hierarchical ni-mo-s nanosheets on carbon fiber …...(ph ~7), for example, current density of 10...

15
NANOMATERIALS 2015 © The Authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. Distributed under a Creative Commons Attribution NonCommercial License 4.0 (CC BY-NC). 10.1126/sciadv.1500259 Hierarchical Ni-Mo-S nanosheets on carbon fiber cloth: A flexible electrode for efficient hydrogen generation in neutral electrolyte Jianwei Miao, 1 Fang-Xing Xiao, 1 Hong Bin Yang, 1 Si Yun Khoo, 1 Jiazang Chen, 1 Zhanxi Fan, 2 Ying-Ya Hsu, 3 Hao Ming Chen, 3 Hua Zhang, 2 * Bin Liu 1 * A unique functional electrode made of hierarchal Ni-Mo-S nanosheets with abundant exposed edges anchored on conductive and flexible carbon fiber cloth, referred to as Ni-Mo-S/C, has been developed through a facile biomolecule-assisted hydrothermal method. The incorporation of Ni atoms in Mo-S plays a crucial role in tuning its intrinsic catalytic property by creating substantial defect sites as well as modifying the morphology of Ni-Mo-S network at atomic scale, resulting in an impressive enhancement in the catalytic activity. The Ni-Mo-S/C electrode exhibits a large cathodic current and a low onset potential for hydrogen evolution reaction in neutral electrolyte (pH ~7), for example, current density of 10 mA/cm 2 at a very small overpotential of 200 mV. Furthermore, the Ni-Mo-S/C electrode has excellent electrocatalytic stability over an extended period, much better than those of MoS 2 /C and Pt plate electrodes. Scanning and transmission electron microscopy, Raman spectroscopy, x-ray diffrac- tion, x-ray photoelectron spectroscopy, and x-ray absorption spectroscopy were used to understand the formation process and electrocatalytic properties of Ni-Mo-S/C. The intuitive comparison test was designed to reveal the su- perior gas-evolving profile of Ni-Mo-S/C over that of MoS 2 /C, and a laboratory-scale hydrogen generator was further assembled to demonstrate its potential application in practical appliances. INTRODUCTION On the journey to pursue clean and sustainable energy resources, elec- trocatalytic and photoelectrocatalytic water splitting have attracted grow- ing attention because of their potential applications in producing hydrogen with ultrahigh purity while generating negligible greenhouse gases (1, 2). Hydrogen evolution reaction (HER), in which protons are reduced to molecular hydrogen, is a critical process for electrocatalytic water split- ting (2). Although platinum (Pt) is the most effective HER electro- catalyst, the scarcity and expensiveness limit its practical applications for large-amount hydrogen generation (3, 4). Therefore, exploration of alternative earth-abundant materials with low cost and high catalytic activity is essential for developing a high-performance electrocatalyst toward large-amount hydrogen production. MoS 2 , a layered transition metal dichalcogenide, has been shown as a versatile functional material for various applications ( 57), including lithium-ion batteries (810), nanoelectronics (11, 12), hydrodesulfur- ization (13), and HER catalysts (14). The earliest observation of hydro- gen evolution on MoS 2 can be dated back to more than 40 years ago (15). Later, in 1991, Sobczynski reported MoS 2 as an HER cocatalyst for water photodecomposition on semiconductors (16). Since then, MoS 2 -based catalysts have been thoroughly studied and widely recog- nized as ideal catalysts for hydrogen generation in acidic medium ( 17). More recently, a number of computational and experimental studies sug- gest that the HER activities of nano-sized MoS 2 particles originate from the edge sites while their basal planes are catalytically inactive (18). To fully exploit the advantageous catalytic properties, well-defined MoS 2 nanostructures with abundant exposed catalytic active sites and enhanced intrinsic properties should be rationally designed. Var- ious strategies have been proposed and reported for improving the catalytic activity of MoS 2 -based HER catalysts. For example, Li et al. reported the synthesis of MoS 2 nanoparticles anchored on graphene through a solvothermal approach. The composite catalyst showed su- perior catalytic performance owing to the abundant exposed edges as well as the strong chemical and electronic coupling between graphene oxide (GO) and MoS 2 (19). Kibsgaard et al. successfully engineered MoS 2 into highly ordered, double-gyroid network, in which active edge sites in MoS 2 are preferentially exposed for electrocatalysis (20). Re- cently, Wang et al. showed the preparation of vertically aligned MoS 2 nanofilms with improved HER activity by electrochemical lithium inter- calation ( 21). Lu et al . demonstrated that the aerophobic surface of na- nostructured MoS 2 is more beneficial for releasing gas bubbles during HER as compared to the planar thin film (22). Besides the aforemen- tioned design and construction of nanostructures, controllable dis- order engineering is another approach that can be used to improve the catalytic activities of MoS 2 . For example, Xie et al. reported the syntheses of defect-rich MoS 2 ultrathin nanosheets through controlling experimental conditions such as precursor concentration and reaction temperature (23, 24). They discovered that the introduction of defect sites could lead to the exposure of additional catalytic centers, thus sig- nificantly improving the catalytic performance of MoS 2 . All of the aforementioned strategies are important and provide us valuable guidelines to further excavate the potential of MoS 2 -based electrocatalysts. However, most of the current studies on MoS 2 -based electrocatalysts are only viable for laboratory-scale applications under ideal experimental conditions, and are not ready to be applied in prac- tical industrial appliances. For the purpose of developing qualified H 2 gasevolving electrodes that can be integrated in large-scale reactors, the following criteria have to be fulfilled. First, the attached HER catalyst must exhibit extraordinary catalytic activity and reasonable stability for long-term operation. Second, the catalysts and their supporting materials should be abundant and cost-effective and able to be mass-produced. 1 School of Chemical and Biomedical Engineering, Nanyang Technological University, Singapore 637459, Singapore. 2 School of Materials Science and Engineering, Nanyang Technological University, Singapore 639798, Singapore. 3 Department of Chemistry, National Taiwan University, Taipei 106, Taiwan. *Corresponding author. E-mail: [email protected] (B.L.); [email protected] (H.Z.) RESEARCH ARTICLE Miao et al. Sci. Adv. 2015;1:e1500259 21 August 2015 1 of 14 on March 19, 2020 http://advances.sciencemag.org/ Downloaded from

Upload: others

Post on 16-Mar-2020

8 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Hierarchical Ni-Mo-S nanosheets on carbon fiber …...(pH ~7), for example, current density of 10 mA/cm2 at a very small overpotential of 200 mV. Furthermore, the Ni-Mo-S/C electrode

R E S EARCH ART I C L E

NANOMATER IALS

1School of Chemical and Biomedical Engineering, Nanyang Technological University,Singapore 637459, Singapore. 2School of Materials Science and Engineering, NanyangTechnological University, Singapore 639798, Singapore. 3Department of Chemistry,National Taiwan University, Taipei 106, Taiwan.*Corresponding author. E-mail: [email protected] (B.L.); [email protected] (H.Z.)

Miao et al. Sci. Adv. 2015;1:e1500259 21 August 2015

2015 © The Authors, some rights reserved;

exclusive licensee American Association for

the Advancement of Science. Distributed

under a Creative Commons Attribution

NonCommercial License 4.0 (CC BY-NC).

10.1126/sciadv.1500259

Hierarchical Ni-Mo-S nanosheets on carbon fibercloth: A flexible electrode for efficient hydrogengeneration in neutral electrolyte

Jianwei Miao,1 Fang-Xing Xiao,1 Hong Bin Yang,1 Si Yun Khoo,1 Jiazang Chen,1 Zhanxi Fan,2 Ying-Ya Hsu,3

Hao Ming Chen,3 Hua Zhang,2* Bin Liu1*

htD

ownloaded from

A unique functional electrode made of hierarchal Ni-Mo-S nanosheets with abundant exposed edges anchoredon conductive and flexible carbon fiber cloth, referred to as Ni-Mo-S/C, has been developed through a facilebiomolecule-assisted hydrothermal method. The incorporation of Ni atoms in Mo-S plays a crucial role in tuning itsintrinsic catalytic property by creating substantial defect sites as well as modifying the morphology of Ni-Mo-Snetwork at atomic scale, resulting in an impressive enhancement in the catalytic activity. The Ni-Mo-S/C electrodeexhibits a large cathodic current and a low onset potential for hydrogen evolution reaction in neutral electrolyte(pH ~7), for example, current density of 10 mA/cm2 at a very small overpotential of 200 mV. Furthermore, theNi-Mo-S/C electrode has excellent electrocatalytic stability over an extended period, much better than those ofMoS2/C and Pt plate electrodes. Scanning and transmission electron microscopy, Raman spectroscopy, x-ray diffrac-tion, x-ray photoelectron spectroscopy, and x-ray absorption spectroscopy were used to understand the formationprocess and electrocatalytic properties of Ni-Mo-S/C. The intuitive comparison test was designed to reveal the su-perior gas-evolving profile of Ni-Mo-S/C over that of MoS2/C, and a laboratory-scale hydrogen generator was furtherassembled to demonstrate its potential application in practical appliances.

tp:/

on M

arch 19, 2020/advances.sciencem

ag.org/

INTRODUCTION

On the journey to pursue clean and sustainable energy resources, elec-trocatalytic and photoelectrocatalytic water splitting have attracted grow-ing attention because of their potential applications in producing hydrogenwith ultrahigh purity while generating negligible greenhouse gases (1, 2).Hydrogen evolution reaction (HER), in which protons are reduced tomolecular hydrogen, is a critical process for electrocatalytic water split-ting (2). Although platinum (Pt) is the most effective HER electro-catalyst, the scarcity and expensiveness limit its practical applicationsfor large-amount hydrogen generation (3, 4). Therefore, exploration ofalternative earth-abundant materials with low cost and high catalyticactivity is essential for developing a high-performance electrocatalysttoward large-amount hydrogen production.

MoS2, a layered transition metal dichalcogenide, has been shown asa versatile functional material for various applications (5–7), includinglithium-ion batteries (8–10), nanoelectronics (11, 12), hydrodesulfur-ization (13), and HER catalysts (14). The earliest observation of hydro-gen evolution on MoS2 can be dated back to more than 40 years ago(15). Later, in 1991, Sobczynski reported MoS2 as an HER cocatalystfor water photodecomposition on semiconductors (16). Since then,MoS2-based catalysts have been thoroughly studied and widely recog-nized as ideal catalysts for hydrogen generation in acidic medium (17).More recently, a number of computational and experimental studies sug-gest that the HER activities of nano-sized MoS2 particles originatefrom the edge sites while their basal planes are catalytically inactive(18). To fully exploit the advantageous catalytic properties, well-definedMoS2 nanostructures with abundant exposed catalytic active sites andenhanced intrinsic properties should be rationally designed. Var-

ious strategies have been proposed and reported for improving thecatalytic activity of MoS2-based HER catalysts. For example, Li et al.reported the synthesis of MoS2 nanoparticles anchored on graphenethrough a solvothermal approach. The composite catalyst showed su-perior catalytic performance owing to the abundant exposed edges aswell as the strong chemical and electronic coupling between grapheneoxide (GO) and MoS2 (19). Kibsgaard et al. successfully engineeredMoS2 into highly ordered, double-gyroid network, in which active edgesites in MoS2 are preferentially exposed for electrocatalysis (20). Re-cently, Wang et al. showed the preparation of vertically aligned MoS2nanofilms with improved HER activity by electrochemical lithium inter-calation (21). Lu et al. demonstrated that the aerophobic surface of na-nostructured MoS2 is more beneficial for releasing gas bubbles duringHER as compared to the planar thin film (22). Besides the aforemen-tioned design and construction of nanostructures, controllable dis-order engineering is another approach that can be used to improvethe catalytic activities of MoS2. For example, Xie et al. reported thesyntheses of defect-rich MoS2 ultrathin nanosheets through controllingexperimental conditions such as precursor concentration and reactiontemperature (23, 24). They discovered that the introduction of defectsites could lead to the exposure of additional catalytic centers, thus sig-nificantly improving the catalytic performance of MoS2.

All of the aforementioned strategies are important and provide usvaluable guidelines to further excavate the potential of MoS2-basedelectrocatalysts. However, most of the current studies on MoS2-basedelectrocatalysts are only viable for laboratory-scale applications underideal experimental conditions, and are not ready to be applied in prac-tical industrial appliances. For the purpose of developing qualified H2

gas–evolving electrodes that can be integrated in large-scale reactors,the following criteria have to be fulfilled. First, the attached HER catalystmust exhibit extraordinary catalytic activity and reasonable stability forlong-term operation. Second, the catalysts and their supporting materialsshould be abundant and cost-effective and able to be mass-produced.

1 of 14

Page 2: Hierarchical Ni-Mo-S nanosheets on carbon fiber …...(pH ~7), for example, current density of 10 mA/cm2 at a very small overpotential of 200 mV. Furthermore, the Ni-Mo-S/C electrode

R E S EARCH ART I C L E

Dow

nloaded from

Third, the catalysts, substrates, and operating environment have to benontoxic and eco-friendly, in accordance with the intention of devel-oping sustainable energy resources. Last, the electrodes should haveexcellent mechanical strength and physical flexibility, which allowthem to be integrated in varied reactors to meet different requirementsand standards. Thanks to the rapid development in the theoretical cal-culation as well as experimental techniques, the progress of developinggas-evolving electrodes for practical applications has been significantlyaccelerated over the past few years (25–28).

Herein, we report a facile synthetic strategy to directly grow nano-structured Ni-Mo-S on three-dimensional (3D) flexible and conduc-tive carbon fiber cloth substrate. We found that the introduction ofnickel ion (Ni2+) in the precursor solution played a crucial role in tailor-ing the morphology and intrinsic properties of MoS2, hence effectivelyaffecting their catalytic behaviors. The optimized Ni-Mo-S/C electrodeshowed pronounced HER electrocatalytic activity in neutral electrolytewith remarkable stability. The as-synthesized Ni-Mo-S catalysts were char-acterized by scanning and transmission electron microscopies (SEM andTEM), Raman spectroscopy, x-ray diffraction (XRD), x-ray photo-electron spectroscopy (XPS), and x-ray absorption spectroscopy. A pro-totype of a laboratory-scale H2 generator was designed to show theapplicability of the Ni-Mo-S/C electrode for future large-scale applications.

http://advances.sciencemag.

RESULTS

Here, our first attempt was to synthesize nanostructured MoS2 ontovarious carbon-based substrates, including carbon fiber cloth (fig. S1)and graphite foil, which can be directly used as functional electrodesfor hydrogen evolution. In these experiments, a biomolecule-assistedhydrothermal synthetic route with Na2MoO4 · 2H2O and L-cysteine (L-Cys)as the Mo and S sources, respectively, was used to synthesize nanostruc-

Miao et al. Sci. Adv. 2015;1:e1500259 21 August 2015

tured MoS2. As one type of amino acid, the L-Cys molecules can beassembled into a polymeric network structure through the formationof peptide bonds and disulfide bonds from the dehydration conden-sation and sulfhydryl (-SH) group oxidation reactions. The as-formed3D network is beneficial for the construction of nanostructured ma-terials. Upon heating to a certain temperature, L-Cys starts to decom-pose and releases H2S that is capable of reducing MoO4

2− to MoS2 onthe basis of the following reactions (8, 29):

HSCH2CHNH2COOH þ H2O → CH3COCOOH þ NH3 þ H2S

o

4MoO2−4 þ 9H2S þ 6CH3COCOOH → 4MoS2 þSO2−

4 þ 6CH3COCOO− þ 12 H2O

However, it was found that MoS2 nanostructures prepared by thismethod could easily coalesce and stack together to form bulky andnon-uniform aggregates on carbon fibers. The coalescence and stack-ing of nanostructures are detrimental for MoS2-based electrocatalystsbecause aggregation leads to a drastic decrease in numbers of exposedactive sites along the edges. The aggregation is attributed to the epitaxialnucleation and growth of layered MoS2, which is further elaborated inthe Supplementary Materials (figs. S2 and S3 and note S1).

To tailor the morphologies and intrinsic properties of MoS2 on car-bon fibers, the synthetic strategy was modified by introducing Ni2+ intothe precursor solution. As shown in Fig. 1A, MoS2 and Ni-Mo-S sam-ples with different Ni-to-Mo precursor ratios, namely, 3:1, 1:1, and 1:3,were directly synthesized on carbon fiber cloth, designated as MoS2/Cand Ni-Mo-S/C (3:1), Ni-Mo-S/C (1:1), and Ni-Mo-S/C (1:3), re-spectively. For comparison purpose, NiSx/C was also synthesized usinga similar hydrothermal approach. Figure 1B shows a photographicimage of a freshly synthesized Ni-Mo-S/C (1:1) with a dimension of5 × 3 cm.

on March 19, 2020

rg/

Fig. 1. Synthesis processes of MoS2/C and Ni-Mo-S/C. (A) Schematic illustration of syntheses of MoS2 and Ni-Mo-S on carbon fiber cloth. (B) Photo-graphic image of freshly prepared Ni-Mo-S/C (1:1).

2 of 14

Page 3: Hierarchical Ni-Mo-S nanosheets on carbon fiber …...(pH ~7), for example, current density of 10 mA/cm2 at a very small overpotential of 200 mV. Furthermore, the Ni-Mo-S/C electrode

R E S EARCH ART I C L E

http://advanceD

ownloaded from

Figure 2 shows the typical low-magnification SEM images of barecarbon fiber from the cloth, NiSx/C, Ni-Mo-S/C, and MoS2/C. Asdisplayed in Fig. 2A, the diameter of a single carbon fiber is ~10 mm.Oriented striations that formed during the manufacturing process arealong the entire surface of fibers. Figure 2B shows the morphology ofas-prepared NiSx/C, where the irregularly shaped NiSx particles withsizes ranging from 100 nm to 1 mm form a shell-like outer layer thatcovers the carbon fiber surface. However, the NiSx particles areloosely attached to the surface of carbon fibers with clear cracksformed between them (fig. S4). The introduction of Ni2+ in the pre-cursor solution plays a critical role in tuning the nanoarchitectures ofNi-Mo-S/C. At a Ni-to-Mo precursor ratio of 3:1, the surface of car-bon fibers was first covered with a rough and tight nanostructured film,followed by attachment of particles with random shape and size (Fig.2C). The nanostructured film, which shows clear visible striations, isthinner as compared to NiSx/C. When the Ni-to-Mo precursor ratio isreduced to 1:1, a uniform and continuous nanostructured film could begrown on the entire surfaces of carbon fibers (Fig. 2D). However, furtherincrease in the Mo content (Ni/Mo = 1:3) in the growth solution leads tothe formation of large aggregates (Fig. 2E). Moreover, if no Ni2+ wasadded in the growth solution, non-uniform MoS2 aggregates wouldcover the entire surface of carbon fibers.

To gain more detailed insights into the morphological evolution,we further compared the high-magnification SEM images of differentNi-Mo-S/C samples. Figure 3A reveals that the nanostructured film inNi-Mo-S/C (3:1) is composed of numerous nanosheets and large ag-gregated nanoparticles, which tightly stack onto the surface of carbon

Miao et al. Sci. Adv. 2015;1:e1500259 21 August 2015

on March 19, 2020

s.sciencemag.org/

fibers. When the Ni-to-Mo ratio in the precursor solution is reducedto 1:1, a uniform nanostructured film composed of abundant edge-exposing nanosheets with an average thickness of 8 to 10 nm is formed,which develops into a nanoscale 3D network (Fig. 3B). The 3D nano-structure of Ni-Mo-S/C (1:1) is expected to benefit its HER application(22). Further increase in the Mo content in the growth solution (Ni/Mo,1:3) results in the formation of bulky aggregates, which significantlydecreases the density of exposed edge structures (Fig. 3C). If no Ni2+

is introduced into the synthesis, freely grownMoS2 can easily form largeaggregates, which folded together and resembled the morphology ofcrumpled paper balls (Fig. 3D). The average thickness of MoS2 nano-sheets is about 20 to 30 nm, which is about three times thicker thanthe edges in Ni-Mo-S (1:1) (Fig. 3B). The morphologies of differentaggregates in different samples were compared (see fig. S5). It was foundthat the aggregates in MoS2/C, Ni-Mo-S/C (1:3), and Ni-Mo-S/C (1:1)are morphologically similar, and are all evolved from crumpled 2D-layered structures. In contrast, although the aggregates in Ni-Mo-S/C(3:1) and NiSx are similar in appearance, both of the aggregates are formedfrom nanoparticles. Obviously, the introduction of Ni2+ during the syn-thesis plays important roles in regulating the development of Ni-Mo-Snanostructures. The added Ni2+ may suppress the MoS2 crystal growthalong the basal planes, thus lowering the probability of formation ofstack-ups and coalescences among the nanosheets. However, overwhelm-ing Ni2+ will prevent the formation of preferable 3D nanostructureson carbon fibers. Therefore, the optimized Ni-to-Mo precursor ratioin our synthesis is 1:1.

Energy-dispersive x-ray spectroscopy (EDS) measurements wereperformed to study the elemental composition of the samples (fig. S6).The spectrum of MoS2/C affirms that the atomic ratio of Mo/S is closeto 1:2, which is consistent with the stoichiometric ratio of MoS2. How-ever, the measured Ni-to-Mo ratios in Ni-Mo-S/C (3:1), Ni-Mo-S/C (1:1),and Ni-Mo-S/C (1:3) are 1:4.5, 1:8.3, and 1:11.3, respectively, which dif-fer from their corresponding precursor ratios. On the basis of these ob-servations, we believe that the predominant constituents in all Ni-Mo-S/Csamples should be Ni-incorporated MoSx. Point EDS was performedto further confirm the composition of aggregates in Ni-Mo-S/C (3:1). Asshown in fig. S7, it is observed that the ratio of Ni/Mo on the aggregateis ~1.61:1, significantly higher than the nonaggregated area (~0.47:1).It suggests that the aggregates in Ni-Mo-S/C (3:1) are Ni-rich, mostprobably due to the presence of NiSx nanoparticles that were adsorbedonto the electrode surface.

To better elucidate the morphological distinction between Ni-Mo-S/C(1:1) and MoS2/C, TEM images were carried out. Figure 4 (A and B)shows TEM images of MoS2/C and Ni-Mo-S/C (1:1). MoS2 has con-tinuous and corrugated layered structures, where restacked and wrinkledparts can be clearly observed. In contrast, the introduction of Ni2+

during the synthesis leads to the formation of interconnected Ni-Mo-Sflake-like structures. In the high-resolution TEM (HRTEM) images(Fig. 4, C and D), lattices with a spacing of 0.27 nm, which can be as-signed to the (100) plane of 2H-MoS2, are identified in both samples.However, as compared to MoS2/C, the crystalline quality of Ni-Mo-S/Cwas significantly reduced. As indicated by dashed circles in Fig. 4D, nu-merous structure defects can be observed in the surface of nanoflakes(see fig. S8). The insets in Fig. 4 (C and D) show the cross-sectionalviews of stacked MoS2/C and Ni-Mo-S/C (1:1), respectively. We foundthat the (002) plane spacing of Ni-Mo-S/C (1:1), which correlates to theinterlayer distance, is significantly enlarged to 0.92 nm from 0.63 nmobserved in MoS2/C. To assess the distribution of Ni across the basal

Fig. 2. Low-magnification SEM images. (A) Bare carbon fiber cloth. (B)NiSx/C. (C) Ni-Mo-S/C (3:1). (D) Ni-Mo-S/C (1:1). (E) Ni-Mo-S/C (1:3). (F) MoS2/C.

All scale bars, 2 mm.

3 of 14

Page 4: Hierarchical Ni-Mo-S nanosheets on carbon fiber …...(pH ~7), for example, current density of 10 mA/cm2 at a very small overpotential of 200 mV. Furthermore, the Ni-Mo-S/C electrode

R E S EARCH ART I C L E

D

plane of Ni-Mo-S/C (1:1), EDS mapping was performed. As shown infig. S9, both Mo and S are uniformly distributed in the entire detectedregion. Meanwhile, the distribution of incorporated Ni atoms on thestructure is quite homogeneous, although its concentration is slightlyhigher near the defect-rich regions.

Porosity is another important property for electrocatalysts becauseporous structures could provide channels that allow fast access to elec-trolytes and efficient transport of reactants and products. The nitrogenadsorption/desorption analyses were conducted to analyze the poros-ity of the as-synthesized Ni-Mo-S/C (1:1) as shown in fig. S10. Thecurve shows typical type IV isotherm hysteresis according to Interna-tional Union of Pure and Applied Chemistry classification (30), af-firming the mesoporous structure of Ni-Mo-S/C (1:1). As indicatedby the inset of fig. S10, the size of most pores formed on Ni-Mo-S/C(1:1) falls in the range of 2 to 10 nm. This conclusion is consistentwith our observation in TEM (Fig. 4 and fig. S8), in which the pores,due to the defect formation, can be clearly observed on the basal plane of

Miao et al. Sci. Adv. 2015;1:e1500259 21 August 2015

Ni-Mo-S/C (1:1). Apart from the mesopores, numerous macroporesranging from 50 to 100 nm can also be formed among the adjacentnanosheets in the structure. However, these macropores could not beprecisely detected using the Brunauer-Emmett-Teller (BET) method.The porous nature of Ni-Mo-S/C (1:1) should significantly benefitthe electrocatalytic hydrogen evolution process.

To understand the crystal structure of Ni-Mo-S/C samples, XRDpatterns were studied in detail. Because the XRD pattern of Ni-Mo-S/C(1:1) shows a predominant diffraction peak at 26° that can be indexedto the (002) plane of carbon (inset of Fig. 5), the XRD pattern of Ni-Mo-S(1:1) was precisely extracted using Bruker EVA software by subtrac-tion of the signal from carbon fiber cloth (Fig. 5). The broadened dif-fraction peaks imply the nanoscale dimensions of Ni-Mo-S flakes onthe carbon fiber cloth. Compared to the reference XRD pattern (JointCommittee on Powder Diffraction Standards card no. 73-1508), the(002) peak of pristine 2H-MoS2 at 14.39° is absent in the Ni-Mo-S/C(1:1) sample. Instead, two new diffraction peaks at 9.22° and 17.64° can

on March 19, 2020

http://advances.sciencemag.org/

ownloaded from

Fig. 3. High-magnification SEM images. (A) Ni-Mo-S/C (3:1). (B) Ni-Mo-S/C (1:1). (C) Ni-Mo-S/C (1:3). (D) MoS2/C. All scale bars, 500 nm.

4 of 14

Page 5: Hierarchical Ni-Mo-S nanosheets on carbon fiber …...(pH ~7), for example, current density of 10 mA/cm2 at a very small overpotential of 200 mV. Furthermore, the Ni-Mo-S/C electrode

R E S EARCH ART I C L E

http://advances.sD

ownloaded from

be observed, which most probably correspond to the first- and second-order reflections from the (002) plane of layered Ni-Mo-S, respec-tively. Our observation is similar to the MoS2 nanosheets with enlargedinterlayer spacing that was recently reported by Xie et al. (24). Inthis case, the spacing between two adjacent Ni-Mo-S layers can becalculated using Bragg’s law (31):

nl ¼ 2dsinq;

where n is the order of reflection, l is the wavelength of incident x-ray(Cu Ka, 0.154 nm), d is the interlayer spacing, and q is the angle be-tween the incident x-ray and the scattering planes. The calculatedspacing between Ni-Mo-S layers is about 0.94 to 0.95 nm, consistentwith our observations in HRTEM images (0.92 nm in inset of Fig.4D). Moreover, two diffraction peaks centered at 32° and 57° can bewell indexed to the (100) and (110) planes of 2H-MoS2, showing thatthe atomic arrangement of Ni-Mo-S (1:1) should be similar to that ofpristine MoS2. In addition, the asymmetric features of these two peaksreflect the turbostratic stacking characteristic of Ni-Mo-S flakes, whichis commonly observed in other layered compounds (32). The absenceof other high-angle diffraction peaks suggests the structure-disorderingfeature of Ni-Mo-S because of the presence of abundant defect sitesacross its basal planes. It is worth mentioning that diffraction peaks thatcorrelate to nickel sulfide and its nonstoichiometric forms were absentin the XRD pattern of Ni-Mo-S/C (1:1).

Figure 6 shows the Raman spectra of Ni-Mo-S/C (1:1) and MoS2/C.Two distinct peaks at ~380 and ~405 cm−1, corresponding to the in-plane E1

2g and out-of-plane A1g vibrational modes of 2H-MoS2, can beclearly identified in the spectra of both MoS2/C and Ni-Mo-S/C (1:1)

Miao et al. Sci. Adv. 2015;1:e1500259 21 August 2015

ciencema

(20, 33, 34). The slightly red shift and broadening of the A1g peak inNi-Mo-S/C (1:1) can be attributed to reduced numbers of stackedlayers along the c axis (33). Besides, the increased width along withreduced intensity of the E1

2g peak in Ni-Mo-S/C (1:1) reveals the presenceof in-plane defect sites (35), which is consistent with our aforemen-tioned experimental observations (Figs. 3 and 4).

To further confirm the chemical composition and elemental chem-ical states of Ni-Mo-S/C (1:1), XPS measurements were carried out. Asshown in Fig. 7A, the survey spectrum recorded from 0 to 1200 eVreveals the presence of Mo, Ni, S, C, and O elements. From the high-resolution XPS spectrum of the Mo 3d scan (Fig. 7B), two major peaks at228.9 and 232 eV are observed and can be assigned to the MoIV 3d5/2andMoIV 3d3/2 in MoS2, affirming the dominance of Mo(IV) in Ni-Mo-S/C(20). Besides the MoIV 3d5/2 signal, there exists a small peak at 266.2 eVthat resulted from the S 2s orbital. Another doublet at relatively higherbinding energy (Mo 3d5/2, 233.1 eV; Mo 3d3/2, 235.9 eV) can be as-signed to the Mo ions in the +6 oxidation state, which may be due tothe inadequate reduction of MoO4

2− species during the hydrothermalsynthesis. The high-resolution S 2p spectrum clearly shows a doublet withthe S 2p3/2 falling on 161.9 eV (Fig. 7C), revealing the −2 oxidationstate of sulfur in MoS2, which is consistent with previous reports (20, 24).The appearance of a shoulder at higher binding energies in the S 2pregion can serve as further evidence for the presence of Mo with higheroxidation states. As shown in Fig. 7D, the intensity of the Ni spectrumis much lower than that of other elements, due to its lower contentin Ni-Mo-S/C. The calculated Ni-to-Mo ratio is about 1:8.56, whichmatches well with the estimated value from EDS measurements(fig. S6). The Ni spectrum shows a major peak and a satellite peakat 854.3 and 861 eV, respectively, attributed to Ni 2p3/2 orbital,which are close to the previously reported values of sulfide Ni-Moalloys (36).

on March 19, 2020

g.org/

Fig. 4. TEM images. (A and B) Low-magnification TEM images of (A)MoS2/C and (B) Ni-Mo-S/C (1:1). (C and D) HRTEM images of (C) MoS2/C

and (D) Ni-Mo-S/C (1:1). Insets: Corresponding cross-sectional HRTEMimages. Dashed circles in (D) indicate the structure defects of Ni-Mo-S/C (1:1).

Fig. 5. XRD analysis. XRD pattern of Ni-Mo-S/C (1:1) after subtraction ofcarbon fiber cloth signal. Inset: Original XRD pattern of Ni-Mo-S/C (1:1). a.u.,

arbitrary unit.

5 of 14

Page 6: Hierarchical Ni-Mo-S nanosheets on carbon fiber …...(pH ~7), for example, current density of 10 mA/cm2 at a very small overpotential of 200 mV. Furthermore, the Ni-Mo-S/C electrode

R E S EARCH ART I C L E

http://advances.sciencemag.o

Dow

nloaded from

To investigate the role of incorporated Ni on the catalytic activityof Ni-Mo-S, electrochemical measurements were designed and carriedout. It is important to mention that all as-fabricated materials weredirectly used as working electrodes for HER without annealing or cal-cination. Neutral phosphate buffer solution, which could minimize theadverse environmental impacts in practical applications (37), was chosenas the primary electrolyte in this work. All electrochemical tests werecarried out using the same setup, with consistent cell configuration,electrode positions, and electrolyte volume. The HER activities of dif-ferent electrodes were investigated by comparing the polarization curvesobtained with linear sweep voltammetry (LSV). In each measurement,the voltage applied on working electrode was linearly swept from 0.2 to−0.5 V versus reversible hydrogen electrode (RHE) at a scan rate of5 mV/s. As shown in Fig. 8A, all Ni-Mo-S/C electrodes have typicalonset overpotentials of 130 to 150 mV, which are apparently smallerthan that of MoS2/C (~200 mV), whereas the bare carbon fiber clothelectrode shows negligible HER catalytic activity. Among all studiedsamples, Ni-Mo-S/C (1:1) shows the best HER catalytic activity witha considerably smaller h10 (overpotential at 10 mA/cm2) of 200 mV.At an applied potential of −0.35 V versus RHE, Ni-Mo-S/C (1:1) canproduce a cathodic current density as large as 36.5 mA/cm2, which istwo times that produced by MoS2/C. During the test, a tremendousamount of H2 bubbles was observed on the surface of Ni-Mo-S/C(1:1) (movie S1). To ensure a fair comparison, we have normalizedthe performance of different electrodes (except for the blank carbon fibercloth) by their corresponding actual catalyst loadings (Table 1) as shownin fig. S11 and found that the Ni-Mo-S/C (1:1) still shows the bestperformance among the samples.

To better understand the different catalytic behaviors of these elec-trodes, their corresponding Tafel plots were obtained from the slow-scan rate (1 mV/s) polarization curves. As shown in Fig. 8B, various Tafelslopes, ranging from 85.3 to 107.2 mV/dec, were found on different

Miao et al. Sci. Adv. 2015;1:e1500259 21 August 2015

rg

electrodes. These values are apparently larger than those recently re-ported for MoS2-based electrocatalysts in acidic electrolytes (39 to65 mV/dec) (18, 23, 24, 38) but are comparable with those values re-ported using neutral electrolytes (37, 39, 40). The increase in Tafelslopes is due to the much lower proton concentration (~1 × 10−7

M, standard conditions) in neutral electrolytes than in acidic solutions,in which electrochemical hydrogen adsorption (Volmer reaction) andelectrochemical desorption (Heyrovsky reaction) processes proceed atcomparable rates (39). Although it is not possible to precisely deter-mine the rate-determining step and exact reaction mechanism in thiscase, Tafel slopes of different electrodes are still important indicatorsfor comparative analyses. Among all tested electrodes, Ni-Mo-S/C (1:1)exhibits the smallest Tafel slope (~85.3 mV/dec), which correlates wellwith its remarkable HER catalytic behavior originating from the abun-dant exposed edge sites and excellent material uniformity. In addition,the defect sites found in the basal plane of Ni-Mo-S/C (1:1) couldalso provide substantial extra active sites that can further enhanceits HER performance. The slightly steeper Tafel slopes observed onMoS2/C (94.4 mV/dec) and Ni-Mo-S/C (1:3) (88.4 mV/dec) may pos-sibly be due to the formation of bulky aggregates, which, in turn, hindersthe interactions between protons and effective active sites to a certainextent. As compared to MoS2/C and other Ni-Mo-S/C samples, mostprobably resulting from their relatively poorer intrinsic catalytic activ-ities, the Ni-Mo-S/C (3:1) and NiSx/C showed much higher Tafelslopes, which are ~103.7 and ~107.2 mV/dec, respectively. Table S1 listedthe exchange current densities (j0) of different samples calculated usingthe Tafel equation. Ni-Mo-S/C (1:1) showed the largest j0 among thetested electrodes, which is 4.89 × 10−2 mA/cm2. The j0 values of othersamples follow the order of Ni-Mo-S/C (3:1) > NiSx > Ni-Mo-S/C(1:3) > MoS2.

For comparison purposes, electrochemical tests were also carriedout in acidic electrolyte, as illustrated in fig. S12, and exhibit a similargeneral trend to that observed in neutral electrolyte (Fig. 8A), in which

on March 19, 2020

/

Fig. 7. XPS. (A) Survey XPS spectrum of Ni-Mo-S/C (1:1). (B to D) High-resolution scans of (B) Mo 3d, (C) S 2p, and (D) Ni 2p.

Fig. 6. Raman spectroscopy. Raman spectra of Ni-Mo-S/C (1:1) andMoS2/C. Inset: Schematic illustration of E2g

1 and A1g vibrational modes in

layered 2H-MoS2.

6 of 14

Page 7: Hierarchical Ni-Mo-S nanosheets on carbon fiber …...(pH ~7), for example, current density of 10 mA/cm2 at a very small overpotential of 200 mV. Furthermore, the Ni-Mo-S/C electrode

R E S EARCH ART I C L E

on March 19, 2020

http://advances.sciencemag.org/

Dow

nloaded from

Ni-Mo-S/C (1:1) still shows the best catalytic performance character-ized by a low onset overpotential (~−0.14 V) and a reasonably smallTafel slope (~48 mV/dec). The Tafel slope values obtained in acidicelectrolyte are closer to those previously reported (17, 19, 41, 42),indicating that electrochemical desorption (Heyrovsky reaction) isthe rate-determining step for HER catalyzed by Ni-Mo-S/C (1:1)through the Volmer-Heyrovsky mechanism.

Moreover, electrochemical impedance spectroscopy (EIS) analyseswere performed to study the electrode kinetics of these samples underthe HER mode (39, 43). All measurements were performed at −0.3 Vversus RHE from 100 K to 0.1 Hz with alternating current (AC) am-

Miao et al. Sci. Adv. 2015;1:e1500259 21 August 2015

plitude of 15 mV in neutral buffer (Fig. 8C). A Randle circuit (inset inFig. 8C), which is commonly used in analyzing charge transport onsingle electrodes, was used to fit our experimental results (44). All elec-trodes show small and similar series resistance (Rs), from 1.9 to 2.2 ohms,mainly originated from uncompensated solution resistance, suggestingthat our experimental setup was consistent. The slight variation of Rsin different electrodes can be partially attributed to the differences inelectrode resistance. The kinetics of electrocatalysis on different electrodescan be reflected from their charge transfer resistance (Rct), and a lowerRct value corresponds to a faster reaction rate (39). The Rct values ofMoS2/C, NiSx/C, and all Ni-Mo-S/C electrodes are summarized in

Fig. 8. Electrochemical performances. (A) Polarization curves of carbon fiber cloth, NiSx/C, MoS2/C, and different Ni-Mo-S/C in neutral electrolyte. GSA,geometric surface area. (B) Corresponding Tafel plots obtained using slow-scan rate polarization curves. (C) Electrochemical impedance spectra ofdifferent electrodes at −0.3 V versus RHE. Inset: Equivalent circuit used for data analyses. CPE, constant-phase element; ZW, Warburg impedance. (D)Plot of charge transport resistance of different samples.

7 of 14

Page 8: Hierarchical Ni-Mo-S nanosheets on carbon fiber …...(pH ~7), for example, current density of 10 mA/cm2 at a very small overpotential of 200 mV. Furthermore, the Ni-Mo-S/C electrode

R E S EARCH ART I C L E

on March 19, 2020

http://advances.sciencemag.org/

Dow

nloaded from

Fig. 8D, and show a similar tendency to that previously observed inHER tests (Fig. 8A), with Ni-Mo-S/C (1:1) exhibiting the lowest Rctof 39.48 ohms.

The onset potentials of different samples were measured using thetangent method. As shown in Fig. 9A, the determined onset potentialsof Ni-Mo-S/C (1:1) and MoS2/C are −0.132 and −0.193 V versus RHE,respectively, in neutral electrolyte. Compared to MoS2/C, the absoluteonset potential value of Ni-Mo-S/C (1:1) is 61 mV smaller, arising fromits significantly improved HER catalytic activity. However, it is not al-ways fair to judge the catalytic behavior on different electrodes solely onthe basis of their differences in onset potential and Tafel slope, especial-ly for the practical gas-evolving reaction. Therefore, we developed asimple strategy that allows us to intuitively compare the onset poten-tials and gas-evolving profiles of different electrodes at the same time.As shown in fig. S13, a special electrode was designed to visualize theHER processes simultaneously taking place on different samples.Single bundles of MoS2/C and Ni-Mo-S/C (1:1) with the same length(2 cm) and similar appearance were first extracted from theircorresponding carbon fiber cloth electrodes and then firmly attachedto the same home-made titanium (Ti) connector. Within the Ti con-nector, the distance between the ends of two samples was fixed at 3 mmfor ease of comparison, whereas the angle between the two bundles wasset at 30° for better observation of the gas-evolving events along theirentire length. The HER study was then carried out using the parallellinked single-bundle electrodes as the working electrode, in which thesynchronized potential bias can be simultaneously applied to MoS2/Cand Ni-Mo-S/C (1:1). In a typical comparative test, the working elec-trode was linearly swept in a potential window of 0 to −0.8 V versusRHE at a scan rate of 10 mV/s using LSV technique, and the wholeprocess was recorded using a digital camera (movie S2). The first ob-servable gas bubbles were found on Ni-Mo-S/C (1:1) and MoS2/C at−0.13 and −0.2 V versus RHE, respectively (fig. S14, A and B), match-ing well with the results obtained from electrochemical tests (Fig. 9A).During the test, H2 bubbles on Ni-Mo-S/C (1:1) were generated fasterand more continuously along the entire bundle. At larger applied biases(−0.8 to −0.4 V versus RHE), it was found that the color of MoS2/Cbundle turned pale white because of the adhesion of numerous freshlyevolved gas bubbles (fig. S14C), whereas the appearance of Ni-Mo-S/Cbundle (1:1) remained consistent throughout the test with a much morefluent gas-evolving profile. The macroscopic differences between thetwo electrodes observed by naked eyes originate from their microscop-ic distinctions. The aggregates formed in MoS2/C, as discussed above,may evidently decrease the exposed active sites, leading to the attenu-ated catalytic activity. Meanwhile, these bulky structures will also hinder

Miao et al. Sci. Adv. 2015;1:e1500259 21 August 2015

the effective proton adsorption and H2 desorption processes becauseof the presence of trapped bubbles within them. Therefore, the micro-scopic gas bubbles have to gain their volume before they can be success-fully released from these trap sites. Furthermore, these accumulatedbubbles on the surface of MoS2/C will make the electrode less aeropho-bic and more hydrophobic in appearance, as observed during the test.In brief, a significant decrease in overall efficiency is expected whenall these unfavorable factors are combined. In contrast, the signifi-cantly improved catalytic performance of Ni-Mo-S/C (1:1) resultsfrom the exposure of abundant uniform edge structures along withadditional active sites arising from the defect sites in the basal plane,whereas the more fluent gas-evolving profile observed on it may benefitfrom its sharper and thinner edges, which will ease the gas-releasing(desorption) process (22).

To affirm that the catalytic enhancement mainly originates fromthe more exposed edge sites and defects formed along the basal planes,it is necessary to normalize the electrode HER activities of MoS2/Cand Ni-Mo-S/C (1:1) to their total electrochemically active surfacearea (EASA). To obtain the total EASA, cyclic voltammetry at differentscan rates was performed to calculate the electrochemical double layercapacitance according to the method discussed by Trasatti and Petrii(45). More recently, this method was also used by Benck et al. to cal-culate the EASA of MoS2-based HER catalysts (46). In our case, thecapacitive current of MoS2/C and Ni-Mo-S/C (1:1) was measured in0.5 M sodium phosphate buffer solution at different scan rates (1 to10, 15, 20, and 25 mV/s) in a potential range from −0.02 to 0.08 Vversus saturated calomel electrode (SCE), because there are no obviouselectrochemical features corresponding to the faradaic current withinthis range. The nonfaradaic capacitive current, which is proportional toboth the scan rate and EASA, can be calculated from these two sets ofCV curves as shown in fig. S15 (A and B). The capacitive current of thetwo electrodes measured at 0.03 V versus SCE was plotted as a functionof scan rate as shown in fig. S15C. We noticed that the dependence ofthe current on the scan rate in this region is linear for both electrodes,which is consistent with the capacitive charging behavior. From theslope of the linear curve, the unit electrode capacitance was calculatedto be 82.78 and 27.74 mF/cm2 for MoS2/C and Ni-Mo-S/C (1:1), re-spectively. In most previously reported studies, the calculated capaci-tance of single-layer/flat MoS2 is ~60 mF/cm

2 (46). With this number,the total EASAs of MoS2/C and Ni-Mo-S/C (1:1) are 1379.67 and462.33 cm2, respectively, on the electrodes with a geometric area of~1 cm2. This result suggests that the total EASA of MoS2/C is threetimes that of Ni-Mo-S/C (1:1). Figure S15D shows the polarizationcurves of MoS2/C and Ni-Mo-S/C (1:1) normalized by their corre-

Table 1. Calculation of actual catalyst loadings in MoS2/C, NiSx/C, and different Ni-Mo-S/C samples.

Sample

Calculated mass loading (mg/cm2)

SD

Test 1 Test 2 Test 3 Average

MoS2

1.64 1.56 1.63 1.61 0.04

Ni-Mo-S/C (Ni/Mo, 1:3)

0.96 0.84 0.87 0.89 0.06

Ni-Mo-S/C (Ni/Mo, 1:1)

0.52 0.54 0.5 0.52 0.02

Ni-Mo-S/C (Ni/Mo, 3:1)

0.35 0.31 0.3 0.32 0.03

NiSx

0.44 0.37 0.42 0.41 0.04

8 of 14

Page 9: Hierarchical Ni-Mo-S nanosheets on carbon fiber …...(pH ~7), for example, current density of 10 mA/cm2 at a very small overpotential of 200 mV. Furthermore, the Ni-Mo-S/C electrode

R E S EARCH ART I C L E

on March 19, 2020

http://advances.sciencemag.org/

Dow

nloaded from

sponding total EASA, confirming that the performance of Ni-Mo-S/C(1:1) is significantly better than its counterpart. These results suggestthat there would be many more active sites on Ni-Mo-S/C (1:1) ascompared to MoS2/C on the electrodes with the same EASA. Besidesthe EASA measurements, the edge-rich feature of Ni-Mo-S/C (1:1)was also confirmed using the irreversible electrochemical oxidationmethod proposed by Bonde et al. (47). (Refer to fig. S16 and noteS2 in the Supplementary Materials.)

Stability is another critical factor for designing gas-evolving electrodesthat are eligible for long-term operation. In our work, the durability ofNi-Mo-S/C (1:1) and MoS2/C electrodes (1 × 1 cm) was assessed usingcontrolled potential electrolysis in neutral electrolyte. The two electrodeswere operated at their corresponding h20 (overpotential at current

Miao et al. Sci. Adv. 2015;1:e1500259 21 August 2015

density of 20 mA/cm2) for 30,000 s. Continuous gas bubbling was ob-served on both Ni-Mo-S/C (1:1) and MoS2/C throughout the test. Asshown in Fig. 9B, Ni-Mo-S/C (1:1) can retain about 97.5% of itsinitial current density after the test, which is apparently better thanthe 85% current density retention of MoS2/C. For comparison, wealso performed the stability of commercial Pt plate electrode (1 ×1 cm) in neutral electrolyte. It was found that only 15% of the initialcurrent density was retained after 1-hour controlled potential elec-trolysis (fig. S17). The rapid deactivation of Pt in neutral buffer elec-trolyte might be due to the poisoning effect of phosphoric acid anions(see fig. S18 and note S3). The excellent stability of Ni-Mo-S/C(1:1) was further affirmed using LSV technique after controlledpotential electrolysis. As shown in Fig. 9C, the polarization curve

Fig. 9. Electrochemical hydrogen evolution tests. (A) Onset comparison between MoS2/C and Ni-Mo-S/C (1:1). (B) Stability tests of Ni-Mo-S/C (1:1) andMoS /C in neutral electrolyte. (C) LSV curves of Ni-Mo-S/C (1:1) before and after stability tests. (D) Comparison of the detected amount of evolved H and

2 2

the theoretical value in Faraday efficiency measurement.

9 of 14

Page 10: Hierarchical Ni-Mo-S nanosheets on carbon fiber …...(pH ~7), for example, current density of 10 mA/cm2 at a very small overpotential of 200 mV. Furthermore, the Ni-Mo-S/C electrode

R E S EARCH ART I C L E

on March 19, 2020

http://advances.sciencemag.org/

Dow

nloaded from

obtained after a stability test was almost identical with the initialscan, without obvious shift in onset potential and curve shapes.In addition, as shown in fig. S19, no obvious changes in the oxida-tion states of Mo, S, and Ni were detected in the XPS measure-ments after the stability test, affirming the excellent stability ofNi-Mo-S/C (1:1).

The Faraday efficiency of Ni-Mo-S/C (1:1) was measured using amultipurpose compact glass photo-/electrocatalysis reactor, integratedwith a single-chamber bulk electrolysis cell. The schematic and photo-graphs of the experimental setup are shown in figs. S20 and S21, respec-tively. Controlled potential electrolysis was performed on Ni-Mo-S/C(1:1) in neutral electrolyte, with a moderate applied bias of −0.2 V ver-sus RHE for 4000 s. The detailed Faraday efficiency calculation stepsare described in the Supplementary Materials (note S4). Figure 9D showsthat the actual detected H2 amount was correlated with the theoreticalvalue calculated on the basis of the charge transferred, resulting in ahigh Faraday efficiency of 98.2%. From the gas chromatography(GC) spectra, H2 and O2 (evolved on graphite foil) were the onlytwo gaseous products detected throughout the test, which affirmsthe capability of Ni-Mo-S/C (1:1) as an HER electrode in producinghigh-purity H2 gas.

To further demonstrate the potential of Ni-Mo-S/C in practical ap-plications, a simple prototype of laboratory-scale hydrogen generatorwas designed. Figures S22 and S23 show the schematic and actualassembly of the laboratory-scale Ni-Mo-S/C hydrogen generator, re-spectively. To prepare this reactor, a piece of Ni-Mo-S/C (1:1) (2 × 4 cm)was first rolled up into a column shape and then firmly fixed on ahomemade Ti holder, which also worked as the lead and electrode con-nector (fig. S23B). A 5-ml sterile round-bottom tube was used as theelectrode housing and the primary gas collector (fig. S23C). The Tiholder together with the clamped Ni-Mo-S/C was loaded into the tubeand fixed using Scotch tape. To integrate the reactor with the gas-evacuation system or secondary gas collector, a round hole with a di-ameter of 1 mm was drilled on its top. Here, a 5-ml syringe, functioningas a simplified gas-evacuation system, was tightly sealed onto the re-actor (fig. S23D). For the practical operation, the H2 generator shouldbe vertically inserted into the electrolyte. The air remaining in the tubewas then evacuated using the syringe, whereas the electrolyte was suckedupward to fill the entire tube. When the system is operated at a mod-erate bias of −0.5 V versus RHE, the 5-ml tube can be filled with H2 in~150 s (movie S3).

On the basis of experimental findings, we propose a hypotheticalgrowth mechanism to explicate the development of a high-performanceNi-Mo-S HER electrocatalyst. Without the incorporation of Ni2+, thegrowth of layered MoS2 crystals proceeds in a fast and uncontrollablefashion. The rapidly grown MoS2 sheets tend to aggregate in solution,thus leading to the formation of bulky coalescences. We noticed thatthe morphological evolution of Ni-Mo-S/C strongly depended on theNi-to-Mo precursor ratio. It is believed that the introduced Ni2+ couldbond to the free sulfur during the growth or nucleation of MoS2nanocrystals, subsequently leading to the formation of substitution-al and Schottky defects that instantly interrupt the regular atomicarrangement in MoS2. The formation of Ni-S bonds was confirmed byx-ray absorption measurement (fig. S24), in which the Ni-S path(~2.23 Å) has a smaller atomic radius than the Mo-S path (~2.39Å). Note that this difference can be attributed to the ionic radialdiscrepancy between Mo4+ (70 pm) and Ni2+ (83 pm), which isconsistent with the observation from XPS (Fig. 7). As a classic hy-

Miao et al. Sci. Adv. 2015;1:e1500259 21 August 2015

drodesulfurization catalyst model, the Co-Mo-S has been thor-oughly studied over the past few decades (13, 48–50). On the basisof the density function theory, Byskov et al. found that the substitu-tional Co atoms in Co-Mo-S resulted in significantly lower sulfurbinding energy as well as shorter bond length near the vacancies(51). In our case, Ni atoms will play similar roles in tuning the Ni-Mo-S framework, that is, the change in bond length and bindingenergy near the incorporated Ni atoms may lead to an extensive dis-tortion across the basal plane, thus hindering the rapid growth of acontinuous MoS2 film, while creating extra edges with more exposedactive sites.

Figure S25 shows the schematic illustration of a MoS2 nanocrystalsurrounded by Ni2+ executing Brownian motion, where Ni could be in-corporated into the existing MoS2 framework upon its successful col-lision with the free sulfur ends. The distribution of Ni2+ around the MoS2nanocrystals is critical in determining the eventually evolved morphol-ogy of Ni-Mo-S structure on carbon fibers, because the frequency ofsuccessful bond formation largely depends on their concentrations,according to the collision theory in chemical kinetics (52). When MoS2nanocrystals are surrounded by excessive Ni2+ ions, vigorous collisionsmay take place, thus accelerating the formation of substitutional andvacancy defects, which, in turn, constrain the continuous growth of thelayered structure of Ni-Mo-S. On the contrary, the inadequate Ni2+

will only introduce random allocated defect sites with lower density,which cannot significantly alter the growth of the layered structuresobserved in MoS2. These assumptions explain the morphological evo-lution that we observed in different Ni-Mo-S/C samples well (Figs. 2and 3). Optimum Ni-to-Mo precursor ratio will lead to the formationof moderate defect sites along the basal plane, which simultaneouslyintroduce additional active sites and constrain the growth rate of Ni-Mo-S layers. On the basis of these assumptions, we constructed themodels of MoS2 and Ni-Mo-S and performed molecular mechanicalsimulation (molecular mechanics + force field) using HyperChem(10). Figure S26 (A and B) shows the molecular models after geometricoptimization for MoS2 and Ni-Mo-S, respectively. Compared to thepristine MoS2, the random defects formed in Ni-Mo-S can cause a sig-nificant geometric distortion on the basal plane because of the changein bond length and the presence of vacancies. The increase in molec-ular complexity along with the distortions across the entire basal planemay further weaken the interaction between the adjacent Ni-Mo-Slayers, thus leading to an enlargement of the interlayer spacing (fig.S26D), which is consistent with the observation in Fig. 4D. On thebasis of these findings, we know that optimizing the Ni-to-Mo precur-sor ratio is important in tailoring the nanostructure of Ni-Mo-S atatomic scale. Apart from the structural effects, the incorporated Niatoms may also function as promoters that further enhance the intrin-sic catalytic activities, especially in neutral electrolyte, similar to thatproposed by Merki et al. (39). Besides, Lin et al. reported the atomicmechanism of the semiconducting (2H)–to–metallic (1T) phasetransition in single-layered MoS2 (10). In our case, the bond-lengthshortening introduced by the Ni atom may also initiate the glidingof a sulfur plane near the defect sites, further resulting in the localizedphase transition that partially converts 2H MoS2 to metastable 1TMoS2. As previously discussed by Lukowski et al., the metallic 1TMoS2 may exhibit enhanced electrocatalytic activity along with the im-proved stability (53). In addition, the incorporation of Ni atoms mayalter the binding energy of pristine MoS2, further improving the reac-tion kinetics.

10 of 14

Page 11: Hierarchical Ni-Mo-S nanosheets on carbon fiber …...(pH ~7), for example, current density of 10 mA/cm2 at a very small overpotential of 200 mV. Furthermore, the Ni-Mo-S/C electrode

R E S EARCH ART I C L E

on March 19, 2020

http://advances.sciencemag.org/

Dow

nloaded from

DISCUSSION

In conclusion, we have developed a facile synthetic strategy to directlygrow nanostructured Ni-Mo-S on carbon fiber cloth as a high-efficiencyfunctional electrode for hydrogen generation in neutral electrolyte.The incorporated Ni atoms play vital roles in constructing Ni-Mo-Snanostructures through forming substantial desirable defect sites aswell as regulating the growth rate of layered MoS2. The optimizedNi-Mo-S/C exhibits excellent HER catalytic performance, character-ized by its low onset potential of −0.132 V versus RHE and a smallTafel slope of 85.3 mV/dec in the neutral electrolyte. The reducedstacking and aggregations in Ni-Mo-S/C effectively prevent the accu-mulation of instantly evolved H2 bubbles during electrolysis, henceimproving the performance and stability of Ni-Mo-S/C in long-termhydrogen production. The remarkable HER performance in acidic andneutral media of Ni-Mo-S/C is comparable with other state-of-the-artHER electrocatalysts (see tables S2 and S3). To further demonstrateits potential application in practical appliances, a laboratory-scale hy-drogen generator with a rolled Ni-Mo-S/C electrode was designed tocarry out the potential controlled electrolysis. Owing to the ultimateflexibility and excellent mechanical strength offered by the carbonfiber cloth, the functional Ni-Mo-S/C electrode developed in this workshould be easily integrated into existing H2 generators with appropriatemodification to meet their specified standards and requirements. To car-ry out more practical application, we also performed HER tests usingNi-Mo-S/C (1:1) in real seawater (see note S5 and figs. S27 and S28 inthe Supplementary Materials). We believe that better electrodeperformance could be achieved through engineering the carbon fibercloth, for example by widening the interstrand spacing and reducingthe packing density of the carbon fiber.

MATERIALS AND METHODS

MaterialsWoven carbon fiber cloth with evenly sized pores was purchased fromGasHub Technology. Sodium molybdate dihydrate (Na2MoO4 · 2H2O),nickel sulfate hexahydrate (NiSO4 · 6H2O), L-Cys (HSCH2CHNH2COOH),sulfuric acid (H2SO4, ≥98%), sodium phosphate monobasic (NaH2PO4),and sodium phosphate dibasic (Na2HPO4) were purchased fromSigma-Aldrich. Hydrogen peroxide (H2O2) was purchased from AlfaAesar, and acetone [(CH3)2CO] was purchased from Acros Organics. Allchemical reagents were directly used without further purification.Milli-Q water (resistivity over 18 megohm·cm) from aMillipore Q waterpurification system was used in all experiments.

Miao et al. Sci. Adv. 2015;1:e1500259 21 August 2015

Syntheses of MoS2 and Ni-Mo-S on carbon fiber clothCarbon fiber cloth substrates were consecutively washed with acetone,H2SO4 (1 M), and deionized water under sonication for 2 hours ineach solution to thoroughly remove organic residues and other impu-rities. The substrates were then soaked in “piranha” solution, that is, amixture of concentrated sulfuric acid and 30% hydrogen peroxidewith a volume ratio of 3:1, overnight to further improve their hydro-philicity. The clean substrates were kept in Milli-Q water to avoid re-gaining hydrophobicity.

The MoS2, NiSx, and Ni-Mo-S were grown on the cleaned carbonfiber cloth using a biomolecule-assisted hydrothermal synthetic route.Briefly, sodium molybdate dihydrate (Na2MoO4 · 2H2O) and nickelsulfate hexahydrate (NiSO4 · 6H2O) were used as Mo and Ni sources,respectively, whereas L-Cys (HSCH2CHNH2COOH) was used as an Ssource and reducing agent. To study the effect of the Ni-to-Mo pre-cursor ratio on the morphology and performance of the as-preparedfunctional electrodes, Na2MoO4 · 2H2O (6.05 mg/ml, 25 mM), NiSO4 ·6H2O (6.57 mg/ml, 25 mM), and L-Cys (16 mg/ml, 132 mM) aqueoussolutions were prepared and mixed according to Table 2, whereas thesum of molar amounts of Ni and Mo was kept constant in the precur-sor solutions for all samples.

The precursor solutions have to be rapidly mixed and subjected tovigorous stirring to prevent the formation of precipitates. Then, thesolution was transferred to a 50-ml Teflon-lined stainless autoclave.Thereafter, a piece of precleaned carbon fiber cloth was verticallyaligned in the growth solution. The autoclave was heated at 200°C for24 hours. After naturally cooling down to room temperature, thesample was thoroughly washed with Milli-Q water and dried usingcompressed airflow at room temperature. The weight of carbon fibercloth electrodes was precisely recorded before and after reaction usinganOhaus analytical balance to calculate the actual catalyst-loading amount.The calculation of actual loadings for MoS2/C, NiSx/C, and differentNi-Mo-S/C samples can be found in Table 1.

Sample characterizationThe morphology and elemental composition of samples were exam-ined using field emission SEM (JEOL JSM-6700F), TEM (JEOL JEM-3010),and EDS (JEOL JED-2300 Analysis Station). BET measurements wereperformed using the Autosorb-1 system (Quantachrome Instruments).XRDmeasurements were conducted on the Bruker AXS D8 Advance andD2 Phaser to study the crystal structures and phase compositions. XPSmeasurements were carried out on an ESCALAB 250 photoelectronspectrometer (Thermo Fisher Scientific) at 2.4 × 10−10 mbar using amonochromatic Al Ka x-ray beam (1486.60 eV). All binding energieswere referenced to the C 1s peak (284.60 eV) arising from the adven-titious hydrocarbons. Raman spectra of the samples were measuredusing a research laser Raman microscope (Renishaw RM1000), witha 514.5-nm excitation laser. The Raman band of Si at 520 cm−1 wasused as the reference to calibrate the spectrometer. A series of extendedx-ray absorption fine structure (EXAFS) measurements were made usingsynchrotron radiation. Measurements were made at the Ni K edge(8333 eV) and the Mo K edge (20,000 eV) with the sample held at roomtemperature. The 01C beamline in the National Synchrotron RadiationResearch Center in Taiwan was designed for such experiments. Thebackscattering amplitude and phase shift functions for specific atompairs were calculated ab initio using the FEFF8 code. X-ray absorp-tion data were analyzed following standard procedures, including pre-and post-edge background subtraction, normalization with respect

Table 2. Preparation of MoS2, NiSx, and Ni-Mo-S on carbon fiber cloth.

Precursor sampleN

a2MoO4·2H2O(25 mM) (ml)

NiSO4·6H2O(25 mM) (ml)

L-Cys(132 mM) (ml)

MoS2

20 0 20

Ni-Mo-S (Ni/Mo, 1:3)

15 5 20

Ni-Mo-S (Ni/Mo, 1:1)

10 10 20

Ni-Mo-S (Ni/Mo, 3:1)

5 15 20

NiSx

0 20 20

11 of 14

Page 12: Hierarchical Ni-Mo-S nanosheets on carbon fiber …...(pH ~7), for example, current density of 10 mA/cm2 at a very small overpotential of 200 mV. Furthermore, the Ni-Mo-S/C electrode

R E S EARCH ART I C L E

on March 19, 2020

http://advances.sciencemag.org/

Dow

nloaded from

to edge height, Fourier transformation, and nonlinear least-squarescurve fitting. The normalized k3-weighted EXAFS spectra, k3x(k),were Fourier-transformed in the k range from 3.1 to 12.7 Å−1 to revealthe contribution of each bond pair on the Fourier transform peak. Theexperimental Fourier-filtered spectra were obtained by performing aninverse Fourier transformation with a Hanning window function withr between 1.1 and 2.4 Å. The S0

2 (amplitude reduction factor) valuesof Mo and Ni atoms were fixed at 0.9 and 0.91 to determine the struc-tural parameters of each bond pair.

Electrochemical measurementsElectrochemical measurements were carried out on an electrochemicalworkstation (CHI 760E) with a three-electrode setup, consisting of anas-prepared electrode as the working electrode, a graphite foil (2 × 3 cm)as the counter electrode, and an SCE (in saturated KCl) as the refer-ence electrode. To assess the catalytic activity in neutral medium,electrochemical tests were primarily carried out in 0.5 M sodiumphosphate buffer solution prepared by mixing 0.5 M NaH2PO4 and0.5 M Na2HPO4 aqueous solutions with an appropriate ratio. The HERperformance was also studied in 0.5 M H2SO4 electrolyte. The electro-lytes were deaerated by purging nitrogen for 30 min before the electro-chemical measurement. The pH of different electrolytes was carefullymeasured using a calibrated pH meter (Mettler Toledo SevenCompact),resulting in 0.18 and 6.94 for 0.5 M H2SO4 and 0.5 M sodium phos-phate buffer solution, respectively.

The reference electrode calibration was performed using a standardPt plate electrode as the working electrode. Cyclic voltammetry scanswere performed at a scan rate of 1 mV/s, and the average of the twopotentials at which the current crossed zero was taken as the experi-mentally determined 0 V versus RHE (or VHOR/HER). On the basis ofthis value, the electrode potential of the SCE that was used in thiswork was obtained by using the following equation:

V 0SCEðVÞ ¼ VHOR=HER − 0:059 V� pH:

The theoretical electrode potential of SCE (in saturated KCl) wasreported to be 0.241 to 0.244 V versus RHE. But the values that we de-termined experimentally, that is, 0.238 V versus RHE in 0.5 M H2SO4

and 0.234 V versus RHE in 0.5 M sodium phosphate buffer solution,slightly differ from the theoretical values.

The catalytic behaviors of different materials were studied andcompared using LSV at a scan rate of 5 mV/s, whereas the stabilitywas studied using amperometric technique. The measured voltagevalues were converted to RHE scale by applying the following calculation:

VRHE ¼ Vmeasured þ V 0SCE þ 0:059� pH;

where VRHE is the converted potential value versus RHE, Vmeasured isthe voltage reading from potentiostat, and V0

SCE is the experimentallydetermined electrode potential of SCE.

To study the electrode kinetics, EIS was performed at −0.3 V versusRHE in a range from 100 K to 0.1 Hz with an AC amplitude of 15 mV,using an electrochemical workstation (Zive SP2, WonATech). Thedata were analyzed and fitted in equivalent circuits using ZViewsoftware, and Nyquist plots were used to study the charge transportproperties.

Here, AC impedance technique was used to obtain the uncom-pensated resistance (Ru) of the working electrodes before the LSV

Miao et al. Sci. Adv. 2015;1:e1500259 21 August 2015

measurements. All polarization curves were iR-corrected with a com-pensation level of 90%.

Faraday efficiency measurementFaraday efficiency measurement was performed inside a multipurposecompact glass photo-/electrocatalysis reactor, integrated with a single-chamber electrolysis cell. The sample line of the reactor was linked toan online GC (Agilent 490), which is equipped with a molecular sievecolumn and a thermal conductivity detector, allowing the real-timedetection of gas products. Inside the electrolysis cell, the Ni-Mo-S/C,graphite foil, and SCE were used as working, counter, and referenceelectrodes, respectively, whereas deaerated 0.5 M sodium phosphatebuffer was used as the electrolyte. Before the measurement, the systemwas thoroughly evacuated using a high-efficiency vacuum pump andthen charged with ultrapure argon (Ar, 99.9995%) gas to 1 atm. Con-stant potential was applied to the working electrode using an electro-chemical workstation (CHI 760E). The compositions and concentrationsof the gas products were analyzed with GC through automaticsampling every 4 min.

SUPPLEMENTARY MATERIALSSupplementary material for this article is available at http://advances.sciencemag.org/cgi/content/full/1/7/e1500259/DC1Fig. S1. SEM image of carbon fiber cloth.Fig. S2. SEM image of MoS2/C synthesized with reduced precursor concentration.Fig. S3. SEM images of MoS2/C synthesized after different durations of hydrothermal treatment.Fig. S4. SEM images of NiSx/C.Fig. S5. SEM images of aggregates.Fig. S6. EDS spectra of different samples.Fig. S7. Point EDS spectra of Ni-Mo-S/C (3:1).Fig. S8. HRTEM images of Ni-Mo-S/C (1:1).Fig. S9. Elemental mapping of Ni-Mo-S/C (1:1).Fig. S10. BET measurements of Ni-Mo-S/C (1:1).Fig. S11. Normalized current density.Fig. S12. HER tests in 0.5 M H2SO4.Fig. S13. Electrode for intuitive comparison test.Fig. S14. Intuitive comparison test for HER.Fig. S15. EASA measurements for MoS2/C and Ni-Mo-S/C (1:1).Fig. S16. Irreversible electrochemical oxidation cyclic voltammetry curves.Fig. S17. HER stability of Pt in neutral electrolyte.Fig. S18. XPS spectra of fresh and deactivated Pt plate electrodes.Fig. S19. High-resolution XPS spectra of Ni-Mo-S/C (1:1) after stability test.Fig. S20. Experimental setup for Faraday efficiency test.Fig. S21. Actual experimental setup for Faraday efficiency measurement.Fig. S22. Laboratory-scale Ni-Mo-S/C hydrogen generator.Fig. S23. Assembly of laboratory-scale Ni-Mo-S/C hydrogen generator.Fig. S24. EXAFS spectra of Ni-Mo-S/C (1:1).Fig. S25. Formation mechanism of Ni-Mo-S nanostructure.Fig. S26. Molecular models.Fig. S27. Collection of seawater.Fig. S28. HER of Ni-Mo-S/C (1:1) in seawater.Note S1. Brief explanation on the aggregation mechanism of MoS2 nanostructures.Note S2. Irreversible electrochemical oxidation of MoS2/C and Ni-Mo-S/C (1:1).Note S3. XPS study of the rapid deactivation of Pt in neutral buffer electrolyte.Note S4. Faraday efficiency calculation.Note S5. Hydrogen evolution on Ni-Mo-S/C (1:1) in real seawater.Table S1. Exchange current densities of MoS2/C, NiSx/C, and different Ni-Mo-S/C samples inneutral electrolyte.Table S2. Comparison of electrocatalysts for HER in neutral or near-neutral electrolytes.Table S3. Comparison of various MoSx-based electrocatalysts for HER.Movie S1. Hydrogen evolution on Ni-Mo-S/C (1:1) electrode in neutral electrolyte.Movie S2. Intuitive comparison test for HER.Movie S3. Laboratory-scale Ni-Mo-S/C (1:1) hydrogen generator.References (54–68)

12 of 14

Page 13: Hierarchical Ni-Mo-S nanosheets on carbon fiber …...(pH ~7), for example, current density of 10 mA/cm2 at a very small overpotential of 200 mV. Furthermore, the Ni-Mo-S/C electrode

R E S EARCH ART I C L E

on March 19, 2020

http://advances.sciencemag.org/

Dow

nloaded from

REFERENCES AND NOTES1. M. G. Walter, E. L. Warren, J. R. McKone, S. W. Boettcher, Q. Mi, E. A. Santori, N. S. Lewis,

Solar water splitting cells. Chem. Rev. 110, 6446–6473 (2010).2. A. Lasia, Hydrogen evolution reaction, in Handbook of Fuel Cells: Fundamentals, Technology

and Applications (Wiley, New York, NY, 2010), vol. 2, pp. 416–440.

3. B. Hinnemann, P. G. Moses, J. Bonde, K. P. Jørgensen, J. H. Nielsen, S. Horch, I. Chorkendorff,J. K. Nørskov, Biomimetic hydrogen evolution: MoS2 nanoparticles as catalyst for hydrogenevolution. J. Am. Chem. Soc. 127, 5308–5309 (2005).

4. B. E. Conway, B. V. Tilak, Interfacial processes involving electrocatalytic evolution and oxidationof H2, and the role of chemisorbed H. Electrochim. Acta 47, 3571–3594 (2002).

5. X. Huang, Z. Zeng, H. Zhang, Metal dichalcogenide nanosheets: Preparation, propertiesand applications. Chem. Soc. Rev. 42, 1934–1946 (2013).

6. X. Huang, C. Tan, Z. Yin, H. Zhang, 25th Anniversary article: Hybrid nanostructures basedon two-dimensional nanomaterials. Adv. Mater. 26, 2185–2204 (2014).

7. H. Li, J. Wu, Z. Yin, H. Zhang, Preparation and applications of mechanically exfoliatedsingle-layer and multilayer MoS2 and WSe2 nanosheets. Acc. Chem. Res. 47, 1067–1075(2014).

8. K. Chang, W. Chen, L-Cysteine-assisted synthesis of layered MoS2/graphene compositeswith excellent electrochemical performances for lithium ion batteries. ACS Nano 5,4720–4728 (2011).

9. X. Cao, Y. Shi, W. Shi, X. Rui, Q. Yan, J. Kong, H. Zhang, Preparation of MoS2-coated three-dimensional graphene networks for high-performance anode material in lithium-ion batteries.Small 9, 3433–3438 (2013).

10. Y.-C. Lin, D. O. Dumcenco, Y.-S. Huang, K. Suenaga, Atomic mechanism of the semi-conducting-to-metallic phase transition in single-layered MoS2. Nat. Nanotechnol. 9,391–396 (2014).

11. B. Radisavljevic, A. Radenovic, J. Brivio, V. Giacometti, A. Kis, Single-layer MoS2 transistors.Nat. Nanotechnol. 6, 147–150 (2011).

12. Z. Yin, H. Li, H. Li, L. Jiang, Y. Shi, Y. Sun, G. Lu, Q. Zhang, X. Chen, H. Zhang, Single-layerMoS2 phototransistors. ACS Nano 6, 74–80 (2012).

13. L. S. Byskov, J. K. Nørskov, B. S. Clausen, H. Topsøe, DFT calculations of unpromoted andpromoted MoS2-based hydrodesulfurization catalysts. J. Catal. 187, 109–122 (1999).

14. X. Huang, Z. Zeng, S. Bao, M. Wang, X. Qi, Z. Fan, H. Zhang, Solution-phase epitaxial growthof noble metal nanostructures on dispersible single-layer molybdenum disulfide nanosheets.Nat. Commun. 4, 1444 (2013).

15. H. Tributsch, J. Bennett, Electrochemistry and photochemistry of MoS2 layer crystals. I. J.Electroanal. Chem. Interfacial Electrochem. 81, 97–111 (1977).

16. A. Sobczynski, Molybdenum disulfide as a hydrogen evolution catalyst for water photo-decomposition on semiconductors. J. Catal. 131, 156–166 (1991).

17. A. B. Laursen, S. Kegnæs, S. Dahl, I. Chorkendorff, Molybdenum sulfides–efficient andviable materials for electro- and photoelectrocatalytic hydrogen evolution. Energy Environ.Sci. 5, 5577–5591 (2012).

18. T. F. Jaramillo, K. P. Jørgensen, J. Bonde, J. H. Nielsen, S. Horch, I. Chorkendorff, Identification ofactive edge sites for electrochemical H2 evolution from MoS2 nanocatalysts. Science 317,100–102 (2007).

19. Y. Li, H. Wang, L. Xie, Y. Liang, G. Hong, H. Dai, MoS2 nanoparticles grown on graphene: Anadvanced catalyst for the hydrogen evolution reaction. J. Am. Chem. Soc. 133, 7296–7299(2011).

20. J. Kibsgaard, Z. Chen, B. N. Reinecke, T. F. Jaramillo, Engineering the surface structure ofMoS2 to preferentially expose active edge sites for electrocatalysis. Nat. Mater. 11, 963–969(2012).

21. H. Wang, Z. Lu, S. Xu, D. Kong, J. J. Cha, G. Zheng, P.-C. Hsu, K. Yan, D. Bradshaw, F. B. Prinz,Y. Cui, Electrochemical tuning of vertically aligned MoS2 nanofilms and its application inimproving hydrogen evolution reaction. Proc. Natl. Acad. Sci. U.S.A. 110, 19701–19706 (2013).

22. Z. Lu, W. Zhu, X. Yu, H. Zhang, Y. Li, X. Sun, X. Wang, H. Wang, J. Wang, J. Luo, X. Lei, L. Jiang,Ultrahigh hydrogen evolution performance of under-water “superaerophobic” MoS2 nano-structured electrodes. Adv. Mater. 26, 2683–2687 (2014).

23. J. Xie, H. Zhang, S. Li, R. Wang, X. Sun, M. Zhou, J. Zhou, X. W. Lou, Y. Xie, Defect-rich MoS2ultrathin nanosheets with additional active edge sites for enhanced electrocatalytic hydrogenevolution. Adv. Mater. 25, 5807–5813 (2013).

24. J. Xie, J. Zhang, S. Li, F. Grote, X. Zhang, H. Zhang, R. Wang, Y. Lei, B. Pan, Y. Xie, Control-lable disorder engineering in oxygen-incorporated MoS2 ultrathin nanosheets for efficienthydrogen evolution. J. Am. Chem. Soc. 135, 17881–17888 (2013).

25. Y. Zheng, Y. Jiao, M. Jaroniec, S. Z. Qiao, Advancing the electrochemistry of the hydrogen-evolution reaction through combining experiment and theory. Angew. Chem. Int. Ed. 54,52–65 (2015).

26. J. Duan, S. Chen, M. Jaroniec, S. Z. Qiao, Porous C3N4 nanolayers@N-graphene films ascatalyst electrodes for highly efficient hydrogen evolution. ACS Nano 9, 931–940 (2015).

27. Y. Zheng, Y. Jiao, Y. Zhu, L. H. Li, Y. Han, Y. Chen, A. Du, M. Jaroniec, S. Z. Qiao, Hydrogenevolution by a metal-free electrocatalyst. Nat. Commun. 5, 3783 (2014).

Miao et al. Sci. Adv. 2015;1:e1500259 21 August 2015

28. S. Chen, J. Duan, M. Jaroniec, S.-Z. Qiao, Nitrogen and oxygen dual-doped carbon hydrogelfilm as a substrate-free electrode for highly efficient oxygen evolution reaction. Adv. Mater.26, 2925–2930 (2014).

29. X.-L. Li, Y.-D. Li, MoS2 nanostructures: Synthesis and electrochemical Mg2+ intercalation. J. Phys.Chem. B 108, 13893–13900 (2004).

30. K. S. W. Sing, D. H. Everett, R. A. W. Haul, L. Moscou, R. A. Pierotti, J. Rouquérol, T. Siemieniewska,Reporting physisorption data for gas/solid systems with special reference to the determinationof surface area and porosity. Pure Appl. Chem. 57, 603–619 (1985).

31. B. D. Cullity, S. R. Stock, Chapter 3: Diffraction I: The directions of diffracted beams, inElements of X-ray Diffraction (Pearson, Upper Saddle River, NJ, 2001), pp. 78–103.

32. L. Cheng, W. Huang, Q. Gong, C. Liu, Z. Liu, Y. Li, H. Dai, Ultrathin WS2 nanoflakes as a high‐performance electrocatalyst for the hydrogen evolution reaction. Angew. Chem. Int. Ed. 53,7860–7863 (2014).

33. C. Lee, H. Yan, L. E. Brus, T. F. Heinz, J. Hone, S. Ryu, Anomalous lattice vibrations of single-and few-layer MoS2. ACS Nano 4, 2695–2700 (2010).

34. B. C. Windom, W. Sawyer, D. W. Hahn, A Raman spectroscopic study of MoS2 and MoO3:Applications to tribological systems. Tribol. Lett. 42, 301–310 (2011).

35. D. Kong, H. Wang, J. J. Cha, M. Pasta, K. J. Koski, J. Yao, Y. Cui, Synthesis of MoS2 and MoSe2films with vertically aligned layers. Nano Lett. 13, 1341–1347 (2013).

36. S. Houssenbay, S. Kasztelan, H. Toulhoat, J. Bonnelle, J. Grimblot, Nature of the differentnickel species in sulfided bulk and alumina-supported nickel-molybdenum hydrotreatingcatalysts. J. Phys. Chem. 93, 7176–7180 (1989).

37. Y. Sun, C. Liu, D. C. Grauer, J. Yano, J. R. Long, P. Yang, C. J. Chang, Electrodeposited cobalt-sulfidecatalyst for electrochemical and photoelectrochemical hydrogen generation from water. J. Am.Chem. Soc. 135, 17699–17702 (2013).

38. Z. Chen, D. Cummins, B. N. Reinecke, E. Clark, M. K. Sunkara, T. F. Jaramillo, Core-shellMoO3-MoS2 nanowires for hydrogen evolution: A functional design for electrocatalyticmaterials. Nano Lett. 11, 4168–4175 (2011).

39. D. Merki, H. Vrubel, L. Rovelli, S. Fierro, X. Hu, Fe, Co, and Ni ions promote the catalyticactivity of amorphous molybdenum sulfide films for hydrogen evolution. Chem. Sci. 3,2515–2525 (2012).

40. P. D. Tran, M. Nguyen, S. S. Pramana, A. Bhattacharjee, S. Y. Chiam, J. Fize, M. J. Field, V. Artero,L. H. Wong, J. Loo, J. Barber, Copper molybdenum sulfide: A new efficient electrocatalyst forhydrogen production from water. Energy Environ. Sci. 5, 8912–8916 (2012).

41. D. J. Li, U. N. Maiti, J. Lim, D. S. Choi, W. J. Lee, Y. Oh, G. Y. Lee, S. O. Kim, Molybdenumsulfide/N-doped CNT forest hybrid catalysts for high-performance hydrogen evolution reac-tion. Nano Lett. 14, 1228–1233 (2014).

42. H. Vrubel, D. Merki, X. Hu, Hydrogen evolution catalyzed by MoS3 and MoS2 particles.Energy Environ. Sci. 5, 6136–6144 (2012).

43. A. J. Bard, L. R. Faulkner, Chapter 10: Techniques based on concepts of impedances, inElectrochemical Methods: Fundamentals and Applications (Wiley, New York, 1980), pp. 368–416.

44. E. Barsoukov, J. R. Macdonald, Chapter 2: Theory, in Impedance Spectroscopy: Theory, Experi-ment, and Applications (Wiley, Hoboken, NJ, 2005), pp. 27–128.

45. S. Trasatti, O. Petrii, Real surface area measurements in electrochemistry. Pure Appl. Chem.63, 711–734 (1991).

46. J. D. Benck, Z. Chen, L. Y. Kuritzky, A. J. Forman, T. F. Jaramillo, Amorphous molybdenumsulfide catalysts for electrochemical hydrogen production: Insights into the origin of theircatalytic activity. ACS Catal. 2, 1916–1923 (2012).

47. J. Bonde, P. G. Moses, T. F. Jaramillo, J. K. Nørskov, I. Chorkendorff, Hydrogen evolution onnano-particulate transition metal sulfides. Faraday Discuss. 140, 219–231 (2009).

48. H. Topsøe, B. S. Clausen, R. Candia, C. Wivel, S. Mørup, In situ Mössbauer emission spectroscopystudies of unsupported and supported sulfided Co-Mo hydrodesulfurization catalysts: Evidencefor and nature of a Co-Mo-S phase. J. Catal. 68, 433–452 (1981).

49. J. Lauritsen, S. Helveg, E. Lægsgaard, I. Stensgaard, B. Clausen, H. Topsøe, F. Besenbacher, Atomic-scale structure of Co-Mo-S nanoclusters in hydrotreating catalysts. J. Catal. 197, 1–5 (2001).

50. J. V. Lauritsen, J. Kibsgaard, G. H. Olesen, P. G. Moses, B. Hinnemann, S. Helveg, J. K. Nørskov,B. S. Clausen, H. Topsøe, E. Lægsgaard, F. Besenbacher, Location and coordination ofpromoter atoms in Co- and Ni-promoted MoS2-based hydrotreating catalysts. J. Catal.249, 220–233 (2007).

51. L. S. Byskov, B. Hammer, J. K. Nørskov, B. Clausen, H. Topsøe, Sulfur bonding in MoS2 andCo-Mo-S structures. Catal. Lett. 47, 177–182 (1997).

52. J. H. Espenson, Chapter 1: Reaction and reaction rates, in Chemical Kinetics and ReactionMechanisms (McGraw, New York, NY, 2002), pp. 1–14.

53. M. A. Lukowski, A. S. Daniel, F. Meng, A. Forticaux, L. Li, S. Jin, Enhanced hydrogen evolutioncatalysis from chemically exfoliated metallic MoS2 nanosheets. J. Am. Chem. Soc. 135, 10274–10277(2013).

54. U. Muller, Chapter 16: Linked polyhedra, in Inorganic Structural Chemistry (Wiley, Chichester,UK, ed. 2, 2007), pp. 166–189.

55. Q. He, X. Yang, W. Chen, S. Mukerjee, B. Koel, S. Chen, Influence of phosphate anion adsorptionon the kinetics of oxygen electroreduction on low index Pt(hkl) single crystals. Phys. Chem.Chem. Phys. 12, 12544–12555 (2010).

13 of 14

Page 14: Hierarchical Ni-Mo-S nanosheets on carbon fiber …...(pH ~7), for example, current density of 10 mA/cm2 at a very small overpotential of 200 mV. Furthermore, the Ni-Mo-S/C electrode

R E S EARCH ART I C L E

http://D

ownloaded from

56. S. M. A. M. Bouwens, J. P. R. Vissers, V. H. J. de Beer, R. Prins, Phosphorus poisoning ofmolybdenum sulfide hydrodesulfurization catalysts supported on carbon and alumina.J. Catal. 112, 401–410 (1988).

57. L. Qingfeng, H. A. Hjuler, N. J. Bjerrum, Oxygen reduction on carbon supported platinumcatalysts in high temperature polymer electrolytes. Electrochim. Acta 45, 4219–4226 (2000).

58. P.-C. Chen, Y.-M. Chang, P.-W. Wu, Y.-F. Chiu, Fabrication of Ni nanowires for hydrogenevolution reaction in a neutral electrolyte. Int. J. Hydrogen Energy 34, 6596–6602 (2009).

59. F. Harnisch, G. Sievers, U. Schröder, Tungsten carbide as electrocatalyst for the hydrogenevolution reaction in pH neutral electrolyte solutions. Appl. Catal. B: Environ. 89, 455–458(2009).

60. C. He, X. Wu, Z. He, Amorphous nickel-based thin film as a Janus electrocatalyst for watersplitting. J. Phys. Chem. C 118, 4578–4584 (2014).

61. S. Cobo, J. Heidkamp, P.-A. Jacques, J. Fize, V. Fourmond, L. Guetaz, B. Jousselme, V. Ivanova,H. Dau, S. Palacin, M. Fontecave, V. Artero, A Janus cobalt-based catalytic material for electro-splitting of water. Nat. Mater. 11, 802–807 (2012).

62. W. Cui, Q. Liu, Z. Xing, A. M. Asiri, K. A. Alamry, X. Sun, MoP nanosheets supported onbiomass-derived carbon flake: One-step facile preparation and application as a novelhigh-active electrocatalyst toward hydrogen evolution reaction. Appl. Catal. B: Environ.164, 144–150 (2015).

63. J. F. Callejas, J. M. McEnaney, C. G. Read, J. C. Crompton, A. J. Biacchi, E. J. Popczun, T. R. Gordon,N. S. Lewis, R. E. Schaak, Electrocatalytic and photocatalytic hydrogen production fromacidic and neutral-pH aqueous solutions using iron phosphide nanoparticles. ACS Nano8, 11101–11107 (2014).

64. N. Jiang, L. Bogoev, M. Popova, S. Gul, J. Yano, Y. Sun, Electrodeposited nickel-sulfide filmsas competent hydrogen evolution catalysts in neutral water. J. Mater. Chem. A 2, 19407–19414(2014).

65. S. Gupta, N. Patel, A. Miotello, D. C. Kothari, Cobalt-boride: An efficient and robust electro-catalyst for hydrogen evolution reaction. J. Power Sources 279, 620–625 (2015).

66. D. Merki, S. Fierro, H. Vrubel, X. Hu, Amorphous molybdenum sulfide films as catalysts forelectrochemical hydrogen production in water. Chem. Sci. 2, 1262–1267 (2011).

Miao et al. Sci. Adv. 2015;1:e1500259 21 August 2015

67. J. Kim, S. Byun, A. J. Smith, J. Yu, J. Huang, Enhanced electrocatalytic properties oftransition-metal dichalcogenides sheets by spontaneous gold nanoparticle decoration.J. Phys. Chem. Lett. 4, 1227–1232 (2013).

68. Y. Tan, P. Liu, L. Chen, W. Cong, Y. Ito, J. Han, X. Guo, Z. Tang, T. Fujita, A. Hirata, M. W. Chen,Monolayer MoS2 films supported by 3D nanoporous metals for high-efficiency electroca-talytic hydrogen production. Adv. Mater. 26, 8023–8028 (2014).

Funding: This work was supported by Nanyang Technological University (M4080977.120.50000),Singapore Ministry of Education under Academic Research Fund (AcRF) Tier 2 (ARC 26/13, no.MOE2013-T2-1-034; ARC 19/15, no. MOE2014-T2-2-093), AcRF Tier 1 (RG 9/12, RG 61/12, RGT18/13, and RG5/13), and Start-Up grant (M4081296.070.500000). This research is also conducted byNanyang Technological University–Hebrew University of Jerusalem–Ben-Gurion University Nano-materials for Energy and Water Management Programme under the Campus for Research Excel-lence and Technological Enterprise, which is supported by the National Research Foundation,Prime Minister’s Office, Singapore. Author contributions: B.L. and H.Z. conceived the idea andsupervised the research work. J.M. designed the experiments and fabricated the electrodes. J.M.conducted the SEM and XRD experiments. F.-X.X. carried out the XPS measurements. H.B.Y. con-ducted the TEM and Raman spectroscopy experiments. S.Y.K. analyzed the EIS data and performedmolecular mechanical simulation. Y.-Y.H. and H.M.C. carried out and analyzed the EXAFSexperiments. B.L. and J.M. wrote the manuscript with critical inputs from J.C., Z.F., and H.Z.Competing interests: The authors declare that they have no competing interests.

Submitted 27 February 2015Accepted 16 July 2015Published 21 August 201510.1126/sciadv.1500259

Citation: J. Miao, F.-X. Xiao, H. B. Yang, S. Y. Khoo, J. Chen, Z. Fan, Y.-Y. Hsu, H. M. Chen,H. Zhang, B. Liu, Hierarchical Ni-Mo-S nanosheets on carbon fiber cloth: A flexible electrodefor efficient hydrogen generation in neutral electrolyte. Sci. Adv. 1, e1500259 (2015).

adv

14 of 14

on March 19, 2020

ances.sciencemag.org/

Page 15: Hierarchical Ni-Mo-S nanosheets on carbon fiber …...(pH ~7), for example, current density of 10 mA/cm2 at a very small overpotential of 200 mV. Furthermore, the Ni-Mo-S/C electrode

hydrogen generation in neutral electrolyteHierarchical Ni-Mo-S nanosheets on carbon fiber cloth: A flexible electrode for efficient

Zhang and Bin LiuJianwei Miao, Fang-Xing Xiao, Hong Bin Yang, Si Yun Khoo, Jiazang Chen, Zhanxi Fan, Ying-Ya Hsu, Hao Ming Chen, Hua

DOI: 10.1126/sciadv.1500259 (7), e1500259.1Sci Adv 

ARTICLE TOOLS http://advances.sciencemag.org/content/1/7/e1500259

MATERIALSSUPPLEMENTARY http://advances.sciencemag.org/content/suppl/2015/08/20/1.7.e1500259.DC1

REFERENCES

http://advances.sciencemag.org/content/1/7/e1500259#BIBLThis article cites 62 articles, 2 of which you can access for free

PERMISSIONS http://www.sciencemag.org/help/reprints-and-permissions

Terms of ServiceUse of this article is subject to the

is a registered trademark of AAAS.Science AdvancesYork Avenue NW, Washington, DC 20005. The title (ISSN 2375-2548) is published by the American Association for the Advancement of Science, 1200 NewScience Advances

Copyright © 2015, The Authors

on March 19, 2020

http://advances.sciencemag.org/

Dow

nloaded from