Optimization of Active Sites of MoS2 Nanosheets Using Nonmetal

Jul 17, 2017 - Transition metal dichalcogenides (TMDs) have emerged as promising nonprecious noble-metal-free catalysts for photocatalytic application...
0 downloads 9 Views 2MB Size
Subscriber access provided by AUSTRALIAN NATIONAL UNIV

Article 2

Optimization of Active Sites of MoS Nanosheets using Non-metal Doping and Exfoliation into Few Layers on CdS nanorods for Enhanced Photocatalytic Hydrogen Production D. Praveen Kumar, Myeong In Song, Sangyeob Hong, Eun Hwa Kim, Madhusudana Gopannagari, D. Amaranatha Reddy, and Tae Kyu Kim ACS Sustainable Chem. Eng., Just Accepted Manuscript • DOI: 10.1021/acssuschemeng.7b00978 • Publication Date (Web): 17 Jul 2017 Downloaded from http://pubs.acs.org on July 17, 2017

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.

ACS Sustainable Chemistry & Engineering 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 26

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

ACS Sustainable Chemistry & Engineering

Optimization of Active Sites of MoS2 Nanosheets using Non-metal Doping and Exfoliation into Few Layers on CdS nanorods for Enhanced Photocatalytic Hydrogen Production D. Praveen Kumar#, Myeong In Song#, Sangyeob Hong, Eun Hwa Kim, Madhusudana Gopannagari, D. Amaranatha Reddy and Tae Kyu Kim* Department of Chemistry and Chemical Institute for Functional Materials, Pusan National University, 2, Busandaehak-ro 63beon-gil, Geumjeong-Gu, Busan 46241, Republic of Korea

KEYWORDS: Photocatalyst, hydrogen production, non-metal doping, exfoliation, MoS2

Corresponding Author *

Email: [email protected]

#

These authors equally contributed to this work

ACS Paragon Plus Environment

1

ACS Sustainable Chemistry & Engineering

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

Page 2 of 26

ABSTRACT

Transition metal dichalcogenides (TMDs) have emerged as promising non-precious noble-metal-free catalysts for photocatalytic applications. Among TMDs, MoS2 has been extensively studied as a cocatalyst due to its exceptional activity for photocatalytic hydrogen evolution. However, the catalytic activity of MoS2 is triggered only by the active S atoms on its exposed edges, whereas the majority of S atoms present on the basal plane are catalytically inactive. Doping of foreign non-metals into the MoS2 system is an appealing approach for activation of the basal plane surface as an alternative for increasing the concentration of catalytically active sites. Herein, we report the development of earth-abundant, few-layered boron-doped MoS2 nanosheets decorated on CdS nanorods (FBMC) employing simple methods and their use for photocatalytic hydrogen evolution under solar irradiation, with lactic acid as a hole-scavenger, under optimal conditions. The FBMC material exhibited a high rate of H2 production (196 mmol·h−1·g−1). The presence of few-layered boron-doped MoS2 (FBM) nanosheets on the surface of the CdS nanorods effectively separated the photogenerated charge carriers and improved the surface shuttling properties for efficient H2 production due to their extraordinary number of active edge sites with superior electrical conductivity. In addition, the observed H2 evolution rate of FBMC was much higher than that for the individual few-layered MoS2-assisted CdS (FMC) and bulk boron-doped MoS2/CdS (BBMC) photocatalysts. To the best of our knowledge, this is the highest H2 production rate achieved with MoS2-based CdS photocatalysts for water splitting under solar irradiation. Considering its low cost and high efficiency, this system has great potential as a photocatalyst for use in various fields.

ACS Paragon Plus Environment

2

Page 3 of 26

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

ACS Sustainable Chemistry & Engineering

INTRODUCTION

Rapid advancements in science and technology and massive population growth have led to serious environmental pollution and escalation of the energy crunch.1 Semiconductor photocatalysts have fascinated scientists in the search for solutions to these problems due to the extensive applicability of such systems.2 Photocatalytic H2 evolution from water on semiconductors is one attractive way to simultaneously combat environmental pollution and the energy crunch as the process is pollution-free and economically viable and the materials are easily stored.3 Numerous photocatalysts have been described for the splitting of water in the presence or absence of sacrificial agents, generating H2 as a clean energy source.4−10 Nevertheless, the majority of these catalysts require ultraviolet (UV) light and their effectiveness is unsatisfactory for practical application. The development of efficient visible-light-accessible photocatalysts is an extremely desirable challenge to achieve effective consumption of solar energy, given that visible light comprises a greater fraction of the solar spectrum (43%) than UV light (4%). Among the various semiconductor photocatalysts, CdS, with a band-gap of ~2.4 eV and an appropriate conduction band potential, has been extensively investigated as a visiblelight-driven photocatalyst for hydrogen generation.11 Nevertheless, investigations on bare CdS have proven it to be been inadequate due to its low photocatalytic efficiency, attributed to rapid recombination of the photogenerated electrons and holes.12 Since the invention of graphene, there has been much interest in auxiliary 2D layered materials due to their distinctive electronic properties and high surface-to-volume ratios, making them prospectively superior platforms for photocatalytic applications.13-16 Among the known 2D layered materials, transition metal dichalcogenides (TMDs)—having extraordinary properties—

ACS Paragon Plus Environment

3

ACS Sustainable Chemistry & Engineering

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

Page 4 of 26

are of particular interest; for instance, photocatalysts can be combined with TMDs as cocatalysts to generate low-cost materials with an edge-terminated structure.17−21 TMDs are a class of materials with the general formula MX2, where M refers to the transition metal (such as Mo, W, Co and etc.) and X indicates the chalcogen (such as S, Se, Te and etc.).22 Among the TMDs, MoS2 is considered an excellent cocatalyst for photocatalytic H2 production from water because it possesses excellent H2 activation ability.23 MoS2 has emerged as the most attractive and promising TMD due to its peculiar properties. MoS2 generally adopts a sandwich structure comprising three stacked atomic layers (S−Mo−S) linked by van der Waals forces and has been used in photocatalysis.24−29 Moreover, density functional calculation of the free energy has shown that the S atoms on the exposed edges of MoS2 bond strongly to H+ in solution and the bound protons are easily reduced to H2 by electrons.30 Theoretically, the catalytic activity of MoS2 is derived only from the active S atoms on its exposed edges, but most S atoms on the basal plane show no activity.31 Therefore, conversion of MoS2 nanosheets into few-layered nanosheets with a large number of exposed active sites increases their activity, thereby making MoS2 a promising cocatalyst as a low-cost alternative to platinum. Despite the large number of exposed active sites in few-layered MoS2 nanosheets, the co-catalytic activity is restricted to some extent due to its inefficient electrical conductivity. The modification of TMDs by doping with transition metals/non-metals can expand their utility in such applications, thereby expanding their prospective for technological applications.32 The comparative quantity of active sites can be increased by morphological control and, in the case of TMDs, this can be accomplished by doping.33 Although the basal plane is not generally as catalytically active as the edge sites, doping offers a strategy for the activation of this surface as an alternative approach for increasing the density of catalytically active sites.34 Atomic doping

ACS Paragon Plus Environment

4

Page 5 of 26

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

ACS Sustainable Chemistry & Engineering

of a material is a technique that can alter the structure and/or properties depending on the dopant and its concentration. Supersession can typically occur as direct supersession of atoms in the lattice (if the dopant is well matched in terms of size, valence, and coordination) or into interstitial sites between subsisting atoms in the lattice. In the context of layered materials, dopant atoms may also be intercalated between layers or alter the structure to engender formation of an incipient phase. In optoelectronic applications, doping can be utilized to control p- or ntype semiconducting behavior by modulating the Fermi level and/or band-gap of the semiconductor.34 Judicious selection of the dopant can yield magnetic semiconductors: a potential technique that can be applied for the development of spintronics, which may form the substructure of solid-state storage devices and computing components.34 Metal-doped TMDs have been extensively studied relative to non-metal doped TMDs, though non-metal doping (with B, N, P and O) is an effective strategy for altering the physical properties of TMDs. Extensive research has been undertaken to increase the number of active S atoms on the exposed edges of MoS2 by conversion of many-layered MoS2 to mono- and few-layered species and conversion from the 2H phase to the 1T phase. However, the separated mono- and few-layer species and 1T phases are unstable than 2H phase. The MoS2 monolayer is known to have two phases: the trigonal prismatic (labelled as 2H, D3h) and octahedral (labelled as 1T, Oh). The 2H phase is relatively stable, but semiconducting and of poor conductivity. The 1T phase is metastable at room temperature, but metallic and of better conductivity.35-38 Hence, herein, we concentrate on increasing the catalytically active sites by activating S atoms on the basal plane; these S species are not generally as catalytically active as the surface edge sites. We report the synthesis of earth-abundant, few-layered, boron-doped MoS2 nanosheets decorated on CdS nanorods (FBMC) by simple methods. The synthesized materials are evaluated for photocatalytic

ACS Paragon Plus Environment

5

ACS Sustainable Chemistry & Engineering

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

Page 6 of 26

hydrogen evolution under solar irradiation with lactic acid as a hole-scavenger under the optimal conditions. The FBMC material exhibits a high rate of H2 production (196 mmol·h−1·g−1). The presence of few-layered boron doped MoS2 (FBM) nanosheets on the surface of the CdS nanorods effectively separates the photogenerated charge carriers and improves the surface shuttling properties for efficient H2 production due to the extraordinary number of active edge sites with superior electrical conductivity. EXPERIMENTAL SECTION Materials. Cadmium acetate dihydrate (Cd(CH3COO)2·2H2O), sodium molybdate dihydrate (Na2MoO4·2H2O), boric acid, and ethanol were purchased from Daejung Chemicals & Metals Co. Ltd., Korea. Thiourea (NH2CSNH2) and thioacetamide (C2H5NS) were obtained from Alfa Aesar. All chemicals were used without further purification. Synthesis of few-layered, boron-doped MoS2/CdS nanorods (FBMC). First, one-dimensional CdS nanorods and bulk boron-doped MoS2 (BBM) nanosheets were synthesized separately using hydrothermal methods; the detailed experimental procedures are provided in the Supporting Information.23, 39 Finally the FBMC composites were synthesized by simple ultrasonication23. Different weight percentages (1−7 %) of BBM were separately dispersed in 10 mL of dimethylformamide (DMF) and ultrasonicated for 3 h at room temperature. The BBM suspension was exfoliated and homogeneously separated into a few layers by the sonication process; the as-prepared CdS nanorods (100 mg) were then added and the mixture was again ultrasonicated for 1 h, followed by 12 h of stirring to improve the interaction between the exfoliated boron-doped MoS2 layers and the CdS nanorods. The black boron-doped MoS2 became greenish; this greenish solid product was washed thoroughly with de-ionized water and

ACS Paragon Plus Environment

6

Page 7 of 26

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

ACS Sustainable Chemistry & Engineering

ethanol many times to eliminate DMF and then dried at 60 oC for 12 h. A schematic representation of the materials preparation is shown in Scheme 1.

Scheme 1. Schematic representation of the preparation procedure of FBMC nanocomposites. RESULTS AND DISCUSSION The solvothermally synthesized CdS nanorods (NRs) were structurally characterized by X-ray diffraction (XRD), which is shown in Fig. 1(a). The diffraction pattern of the bare CdS NRs was clearly indexed to the pure hexagonal phase of CdS, which is in good agreement with literature values (JCPDS card #: 41-1049). Although the boron doped bulk MoS2/CdS (BBMC) and FBMC nanocomposites contained 6 % boron doped MoS2 on the CdS substrate as a cocatalyst, no diffraction peaks of boron doped MoS2 were observed, which may be ascribed to the relatively low amount and low diffraction intensity of boron doped MoS2.23 Further, the formation of FBM by exfoliation was confirmed by the differences in the respective diffraction patterns of FBM and BBM, such as the shift in the diffraction peak at 2θ = 14.5° to a value of 10.0° upon exfoliation, providing a clear evidence of an increase in the space in-between the boron doped MoS2 layers (Fig. S1 in Supporting Information (SI)).11 Optical absorption of as-synthesized the CdS NRs and the composite materials were studied by UV/Vis diffuse reflectance spectroscopy (DRS), as shown in Fig. 1(b). The bare CdS

ACS Paragon Plus Environment

7

ACS Sustainable Chemistry & Engineering

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

Page 8 of 26

NRs had a prominent absorption with an absorption edge at about 520 nm, which corresponds to the intrinsic band-gap absorption of CdS NRs.11 The optical band-gap of the CdS semiconductor was calculated to be 2.38 eV by utilizing the Kubelka–Munk method, which is consistent with the reported value.13 Decoration of boron doped MoS2 on the CdS NRs gave rise to the same absorption edge as that of the bare CdS NRs, indicating that the band-gap of CdS was unchanged. The deposition of boron doped MoS2 on the surface of the CdS NRs did not alter the band structure of CdS. However, the broad background absorption of the composite was extended in the range of 530–750 nm because black boron doped MoS2 can lead to a large increase in the opacity.

Figure 1. (a) XRD patterns and (b) UV/Vis diffuse reflectance spectra of CdS and related composites (BBMC and FBMC). The morphology of the FBMC composite was analyzed by transmission electron microscopy (TEM) and energy dispersive X-ray spectral (EDS) mapping analysis, as displayed in Fig. 2. Figs. 2(a)–(d) show FBM dispersed on the CdS nanorods, with CdS exhibiting a rod-

ACS Paragon Plus Environment

8

Page 9 of 26

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

ACS Sustainable Chemistry & Engineering

like 1-D morphology decorated with FBM nanosheets, as indicated by the red circles; the highresolution TEM (HRTEM) images of the composite (Figs. 2(c) and 2(d)) also clearly confirm the presence of CdS lattice fringes with a spacing of 0.33 nm and FBM, respectively. The elemental composition was determined by EDS mapping analysis, providing good evidence of the presence of FBM on the CdS nanorods, as shown in Figs. 2(e)–2(h). Fig. 2(e) confirms the presence of three elements, Cd, S and Mo. Figs. 2(f)–(h) show the individual elemental mapping results. From EDS, it was confirmed that Mo was well-dispersed, but at a much lower percentage than Cd and S; this was due to the much lower weight percentage of FBM used in the synthesis of FBMC. We were unable to detect boron in the EDS mapping due to its smaller atomic size and significantly lower percentage compared to FBMC.38 FESEM and TEM image of CdS NRs and TEM images of bulk MoS2/CdS and few layered MoS2/CdS are shown in Fig. S2, confirming high agglomeration of MoS2 layers on CdS in bulk MoS2/CdS and few separated layers on CdS in few layered MoS2/CdS.

ACS Paragon Plus Environment

9

ACS Sustainable Chemistry & Engineering

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

Page 10 of 26

Figure 2. (a)–(b) TEM and HRTEM images of FBMC, respectively, (c)–(d) HRTEM images of CdS and boron-doped MoS2, respectively, and (e)–(h) elemental analysis of FBMC (Cd, S and Mo elements).

Figure 3. (a)–(e) XPS spectra of FBMC composite showing expanded Cd, S, Mo and B regions. (f) Photoluminescence (PL) spectra of CdS, BBMC and FBMC. The chemical states and elemental compositions at the surface of FBMC nanocomposite were analyzed by X-ray photoelectron spectroscopy (XPS), as shown in Figs. 3(a)–(e).

ACS Paragon Plus Environment

10

Page 11 of 26

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

ACS Sustainable Chemistry & Engineering

According to XPS survey spectrum, we found the existence of B, Cd, Mo, and S and trace amounts of C and O elements in the FBMC. The B element peak in XPS clearly evidenced boron doping in FBM. Furthermore, the chemical states of the elements were confirmed from the highresolution B 1s, Cd 3d, Mo 3d and S 2p XPS spectra, showing peaks of the B 1s at binding energies of 190.6 eV and of Cd 3d at 404.6 and 411.7 eV, which correspond to Cd 3d3/2 and Cd 3d5/2, respectively. The binding energy difference indicated the chemical state of Cd in the nanocomposite was +2. The binding energy of the Mo 3d3/2 (at 231.8 eV) and those of the S 2p3/2 (at 161.5 eV) clearly confirms the presence of boron doped MoS2 in FBMC composite.11, 13, 39 To confirm the effect of boron doping in FBMC, we compared XPS results of FMC (Fig. S6). Fig. S6(a) shows the Cd 3d binding energies at 404.9 (3d5/2) and 411.5 eV (3d3/2), confirming the +2 oxidation state of Cd in CdS.29 In addition, the S 2p peaks are present at binding energies of 161.2 and 162.5 eV. The binding energies of the Mo 3d spectrum appear at 228.3 (Mo 3d5/2) and 232.0 eV (Mo 3d3/2), which indicates that Mo4+ is the dominant oxidation state. The binding energies of Mo and S are slightly differing in FBMC and FMC, which could be ascribed to the presence of boron in FBMC. Photoluminescence (PL) experiments were performed to evaluate the efficiency of trapping and recombination of the photogenerated charge carriers. The PL spectra of the CdS, BBMC and FBMC were measured after 380 nm excitation (Fig. 3(f)). A near-band-edge emission around 550 nm was observed for all samples, which is due to the exciton emission of CdS.13 In contrast, the decreased PL intensity of BBMC and FBMC indicated facilitated transport and separation of photogenerated electrons and holes. Notably, the BBMC material also exhibited an obvious decrease in the PL intensity, which can be ascribed to the lower probability of recombination of the photogenerated carriers. Moreover, the lower intensity of the emission

ACS Paragon Plus Environment

11

ACS Sustainable Chemistry & Engineering

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

Page 12 of 26

bands indicates that most of the surface states were passivated, thereby making the electrons available for photocatalytic H2 evolution. The emission bands of FBMC were much less intense than those of BBMC, which implies that FBM is favorable for the charge carrier transfer processes (on a faster time-scale) because of the increased number of surface edge sites and separated layers. Furthermore, the obtained PL results were compared with few layered MoS2/CdS (FMC) and bulk MoS2/CdS (BMC), as in Fig. S5. Fig. S5 clearly explains that PL intensity of FBMC and BBMC is much lower than other nanocomposites (FMC and BMC), thus indicating that effective charge carrier separations in FBMC are due to boron doping in the nanocomposites. To

further evaluate

the separation

efficiency and

charge carrier transport,

photoelectrochemical (PEC) analysis was carried out by using indium-tin-oxide (ITO) electrodes coated with CdS NRs and BBMC and FBMC nanocomposites under simulated sunlight irradiation by employing an electrochemical potentiostat. Fig. 4(a) shows the time-dependent photocurrent responses of CdS and the BBMC and FBMC composites with 30 s simulated sunlight on/off cycles. As expected, the FBMC nanocomposites showed the highest photocurrent intensity relative to the CdS NRs and BBMC, suggesting a higher separation efficiency of the photoexcited electron–hole pairs in the former. Consequently, more electrons could be captured by protons to form H2. Moreover, the photocurrent responses were highly reproducible for several on–off cycles and remained stable. This indicates that FBM on CdS can effectively prevent photocorrosion. The photocurrent data are consistent with PL results. Based on these results, it is evident that the FBM nanostructures enhance the photocatalytic activity of the CdS nanostructures.

ACS Paragon Plus Environment

12

Page 13 of 26

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

ACS Sustainable Chemistry & Engineering

The presence of FBM on CdS nanorods plays a crucial role on enhanced photocatalytic H2 production rate. Structural measurements indicate effective separations of few-layered MoS2 nanosheets, which is dissimilar from those of BBM. The effective layer separation into few layers increases the number of active sites with higher surface area. In addition, morphological properties clearly show that the better deposition of FBM on CdS NRs, which leads to better interactions between three components, benefiting effective charge transportations. The optical properties of the FBMC provide evidences for a more effective solar light utilization compared to BBMC and CdS, based on its higher absorption coefficient in the visible region. Charge carrier transfer studies (PL analyses) also confirm the fast migration of carriers in FBMC reducing charge carriers recombination. PEC studies support the fact that the separation and transportation of charge carriers in FBMC occur at a higher current density than those in BBMC and CdS. All these findings strongly indicate FBM greatly influences the H2 production rate of CdS NRs under solar light irradiation. Photocatalytic H2 generation experiments employing CdS NRs and FBMC were performed under solar simulator irradiation using lactic acid aqueous solution. It has been proven that lactic acid is a good hole scavenger for suppressing the recombination of photogenerated charge carriers. The data show that the amount of H2 gas generated increased linearly with respect to the irradiation time; the loading of FBM as a cocatalyst on the CdS NRs remarkably influenced the rate of H2 production (Fig. 4(b)). The undecorated CdS NRs exhibited significantly lower activity (5 mmol·h−1·g−1). The loading of FBM was varied from 0 to 7.0 wt.%. The composite catalyst showed a significantly higher rate of H2 generation than the CdS NRs; at a loading of 6% FBM, the rate of H2 production reached 196 mmol·h−1·g−1, which is 38fold higher than that of the bare CdS NRs, indicating that FBM is an excellent cocatalyst. Above

ACS Paragon Plus Environment

13

ACS Sustainable Chemistry & Engineering

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

Page 14 of 26

and below this optimum loading of FBM on the CdS nanorods, the rate of H2 production was lower. This is ascribed to the lower number of catalytically active sites at low FBM loading, whereas higher loading leads to obstruction of the incident light and inhibits the generation of electrons from the CdS NRs. In addition, higher loading leads to a greater number of S atoms on the basal planes; these S atoms are inactive, resulting in reduced photocatalytic activity as the cocatalytic activity of FBM depends strongly on S atoms on the surface exposed edges.

ACS Paragon Plus Environment

14

Page 15 of 26

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

ACS Sustainable Chemistry & Engineering

Figure 4. (a) Photoelectrochemical characterization of CdS and related composites and photocatalytic H2 production studies, (b) effects of FBM loading ratio on CdS NRs, (c) recyclability test for FBMC, (d) long term stability test of FBMC with 1 mg of catalyst dispersed in 15 mL of 20 vol.% aqueous lactic acid solution under simulated solar light irradiation, (e) effect of lactic acid concentration on performance of FBMC in aqueous solution with 1 mg of catalyst, (f) effect of catalyst loading in 15 mL of 20 vol.% aqueous lactic acid solution on H2 production. To assess the stability of catalyst, we evaluated the time course of photocatalytic hydrogen generation using the optimized FBMC catalyst with lactic acid as a scavenger as shown in Fig. 4(c). Five reaction cycles, each lasting or 5 h were performed; the rate of hydrogen production remained almost unchanged over the five cycles. However, a marginal decrease was observed after three reaction cycles, at which point an additional 3 mL of lactic acid was added, followed by two more reaction cycles. The performance was similar to that in the first cycle, which clearly indicates that the composite material is stable. To evaluate the stability and durability of the optimized FBMC photocatalysts, we also performed photocatalytic H2 generation experiments for an extended period of 60 h (Fig. 4(d)). The rate of H2 evolution improved with respect to time for the first 50 h and then decreased slightly. We confirmed that the decrease in the production rate was due to the photoreactor being filled with H2 gas, resulting in a sluggish rate of H2 evolution. Thus, the generated gas was removed from the reactor and the same experiment was performed for a further 10 h. The rate of H2 production remained similar to that in the first 10 h. The effect of the scavenger concentration on the rate of H2 production using the optimized FBMC catalyst was also evaluated (Fig. 4(e)). Initially, the rate of H2 production (210

ACS Paragon Plus Environment

15

ACS Sustainable Chemistry & Engineering

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

Page 16 of 26

mmol·h−1·g−1) was proportional to the concentration of lactic acid up to 4 mL in 15 mL of reaction solution, and thereafter, was indirectly proportional to the increase in the scavenger concentration. The decreased rate of H2 production is due to the high formation of intermediates at a high concentration of lactic acid. Photoexperiments were conducted with different amounts (1 to 20 mg) of the optimized FBMC catalyst in the reaction solution (in 15 mL), demonstrating that the rate of H2 production differed significantly with respect to the amount catalyst in the reaction solution. At loadings of 1−5 mg, the rate of H2 production increased up to 277 µmol·h−1 with respect to the amount of catalyst. The H2 production rate became saturated at loadings exceeding 5 mg in 15 mL of reaction solution due to the large amount of catalyst, causing a shielding effect for the suspended catalyst particles (Fig. 4(f)). In the comparative assessments of the H2 production rate for CdS, boron-doped MoS2, pristine few bulk MoS2/CdS (BMC), few layered MoS2/CdS (FMC), BBMC and FBMC, the H2 evolution rate of the optimized 6 wt.% FBMC (196 mmol·h−1 ·g−1) is 38- and 2.7-fold greater than that of CdS (5 mmol·h−1 ·g−1) and BBMC (72 mmol·h−1 ·g−1), respectively (Fig. S3(b)). We could not observe hydrogen production from boron-doped MoS2. The rate of H2 production of FMC and BMC are 125 mmol·h−1 ·g−1 and 60 mmol·h−1 ·g−1 respectively. We also studied effects of scavengers for 6 wt.% FBMC catalyst (Fig. S3(c)). The apparent quantum efficiency (QE) of the optimized FBMC nanocomposite under visible-light irradiation using a 150 W Xe lamp with a 425 nm band-pass filter was determined. The quantum yield was estimated to be around 31.5% (the rate of H2 production was 84 mmol·h−1·g−1); a comprehensive description of the experimental procedure for the QE measurement and calculations is provided in the SI. A plausible mechanism for the efficient production of H2 under solar irradiation using lactic acid as a hole scavenger and FBMC as a photocatalyst is depicted in Fig. 5. Excitation of

ACS Paragon Plus Environment

16

Page 17 of 26

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

ACS Sustainable Chemistry & Engineering

the semiconducting CdS in FBMC by solar light generates electron–hole pairs at the conduction band (CB) and valence band (VB), respectively. The FBM nanosheets effectively separate the photogenerated charge carriers and improve the surface shuttling properties; the transfer is followed by reduction of protons to produce molecular hydrogen. The sacrificial agent (i.e., lactic acid) is consumed by the photogenerated holes in the VB of CdS to generate pyruvic acid. However, FBMC exhibits superior activity relative to that of the BBMC and FMC. As supported by all experimental data, this improvement is due to the few-layered nature of boron doped MoS2 and non-metal doping into the MoS2 system, which strongly hinder the recombination of the photogenerated charge carriers and enhance the surface shuttling properties by increasing the surface active sites and thus the electrical conductivity. Mott-Schottky (Fig. S4 in SI) and photocurrent measurements are good supports for the proposed reaction mechanism. In addition, the conduction band potential of FBM is very close to the CdS conduction band potential, which leads to very fast migration of electrons from FBM to CdS.

Figure 5. Schematic representation of the proposed mechanism of reaction of the FBMC nanocomposite.

ACS Paragon Plus Environment

17

ACS Sustainable Chemistry & Engineering

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

Page 18 of 26

CONCLUSIONS The study demonstrates that earth-abundant, noble-metal-free, few-layered boron-doped MoS2 nanosheets can be used as an efficient co-catalyst for CdS nanorods, leading to extraordinary photocatalytic H2 production under simulated solar light irradiation. The FBMC material exhibits a high rate of H2 production (196 mmol·h−1 ·g−1). The presence of the FBM nanosheets on the surface of the CdS nanorods effectively separates the photogenerated charge carriers and improves the surface shuttling properties for efficient H2 production due to the extraordinary number of active edge sites with superior electrical conductivity. Furthermore, the observed H2 evolution rate is much higher than those for FMC and the BBMC. Moreover, to the best of our knowledge, this is the highest H2 production rate achieved with a MoS2-based CdS photocatalyst for water splitting under solar irradiation. Considering its low cost and high efficiency, this system has great potential for the development of highly efficient photocatalysts for use in various fields. Hence, this system provides significant motivation for further research on the conversion of solar energy to chemical fuels. ASSOCIATED CONTENT Supporting Information. Preparation methods of CdS NRs, BM and BBMC, characterization techniques details and complete details of photocatalytic activity test, PEC and Mott-Schottky analyses. XRD pattern of BM, few-layered MoS2 (FM), BBM and FBM. Elemental analyses, FESEM and TEM image of CdS NRs, FMC and BMC. Photocatalytic H2 production assessments of studies of CdS, BMC, FMC, BBMC and FBMC. Mott−Schottky plots of BBM, FBM and CdS NRs. PL spectra of all synthesized materials and XPS analysis of FMC.

ACS Paragon Plus Environment

18

Page 19 of 26

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

ACS Sustainable Chemistry & Engineering

Comparison of photocatalytic H2 evolution rate of MoS2/CdS based photocatalysts. This material is available free of charge via the Internet at http://pubs.acs.org. ACKNOWLEDGMENT This work was supported by National Research Foundation of Korea (NRF) grants, funded by the Korean Government (MSIP) (2014R1A4A1001690 and 2016R1E1A1A01941978).

REFERENCES (1) Chu, S.; Majumdar, A. Opportunities and challenges for a sustainable energy future. Nature 2012, 488, 294−303. DOI:10.1038/nature11475 (2) Tran, P. D.; Wong, L. H.; Barber, J.; Loo, J. S. C. Recent advances in hybrid photocatalysts for

solar

fuel

production.

Energy

Environ.

Sci.

2012,

5,

5902–5918.

DOI:

10.1039/C2EE02849B (3) Walter, M. G.; Warren, E. L.; Mc Kone, J. R.; Boettcher, S. W.; Santori, E. A.; Lewis, N. S. Solar water splitting cells. Chem. Rev. 2010, 110, 6446−6473. DOI: 10.1021/cr1002326 (4) Babu, S. G.; Vinoth, R.; Kumar, D. P.; Shankar, M. V.; Chou, H. L.; Vinodgopal, K.; Neppolian, B. Influence of electron storing, transferring and shuttling assets of reduced graphene oxide at the interfacial copper doped TiO2 p-n hetero-junction for the increased hydrogen production. Nanoscale 2015, 7, 7849–7857. DOI: 10.1039/C5NR00504C (5) Kumar, D. P.; Shankar, M. V.; Kumari, M. M.; Sadanandam, G.; Srinivas, B.; Durgakumari, V. Nano-size effects on CuO/TiO2 catalysts for highly efficient H2 production under

solar

light

irradiation.

Chem.

Comm.

2013,

49,

9443−9445.

DOI:

10.1039/C3CC44742A

ACS Paragon Plus Environment

19

ACS Sustainable Chemistry & Engineering

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

Page 20 of 26

(6) Kumar, D. P.; Reddy, N. L.; Kumari, M. M.; Srinivas, B.; Durgakumari, V.; Sreedhar, B.; Roddatis, V.; Bondarchuk, O.; Karthik, M.; Neppolian, B.; Shankar, M. V. Cu2O-sensitized TiO2 nanorods with nanocavities for highly efficient photocatalytic hydrogen production under solar irradiation. Sol. Energy Mater. Sol. Cells 2015, 136, 157–166. DOI: 10.1016/j.solmat.2015.01.009 (7) Kumar, D. P.; Reddy, N. L.; Srinivas, B.; Durgakumari, V.; Roddatis, V.; Bondarchuk, O.; Karthik, M.; Ikuma, Y.; Shankar, M. V. Stable and active CuxO/TiO2 nanostructured catalyst for proficient hydrogen production under solar light irradiation. Sol. Energy Mater. Sol. Cells 2016, 146, 63–71. DOI: 10.1016/j.solmat.2015.11.030 (8) Kumar, D. P.; Reddy, N. L.; Karthik, M.; Neppolian, B.; Madhavan, J.; Shankar, M. V. Solar light sensitized p-Ag2O/n-TiO2 nanotubes heterojunction photocatalysts for enhanced hydrogen production in aqueous-glycerol solution. Sol. Energy Mater. Sol. Cells 2016, 154, 78–87. DOI: 10.1016/j.solmat.2016.04.033 (9) Kumar, D. P.; Kumari, V. D.; Karthik, M.; Sathish, M.; Shankar, M. V. Shape dependence structural, optical and photocatalytic properties of TiO2 nanocrystals for enhanced hydrogen production by photoinduced glycerol reforming. Sol. Energy Mater. Sol. Cells 2017, 163, 113–119. DOI: 10.1016/j.solmat.2017.01.007 (10) Kumar, D. P.; Choi, J.; Hong, S.; Reddy, D. A.; Lee, S.; Kim, T. K. Rational synthesis of MOF-derived noble metal-free nickel phosphide nanoparticles as a highly efficient cocatalyst for photocatalytic hydrogen evolution. ACS Sustain. Chem. Eng. 2016, 4, 7158– 7166. DOI: 10.1021/acssuschemeng.6b02032 (11) Kumar, D. P.; Hong, S.; Reddy, D. A.; Kim, T. K. Noble metal-free ultrathin MoS2 nanosheet-decorated CdS nanorods as an efficient photocatalyst for spectacular hydrogen

ACS Paragon Plus Environment

20

Page 21 of 26

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

ACS Sustainable Chemistry & Engineering

evolution under sun light irradiation. J. Mater. Chem. A 2016, 4, 18551–18558. DOI: 10.1039/C6TA08628D (12) Marschall, R. Semiconductor composites: strategies for enhancing charge carrier separation to improve photocatalytic activity. Adv. Funct. Mater. 2014, 24, 2421−2440. DOI: 10.1002/adfm.201303214 (13) Reddy, D. A.; Park, H.; Ma, R.; Kumar, D. P.; Lim, M.; Kim, T. K. Heterostructured WS2MoS2 ultrathin nanosheets integrated on CdS nanorods for promoting charge separation and migration to improve solar-driven photocatalytic hydrogen evolution. ChemSusChem 2017, 10, 1563−1570. DOI: 10.1002/cssc.201601799 (14) Reddy, D. A.; Kim, H. K.; Kim, Y.; Lee, S.; Choi, J.; Islam, M. J.; Kumar, D. P.; Kim, T. K. Multicomponent transition metal phosphides derived from layered double hydroxide double-shelled nanocages as an efficient non-precious co-catalyst for hydrogen production. J. Mater. Chem. A 2016, 4, 13890–13898. DOI: 10.1039/C6TA05741A (15) Lu, Q.; Yu, Y.; Ma, Q.; Chen, B.; Zhang, H. 2D Transition metal dichalcogenide nanosheet based composites for photocatalytic and electrocatalytic hydrogen evolution reactions. Adv. Mater. 2016, 28, 1917–1933. DOI: 10.1002/adma.201503270 (16) Zhang, X.; Lai, Z.; Tan, C.; Zhang, H. Solution processed two dimensional MoS2 nanosheets: Preparation, hybridization, and applications. Angew. Chem. Int. Ed. 2016, 55, 8816–8838. DOI: 10.1002/anie.201509933 (17) Chen, J.; Wu, X. J.; Yin, L.; Li, B.; Hong, X.; Fan, Z.; Chen, B.; Xue, C.; Zhang, H. Onepot synthesis of CdS nanocrystals hybridized with single-layer transition metal dichalcogenide nanosheets for efficient photocatalytic hydrogen evolution. Angew. Chem. Int. Ed. 2015, 54, 1210–1214. DOI: 10.1002/anie.201410172

ACS Paragon Plus Environment

21

ACS Sustainable Chemistry & Engineering

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

Page 22 of 26

(18) Zhou, W.; Yin, Z.; Du, Y.; Huang, X.; Zeng, Z.; Fan, Z.; Liu, H.; Wang, J.; Zhang, H. Synthesis of few layer MoS2 nanosheet coated TiO2 nanobelt heterostructures for enhanced photocatalytic activities. Small, 2013, 9, 140–147. DOI: 10.1002/smll.201201161 (19) Xiang, Q. J.; Yu, J. G.; Jaroniec, M. Graphene-based semiconductor photocatalysts. Chem. Soc. Rev. 2012, 41, 782–796. DOI:10.1039/c1cs15172j (20) Xiang, Q. J.; Yu, J. G. Graphene-based photocatalysts for hydrogen generation. J. Phys. Chem. Lett. 2013, 4, 753–759. DOI: 10.1021/jz302048d (21) Low, J. X.; Cao, S. W.; Yu, J. G.; Wageh, S. Two-dimensional layered composite photocatalysts. Chem. Comm. 2014, 50, 10768–10777. DOI: 10.1039/C4CC02553A (22) Ambrosi, A.; Sofer, Z.; Pumera, M. 2H to 1T phase transition and hydrogen evolution activity of MoS2, MoSe2, WS2 and WSe2 strongly depends on the MX2 composition. Chem. Comm. 2015, 51, 8450–8453. DOI: 10.1039/C5CC00803D (23) He, J.; Chen, L.; Wang, F.; Liu, Y.; Chen, P.; Au, C. T.; Yin, S. F. CdS nanowires decorated with ultrathin MoS2 nanosheets as an efficient photocatalyst for hydrogen evolution. ChemSusChem 2016, 9, 624–630. DOI: 10.1002/cssc.201501544 (24) Li, Y. G.; Wang, H. L.; Xie, L. M.; Liang, Y. Y.; Hong, G. S.; Dai, H. J. MoS2 nanoparticles grown on graphene: an advanced catalyst for the hydrogen evolution reaction. J. Am. Chem. Soc. 2011, 133, 7296–7299. DOI: 10.1021/ja201269b (25) Choi, J.; Reddy, D. A.; Han, N. S.; Hong, S.; Kumar, D. P.; Song, J. K.; Kim, T. K. Modulation of charge carrier pathways in CdS nanospheres by integrating MoS2 and Ni2P for improved migration and separation toward enhanced photocatalytic hydrogen evolution. Catal. Sci. Technol. 2016, 7, 641–649. DOI: 10.1039/C6CY02145J

ACS Paragon Plus Environment

22

Page 23 of 26

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

ACS Sustainable Chemistry & Engineering

(26) Karunadasa, H. I.; Montalvo, E.; Sun, Y. J.; Majda, M.; Long, J. R.; Chang, C. J. A molecular MoS₂ edge site mimic for catalytic hydrogen generation. Science 2012, 335, 698–702. DOI: 10.1126/science.1215868 (27) Kibsgaard, J.; Chen, Z.; Reinecke, B. N.; Jaramillo, T. F. Engineering the surface structure of MoS2 to preferentially expose active edge sites for electrocatalysis. Nat. Mater. 2012, 11, 963–969. DOI: 10.1038/nmat3439 (28) Xie, J. F.; Zhang, H.; Li, S.; Wang, R. X.; Sun, X.; Zhou, M.; Zhou, J. F.; Lou, X. W.; Xie, Y. Defect-rich MoS2 ultrathin nanosheets with additional active edge sites for enhanced electrocatalytic

hydrogen

evolution.

Adv.

Mater.

2013,

25,

5807–5813.

DOI:

10.1002/adma.201302685 (29) Kong, D. S.; Wang, H. T.; Cha, J. J.; Pasta, M.; Koski, K. J.; Yao, J.; Cui, Y. Synthesis of MoS2 and MoSe2 films with vertically aligned layers. Nano Lett. 2013, 13, 1341–1347. DOI: 10.1021/nl400258t (30) Jaramillo, T. F.; Jorgensen, K. P.; Bonde, J.; Nielsen, J. H.; Horch, S.; Chorkendorff, I. Identification of active edge sites for electrochemical H2 evolution from MoS2 nanocatalysts. Science 2007, 317, 100–102. DOI: 10.1126/science.1141483 (31) Hinnemann, P. G.; Moses, P. G.; Bonde, J.; Jorgensen, K. P.; Nielsen, J. H.; Horch, S.; Chorkendorff, I.; Norskov, J. K. Biomimetic hydrogen evolution:  MoS2 nanoparticles as catalyst for hydrogen evolution. J. Am. Chem. Soc. 2005, 127, 5308−5309. DOI: 10.1021/ja0504690 (32) He, J.; Wu, K.; Sa, R.; Li, Q.; Wei, Y. Magnetic properties of nonmetal atoms absorbed MoS2 monolayers. Appl. Phys. Lett. 2010, 96, 082504. DOI: 10.1063/1.3318254

ACS Paragon Plus Environment

23

ACS Sustainable Chemistry & Engineering

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

Page 24 of 26

(33) Yue, Q.; Chang, S.; Qin, S.; Li, J. Functionalization of monolayer MoS2 by substitutional doping: a first-principles study. Phys. Lett. A 2013, 377, 1362–1367. DOI: 10.1016/j.physleta.2013.03.034 (34) Tedstone, L. A.; Lewis, D. J.; Brien, P. O. Synthesis, properties, and applications of transition metal-doped layered transition metal dichalcogenides. Chem. Mater. 2016, 28, 1965−1974. DOI: 10.1021/acs.chemmater.6b00430 (35) Voiry, D.; Yamaguchi, H.; Li, J.; Silva, R.; Alves, D. C. B.; Fujita, T.; Chen, M.; Asefa, T.; Shenoy, V. B.; Chhowalla. G. M. Enhanced catalytic activity in strained chemically exfoliated WS2 nanosheets for hydrogen evolution. Nat. Mater. 2013, 12, 850–855. DOI: 10.1038/nmat3700 (36) Wang, T.; Liu, L.; Zhu, Z.; Papakonstantinou, P.; Hu, J.; Liu, H.; Li, M. Enhanced electrocatalytic activity for hydrogen evolution reaction from self-assembled monodispersed molybdenum sulfide nanoparticles on an Au electrode. Energy Environ. Sci. 2013, 6, 625– 633. DOI: 10.1039/C2EE23513G (37) Calandra, M. Chemically exfoliated single layer MoS2: Stability, lattice dynamics, and catalytic adsorption from first principles. Phys. Rev. B 2013, 88, 245428. DOI: 10.1103/PhysRevB.88.245428 (38) Kan, M.; Wang, J. Y.; Li, X. W.; Zhang, S. H.; Li, Y. W.; Kawazoe, Y.; Sun, Q.; Jena, P. Structures and phase transition of a MoS2 monolayer. J. Phys. Chem. C 2014, 118, 1515– 1522. DOI: 10.1021/jp4076355. (39) Liu, X.; Li, L.; Wei, Y.; Zheng, Y.; Xiao, Q.; Feng, B. Facile synthesis of boron- and nitride-doped MoS2 nanosheets as fluorescent probes for the ultrafast, sensitive, and labelfree detection of Hg2+. Analyst 2015, 140, 4654–4661. DOI: 10.1039/C5AN00641D

ACS Paragon Plus Environment

24

Page 25 of 26

ACS Sustainable Chemistry & Engineering

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 ACS Paragon Plus Environment

25

ACS Sustainable Chemistry & Engineering

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

Page 26 of 26

For Table of Contents Use Only Synopsis The earth abundant few layered boron doped MoS2 nanosheets decorated CdS nanorods (FBMC) by simple methods. The doping of foreign nonmetal into MoS2 system offers a strategy for the activation of basal plan surface as alternative approach of increasing the density of catalytically active sites.

ACS Paragon Plus Environment

26