<<

Titania - Supported transition metals as photocatalysts for hydrogen production from propan-2-ol and methanol C. Maheu, Eric Puzenat, Christophe Geantet, Luis Cardenas, Pavel Afanasiev

To cite this version:

C. Maheu, Eric Puzenat, Christophe Geantet, Luis Cardenas, Pavel Afanasiev. Titania - Supported transition metals sulfides as photocatalysts for hydrogen production from propan-2-ol and methanol. International Journal of Hydrogen Energy, Elsevier, 2019, ￿10.1016/j.ijhydene.2019.05.080￿. ￿hal- 02935057￿

HAL Id: hal-02935057 https://hal.archives-ouvertes.fr/hal-02935057 Submitted on 10 Sep 2020

HAL is a multi-disciplinary open access L’archive ouverte pluridisciplinaire HAL, est archive for the deposit and dissemination of sci- destinée au dépôt et à la diffusion de documents entific research documents, whether they are pub- scientifiques de niveau recherche, publiés ou non, lished or not. The documents may come from émanant des établissements d’enseignement et de teaching and research institutions in France or recherche français ou étrangers, des laboratoires abroad, or from public or private research centers. publics ou privés. Version of Record: https://www.sciencedirect.com/science/article/pii/S0360319919319366 Manuscript_e2d3289b89e0d9fb215783198a4bf081

1 Titania - supported transition metals sulfides as photocatalysts for hydrogen production 2 from propan-2-ol and methanol

3 Clement Maheu*, Eric Puzenat, Christophe Geantet, Luis Cardenas, Pavel Afanasiev*

4 Institut de Recherches sur la Catalyse et l’Environnement de Lyon IRCELYON, UMR 5256, 5 CNRS – Université Lyon 1, 2 av A. Einstein 69626 Villeurbanne Cedex (France); Fax: 33 04 6 7244 5399; Tel: 33 04 72 44 5466; 7 8 9 E-mail: [email protected] 10 [email protected] 11 12

13 Abstract

14 A series of transition metals sulfides deposited on anatase titania (MSx/TiO2) were prepared 15 by precipitation of transition metals salts with thioacetamide in aqueous medium under reflux. 16 The solids were characterized by XRD, XPS, temperature programmed reduction and 17 transmission electron microscopy. The properties of as obtained catalysts were compared for 18 the photocatalytic hydrogen evolution reaction (PHER) in pure methanol and water- 19 isopropanol mixture. The sequences of PHER activity were compared with electrochemical 20 HER and hydrodesulfurization (HDS) activity of the corresponding sulfides 21 prepared by the same technique. For PHER, in both alcohols the most active photocatalysts 22 contain hydrogenating sulfides of Co and Ru. However the PHER activity does not follow the 23 same trend as electrocatalytic HER and thiophene HDS. Some sulfides, such as HgS or CuS, 24 show poor activity in HDS and electrocatalytic HER, but have the PHER activity comparable 25 with that of the best samples. This difference suggests that the PHER rate is not merely 26 related to the hydrogen activating properties of the co-catalyst, but is enhanced by the transfer 27 of photogenerated electrons from TiO2 towards the . The ranking of the co-catalysts 28 and the PHER activity depend also on the nature of the alcohol molecule, the overall PHER 29 rates in water-isopropanol mixture being lower than in methanol.

30

31 Keywords:

32 Hydrogen evolution reaction; transition metal sulfides; photocatalysts; titanium oxide

33

1 © 2019 published by Elsevier. This manuscript is made available under the Elsevier user license https://www.elsevier.com/open-access/userlicense/1.0/ 1 1. Introduction

2 Direct conversion of solar energy to chemical energy of hydrogen is considered to be a very 3 promising strategy to mitigate the global warming and to address the problem of alternative

4 energy sources [1,2]. Titanium dioxide TiO 2 is by far the most studied semiconductor for the 5 artificial photosynthesis design [3,4]. Versatile aspects of surface chemistry and bulk 6 properties are important for photocatalytic hydrogen evolution reaction (PHER) using titania, 7 such as surface hydroxylation, type of exposed crystallographic facets [5], or carrier dynamics 8 [6,7]. Intense current research efforts are directed to the design of efficient co-catalysts,

9 present on the surface of TiO 2 in order to enhance the utilization of photogenerated electrons 10 and holes in redox processes involved in the targeted reactions. PHER catalysts are widely 11 studied that include proton reduction by photogenerated electrons and oxidation of organic 12 molecules by photogenerated holes. Association of titania with electron accepting and/or 13 hydrogenating materials, such as noble metals, proves favorable for PHER, one of the best

14 benchmark reference PHER catalyst being Pt/TiO 2. The relationships between the PHER 15 performance and the properties of co-catalysts are widely studied but not fully understood. 16 The improvement of PHER performance in composite catalysts vs. bare titania was attributed 17 to increased charge separation [8], plasmonic effects and increased rate of hydrogen atoms 18 recombination [9]. The impact of a co-catalyst is complex as it might simultaneously 19 influence several thermodynamic or kinetic parameters in the system, including the lifetimes 20 of charge carriers or adsorbed intermediates, which are usually unknown. Beside the efforts of 21 fundamental understanding, great number of research works report on the empirical studies of 22 novel catalysts combining titania and nanoparticles of various co-catalysts, aiming to enhance 23 PHER performance. Much attention has been recently paid to the studies of layered sulfides

24 on titania, particularly to the MoS 2/TiO 2 composites [10], which has been recently reviewed

25 [11]. MoS2 is considered as a cheap alternative to platinum, both for electrochemical HER

26 and PHER. However the specific PHER rates even for the best MoS 2/TiO 2 systems are still

27 lover than for Pt/TiO 2.

28 Other sulfides were also investigated as co-catalysts with TiO 2, though less extensively

29 studied than MoS 2. Tungsten sulfide WS 2 demonstrated PHER performance comparable to

30 MoS 2 [12,13]. Copper sulfide on Aeroxide® P25 titania (0.5-15% wt. CuS) has been prepared 31 hydrothermally using and showed good activity for methanol reforming, the best

32 catalyst containing 1wt%. CuS [14]. Core-shell composite of NiS and anatase TiO 2 was

2

1 applied for PHER using methanol as an organic substrate and demonstrate promising activity

2 [15]. quantum dots on TiO 2 nanoparticles prepared with a precipitation- 3 deposition method were tested using ethanol as a sacrificial reagent and showed PHER rate

4 exceeding that of the pure TiO 2 by more than 35 times [16]. CdS had been 5 widely studied for PHER, but rather as a semiconductor than as a co-catalyst [17]. Moreover 6 CdS as a semiconductor could be promoted by other TM sulfide and non-sulfide co-catalysts 7 [18,19]. However CdS value for photocatalytic applications is limited because of its 8 solubilisation by photocorrosion, releasing toxic cadmium ions. Composite materials

9 containing bismuth [20], zinc [21] and indium [22] sulfides in combination with TiO 2 showed 10 promising HER activities, as well as bimetallic and trimetallic sulfide materials as CuS/ZnS

11 [23] or ZnS–In 2S3–CuS [24]. Multi-metal titania–based composite materials have been 12 applied as photocatalysts, containing up to four metals chalcogenides such as, for instance,

13 ZnS/CdS-Mn/MoS 2/CdTe/TiO 2 [25]. Comparison of PHER efficiency measured in different 14 laboratories is always difficult because the conditions of the experiments (preparation 15 methods, reactor geometry, light source, oxidized molecule, support...) are strongly varied.

16 Despite a great number of publications on the sulfide-containing photocatalysts, systematic 17 comparison of PHER performance for a series of sulfides on titania prepared by a uniform 18 technique have never been reported. The present work aims to study an extended series of 19 sulfides prepared by means of thioacetamide precipitation technique, in an attempt to provide 20 a qualitative insight into the key properties defining their relative PHER performance.

21

22 2. Experimental

23 2.1 Preparation of catalysts

24 The series of MS x/TiO 2 supported catalysts (M= Ag, Co, Cu, Hg, Mo, Ni, Ru) were prepared 25 with the aqueous deposition method under the conditions similar to those reported by Girel et 26 al . [26] and with the same molar amount of co-catalyst metals (1.3 at % of metal). Typically, TM 27 2.0 g of commercial TiO 2 (CristalACTiV PC500) and 900 mg of thioacetamide were 28 suspended in 100 mL of deionized water. For each synthesis, transition metal precursor salt 29 was added, in an amount containing 350 µmol equivalent of metal. Then, aqueous suspension 30 was refluxed under stirring for 1 h. In the synthesis using ammonium heptamolybdate, 2.0 ml 31 of 2.5 M HCl was added to accelerate precipitation. The solid materials were isolated by

32 centrifugation, washed and dried under N 2.

3

1 Bulk transition metal sulfides were prepared in the same conditions as titania-deposited

2 catalysts, but with a tenfold increase of the precursor amount and without addition of TiO 2. 3 They were used as reference samples for characterization and phase identification of sulfide 4 phases as well as for comparison of the electrochemical HER activity. A reference catalyst

5 containing 1 wt.% Pt on P25 TiO 2 was prepared by impregnation, drying and reduction with

6 H2, as described earlier [27]. 7 8 2.2. Characterizations

9 Temperature-programmed reduction (TPR) was carried out in a quartz reactor. The samples of −1 10 sulfides (0.05–0.1 g) were heated under H 2 flow (50 ml min ) from room temperature to −1 11 1050 °C at a rate 5° min . The H 2S produced in the reduction reaction was detected by a

12 Thermo Prolab quadrupole mass-spectrometer. The amount of H 2S released from the solid

13 was quantified after calibration of the quadrupole detector with a known H 2S content gas 14 mixture. Transmission electron microscopy (TEM) was carried out on a JEOL 2010 device 15 with an accelerating voltage 200 keV. The samples were dispersed in ethanol by ultrasound, 16 and then put onto a lacey carbon on a copper grid. In order to protect them from oxidation by 17 air, the samples still covered with liquid hexane were immediately introduced into the TEM 18 vacuum chamber. The X-ray diffraction (XRD) patterns were obtained on a Bruker D8 19 Advance diffractometer with Cu-Kα emission and identified using standard JCPDS files. The 20 metal content in the solids was determined by plasma-coupled atomic emission spectroscopy

21 (ICP-OES Activa Jobin Yvon) after dissolution in a HNO 3/H 2SO 4 mixture. Elemental analysis 22 of light elements (CHNS) was performed on an analyzer Thermo Fisher Flash 2000. X-ray 23 photoelectron spectroscopy (XPS) measurements were performed using a commercial Axis

24 Ultra DLD spectrometer (Kratos Analytical). A monochromatic Al-Kα X-ray source at 25 1486.6 eV was used and a flood electron gun to minimize surface charging. Energy 26 calibration was made fixing both the binding energy of the C 1s intense peak assigned to C-C

27 or C-H bonds at 284.5 eV and the binding energy of the Ti 2p 3/2 state at 458.5 eV. Data 28 analysis was performed using CasaXPS, after a Shirley background subtraction. The C 1s and 29 S 2p signals were decomposed into a combination of Voigt functions. For Ru 3d an 30 asymmetric lineshape was used: product of a Doniach Sunjic with a Gaussian/Lorentzian 31 function. UV-Vis diffuse reflectance was measured at room temperature with an optical fiber 32 set-up. Six optical fibers illuminate the powder sample with a Deuterium-Halogen light source 33 (AvaLight-DHS, Avantes). A seventh fiber collects the diffused light which is then analyzed

4

1 by a CCD dectector. Pure BaSO 4 was used as a reference and to dilute TiO 2-based powders. 2 Kubleka-Munk (KM) function was applied to interpret spectra. Note that KM theory is valid

3 only for thick samples (here 2 mm) with weak absorption ( > 0.6). To do so we diluted the

4 samples in BaSO 4 (0.1 wt.% MS x/TiO 2 in BaSO 4) [28].

5 6 2.3 Photocatalytic tests 7 Photocatalytic reaction was conducted in liquid phase, in a double-wall semi-batch slurry 8 reactor. Most of the catalytic tests involved 50 mg of catalyst dispersed either in pure 9 methanol or in 50 mL of 50 vol. % mixture of propan-2-ol (IPA) and distilled water. Before 10 irradiation, the reactor was purged under an argon flow for 30 min to conduct the test under 11 anaerobic conditions. Gases were analyzed by micro gas chromatography with thermal 12 conductivity detectors (Agilent Technologies, 300 A). It quantified several gases including 13 oxygen, hydrogen, hydrogen sulphide, propene and propane. Photoreactor temperature was 14 regulated at 20°C by thermostated water recirculation in the double-wall.

15 A 125 W high-pressure lamp (Philips HPK 125) was used to illuminate the reactor 16 through a 20 cm 2 area optical window. The reactor Pyrex glassware cut light below 290 nm 17 wavelength. The number of incident photons, likely to be absorbed by anatase, was measured 18 between 290 and 390 nm with a radiometer device, a spectrometer (AvaSpec-2018, Avantes) 19 coupled with one fiber optic probe (AvaSpec, FCR-7UV-400-2-ME) equipped with a cosinus 20 corrector. The intensity value was 1370 µE/h for the screening test in water-IPA mixture and 21 1550 µE/h for the methanol PHER test. Irradiance spectrum is provided in the Supplementary 22 Information (SI), Fig. S1. The photonic efficiency (PE) has been calculated as hydrogen 23 production rate divided by this photonic intensity, following IUPAC recommendations [29]. 24 Because of the uncertainty on the amount of hydrogen produced and on the incident amount 25 of photons determination, the error on the PE was estimated to be close to 10%. As the PE 26 depends on the slurry concentration, reference curve of normalized PHER rate as a function of 27 catalyst mass (slurry concentration) was measured (Fig. S2). This curve was applied to 28 compare PHER rates measure in this work with literature data.

29

30 2.4. Electrochemical measurements

31 To prepare the electrodes, 5 mg of a solid were dispersed by ultrasound in a solvent 32 containing 800 µl of ethanol and 400 µl of 0.5% Nafion solution, optionally containing 1 mg

5

1 of acetylene black carbon. Then, 10 µl of suspension was dropped onto rotating glassy carbon 2 electrode (0.0706 cm 2) and then dried in argon. The catalyst loading was 0.65 mg/cm 2. The 3 electrochemical measurements were performed on a Voltalab three-electrode potentiostat 4 using a Pt counter electrode and SCE reference electrode at 25 °C in purged solutions of

5 0.1 M H 2SO 4, by bubbling argon. Cyclic voltammograms were obtained in the potential range 6 from 0.5 to −0.1 V vs. SCE at scan rates ranging from 20 to 200 mV s−1 . Linear Sweep 7 Voltammetry (LSV) scans were carried out at 2 mV/s rate.

8

9 2.5. Thiophene hydrodesulfurization (HDS).

10 Catalytic activity in thiophene HDS (Scheme 1) was measured at atmospheric pressure in a 11 fixed-bed flow microreactor. Prior to catalytic tests, all samples were re-activated under the

12 standard conditions by means of treatment under a flow of 10% H 2S/H 2 at 350°C, for 2h.

HYD S

S DDS 13

14 Scheme 1. Thiophene HDS pathways

15 Reaction was carried out in the temperature range 280–340 °C, under 50 ml/min gas flow, 16 using 100–200 mg of catalyst and partial pressure of thiophene 2.7 kPa. The plug-flow reactor 17 model was used to calculate the specific reaction rate, Vs, according to the equation

18 Vs = − (F/m)ln(1 − x)

19 where F is thiophene molar flow (mol/s), m is the catalyst mass (g), and x is thiophene 20 conversion. Catalytic activity was estimated at steady state conversion, after at least 16 h on- 21 stream.

22

23 3. Results and discussion

6

1 3.1. Synthesis and characterizations of solid materials (TEM, XRD, TPR)

2 The primary goal in this work was to study the PHER activity for a series of titania-supported 3 sulfides obtained with the same preparation technique, as a function of the nature of transition 4 metal (TM) sulfide. First, the composition and the morphology of sulfides deposited with this 5 technique were investigated. To this aim, the samples of supported catalysts and/or the 6 corresponding reference bulk sulfides have been characterized using elemental analysis, 7 TEM, XRD, XPS and TPR.

8 The set of available sulfides is limited by the possibilities of the preparation technique. 9 Stability of TM sulfides vs. the corresponding oxides increases from the left to the right of the 10 periodic table. Sulfides of oxophilic early TM (Ti, Nb or Zr) could not be precipitated using 11 aqueous thioacetamide reaction. We studied the reactions of 3d metals from Mn to Zn. 12 Additionally we tried several 4d and 5d TMs (Mo, W, Ru, Pt, Ag, Hg).

13 Thioacetamide has been widely used as source for synthetic and analytic purposes. In 14 various conditions, reaction of metals precursors with thioacetamide has been applied to 15 prepare sulfides of Mo [30], Co and Ni [31], Cd [32], Sn [33] and other metals. Here we use 16 uniform aqueous reflux synthesis conditions. Already by visual observation we could identify 17 the cases where TM sulfides were probably not formed. An inspection of the color change 18 after the reaction suggested that for Mn, Fe, W and Pt the corresponding (yellow or dark 19 brown) sulfides were not formed, but the reactants probably remained in the oxide form. 20 Platinum-containing product remained yellow as is the parent hexachloroplatinate. Zinc 21 nitrate if put into the reaction with thioacetamide does not give any precipitate under the 22 conditions applied (though ZnS could be prepared using thioacetamide in different conditions, 23 [34]). Therefore, W, Pt, Mn, Fe and Zn- containing solids were discarded. From twelve metals 24 precursors initially tried, seven sulfides were obtained and retained for further 25 characterizations and catalytic tests. The TM sulfides for which the thioacetamide technique 26 failed could be prepared using different methods, such as hydrothermal or microwave 27 assisted, but in this work we follow a uniform preparation protocol.

28 Chemical analysis of the solids after reaction shows quantitative deposition of the transition 29 metals species onto titania (Table S1). For four metals, the chemical form of supported 30 species might be inferred from the XRD of the reaction products of individual precursors.

31 Indeed, Ag 2S, CuS, NiS 2 and HgS crystalline sulfides are formed upon reflux (Fig. 1-4). Other 32 three sulfides (Mo, Ru, Co) form sulfur-rich amorphous precipitates with no crystalline

7

1 phases detectable by XRD. Secondary reaction of slow oxidation of thioacetamide upon

2 reflux leads to the formation of crystalline sulfur impurities in CoS x, MoS x and RuS x (Fig. 5).

3 As discussed in [26], molybdenum precipitates as amorphous MoS x phase with composition

4 close to MoS 3. Ruthenium and cobalt give dark XRD-amorphous precipitates but due to 5 considerable sulfur impurity the exact composition of amorphous metal sulfides is unclear.

(120) (022) (-113) (-122)

(111) (013) (102)

AgS x (-112) β− Ag 2S: PDF 01-080-5476

(-102) & (110) (-102) & (-104) (-210) (-121) (031)

(-202) (-114) & (122) (211) & (211)& (220) (012) (100) (133) (210) (-104) (-111)

10 20 30 40 50 60 70 80 6 angle (2 θ)

7 Figure 1. XRD pattern of bulk AgSx (black) and the matching pattern of β-Ag 2S (PDF 01- 8 080-5476) (blue). (103) (110) (102) (006) CuS x CuS: PDF 00-006-0464 (101) (116) (108) (100) (002) (105) (1011) (208) (213)

10 20 30 40 50 60 70 80 θ 9 angle (2 ) 10 Figure 2: XRD pattern of bulk CuSx (black) and the matching pattern of CuS (PDF 00-006- 11 0464) (blue).

8

(200)

NiS x

NiS 2: PDF 04-003-1992 S: PDF 04-012-7311 (111) (311) (210) (211) (220) (222) (026) (023) (321) (113) (222) (313) (331)

10 20 30 40 50 60 70 80 angle (2 θ) 1

2 Figure3: XRD pattern of bulk NiS x (black) and the matching pattern of NiS 2 (PDF 04-003- 3 1992) (blue); elemental orthorhombic sulfur (04-012-7311) is also observed. 4 (101)

HgS x

(012) HgS: PDF 04-002-6787 (111) (113) (110) (202) & (015) &(202) (021) (122) & (205) &(122) (003) (114) & (203) &(114) (104) (211) (100) (103) (116) (024) (006)

10 20 30 40 50 60 70 80 θ 5 angle (2 )

6 Figure 4. XRD pattern of bulk HgS x (black) and the matching pattern of HgS (PDF 04-002- 7 6787) (blue).

CoS x

RuS x

MoS x

0 10 20 30 40 50 60 70 80 angle (2 θ) 8

9

1 Figure 5. XRD patterns of CoS x, RuS x and MoS x amorphous phases. Narrow peaks 2 correspond to the common impurity of elemental rhombohedral sulfur (00-024-1206) and

3 elemental monoclinic sulfure (04-007-3019) only observed for CoS x.

4

5 Fig. 6 depicts TEM images of the supported sulfide catalysts under study. TEM reveals that 6 sulfides dispersion varies depending on its nature. For Ag and Ni, 100 nm - range particles of

7 Ag 2S and NiS 2 were observed, so the dispersion is rather poor. Hg and Mo give relatively

8 well-dispersed 10 nm - range particles. MoS x amorphous phase is fragile and rapidly

9 transforms under the electron beam into the MoS 2 fringes. For the supported Co and Cu 10 sulfides, we could not identify the sulfide species by means of HREM and STEM. The TM 11 species are ultradispersed, of subnanometer size and are homogeneously smeared over the 12 titania, as attested by the EDS scans. For both cobalt and copper (Cu sample is shown in Fig. 13 S3) only titania grains were observed. Similarly, ruthenium sulfide while detected by EDS 14 forms highly dispersed sulfide species not detectable by HREM (Fig. S4). However, due to 15 the relatively high mass of Ru, STEM contrast was sufficient to detect 1-1.5 nm size clusters

16 of RuS x (Fig. 6).

17 Additional characterization has been carried out using temperature programmed reduction 18 (TPR), in which we detected released upon linear heating of the samples up 19 to 1050°C (Fig. S5-S11). TPR study allows correlating the amount of sulfur released and the 20 temperature of TPR peaks with the chemical nature of sulfide species. In particular the TPR 21 characterization is important to give evidence of the sulfides formation in the case of 22 amorphous species, not detectable by TEM.

23 For the metals that form large sulfide particles (Ag, Ni), TPR of supported catalysts gives the 24 main reduction peaks in the same temperature range as for the corresponding bulk sulfides 25 (Fig. S5, S6). The shape of the ascending part in the high temperature TPR curve region is 26 related to the sulfide formation enthalpy, and therefore is specific for the corresponding metal 27 sulfide phase, as discussed earlier [35]. Beside the sulfide core reduction, additional peaks are 28 observed at low temperatures due to the weakly bonded sulfur species located at the surface of

29 TiO 2 and/or at the surface of sulfide particles. The higher is sulfide dispersion, the higher is 30 the area of low-temperature peak. For highly dispersed supported species TPR proves the 31 presence of sulfides not detectable by TEM (Cu, Co, Fig. S7, S8). Beside the low-temperature 32 peaks, TPR attests formation of transition metal sulfides due to systematic presence of

10

1 relatively high temperature feature, related to the core reduction (Fig. S8, Fig. S9). Thus, 2 supported ruthenium sulfide TPR shows a large pic of weakly bonded sulfur at low 3 temperature but also the high temperature core reduction peak, which attests the presence of

4 RuS 2 particles, having size of the order of 1 nm (Fig. S9).

5 The presence of RuS x clusters (x ≈2) was also confirmed by XPS measurement, as shown in

6 Fig. S12. Ru 3d, Ru 3p and S 2p signals were analyzed for supported and bulk RuS x. The Ru 7 3p and Ti 2p signals overlap, which makes the interpretation difficult. The state at 279.9 eV 8 (Fig. S12 C) can be either metallic Ru or its sulfide [36,37]. A second state, probably an 9 oxide, is also observed in low surface concentration. The contribution located at 161.9 eV 2- 10 (Fig. S12 D) is attributable to type sulfur (S-S) . It confirms that RuS x is deposited at 11 the surface of TiO 2. The molar ratio / Ru was found close to 1.4 for the supported 12 sulfides. We wanted to study sulfides-based material in realistic conditions, as it is present in 13 the catalytic tests. Therefore, no particular precaution was taken to load the photoreactor 14 without air exposure. It explains why some sulfate is observed by XPS (167 eV, Fig. S12 B 15 and D) [38].

16 Finally, BET surface areas of the supported catalysts are high and remain similar to that of 17 bare titania (350 m 2/g) and are all greater than 300 m 2/g. For example specific surface area is 2 2 18 of 321 m /g for MoS x/TiO 2 and 328 m /g for CoS x/TiO 2. The morphology of supported

19 catalysts is dominated by TiO 2 support because the TM sulfides loadings are relatively small,

20 whereas no thermal treatments were involved during preparation, which could lead to TiO 2 21 sintering.

22 Overall, for seven metals under study, sulfide species deposited on the PC500 anatase TiO 2 23 were prepared by thioacetamide method. However the crystallinity and the dispersion of the 24 sulfides as obtained are different.

25

26 3.2. Photocatalytic HER activity

27 The two studied reactions are:

28 - CH 3OH ‰ HCHO + H 2 (1)

29 - CH 3CHOHCH 3 ‰ CH 3COCH 3 + H 2 (2)

11

1 The initial values of PHER activity were compared after several hours of test, in order to 2 characterize the intrinsic ability of TM sulfides to co-catalyze the reactions (1, 2). In all cases, 3 a short induction period (15-20 min) was observed. Then, for the duration of the experiment 4 (3-6 hours) the PHER rate was stable.

5 The PHER rate evolution was also checked on a longer time scale. We studied stability for 6 several days on-stream for the semi-continuous flow IPA test, or for several repetitive runs for 7 the batch methanol test (in the last case - with refreshing the methanol and carrying out purge 8 with argon between the runs). In the water-IPA test we measured PHER rate at the 9 temperatures stepwisely changed between 10 and 45 °C, in order to determine the apparent 10 activation energy (which is beyond the scope of the present work and will be discussed 11 elsewhere).

12 After the initial stabilization period, during several hours on-stream in isothermal conditions 13 the relative variations of PHER rate were less than 5%; an apparent steady state was achieved. 14 However the PHER rates slowly evolved on the several-days scale for the continuous flow 15 test (Fig. S13, S14). In most cases (Co, Cu, Hg, Mo, Ni, Ru) slow activation occurred,

16 probably due to the cleaning of TiO 2 surface from the residuals of thioacetamide used in the

17 preparation. In the case of Ag 2S slow deactivation occurs, which might be attributed to photo 18 corrosion. The steady state rate measured at 20 °C in the first or in the second temperature 19 cycle was taken to compare the PHER activity (Fig. S13). Batch PHER test in the methanol 20 medium provides good stability for several (4-5) repetitive 3h runs (Fig. S15). In bothcases, 21 the variations of activity with time are lesser than the differences between various TM 22 sulfides and do not influence the conclusions of this work.

23 Post-mortem XPS analysis of the solids after water-IPA tests showed that both metal species 24 and sulfur species are partially transformed to the higher oxidation states (Fig. S16). In 25 agreement with the PHER test results, the most important degree of corrosion was observed

26 for Ag 2S/TiO 2 sample. After 70 hours of test, surface layers were tatally transformed to 27 sulfate. Note that the stability study was not in the focus of this work, because we are 28 interested in the understanding of titania combined with fresh sulfides.

29 The PHER photon efficiency measured in pure methanol and water-IPA 50/50 vol. mixture is 30 presented in Fig. 7. The absolute and mass - specific PHER rates are given in Table S2. For 31 the water–IPA mixture, the performance changes in the sequence Ru > Ni > Co > Hg > Ag> 32 Mo > Cu. In methanol the sequence is Co > Ni > Ru = Mo = Cu > Ag > Hg. In methanol even

12

1 the least active HgS/TiO 2 produced 5 times more hydrogen than bare TiO 2. In IPA the PHER

2 activity of the least active CuS/TiO 2 was 12 times higher than of bare titania which undergone 3 the thioacetamide reflux procedure.

4 The photocatalytic properties of HgS, Ag 2S and RuS x supported on TiO 2 are reported for the 5 first time. Moreover, no transition metal sulfides-based photocatalysts were studied 6 previously for IPA dehydrogenation. Therefore no literature references exist for these cases. 7 In methanol, the PHER performance of our most active catalysts (Co and Ru sulfides) is at the 8 level of the best sulfide catalysts reported in the literature (Table S2). As far as comparison is 9 possible, our absolute PHER values are lower than the literature values for copper and nickel

10 sulfides [14,39]. Our values for molybdenum sulfide are lower than for amorphous MoS x

11 reported in [10] but higher than observed earlier for the MoS 2 nanosheets [40]. However if 12 approximately re-calculated for the same conditions (lamp power and slurry concentration, 13 Table S2, Fig. S2), the best values reported in the literature for Co, Cu and Mo sulfides are in 14 a remarkable agreement with our data (Table S2). Finally, our values measured in methanol

15 for cobalt and nickel co-catalysts are higher than observed in ethanol for CoS x and in lactic 16 acid for NiS, but since the organic molecules are not the same, such comparison makes lesser 17 sense [39,40,16,41].

18 Comparison with the existing literature is only approximate as the conditions applied from 19 one study to another are never the same and recommendations for reporting the activities 20 [42,43] are not always followed. Hence a more reliable activity assessment might be achieved

21 by comparison with a well-known Pt/TiO 2 benchmark reference (Table S2). The overall 22 activity in methanol is much higher than in water-IPA, showing the importance of organic 23 substrate for this type of studies. In methanol, the best cobalt sulfide promoted catalyst shows

24 approximately 30 % PHER rate of the 1% Pt/TiO 2 reference (Table S2). In water-IPA the

25 relative activity of TM sulfides attains, in the best case of RuSx, only 16 % of activity of the

26 Pt/TiO 2 reference. The observed PHER activity changes depending on the alcohol, in

27 agreement with previous works [44]. With metal supported on TiO 2 the activity is often 28 higher in methanol than in IPA [45,46].

29 The observed PHER trends show both expected and unexplained features. The best catalysts 30 in methanol and water-IPA media are, respectively, Co and Ru sulfides, which are known as

31 active hydrogen production catalysts [47,48]. On the other hand, MoS x, which is by far the 32 most studied sulfide for PHER, is not among the best co-catalysts in both media. We closely

13

1 reproduce the PHER activity in methanol of MoS x/TiO 2, observed in the previous work, which

2 used similar thioacetamide preparation [26]. However MoS x/TiO 2 activity in methanol is

3 almost the same as for CuS/TiO 2, which is not among the front-running PHER catalysts.

4 Moreover, MoS x/TiO 2 shows poor performance in water-IPA, where it appears less active

5 than HgS/TiO 2. Surprisingly, HgS is a reasonably good co-catalyst for PHER,

6 To understand the observed trends of PHER, various properties of the systems under study 7 were compared such as the morphology, the electrocatalytic HER activity and their thiophene

8 Hydrodesulfurization (HDS) activity. Deposition of sulfides at the surface of TiO 2 does not 9 alter the light absorption properties, important for PHER. The observed color change is due to 10 light absorption by sulfides in visible region, which does not interfere with electron-hole 11 generation in titania. Indeed, Kubelka-Munk transforms of UV-vis spectra (Fig. S17) are

12 similar between 5.5 and 3.2 eV. The band gap (E g) was determined as the intersection

13 between the x-axis in the Tauc plots. The Eg for seven samples under study is included within

14 the region between 3.2 and 3.35 eV. The Eg of seven catalysts and pure PC500 titania is the 15 same within the accuracy of the experimental method and corresponds to the literature value 16 for PC500 anatase [49]. Therefore the amount of (useful) photons absorbed by the samples is 17 similar, so other properties are responsible for the differences of observed PHER activity.

18 The dispersion degree of the co-catalyst species seems to be less important than the nature of

19 the TM sulfide. Indeed, two most active co-catalysts in the methanol medium are CoS x and

20 NiS 2. However, CoS x species are too finely dispersed to be observed by conventional TEM,

21 whereas the NiS 2 particles are large (near 100 nm). On the other hand, poorly dispersed Ag 2S

22 shows better activity in water-IPA than well - dispersed MoS x. We already noted in the 23 previous works that for PHER in methanol, the dispersion of sulfide co-catalyst on titania 24 does not seem to be a crucial parameter [26].

25 We are compelled to conclude that the differences of PHER rates are related to the intrinsic 26 properties of sulfides deposited onto titania. The property of a co-catalyst obviously relevant 27 to the PHER performance should be its ability to reversibly activate hydrogen. This ability is 28 revealed in the heterogeneous catalysis or in the electrochemical HER. Seven TM sulfides 29 considered here are known to possess very different ability to activate hydrogen and to 30 catalyze hydrogenation or hydrogenolysis of organic substrates. Ru and Mo sulfides are well- 31 known highly active hydrogenating catalysts [36,38,50]. Co and Ni are used as promoters in 32 sulfide heterogeneous catalysts but the activity of Co and Ni binary sulfides is low, though

14

1 non-negligible. Finally, CuS, HgS and Ag 2S have never been shown to act as hydrogen 2 activators or hydrogenating heterogeneous catalysts. To check in what extent the hydrogen 3 activation ability of TM sulfides is important, we compare the trends in HDS and in the 4 electrochemical HER with the observed PHER trend. The corresponding literature data are 5 dispersed; the preparation techniques and the reaction conditions are very different from one 6 work to another. To provide a reliable series of activity, we measured the electrochemical 7 HER performances and HDS activity for the catalysts studied above in PHER, as follows.

8

9 Figure 6 TEM and STEM HAADF images of selected solids. The transition metals forming co- 10 catalyst sulfides are indicated on the images. White spots in the STEM images correspond to the 11 TM sulfides particles.

15

4 isopropanol - water methanol

3

2

1 Photon Efficiency (%) Efficiency Photon

0 1 CuS MoSx Ag2S HgS CoSx NiS2 RuS2

2 Figure 7. PHER activity of titania-deposited sulfides in methanol and in water-IPA mixture.

3

4 3.3. Electrocatalytic HER and thiophene HDS activity

5 PHER via alcohols dehydrogenation certainly follows a complex mechanism. To get a better 6 insight, we used model HER and HDS reactions, in order to disentangle the intrinsic 7 hydrogen-activating properties of TM sulfides from their influence on the titania 8 semiconductor. HER and thiophene HDS reactions include respectively the H-H bond 9 formation and breaking on the sulfide surface. These properties are correlated, because of the 10 microreversibility principle. So HDS and HER both estimate the ability of a catalyst to

11 activate H2 and as such they mimic the last PHER step, where molecular H2 is formed . 12 However, these reactions occur under different conditions and may involve different species. 13 Thus, HER proceeds with a participation of hydrated protons, whereas HDS obviously cannot 14 involve such species.

15 TM sulfides and in particular MoS 2 recently emerged as HER electro catalysts with high 16 activity and good stability in acidic medium [51] . The HER performance of sulfides strongly 17 depends on the nature of TM metal and on the preparation technique. To understand how the 18 nature of TM sulfides obtained by thioacetamide technique impacts its electrochemical HER 19 performance, we compared the initial activities of seven sulfides as prepared. Such initial 20 HER activities are provided by steady–state characteristics which are achieved after several 21 LSV cycles. Since pure titania is inactive, the electrocatalytic properties of diluted sulfides 22 species supported on inert PC500 are difficult to measure. By this reason we proceeded with

16

1 the HER measurements for bulk TM sulfides obtained from the thioacetamide reaction. Here

2 we assume that similar TM sulfide species are formed with and without TiO 2 and they have 3 similar electrochemical properties.

4 The Tafel plots and LSV curves are presented in Fig 8a, b. The numeric values of Tafel slopes

5 and η10 are summarized in Table 1. Low Tafel slope and low values of η10 (overpotential 6 necessary to attain the 10 mA/cm 2 current) are characteristics of high HER performance. The 7 overall ranking of sulfides HER performance as a function of metal is Mo > Ru ≈ Ag > Cu > 8 Co > Ni > Hg. Worth emphasizing that we do not claim this ranking to be a general trend, 9 transposable to any sulfides of these metals, but merely a particular sequence for the materials 10 prepared using thioacetamide method.

11 The specific rate of thiophene HDS reaction (Scheme 1) could be considered as another 12 descriptor of hydrogen-activation and hydrogenating activity. With the exception of Ag, we 13 see that highly hydrogenating sulfides (Ru, Mo) demonstrate the highest HER activity and 14 overall correlation between the electrocatalytic HER and HDS performances is good (Table 15 1). The activities per mole of sulfided metal of Cu, Co and Ni sulfides are usually found to be 16 low, whereas the activities of Ru and Mo sulfides are high. This is true for both hydrogenation 17 and HDS reactions [52] and for binary [53] or ternary sulfides including these elements [54] .

18 Ag 2S was recently studied as HER electrocatalyst and shown moderate but non- 19 negligible performance [55] . However, its HDS activity appeared negligible, probably 20 because relatively harsh HDS conditions modify the initial sulfide. Finally, mercury sulfide 21 HgS was never studied in HDS/HER and here we attest its zero HDS activity as well as poor 22 electrochemical performance.

23 Confronting the results of photocatalytic HER, electrocatalytic HER and thiophene HDS, we 24 see that if the two latter activities are well correlated, there is no clear relationship between 25 any of them and observed PHER. Therefore, photocatalytic performance depends more 26 strongly on other parameters influenced by the co-catalysts (charges separation, lifetimes etc.) 27 than by their hydrogen-activating ability.

17

-100 0 HgS (a) RuS x 124 NiS -200 MoS 87 -4 2 (b) x CoS x ) -300 2 -8 CuS

CoS x 170 Ag 2S 127 RuS Ag 2S x -400 -12 J (mA/cm J

MoS x CuS 134 -500 HgS 209 -16

Overmpotential vs. (mV) vs. RHEOvermpotential NiS 2 240 -600 -20 -0,8 -0,3 0,2 -600 -400 -200 0 Log J (mA/cm 2) 1 Overmpotential vs. RHE (mV)

2 Figure 8. Electrochemical HER activity of TM sulfides: Tafel plots (a) and LSV curves (b).

3 Table 1 Overall performance in the PHER, thiophene HDS and electrochemical HER.

property\sulfide Ag 2S CoS x CuS HgS MoS x NiS 2 RuS 2 Thiophene HDS a 0.1 2.1 0.3 0.0 16 0.5 91 Electro HER Tafel b 127 170 134 209 87 240 124 c f f f Electro HER η10 460 670 580 910 433 770 445 Photo HER, methanol d 1.50 4.25 2.50 0.50 2.45 4.00 2.70 e Photo HER, IPA-H2O 0.30 0.40 0.01 0.35 0.10 0.50 1.90 4 a reaction rate,10 -8 mol.g -1.s -1 measured at 300 °C (per gram of metal sulfide); b mV/decade; c 5 -mV; d,e photon efficiency, %; f extrapolation.

6 3.4. Mechanisctic considerations

7 The observed HDS and HER trends have many common points, which could be expected. In 8 the most general case, the overall (HER, HDS) performance of a sulfide material is a product 9 of two terms. The first term is the intrinsic chemical performance and the second is 10 morphology –related term describing the availability of a sulfide surface to the reactants. 11 Beside these factors, the electric conductivity of sulfide plays an important role for the 12 electrochemical HER, though in some extent it is leveled off by applying conductor binders 13 (usually Nafion and carbon). While morphology of sulfides can be extremely versatile 14 depending on the preparation ways, the intrinsic chemical trends are relatively well 15 understood on the basis of generalized Sabatier principle [56–58].The energies of S - vacancy 16 creation and those of hydrogen dissociative adsorption and desorption govern both the HDS 17 and HER trends. Thus, M-S bond energy and hydrogen adsorption energy are recognized to 18 be useful descriptors for HDS activity. So-called volcano plots allow predicting which

19 sulfides could be potentially active HDS (HER) catalysts [59,60].

18

1 With this respect Mo and Ru sulfides have the M-S bond energies are close to the optimal 0 2 value [60], whereas Ag 2S and HgS have it too weak (∆Hf values are −31.8 and −58.2. 3 KJ/mol , respectively [61]. The observed HDS trend is therefore in agreement with the 4 existing volcano plots.

5 The HDS and HER properties are probably related to the electronic band structures of the 6 corresponding sulfides via the M-S bond energy. However there are neither clear rules for 7 such relationship, nor any features of the band structure could be potential descriptors for the 8 HDS and/or HER activity. The HDS and HER rates for the sulfides having similar electronic

9 bands vary by many orders or magnitude. Thus, semiconducting sulfides with E g between 1

10 and 2 eV can be both highly active (MoS 2, RuS 2) and very poor (HgS) HDS catalysts.

11 On the other hand the PHER rate sequence vs. the nature of sulfide and its dispersion has 12 much lesser variations. All tested sulfides induce a significant increase of the PHER rate on 13 titania, even if the ranking is not the same for two alcohols. This observation suggests that the 14 action of different sulfides might be explained on a common basis. For PHER, not the 15 hydrogen-activating power but the electronic structure of transition metal sulfide is the most 16 important property.

17 Consideration of relative positions of electronic bands shows that similar alignment between

18 TiO 2 and all sulfides should occur. The electron affinity of all sulfides under study is higher

19 than for titania. The Fermi levels are slightly higher than Fermi level of TiO 2. Position of VB 20 slightly varies from one sulfide to another, but in all cases the top of VB is due to the states 21 dominated by S 3p electrons. After the band alignment the bottom of CB for semiconductor

22 sulfides (E f for metallic sulfides) will be always below the bottom of CB for TiO 2 (Scheme 2).

23

24 Scheme 2. Relative bands positions of titania and seven sulfides studied in this work; values 25 adopted from [62].

19

1 Therefore the photo generated electrons from the bottom CB of titania (or shallow trapped 2 electrons) wouldtransferred to the sulfides CB. That means, photo generated electrons will be

3 attracted towards the interface between sulfide and TiO 2. If a sulfide possess strong hydrogen 4 - activating power then the overall PHER rate might be additionally increased due to the

5 catalytic action of sulfide surface (similarly to the well-known Pt/TiO 2 reference case). 6 However, even if the sulfide possesses no hydrogen-activating properties at all, the PHER rate 7 is still higher than for bare titania, due to an enhance charge carrier separation. In that case,

8 TiO 2 is the active site for both oxidation and reduction half reactions.

9 The second important finding to be explained is that the dispersion of TM sulfide particles

10 does not play a crucial role. We suppose that due to polarization of TiO 2 agglomerate by 11 adjacent sulfide particle, the photo generated electrons which have high mobility and

12 relatively long lifetime, might be collected towards the interface between TiO 2 and sulfide via 13 an antenna-like mechanism [63], as depicted in Scheme 3.

14

15 Scheme 3 Harvesting of photo generated electrons from distant regions of a TiO 2 agglomerate by 16 a TM sulfide particle.

17 Finally there are two distinct groups of co-catalysts with respect to their relative activity in 18 water-IPA and in methanol media. In the first group, water-IPA and methanol tests give the 19 PHER rates ratio close to 1. This group includes Ru, Hg sulfides and Pt metal co-catalysts 20 (Table S2). For the second group including five of seven tested sulfides (Co, Cu, Mo, Ag and 21 Ni), the PHER rate in methanol is at least an order of magnitude greater than in water-IPA. 22 Both groups include relatively poor and highly active catalysts in PHER as well as in HER 23 and HDS. Both groups include relatively poorly dispersed and highly dispersed sulfides. No

20

1 property descriptor correlated with this activity difference could be identified for the time 2 being. However such a striking difference of behavior suggests that some fundamental 3 mechanistic difference exists between two types of co-catalysts, though it remains yet 4 unexplained.

5 4. Conclusions 6 In this work we prepared a series of transition metals sulfides on PC500 titania using 7 thioacetamide solution reaction. The properties of obtained catalysts were compared in PHER

8 in methanol and water-isopropanol media, as well as electrocatalytic HER in 0.5 M H 2SO 4 9 and heterogeneous gas-phase thiophene HDS reaction. The following conclusions can be 10 drawn from our work.

11 (i) Several transition metals sulfides of the first and the second raw show high PHER activity

12 in the methanol and water-isopropanol media, superior to that of frequently studied MoS 2.

13 (ii) Commonly known hydrogen-activating sulfides (Ru, Co, Ni) show the best photocatalytic 14 HER performance. However, other metals sulfides such as Hg or Cu have a poor activity for 15 HDS but show a PE many times higher than bare titania. Similarly, the trends established for 16 the seven sulfides differ in photocatalytic and electrocatalytic HER. Therefore proton 17 reduction or hydrogen activation by the co-catalyst, albeit favorable, is not a critical parameter 18 in the PHER.

19 (iii) Dispersion of the sulfide species has no determining effect on the PHER performance.

20 Porly dispersed NiS 2 and highly dispersed CoS x species show comparable high PHER 21 activity. In agreement with previous observations, sulfide particles, even moderately 22 dispersed, can efficiently harvest electrons from extended areas of adjacent titania.

23 (iv) The observed PHER rates strongly depend on both the nature of sulfide and that of the 24 organic substrate. However, this observation does not allow concluding on which step in the 25 multistep PHER scheme is rate limiting, because the organic molecule might interfere with 26 both electron and hole evolution. In order to understand this issue, integrated models should 27 be developed with rate equation including all the essential steps of the process. A better 28 insight can be obtained further measuring apparent activation energy of the process.

29 Supplementary data

21

1 Supplementary material is available including experimental conditions, samples 2 compositions, photocatalytic results, EDS analyses,TPR patterns, XPS spectra and UV-Vis 3 absorption spectra.

22

1 References

2 [1] Le Goff A, Artero V, Jousselme B, Tran PD, Guillet N, Metaye R, et al. From Hydrogenases to

3 Noble Metal-Free Catalytic Nanomaterials for H 2 Production and Uptake. Science 4 2009;326:1384–7. doi:10.1126/science.1179773. 5 [2] Lewis NS, Nocera DG. Powering the planet: Chemical challenges in solar energy utilization. 6 Proceedings of the National Academy of Sciences 2006;103:15729–35. 7 doi:10.1073/pnas.0603395103.

8 [3] Schneider J, Matsuoka M, Takeuchi M, Zhang J, Horiuchi Y, Anpo M, et al. Understanding TiO 2 9 Photocatalysis: Mechanisms and Materials. Chemical Reviews 2014;114:9919–86. 10 doi:10.1021/cr5001892. 11 [4] Camposeco R, Castillo S, Navarrete J, Gomez R. Synthesis, characterization and photocatalytic

12 activity of TiO 2 nanostructures: Nanotubes, nanofibers, nanowires and nanoparticles. Catalysis 13 Today 2016;266:90–101. doi:10.1016/j.cattod.2015.09.018.

14 [5] Yu J, Qi L, Jaroniec M. Hydrogen Production by Photocatalytic Water Splitting over Pt/TiO 2 15 Nanosheets with Exposed (001) Facets. The Journal of Physical Chemistry C 2010;114:13118– 16 25. doi:10.1021/jp104488b.

17 [6] Tang J, Durrant JR, Klug DR. Mechanism of Photocatalytic Water Splitting in TiO 2. Reaction of 18 Water with Photoholes, Importance of Charge Carrier Dynamics, and Evidence for Four-Hole 19 Chemistry. Journal of the American Chemical Society 2008;130:13885–91. 20 doi:10.1021/ja8034637. 21 [7] Kandiel TA, Dillert R, Robben L, Bahnemann DW. Photonic efficiency and mechanism of 22 photocatalytic molecular hydrogen production over platinized titanium dioxide from aqueous 23 methanol solutions. Catalysis Today 2011;161:196–201. doi:10.1016/j.cattod.2010.08.012. 24 [8] Kumar SG, Rao KSRK. Comparison of modification strategies towards enhanced charge carrier

25 separation and photocatalytic degradation activity of metal oxide semiconductors (TiO 2, WO 3 26 and ZnO). Applied Surface Science 2017;391:124–48. doi:10.1016/j.apsusc.2016.07.081.

27 [9] Gomathi Devi L, Kavitha R. A review on plasmonic metal-TiO 2 composite for generation, 28 trapping, storing and dynamic vectorial transfer of photogenerated electrons across the 29 Schottky junction in a photocatalytic system. Applied Surface Science 2016;360:601–22. 30 doi:10.1016/j.apsusc.2015.11.016.

31 [10] Du F, Lu H, Lu S, Wang J, Xiao Y, Xue W, et al. Photodeposition of amorphous MoS x cocatalyst 32 on TiO 2 nanosheets with {001} facets exposed for highly efficient photocatalytic hydrogen 33 evolution. International Journal of Hydrogen Energy 2018;43:3223–34. 34 doi:10.1016/j.ijhydene.2017.12.181.

35 [11] Chen B, Meng Y, Sha J, Zhong C, Hu W, Zhao N. Preparation of MoS 2/TiO 2 based 36 nanocomposites for photocatalysis and rechargeable batteries: progress, challenges, and 37 perspective. Nanoscale 2018;10:34–68. doi:10.1039/C7NR07366F.

38 [12] Jing D, Guo L. WS 2 sensitized mesoporous TiO 2 for efficient photocatalytic hydrogen production 39 from water under visible light irradiation. Catalysis Communications 2007;8:795–9. 40 doi:10.1016/j.catcom.2006.09.009.

41 [13] Afanasiev P. Synthesis of BaW 2O7-ethylene glycol inorganic–organic hybrid and its topochemical 42 transformation to thin WS 2 nanoplates. Journal of Solid State Chemistry 2018;258:656–63. 43 doi:10.1016/j.jssc.2017.11.036. 44 [14] Wang Q, An N, Bai Y, Hang H, Li J, Lu X, et al. High photocatalytic hydrogen production from

45 methanol aqueous solution using the photocatalysts CuS/TiO 2. International Journal of 46 Hydrogen Energy 2013;38:10739–45. doi:10.1016/j.ijhydene.2013.02.131.

47 [15] Xu F, Zhang L, Cheng B, Yu J. Direct Z-Scheme TiO 2/NiS Core–Shell Hybrid Nanofibers with 48 Enhanced Photocatalytic H 2-Production Activity. ACS Sustainable Chem Eng 2018;6:12291–8. 49 doi:10.1021/acssuschemeng.8b02710.

23

1 [16] Yu Z, Meng J, Xiao J, Li Y, Li Y. Cobalt sulfide quantum dots modified TiO 2 nanoparticles for 2 efficient photocatalytic hydrogen evolution. International Journal of Hydrogen Energy 3 2014;39:15387–93. doi:10.1016/j.ijhydene.2014.07.165.

4 [17] Zong X, Han J, Ma G, Yan H, Wu G, Li C. Photocatalytic H 2 evolution on CdS loaded with WS 2 as 5 cocatalyst under visible light irradiation. The Journal of Physical Chemistry C 2011;115:12202–8. 6 doi:10.1021/jp2006777. 7 [18] Yao W, Huang C, Muradov N, T-Raissi A. A novel Pd–Cr2O3/CdS photocatalyst for solar 8 hydrogen production using a regenerable sacrificial donor. International Journal of Hydrogen 9 Energy 2011;36:4710–5. doi:10.1016/j.ijhydene.2010.12.124.

10 [19] Cui H, Li B, Zhang Y, Zheng X, Li X, Li Z, et al. Constructing Z-scheme based CoWO 4/CdS 11 photocatalysts with enhanced dye degradation and H 2 generation performance. International 12 Journal of Hydrogen Energy 2018;43:18242–52. doi:10.1016/j.ijhydene.2018.08.050. 13 [20] Kim J, Kang M. High photocatalytic hydrogen production over the band gap-tuned urchin-like

14 Bi 2S3-loaded TiO 2 composites system. International Journal of Hydrogen Energy 2012;37:8249– 15 56. doi:10.1016/j.ijhydene.2012.02.057.

16 [21] Zhang C, Zhou Y, Bao J, Fang J, Zhao S, Zhang Y, et al. Structure regulation of ZnS@g-C3N4/TiO 2 17 nanospheres for efficient photocatalytic H 2 production under visible-light irradiation. Chemical 18 Engineering Journal 2018;346:226–37. doi:10.1016/j.cej.2018.04.038.

19 [22] Jia F, Yao Z, Jiang Z. Solvothermal synthesis ZnS–In 2S3–Ag 2S solid solution coupled with TiO 2-xSx 20 nanotubes film for photocatalytic hydrogen production. International Journal of Hydrogen 21 Energy 2012;37:3048–55. doi:10.1016/j.ijhydene.2011.11.012.

22 [23] Zhang J, Yu J, Zhang Y, Li Q, Gong JR. Visible Light Photocatalytic H 2-Production Activity of 23 CuS/ZnS Porous Nanosheets Based on Photoinduced Interfacial Charge Transfer. Nano Letters 24 2011;11:4774–9. doi:10.1021/nl202587b.

25 [24] Li Y, Chen G, Wang Q, Wang X, Zhou A, Shen Z. Hierarchical ZnS-In 2S3-CuS Nanospheres with 26 Nanoporous Structure: Facile Synthesis, Growth Mechanism, and Excellent Photocatalytic 27 Activity. Advanced Functional Materials 2010;20:3390–8. doi:10.1002/adfm.201000604.

28 [25] Feng H, Tang N, Zhang S, Liu B, Cai Q. Fabrication of layered (CdS-Mn/MoS 2/CdTe)-promoted 29 TiO 2 nanotube arrays with superior photocatalytic properties. Journal of Colloid and Interface 30 Science 2017;486:58–66. doi:10.1016/j.jcis.2016.09.048. 31 [26] Girel E, Puzenat E, Geantet C, Afanasiev P. On the photocatalytic and electrocatalytic hydrogen

32 evolution performance of molybdenum sulfide supported on TiO 2. Catalysis Today 33 2017;292:154–63. doi:10.1016/j.cattod.2016.09.018. 34 [27] Kachina A, Puzenat E, Ould-Chikh S, Geantet C, Delichere P, Afanasiev P. A New Approach to the 35 Preparation of Nitrogen-Doped Titania Visible Light Photocatalyst. Chemistry of Materials 36 2012;24:636–42. doi:10.1021/cm203848f. 37 [28] Kortüm G, Braun W, Herzog G. Principles and Techniques of Diffuse-Reflectance Spectroscopy. 38 Angewandte Chemie International Edition 1963;2:333–341. 39 [29] Braslavsky SE, Braun AM, Cassano AE, Emeline AV, Litter MI, Palmisano L, et al. Glossary of 40 terms used in photocatalysis and radiation catalysis (IUPAC Recommendations 2011). Pure and 41 Applied Chemistry 2011;83:931–1014. doi:10.1351/PAC-REC-09-09-36.

42 [30] Cui Y, He J, Yuan F, Xue J, Li X, Wang J. Preparation of MoS 2 microspheres through surfactants- 43 assisted hydrothermal synthesis using thioacetamide as reducing agent. Hydrometallurgy 44 2016;164:184–8. doi:10.1016/j.hydromet.2016.05.002. 45 [31] Li Y, Li J, Wang M, Liu Y, Cui H. High rate performance and stabilized cycle life of Co 2+ -doped 46 nanosheets synthesized by a scalable method of solid-state reaction. Chemical 47 Engineering Journal 2019;366:33–40. doi:10.1016/j.cej.2019.02.066. 48 [32] Iqbal M, Ali A, Nahyoon NA, Majeed A, Pothu R, Phulpoto S, et al. Photocatalytic degradation of 49 organic pollutant with nanosized cadmium sulfide. Materials Science for Energy Technologies 50 2019;2:41–5. doi:10.1016/j.mset.2018.09.002.

24

1 [33] Ma T, Sun L, Niu Q, Xu Y, Zhu K, Liu X, et al. N-doped carbon-coated Tin sulfide/graphene 2 nanocomposite for enhanced lithium storage. Electrochimica Acta 2019;300:131–7. 3 doi:10.1016/j.electacta.2019.01.104. 4 [34] Chen F, Jin X, Cao Y, Jia D, Liu A, Wu R, et al. Effects of the synthesis conditions on the 5 photocatalytic activities of sulfide-graphene oxide composites. Dyes and Pigments 6 2019;162:177–88. doi:10.1016/j.dyepig.2018.09.054. 7 [35] Afanasiev P. On the interpretation of temperature programmed reduction patterns of 8 transition metals sulphides. Applied Catalysis A: General 2006;303:110–5. 9 doi:10.1016/j.apcata.2006.02.014. 10 [36] Mitchell P, Scott C, Bonnelle J-P, Grimblot J. Ru/Alumina and Ru-Mo/Alumina Catalysts: An XPS 11 Study. Journal of Catalysis 1987;107:482–9. doi:10.1016/0021-9517(87)90312-5. 12 [37] Morgan DJ. Resolving ruthenium: XPS studies of common ruthenium materials: XPS studies of 13 common ruthenium materials. Surface and Interface Analysis 2015;47:1072–9. 14 doi:10.1002/sia.5852. 15 [38] De Los Reyes JA. Ruthenium sulfide supported on alumina as hydrotreating catalyst. Applied 16 Catalysis A: General 2007;322:106–12. doi:10.1016/j.apcata.2007.01.004.

17 [39] Huang J, Shi Z, Dong X. Nickel sulfide modified TiO 2 nanotubes with highly efficient 18 photocatalytic H 2 evolution activity. Journal of Energy Chemistry 2016;25:136–40. 19 doi:10.1016/j.jechem.2015.11.007.

20 [40] Li H, Wang Y, Chen G, Sang Y, Jiang H, He J, et al. Few-layered MoS 2 nanosheets wrapped 21 ultrafine TiO 2 nanobelts with enhanced photocatalytic property. Nanoscale 2016;8:6101–9. 22 doi:10.1039/C5NR08796A.

23 [41] Zhang L, Tian B, Chen F, Zhang J. Nickel sulfide as co-catalyst on nanostructured TiO 2 for 24 photocatalytic hydrogen evolution. International Journal of Hydrogen Energy 2012;37:17060–7. 25 doi:10.1016/j.ijhydene.2012.08.120. 26 [42] Buriak JM, Kamat PV, Schanze KS. Best Practices for Reporting on Heterogeneous 27 Photocatalysis. ACS Applied Materials & Interfaces 2014;6:11815–6. doi:10.1021/am504389z. 28 [43] Kisch H, Bahnemann D. Best Practice in Photocatalysis: Comparing Rates or Apparent Quantum 29 Yields? The Journal of Physical Chemistry Letters 2015;6:1907–10. 30 doi:10.1021/acs.jpclett.5b00521. 31 [44] Schneider J, Bahnemann DW. Undesired Role of Sacrificial Reagents in Photocatalysis. The 32 Journal of Physical Chemistry Letters 2013;4:3479–83. doi:10.1021/jz4018199. 33 [45] Bowker M, Morton C, Kennedy J, Bahruji H, Greves J, Jones W, et al. Hydrogen production by

34 photoreforming of biofuels using Au, Pd and Au–Pd/TiO 2 photocatalysts. Journal of Catalysis 35 2014;310:10–5. doi:10.1016/j.jcat.2013.04.005. 36 [46] Bahruji H, Bowker M, Davies PR, Pedrono F. New insights into the mechanism of photocatalytic

37 reforming on Pd/TiO 2. Applied Catalysis B: Environmental 2011;107:205–9. 38 doi:10.1016/j.apcatb.2011.07.015. 39 [47] Yu J, Guo Y, Miao S, Ni M, Zhou W, Shao Z. Spherical Ruthenium Disulfide-Sulfur-Doped 40 Graphene Composite as an Efficient Hydrogen Evolution Electrocatalyst. ACS Applied Materials 41 & Interfaces 2018;10:34098–107. doi:10.1021/acsami.8b08239. 42 [48] Faber MS, Dziedzic R, Lukowski MA, Kaiser NS, Ding Q, Jin S. High-Performance Electrocatalysis

43 Using Metallic Cobalt Pyrite (CoS 2) Micro- and Nanostructures. Journal of the American 44 Chemical Society 2014;136:10053–61. doi:10.1021/ja504099w. 45 [49] Maheu C, Cardenas L, Puzenat E, Afanasiev P, Geantet C. UPS and UV spectroscopies combined

46 to position the energy levels of TiO 2 anatase and rutile nanopowders. Physical Chemistry 47 Chemical Physics 2018;20:25629–37. doi:10.1039/C8CP04614J.

48 [50] Ho TC. Hydrodesulfurization with RuS 2 at low hydrogen pressures. Catalysis Letters 2003;89:5. 49 [51] Jaramillo TF, Jorgensen KP, Bonde J, Nielsen JH, Horch S, Chorkendorff I. Identification of Active

50 Edge Sites for Electrochemical H 2 Evolution from MoS 2 Nanocatalysts. Science 2007;317:100–2. 51 doi:10.1126/science.1141483.

25

1 [52] Lacroix M, Boutarfa N, Guillard C, Vrinat M, Breysse M. Hydrogenating properties of 2 unsupported transition metal sulphides. Journal of Catalysis 1989;120:473–7. 3 doi:10.1016/0021-9517(89)90287-X. 4 [53] Quartararo J, Mignard S, Kasztelan S. Hydrodesulfurization and hydrogenation activities of 5 alumina-supported transition metal sulfides. Journal of Catalysis 2000;192:307–15. 6 doi:10.1006/jcat.2000.2825. 7 [54] Afanasiev P, Bezverkhyy I. Ternary transition metals sulfides in hydrotreating catalysis. Applied 8 Catalysis A: General 2007;322:129–41. doi:10.1016/j.apcata.2007.01.015.

9 [55] Basu M, Nazir R, Mahala C, Fageria P, Chaudhary S, Gangopadhyay S, et al. Ag 2S/Ag 10 Heterostructure: A Promising Electrocatalyst for the Hydrogen Evolution Reaction. Langmuir 11 2017;33:3178–86. doi:10.1021/acs.langmuir.7b00029. 12 [56] Laursen AB, Varela AS, Dionigi F, Fanchiu H, Miller C, Trinhammer OL, et al. Electrochemical 13 Hydrogen Evolution: Sabatier’s Principle and the Volcano Plot. J Chem Educ 2012;89:1595–9. 14 doi:10.1021/ed200818t. 15 [57] Medford AJ, Vojvodic A, Hummelshøj JS, Voss J, Abild-Pedersen F, Studt F, et al. From the 16 Sabatier principle to a predictive theory of transition-metal heterogeneous catalysis. Journal of 17 Catalysis 2015;328:36–42. doi:10.1016/j.jcat.2014.12.033. 18 [58] Hellstern TR, Kibsgaard J, Tsai C, Palm DW, King LA, Abild-Pedersen F, et al. Investigating 19 Catalyst–Support Interactions To Improve the Hydrogen Evolution Reaction Activity of

20 Thiomolybdate [Mo 3S13 ]2– Nanoclusters. ACS Catal 2017;7:7126–30. 21 doi:10.1021/acscatal.7b02133. 22 [59] Chianelli RR, Berhault G, Raybaud P, Kasztelan S, Hafner J, Toulhoat H. Periodic trends in 23 hydrodesulfurization: in support of the Sabatier principle. Applied Catalysis A: General 24 2002;227:83–96. doi:10.1016/S0926-860X(01)00924-3. 25 [60] Raybaud P. Understanding and predicting improved sulfide catalysts: Insights from first 26 principles modeling. Applied Catalysis A: General 2007;322:76–91. 27 doi:10.1016/j.apcata.2007.01.005. 28 [61] Lange NA, Dean JA. Lange’s handbook of chemistry. 12th ed., New York: McGraw-Hill; 1979. 29 [62] Xu Y, Schoonen MAA. The absolute energy positions of conduction and valence bands of 30 selected semiconducting minerals. American Mineralogist 2000;85:543–56. doi:10.2138/am- 31 2000-0416. 32 [63] Wang C, Pagel R, Dohrmann JK, Bahnemann DW. Antenna mechanism and deaggregation 33 concept: novel mechanistic principles for photocatalysis. Comptes Rendus Chimie 2006;9:761– 34 73. doi:10.1016/j.crci.2005.02.053.

35

26

Graphical abstract for :

Titania - supported transition metals sulfides as photocatalysts for hydrogen production from propan-2-ol and methanol

Clement Mahieu*, Eric Puzenat, Christophe Geantet, Luis Cardenas, Pavel Afanasiev*

Institut de Recherches sur la Catalyse et l’Environnement de Lyon IRCELYON, UMR 5256, CNRS – Université Lyon 1, 2 av A. Einstein 69626 Villeurbanne Cedex (France); Fax: 33 04 7244 5399; Tel: 33 04 72 44 5466;

E-mail: [email protected] clement.mahieu @ircelyon.univ-lyon1.fr