Metallic Iron–Nickel Sulfide Ultrathin Nanosheets ... - ACS Publications

Sep 4, 2015 - Metallic Iron−Nickel Sulfide Ultrathin Nanosheets As a Highly Active. Electrocatalyst for Hydrogen Evolution Reaction in Acidic Media...
1 downloads 0 Views 522KB Size
Subscriber access provided by UNIV OF NEBRASKA - LINCOLN

Communication

Metallic Iron-Nickel Sulfide Ultrathin Nanosheets As a Highly Active Electrocatalyst for Hydrogen Evolution Reaction in Acidic Media Xia Long, Guixia Li, Zilong Wang, Houyu Zhu, Teng Zhang, Shuang Xiao, Wenyue Guo, and Shihe Yang J. Am. Chem. Soc., Just Accepted Manuscript • DOI: 10.1021/jacs.5b07728 • Publication Date (Web): 04 Sep 2015 Downloaded from http://pubs.acs.org on September 4, 2015

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Journal of the American Chemical Society is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 6

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society

Metallic Iron-Nickel Sulfide Ultrathin Nanosheets As a Highly Active Electrocatalyst for Hydrogen Evolution Reaction in Acidic Media Xia Long1, Guixia Li2, Zilong Wang1, HouYu Zhu2, Teng Zhang1, Shuang Xiao1, Wenyue Guo2, Shihe Yang1* 1. Department of Chemistry, William Mong Institute of Nano Science and Technology, The Hong Kong University of Science and Technology, Clear Water Bay, Kowloon, Hong Kong 2. College of Science, China University of Petroleum, Qingdao, PR China Supporting Information ABSTRACT: Developing efficient but low cost electrocatalysts for hydrogen evolution reaction (HER) is critical to sustainable energy conversion technologies. Herein we report on the synthesis of iron-nickel sulfide (INS) ultrathin nanosheets by topotactic conversion from a hydroxide precursor. When tested as an HER electrocatalyst, such nanosheets exhibit excellent activity and stability in strong acidic solutions, lending an attractive alternative to the precious Pt benchmark catalyst and other transition metal HER catalysts that have been extensively investigated. With the metallic α-INS nanosheets, we obtained an even lower over2 potential of 105 mV at 10 mA/cm and a smaller Tafel slope of 40 mV/dec. With the help of DFT calculations, the high specific surface area, facile ion transport and charge transfer, + abundant electrochemical active sites, suitable H adsorption and H2 formation kinetics and energetics are proposed to contribute to the high activity of the INS ultrathin nanosheets towards HER.

The growing demand for energy and environment concerns have stimulated global efforts to explore alternative energy sources to fossil fuels. As an ideal candidate for replacing fossil fuels, hydrogen produced from water splitting [1] is especially attractive. The Pt group metals are most active in catalyzing hydrogen evolution reaction (HER), but the high cost and elemental scarcity greatly hinder their wide[2] spread applications. Therefore, the attention has been turned to earth abundant and low cost materials. Recently, transition metal dichalcogenides (TMD) such as WS2, MoS2, MoSe2, etc. have been surveyed as HER catalysts because of the low cost and high stability on top of the gen[3] eral interest in two-dimensional (2D) materials. Surprisingly, nickel sulfides such as NiS, NiS2, Ni3S4, etc. have rarely been studied as HER catalysts in spite of their extensive use as electrode materials for supercapacitors and Li-ion batter[4] ies. Insomuch as Fe impurity could greatly improve the oxygen evolution reaction (OER) performance of nickel hydroxides/oxides catalysts by changing the local electronic [5] structure, it is reasonable to expect a similar influence of Fe on the HER catalytic activity of nickel sulfides. In the present work, we show that with Fe incorporation, both the HER

catalytic activity and stability of nanosized nickel sulfide (iron-nickel sulfide, INS) ultrathin nanosheets are greatly improved, outperforming most of the known transition metal based HER catalysts in acidic solutions. In particular, the INS nanosheets put at a disadvantage the more widely investigat[3a-f,6] which ed TMD electrocatalysts, such as MoS2 and WS2, exhibit good HER performance only after transformation from the thermodynamically favored semiconducting 2H [3b,3f,6c] phase to the metallic 1T polymorph. By annealing treatment, we obtained metallic α-INS nanosheets, which showed further improved HER performance with a low over2 potential of 105 mV at 10 mA/cm and a small Tafel slope of 40 mV/dec. By combining with theoretical calculations, we find that besides the high specific surface area, fast ionic and electronic transfer/transport, and abundant redox reaction centers resulting from the porous ultrathin nanosheet struc+ ture, the suitable H adsorption and H2 formation kinetics and energetics are responsible for the advanced catalytic performance of the HER catalysts. The β-INS ultrathin nanosheets were synthesized by a simple topotactic conversion reaction, starting from the FeNi layered double hydroxide (LDH) nanosheets (Figure 1A and [7] S1). Scanning electron microscopy (SEM) (Figure 1B) and transmission electron microscopy (TEM) (Figure 1D) images indicate that the nanosheet morphology of the LDH precursor with a lateral size of several micrometers was well preserved after the sulfidation treatment. The corresponding energy dispersive X-ray (EDX) elemental mapping images in Figure S2 verify the existence of Fe, Ni and S and show a uniform distribution of these elements in the as-formed nanosheets with an Fe/Ni atomic ratio of ~1/10, consistent with the XPS result (Table S1). The high resolution TEM (HRTEM) image in Figure 1F shows lattice fringes of (300), (101), and (110) planes, which correspond to d = 0.308, 0.325 and 0.503 nm, respectively, slightly larger than those of rhombohedral β-NiS due to the substitution of Fe for Ni. From the atomic force microscopic (AFM) height images, the thickness of β-INS nanosheet was measured to be around 2 nm (Figure S3A, B), a little larger than that of the LDH precursor (Figure S3C, D). In contrast to the smooth surfaces of the FeNi LDH nanosheets precursor (Figure S1), the β-INS

ACS Paragon Plus Environment

Journal of the American Chemical Society 2

mA/cm was as low as 117 mV (Figure 2A, blue curve), in fact lower than most non-noble metals based HER catalysts reported so far (Table S2). Clearly, the β-INS ultrathin nanosheets are a high performance HER catalyst in acidic solutions. We believe that the high HER activity stems from the rich redox reaction centers of ultrathin nanosheets as [13] well as synergetic effects of the bi-transition metals.

X-ray photoelectron spectroscopy (XPS) showed the existence of Fe, Ni and S in β-INS ultrathin nanosheets (Figure S4B). The Ni 2p XPS spectrum was best fitted with one spin2+ orbit doublet (Figure S5), which is characteristic of Ni (ma[9] genta curves) and two shakeup satellites (blue curves). The Fe 2p high resolution XPS spectrum is deconvoluted into four peaks: two fitted in the Fe 2p1/2 region and two in the Fe 2p3/2 2+ region, which could be accounted for the coexistence of Fe 3+ [10] (magenta curves) and Fe (blue curves). The activity of HER catalyst was reported to be sensitive to the valence state [11] and coordination environment of the metal centers. From the deconvoluted XPS peaks of metal ions, both the Ni and Fe peaks are positively shifted, suggesting strong interactions [12] involving the ions. This may have important implications in modulating electronic environments of the metal centers in β-INS ultrathin nanosheets and thus promoting the HER catalysis. The electrochemical performance of β-INS on HER was investigated in 0.5 M H2SO4 by using a standard threeelectrode setup. Ag/AgCl and platinum (Pt) wire were used as reference and counter electrode, respectively. β-INS catalyst coated on glass carbon (GC) was applied as working electrode. Figure 2A shows the polarization curves of catalyzed HER. As expected, Pt wire showed a superior HER activity with a negligible overpotential (Figure 2A, black curve), whereas bare GC showed no HER activity with a negligible current density even at -0.3 V vs RHE (Figure 2A, purple curves). The β-NiS nanosheets (Figure S6A, B) exhibited a moderate HER activity with an overpotential of 202 mV at 10 2 mA/cm (magenta curves), but the β-INS nanosheets showed a much lower onset potential and the overpotential at 10

-6 -8

-40

-10 -0.3

-0.2

2

α-INS

-30

J (mA/cm )

bare GC

-4

iS

-2

50 mv/s

3 10 mv/s

-0.1

Pt

0.0

-50 -1.2 -1.0 -0.8 -0.6 -0.4 -0.2

2 1 0 -1 -2

0.0

0.10

0.12

E (V vs RHE)

C

D 5

1.5 50 mV/s

0.16

0.18

0.20

5 mV/s

50

60

β-NiS β-INS

4

2

2

1.0

0.14

E (V vs RHE)

∆ J (mA/cm )

The X-ray diffraction (XRD) patterns of the LDH precursor and the β-INS nanosheets are shown in Figure S4A. The diffraction peaks of the LDH precursor are ascribed to (003), (006) and (012) planes of β-Ni(OH)2, positively shifted due to [7,8] the substitution of Fe for Ni. After sulfidation, the β-INS nanosheets show diffraction peaks of (110), (101) and (300) planes of rhombohedral NiS with similar positive shifts, which are consistent with the HRTEM result (Figure 1F) and thus confirm the successful conversion of FeNi LDH to βINS.

-20

0

β-INS

2

J (mA/cm )

-10

Figure 1. Morphology and structure characterizations. SEM images of (A) FeNi LDH precursor, (B) β-INS nanosheets and (C) metallic α-INS nanosheets. TEM images (D, G), HRTEM images (F, I) and the corresponding Fourier transformed patterns (E, H) of the β-INS (D, E, F) and the α-INS (G, H, I) nanosheets. The Fourier transformed patterns (E, H) are rhombohedral and hexagonal, respectively, for the β-INS and α-INS nanosheets.

B 4

0

β-N

A

J (mA/cm )

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

nanosheets are porous with rough surfaces (Figure 1B), indicating the successful transformation from LDH to β-INS.

Page 2 of 6

0.5 0.0 -0.5 -1.0 0.10

3 2 1 0

0.12

0.14

0.16

E (V vs RHE)

0.18

0.20

0

10

20

30

40

Scan rate (mV/s)

Figure 2. Electrochemical performances and electrochemical surface area (ECSA) tests of catalysts towards HER in the acidic electrolytes of 0.5 M H2SO4. (A) Linear sweep voltammetry (LSV) curves of β-NiS, β-INS and α-INS catalyzed HER, (B) CV curves of β-INS and (C) β-NiS nanosheets with various scan rates, (D) charging current density differences plotted against scan rates. The linear slope, equivalent to twice of the double-layer capacitance Cdl, was used to represent the ECSA.

High surface area is usually thought to be one of the most important factors for an advanced electrocatalyst. For the βINS ultrathin nanosheets, the nitrogen adsorptiondesorption isotherm (Figure S7A) measured a Brunauer2 Emmett-Teller (BET) surface area at 109.83 m /g, much high2 er than that of FeNi LDH precursor (54.68 m /g, Figure S7B) 2 and β-NiS nanosheets (80.44 m /g, Figure S7C). We further tested the electrochemical double layer capacitances (Cdl) of [6c,14] the catalysts by a simple cyclic voltammetry method (Figure 2B, 2C) to relate the catalytic activity with the electrochemical surface area (ECSA). From Figure 2D, it is clear that the Cdl of β-INS at 46.4 µF is about three times higher than that of β-NiS (15.5 µF). The increase of ECSA is attributed to the enhanced anion exchangeability between the electrolyte and catalytic active sites. In return, the increased catalytic active sites and enhanced anion accessibility enabled by the porous nanosheet structure resulted in the significant improvement of catalytic performance (vide infra). In order to further improve the HER kinetics, the rhombohedral β-INS nanosheets were annealed in an Ar atmosphere at 450 ºС for 5 h and transformed to hexagonal α-INS, which have metallic properties. From the XRD peaks of α-INS (Figure S4A blue curve), the d spacings are calculated to be 2.99 nm, 2.71 nm, 2.63 nm and 2.02 nm, which are consistent with the HRTEM observation (Figure 1I) and can be ascribed to (100), (002), (101) and (102) planes of α-NiS, respectively, thereby confirming the successful transformation of β-INS to α-INS. The as transformed α-INS still kept the nanosheet structure of β-INS on comparing the SEM (Figure 1C) and TEM (Figure 1F) images. Next, the resistivities of β-INS and

ACS Paragon Plus Environment

Page 3 of 6

Journal of the American Chemical Society -3

A -2.0 -2.2

α-INS

Pt 33 mV/dec

B -70 η=100 mV -60 -50

-2.4

Z"/ohm

log (j/A)

40 mV/dec -2.6 β−INS -2.8 48 mV/dec

-0.15

-10

E (V vs RHE)

-1.0

-0.18 -0.21 -0.24

-1.1

-0.27 0

0

1000

Time (h) 2000 3000

Time (s)

39

4000

0 20 40 60 80 100 120 140 160 180

After

2

-0.12

α-INS β−INS

D 0

J (mA/cm )

-0.09

-20

Z'/ohm

E (V vs RHE) 5 mA 10 mA 20 mA

-0.06

Rct

-30

0

-0.12 -0.10 -0.08 -0.06 -0.04 -0.02 0.00

C

C

Rs

-40

-10

-3.0

E (V vs RHE)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

α-INS were measured at room temperature to be 1.18 × 10 -5 ohm·cm and 5.15 × 10 ohm·cm, respectively, differing by as much as about two orders of magnitude. Plausibly, the α-INS -5 behaves as a metal (~ 10 ohm·cm). Advantageously, the intrinsic metallic character of the α-INS nanosheets conduces to the HER catalysis. First, the intrinsic high conductivity allows to supply charges on demand to sustain and facilitate the HER process. Second, it ensures an ohmic contact at the catalyst-GC electrode and catalyst-electrolyte interfaces for fast charge transfer since no additional overpotential is need[15] ed to overcome the Schottky barrier commonly encountered in HER electrocatalysis. Indeed, from the LSV curve of α-INS catalyzed HER (Figure 2A, red curves), it is clear that the potential needed to achieve the current density of 10 2 mA/cm is positively shifted from -117 mV for β-INS to -105 mV for α-INS, confirming the higher HER activity of α-INS than β-INS nanosheets. It is noteworthy that the electrochemical performance of metallic α-NiS is also enhanced (Figure S8). The overpotential needed to achieve a specific current density was further decreased in our favor with increasing iR compensation level (Figure S9A). Of note, the LSV curves for α-INS nanosheets by using Pt (black curve) and graphite rod (red curve) as the counter electrode are quite similar, except for only a slightly increased overpotential that is required to achieve the same current density for the latter (Figure S9B), indicating that the high HER performance of α-INS nanosheets indeed reflects the intrinsic high catalytic activity of α-INS nanosheets instead of Pt contamination.

5000

-20

Initial -30 -40 -0.18 -0.15 -0.12 -0.09 -0.06 -0.03 0.00

E (V vs RHE)

Figure 3. (A) Tafel plots of Pt, β-INS, and α-INS catalyzed HER, (B) Electrochemical impedance spectra (EIS) of β-INS and α-INS nanosheets catalyzed HER at overpotential of 100 mV, (C) Chronopotentiometry test of α-INS at the current 2 2 2 density of 5 mA/cm , 10 mA/cm and 50 mA/cm , inset: long 2 time CP testing of α-INS at the current density of 10 mA/cm , (D) LSV curves of α-INS nanosheets on HER before and after 2 the CP testing at 10 mA/cm for nearly 40 h. The Tafel slope was determined to evaluate HER kinetics. As shown in Figure 3A, the measured Tafel slope for Pt is about 33 mV/dec, close to the literature values.[16] The Tafel slope of β-INS is around 48 mV/dec, which is further decreased to as low as 40 mV/dec for α-INS. As far as we know, this value is the smallest one to date for nickel sulfide based HER catalysts and comparable to or even smaller than that of many Mo- and W-based HER catalysts shown in Table S2. The low Tafel slope of 40 mV/dec is indicative of the Volmer-Heyrovsky mechanism,[17] which com-

prises fast discharge of a proton and a slow combination of the discharged proton with an additional proton [Had + H3O+ + e-→ H2 + H2O]. The fast electrode reaction kinetics of α-INS is made more evident by electrochemical impedance spectroscopy (EIS). Figure 3B shows the Nyquist plots of α-INS (black curve) and βINS (red curve) nanosheets at the overpotential of 100 mV. The EIS data reveals a much smaller charge transfer resistance (Rct, 63.16 Ω) of the α-INS electrode than that of β-INS (154.5 Ω), suggesting a much faster electron transfer and a high Faradic efficiency during HER. This apparently benefits from the unique electronic properties and the low resistance Ohmic contact of the α-INS nanosheets. Moreover, the BET specific surface area of α-INS is close to that of β-INS nanosheets (Figure S10), while the Cdl of α-INS (56.23 mF, Figure S11) is larger than that of β-INS catalyst, reflecting a relatively larger ECSA associated with more exposed catalytic active sites for the α-INS nanosheets, leading to a larger exchange current density of α-INS than that of β-INS (Figure S12).

Yet another important criterion for an HER catalyst is its operational stability. For this purpose, we chose α-INS nanosheet, and tested it by chronopotentiometry (Figure 3C). The potential needed to achieve the current density of 5 2 2 2 mA/cm , 10 mA/cm and 20 mA/cm was around -0.082 V, 0.105 V, and -0.13 V, respectively, which declined slightly over the 3500 second period, indicating good stability of the catalyst at high current densities. When the chronopotentiome2 try testing at the current density of 10 mA/cm was prolonged to nearly 40 h (Figure 4C inset), the potential showed a negligible change as well. Moreover, after the chronopotentiometry test, the polarization curve showed a very small negative shift (Figure 3D, red curve) compared with the initial one (Figure 4D, black curve), further confirming the long-term superior stability of the α-INS ultrathin towards HER. To understand the catalytic efficacies arising from the Fe incorporation, the detailed HER pathways on both the α-NiS and α-INS catalysts were investigated by DFT calculations (Figure 4, also see SI for details). In the DFT calculation, a proton was firstly absorbed in a surface interstice surrounded + by S, Fe and Ni atoms (Figure 4A1). Then another proton (H ) gets close to the absorbed proton (Had) (Figure 4A2) to form an absorbed H2. After migrating to a metal site (Figure 4A3), the absorbed H2 finally makes its way to desorption. Note that while the product H2 is formed on the Ni site on the surface of the NiS catalyst, it is more preferable to form on the Fe site of the INS catalyst. Importantly, the fluent product migration rapidly vacates the active site for the next HER turn-over cycle, promoting the overall catalytic activity. It thus appears that the change of electronic structure of the catalytically active center arising from the Fe-incorporation can modulate and facilitate the HER process. Therefore, our INS catalysts work in a metal-ligand cooperative catalysis mode, similar to that of metal complex catalysts and hydrogenases towards HER. For a metal complex catalyst, it incorporates proton that relays from pendant acid-base groups that position close to the metal center where hydrogen pro[18] duction occurs, and the active sites of hydrogenase feature [19] pendant base proximate to the metal centers as well. In our case, with pendant bases S close to the metal centers of Fe and Ni, the INS nanosheets have a similar structure and chemical composition to those of NiFeS2(μ-H) core in the active site of [NiFe]-H2ases hydrogenase and NiFe thiolato [18a,20] hydride complex. Therefore, it is reasonable that the INS ultrathin nanosheets might share a similar catalytic

ACS Paragon Plus Environment

Journal of the American Chemical Society

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

mechanism to the metal complex catalysts and hydrogenases toward the HER.

Page 4 of 6

Experimental procedures and supporting data including SEM, TEM, HRTEM, AFM, EDX of catalysts. This material is available free of charge via the Internet at http://pubs.acs.org. Corresponding Author [email protected] Notes The authors declare no competing financial interests.

ACKNOWLEDGMENT This work was supported by the NSFC/RGC Joint Research Scheme (N_HKUST610/14) and the HK-RGC General Research Funds (HKUST 606511). Figure 4. (A) Schematic reaction pathway of HER on α-INS ultrathin nanosheets in acid environment. (B) The kinetic energy barrier profiles of HER on α-INS and α-NiS nanosheets. The yellow, blue and red spheres in A1-A3 represent S, Ni, and Fe atoms, respectively.

In general, the activity of an HER catalyst is related to the kinetic barrier of the rate-determining H2 evolution pathway. To shed more light on the Fe-enhanced catalytic activity, we further investigated the energy profiles along the HER pathway on α-INS and α-NiS slabs by DFT (Figure 4B). As mentioned above, the rate-determining step is thought to be the + combination of Had with H based on the Tafel slope of ~40 mV/dec. Significantly, the energy barrier for this step is calculated to be only 0.01 eV on the α-INS slab, which is much lower than that on NiS (0.19 eV). Moreover, the released energy for the formation of H2 on α-INS slab is 2.39 eV, which is, also advantageously, larger than that on NiS slab (2.19 eV). Therefore, drawing on the simulation results, the higher catalytic activity of the INS nanosheets than the pristine NiS nanosheets on HER in acidic solution is attributed to the + lower energy barrier for H adsorption and the higher exothermicity for H2 formation. For comparison, the HER pathways on β-INS and β-NiS catalysts were also investigated (Figure S13), which also accord well with the experimental results. In summary, β-INS ultrathin nanosheets have been successfully synthesized by topotactic sulfidation of FeNi LDH precursors and proved to be excellent HER catalysts in strong acidic solutions. What is more, metallic α-INS ultrathin nanosheets obtained by annealing treatment of the β-INS show further improved HER performance with a low overpo2 tential of 105 mV for 10 mA/cm , a small Tafel slope of 40 mV/dec, and long time stability. Some important features of the ultrathin HER catalyst are believed to have played a key role: (1) the rich and accessible catalytic active sites, (2) the felicitous electronic environments of the bi-transition metal system that enhance HER catalytic activity and durability, and (3) the intrinsic metallic character of α-INS, which further hastens charge transfer and forestalls any possible overpotential from Schottky barriers. Our DFT calculations have + shown that the suitable kinetics and energetics for the H adsorption and H2 formation contribute to the advanced HER catalytic activity. To our knowledge, this is the first study on multi-transition metal based sulfide nanosheets with atomic thickness as HER catalysts with a remarkable electrocatalytic performance, opening up an exciting avenue to expand the family of novel low cost HER catalysts with high activity and long term stability.

REFERENCES [1] a). Dresselhaus M.; Thomas I. Nature 2001, 414, 332. b). Cook T. R.; Dogutan D. K.; Reece S. Y.; Surendranath Y.; Teets T. S.; Nocera D. G. Chem. Rev. 2010, 110, 6474. c). Walter M. G.; Warren E. L.; McKone J. R.; Boettcher S. W.; Mi Q.; Santori E. A.; Lewis N. S. Chem. Rev. 2010, 110, 6446. [2] Dasgupta N. P.; Liu C.; Andrews S.; Prinz F. B.; Yang P. J. Am. Chem. Soc. 2013, 135, 12932. [3] a). Wang H.; Lu Z.; Xu S.; Kong D.; Cha J. J.; Zheng G.; Hsu P.-C.; Yan K.; Bradshaw D.; Prinz F. B.; Cui Y. P. Natl. Acad. Sci. USA 2013, 110, 19701. b). Voiry D.; Yamaguchi H.; Li J.; Silva R.; Alves D. C.; Fujita T.; Chen M.; Asefa T.; Shenoy V. B.; Eda G.; Chhowalla M. Nat. Mater. 2013, 12, 850. c). Wang H.; Lu Z.; Kong D.; Sun J.; Hymel T. M.; Cui Y. ACS nano 2014, 8, 4940. d). Jaramillo T. F.; Jørgensen K. P.; Bonde J.; Nielsen J. H.; Horch S.; Chorkendorff I. Science 2007, 317, 100. e). Hinnemann B.; Moses P. G.; Bonde J.; Jørgensen K. P.; Nielsen J. H.; Horch S.; Chorkendorff I.; Nørskov J. K. J. Am. Chem. Soc. 2005, 127, 5308. f). Jin S.; Lukowski M. A.; Daniel A. S.; English C. R.; Meng F.; Forticaux A.; Hamers R. Energ. Environ. Sci. 2014, 7, 2608. g). Wang H.; Kong D.; Johanes P.; Cha J. J.; Zheng G.; Yan K.; Liu N.; Cui Y. Nano letters 2013, 13, 3426. [4] a). Yang J.; Duan X.; Qin Q.; Zheng W. J. Mater. Chem. A 2013, 1, 7880. b). Olivas A.; Cruz-Reyes J.; Petranovskii V.; Avalos M.; Fuentes S. J. Vac. Sci. Technol., A 1998, 16, 3515. c). Hou L.; Yuan C.; Li D.; Yang L.; Shen L.; Zhang F.; Zhang X. Electrochimica Acta 2011, 56, 7454. d). Zhu T.; Wang Z.; Ding S.; Chen J. S.; Lou X. W. D. RSC Adv. 2011, 1, 397. e). Pan Q.; Xie J.; Liu S.; Cao G.; Zhu T.; Zhao X. RSC Adv. 2013, 3, 3899. f). Ni S.; Yang X.; Li T. J. Mater. Chem. 2012, 22, 2395. g). Sun H.; Qin D.; Huang S.; Guo X.; Li D.; Luo Y.; Meng Q. Energ. Environ. Sci. 2011, 4, 2630. [5] a). Trotochaud L.; Young S. L.; Ranney J. K.; Boettcher S. W. J. Am. Chem. Soc. 2014, 136, 6744. b). Hunter B. M.; Blakemore J. D.; Deimund M.; Gray H. B.; Winkler J. R.; Müller A. M. J. Am. Chem. Soc. 2014, 136, 13118. c). Friebel D.; Louie M. W.; Bajdich M.; Sanwald K. E.; Cai Y.; Wise A. M.; Cheng M.-J.; Sokaras D.; Weng T.-C.; AlonsoMori R. Davis R. C.; Bargar J. R.; Nørskov J. K.; Nilsson A.; Bell A. T. J. Am. Chem. Soc. 2015, 137, 1305. d). Smith A. M.; Trotochaud L.; Burke M. S.; Boettcher S. W. Chem. Commun. 2015, 51, 5261. [6] a). Cheng L.; Huang W.; Gong Q.; Liu C.; Liu Z.; Li Y.; Dai H. Angew. Chem. Int. Ed. 2014, 53, 7860. b). Lin J.; Peng Z.; Wang G.; Zakhidov D.; Larios E.; Yacaman M. J.; Tour J. M. Adv. Energ. Mater. 2014, 4, 1301875. c). Lukowski M. A.; Daniel A. S.; Meng F.; Forticaux A.; Li L.; Jin S. J. Am. Chem. Soc. 2013, 135, 10274. d). Xie J.; Zhang H.; Li S.; Wang R.; Sun X.; Zhou M.; Zhou J.; Lou X. W. D.; Xie Y. Adv. Mater. 2013, 25, 5807. [7] a). Long X.; Li J.; Xiao S.; Yan K.; Wang Z.; Chen H.; Yang S. Angew. Chem. 2014, 126, 7714. b). Long X.; Xiao S.; Wang Z.; Zheng X.; Yang S. Chem. Commun. 2015, 51, 1120. [8] Gong M.; Li Y.; Wang H.; Liang Y.; Wu J. Z.; Zhou J.; Wang J.; Regier T.; Wei F.; Dai H. J. Am. Chem. Soc. 2013, 135, 8452. [9] a). Huang L.; Chen D.; Ding Y.; Wang Z. L.; Zeng Z.; Liu M. ACS Appl. Mater. Inter. 2013, 5, 11159. b). Yuan C.; Li J.; Hou L.; Zhang X.; Shen L.; Lou X. W. D. Adv. Funct. Mater. 2012, 22, 4592. c). Shen L.; Che Q.; Li H.; Zhang X. Adv. Funct. Mater. 2014, 24, 2630.

Supporting Information

ACS Paragon Plus Environment

Page 5 of 6

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society

[10] Thomas J. E.; Skinner W. M.; Smart R. S. Geochim. Cosmochim. Ac. 2003, 67, 831. [11] a). Cao B.; Veith G. M.; Neuefeind J. C.; Adzic R. R.; Khalifah P. G. J. Am. Chem. Soc. 2013, 135, 19186. b). Kong D.; Wang H.; Lu Z.; Cui Y. J. Am. Chem. Soc. 2014, 136, 4897. [12] Zhao Z.; Wu H.; He H.; Xu X.; Jin Y. Adv. Funct. Mater. 2014, 24, 4698. [13] a). Liu Q.; Jin J.; Zhang J. ACS Appl. Mater. Inter. 2013, 5, 5002. b). Peng S.; Li L.; Li C.; Tan H.; Cai R.; Yu H.; Mhaisalkar S.; Srinivasan M.; Ramakrishna S.; Yan Q. Chem. Commun. 2013, 49, 10178. c). Wang D-Y.; Gong M.; Chou G-L.; Pan C-J.; Chen H-A.; Wu Y.; Lin MC.; Guan M.; Yang J.; Chen C-W.; Wang Y-L.; Hwang B-J.; Chen C-C.; Dai H. J. Am. Chem. Soc. 2015, 137, 1587. [14] Merki D.; Vrubel H.; Rovelli L.; Fierro S.; Hu X. Chem. Sci. 2012, 3, 2515.

[165] Xu K.; Chen P.; Li X.; Tong Y.; Ding H.; Wu X.; Chu W.; Peng Z.; Wu C.; Xie Y. J. Am. Chem. Soc. 2015, 137, 4119. [16] a). Tian J.; Liu Q.; Asiri A. M.; Sun X. J. Am. Chem. Soc. 2014, 136, 7587. b). Popczun E. J.; McKone J. R.; Read C. G.; Biacchi A. J.; Wiltrout A. M.; Lewis N. S.; Schaak R. E. J. Am. Chem. Soc. 2013, 135, 9267. [17] Conway B.; Tilak B. Electrochim. Acta 2002, 47, 3571. [18] a). Barton B. E.; Rauchfuss T. B. J. Am. Chem. Soc. 2010, 132, 14877. b). Wilson A. D.; Shoemaker R.; Miedaner A.; Muckerman J.; DuBois D. L.; DuBois M. R. P. Natl. Acad. Sci. USA 2007, 104, 6951. [19] Nicolet Y.; de Lacey A. L.; Vernede X.; Fernandez V. M.; Hatchikian E. C.; Fontecilla-Camps J. C. J. Am. Chem. Soc. 2001, 123, 1596. [20] Barton B. E.; Whaley C. M.; Rauchfuss T. B.; Gray D. L. J. Am. Chem. Soc. 2009, 131, 6942.

ACS Paragon Plus Environment

Journal of the American Chemical Society

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

SYNOPSIS

Page 6 of 6 TOC

ACS Paragon Plus Environment