ArticlePDF Available

Hollow cobalt phosphide octahedral pre-catalysts with exceptionally high intrinsic catalytic activity for electro-oxidation of water and methanol

Authors:
  • Dalian Institute of Chemical Physics, Chinese Academy of Sciences, China

Abstract and Figures

Microstructural engineering is an effective approach to improving electrocatalytic activity of a catalyst. Shape-controlled hollow nanostructures represent a class of interesting architectures for use in electrocatalysis, given that they may offer large surface area, preferably exposed active sites, reduced diffusion pathways for both charge and mass transport, as well as enhanced catalytic activity due to the nano-cavity effect. Herein, we for the first time report the synthesis of hollow cobalt phosphide nanoparticles with well-defined octahedral shape (CoP OCHs) via multi-step reactions. When used as pre-catalysts for the oxygen evolution reaction (OER) in alkaline solution, the as-obtained CoP OCHs not only show high apparent catalytic activity requiring an overpotential of only 240 mV to deliver the benchmark current density of 10 mA cm-2, but also exhibit an exceptionally high catalyst surface-area-based turnover frequency (TOF) of 17.6 s-1 and a catalyst mass-based TOF of 0.072 s-1 at a low overpotential of 300 mV, demonstrating excellent intrinsic catalytic activity. Moreover, the CoP OCHs also show high electrocatalytic activity for the methanol oxidation reaction (MOR), outperforming most metal phosphide based MOR catalysts reported so far. The synthetic strategy reported here can be readily extended to prepare other hollow shape-controlled metal phosphide catalysts.
Content may be subject to copyright.
Hollow cobalt phosphide octahedral pre-catalysts
with exceptionally high intrinsic catalytic activity
for electro-oxidation of water and methanol
Junyuan Xu,
a
Yuefeng Liu,
b
Junjie Li,
a
Isilda Amorim,
a
Bingsen Zhang,
c
Dehua Xiong,
d
Nan Zhang,
a
Sitaramanjaneya Mouli Thalluri,
a
Juliana P. S. Sousa
a
and Lifeng Liu *
a
Microstructural engineering is an eective approach to improving
electrocatalytic activity of a catalyst. Shape-controlled hollow nano-
structures represent a class of interesting architectures for use in
electrocatalysis, given that they may oer large surface area, prefer-
ably exposed active sites, reduced diusion pathways for both charge
and mass transport, as well as enhanced catalytic activity due to the
nano-cavity eect. Herein, we for the rst time report the synthesis of
hollow cobalt phosphide nanoparticles with a well-dened octahedral
shape (CoP OCHs) via multi-step reactions. When used as pre-
catalysts for the oxygen evolution reaction (OER) in an alkaline
solution, the as-obtained CoP OCHs not only show high apparent
catalytic activity requiring an overpotential of only 240 mV to deliver
the benchmark current density of 10 mA cm
2
, but also exhibit an
exceptionally high catalyst surface-area-based turnover frequency
(TOF) of 17.6 s
1
and a catalyst mass-based TOF of 0.072 s
1
at a low
overpotential of 300 mV, demonstrating excellent intrinsic catalytic
activity. Moreover, the CoP OCHs also show high electrocatalytic
activity for the methanol oxidation reaction (MOR), outperforming
most metal phosphide based MOR catalysts reported so far. The
synthetic strategy reported here can be readily extended to prepare
other hollow shape-controlled metal phosphide catalysts.
Introduction
Electrochemical oxidation of water or alcoholic molecules is
a key reaction in many energy conversion devices (e.g. water
electrolyzers,
1
fuel cells,
2
and metalair batteries
3
) and indus-
trial processes.
4
Conventionally, platinum group metal (PGM)
catalysts are needed for water and alcohol oxidation to achieve
practically high rates. However, the prohibitive cost and low
natural abundance of PGMs substantially limit the widespread
deployment of the above-mentioned energy conversion devices.
In this context, considerable eort has recently been devoted to
developing inexpensive and earth-abundant electrocatalysts for
the oxygen evolution reaction (OER)
1
and alcohol electro-
oxidation reaction (EOR),
58
respectively, aiming to achieve
electrochemical performance comparable to or exceeding that
of PGM catalysts. In particular, transition metal phosphides
(TMPs) have emerged as promising alternative candidates to
PGMs, owing to their metalloid characteristics, high intrinsic
catalytic activity and reasonably good chemical stability, which
are of great benet for electrocatalysis.
9,10
Given that the activity of electrocatalysts is largely dependent
on their surface properties, a lot of research studies have
focused on structural engineering of catalysts to preferably
expose as many catalytically active sites as possible. Specically,
TMPs with various nano-architectures including nano-
particles,
1013
nanowires,
1418
nanotubes,
1921
nanorods,
6,2224
nanosheets,
2528
sea-urchin,
29
polyhedra,
30,31
coreshell,
32,33
porous
7,34,35
and hollow structures
3640
have been reported, and
they mostly demonstrated good electrocatalytic performance for
the OER
10,1223,2527,2940
and EOR.
57
Hollow structured catalysts,
among others, are of particular interest for electrocatalysis, in
that they can oer large surface area, abundant catalytically
active sites, reduced diusion lengths for both mass and charge
transport, and are able to eectively prevent agglomeration and
surface area loss during the reaction.
41
Moreover, it has been
demonstrated that electrocatalytic reactions can be enhanced in
the nanoscale cavities of hollow catalysts due to the conne-
ment eect.
42,43
However, hollow TMP-based OER and EOR
electrocatalysts, especially those having a well-dened shape,
have been rarely studied so far.
30,3640
In this work, we report the synthesis of hollow cobalt phos-
phide (CoP) nanoparticles (NPs) with a well-dened octahedral
shape. This involves hydrothermal synthesis of cobalt
monoxide (CoO) nano-octahedra (OCHs), followed by surface
a
International Iberian Nanotechnology Laboratory (INL), Av. Mestre Jose Veiga,
4715-330 Braga, Portugal. E-mail: lifeng.liu@inl.int
b
Dalian National Laboratory for Clean Energy, Dalian Institute of Chemical Physics,
Chinese Academy of Sciences, 116023, Dalian, China
c
Shenyang National Laboratory for Materials Science and Institute of Metal Research,
Chinese Academy of Sciences, 110016, Shenyang, China
d
State Key Laboratory of Silicate Materials for Architectures, Wuhan University of
Technology, 430070, Wuhan, China
Electronic supplementary information (ESI) available. See DOI:
10.1039/c8ta07958g
Cite this: DOI: 10.1039/c8ta07958g
Received 16th August 2018
Accepted 2nd October 2018
DOI: 10.1039/c8ta07958g
rsc.li/materials-a
This journal is © The Royal Society of Chemistry 2018 J. Mater. Chem. A
Journal of
Materials Chemistry A
COMMUNICATION
Published on 02 October 2018. Downloaded by International Iberian Nanotechnology Laboratory on 10/16/2018 9:42:04 AM.
View Article Online
View Journal
oxidation and phosphorization treatment and a subsequent
chemical etching process. We demonstrate that the as-obtained
CoP OCHs can serve as highly ecient pre-catalysts for the OER
in an alkaline electrolyte, exhibiting outstanding apparent
activity with an overpotential of h
10
¼240 mV to deliver
10 mA cm
2
, exceptionally high surface-specic turnover
frequency (TOF) of 17.6 s
1
at h¼300 mV, and very good
catalytic stability. Furthermore, the CoP OCHs are also capable
of catalyzing the methanol oxidation reaction (MOR), showing
catalytic performance better than that of other TMP-based MOR
catalysts reported before.
57
Results and discussion
The synthesis of hollow CoP OCHs is schematically illustrated
in Scheme 1. CoO NPs with a uniform size and a well-dened
octahedral shape were rst obtained through a simple hydro-
thermal method. The collected CoO OCHs were then annealed
at 300 C in air for 2 h, which resulted in the formation of
a CoO@Co
3
O
4
core/shell nanostructure. Subsequently, the
resulting CoO@Co
3
O
4
OCHs were phosphorized at the same
temperature using sodium hypophosphite (NaH
2
PO
2
) as the
phosphorus source and high-purity nitrogen as the carrier gas,
during which the surface oxide layer was converted to a phos-
phide shell (denoted as CoO@CoP). In addition, during phos-
phorization voids will be created due to unequal diusion rates
of CoP and CoO at their interface (i.e. the Kirkendall eect), and
some cobalt phosphate would form because of the inter-
diusion of phosphor and oxygen elements. In the nal
etching step, hydrochloric acid (HCl, 0.5 M) can penetrate into
the OCHs through either the voids or porous channels gener-
ated by the dissolution of cobalt phosphate, leaching the CoO
cores out and eventually leading to the formation of hollow
porous CoP OCHs. It is worth mentioning that surface oxidation
of CoO OCHs is a necessary step to obtain well-dened hollow
CoP OCHs. Direct phosphorization of CoO would lead to the
formation of non-uniform CoO@CoP particles with an ill-
dened shape and to signicant agglomeration, as shown in
Fig. S1 (ESI).It is believed that the Co
3
O
4
layer plays a crucial
role in stabilizing the octahedral structure during the phos-
phorization treatment. The post-annealing temperature
(i.e. 300 C) of CoO OCHs was chosen according to our
thermogravimetric-dierential scanning calorimetry in air
(TG-DSC, Fig. S2, ESI), where we found that CoO starts to get
oxidized above 250 C. An attempt to perform surface oxidation
at a higher temperature, e.g., 350 C, was also made. However,
no phase-pure CoP OCHs could be obtained in this case, and
a certain amount of Co
3
O
4
still remained aer the nal
chemical etching process in 0.5 M HCl, albeit the nal product
also exhibited a hollow octahedral morphology (Fig. S3, ESI).
The initial CoO and nal hollow CoP OCHs as well as the two
intermediate core/shell nanostructures were comprehensively
characterized and analyzed by X-ray diractometry (XRD),
scanning and transmission electron microscopy (SEM & TEM)
equipped with energy-dispersive X-ray spectroscopy (EDX), and
X-ray photoelectron spectroscopy (XPS).
XRD examination conrmed that the octahedral NPs ob-
tained upon the hydrothermal synthesis consist exclusively of
cubic phase CoO (ICDD no. 00-048-1719, Fig. S4, ESI). Surface
oxidation at 300 C resulted in the formation of a cubic Co
3
O
4
(ICDD no. 00-043-1003) surface layer, and the surface layer was
completely converted to orthorhombic CoP (ICDD no. 00-029-
0497) aer phosphorization, as unambiguously illustrated in
the XRD patterns of CoO@Co
3
O
4
and CoO@CoP (Fig. S4, ESI).
Aer chemical etching, all diraction peaks resulting from CoO
disappeared and the nal product can be assigned to phase-
pure CoP.
The morphology, microstructure and composition of CoO,
CoO@Co
3
O
4
, CoO@CoP and hollow CoP OCHs were further
examined by SEM and TEM. SEM observations revealed that
CoO NPs with an average size (vertex to vertex) of 170 nm and
a well-dened octahedral shape can be obtained on a large scale
(Fig. S5a, ESI). These octahedral NPs are single-crystalline, as
conrmed by the electron diraction (ED) pattern and the high-
resolution TEM (HRTEM) image shown in Fig. 1b and c,
respectively. TEM elemental analysis illustrated that both Co
and O are evenly distributed over a single OCH (Fig. 1d). The
surface oxidation and subsequent phosphorization treatment
do not markedly change the morphology and bulk crystallinity
of the octahedral NPs (Fig. S5b, S5c, S6 and S7, ESI), as
Scheme 1 Schematic illustration of the synthesis procedure of hollow CoP OCH pre-catalysts.
Fig. 1 TEM characterization of CoO OCHs. (a) TEM image, (b) SAED
pattern, (c) HRTEM image, and (d) HAADF image and elemental
mappings of Co, O and their overlap.
J. Mater. Chem. A This journal is © The Royal Society of Chemistry 2018
Journal of Materials Chemistry A Communication
Published on 02 October 2018. Downloaded by International Iberian Nanotechnology Laboratory on 10/16/2018 9:42:04 AM.
View Article Online
evidenced by the fact that the diractions resulting from CoO
still dominate in the ED pattern acquired from a single
CoO@Co
3
O
4
or CoO@CoP (Fig. S6b and S7b, ESI). However,
HRTEM imaging indeed showed the formation of a 34nm
Co
3
O
4
layer aer post-annealing of CoO NPs in air and the
conversion of Co
3
O
4
to a porous layer aer the phosphorization
(Fig. S6c and S7c, ESI). The latter was also unambiguously
conrmed by the TEM elemental mapping (Fig. S7d, ESI).
Moreover, according to HRTEM imaging (Fig. S7c, ESI) and
EDX line scan (Fig. S8, ESI) analyses of CoO@CoP, there is an
amorphous phase between the CoO core and the formed CoP
layer, which likely consists of cobalt phosphate that results from
the outward diusion of oxygen and inward diusion of phos-
phorus during the phosphorization treatment.
19
Hollow CoP inheriting the well-dened octahedral shape of
the initial CoO OCHs was obtained aer chemical etching of
CoO@CoP. Both SEM and TEM observations conrmed that the
CoO cores were completely gone aer etching (Fig. S5dand
2a). ED and HRTEM examinations showed that hollow
OCHs comprise exclusively orthorhombic CoP (Fig. 2b and c),
consistent with the XRD result. EDX analyses revealed that aer
chemical etching the P signal was substantially increased while
the O signal signicantly reduced, indicating that pure CoP was
obtained (Fig. S9, ESI). The weak O peak might result from the
surface oxidation upon exposing the sample to air.
Furthermore, we prepared porous CoP nanospheres (NSs)
comprising randomly oriented ne crystallites and used them
as a control pre-catalyst when assessing the electrocatalytic
performance of hollow CoP OCHs, in order to scrutinize the
inuence of the microstructure on the catalytic activity. These
porous CoP NSs were synthesized by phosphorization of the
Co-glycorate precursor (Fig. S10 and S11, ESI).
44
They have an
average diameter of ca. 240 nm and consist exclusively of
orthorhombic CoP, the same as that of hollow CoP OCHs.
Moreover, the porous CoP NSs have similar crystallinity to that
of hollow CoP OCHs (1.06% vs. 1.15%), according to our
quantitative analysis using Highscore soware (Fig. S12, ESI).
In addition, XPS analyses revealed that both hollow CoP OCHs
and porous CoP NSs have essentially similar surface chemistry
(Fig. S13, ESI).
The OER activity of hollow CoP OCHs and porous CoP NSs
was investigated in 1.0 M KOH electrolyte using cyclic voltam-
metry (CV), electrochemical impedance spectroscopy (EIS) and
chronopotentiometry (CP). For comparison, the electrocatalytic
performance of CoO OCHs, CoO@Co
3
O
4
, CoO@CoP and
commercial RuO
2
NP control catalysts was also measured under
the same conditions. All catalysts were loaded on carbon paper
(CP) substrates, and the loading density was optimized to be
0.5 mg cm
2
(Fig. S14, ESI). Prior to the catalytic test, pre-
activation was carried out by repetitive CV scans at 5 mV s
1
in the potential range of 1.01.7 V vs. reversible hydrogen
electrode (RHE) until a steady state CV curve was obtained
(Fig. S15, ESI). Fig. 3a shows the cathodic branches of iR-cor-
rected CV curves of all samples. A bare CP substrate only
generates negligible current density, suggesting that it's not
catalytically active towards the OER. The overpotential (h
10
)
needed to deliver the benchmark current density of 10 mA cm
2
is broadly used as an indicator to compare the apparent cata-
lytic activity. The hollow CoP OCHs only need a h
10
of 240 mV to
deliver 10 mA cm
2
, substantially lower than that of porous CoP
NSs (h
10
¼280 mV) and commercial RuO
2
NPs (h
10
¼310 mV);
moreover, they can aord a high current density of
290 mA cm
2
at h¼320 mV. In addition, the hollow CoP OCHs
show OER activity signicantly better than that of CoO@CoP,
CoO and CoO@Co
3
O
4
control catalysts (Fig. S15, ESI), indi-
cating that phosphorization and microstructure engineering
indeed substantially improve the catalytic performance.
Furthermore, the apparent OER activity of hollow CoP OCHs
also outperforms that of many other mono-metallic phosphide
catalysts reported recently in the literature, such as CoP (h
10
¼
248,
14
290,
45
320 mV
46
), Ni
2
P(h
10
¼290
15
or 280 mV
47
) and FeP
(h
10
¼280,
19
288
20
or 350 mV
25
), and remarkably it is even
superior to that of some bi- or tri-metallic phosphides that are
supposed to have better performance than mono-metallic
phosphides due to the synergy between transition
metals.
12,13,17,21,23,25,2932,3440,4852
A detailed comparison between
the hollow CoP OCHs and other catalysts is summarized in
Table S1 (ESI).
While the apparent OER activity is heavily dependent on the
loading mass, mass activity and specic activity, which are
obtained through normalizing the catalytic current by the mass
of catalysts and the real catalytically-active surface area,
respectively, can better reect the utilization of catalysts and the
intrinsic activity of catalytic materials. To this end, the mass
activities of hollow CoP OCHs and control catalysts are
compared (Fig. 3b), and hollow CoP OCHs are found to exhibit
a mass activity of 581 A g
CoP
1
at h¼320 mV, substantially
higher than those of porous CoP NSs (91 A g
CoP
1
) and RuO
2
NPs (31 A g
RuO
2
1
) at the same overpotential. Furthermore,
the specic activities of all catalysts were calculated upon
normalizing the catalytic current by their corresponding
electrochemically-accessible surface area (i.e. ECSA, see
Experimental details, Fig. S16, ESI).
53,54
Hollow CoP OCHs
exhibit an ECSA of 67.5 cm
catalyst2
, similar to that of RuO
2
NPs
(75.0 cm
catalyst2
) but remarkably higher than that of porous CoP
NSs (40.0 cm
catalyst2
). Nevertheless, aer normalization, hollow
CoP OCHs still show the best specic activity, able to deliver
Fig. 2 TEM characterization of hollow CoP OCHs. (a) TEM image, (b)
SAED pattern, (c) HRTEM image, and (d) HAADF-STEM image of CoP
OCHs and elemental mappings of Co, P and their overlap.
This journal is © The Royal Society of Chemistry 2018 J. Mater. Chem. A
Communication Journal of Materials Chemistry A
Published on 02 October 2018. Downloaded by International Iberian Nanotechnology Laboratory on 10/16/2018 9:42:04 AM.
View Article Online
4.3 mA cm
catalyst
2
at h¼320 mV (Fig. 3c), ca. 4 and 20 times
more active than porous CoP NSs and RuO
2
NPs, respectively. It
is worth noting that hollow CoP OCHs and porous CoP NSs have
essentially the same crystal phase (Fig. S12, ESI) and surface
chemistry (Fig. S13, ESI), and similar crystallinity and electrical
conductivity (as illustrated by the equivalent series resistance R
s
of 0.45 Ucm
2
for hollow CoP OCHs vs. 0.52 Ucm
2
for porous
CoP NSs, Fig. S17, ESI), and therefore it is assumed that the
high intrinsic (specic) OER activity of hollow CoP OCHs results
from their unique microstructure where more catalytically
active sites are likely preferably exposed and better mass
transport can be achieved.
The OER activity of all catalysts was further assessed using
the turnover frequency values calculated based on the catalyst
mass (TOF
mass
) and surface active sites (TOF
surface
)ath¼260,
280, 300 and 320 mV, respectively. The details about the
calculation are presented in the ESI.Assuming that all metal
species are catalytically active (i.e. the lower limit), the TOF
mass
value of hollow CoP OCHs is 0.072 s
1
at h¼300 mV, signi-
cantly higher than that of many non-precious OER catalysts
including oxides,
55,56
(oxy)hydroxides,
57,58
carbides,
59,60
chalco-
genides,
61
nitrides
62
and phosphides
12,13,15,21,22,2932,35,36,3840,5052
(Table S1, ESI). If only surface active sites are taken into
account (Fig. S18, ESI),
63,64
the TOF
surface
value of hollow CoP
OCHs can be even higher, reaching 17.6 s
1
at h¼300 mV
(Fig. 3e), substantially outperforming both porous CoP NSs and
RuO
2
NPs and indicating that the hollow CoP OCHs are
intrinsically more active for the OER.
The OER kinetics of all catalysts was investigated by Tafel
analysis. As shown in Fig. 3f, the Tafel slope of hollow CoP OCHs
is only 38 mV dec
1
, smaller than that of both porous CoP NSs
(45 mV dec
1
) and commercial RuO
2
NPs (53 mV dec
1
), sug-
gesting a more favorable OER rate at the hollow CoP OCH pre-
catalysts. The fast OER kinetics of hollow CoP OCHs was also
veried by EIS analysis (Fig. S17, ESI), where the charge transfer
resistance (R
ct
¼0.80 Ucm
2
) of hollow CoP OCHs measured at
h¼260 mV is remarkably lower than that of porous CoP NSs
(6.4 Ucm
2
) and commercial RuO
2
NPs (10.6 Ucm
2
).
Furthermore, the catalytic stability as a crucial performance
indicator was assessed using chronopotentiometry at a constant
Fig. 3 OER performance of hollow CoP OCH pre-catalysts measured in 1.0 M KOH electrolyte at room temperature. The OER performance of
porous CoP NSs and commercial RuO
2
NPs is given for comparison (a) iR-corrected polarization curves (i.e. the cathodic branches of the CV
curves shown in Fig. S14) recorded at a scan rate of 5 mV s
1
in the potential range of 1.01.7 V vs. RHE. (b) Mass activity. (c) Specic activity. (d)
Mass-based and (e) surface-charge-based TOF values calculated at h¼260, 280, 300 and 320 mV. (f) Tafel analysis. (g) Chronopotentiometric
curves recorded at a constant current density of 10 mA cm
2
.
J. Mater. Chem. A This journal is © The Royal Society of Chemistry 2018
Journal of Materials Chemistry A Communication
Published on 02 October 2018. Downloaded by International Iberian Nanotechnology Laboratory on 10/16/2018 9:42:04 AM.
View Article Online
current density of 10 mA cm
2
(Fig. 3g). Both hollow CoP OCHs
and porous CoP NSs showed very good stability, and could
sustain at 10 mA cm
2
with little degradation for at least 60 h. In
contrast, the overpotential needed for RuO
2
NPs to maintain
10 mA cm
2
was increased by 33 mV in 60 h, exhibiting gradual
degradation.
Besides the OER, the catalytic performance of hollow CoP
OCH pre-catalysts towards the MOR was also investigated and
compared to that of other non-noble metal based catalysts.
58
Fig. 4a shows apparent and mass-specic MOR activities of
hollow CoP OCHs and porous CoP NSs measured in a mixed
solution of 1.0 M KOH and 1.0 M CH
3
OH (Fig. S19, ESI). A
remarkable decrease in the onset potential (dened as the
potential at 1 mA cm
2
)ofca. 170 mV was observed for both
OCH and NS pre-catalysts, suggesting that they are catalytically
active for the MOR.
57
The mass activity of hollow CoP OCHs is
206.2 A g
CoP
1
at 0.4 V vs. saturated calomel electrode (SCE),
much higher than that of porous CoP NSs (69.7 A g
CoP
1
) and
other CoP catalysts reported previously.
57
Furthermore, hollow
CoP OCHs show a specic activity of 1.53 mA cm
2
at 0.4 V vs.
SCE, almost two times higher than that of porous CoP NSs
(0.87 mA cm
2
), implying that the hollow octahedral CoP is
intrinsically more active for the MOR. A detailed comparison
between hollow CoP OCHs and other non-precious MOR cata-
lysts reported in the literature is summarized in Table S2 (ESI).
As far as the MOR kinetics is concerned, the Tafel slope of
hollow OCHs is smaller than that of porous NSs, showing
a favorable reaction rate (Fig. 4c). This was also corroborated by
the EIS analysis (Fig. 4d), where the R
ct
of hollow OCHs
(1.2 Ucm
2
)isca. 3 times smaller than that of porous NSs
(3.1 Ucm
2
), suggesting that the MOR occurs faster on OCH
pre-catalysts. The catalytic stability of hollow CoP OCHs and
porous CoP NSs was investigated by continuous CV scans in the
potential range of 00.35 V vs. SCE at a scan rate of 100 mV s
1
.
Upon a given number of CV cycles, the catalytic current density
at 0.44 V vs. SCE and the R
ct
value extracted from the EIS Nyquist
plot (Fig. S20, ESI) were compared to the initial values (Fig. 4a
and d). As shown in Fig. 4e and f, almost no decay in the current
density and R
ct
aer 2500 cycles are observed for both hollow
CoP OCHs and porous CoP NSs, demonstrating very good
stability of these catalysts. A further increase in cycle number to
4000 results in a decrease in the current density (to 70 and 72%
of its initial value for OCHs and NSs, respectively) and an
increase in R
ct
; however, this can be recovered to a large extent
by replenishing the electrolyte (Fig. S20, ESI), indicating that
the observed performance decay may be mostly due to the
consumption of methanol.
The composition and morphology of hollow CoP OCHs
subjected to extended OER and MOR stability tests were
examined (Fig. S21S24). The P 2p XPS spectrum shows that
the binding energy (BE) peak originating from phosphide
disappears, and only a weak PO peak remains. In the Co 2p
3/2
spectrum, a predominant peak located at around 780 eV is
observed, which can be assigned to the BE of Co in
CoOOH.
13,53,65
Moreover, the CoO and OH components
appearing in the deconvoluted O 1s XPS spectrum represent the
oxygen bonds in CoOOH.
13,53,65
The XPS analyses conrm that
the initial CoP has been mostly converted to CoOOH upon the
extended OER and MOR electrolysis. This agrees well with
previous reports on phosphide-based OER catalysts where the
Fig. 4 MOR performance of the CoP pre-catalysts. (a) Apparent and mass activities of hollow CoP OCHs (orange) and porous CoP NSs (olive)
recorded at a scan rate of 5 mV s
1
in 1.0 M KOH + 1.0 M methanol (solid lines) and 1.0 M KOH (dotted lines), respectively. (b) Specic activity. (c)
Tafel analysis. (d) Nyquist plots measured at 0.35 V vs. SCE. The scattered open circles are experimental data and the dotted lines are tting
curves. The inset shows the equivalent circuit model. The catalytic current densities recorded at 0.44 V vs. SCE and R
ct
values as a function of the
number of CV scans (0 to 0.35 V vs. SCE at 100 mV s
1
) for (e) hollow CoP OCH and (f) porous CoP NS pre-catalysts.
This journal is © The Royal Society of Chemistry 2018 J. Mater. Chem. A
Communication Journal of Materials Chemistry A
Published on 02 October 2018. Downloaded by International Iberian Nanotechnology Laboratory on 10/16/2018 9:42:04 AM.
View Article Online
progressive in situ transformation of phosphide to corre-
sponding (oxy)hydroxide under alkaline OER conditions had
been observed by many groups.
10,14,15,4552
Our SEM-EDX anal-
yses further corroborate the leach of P and substantial increase
of the O signal for the tested catalysts. Although the octahedral
shape is not visible anymore, the catalysts remain separate
without signicant aggregation and interwoven layered ne
structures are found to form. The composition and morphology
changes were also observed for porous CoP NS pre-catalysts, as
shown in Fig. S25S28 (ESI).
It is worth stressing that notwithstanding the composition
and morphology transformation upon extended electro-
oxidation, the long-term electrocatalytic performance is
predominantly governed by the initial performance of TMP
pre-catalysts. Namely, once the pre-catalystdemonstrates
outstanding catalytic activity, this catalytic performance will be
retained during the subsequent long-term electro-oxidation
process, even if the morphology, crystal structure and compo-
sition of pre-catalystswill signicantly change in this process.
This phenomenon has already been widely observed for non-
oxide based OER catalysts,
10,14,15,4449,53,6671
though how the
catalytic activity is balanced during the very complex morpho-
logical, structural and compositional transformation has been
unclear so far (A compensation mechanism might exist). In this
sense, developing high-performance pre-catalysts, like the
hollow CoP OCHs presented in this work, remains signicant.
Conclusions
In summary, we have for the rst time synthesized hollow cobalt
phosphide octahedral nanoparticles and utilized them as pre-
catalysts for electro-oxidation of water and methanol.
Beneting from the unique structural features, the hollow cobalt
phosphide nano-octahedron pre-catalysts show both high
apparent and high intrinsic catalytic activities for the oxygen
evolution and methanol oxidation reactions, in comparison to
cobalt phosphide nanospheres having the same crystal structure
and surface chemistry as well as similar feature sizes, crystallinity
and electrical conductivity. Although the morphology, structure
and composition of the hollow cobalt phosphide octahedra
signicantly changed upon the extended stability test, the
outstanding catalytic performance was inherited which endows
the transformed catalysts to drive the oxygen evolution and
methanol oxidation reactions for a long time without remarkable
degradation. The hollow cobalt phosphide nano-octahedra hold
a substantial promise for eciently catalyzing the electro-
oxidation of water and methanol in water electrolyzers and
direct methanol fuel cells, respectively.
Conicts of interest
There are no conicts to declare.
Acknowledgements
This work was nancially supported by the European Horizon
2020 project CritCatunder the grant agreement number
686053. L. F. L. acknowledges the nancial support from the
Portuguese Foundation of Science and Technology (FCT) under
the projects IF/2014/01595and PTDC/CTM-ENE/2349/2014
(grant agreement No. 016660).
Notes and references
1 N. T. Suen, S. F. Hung, Q. Quan, N. Zhang, Y. J. Xu and
H. M. Chen, Chem. Soc. Rev., 2017, 46, 337.
2 B. C. Ong, S. K. Kamarudin and S. Basri, Int. J. Hydrogen
Energy, 2017, 42, 10142.
3 Z. L. Wang, D. Xu, J. J. Xu and X. B. Zhang, Chem. Soc. Rev.,
2014, 43, 7746.
4 A. Anglada, A. Urtiaga and I. Ortiz, J. Chem. Technol.
Biotechnol., 2009, 84, 1747.
5 D. N. Liu, W. B Liu, K. Y. Wang, G. Du, A. M. Asiri, Q. Lu and
X. P. Sun, Nanotechnology, 2016, 27, 44LT02.
6 J. Zhou, Y. B. Dou, A. Zhou, R. M. Guo, M. J. Zhao and J. R. Li,
Adv. Energy Mater., 2017, 7, 1602643.
7 C. Z. Zhu, S. F. Fu, B. Z. Xu, J. H. Song, Q. R. Shi,
M. H. Engelhard, X. L. Li, S. P. Beckman, J. M. Sun, D. Du
and Y. H. Lin, Small, 2017, 13, 1700796.
8 X. Zou, A. Goswami and T. Asefa, J. Am. Chem. Soc., 2013, 135,
17242.
9 S. Carenco, D. Portehault, C. Boissi`
ere, N. M´
ezailles and
C. Sanchez, Chem. Rev., 2013, 113, 7981.
10 J. Y. Xu, J. J. Li, D. H. Xiong, B. S. Zhang, Y. F. Liu, K. H. Wu,
I. Amorim, W. Li and L. F. Liu, Chem. Sci., 2018, 9, 3470.
11 E. J. Popczun, C. G. Read, C. W. Roske, N. S. Lewis and
R. E. Schaak, Angew. Chem., Int. Ed., 2014, 53, 5427.
12 A. Mendoza-Garcia, H. Y. Zhu, Y. S. Yu, Q. Li, L. Zhou, D. Su,
M. J. Kramer and S. H. Sun, Angew. Chem., Int. Ed., 2015, 54,
9642.
13 D. Li, H. Baydoun, C. N. Vernai and S. L. Brock, J. Am. Chem.
Soc., 2016, 138, 4006.
14 W. Li, X. F. Gao, D. H. Xiong, F. Xia, J. Liu, W. G. Song,
J. Y. Xu, S. M. Thalluri, M. F. Cerqueira, X. L. Fu and
L. F. Liu, Chem. Sci., 2017, 8, 2952.
15 L. A. Stern, L. G. Feng, F. Song and X. L. Hu, Energy Environ.
Sci., 2015, 8, 2347.
16 W. Li, X. F. Gao, X. G. Wang, D. H. Xiong, P. P. Huang,
W. G. Song, X. Q. Bao and L. F. Liu, J. Power Sources, 2016,
330, 156.
17 J. J. Duan, S. Chen, A. Vasileand S. Z. Qiao, ACS Nano, 2016,
10, 8738.
18 W. Li, D. H. Xiong, X. F. Gao, W. G. Song, F. Xia and L. F. Liu,
Catal. Today, 2017, 287, 122.
19 J. Y. Xu, D. H. Xiong, I. Amorim and L. F. Liu, ACS Appl. Nano
Mater., 2018, 1, 617.
20 Y. Yan, B. Y. Xia, X. M. Ge, Z. L. Liu, A. Fisher and X. Wang,
Chem.Eur. J., 2015, 21, 18062.
21 L. T. Yan, L. Cao, P. C. Dai, X. Gu, D. D. Liu, L. J. Li, Y. Wang
and X. B. Zhao, Adv. Funct. Mater., 2017, 27, 1703455.
22 D. H. Xiong, X. G. Wang, W. Li and L. F. Liu, Chem. Commun.,
2016, 52, 8711.
23 D. Li, H. Baydoun, B. Kulikowski and S. L. Brock, Chem.
Mater., 2017, 29, 3048.
J. Mater. Chem. A This journal is © The Royal Society of Chemistry 2018
Journal of Materials Chemistry A Communication
Published on 02 October 2018. Downloaded by International Iberian Nanotechnology Laboratory on 10/16/2018 9:42:04 AM.
View Article Online
24 X. G. Wang, Y. V. Kolen'ko and L. F. Liu, Chem. Commun.,
2015, 51, 6738.
25 X. G. Wang, W. Li, D. H. Xiong and L. F. Liu, J. Mater. Chem.
A, 2016, 4, 5639.
26 X. F. Xiao, C. T. He, S. L. Zhao, J. Li, W. S. Lin, Z. K. Yuan,
Q. Zhang, S. Y. Wang, L. M. Dai and D. S. Yu, Energy
Environ. Sci., 2017, 10, 893.
27 Y. J. Li, H. C. Zhang, M. Jiang, Q. Zhang, P. L. He and
X. M. Sun, Adv. Funct. Mater., 2017, 27, 1702513.
28 X. G. Wang, Y. V. Kolen'ko, X. Q. Bao, K. Kovnir and L. F. Liu,
Angew. Chem., Int. Ed., 2015, 54, 8188.
29 A. Mendoza-Garcia, D. Su and S. H. Sun, Nanoscale, 2016, 8, 3244.
30 M. J. Liu and J. H. Li, ACS Appl. Mater. Interfaces, 2016, 8,
21582165.
31 F. Li, Y. F. Bu, Z. J. Lv, J. Mahmood, G. F. Han, I. Ahmad,
G. Kim, Q. Zhong and J. B. Baek, Small, 2017, 13, 1701167.
32 Y. Pan, K. A. Sun, S. J. Liu, X. Gao, K. L. Wu, W. C. Cheong,
Z. Chen, Y. Wang, Y. Li, Y. Q. Liu, D. S. Wang, Q. Peng,
C. Chen and Y. D. Li, J. Am. Chem. Soc., 2018, 140, 2610.
33 Y. H. Hou, Y. P. Liu, R. Q. Gao, Q. J. Li, H. Z. Guo,
A. Goswami, R. Zboril, M. B. Gawande and X. X. Zou, ACS
Catal., 2017, 7, 7038.
34 Y. W. Tan, H. Wang, P. Liu, Y. H. Shen, C. Cheng, A. Hirata,
T. Fujita, Z. Tang and M. W. Chen, Energy Environ. Sci., 2016,
9, 2257.
35 S. F. Fu, C. Z. Zhu, J. H. Song, M. H. Engelhard, X. L. Li, D. Du
and Y. H. Lin, ACS Energy Lett., 2016, 1, 792.
36 P. L. He, X. Y. Yu and X. W. Lou, Angew. Chem., Int. Ed., 2017,
56, 3897.
37 X. P. Zhang, L. Huang, Q. Q. Wang and S. J. Dong, J. Mater.
Chem. A, 2017, 5, 18839.
38 X. Y. Yu, Y. Feng, B. Y. Guan, X. W. Lou and U. Paik, Energy
Environ. Sci., 2016, 9, 1246.
39 H. H. Zou, C. Z. Yuan, H. Y. Zou, T. Y. Cheng, S. J. Zhao,
Y. Qazi, S. L. Zhong, L. Wang and A. W. Xu, Catal. Sci.
Technol., 2017, 7, 1549.
40 L. Zhang, C. Chang, C. W. Hsu, C. W. Chang and S. Y. Lu,
J. Mater. Chem. A, 2017, 5, 19656.
41 L. Yu, H. Hu, H. B. Wu and X. W. Lou, Adv. Mater., 2017, 29,
1604563.
42 M. A. Mahmoud, F. Saira and M. A. El-Sayed, Nano Lett.,
2010, 10, 3764.
43 M. A. Mahmoud and M. A. El-Sayed, Nano Lett., 2011, 11, 946.
44 H. Liu, F. X. Ma, C. Y. Xu, L. Yang, Y. Du, P. P. Wang, S. Yang
and L. Zhen, ACS Appl. Mater. Interfaces, 2017, 9, 11634.
45 Y. P. Zhu, Y. P. Liu, T. Z. Ren and Z. Y. Yuan, Adv. Funct.
Mater., 2015, 25, 7337.
46 J. F. Chang, Y. Xiao, M. L. Xiao, J. J. Ge, C. P. Liu and W. Xing,
ACS Catal., 2015, 5, 6874.
47 J. Y. Xu, X. K. Wei, J. D. Costa, J. L. Lado, B. Owens-Bair,
L. P. L. Goncalves, S. P. S. Fernandes, M. Heggen,
D. Y. Petrovykh, R. E. Dunin-Borkowski, K. Kovnir and
Y. V. Kolen'ko, ACS Catal., 2017, 7, 5450.
48 H. F. Liang, A. N. Gandi, D. H. Anjum, X. B. Wang,
U. Schwingenschl¨
ogl and H. N. Alshareef, Nano Lett., 2016,
16, 7718.
49 J. Yu, Q. Q. Li, Y. Li, C. Y. Xu, L. Zhen, V. P. Dravid and
J. S. Wu, Adv. Funct. Mater., 2016, 26, 7644.
50 Z. Zhang, J. H. Hao, W. S. Yang and J. L. Tang, RSC Adv.,
2016, 6, 9647.
51 G. Zhang, G. C. Wang, Y. Liu, H. J. Liu, J. H. Qu and J. H. Li,
J. Am. Chem. Soc., 2016, 138, 14686.
52 J. Y. Li, M. Yan, X. M. Zhou, Z. Q. Huang, Z. M. Xia,
C. R. Chang, Y. Y. Ma and Y. Q. Qu, Adv. Funct. Mater.,
2016, 26, 6785.
53 W. Li, X. F. Gao, D. H. Xiong, F. Wei, W. G. Song, J. Y. Xu and
L. F. Liu, Adv. Energy Mater., 2017, 7, 1602579.
54 C. C. L. McCrory, S. Jung, I. M. Ferrer, S. M. Chatman, J. C. Peters
and T. F. Jaramillo, J. Am. Chem. Soc., 2015, 137, 4347.
55 H. Y. Jin, J. Wang, D. F. Su, Z. Z. Wei, Z. Z. Pang and Y. Wang,
J. Am. Chem. Soc., 2015, 137, 2688.
56 Y. S. Jin, H. T. Wang, J. J. Li, X. Yue, Y. J. Han, P. K. Shen and
Y. Cui, Adv. Mater., 2016, 28, 3785.
57 K. Fan, H. Chen, Y. F. Ji, H. Huang, P. M. Claesson,
Q. Daniel, B. Philippe, H. Rensmo, F. S. Li, Y. Luo and
L. C. Sun, Nat. Commun., 2016, 7, 11981.
58 H. H. Shi, H. F. Liang, F. W. Ming and Z. C. Wang, Angew.
Chem., Int. Ed., 2017, 56, 573.
59 K. Xu, H. Ding, H. F. Lv, P. Z. Chen, X. L. Lu, H. Cheng,
T. P. Zhou, S. Liu, X. J. Wu, C. Z. Wu and Y. Xie, Adv.
Mater., 2016, 28, 3326.
60 H. S. Fan, H. Yu, Y. F. Zhang, Y. Zheng, Y. B. Luo, Z. F. Dai,
B. Li, Y. Zong and Q. Y. Yan, Angew. Chem., Int. Ed., 2017, 56,
12566.
61 L. L. Feng, G. T. Yu, Y. Y. Wu, G. D. Li, H. Li, Y.H. Sun, T. Asefa,
W. Chen and X. M. Zou, J. Am. Chem. Soc., 2015, 137, 14023.
62 X. D. Jia, Y. F. Zhao, G. B. Chen, L. Shang, R. Shi, X. F. Kang,
G. I. N. Waterhouse, L. Z. Wu, C. H. Tung and T. R. Zhang,
Adv. Energy Mater., 2016, 6, 1502585.
63 D. Meki, S. Fierro, H. Vrubel and X. L. Hu, Chem. Sci., 2011, 2,
1262.
64 J. Y. Xu, T. F. Liu, J. J. Li, B. Li, Y. F. Liu, B. S. Zhang,
D. H. Xiong, I. Amorim, W. Li and L. F. Liu, Energy
Environ. Sci., 2018, 11, 1819.
65 M. S. Burke, M. G. Kast, L. Trotochaud, A. M. Smith and
S. W. Boettche, J. Am. Chem. Soc., 2015, 137, 3638.
66 S. Jin, ACS Energy Lett., 2017, 2, 1937.
67 X. Zhang, X. Zhang, H. M. Xu, Z. S. Wu, H. L. Wang and
Y. Y. Liang, Adv. Funct. Mater., 2017, 27, 1606635.
68 Y. Y. Wu, G. D. Li, Y. P. Liu, L. Yang, X. R. Lian, T. Asefa and
X. X. Zou, Adv. Funct. Mater., 2016, 26, 4839.
69 X. Xu, F. Song and X. L. Hu, Nat. Commun., 2016, 7, 12324.
70 Y. Q. Zhang, B. Ouyang, J. Xu, G. C. Jia, S. Chen, R. S. Rawat
and H. J. Fan, Angew. Chem., Int. Ed., 2016, 55, 8670.
71 D. H. Xiong, Q. Q. Zhang, S. M. Thalluri, J. Y. Xu, W. Li,
X. L. Fu and L. F Liu, Chem.Eur. J., 2017, 23, 8749.
This journal is © The Royal Society of Chemistry 2018 J. Mater. Chem. A
Communication Journal of Materials Chemistry A
Published on 02 October 2018. Downloaded by International Iberian Nanotechnology Laboratory on 10/16/2018 9:42:04 AM.
View Article Online
... In addition to geometric current density, the ECSA normalized current density also show the high value of as-synthesized electrocatalyst compared with other study. This suggests that the PtRu/Ti 3 C 2 T x has a higher intrinsic activity compared with other Pt-based electrocatalyst, indicating that there are more exposed catalytically active sites and better mass transport during the MOR [56]. The comparison of peak potential value with other electrocatalysts is also evaluated. ...
... Geometrical structure and chemical composition are the two important parameters that have a critical influence on the properties of functional materials [1][2][3][4]. NPs with zero-/one-/two-dimensional (0D/1D/2D) structures have received considerable attention in recent decades, and the fabrication of all kinds of nanostructures has been of central interest in this field [5][6][7][8][9][10][11][12][13][14][15][16][17]. Gold (Au) NPs, as one of the promising candidates for catalytic, electronic, plasmonic, molecular diagnostic and sensing applications, are well known for their tailorable size and surface chemistry, high stability and good biocompatibility [18][19][20][21][22][23]. ...
Article
Full-text available
Crystallization plays a critical role in determining crystal size, purity and morphology. Therefore, uncovering the growth dynamics of nanoparticles (NPs) atomically is important for the controllable fabrication of nanocrystals with desired geometry and properties. Herein, we conducted in situ atomic-scale observations on the growth of Au nanorods (NRs) by particle attachment within an aberration-corrected transmission electron microscope (AC-TEM). The results show that the attachment of spherical colloidal Au NPs with a size of about 10 nm involves the formation and growth of neck-like (NL) structures, followed by five-fold twin intermediate states and total atomic rearrangement. The statistical analyses show that the length and diameter of Au NRs can be well regulated by the number of tip-to-tip Au NPs and the size of colloidal Au NPs, respectively. The results highlight five-fold twin-involved particle attachment in spherical Au NPs with a size of 3–14 nm, and provide insights into the fabrication of Au NRs using irradiation chemistry.
Article
Full-text available
Efficient electrocatalysts, with high tolerance to methanol oxidation, good stability, and acceptable cost are the main requisites for promising direct methanol fuel cell (DMFC) electrode materials. This target can be achieved by the integration of different active materials with unique structures. In this work, a cobalt metal–organic framework (Co-MOF) flower structure was prepared by a hydrothermal method, and then a simple ultrasonication method was employed to anchor carbon nanotubes (CNTs) in between the MOF flower petals and fabricate a Co-MOF/CNT hybrid composite. Different ratios of CNTs were used in the composite preparations, namely 25, 50, and 75 wt% of the composite. The nanocomposites were entirely investigated using different characterization techniques, such as XRD, FTIR, SEM, TEM, and XPS. Comparative electrochemical measurements confirmed that due to the integration of highly conductive CNTs with the porous active fascinating structure of Co-MOF, Co-MOF/50% CNTs exhibited improved electrocatalytic activity with a current density of 35 mA cm⁻² at a potential of 0.335 V and a scan rate of 50 mV s⁻¹. The excellent electrochemical activity and stability could be due to the synergy between Co-MOF and the CNTs that conferred adequate active sites for methanol electro-oxidation and a lower equivalent series resistance, as revealed from the electrochemical impedance spectroscopy study. This study opens a new avenue to decrease the utilization of platinum and increase the methanol oxidation activity using low-cost catalysts.
Article
For electrocatalysts in water splitting, activity and stability is at high are pivotal current density. Thus, transition metal phosphides were prepared using ionic liquid as precursor on nickel foam, where Fe was doped and this catalyst was covered by carbon. This Fe‐CoNiP−C/NF exhibits excellent OER activity, which only needs overpotentials of 355 and 424 mV to reach the current densities of 50 and 100 mA cm ⁻² , respectively. Moreover, it obtains a good stability for more than 30 hours at 100 mA cm ⁻² . Experimental results demonstrate part of this TMPs‐based catalyst is oxidized to metal oxyhydroxide species after phosphidation, which can significantly improve the activity for OER. Also, the low Fe doping and outer carbon enhance the OER activity and stability.
Article
Controllable designing of well-defined heterojunction nanostructures provides an insightful strategy for accelerating the kinetics of the hydrogen and oxygen evolution reactions (HER/OER), but such task is still challenging. Herein, we proposed a protocol of heterojunction interface editing (HIE) strategy by oxygen atoms decoration for synergistic boosting electrocatalytic HER and OER performances. A novel Co/NiCoP nanospheres (NSs) heterojunction was synthesized by crystal seed template transformation method with Ni5P4 microspheres as seeds. The effective oxygen atoms interface editing increased the oxidation state of Co atoms and prolonged the Co-P bond length of Co/NiCoP NSs heterojunction, thus the electron localization on P sites was enhanced, leading to the dramatically elevated HER and OER performances simultaneously. The as-constructed O-Co/NiCoP NSs show excellent electrocatalytic activity with 361 and 430 mV vs. reversible hydrogen electrode (RHE) to arrive high current density of 300 mAcm−2 for HER and OER in 1 M KOH as well as good stability. The proposed HIE concept could provide a new perspective on the catalyst design for energy conversion systems.
Article
With the continuously increasing global energy demand, there is an urgent requirement to find efficient methanol oxidation reaction (MOR) catalysts that can replace precious metals. In this work, we have elaborately integrated 5,10,15,20-tetrakis(4-carboxyphenyl) porphyrin (H2TCPP) with copper cobaltate (CuCo2O4), which possesses efficient separation of photogenerated charges and increased active sites. The mass activity of H2TCPP/CuCo2O4 (534.75 mA mg-1) toward MOR is higher than that of pure CuCo2O4 (291.75 mA mg-1) under light. In addition, H2TCPP/CuCo2O4 can catalyze the oxidation of other alcohols, such as ethanol, ethanediol, isopropanol, and glycerol. This study demonstrates that it is feasible to enhance the MOR activity by the modification of bimetallic transition metal oxides with porphyrins.
Article
Developing a fast and highly active oxygen evolution reaction (OER) catalyst to change energy kinetics technology is essential for making clean energy. Herein, we prepare three‐dimensional (3D) hollow Mo‐doped amorphous FeOOH (Mo‐FeOOH) based on the precatalyst MoS2/FeC2O4 via in situ reconstruction strategy. Mo‐FeOOH exhibits promising OER performance. Specifically, it has an overpotential of 285 mV and a durability of 15 h at 10 mA cm‐2. Characterizations indicate that Mo was included inside the FeOOH lattice, and it not only modifies the electronic energy levels of FeOOH but also effectively raises the inherent activity of FeOOH for OER. Additionally, in situ Raman analysis indicates that FeC2O4 gradually transforms into the FeOOH active site throughout the OER process. This study provides ideas for designing in situ reconstruction strategies to prepare heteroatom doping catalysts for high electrochemical activity.
Article
Full-text available
Transition metal phosphides (TMPs) have recently emerged as a new class of pre-catalysts that can efficiently catalyze the oxygen evolution reaction (OER). However, how the OER activity of TMPs varies with the catalyst composition has not been systematically explored. Here, we report the alkaline OER electrolysis of a series of nanoparticulate phosphides containing different equimolar metal (M = Fe, Co, Ni) components. Notable trends in OER activity are observed, following the order of FeP < NiP < CoP < FeNiP < FeCoP < CoNiP < FeCoNiP, which indicate that the introduction of a secondary metal(s) to a mono-metallic TMP substantially boosts the OER performance. We ascribe the promotional effect to the enhanced oxidizing power of bi- and tri-metallic TMPs that can facilitate the formation of MOH and chemical adsorption of OH groups, which are the rate-limiting steps for these catalysts according to our Tafel analysis. Remarkably, the tri-metallic FeCoNiP pre-catalyst exhibits exceptionally high apparent and intrinsic OER activities, requiring only 200 mV to deliver 10 mA cm-2 and showing a high turnover frequency (TOF) of ≥ 0.94 s-1 at the overpotential of 350 mV.
Article
Full-text available
We present a facile synthetic method that yields [email protected]xP core-shelltype heterogeneous nanostructures with excellent oxygen evolution reaction (OER) activity. This nanocatalyst can deliver a current density of 10 mA/cm² at a small overpotential of 310 mV and exhibits high catalytic stability. Additionally, the catalytic activity of [email protected]xP is 8 times higher than that of the Co2P nanoparticles, owing primarily to the strong electronic interaction between the Ag core and the CoxP shell.
Article
Full-text available
Uniform Ni3C nanodots dispersed in ultrathin N-doped carbon nanosheets were successfully prepared by carburization of the two dimensional (2D) nickel-cyanide coordination polymer precursors. The Ni3C based nanosheets have lateral length of about 200 nm and thickness of 10 nm. When doped with Fe, the Ni3C based nanosheets exhibited outstanding electrocatalytic properties for both hydrogen evolution reactions (HER) and oxygen evolution reactions (OER). For example, 2 at% Fe (atomic percent) doped Ni3C nanosheets depict a low overpotential (292 mV) and a small Tafel slope (41.3 mV dec-1) for HER in KOH solution. An outstanding OER catalytic property is also achieved with a low overpotential of 275 mV and a small Tafel slope of 62 mV dec-1 in KOH solution. Such nanodots incorporated 2D hybrid structures can serve as an efficient bifunctional electrocatalyst for overall water splitting.
Article
Full-text available
The rational design and engineering of metal-organic frameworks (MOFs) with hollow structure and complex compositions has attracted increasing research interest due to their importance in various applications. Herein, we develop a facile way for in situ transformation of cobalt-based homobimetallic zeolitic imidazolate framework polyhedra (NiCo-ZIF PH) into nickel-cobalt phosphide/nitrogen-doped carbon polyhedral nanocages (NiCoP/NC PHCs) via tannic acid etching as well as calcination and phosphidation process. Impressively, the hollow structure of polyhedral nanocages formed after etching process is robust enough to endure the subsequent thermal treatment. The as-obtained NiCoP/NC PHCs are endowed with large surface area and abundant pore volume, leading to the exposure of more active sites and rapid mass transfer. Besides, the guest metal Ni is homogeneously incorporated into the host CoP crystal lattice without phase segregation, which effectively modulates the electronic structures. Benefiting from the above advantages, the NiCoP/NC PHCs exhibit obviously improved electrocatalytic activity towards OER, with a low overpotential of 297 mV to afford the current density of 10 mA cm-2 and excellent long-term stability. Moreover, the present synthetic route is easy to scale up, which is especially appeal to practical application.
Article
Full-text available
Iron (Fe)-doped porous cobalt phosphide polyhedrons are designed and synthesized as an efficient bifunctional electrocatalyst for both hydrogen evolution reaction (HER) and oxygen evolution reaction (OER). The synthesis strategy involves one-step route for doping foreign metallic element and forming porous cobalt phosphide polyhedrons. With varying doping levels of Fe, the optimized Fe-doped porous cobalt phosphide polyhedron exhibits significantly enhanced HER and OER performances, including low onset overpotentials, large current densities, as well as small Tafel slopes and good electrochemical stability during HER and OER.
Article
In this article, we for the first time report the synthesis and electrocatalytic properties of ruthenium cobalt phosphide hybrid clusters for the hydrogen evolution reaction (HER). Two types of catalysts are investigated: wet chemical reduction of Ru3+ on pre-formed cobalt phosphide (CoP) nanoparticles results in ruthenium-cobalt phosphide side-by-side structures (Ru/CoP); while phosphorizing chemically reduced RuCo alloy leads to the formation of hybrid ruthenium cobalt phosphide (RuCoP) clusters. Compared to pristine Ru clusters, both Ru/CoP and RuCoP show significantly improved HER performance in both acidic and alkaline solutions. In particular, the hybrid RuCoP clusters demostrate a considerably low overpotential (η10) of 11 mV at -10 mA cm-2 and a high turnover frequency (TOF) of 10.95 s-1 at η = 100 mV in 0.5 M H2SO4. Even in 1.0 M KOH the excellent HER activity of RuCoP clusters remains, with a very low η10 of 23 mV and exceptionally high TOF value of 7.26 s-1 at η = 100 mV. Moreover, the RuCoP catalysts can sustain galvanostatic electrolysis in both acidic and alkaline solutions at -10 mA cm-2 for 150 h with little degradation, showing better catalytic stability than the state-of-the-art commercial Pt/C catalysts. Our density functional theory (DFT) calculations indicate that the RuCoP hybrid exhibits a hydrogen adsorption energy very close to that of Pt and water and OH adsorption energies distinct from pristine Ru, which reasonably explain the experimentally observed excellent HER activities and highlight the importance of synergistic coupling with cobalt phosphide to boost the HER performance of ruthenium.
Article
The oxygen evolution reaction (OER) is a half-cell reaction that is of importance to many electrochemical processes, especially for electrochemical and photoelectrochemical water splitting. Developing efficient, durable and low-cost OER electrocatalysts comprising earth-abundant elements has been in the focus of electrocatalysis research. Herein, we report a cost-effective, scalable and template-free approach to the fabrication of hollow iron phosphide-phosphate (FeP-FePxOy) composite nanotubes (NTs), which is realized by hydrothermal growth of iron oxy-hydroxide nanorods (NRs) and a subsequent post-phosphorization treatment. The hollow interior of NTs results from the Kirkendall effect occurring upon phosphorization. When used to catalyze the OER in basic medium, the as-synthesized FeP-FePxOy composite NTs exhibit excellent catalytic activity, delivering the benchmark current density of 10 mA cm-2 at a low overpotential of 280 mV and showing a small Tafel slope of 48 mV dec-1 and a high turnover frequency of 0.10 s-1 at the overpotential of 350 mV. Moreover, the composite NTs demonstrate outstanding long-term stability, capable of catalyzing the OER at 10 mA cm-2 for 42 h without increasing the overpotential, holding substantial potential for use as active and inexpensive anode catalysts in water electrolyzers.
Article
The construction of highly active and stable non-noble metal electrocatalysts for hydrogen and oxygen evolution reactions electrocatalysts is a major challenge for overall water splitting. Herein, we report a novel hybrid nanostructure with CoP na-noparticles (NPs) embedded in N-doped carbon nanotube hollow polyhedron (NCNHP) through a pyrolysis-oxidation-phosphidation strategy derived from core-shell ZIF-8@ZIF-67. Benefiting from the synergistic effects between highly active CoP NPs and NCNHP, the CoP/NCNHP hybrid exhibited outstanding bifunctional electrocatalytic performances. When the CoP/NCNHP was employed as both anode and cathode for overall water splitting, a potential as low as 1.64 V was needed to achieve the current density of 10 mA·cm-2, and it still exhibited superior activity after continuously working for 36 h with nearly negligible decay in potential. Density functional theory calculations indicated that the electron transfer from NCNHP to CoP could increase the electronic states of Co d-orbital around the Fermi level, which could increase binding strength with H, and therefore improve the electrocatalytic performance. The strong stability is attributed to high oxidation resistance of CoP surface protected by the NCNHP.