark conditionsat room temperature to give a clear aqueous Ptsingle-atom solution. Then mesoporous carbon so-lution and Pt single-atom solution were mixed, fil-tered, and dried at room temperature. Finally, theice lattice naturally confines the dispersed ions andatoms to affect the photochemical reduction prod-ucts and further prevent the aggregation of atoms.To test the generality of this concept, they also fab-ricated isolated Pt atoms deposited on different sup-ports, including mesoporous carbon, graphene, car-bon nanotubes, TiO2 nanoparticles and zinc oxidenanowires. Among them, the isolated Pt atoms sup-ported onmesoporous carbon exhibited exceptionalcatalytic performance for hydrogen evolution reac-tion, as well as an excellent long-time durability, out-performing the commonly employed Pt/carbon cat-alyst. This iced-photochemical reduction approachprovides a promising avenue for the green synthe-sis of SAs and sub-nanometer clusters, and opensup possibilities for fine-tuning the nucleation andgrowth of nanocrystals in wet chemistry.

Page 59: Special Topic: Single-Atom Catalysts - Western Engineering...Predictive approach of heterogeneous catalysis Metal-organic frame-work assisted synthesis of single-atom catalysts for

680 Natl Sci Rev, 2018, Vol. 5, No. 5 REVIEW

Other techniques-assisted adsorption

Iced-photochemical reduction strategy(b)

H2PtCl6 solution Pt nanoparticles

Freeze Pt single atoms

UV light

UV light

Ball milling strategy(c)

(1) Fresh Fe SiO2C

(2) In-situ Fe SiO2C

(3) In-situ Fe/SiO2

(1) Fresh Fe SiO2C

(2) In-situ Fe SiO2C

(3) In-situ Fe/SiO2

Nor

mal

ized

abs

orba

nce

(a.u

.)

7100 7110 7120 7130 7140 7150 7160 7170Energy (eV)

Fe foilFeSi2Fe2O3

Fe foil

FeSi2Fe2O3

Fe-Fe

Fe-Si

Fe-O

Fe-C

FT k

3 χ(k

) (A

-3)

×2

×2

×2

-1 0 1 2 3 4R (A)

Photochemical strategy(a)

Pd1/TiO2

FittingPd foil

FT χ

(k)* k

2

Pd-O

Pd-Pd

1 2 3 4 5R (A)

C

ALD strategy(d)

Graphene

o o

Graphene

H3C H3C CH3CH3

CH3 CH3

PtPtMeCpPtMe3

MeCpPtMe3MeCpPtMe3

H3CCH3

CH3

CH3

Pt

Graphene

o o

Pt Pt

Graphene

Pt PtH3C CH3

CH3

PtH3C CH3

CH3

Pt

MeCpPtMe3

O2

O2

O2ALD Pton GNS

x1/3

Figure 4. (a) Structural characterizations of Pd1/TiO2 catalyst. Adapted with permission from [82]. (b) Schematic illustration of the iced-photochemical

process comparedwith the conventional photochemical reduction of H2PtCl6 aqueous solution. Adaptedwith permission from [83]. (c) Structural features

of 0.5% Fe C©SiO2. Adapted with permission from [39]. (d) Schematic illustrations of the Pt ALD mechanism on graphene nanosheets. Adapted with

permission from [86].

Recently, high-energy bottom-up ball-millingsynthesis has been proved as a powerful methodto break and reconstruct chemical bonds of ma-terials with high efficiency. Such an approach wastaken by Bao et al., who reported a lattice-confinedsingle iron site catalyst embedded within a sil-ica matrix by a solid fusion method. Briefly, com-mercial SiO2 and Fe2SiO4 were mixed and sub-jected to ball milling under argon and fused inthe air [39]. As expected, the unsaturated singleFe sites served as active centers (Fig. 4c) to effi-ciently enable the direct, non-oxidative conversion

of methane, exclusively to ethylene and aromat-ics. The presence of single Fe sites effectively pre-vented catalytic C-C coupling, oligomerization andcoke deposition. In addition, this catalyst showedextremely stable performance, with no deactivationobserved during long-term testing, and the selec-tivity for total carbon of the three products wasretained. Subsequently, the group used the samemethod to construct single-atom iron sites by em-bedding highly dispersed FeN4 centers in graphenematrix via high-energy ball milling of iron phthalo-cyanine and graphene nanosheets [84]. In this

Page 60: Special Topic: Single-Atom Catalysts - Western Engineering...Predictive approach of heterogeneous catalysis Metal-organic frame-work assisted synthesis of single-atom catalysts for

REVIEW Li et al. 681

(a)(b)

Galvanic replacement

Pt foil

Pt0.1Cu14/Al2O3

Pt0.2Cu12/Al2O3

Pt2Cu6/Al2O3

0 2 4 6R (A)

χ (A

-4)

Figure 5. (a) Characterization of Pt/Cu SAANPs. Adapted with permission from [50]. (b)

Scanning tunneling microscope image of a 0.01 ML Pt/Cu(111) SAA surface. Adapted

with permission from [90].

system, the FeN4 center is highly dispersed and wellstabilized by the graphene matrix. The formation ofthe Fe = O intermediate is important in promotingthe conversion of benzene to phenol. Remarkably,this reaction can proceed efficiently at mild condi-tions such as room temperature or even as low as0◦C. DFT calculations confirm that the catalytic ac-tivity stems from the confined iron sites, along withmoderate activation barriers for the reaction thatproceeded at room temperature. Both studies clearlyshow the potential of the highly efficient ball-millingmethod for the fabrication of SACs for use in cataly-sis areas.

The atomic layer deposition (ALD) technique isa gas-phase chemical process and commonly usedto deposit a thin layer of film in a bottom-up fash-ion with near-atomic precision on the substrateby repeated exposure of separate precursors [85].This technique offers the feasibility of precise con-trol of the catalyst size from a single-atom, sub-nanometer cluster to the nanoparticle. It is expectedthat ALD would potentially provide a powerful ap-proach for the construction of intriguing SACs.Thisapproach was first demonstrated by Sun et al. in2013, who reported a practical synthesis of isolatedsingle Pt atoms on graphene nanosheets using theALD technique (Fig. 4d) [86]. In this work, Ptwas deposited on graphene supports by the ALDmethod using MeCpPtMe3 and oxygen as precur-sors and nitrogen as a purge gas. The resulting PtSAC showed improved catalytic activity comparedwith the commercial Pt/C catalyst. X-ray absorptionfine structure (XAFS) analyses show that the low-coordination and partially unoccupied 5d orbital ofPt atoms are responsible for the excellent catalyticperformance.

In 2015, Lu et al. described a single-atomPd1/graphene catalyst prepared by the ALDmethod with excellent performance in the selectivehydrogenation of 1,3-butadiene [87]. First, theanchor sites were created by an oxidation process onpristine graphene nanosheets, followed by a reduc-tion process via thermal de-oxygenation to controlthe surface oxygen functional groups. After an an-nealing step, phenolic oxygenwas observed to be thedominated oxygen species on the graphene support.ALD was then performed on the reduced grapheneto give a single-atom Pd catalyst by alternately ex-posing Pd(hfac)2 and formalin.This catalyst showedsuperior catalytic performance in the selective hy-drogenation of 1,3-butadiene, affording nearly100% butenes selectivity, and ∼70% selectivity for1-butene at a conversion ratio of 95% under mildconditions. They speculate that both the mono-π -adsorptionmode of 1,3-butadiene and the enhancedsteric effect induced by 1,3-butadiene adsorption onthe isolated Pd atoms contribute to the improvedselectivity of butenes. In addition, the Pd1/grapheneshowed remarkable durability against deactivationvia either metal atom aggregation or coking duringa 100-h reaction time on stream.

Using the same strategy, Sun and co-workersdescribed the preparation of isolated single Ptatoms and clusters on nitrogen-doped graphenenanosheets (NGNs) [88]. Here, Pt was first de-posited on the NGNs by the ALD technique usingMeCpPtMe3 andO2 as precursors andN2 as a purg-ing gas and a carrier gas. The size, density and dis-tribution of the Pt atoms on the NGNs or graphenenanosheets (GNs) canbeprecisely controlled by theALD cycles. As expected, the isolated Pt atoms andclusters on the NGNs have been demonstrated toshow superior catalytic activity and stability for thehydrogen evolution reaction (HER) compared withthe conventional Pt NP catalysts. This can be ex-plained by the small size and the special electronicstructure of the adsorbed single Pt atoms on NGNs.Together, the use of the ALD technique has showngreat promise for large-scale synthesis of highly ac-tive and stable single-atom and cluster catalysts.

The galvanic-replacement methodGalvanic replacement is a highly versatile and effec-tive approach for the construction of a variety ofnanostructures, with the ability to control the sizeand shape, composition, internal structure andmor-phology [24,57,89]. It is an electrochemical processthat consists of oxidation of one metal, termed as asacrificial template, by other metal ions that have ahigher reduction potential. When they are exposed

Page 61: Special Topic: Single-Atom Catalysts - Western Engineering...Predictive approach of heterogeneous catalysis Metal-organic frame-work assisted synthesis of single-atom catalysts for

682 Natl Sci Rev, 2018, Vol. 5, No. 5 REVIEW

to each other in solution, the sacrificial metal tem-plate will be preferably oxidized and dissolved intothe solution, while the ions of the second metalwill be reduced and deposited onto the templatesurface.

In 2015, Sykes et al. demonstrated that lowconcentrations of isolated Pt atoms in the Cu(111)surface (Fig. 5a) canbepreparedbygalvanic replace-ment on pre-reduced CuNPs to catalyse the butadi-enehydrogenationwith remarkable activity andhighselectivity to butenes [50]. In this case, CuNPswerefirst prepared and supported on γ -Al2O3 followedby calcination in air. The galvanic-replacement re-action was then carried out in an aqueous so-lution under nitrogen protection with constantstirring and refluxing. A desired amount of Pt pre-cursor was introduced to a suspension of Cu NPs inan aqueous solution containing HCl. The resultingmaterial was filtered, washed and dried to yield thecatalyst.They notice that, at low Pt loadings, the iso-lated Pt atoms can substitute into the Cu(111) sur-face to activate the dissociation and spillover of Hto Cu.Theweak binding between butadiene and Cuwould facilitate the highly selective hydrogenationreaction to butenes, without decomposition or poi-soning of the catalysts. This catalyst, with less thanone Pt atom per 100 copper atoms, also binds COmore weakly than metallic Pt, which is particularlyimportant for use in many Pt-catalysed chemicalreactions.

In a follow-up report, the Sykes group used thesame approach to construct Pt/Cu single-atom al-loys (SAAs) to examine C–H activation in differ-ent systems, including methyl groups, methane andbutane [90]. They observed that the Pt atoms weredistributed over the Cu surface and across both ter-races and at regions near step edges (Fig. 5b). Theresults show the Pt/Cu SAAs activate C–H bondsmore efficiently than Cu, along with superior stabil-ity under realistic operating conditions, effectivelyavoiding the coking problem that typically occurredwith Pt. Both pieces of work from the Sykes groupdemonstratedhowSAs canbedepositedon alloys—an important future direction for this field.

Though a variety of SACs have been developedby the bottom-up strategy, the downside of themethods described here is that it is still challengingto access SACs with high metal loading and a ho-mogeneous coordination environment for the activesites used in the catalytic process.This would lead tolimited selectivity and stability of the SACs for theirpractical use in various industrial fields. In addition,although ground-breaking, some of these methodsdo require specific/sophisticated preparation proce-dures that might not be compatible with all kinds ofSACs and ideal from practical perspectives.

TOP-DOWN SYNTHETICMETHODOLOGIES FOR THECONSTRUCTION OF SACSThe top-down strategy is based on the dissolutionof ordered nanostructures into smaller pieces togive desired properties and intriguing performances[59,91]. Extensive research efforts have pursued thisstrategy with the overarching aim of synthesizingSACs with unprecedented chemical and physicalproperties and understanding the complex mecha-nisms for catalysis that occur at the atomic level.This strategy has proven particularly useful in theformation of SACs with accurate control over themicro- or nanostructures [92].The precise structure(such as coordination number, dispersion tenden-cies and binding mode) of metal SAs synthesized bythe top-down methods has shown great promise inindustrially important applications [9,89,93,94]. Ef-forts to further understand the underlying featuresandmechanisms are required for thedevelopmentofnew methods for the construction of SACs and rep-resent a fertile area for future studies.

The high-temperature pyrolysis methodHigh-temperature pyrolysis has become one of thefascinating methods for synthesizing nanomaterialson different supports. Particularly, the developmentof a template-sacrificial approach via acid leaching oroxidative calcination has offered an alternative wayto generate SACs. Of note is that an appropriate py-rolysis temperature is critically important to give thedesired properties.

MOFs and zeolitic imidazolate frameworks(ZIFs) have interconnected 3D molecular-scalecages that make them highly accessible throughsmall apertures. Importantly, they can serve astemplates to obtain nitrogen-doped porous carbonwith abundant active nitrogen sites. Very recently,Wu et al. took advantage of theMOFs and originallydeveloped an effective strategy for accessing singleCo atoms supported on nitrogen-doped porouscarbon with a particularly high metal loading ofover 4 wt% via the pyrolysis of bimetallic Zn/CoMOFs [57]. This is pioneering work in this fieldand the strategy is particularly applicable to accesshigh-loading metal SACs that would otherwise bedifficult to produce. It should be noted that theenhancement of metal loading for preparing SACsin the present study is a significant breakthroughin this area, highlighting the specific requirementof SACs for practical applications. Importantly, theintroduction of Zn atoms into MOFs is critical andacts as an elegant approach to efficiently manipulatethe adjacent spatial distance between Co atoms,

Page 62: Special Topic: Single-Atom Catalysts - Western Engineering...Predictive approach of heterogeneous catalysis Metal-organic frame-work assisted synthesis of single-atom catalysts for

REVIEW Li et al. 683

High-temperature pyrolysis

CarbonizationReduction

Evaporation

dCo-Co

C N Zn Co(a) (b)C

N

O

Zn

Ni

Ni2+

Absorption

Ionic

Exchange

24 h, DMFPyrolysis

LeachingH2/ArNaOH

C Co BrNSiO2

(d)Fe(acac)3

Fe-ISAs/CN

O2OH-

Pyrolysis

Zn(NO3)22-Methylimidazole

Fe(acac)3@ZIF-8

FeCNOH

(c)

NFe

DopamineTris HCI PH 8.5

Carbonization Acid leaching

α-FeOOH nanorod α-FeOOH@PDA Fe/FeO@CN SA-Fe/CN

(e)

Adsorption

Pyrolysis

KirkendallEffect

Graphitization

Volatilization

C N Co Zn Fe Cl

(f)

Figure 6. (a) Schematic illustration of the construction of Co SAs/N–C. Adapted with permission from [57]. (b) Schematic illustration of the construction

of Ni SAs/N–C. Adapted with permission from [95]. (c) Schematic illustrations of the construction of Fe-ISAs/CN. Adapted with permission from [97].

(d) Schematic illustration of the construction of ISAS-Co/HNCS. Adapted with permission from [99]. (e) Schematic illustration of the construction of

SA-Fe/CN. Adapted with permission from [103]. (f) Schematic illustration of the construction of (Fe, Co)/N–C. Adapted with permission from [104].

thereby effectively preventing the formation of CoNPs (Fig. 6a). The Zn atoms, with a low boilingpoint of 907◦C, can be evaporated in the high-temperature pyrolysis process, providing abundantN sites. The Co nodes can be reduced in situ bycarbonization of the organic linkers in MOFs andanchored on the as-obtained N-doped porous car-bon support. Assuming the MOF as an integratedsystem, using this high-temperature pyrolysis ofMOF to access unsaturated SAs anchored on theN-doped porous carbon support can be catego-rized into the top-down approach. Control testingdemonstrated that the aggregated Co atoms wereformed for Co-containing MOF (ZIF-67) after apyrolysis treatment. HAADF-STEM and EXAFSverified the presence of isolated Co atoms dispersedon the N-doped porous carbon support. The result-ing Co SAC shows exceptional oxygen-reductionreaction (ORR) catalytic performance with a half-wave potential more positive than the commercialPt/C and most of the reported non-precious metal

catalysts. Robust chemical stability during electro-catalysis and thermal stability that resists sinteringat a high temperature of 900◦C have also been con-firmed, as little evidence of catalyst degradation wasobserved during the catalytic cycles. This work hasunderlined the significant importance of employingMOFs as an ideal carbon support for stabilizing sin-gle metal atoms at the atomic scale.

Subsequently, an ionic exchange strategy was de-veloped by the Wu group to assist in the construc-tion of a single Ni atom catalyst (Fig. 6b) betweenZn nodes and adsorbed Ni ions within the cavitiesof the MOF [95]. In this case, ZIF-8 was first dis-persed in n-hexane under ultrasound until a homo-geneous solution was formed. Then a small amountof Ni(NO3)2 aqueous solution was introduced, andthe mixed solution was vigorously stirred to causethe Ni ions to be absorbed completely. Then thesample was centrifuged and dried, followed by ahigh-temperature heating process in an argon at-mosphere to yield Ni SAC. This Ni SAC, with a

Page 63: Special Topic: Single-Atom Catalysts - Western Engineering...Predictive approach of heterogeneous catalysis Metal-organic frame-work assisted synthesis of single-atom catalysts for

684 Natl Sci Rev, 2018, Vol. 5, No. 5 REVIEW

metal weight loading of 1.53 wt%, delivered an ex-cellent turnover frequency forCO2 electroreductionof 5273 h−1, along with a maximum Faradaic effi-ciency for CO production of 71.9% and a high cur-rent density of 10.48 mA cm−2. This work, for thefirst time, demonstrates the great potential of usingMOF-basedmaterials to access SACs for use inCO2electroreduction.

To investigate the relationship betweencoordination numbers and CO2 electroreductioncatalytic performance, the Wu group sequentiallyprepared a series of Co SACs with different Ncoordination environments treated at differenttemperatures [96]. Bimetallic Co/Zn ZIFs weretreated by a pyrolysis process, during which theZn was evaporated away and the Co was reducedby carbonized organic linkers, generating isolatedCo atoms stabilized on nitrogen-doped carbon. Bycontrolling the pyrolysis temperatures, three CoSACs with different Co–N coordination numberswere obtained, being Co–N4 (800◦C), Co–N3(900◦C), and Co–N2 (1000◦C), respectively.The catalytic performance of these samples wasexamined, and the results show that the isolatedCo atom with two coordinated nitrogen atoms(prepared at 1000◦C) can afford significantlyhigher selectivity and superior activity, resulting ina CO formation Faradaic efficiency of 94% and acurrent density of 18.1mA cm−2 at an overpotentialof 520 mV. Importantly, this catalyst achieved aturnover frequency for CO formation of 18 200 h−1,outperforming most of the reported metal-basedcatalysts under comparable conditions. DFT cal-culation reveals that the decreased N coordinationenvironment leads to more unoccupied 3d orbitalsfor Co atoms, thereby facilitating adsorption ofCO2

�− and increasing CO2 electroreduction per-formance. This study demonstrates the significanteffect of N coordination environments on SACs forcatalytic performance.

The above studies further confirm the great po-tential of high-temperature pyrolysis of MOFs as apromising strategy to access SACs for different de-manding industrial applications.

With these attractive features, Li and co-workersprepared a highly stable isolated Fe atom catalyst,with Fe loading up to 2.16 wt%, that showed excel-lent ORR reactivity via a cage-encapsulated precur-sor pyrolysis approach [97]. This method is highlyeffective to access SACs because the precursors canbe encapsulated inside the ZIF pores, thereby pre-venting them from aggregating into nanoparticles(Fig. 6c). In this study, Fe(acac)3 was mixed withZIF-8, and the molecular-scale cages were formedwith the assembly of Zn2+ and 2-methylimidazole,with one Fe(acac)3 molecule trapped in one cage.

After a pyrolysis step, the ZIF-8 was transformedinto nitrogen-doped porous carbon, whereas theFe(acac)3 within the cagewas reducedby carboniza-tion of the organic linker, resulting in the formationof isolated iron atoms anchored on nitrogen species.The catalyst has been demonstrated to show excep-tional ORR catalytic activity, good methanol toler-ance and impressive stability. Importantly, the ORRcatalytic activity of this SAC outperforms those ofrecently reported Fe-bases materials and other non-precious metal materials. Experimental results andDFT calculations reveal the excellent ORR perfor-mance stems from the formation of atomically iso-lated iron atoms coordinated with four N atoms andone O2 molecule adsorbed end-on.

Using a similar approach, Li et al. describedthe synthesis of atomically dispersed Ru3 clus-ters via a cage-separated precursor pre-selectionand pyrolysis strategy [98]. Generally, two stepsare involved: (i) encapsulation and separation ofpreselected metal cluster precursors followed by(ii) a pyrolysis treatment. The resulting catalyst wascharacterized by HAADF-STEM and XAFS, andthe catalytic performance was tested for the oxida-tion of 2-amino-benzyl alcohol. The results showthat this Ru3/nitrogen-doped carbon (CN) catalystpossesses 100% conversion, 100% selectivity andan unexpectedly high turnover frequency (TOF),outperforming those of Ru SACs and small-sized Ruparticle catalysts.

An alternative approach to the thermal treatmentof MOFs for achieving SACs has been employed byLi et al., who used SiO2 as a template to access a hol-lowN-doped carbon spherewith isolatedCo atomicsites (Fig. 6d) [99]. Briefly, the SiO2 template wasdispersed in Co–TIPP/TIPP solution before intro-ducing another monomer. The collected powderwas thermally treated under a flowing H2/Ar andthen etched with sodium hydroxide to remove theSiO2 template to yield the Co SAC. Its ORR per-formance was investigated and the results demon-strate that exceptional catalytic activity was origi-nated from the single Co sites that can significantlyfacilitate the proton and charge transfer to the ad-sorbed ∗OHspecies.Using the same approach, aMoSAC was prepared by the Li group using sodiummolybdate and chitosan as precursors and showedexcellent HER performance [100]. Further studiesof the structure of the catalyst were supported byAC-STEMandXAFS, which confirmed that theMoatomwas anchored with one nitrogen atom and twocarbon atoms (Mo1N1C2).

In 2016,Zhang et al.described a similar template-sacrificial approach to create a self-supportingCo–N–C catalyst with single-atom dispersionand showed excellent catalytic activity for the

Page 64: Special Topic: Single-Atom Catalysts - Western Engineering...Predictive approach of heterogeneous catalysis Metal-organic frame-work assisted synthesis of single-atom catalysts for

REVIEW Li et al. 685

chemoselective hydrogenation of nitroarenes toyield azo compounds under mild conditions [101].In this study, the Co(phen)2(OAc)2 complex wassupported on Mg(OH)2 and then subjected to apyrolysis process. This was followed by the removalof the MgO support by an acid-leaching treatment.The merit of employing Mg(OH)2 is that it canessentially prevent the aggregation of cobalt atoms.This is because of the moderate interaction betweenMg(OH)2 and theCo species, as well as its inertnesstowards the reaction with Co during the pyrolysisprocess. After the acid-leaching step, the supportmaterial was removed to give a self-supportingCo–N–C material. X-ray absorption spectroscopywas tested and the exact structure of the catalyst wasconfirmed to be CoN4C8–1-2O2. Specifically, theCo single atom was coordinated with four pyridinicnitrogen atoms on the graphitic layer, along withoxygen atoms weakly adsorbed on the Co atomsperpendicular to the Co–N4 plane.

Using the same approach, Zhang et al. preparedan atomically dispersed Fe−N−C catalyst, whichexhibited exceptional activity and excellent reusabil-ity for the selective oxidation of the C−H bond,along with tolerance for a wide scope of substrates[102]. Briefly, the Fe(phen)x complex supportedon the nano-MgO template was pyrolysed at differ-ent temperatures underN2 atmosphere, followed byan acid-leaching step to remove the MgO template.They observed that the properties of the Fe specieswere dependent on the pyrolysis temperature, withmore metallic Fe particles formed at higher tem-peratures. The critical role of the Fe−Nx sites incatalysis was further confirmed by potassium thio-cyanate titration experiments and Mossbauer spec-troscopy.

An effective core–shell strategy has been in-troduced by the Li group using metal hydroxidesor oxides coated with polymers followed by high-temperature pyrolysis and acid-leaching steps,to synthesize single metal atoms anchored onthe inner wall of hollow CN materials [103]. Byemploying different metal precursors or polymers,they have successfully synthesized a series ofmetal SAs dispersed on CN materials (Fig. 6e).In detail, α-FeOOH nanorods were first pre-pared by a hydrothermal method, followed byself-polymerizing dopamine monomers to generateα-FeOOH@PDA. Then it was thermally treatedunder an inert atmosphere, during which thepolydopamine (PDA) layers were converted intothe CN shell and α-FeOOH was reduced to iron,giving rise to the strong interaction between the Featoms and the CN shell. Finally, acid leaching wascarried out to generate Fe SAs on the inner wall ofthe CN materials. The obtained SA-Fe/CN catalyst

showed a high conversion of 45% and an excellentselectivity of 94% for the hydroxylation of benzeneto phenol, outperforming Fe nanoparticles/CN.

Notably, in a most recent research, Wu et al.originally developed a host–guest strategy basedon MOFs to construct a Fe–Co dual-sites cata-lyst embedded in N-doped porous carbon support[104]. It involves binding betweenConodes and ad-sorbed Fe ions within the confined space of MOFs(Fig. 6f). Specifically, Zn/Co bimetallic MOF wasemployed as a host to encapsulate FeCl3 withinthe cavities by a double-solvents method. The Fe3+

specieswere reducedby the as-generated carbonandbond with the neighboring Co atoms. Meanwhile,the adsorbed Fe3+ species can accelerate the de-composition of metal–imidazolate–metal linkagesand generate voids inside the MOF. EXAFS andMossbauer spectroscopic analyses were performedto investigate the coordination environment of theFe–Codual sites.Theexperimental results show thatFeCoN6 is the active site for the (Fe, Co)/N–C cat-alyst and has been demonstrated to endow excel-lentORRperformance in an acidic electrolyte, alongwith comparable onset potential and half-wave po-tential to those of the commercial Pt/C. DFT calcu-lation reveals that the activation of O–O is favoredon the dual sites, which is important for the four-electronoxygen-reductionprocess.The fuel cell test-ing revealed that this catalyst outperforms most ofthe reported Pt-free catalysts in H2/O2 and H2/airconditions. In addition, this cathodecatalyst is ratherrobust in long-term operation for electrode mea-surement and H2/air single cell testing. Of noteis that, despite the fact that SACs generally confergreater activity than the corresponding nanoparti-cles, it is still important to be aware of the poten-tial aggregation pathways available to them. This isespecially crucial in cases where higher operationaltemperatures were applied. Therefore, the superiorcatalytic activity, selectivity, stability and the ease offabricationof thedual-sitesFe–Cocatalystmake thistype of SAC truly remarkable. Importantly, themainadvantages of this host–guest strategy include theability to incorporate different metal atoms and topermit the catalyst to be operated in awider dynamicrange. This study is expected to provide avenuesfor the synthesis of high-performance dual-sites cat-alysts with unique properties for use in chemicaltransformations.

Overall, these studies have shown that the high-temperature pyrolysis method is capable of produc-ing SACs with precisely controlled structures andmorphologies. Additionally, this unique approachhas been seen as a significant opportunity to enablethe efficient constructionof high-performance SACsfor use in various reactions.

Page 65: Special Topic: Single-Atom Catalysts - Western Engineering...Predictive approach of heterogeneous catalysis Metal-organic frame-work assisted synthesis of single-atom catalysts for

686 Natl Sci Rev, 2018, Vol. 5, No. 5 REVIEW

High-temperature atomic migrationFresh Aged

Figure 7. Schematic illustration of Pt nanoparticle sintering, showing how ceria can

trap the mobile Pt to suppress sintering. Adapted with permission from [105].

The high-temperature atomic-migrationmethodHigh temperatures are generally detrimental to cat-alysts’ activities. Although the SAs are homoge-neously dispersed on the support materials, theyhave a high propensity to move and aggregateinto nanoparticles when heated at high tempera-tures. Datye and co-workers take advantage of thephenomenon that metal nanoparticles can emit mo-bile species to prepare atomically dispersed metalcatalysts [105]. In this study, a Pt/La-Al2O3 cat-alyst was physically mixed with different types ofceria powders followed by a thermal treatment inflowing air. Because of the strong interaction be-tween PtO2 and ceria powders, the Pt species emit-ted from the alumina were trapped on the CeO2,forming thermally stable Pt1/CeO2 SACs (Fig. 7).The performance of the resulting SAC was testedfor CO oxidation, and the results suggest that it canserve as a highly effective sintering-resistant CO-oxidation catalyst at high temperature. They believethat this atom-trapping approach is potentially appli-cable and might provide exciting possibilities to ac-cess a variety of high-performance SACs. This workrepresents a novel strategy and has been demon-strated as being particularly effective in fabricatingSACs and connecting the relationship between thenanoparticles and SAs.

CONCLUSIONS AND PERSPECTIVEOver just a few years, there has been remarkableprogress in the development of various methods for

the synthesis of SACs. In this review, we summarizethe progress, bring new insights from recent yearsand pointed the way to the synthesis of SACs.

Currently, two general approaches have beenemployed for accessing SACs: bottom-up and top-down. Though still being developed, SACs haveemerged as an exceptional advancement in the de-velopment of highly efficient heterogeneous cata-lysts. The researchers have shown evidence that thesize of the nanomaterials does affect catalytic effi-ciency in the catalysis process. A noteworthy resultis that, by reducing the size of nanostructures fromthe nano- to the sub-nano scale and finally to SAs inatomic dimensions, catalytic performance has beenobserved to change drastically. This results from thelow-coordination environment, quantum size effectand enhanced metal–support interactions. More-over, the homogeneously and isolated metal activesites canmaximizemetal utilization, giving rise to theimpressively enhanced catalytic performance.

Recent experimental and theoretical progress hasunambiguously validated the strong evidence for thehigh activity, selectivity and stability of the high-performance SACs. These intriguing properties ofSACs are believed to endow great potential for ap-plications in heterogeneous catalysis. Importantly,SACs can act as an ideal platform to serve as abridge to connect hetero- and homogeneous catal-ysis. Thus, SACs are thought to have the potentialto overcome the difficulty encountered in homoge-neous catalysis.

As discussed previously, a major limiting fac-tor in the development of SACs is the lack of gen-eral methods to directly and efficiently access high-performance SACs. The construction of SACs foruse in catalysis represents an important challenge,highlighting the need for more fundamental re-search into detailed mechanisms. Along with theemergence of new characterization and computa-tional modeling techniques, single-atom active sitescan be investigated further. More advanced, directand effective in situ spectroscopic and microscopictechniques become particularly important to offernew insights into the chemical reactions involvedin SACs. Elucidating the important role of metalprecursors, support materials and experimental con-ditions and understanding the prerequisites forcatalytic activity of a given catalytic system are cru-cial for developing effective strategies for the syn-thesis of SACs. Several aspects should also be givenenough attention: first, the development of novel,controllable and facile synthesis methods for ac-cess high-loading SACs with finely and densely dis-persed single atoms; second, the construction ofsingle metal atoms with robust stabilization onthe support for use in practical conditions; third,

Page 66: Special Topic: Single-Atom Catalysts - Western Engineering...Predictive approach of heterogeneous catalysis Metal-organic frame-work assisted synthesis of single-atom catalysts for

REVIEW Li et al. 687

detailed experimental and theoretical work shouldbe done to comprehensively understand SACs-support effects. The top-down strategy has showngreat promise and significantly contributed to thesimplified synthesis routes for SACs with excep-tional activity and stability. Moreover, the metalloading can be markedly increased from 1% to 5%,and the coordination environments can be elabo-rately controlled.Thiswill definitely facilitate the de-velopment of general protocols for accessing SACsand underpin the exploration of other intriguing ap-plications.

Together, the field of SAs is expansive and rapidlydeveloping towards different applied research fields.The continued development of SACs represents animportant advancement in heterogeneous catalysisand will surely be the important focus of extensiveresearch efforts and a thriving field for various appli-cations for years to come.

FUNDINGThis work was supported by the National Key R&D Programof China (2017YFA0208300) and the National Natural Sci-ence Foundation of China (21522107, 21671180, 21521091,21390393, U1463202).

REFERENCES1. Karkas MD, Verho O and Johnston EV et al. Artificial photosyn-

thesis: molecular systems for catalytic water oxidation. Chem

Rev 2014; 114: 11863–2001.2. Tao F and Crozier PA. Atomic-scale observations of cata-

lyst structures under reaction conditions and during catalysis.

Chem Rev 2016; 116: 3487–539.3. Gilroy KD, Ruditskiy A and Peng HC et al. Bimetallic nanocrys-

tals: syntheses, properties, and applications. Chem Rev 2016;

116: 10414–72.4. Blakemore JD, Crabtree RH and Brudvig GW. Molecular cata-

lysts for water oxidation. Chem Rev 2015; 115: 12974–3005.5. Wang X, Ruditskiy A and Xia Y. Rational design and synthesis

of noble-metal nanoframes for catalytic and photonic applica-

tions. Natl Sci Rev 2016; 3: 520–33.6. Seh ZW, Kibsgaard J and Dickens CF et al. Combining theory

and experiment in electrocatalysis: insights into materials de-

sign. Science 2017; 355: 6321–32.7. Vajda S and White MG. Catalysis applications of size-selected

cluster deposition. ACS Catal 2015; 5: 7152–76.8. GawandeMB, Goswami A and Felpin FX et al. Cu and Cu-based

nanoparticles: synthesis and applications in catalysis. Chem

Rev 2016; 116: 3722–811.9. Liu J. Catalysis by supported single metal atoms. ACS Catal

2017; 7: 34–59.10. Yang XF, Wang A and Qiao B et al. Single-atom catalysts: a

new frontier in heterogeneous catalysis. Acc Chem Res 2013;

46: 1740–8.

11. Boles MA, Ling D and Hyeon T et al. The surface science of

nanocrystals. Nat Mater 2016; 15: 141–53.12. Xia BY, Wu HB and Wang X et al. One-pot synthesis of cubic

PtCu3 nanocages with enhanced electrocatalytic activity for

the methanol oxidation reaction. J Am Chem Soc 2012; 134:13934–7.

13. Christopher P, Xin H and Linic S. Visible-light-enhanced cat-

alytic oxidation reactions on plasmonic silver nanostructures.

Nat Chem 2011; 3: 467–72.14. Wu Y,Wang D and Li Y. Nanocrystals from solutions: catalysts.

Chem Soc Rev 2014; 43: 2112–24.15. Lu Q, Wang AL and Gong Y et al. Crystal phase-based epitax-

ial growth of hybrid noble metal nanostructures on 4H/fcc Au

nanowires. Nat Chem 2018; 10: 456–61.16. Fan Z and Zhang H. Crystal phase-controlled synthesis, proper-

ties and applications of noble metal nanomaterials. Chem Soc

Rev 2016; 45: 63–82.17. Chen Y, Fan Z and Luo Z et al. High-yield synthesis of crystal-

phase-heterostructured 4H/fcc Au@Pd core-shell nanorods

for electrocatalytic ethanol oxidation. Adv Mater 2017; 29:1701331–5.

18. Fan Z and Zhang H. Template synthesis of noble metal

nanocrystals with unusual crystal structures and their catalytic

applications. Acc Chem Res 2016; 49: 2841–50.19. Liu P, Qin R and Fu G et al. Surface coordination chem-

istry of metal nanomaterials. J Am Chem Soc 2017; 139:2122–31.

20. Boronat M, Leyva-perez A and Corma A. Theoretical and exper-

imental insights into the origin of the catalytic activity of sub-

nanometric gold clusters: attempts to predict reactivity with

clusters and nanoparticles of gold. Acc Chem Res 2014; 47:834–44.

21. Abbet S, Sanchez A and Heiz U et al. Acetylene cyclotrimer-

ization on supported size-selected Pdn clusters (1 ≤ n ≤30): one atom is enough! J Am Chem Soc 2000; 122:3453–7.

22. KadenWE,Wu T and KunkelWA et al. Electronic structure con-

trols reactivity of size-selected Pd clusters adsorbed on TiO2

surfaces. Science 2009; 326: 826–9.23. Sun D and Li Z. Double-solvent method to Pd nanoclusters

encapsulated inside the cavity of NH2 Uio-66(Zr) for efficient

visible-light-promoted Suzuki coupling reaction. J Phys Chem

C 2016; 120: 19744–50.24. Liu L, Diaz U and Arenal R et al. Generation of subnanomet-

ric platinum with high stability during transformation of a 2D

zeolite into 3D. Nat Mater 2017; 16: 132–8.25. Chakraborty I and Pradeep T. Atomically precise clusters of no-

ble metals: emerging link between atoms and nanoparticles.

Chem Rev 2017; 117: 8208–71.26. Jin R, Zeng C and Zhou M et al. Atomically precise colloidal

metal nanoclusters and nanoparticles: fundamentals and op-

portunities. Chem Rev 2016; 116: 10346–413.27. Fortea-Perez FR, Mon M and Ferrando-Soria J et al. The MOF-

driven synthesis of supported palladium clusters with catalytic

activity for carbene-mediated chemistry. Nat Mater 2017; 16:760–6.

Page 67: Special Topic: Single-Atom Catalysts - Western Engineering...Predictive approach of heterogeneous catalysis Metal-organic frame-work assisted synthesis of single-atom catalysts for

688 Natl Sci Rev, 2018, Vol. 5, No. 5 REVIEW

28. Yao S, Zhang X and Zhou W et al. Atomic-layered Au clusters on α-MoC as

catalysts for the low-temperaturewater-gas shift reaction. Science 2017; 357:389–93.

29. Maschmeyer T, Rey F and Sankar G et al. Heterogeneous catalysts obtained

by grafting metallocene complexes onto mesoporous silica. Nat Catal 1995;

378: 159–62.30. Fu Q, Saltsburg H and Flytzani-Stephanopoulos M. Active nonmetallic Au and

Pt species on ceria-based water-gas shift catalysts. Science 2003; 301: 935–8.

31. Hackett SFJ, Brydson RM and Gass MH et al. High-activity, single-site meso-

porous Pd/Al2O3 catalysts for selective aerobic oxidation of allylic alcohols.

Angew Chem 2007; 119: 8747–50.32. Zhu C, Fu S and Shi Q et al. Single-atom electrocatalysts. Angew Chem Int Ed

2017; 56: 13944–60.33. Zheng Y, Jiao Y and Zhu Y et al.Molecule-level g-C3N4 coordinated transition

metals as a new class of electrocatalysts for oxygen electrode reactions. J Am

Chem Soc 2017; 139: 3336–9.34. Zhang H, Hwang S and Wang M et al. Single atomic iron catalysts for oxygen

reduction in acidic media: particle size control and thermal activation. J Am

Chem Soc 2017; 139: 14143–9.35. Ma XL, Liu JC and Xiao H et al. Surface single-cluster catalyst for N2-to-NH3

thermal conversion. J Am Chem Soc 2018; 140: 46–9.36. Zhang W and Zheng W. Single atom excels as the smallest functional mate-

rial. Adv Funct Mater 2016; 26: 2988–93.37. Wang D and Astruc D. The golden age of transfer hydrogenation. Chem Rev

2015; 115: 6621–86.38. Zhang X, Sun Z and Wang B et al. C–C coupling on single-atom-based hetero-

geneous catalyst. J Am Chem Soc 2018; 140: 954–62.39. Guo X, Fang G and Li G et al. Direct, nonoxidative conversion of methane to

ethylene, aromatics, and hydrogen. Science 2014; 344: 616–9.40. Lang R, Li T and Matsumura D et al. Hydroformylation of olefins by a rhodium

single-atom catalyst with activity comparable to RhCl(PPh3)3. Angew Chem

Int Ed 2016; 55: 16054–8.41. Fei H, Dong J and Arellano-JimenezMJ et al.Atomic cobalt on nitrogen-doped

graphene for hydrogen generation. Nat Commun 2015; 6: 8668–75.42. Liang S, Hao C and Shi Y. The power of single-atom catalysis. ChemCatChem

2015; 7: 2559–67.43. Chen Y, Huang Z and Ma Z et al. Fabrication, characterization, and

stability of supported single-atom catalysts. Catal Sci Technol 2017; 7:4250–8.

44. Fei H, Dong J and Feng Y et al. General synthesis and definitive structural

identification of MN4C4 single-atom catalysts with tunable electrocatalytic

activities. Nat Catal 2018; 1: 63–72.45. Xu H, Cheng D and Cao D et al. A universal principle for a rational design of

single-atom electrocatalysts. Nat Catal 2018; doi: 10.1038/s41929-018-0063-

z.

46. Qiao B, Wang A and Yang X et al. Single-atom catalysis of CO oxidation using

Pt1/FeOx. Nat Chem 2011; 3: 634–41.47. Yang S and Lee H. Atomically dispersed platinum on gold nano-octahedrawith

high catalytic activity on formic acid oxidation. ACS Catal 2013; 3: 437–43.48. Kyriakou G, Boucher MB and Jewell AD et al. Isolated metal atom geometries

as a strategy for selective heterogeneous hydrogenations. Science 2012; 335:1209–12.

49. Shi Y, Zhao C and Wei H et al. Single-atom catalysis in mesoporous

photovoltaics: the principle of utility maximization. Adv Mater 2014; 26:8147–53.

50. Lucci FR, Liu J and Marcinkowski MD et al. Selective hydrogenation of 1,3-

butadiene on platinum-copper alloys at the single-atom limit. Nat Commun

2015; 6: 8550–7.51. Serna P and Gates BC. Molecular metal catalysts on supports: organometallic

chemistry meets surface science. Acc Chem Res 2014; 47: 2612–20.52. Thomas JM. The concept, reality and utility of single-site heterogeneous cat-

alysts (SSHCs). Phys Chem Chem Phys 2014; 16: 7647–61.53. Choi CH, KimM and Kwon HC et al. Tuning selectivity of electrochemical reac-

tions by atomically dispersed platinum catalyst.Nat Commun 2016; 7: 10922–30.

54. Zhang H, Watanabe T and Okumura M et al. Catalytically highly active top

gold atom on palladium nanocluster. Nat Mater 2012; 11: 49–52.55. Wang X, Maeda K and Thomas A et al. A metal-free polymeric photocatalyst

for hydrogen production from water under visible light. Nat Mater 2009; 8:76–80.

56. Li X, Bi W and ChenM et al. Exclusive Ni-N4 sites realize near-unity CO selec-

tivity for electrochemical CO2 reduction. J Am Chem Soc 2017; 139: 14889–92.

57. Yin P, Yao T and Wu Y et al. Single cobalt atoms with precise N-coordination

as superior oxygen reduction reaction catalysts. Angew Chem Int Ed 2016; 55:10800–5.

58. Chung HT, Cullen DA and Higgins D et al. Direct atomic-level insight into the

active sites of a high-performance PGM-free ORR catalyst. Science 2017; 357:479–84.

59. Wu Y, Wang D and Li Y. Understanding of the major reactions in solution

synthesis of functional nanomaterials. Sci China Mater 2016; 59: 938–96.60. Wei H, Liu X and Wang A et al. FeOx-supported platinum single-atom and

pseudo-single-atom catalysts for chemoselective hydrogenation of function-

alized nitroarenes. Nat Commun 2014; 5: 5634–41.61. Lin J, Wang A and Qiao B et al. Remarkable performance of Ir1/FeOx single-

atom catalyst in water gas shift reaction. J Am Chem Soc 2013; 135: 15314–7.62. Zhang C, Oliaee SN and Hwang SY et al. A generic wet impregnation method

for preparing substrate-supported platinum group metal and alloy nanoparti-

cles with controlled particle morphology. Nano Lett 2016; 16: 164–9.63. Iqbal S, Kondrat SA and Jones DR et al. Ruthenium nanoparticles supported

on carbon: an active catalyst for the hydrogenation of lactic acid to 1,2-

propanediol. ACS Catal 2015; 5: 5047–59.64. Moses-DeBusk M, Yoon M and Allard LF et al. CO oxidation on supported

single Pt atoms: experimental and ab initio density functional studies of CO

interaction with Pt atom on θ -Al2O3(010) surface. J Am Chem Soc 2013; 135:12634–45.

65. Narula CK, Allard LF and Stocks GM et al. Remarkable NO oxidation on single

supported platinum atoms. Sci Rep 2015; 4: 7238–43.66. Wang L, Zhang S and Zhu Y et al. Catalysis and in situ studies of Rh1/Co3O4

nanorods in reduction of NO with H2. ACS Catal 2013; 3: 1011–9.67. Gu XK, Qiao B and Huang CQ et al. Supported single Pt1/Au1 atoms for

methanol steam reforming. ACS Catal 2014; 4: 3886–90.68. Nie L, Mei D and Xiong H et al. Activation of surface lattice oxygen in single-

atom Pt/CeO2 for low-temperature CO oxidation. Science 2017; 358: 1419–23.69. Zhang Z, Zhu Y and Asakura H et al. Thermally stable single atom Pt/m-Al2O3

for selective hydrogenation and CO oxidation.Nat Commun 2017; 8: 16100–9.70. Moliner M, Gabay JE and Kliewer CE et al. Reversible transformation of Pt

nanoparticles into single atoms inside high-silica chabazite zeolite. J Am

Chem Soc 2016; 138: 15743–50.71. Lu J, Aydin C and Browning ND et al. Imaging isolated gold atom catalytic

sites in zeolite NaY. Angew Chem Int Ed 2012; 51: 5842–6.

Page 68: Special Topic: Single-Atom Catalysts - Western Engineering...Predictive approach of heterogeneous catalysis Metal-organic frame-work assisted synthesis of single-atom catalysts for

REVIEW Li et al. 689

72. Yang M, Li S and Wang Y et al. Catalytically active Au-O(OH)x- species sta-

bilized by alkali ions on zeolites and mesoporous oxides. Science 2014; 346:1498–501.

73. Yang S, Kim J and Tak YJ et al. Single-atom catalyst of platinum supported

on titanium nitride for selective electrochemical reactions. Angew Chem Int

Ed 2016; 55: 2058–62.74. Chen Z, Zhang Q and ChenW et al. Single-site Au1 catalyst for silane oxidation

with water. Adv Mater 2018; 30: 1704720–7.75. Lin L, Zhou W and Gao R et al. Low-temperature hydrogen production from

water and methanol using Pt/α-MoC catalysts. Nature 2017; 544: 80–3.76. Wang X, Chen W and Zhang L et al. Uncoordinated amine groups of metal-

organic frameworks to anchor single Ru sites as chemoselective catalysts to-

ward the hydrogenation of quinoline. J Am Chem Soc 2017; 139: 9419–22.77. Tjoa V, Chua J and Pramana SS et al. Facile photochemical synthesis of

graphene-Pt nanoparticle composite for counter electrode in dye sensitized

solar cell. ACS Appl Mater Interfaces 2012; 4: 3447–52.78. Xu D, Diao P and Jin T et al. Iridium oxide nanoparticles and iridium/iridium

oxide nanocomposites: photochemical fabrication and application in cat-

alytic reduction of 4-nitrophenol. ACS Appl Mater Interfaces 2015; 7:16738–49.

79. ZhanWW, Zhu QL and Dang S et al. Synthesis of highly active sub-nanometer

Pt@Rh core-shell nanocatalyst via a photochemical route: porous titania

nanoplates as a superior photoactive support. Small 2017; 13: 1603879–84.80. Fageria P, Uppala S and Nazir R et al. Synthesis of monometallic (Au and

Pd) and bimetallic (AuPd) nanoparticles using carbon nitride (C3N4) quantum

dots via the photochemical route for nitrophenol reduction. Langmuir 2016;

32: 10054–64.81. Yang M, Allard LF and Flytzani-Stephanopoulos M. Atomically dispersed Au-

(OH)x species bound on titania catalyze the low-temperature water-gas shift

reaction. J Am Chem Soc 2013; 135: 3768–71.82. Liu P, Zhao Y and Qin R et al. Photochemical route for synthesizing atomically

dispersed palladium catalysts. Science 2016; 352: 797–800.83. Wei H, Huang K andWang D et al. Iced photochemical reduction to synthesize

atomically dispersed metals by suppressing nanocrystal growth.Nat Commun

2017; 8: 1490–7.84. Deng D, Chen X and Yu L et al. A single iron site confined in a graphene matrix

for the catalytic oxidation of benzene at room temperature. Sci Adv 2015; 1:e1500462.

85. O’Neill BJ, Jackson DHK and Lee J et al. Catalyst design with atomic layer

deposition. ACS Catal 2015; 5: 1804–25.86. Sun S, Zhang G and Gauquelin N et al. Single-atom catalysis using

Pt/graphene achieved through atomic layer deposition. Sci Rep 2013; 3: 1775–83.

87. Yan H, Cheng H and Yi H et al. Single-atom Pd1/graphene catalyst achieved by

atomic layer deposition: remarkable performance in selective hydrogenation

of 1,3-butadiene. J Am Chem Soc 2015; 137: 10484–7.88. Cheng N, Stambula S and Wang D et al. Platinum single-atom and cluster

catalysis of the hydrogen evolution reaction. Nat Commun 2016; 7: 13638–46.

89. Ren Q, Wang H and Lu XF et al. Recent progress on MOF-derived heteroatom-

doped carbon-based electrocatalysts for oxygen reduction reaction. Adv Sci

2018; 5: 1700515–35.90. Marcinkowski MD, Darby MT and Liu J et al. Pt/Cu single-atom alloys as

coke-resistant catalysts for efficient C–H activation. Nat Chem 2018; doi:

10.1038/NCHEM.2915.

91. Wu Y, Wang D and Zhou G et al. Sophisticated construction of Au islands

on Pt-Ni: an ideal trimetallic nanoframe catalyst. J Am Chem Soc 2014; 136:11594–7.

92. Qin R, Liu P and Fu G et al. Strategies for stabilizing atomically dispersedmetal

catalysts. Small Methods 2018; 2: 1700286–306.93. Jiao L, Wang Y and Jiang HL et al. Metal-organic frameworks as platforms

for catalytic applications. Adv Mater 2017; doi: 10.1002/adma.201703663.

94. Liang Z, Qu C and Guo W et al. Pristine metal-organic frameworks and

their composites for energy storage and conversion. Adv Mater 2017; doi:

10.1002/adma.201702891.

95. Zhao C, Dai X and Yao T et al. Ionic exchange of metal-organic frameworks

to access single nickel sites for efficient electroreduction of CO2. J Am Chem

Soc 2017; 139: 8078–81.96. Wang X, Chen Z and Zhao X et al. Regulation of coordination number over

single Co sites triggers the efficient electroreduction of CO2. Angew Chem Int

Ed 2018; 57:1–6.97. Chen Y, Ji S and Wang Y et al. Isolated single iron atoms anchored on N-

doped porous carbon as an efficient electrocatalyst for the oxygen reduction

reaction. Angew Chem Int Ed 2017; 56: 6937–41.98. Ji S, Chen Y and Fu Q et al. Confined pyrolysis within metal-organic frame-

works to form uniform Ru3 clusters for efficient oxidation of alcohols. J Am

Chem Soc 2017; 139: 9795–8.99. Han Y, Wang YG and Chen W et al. Hollow N-doped carbon spheres with

isolated cobalt single atomic sites: superior electrocatalysts for oxygen re-

duction. J Am Chem Soc 2017; 139: 17269–72.100. Chen W, Pei J and He CT et al. Rational design of single molybdenum

atoms anchored on N-doped carbon for effective hydrogen evolution reaction.

Angew Chem Int Ed 2017; 56: 16086–90.101. Liu W, Zhang L and Yan W et al. Single-atom dispersed Co–N–C catalyst:

structure identification and performance for hydrogenative coupling of ni-

troarenes. Chem Sci 2016; 7: 5758–64.102. LiuW, Zhang L and Liu X et al.Discriminating catalytically active FeNx species

of atomically dispersed Fe–N–C catalyst for selective oxidation of the C–H

bond. J Am Chem Soc 2017; 139: 10790–8.103. Zhang M, Wang YG and Chen W et al. Metal (Hydr)oxides@Polymer core–

shell strategy to metal single-atom materials. J Am Chem Soc 2017; 139:10976–9.

104. Wang J, Huang Z and Liu W et al. Design of N-coordinated dual-metal sites:

a stable and active Pt-free catalyst for acidic oxygen reduction reaction. J Am

Chem Soc 2017; 139: 17281–4.105. Jones J, Xiong H and DeLaRiva AT et al. Thermally stable single-

atom platinum-on-ceria catalysts via atom trapping. Science 2016; 353:150–4.

Page 69: Special Topic: Single-Atom Catalysts - Western Engineering...Predictive approach of heterogeneous catalysis Metal-organic frame-work assisted synthesis of single-atom catalysts for

INTERVIEW National Science Review

5: 690–693, 2018

doi: 10.1093/nsr/nwy043

Advance access publication 26 April 2018

Special Topic: Single-Atom Catalysts

Single-atom catalysis: a new field that learns from traditionBy Philip Ball

Much of industrial chemical processing (in the petrochemicals industry, for example), and a great deal of laboratory chemical synthesis,involves catalysts that both lower the energy barrier to reaction and may help steer a reaction along a particular path. Traditionally,catalysts have come in two classes: heterogeneous, typically meaning that the catalyst is an extended solid; and homogeneous, where thecatalyst is a small molecule that shares a solvent with the reactants. In heterogeneous catalysis, the reaction generally takes place on asurface, involving molecules attached there by covalent bonds. Homogeneous catalysts are often organometallic compounds, in which ametal atom or small cluster of atoms supplies the active site for reaction.

In recent years, these distinctions have become somewhat blurred thanks to the advent of single-atom catalysis, where the catalytic siteconsists of a single atom (as in many homogeneous catalysts) attached to or embedded in a surface. The emergence of this field might beregarded as the logical conclusion of the use of ‘supported metal clusters’—small metal particles of nanometer scale and below, containingperhaps hundreds, tens or just a few atoms. It has became clear that such clusters can sometimes provide greater product selectivity andactivity than macro-sized particles or powders of the same metal, partly because the active sites might be atoms at particular locations (suchas edges and corners) in the nanoscale particles. By reducing their scale down to the level of single atoms, one can optimize these properties.At the same time, the potential uniformity of the atoms’ environments makes such catalysts more amenable to rational design and modelingto understand mechanism.

This field represents an appealing blend of fundamental chemistry and physics—from the quantum-mechanical level upwards—andapplied research aimed at producing many of the products vital to society, such as fuels and materials. Researchers in China have beenstrongly active in this field in recent years (see, for example, refs [1–5]). Jean-Marie Basset of the King Abdullah University of Science andTechnology inThuwal, Saudi Arabia, is one of the leading practitioners in the area, and National Science Review spoke to him about thedevelopment and prospects of the field.

NSR:When was it first appreciated that catalysis by supportedmetals could involve single atoms?Basset: In work that my team and I published in 1998 [6] wediscovered that when we fully characterize a platinum parti-cle modified with tin deposited by the surface organometallicchemistry (SOMC) technique, we found that each Pt atoms issurrounded by tin atoms, and this increased the selectivity ofthe surface for catalysing dehydrogenation of isobutane to al-most 100%. At that time we said that this increase in selectiv-ity could be explained by a ‘site isolation effect’—that the im-portant factor was that the platinum atoms were individuallyisolated.

At about the same time, we discovered that a single zirco-nium atom attached to a silica surface by a triple bond to a Si-O- group could achieve low-temperature hydrogenolysis (split-ting of the carbon backbone using hydrogen) of alkanes. Themethod we used to prepare the lone Zr atoms started with anorganometallic alkyl precursor, and created Zr atoms with a hy-drogen attached [7]. Hydrogenolysis of alkane was a knownreaction in heterogeneous catalysis (for example, on nickelparticles) but here the temperature was much lower (closeto room temperature) than in heterogeneous catalysis (above200◦C). Besides that, the mechanism could be unambiguouslydetermined, and shown to occur on a single Zr atom—it was

Jean-Marie Basset, distinguished professor at King Abdullah University

of Science and Technology, Saudi Arabia. (Courtesy of Prof. Basset)

no longer necessary to invoke any ‘ensemble effect’, a commonnotion in heterogeneous catalysis on small metal particles.NSR: I understand that one of the difficulties with supportedmetal nanoparticles is that they are often inhomogeneous. Is oneof the attractions of single-atom catalysis that homogeneity be-comes possible again?Basset: That’s the right question, and the answer is yes. Withsingle atoms, we can have a situation where all the active

C©TheAuthor(s) 2018. Published by Oxford University Press on behalf of China Science Publishing &Media Ltd. All rights reserved. For permissions, please e-mail:[email protected]

Page 70: Special Topic: Single-Atom Catalysts - Western Engineering...Predictive approach of heterogeneous catalysis Metal-organic frame-work assisted synthesis of single-atom catalysts for

INTERVIEW Ball 691

sites are almost identical. The discovery of low-temperaturehydrogenolysis of alkanes was part of the origin of surfaceorganometallic chemistry. Since then, a huge variety of metalatoms have been attached homogeneously on the surfaces of ox-ides, with the same structure for each atom. SOMC is not par-ticularly familiarwithin the heterogeneous catalysis community,but has led to the discovery of new catalytic reactions such asZiegler-Nattadepolymerization [8], alkanemetathesis [9], non-oxidative coupling of methane [10] and cyclo-alkane metathe-sis [11]. Furthermore it has improved the activity, selectivityor lifetime of known reactions such as alkene metathesis andepoxidation, and imine metathesis. In these cases, the majorityif not all of the active sites are identical. Because the structureof these grafted atoms are known at the atomic and molecularlevel, we can use the familiar concepts of molecular chemistry(organic, organometallic, coordination chemistry) to explainhow bonds can be broken and reformed. The reactivity of sur-face organometallic fragments (SOMF) or surface coordinationfragments (SCF) is pivotal to the outcomes.NSR: In your own work, you have promoted the idea of SOMC,which seems to aim at applying the concepts of homogeneousorganometallic chemistry to surface-bound species. Can you ex-plain more about what this entails?What advantages does a sur-face provide, relative to homogeneous catalysts?Basset: As I said earlier, SOMC is not really an extension of ho-mogeneous catalysis. Rather, it is a new discipline of heteroge-neous catalysis. It uses organometallic compounds to preparewell-defined heterogeneous catalysts in which a single atom islinked to a surface. It is quite distinct from homogeneous cataly-sis, in which typically the catalysts are metal atoms with ligandsattached, because here the ‘ligand’ is a rigid surface, which cre-ates completelydifferent reactivity.Theconcept is actuallymuchcloser to heterogeneous catalysis, because it is like a supportedmetal on an oxide. Or one might better say, it is closest to theconcept of surface-active catalysts (SAC).

You can see the strong difference with classical homoge-neous catalysis, and the consequent benefits, in the way thatmany new reactions which do not exist in homogeneous catal-ysis (or for that matter in heterogeneous catalysis) can beachieved in SOMC.

Here’s a simple way of picturing the comparison between aheterogeneous SOMF used in SOMC and a surface-active cata-lyst:

M'

OO

O

O

Mx

M'

OO

O

O

M

X

B

A

x

A + B

The comparison between SOMF and SAC.

The image on the left shows a SACwhere a metal atomM’ islinked to an oxide. On the right is a SOMF. In SOMC we havea ‘preconceived’mechanism, and the fragments are just possibleintermediates in the catalytic cycle. In SAC it is intuitively as-sumed that the metal will, under the influence of the reagents Aand B, adopt the right coordination sphere. But in SOMC thefragments A and B are components of the SOMF, already at-tached to themetal atom before grafting. Another difference be-tween SOMCand SAC is the presence of a predetermined spec-tator ligand X in the former to tune the coordination number,the electron density and ultimately the steric control around themetal atom.The concepts ofmolecular chemistry are used to de-termine A, B and X, as in homogeneous catalysis—but the sur-face acts as a ligand that brings rigidity, pincer properties, acid-base and redox properties.NSR:Howeasy is it to prepare and characterizewell-defined cat-alysts of this type?Howmuch dowe know, and not know, aboutthe precise environment of the metal atoms?Basset:The preparation of SOMC is becoming ever easier. Wehave improved themethods.At thebeginning itwasnecessary touse fragile organometallic components under a well controlledatmosphere, but now there are techniques to adsorb simple co-ordination complexes on oxides, as was done in many cases inclassical heterogeneous catalysis (for example, WCl6 and TiCl4on silica) and then to alkylate (say) in situ to achieve the rightcoordination sphere.

The characterization is becoming easier, because the sites aremostly identical.The classical techniques can be applied, such assurface microanalysis, IR and UV spectroscopy, extended X-rayabsorption fine structure spectroscopy (EXAFS), X-ray absorp-tion near-edge structure (XANES) and density functional the-ory for calculations.Themost usefulmethod is solid-stateNMR,which has played a decisive role in identifying the SOMFs withthe accuracy of molecular chemistry. When we write a formulaon a surface, it is no longer a cartoon but is very close to the realstructure of most of the sites.NSR:That singlemetal atoms can be important and versatile cat-alytic centers is of course a well-established idea in bioinorganicchemistry too. Is there any overlap with this field in terms of anunderstanding of the mechanisms involved?Basset: SOMC can let us create well-defined SOMFs, thanksto the conceptual overlap with organometallic chemistry. But itcan also lead to well-defined surface coordination compoundsthanks to theoverlapwithbioinorganic chemistry, inorganic andcoordination chemistry.This is thedirection inwhichweare cur-rently obtaining the most spectacular results—unpublished asyet!NSR:What kinds of reactions can be catalysed by these surfaceorganometallic single-atom systems?What are someof themostuseful and/or important?Basset: See the scheme below. Some recent advances includeusing CO2 to make cyclic carbonates, alkane metathesis, con-verting methane to ethane and hydrogen and to aromatics, oxi-dation chemistry of epoxides and aldehydes, and direct transfor-mation of ethylene to propylene. The case of alkane metathesisis particularly important. When we discovered this reaction the

Page 71: Special Topic: Single-Atom Catalysts - Western Engineering...Predictive approach of heterogeneous catalysis Metal-organic frame-work assisted synthesis of single-atom catalysts for

692 Natl Sci Rev, 2018, Vol. 5, No. 5 INTERVIEW

Reactions catalysed by surface organometallic single-atom systems.

turnover number [number of times the catalyst can repeat thereaction] was 60 using hydrogenated tantalum atoms. Now wecan reach turnover numbers of 20 000 with bimetallic systems(tungsten/titanium) [12]. Our target is 100 000, which is whatwe need for such a process to become commercial.NSR:Howpredictable and amenable to rational design are thesesystems? Dowe have the computational methods that we need?Basset:We use the scheme below to do ‘catalysis by design’ anddiscover new reactions or to improve existing ones. The most

Catalysis by design.

important aspect is to transfer the elementary steps known in or-ganic, organometallic and coordination chemistry to write a pri-ori a catalytic cycle. Based on these elementary steps, we choosethe metal, the support and the ligands X, A and B as mentionedearlier, and fully characterize the coordination sphere.NSR:How did your own interest in this field evolve?Basset: My first experiment in this area was to chemisorb aniron carbonyl complex Fe3(CO)12 on alumina in order to makeiron nanoparticles. We were surprised to find formation of thespecies [HFe3(CO)11]− Al+ [13]. This was a shock for us. Itseemed that a new kind of chemistry was emerging from theoverlap between organometallic and surface chemistry.We pro-gressively developed this chemistry in many directions, withmany metals and diverse supports: porous, non-porous, acidic,redox and so on. Then we discovered that this field could alsoapply to nanoparticles of zero-valent metals and we adapted thetools to characterize such materials.

The first catalytic reaction was also a second shock: the low-temperature hydrogenolysis of alkanes with atomic Zr on hy-drogenated silica, which I mentioned earlier. This opened theway to predict Ziegler-Natta depolymerization. Moving from

Page 72: Special Topic: Single-Atom Catalysts - Western Engineering...Predictive approach of heterogeneous catalysis Metal-organic frame-work assisted synthesis of single-atom catalysts for

INTERVIEW Ball 693

�This area has a fantastic future because it allows us toget out of the ‘black box’.

—Jean-Marie Basset

�Zr to Ta, we discovered how to conduct metathesis of alkanes[7] and plenty of new reactions too. Density functional theorywas a crucial tool to understand what was going on, in particu-lar using the tools developed by Luigi Cavallo here atKAUST.The strategy to develop new catalytic reactions slowly emerged,and the concepts are still evolving as we discover new catalyticreactions.NSR:There seems to be a strong interest in this topic in China.Fromwhere do you think themost important results are emerg-ing in China? Do you feel they are building on a strong traditionof inorganic chemistry in China?Basset: Catalysis requires diversified approaches. I’m not surethat theSOMCconcepthas yet been somuchexplored inChina.But Xuxu Wang from Fuzhou, one of my former students inLyon, has made big impact in photocatalysis via SOMC. Nev-ertheless, there have been some great advances in catalysis inChina in the last 20 years or so. Many impressive homoge-neous and heterogeneous systems have been developed: het-erogeneous by Tao Zhang, Can Li, Xinhe Bao, Yuhan Sun, WeiWei andothers; homogeneous byXiaomingFeng,KuilingDing,Zhenfeng Xi, Qilin Zhou, Zhangjie Shi, Aiwen Lei, GuoshengLiu, Shuli You, Zhixiang Yu and many others. However, I’d liketo see more emphasis given to fundamental understanding atthe molecular level. For example, studies on the reactivities oforganometallic species seem less popular, but they build the ba-sis for catalytic applications. There is, however, very good workin this area frompeople likeZuoweiXie, ShaowuWang,YaofengChen, Ming-Hua Zeng and many others.NSR: Where do you feel the field is now heading? Are therepotential types of analytical/characterization techniques thatwould make a big difference to our fundamental understandingof the processes involved?Basset: I feel that this area has a fantastic future because it allowsus to get out of the so-called ‘black box’. This is due to the factthat we have the conceptual and experimental tools to predictany reaction, just by transferring concepts frommolecular chem-istry to surfaces. The science of molecular chemistry, whether itis inorganic, organometallic or organic, teaches us how to createor cleave bonds.Then the choice of metals, ligands, surfaces andso on is becoming more understood.We have some spectacular

new results thatwill explainmyoptimismwhenwepublish themin the near future.NSR:Doyou think that this is one area of chemistry inwhich thelinks between fundamental research and industrial applicationsare particularly strong?Basset: I believe that CO2 chemistry, photocatalytic dissocia-tion of water, CH4 and alkane chemistry, and oxidation are theareas where industry will benefit most from SOMC.NSR: Who were your own key influences in your early career,and why?Basset:When Iwas in France, I recruited YvesChauvinwhen heretired from the French Petroleum Institute [where he workedfrom 1960 to 1995]. Nine years later he was awarded the No-bel Prize in chemistry. Not only was he a friend but I learned alot from his broad knowledge of homogeneous catalysis and in-dustrial processes. I want tomention his strong influence onme,and Iwill always be thankful tohim:hewas amodest, curious butfantastic scientist.

Besides Yves Chauvin, I would like to mention Renato Ugofrom Milan, who in the 1980s was developing analogies be-tween homogeneous and heterogeneous catalysis; Paolo Chini,also from Milan, who made me dream about large clusters;Bob Grubbs from Caltech for his work on olefin metathesis;Wolfgang Herrmann for collaboration on SOMC; and all thecommunity in homogeneous and heterogeneous catalysis, fromwhom I have learnt a lot in two disciplines that have tended todevelop their own concepts separately.

Philip Ball writes for NSR from London.

REFERENCES1. Qiao B, Wang A and Yang X et al. Nat Chem 2011; 3: 634–41.2. Wang X, Chen Z and Zhao X et al. Angew Chem Int Ed 2018; 57: 1944–8.3. Guo X, Fang G and Li G et al. Science 2014; 344: 616–9.4. Liu P, Zhao Y and Qin R et al. Science 2016; 352: 797–800.5. Cao Y, Chen S and Luo Q et al. Angew Chem Int Ed 2017; 56: 12191–6.6. Humblot F, Candy JP and Peltier FL et al. J Catal 1998; 179: 459–68.7. Corker J, Lefebvre F and Lecuyer C et al. Science 1996; 271: 966–9.8. Dufaud V and Basset JM. Angew Chem Int Ed 1998; 37: 806–10.9. Vidal VV, Theolier A and Thivolle-Cazat J et al. Science 1997; 276: 99–102.10. Soulivong D, Norsic S and Taoufik M et al. J Am Chem Soc 2008; 130: 5044–5.11. Riache N, Callens E and Samantaray MK et al. Chem Eur J 2014; 20: 15089–94.12. Samantaray MK, Kavitake S and Morlanes N et al. J Am Chem Soc 2017; 139:

3522–7.

13. Hugues F, Basset JM and Taarit YB et al. J Am Chem Soc 1982; 104:7020–4.