Você está na página 1de 17

Supporting Information

Spherical Ruthenium Disulfide-Sulfur-Doped


Graphene Composite as An Efficient Hydrogen
Evolution Electrocatalyst
Jie Yu,† Yanan Guo,† Shuanshuan Miao,† Meng Ni,‡ Wei Zhou,*,† and Zongping
Shao*,†,§


Jiangsu National Synergetic Innovation Center for Advanced Materials (SICAM),

State Key Laboratory of Materials-Oriented Chemical Engineering, College of

Chemical Engineering, Nanjing Tech University, No. 5, Xin Mofan Road, Nanjing

210009, P.R. China. *E-mail: zhouwei1982@njtech.edu.cn (Wei Zhou);

shaozp@njtech.edu.cn (Zongping Shao).


Building Energy Research Group, Department of Building and Real Estate, The

Hong Kong Polytechnic University, Hung Hom, Kowloon, 999077, Hong Kong,

China.

§
Department of Chemical Engineering, Curtin University, Perth, Western Australia
6845, Australia.

S-1
Experimental Section

Materials synthesis

Synthesis of s-RuS2/S-rGO

First, graphite oxide powders, as the raw material for graphene in this study, was

purchased from Nanjing XFNANO Materials Tech Co., Ltd. In the typical synthesis

of s-RuS2/S-rGO composites, 50 mg of graphite oxide powders was sonicated in

53.3 ml of a mixed solution with ethylene glycol and DI water (1:3 by volume) for

about 1 h to form a homogeneous graphene oxide (GO) dispersion. Then, 83 mg of

ruthenium chloride (RuCl3) and 240 mg of L-cysteine were introduced into the above

solution under vigorous stirring for 1 h. After that, the well-mixed precursor solution

was transferred to a 100 ml Teflon-lined autoclave for the hydrothermal reaction at

433 K for 9 h. When it cooled down to room temperature, the s-RuS2/S-rGO

composite precursor (which is denoted s-RuS2/S-rGO-p) was collected by being

washed with water and then lyophilized. Finally, the precursor was treated at 650 °C

for 2 h in an argon atmosphere to obtain the s-RuS2/S-rGO hybrid.

Synthesis of s-RuS2, S-rGO, s-RuS2+S-rGO, rGO, Ru/rGO

As control experiments, pure spherical RuS2 particles (s-RuS2) and S-rGO were

prepared by the same procedure without the introduction of GO and RuCl3,

respectively. Subsequently, s-RuS2+S-rGO was prepared by physically mixing s-RuS2

and S-rGO in an appropriate proportion. In addition, a similar experimental process

was also followed to prepare rGO simultaneously without RuCl3 and L-cysteine.

S-2
Ru/rGO was also developed by the identical processes to those described above but

without the sulfur source and the argon atmosphere was changed to a hydrogen-argon

atmosphere.

Characterizations

In order to identify the phase and purity of the synthesized catalysts, powder X-ray

diffraction (XRD) measurements were conducted on a Rigaku Smartlab

diffractometer using filtered Cu-Kα radiation (λ = 1.5418 Å) with a 2θ range of

10-90°. The micro-morphologies of the samples were observed by a HITACHI-S4800

field-emission scanning electron microscope (FE-SEM) and an FEI Tecnai G2T20

transmission electron microscope (TEM). Additionally, the corresponding scanning

TEM (STEM)-EDX line scan and element mapping were taken on an FEI Tecnai G2

F30 STWIN field-emission transmission electron microscope equipped with an EDX

analyzer at 200 kV. The chemical compositions and surface element states of the

samples were also probed by X-ray photoelectron spectroscopy (XPS) measurements

on a PHI5000 VersaProbe spectrometer equipped with an Al-Kα X-ray source, and

these data were fitted by the public software package XPSPEAK. The Raman spectra

of the samples were measured to determine the degree of graphitization using an

HR800 UV Raman microspectrometer (JOBIN YVON, France) with the green line of

an argon laser as the excitation source. Based on the N2-adsorption-desorption

isotherms, the specific surface areas and pore size distributions were obtained by

using the Brunauer-Emmett-Teller (BET) and Barrett-Joyner-Halenda (BJH) methods.

The mass ratio of the total metal was approximately determined via thermo

S-3
gravimetric analysis (TGA) on a thermo-balance (STA 449 F3 Jupiter®, NETZSCH).

Electrode preparation and electrochemical measurements

All electrochemical measurements were performed on a CHI 760E bipotentiostat with

a standard three-electrode electrochemical cell consisting of a glassy carbon/carbon

cloth (stability testing-chronopotentiometric (CP) measurements) as the working

electrode substrate, a graphite rod as the counter electrode, and Ag|AgCl (3.5 M KCl)

as the reference electrode. The HER performance was also studied in a wide pH

environment by using 1 M KOH, 0.5 M H2SO4 and 1 M PBS solutions as electrolytes,

respectively. To prepare the working electrode, 10 mg of each catalyst sample and 50

µL of 5 wt. % Nafion solution were dispersed in 500 µL of ethanol by at least 1 h of

mild sonication to form a homogeneous working electrode ink. Subsequently, 5 µL of

the suspension was drop-dried onto a glassy carbon substrate with 5 mm in diameter

(a loading of 0.464 mg cm-2). When the carbon cloth was used for stability testing, the

catalyst loading was 2 mg cm-2.

For approximately 30 min prior to the start of each test till the end of the test, the

electrolyte was continuously bubbled with Ar. Hundreds of potential cycles were first

carried out to make pretreatment at different potential regions versus Ag|AgCl for

HER tests in all pH values. Then, HER polarization curves obtained from the linear

sweep voltammetry (LSV) were recorded on a rotating disk electrode (RDE) at a scan

rate of 5 mV s-1 from different potential regions versus Ag|AgCl in diverse pH media

with constant rotation speeds of 1600 rpm to eliminate the bubbles. All potentials

were calibrated against and converted to the reversible hydrogen electrode (RHE)

S-4
after iR-corrected: ERHE = EAg/AgCl +0.197 + 0.059×pH. The polarization curves were

replotted as the overpotential (η) versus the logarithm of the current density (log |j|) to

obtain the Tafel plots. The electrochemical double-layer capacitance (EDLCs), Cdl,

was calculated from the cyclic voltammetry curves (CVs), which were recorded in a

potential range with no faradic current at different rates from 20 to 160 mV s-1. To

study the electrode kinetics, electrochemical impedance spectra (EIS) measurements

were performed over a frequency range from 100 kHz to 0.1 Hz under the influence

of an AC voltage of 5 mV. The turnover frequency (TOF) values of various

electrocatalysts were calculated based on a previously reported method.1,2 Accelerated

stability tests of the catalysts were conducted by continuous potential cycling at a

sweep rate of 100 mV s-1 for a given number of cycles. Long-term

chronopotentiometric (CP) measurements were studied at three fixed current densities

of 10 mA cm-2, 20 mA cm-2, and 50 mA cm-2, respectively, for 100 h. The overall

water splitting tests were conducted in a homemade two-electrode system with

catalysts loaded on Ni foams (mass loading: 5 mg cm-2). The polarization curves were

obtained with a scan rate of 5 mV s-1 and further iR-corrected.

Figure S1. XRD patterns of (a) the as-prepared s-RuS2/S-rGO-p composite, and (b) other contrast
samples including pure s-RuS2, metal-free S-rGO, and Ru/rGO catalysts.
S-5
Figure S2. The particle size distribution histogram of RuS2.

Figure S3. SEM images of the spherical RuS2 (a) and Ru/rGO (b) catalysts.

Figure S4. Thermogravimetric analysis (TGA) curve of the s-RuS2/S-rGO hybrid.

Figure S5. The overview survey XPS spectrum of the s-RuS2/S-rGO hybrid.
S-6
Figure S6. High-resolution XPS spectra of Ru 3p for the s-RuS2/S-rGO and s-RuS2+S-rGO
composites.

Figure S7. Raman spectra of the s-RuS2/S-rGO and S-rGO catalysts.

Figure S8. Overpotentials at the current densities of 10 (a) and 50 (b) mA cm-2 for the
s-RuS2/S-rGO hybrid, pure s-RuS2, metal-free S-rGO, s-RuS2+S-rGO, and s-RuS2/S-rGO-p in 1
M KOH. Error bars represent standard deviations from at least three independent measurements.

S-7
Figure S9. HER polarization curves of the S-rGO and rGO samples in 1 M KOH solution at a
scan rate of 5 mV s-1.

Figure S10. EIS Nyquist plots of s-RuS2/S-rGO and s-RuS2 catalysts measured at the potential of
-1.2 V vs. Ag/AgCl.

Figure S11. (a, b) Cyclic voltammetry profiles for s-RuS2/S-rGO (a) and s-RuS2+S-rGO (b) at
different scan rates (20-160 mV s-1) in 1 M KOH, and c) the corresponding linear fitting plots of
the capacitive currents versus scan rates.

S-8
Figure S12. Tafel plots of these s-RuS2/S-rGO, s-RuS2, S-rGO, s-RuS2+S-rGO, and
s-RuS2/S-rGO-p samples obtained from the polarization curves in Figure 3a.

Figure S13. Thermogravimetric analysis (TGA) curve of the Ru/rGO hybrid.

Figure S14. HER polarization curves of the s-RuS2/S-rGO hybrid, commercial RuO2, and IrO2
catalysts in 1 M KOH (a), 0.5 M H2SO4 (b), and 1 M PBS (c) solutions, respectively.

S-9
Figure S15. Calculated price activities of s-RuS2/S-rGO and two commercial Pt/C catalysts (20
wt.% and 40 wt.% Pt/C) in the 1 M KOH aqueous solution.

Figure S16. Time-dependent potential changes at a fixed current density of 10 mA cm-2 for the
s-RuS2/S-rGO sample in alkaline media.

S-10
Figure S17. (a) HER polarization curves in 0.5 M H2SO4 solution at a scan rate of 5 mV s-1, and
(b) the corresponding Tafel plots of the s-RuS2/S-rGO hybrid, pure RuS2, metal-free S-rGO,
s-RuS2+S-rGO, s-RuS2/S-rGO-p, Ru/rGO, and the commercial two Pt/C catalysts (20 wt.% Pt/C
and 40 wt.% Pt/C). (c, d) Overpotentials at the current densities of 10 (c) and 50 (d) mA cm-2 for
the s-RuS2/S-rGO hybrid, pure RuS2, metal-free S-rGO, s-RuS2+S-rGO, s-RuS2/S-rGO-p, Ru/rGO,
and the commercial two Pt/C catalysts (20 wt.% Pt/C and 40 wt.% Pt/C) in 0.5 M H2SO4 solution.
Error bars represent standard deviations from at least three independent measurements.

S-11
Figure S18. (a) HER polarization curves in 1 M PBS solution at a scan rate of 5 mV s-1, and (b)
the corresponding Tafel plots of the s-RuS2/S-rGO hybrid, pure RuS2, metal-free S-rGO,
s-RuS2+S-rGO, s-RuS2/S-rGO-p, Ru/rGO, and the commercial two Pt/C catalysts (20 wt.% Pt/C
and 40 wt.% Pt/C). (c, d) Overpotentials at the current densities of 10 (c) and 50 (d) mA cm-2 for
the s-RuS2/S-rGO hybrid, pure RuS2, metal-free S-rGO, s-RuS2+S-rGO, s-RuS2/S-rGO-p, Ru/rGO,
and the commercial two Pt/C catalysts (20 wt.% Pt/C and 40 wt.% Pt/C) in 1 M PBS solution.
Error bars represent standard deviations from at least three independent measurements.

Figure S19. (a) The energy diagram of the water dissociation on the Pt (111), RuO2 (110), and Ru
(001) surfaces. (b) The adsorption free energy of hydrogen intermediates adsorption (∆GH*) on the
Pt (111), RuO2 (110), and Ru (001) surfaces.

S-12
Table S1. Specific surface areas for different samples.
Sample SBET[a] (m2 g-1)
s-RuS2/S-rGO 53
s-RuS2 45
[a] SBET: specific surface area from BET method.

Table S2. Summary of some recently reported representative HER electrocatalysts in 1 M KOH.
Catalyst Loading η10 (mV) Tafel slope Electrolyte References
(mg cm-2) (mV dec-1)
s-RuS2/S-rGO 0.464 25 29 1 M KOH This work
MoCx 0.8 151 59 1 M KOH Nat. Commun.
nano-octahedrons 2015, 6, 6512.
CP/CTs/Co-S 0.32 190 131 1 M KOH ACS Nano 2016,
10, 2342.
CoMoP@C 0.354 81 55.53 1 M KOH Energy Environ.
Sci. 2017, 10, 788.
RuP2@NPC 1 ~ 52 69 1 M KOH Angew. Chem. Int.
Ed. 2017, 56,
11559.
Ni-BDT-A 0.3 80 70 1 M KOH Chem 2017, 3, 122.
RuCo@NC 0.275 28 31 1 M KOH Nat. Commun.
2017, 8, 14969.
Ru@C2N 0.285 17 38 1 M KOH Nat. Nanotech.
2017, 12, 441.
N-doped Ni3S2 0.6 155 113 1 M KOH Adv. Energy Mater.
2018, 1703538.
Pt3Ni2 NWs-S/C --- ~ 45 --- 1 M KOH Nat. Commun.
2017, 8, 14580.
Ni@Ni2P−Ru 0.286 31 41 1 M KOH J. Am. Chem. Soc.
HNRs 2018, 140, 2731.
IrCo@NC-500 0.277 45 80 1 M KOH Adv. Mater. 2018,
1705324
np-(Co0.52Fe0.48)2P --- > 50 40 1 M KOH Energy Environ.
Sci. 2016, 9, 2257.

Table S3. Summary of some recently reported representative HER electrocatalysts in 0.5 M
H2SO4 and 1 M PBS.
Catalyst Loading η10 (mV) Tafel slope Electrolyte References
(mg cm-2) (mV dec-1)
s-RuS2/S-rGO 0.464 69 64 0.5 M H2SO4 This work
Ni2P@NPCNFs 0.337 63.2 56.7 0.5 M H2SO4 Angew. Chem. Int.
Ed. 2018, 57, 1.
RuP2@NPC 1 38 38 0.5 M H2SO4 Angew. Chem. Int.
S-13
Ed. 2017, 56,
11559.
Mn-5% 0.34 43 34 0.5 M H2SO4 ACS Energy Lett.
2018, 3, 779.
MoCx 0.8 142 53 0.5 M H2SO4 Nat. Commun.
nano-octahedrons 2015, 6, 6512.
CoP/CC 0.92 209 129 0.5 M H2SO4 J. Am. Chem. Soc.
2014, 136, 7587.
CoNx/C 2 133 57 0.5 M H2SO4 Nat. Commun.
2015, 6, 7992.
s-RuS2/S-rGO 0.464 93 41 1 M PBS This work
CoNx/C 2 247 --- 1 M PBS Nat. Commun.
2015, 6, 7992.
RuP2@NPC 1 57 87 1 M PBS Angew. Chem. Int.
Ed. 2017, 56,
11559.
CoP@BCN-1 0.4 122 59 1 M PBS Adv. Energy
Mater. 2017, 7,
1601671.
Ni2P@NPCNFs 0.337 185.3 230.3 1 M PBS Angew. Chem. Int.
Ed. 2018, 57, 1.
Ni0.89Co0.11Se2 2.62 82 78 1 M PBS Adv. Mater. 2017,
MNSN/NF 29, 1606521.

Note S1. TOF calculation of the catalysts.


In order to obtain the turnover frequency (TOF) per site for our developed catalyst,
the formula was used as follows:
# 
 /  
TOF = (1)
#  /  

The total number of hydrogen turn overs was calculated from the current density
according to:
!"
 $
' $ 
0 ).%×$%2 

 0
#H = j   #
& ()*+,../  
'   = 3.12 ×
$%%% $ 
0

8 /9
10$, 7 :;

 (2)
As previous report,1, 2 due to the ambiguous hydrogen binding sites, the number of
active sites was conservatively estimated per surface atom (for RuS2, both Ru and S)
using the crystal structure (unit cell) of each sample. Actual surface area values were
obtained from the BET measurements. Owing to the small BET surface area
difference between s-RuS2/S-rGO and pure s-RuS2, the specific surface area of 53 m2
g-1 for s-RuS2/S-rGO was used to roughly estimate active sites.
For s-RuS2/S-rGO, surface atoms per testing area:

S-14
12 atoms/unit cell U UV
# active sites = ( )/. × 53 × 0.464  × 0.38 =
5.6095. Å. /MNOP QRSS V QU
Z[\;9
1.56× 10$Y (3)
:;

Turnover frequency (per surface atom) of s-RuS2/S-rGO in alkaline solutions:


^ /# `b `b
..$×$%!]   a  ×, 
At η = 40 mV TOF = _` _` _`
def`# = 0.5 s g$ atomg$
$.,)×$%!c
_`

^
`b `b
..$×$%!] #  a  ×,% 
At η = 56 mV TOF = _` _` _`
def`# = 1 sg$ atomg$
$.,)×$%!c
_`

Note S2. DFT Methods and Models.


The crystal bulk structures of RuS2, RuO2, Ru and Pt were obtained from the
Crystallography Open Database (COD). The optimized lattice parameters are quite
similar to experimental results for all four materials. The RuS2 (200) 2×2 supercell
with periodically repeated four-atomic layers, the RuO2 (110) 2×1 supercell with
five-atomic layers, the Ru (001) 3×3 supercell with three-atomic layers, and the

√3 × √3 reconstructed Pt (111) 4×4 supercell with three-atomic layers were


modelled. All DFT calculations were performed using the Amsterdam Density
Functional (ADF) package (version 2017). 3, 4 The revised Perdew-Burke-Ernzerhof
(revPBE) functional 5 within the generalized gradient approximation (GGA) was
applied to deal with the electron exchange-correlation. 6 The Grimme’s correction
employing BJ-damping (-D3(BJ)) was used to account for the dispersion interaction. 7
The atomic orbitals with double-zeta quality and one polarization function (DZP) was
adopted. 8 The relativistic effects of heavy atoms were taken into account by
employing the scalar zero order regular approximation (ZORA). 9 The core electrons
of atoms were frozen to speed up the calculation. The Quasi Newton optimization
method was employed throughout the study. The geometry optimization was
considered to be converged when all five criteria are met: (i) the change in bond
energy is less than 1×10-3 Hartree; (ii) the maximum Cartesian nuclear gradient is
small than 1×10-2 Hartree/Angstrom; (iii) the root mean square (RMS) of the
Cartesian nuclear gradients is smaller than 2/3×10-2 Hartree/Angstrom; (iv) the
maximum Cartesian step is smaller than 1×10-2 Angstrom; (v) the RMS of the
Cartesian steps is smaller than 2/3×10-2 Angstrom. The two-dimensional periodic
boundary condition was applied. The k-point over the Brillouin zone for all structures
was set to 3×3×1, which was sampled based on the Wiesenekker-Baerends scheme. 10
In geometry optimizations of the bulk cells and the individual surfaces, all atoms were
allowed to relax. Relatively, in the case of H2O/H adsorbed surfaces, the top three
layers for RuS2 and RuO2 and the top two layers for Ru and Pt were relaxed, while the
rest of the atoms were fixed in their equilibrium positions. Three H adsorption sites on
the RuS2 (200) surface were taken into account, i.e. the Ru atom top (Ru-T), the S
S-15
atoms top (S-T) and the hollow (H, see Fig. 6b) sites. The calculated ∆GH* vales at the
Ru-T (0.43 eV) and the S-T (0.58 eV) sites are larger than that at the H site (0.16 eV).
In the main text, the ∆GH* at the H site was used for the discussion. For RuO2 (110),
Ru (001) and Pt (111), various H adsorption sites for each surface were considered.
The site model which has a positive H binding energy was eliminated. Among the
remaining site models, the smallest ∆GH* was employed for the discussion.
The linear transit method 11 was employed to search for the transition state for
water dissociation. The H adsorption free energy was calculated through the following
equation as in previous calculations: 12, 13
∆GH* = ∆Eads +∆EZPE -T∆SH (4)
where ∆Eads is the H adsorption energy, ∆EZPE the zero point energy difference of H
in the adsorbed state and gas phase state, ∆SH the entropy difference of H in the
adsorbed state and gas phase state. The approximate value of (∆EZPE -T∆SH) is 0.24
eV at 300 K, therefore
∆GH* = ∆Eads + 0.24 eV (5)
The H adsorption energy was computed by
$ k
ΔE  = k lE(surface + Np) − E(surface) −  E(H )r (6)
in which n is the number of H atoms in the model. 12

References

(1) Popczun, E. J.; McKone, J. R.; Read, C. G.; Biacchi, A. J.; Wiltrout, A. M.; Lewis,
N. S.; Schaak, R. E. Nanostructured Nickel Phosphide as An Electrocatalyst for the
Hydrogen Evolution Reaction. J. Am. Chem. Soc. 2013, 135, 9267-9270.
(2) Laursen, A. B.; Patraju, K. R.; Whitaker, M. J.; Retuerto, M.; Sarkar, T.; Yao, N.;
Ramanujachary, K. V.; Greenblatt, M.; Dismukes, G. C. Nanocrystalline Ni5P4: A
Hydrogen Evolution Electrocatalyst of Exceptional Efficiency in Both Alkaline and
Acidic Media. Energy Environ. Sci. 2015, 8, 1027-1034.
(3) Philipsen, P. H. T.; Velde, G. T.; Baerends, E. J.; Berger, J. A.; Boeij, P. L. D.;
Franchini, M.; Groeneveld, J. A.; Kadantsev, E. S.; Klooster, R.; Kootstra, F.;
Romaniello, P.; Skachkov, D. G.; Snijders, J. G.; Verzijl, C. J. O.; Wiesenekker, G.;
Ziegler, T. BAND2017, SCM, Theoretical Chemistry, Vrije Universiteit, Amsterdam,
The Netherlands, http://www.scm.com.
(4) Velde, G. T.; Baerends, E. J. Precise Density-Functional Method for Periodic
Structures. Phys. Rev. B, 1991, 44, 7888-7903.
(5) Zhang, Y.; Yang, W. Comment on “Generalized Gradient Approximation Made
Simple”. Phys. Rev. Lett. 1998, 80, 890-890.
(6) Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation
Made Simple. Phys. Rev. Lett. 1996, 77, 3865-3868.
(7) Grimme, S.; Ehrlich, S.; Goerigk, L. Effect of the Damping Function in
Dispersion Corrected Density Functional Theory. J. Comput. Chem. 2011, 32,
1456-1465.
(8) Van Lenthe, E.; Baerends, E. J. Optimized Slater-Type Basis Sets for the Elements
1-118. J. Comput. Chem. 2003, 24, 1142-1156.
S-16
(9) Philipsen, P. H. T.; Van Lenthe, E.; Snijders, J. G.; Baerends, E. J. Relativistic
Calculations on the Adsorption of CO on the (111) Surfaces of Ni, Pd, and Pt within
the Zeroth-Order Regular Approximation. Phys. Rev. B: Condens. Mater. Phys. 1997,
56, 13556-13562.
(10) Zaharioudakis, D. Quadratic and Cubic Tetrahedron Methods for Brillouin Zone
Integration. Comput. Phys. Commun. 2005, 167, 85-89.
(11) ADF manual, release 2017, 125-127, https://www.scm.com/doc/ADF/index.html.
(12) Nørskov, J. K.; Bligaard, T.; Logadottir, A.; Kitchin, J. R.; Chen, J. G.; Pandelov,
S.; Stimming, U. Trends in the Exchange Current for Hydrogen Evolution.
J. Electrochem. Soc. 2005, 152, J23-J26.
(13) Mir, S. H.; Chakraborty, S.; Wärnå, J.; Narayan, S.; Jha, P. C.; Jha, P. K.; Ahuja,
R. A Comparative Study of Hydrogen Evolution Reaction on Pseudo-Monolayer WS2
and PtS2: Insights Based on the Density Functional Theory. Catal. Sci. Technol. 2017,
7, 687-692.

S-17

Você também pode gostar