You are on page 1of 28

Review

pubs.acs.org/cm

Synthesis, Characterization, and Properties of Metal Phosphide


Catalysts for the Hydrogen-Evolution Reaction
Juan F. Callejas,†,§ Carlos G. Read,†,§ Christopher W. Roske,‡ Nathan S. Lewis,*,‡
and Raymond E. Schaak*,†

Department of Chemistry and Materials Research Institute, The Pennsylvania State University, University Park, Pennsylvania 16802,
United States

Division of Chemistry and Chemical Engineering, California Institute of Technology, Pasadena, California 91125, United States
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

ABSTRACT: Hydrogen gas obtained by the electrolysis of


water has long been proposed as a clean and sustainable
alternative to fossil fuels. Noble metals such as Pt are capable
of splitting water at low overpotentials, but the implementa-
Downloaded via BONN LIBRARIES on April 28, 2023 at 11:15:47 (UTC).

tion of inexpensive solar-driven water-splitting systems and


electrolyzers could benefit from the development of robust,
efficient, and abundant alternatives to noble metal catalysts.
Transition metal phosphides (MxPy) have recently been
identified as a promising family of Earth abundant electro-
catalysts for the hydrogen-evolution reaction (HER) and are
capable of operating with low overpotentials at operationally
relevant current densities while exhibiting stability under
strongly acidic conditions. In this review, we highlight the
progress that has been made in this field and provide insights into the synthesis, characterization, and electrochemical behavior of
transition metal phosphides as HER electrocatalysts. We also discuss strategies for the incorporation of metal phosphides into
integrated solar-driven water-splitting systems and highlight key considerations involved in the testing and benchmarking of such
devices.

1. INTRODUCTION only water upon combustion. The clean, scalable, and


The development of clean, affordable, and sustainable affordable production of hydrogen is also an important
approaches to fuel generation and utilization is a critical global requirement for the implementation of fuel-cell technologies
challenge. With a rapidly rising world population the global on a global scale.3 Functional hydrogen-based fuel-cell modules
primary energy-consumption rate is expected to increase from are commercially available but are economically disfavored
17 TW in 2010 to 27 TW by 2040.1 Because of their high relative to traditional combustion engines and batteries because
energy density and ease of combustion, fossil fuels have of high costs, storage issues, and limited access to H2 fuel. The
remained the primary global energy carriers for the past two combination of fuel-cell technologies with the clean, wide-
centuries and have played a pivotal role in worldwide industrial spread, and on-demand production of H2 thus has the potential
and technological development. Even though coal and natural to significantly impact the transportation and industrial energy
gas could continue to meet the world’s energy demand for the sectors.
foreseeable future, environmental concerns over the extraction Currently, most hydrogen is produced through industrial
and inefficient combustion of nonrenewable fossil fuels has reforming methods.3 For example, steam−methane reforming
motivated the search for cleaner and more sustainable energy involves the reaction between steam (water vapor) and
platforms.1 methane over a nickel-based catalyst at temperatures above
Solar, wind, and other renewable energy technologies have 700 °C to yield H2 and CO. The obtained CO is then further
emerged as promising alternatives to conventional energy reacted with more steam to finally produce CO2 and more H2
sources. These renewable resources are often intermittent and through the water−gas shift reaction. In addition to high
depend on the time of day and/or weather, requiring batteries reaction temperatures, industrial reforming requires large
and/or other storage technologies to compensate for the amounts of natural gas and adds significantly to rising
intermittency of the resource. The storage of energy as atmospheric CO2 levels. Additionally, hydrogen produced by
chemical bonds in molecules is a promising approach to this method often carries sulfur-containing impurities that are
facilitate long-term storage and additionally to serve the
demands of transportation systems.2 Molecular hydrogen, H2, Received: May 27, 2016
has a considerably higher specific energy than most hydro- Revised: July 28, 2016
carbons and is a well-known zero-emission fuel that liberates Published: August 7, 2016

© 2016 American Chemical Society 6017 DOI: 10.1021/acs.chemmater.6b02148


Chem. Mater. 2016, 28, 6017−6044
Chemistry of Materials Review

of significant environmental concern and that can readily


poison fuel-cell catalysts.4 Devices that facilitate water
electrolysis, including electrochemical and photoelectrochem-
ical cells (PEC), are emerging technologies that have the
potential to renewably generate clean hydrogen fuel from water
without fossil fuels or harmful byproducts.
Overall water “splitting” is the electrochemical reaction that
separates water into molecular hydrogen, H2(g), and molecular
oxygen, O2(g). With ΔG = 237.2 kJ/mol under standard
conditions, the water-splitting reaction is highly endothermic
and requires 1.23 V per electron transferred. While water
splitting can be facilitated in acidic, alkaline, or neutral aqueous
solutions, each has unique advantages, disadvantages, and
challenges.5 Water splitting is favored in strong electrolytes
because of their high ionic conductivities. Efficient, safe
electrolyzers are constructed in either acidic or alkaline media
so that either protons or hydroxide ions can be transferred
between the anolyte and catholyte to avoid a substantial
increase in pH gradients as a result of the electrolysis reaction.6
Electrolysis under pH-neutral conditions is impacted by the
formation of substantial pH gradients that can impede the
water-splitting process.7 The formation of pH gradients can be
mitigated by use of a single compartment reactor with sufficient
mixing, but such systems result in the formation of explosive
mixtures of O2 and H2 gases and/or are inefficient. Acidic
electrolytes are particularly well suited for the HER, because the
reduction of a positively charged proton is more energetically
favorable than reduction of a neutral water molecule.
Additionally, the extensive availability of highly efficient Figure 1. (A) Design schemes for solar-driven electrochemical cells.
proton-exchange membranes favors a reactor design in acidic Adapted with permission from ref 5. Copyright 2014 American
conditions. The water-splitting process can be described by two Chemical Society. (B) Schematic of an integrated PEC water-splitting
separate half-reactions: the hydrogen-evolution reaction device, showing light-absorbing microwire semiconductor arrays with
(HER), which involves proton reduction and occurs at the surface-anchored HER and OER catalysts. The anode (top) and
cathode, and the oxygen-evolution reaction (OER), which cathode (bottom) compartments are connected by a proton-
involves water oxidation and occurs at the anode. Shown below permeable membrane. Reproduced with permission from E. A. Santori.
are the HER and OER in acidic aqueous solutions, which is the
primary emphasis of this review because it is relevant to the and a membrane that separates the two compartments while
conditions under which polymer electrolyte membrane based allowing selective proton transport. High-energy photons (>1.8
water electrolysis devices operate. eV) are absorbed at the photoanode (shown in red), and the
Overall water-splitting: OER catalysts attached to the photoanode’s surface oxidize
water and release O2(g) and protons. Protons then move across
2H 2O → 2H 2 + O2 E°cell = 1.23 V vs NHE a proton-permeable membrane toward the photocathode,
Oxygen-evolution reaction: where lower-energy (<1.2 eV) solar photons are absorbed.
Catalysts that facilitate the HER are attached to the
2H 2O → O2 + 4H+ + 4e− E° = 1.23 V vs NHE photocathode’s surface where the protons combine with
electrons to produce H2(g). The membrane that separates
Hydrogen-evolution reaction: the two compartments shuttles protons from the photoanode
2H+ + 2e− → H 2 E° = 0.00 V vs NHE to the photocathode while keeping the H2(g) and O2(g)
products in separate compartments. This avoids the formation
Several different design schemes have been proposed for of an explosive mixture of H2(g) and O2(g) and also prevents
solar-driven electrochemical cells (Figure 1).5 The simplest, the oxidation of H2(g) at the anode and reduction of O2(g) at
which consists of a photovoltaic cell or module connected to a the cathode and recombination of products within the
water electrolyzer, indirectly converts solar energy into electrolyte. While effective integration of all of the components
chemical fuel. Integrated photoelectrochemical cells are is required to achieve optimal solar water-splitting performance,
attractive options for direct solar fuel production because of the materials themselves are critical since they directly impact
their projected lower costs and potentially high efficiencies as the overall efficiency, stability, scalability, and cost of the device.
compared to indirect schemes.8 One possible configuration, in The widespread implementation of water-splitting technologies
which several key components are integrated and work therefore requires the discovery, development, and integration
synergistically to facilitate overall sunlight-driven water of robust and Earth-abundant materials for each of these
electrolysis, is shown in Figure 1.9 The general layout includes individual components. This review focuses specifically on
two distinct arrays of semiconductor microwires that absorb catalytic materials for the HER, which is one of the key
different portions of the incoming solar spectrum, catalysts that components of the cathode in a full solar-driven water-splitting
decorate the microwire arrays to facilitate the OER and HER, device.
6018 DOI: 10.1021/acs.chemmater.6b02148
Chem. Mater. 2016, 28, 6017−6044
Chemistry of Materials Review

The HER can be facilitated by a diverse range of catalytic Besenbacher28 have developed [Mo3S13]2− and [Mo3S4]4+
systems. In nature and under mild, benign, and pH-neutral clusters, respectively, while Chang27 has developed molecular-
operating conditions, the HER can be carried out by several based systems, which all aim to mimic the HER-active MoS2
biological catalysts, including the [FeFe],10,11 [FeNi],12,13 and edge sites. While such studies of molecular mimics of solid-state
[Fe]-only hydrogenases14,15 as well as by the [FeMo] catalysts are important for active-site design and activity
nitrogenase.16,17 These enzymes often have metal−sulfur optimization, they also emphasize the structural and chemical
clusters as active sites embedded in a complex biological cavity interrelationships among heterogeneous, homogeneous, and
that provides a suitable chemical environment for the HER. biological catalytic systems.
Substantial work has been directed toward developing The most highly studied molybdenum-based HER catalysts,
molecular mimics of these enzyme active sites, as well as including NiMo, Mo2C, and MoS2, are also well-known
other homogeneous molecular catalysts that exhibit comparable catalysts for hydrodesulfurization (HDS).44 HDS is the catalytic
or even higher rates for HER catalysis than the natural process by which sulfur impurities are removed from
enzymes.18 For example, several nickel phosphine complexes hydrocarbon fuels and feedstocks. Despite being distinct
with proximal amine groups have been shown to facilitate the chemical processes, both HDS and HER are regulated by the
HER at very high rates, albeit in acetonitrile solutions.19−22 reversible and dissociative binding of hydrogen molecules on
Other examples of HER molecular catalysts include diiron,23 the surface of a catalyst. Computational studies have indicated
iron diglyoxime,24 and cobalt diglyoxime25 complexes, as well as that both HER and HDS catalysts have active sites that bind
thiomolybdate clusters.26−28 Interestingly, model compounds atomic hydrogen with intermediate strengths, such that the free
of enzyme active sites often are inactive or are significantly less energy of adsorbed hydrogen is closely matched to the free
active than desired for catalytic hydrogen production. When energy of the products, leading to ΔG°H* ≈ 0.45 Hydrogen
inserted into the proper biological cavity, certain inactive adsorption energies that are too high (e.g., strong hydrogen
synthetic complexes exhibit catalytic activities comparable to adsorption) would prevent the release of products, which
those of naturally occurring enzymes.29 From a device include H2 for the HER and H2S for HDS. In contrast,
perspective, biological and homogeneous systems with excep- hydrogen adsorption energies that are too low (e.g., weak
tional catalytic properties face challenges involving anchoring to hydrogen adsorption) will result in slow electron-transfer rates.
and integrating with solid-state systems, as well as stability in Both strong and weak hydrogen adsorption, therefore, result in
chemically harsh, non-neutral electrolytic environments, low catalytic rates. Because HDS catalysts have intermediate
including many of the proposed devices that require strongly hydrogen adsorption energies, it has been proposed, by our
acidic or basic conditions to function efficiently.30 One key goal group and others, that HDS catalyst systems may be fertile
of HER-catalyst development is therefore finding heteroge- ground for the discovery and development of new Earth-
neous systems that can combine the high activity of biological abundant HER catalysts.
and molecular catalysts with the superior stability and Among the most highly studied and active HDS catalysts are
integration capabilities of solid-state materials. Ni2P46,47 and related transition metal phosphides, including
Platinum is the most widely used heterogeneous catalyst for CoP, Fe2P, MoP, and WP.48,49 Given the potential mechanistic
the HER, due to the high catalytic activity and durability of Pt analogy between the HDS and HER catalytic processes, we
under harsh operating conditions.31 The low terrestrial hypothesized that Ni2P and other metal phosphides may indeed
abundance and cost of mining Pt has motivated the search be active and Earth-abundant HER catalysts. Additionally, in
for Earth-abundant alternatives.5,9 Molybdenum-based materi- 2005, Rodriguez and co-workers suggested, based on density
als have been at the forefront of Earth-abundant hydrogen- functional theory (DFT) calculations, that the (001) surface of
evolution catalysis for decades. NiMo alloys were reported by Ni2P combines the favorable H binding present in hydrogenase
Fogarty and co-workers as highly active HER catalysts in systems with the thermostability of a heterogeneous catalyst,
alkaline aqueous solutions.32 Other related alloys including making it a very promising alternative to Pt for catalyzing the
CoMo,33 FeMo,33 and NiMoZn34 have also been reported to HER.45 In 2013, we experimentally validated this prediction,
be active catalysts for the HER. However, despite their high showing that Ni2P was indeed a highly active HER catalyst in
catalytic activity and stability under alkaline conditions, these acidic aqueous solution.50 Since then, our group and others
alloys quickly corrode in acidic environments.35 Other Mo- have demonstrated that the HDS-active metal phosphides
based HER catalysts, including Mo 2 C, 3 6,37 MoB, 36 comprise a new class of highly active and acid-stable HER
Co6Mo1.4N2,38 and NiMoNx,39 have been investigated, and catalysts. The field of metal phosphide HER catalysts has
many of these catalysts exhibit extended stability in acidic rapidly expanded to include a growing number of catalytic
aqueous solutions. systems and preparation methods, demonstrations of integra-
Using theoretical and experimental methods, Hinnemann tion into functional photocathode systems, mechanistic insights
and co-workers showed that the edge sites of MoS2, which are in the catalytic reactions, and guidelines for designing new
chemically and structurally distinct from the Mo-based alloys, catalysts and improving the performance of existing catalysts.
have chemical environments that can facilitate the HER.40 This Review highlights recent developments in transition
Accordingly, MoS2 has been the leading Earth-abundant metal phosphides as an emerging family of highly active and
alternative to platinum for catalyzing the HER in acidic Earth-abundant catalysts for the HER, primarily in acidic
aqueous solutions. The HER-active edge sites of MoS2 have conditions that are relevant to proton-exchange-membrane
structural commonalities with the active-site clusters in some electrolysis systems. The HER performance in pH-neutral and
hydrogenase and nitrogenase enzymes.40 Extensive research alkaline aqueous solutions is also highlighted due to potential
efforts have been directed toward understanding and max- relevance to alternative water-electrolysis systems. We include
imizing the number of exposed active sites in MoS2, and this in this Review a survey of how transition metal phosphides are
has led to the development of improved MoS2-based HER synthesized across multiple platforms such as bulk crystals,
catalysts that are highly active and acid stable.41−43 Ooi26 and films, and nanoparticles, because collectively these techniques
6019 DOI: 10.1021/acs.chemmater.6b02148
Chem. Mater. 2016, 28, 6017−6044
Chemistry of Materials Review

and the materials they produce are relevant for exploratory strongly acidic or alkaline solutions. More ionic phosphides,
synthesis and catalyst discovery, optimization of catalytic such as Ca3P2 and Zn3P2, however, readily decompose in water
performance through active-site exposure and surface area to produce highly pyrophoric and toxic gases, such as
maximization, integration into devices, and understanding phosphines and diphosphines.
mechanistic details of the catalytic reactions. Additionally, we In contrast to the metal-rich and stoichiometric phosphides,
discuss aspects of materials and electrochemical character- phosphorus-rich transition metal phosphides (y > x in MxPy)
ization that are crucial for fully understanding the materials exhibit significant phosphorus−phosphorus bonding, due to the
being studied, accurately attributing catalytic activities to the ability of phosphorus to bond with itself to form various
correct materials features, and benchmarking performance oligomers and clusters. For example, a number of MP2
metrics with related systems. We then provide an overview of compounds, such as NiP2 and SiP2, adopt the pyrite-type
the properties and performance metrics of transition metal structure in which the phosphorus atoms are arranged in P−P
phosphides for HER catalysis, including their integration with dimers. Other polyphosphides contain various phosphorus
light-absorber materials as an important step toward building a oligomers, clusters, chains, and planes. These so-called
practical solar-driven water-splitting device. polyphosphides exhibit characteristics that are markedly
different from their metal-rich or stoichiometric counterparts,
2. OVERVIEW OF METAL PHOSPHIDES including lower thermal stabilities, higher reactivities, and softer
Metal phosphides, represented by the general formula MxPy, are materials properties that can be classified as significantly less
solid-state compounds formed from the combination of refractory. As a result, many phosphorus-rich phosphides are
metallic or semimetallic elements with phosphorus. The crystal thermally unstable, disproportioning at high temperatures to
structures adopted by the large number of known binary, elemental phosphorus and more metal-rich phases.
ternary, and higher-order metal phosphides are diverse (Figure
2), spanning simple high-symmetry ionic structures such as 3. SYNTHESIS OF METAL PHOSPHIDES
Metal phosphides can be synthesized using a variety of methods
and in various forms, including single crystals,54−56 bulk
polycrystalline powders,57 films,58,59 and nanostructured
solids60−66 (Figure 3). A comprehensive list of phosphide
materials that have been tested for the HER is included in later
sections. Bulk metal phosphides can be prepared through
traditional solid-state strategies by direct combination of the
elements at high temperatures in an inert atmosphere or under
vacuum. Using this approach, many phosphide phases can be
routinely accessed in high purity and on a large scale. As is
typical for bulk-scale solid-state reactions, high reaction
temperatures (>900 °C) and long reaction times (1−10
days) are generally required. For example, in a representative
synthesis of bulk FeP,67 stoichiometric amounts of iron metal
and red phosphorus are sealed in an evacuated silica tube,
which is then heated to 900 °C for approximately 8 days. Red
phosphorus is often used in these direct high-temperature solid-
Figure 2. Crystal structures of representative types of metal state reactions, although the more reactive white phosphorus
phosphides: NaCl-type LaP, skutterudite-type LaRu4P12, ThCr2Si2- allotrope can also be used, as can certain reactive metal
type LaRu2P2, and MgAs4-type ZnP4. phosphides. For example, various phosphide phases, including
AlP68 and NbP,69 have been accessed by high-temperature
NaCl-type LaP to more complex structures such as ThCr2Si2- reactions between a metal phosphide of a lower stability, such
type LaRu2P2 and skutterudite-type LaRu4P12. The bonding in as Ca3P2 or Zn3P2, and the appropriate metal powder (>1000
metal phosphides is also diverse and, depending on the °C). Because the high-temperature solid-state reactions can
composition and constituent elements, can be described as produce highly reactive and pyrophoric byproducts, including
ionic, covalent, or metallic. Metal-rich (x > y in MxPy) or P4 and phosphine, properly trained personnel must work under
stoichiometric (x = y = 1 in MxPy) metal phosphides are often rigorously air-free conditions to perform the reactions safely as
semiconducting and in some cases even metallic or super- well as to isolate the products.
conducting due to the presence of significant metal−metal To lower the temperatures required by direct reactions and
bonding. For instance, TiP and Fe2P exhibit metallic behavior, to expand the palette of accessible phases, molten fluxes have
whereas GaP and InP are well-known semiconductors. been used extensively in the synthesis of metal phosphides.55 In
Superconducting properties have been observed in various this approach, a nominally unreactive and low-melting metal,
metal-rich phases such as Mo3P51 and LaRu2P2.52 such as Sn or Pb, is mixed with the precursor elements and
The relatively strong M−P bonds can impart transition metal used as a high temperature solvent to enhance the diffusion rate
phosphides with high thermal stability and hardness, as well as of the solid reagents.55 After the reaction, the metal matrix must
resistance to oxidation and chemical attack. For example, the be separated from the products either mechanically or by
phosphides of various metals, such as Ti, Ta, Mo, and W, are of dissolution in acid. For example, in a representative flux
interest as oxidation-resistant coatings for high-temperature synthesis of RuP2,70 stoichiometric amounts of ruthenium and
applications.53,54 Importantly for applications such as HER red phosphorus powders are placed along with excess tin in an
electrocatalysis, many transition metal phosphides are imper- evacuated silica tube, which is sealed and heated to 1200 °C for
vious to dilute acids and bases, and some are unaffected even by approximately 3 days. After the reaction, the phosphide product
6020 DOI: 10.1021/acs.chemmater.6b02148
Chem. Mater. 2016, 28, 6017−6044
Chemistry of Materials Review

Figure 3. Representative types of metal phosphide crystals, films, and nanoparticles. (A) A crystal of NdFe4P12 with the cubic LaFe4P12-type structure
grown from a tin flux. Adapted with permission from ref 55. Copyright 2005 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim. (B) Cu metal
wire and foil and the wire and foil with a thick Cu3P coating made by refluxing in trioctylphosphine. Adapted with permission from ref 57. Copyright
2007 American Chemical Society. (C) SEM images of a Ni2P film on Ni. Adapted with permission from ref 58. Copyright 2016 American Chemical
Society. (D) SEM image of a thin film showing a mixture of Ni2P and Ni0.85Se deposited using CVD. Adapted with permission from ref 59.
Copyright 2008 American Chemical Society. (E) TEM image of Zn3P2 nanoparticles. Adapted with permission from ref 61. Copyright 2014
American Chemical Society. (F, G) TEM images of Ni2P nanoparticles. Adapted with permission from refs 62 and 63. Copyright 2007 and 2009
American Chemical Society. (H) TEM image of FexNi2−xP nanoparticles. Adapted with permission from ref 64. Copyright 2015 American Chemical
Society. (I) TEM image of Rh2P nanoparticles. Adapted with permission from ref 65. Copyright 2015 American Chemical Society. (J) TEM image of
CoP nanoparticles. Adapted with permission from ref 66. Copyright 2011 American Chemical Society.

is then recovered from the flux by dissolving the tin in hot phosphides for catalytic applications such as hydrodesulfuriza-
concentrated HCl. In many instances, the flux method yields tion or hydrodenitrogenation, the strategy has recently been
high-quality phosphide crystals (Figure 3) and provides access extended to the production of phosphide materials directly on
to metastable and low-temperature phases that are inaccessible the surfaces of electroactive substrates, such as conductive
by the use of higher-temperature solid-state reactions, which carbon paper78 and metal foams.79 Electrochemical and
tend to favor the formation of more thermodynamically stable electroless deposition methods have also been explored as a
products. Moreover, phosphide phases like CrP4,71 MnP4,72,73 way to directly coat electrode surfaces with metal phos-
and Re2P574 are challenging to obtain through alternative phides.80−86 However, these approaches tend to yield
methods without the use of high pressures. amorphous Co−P and Ni−P alloys with a wide range of
Another approach that has been extensively used in the phosphorus contents.
synthesis of metal phosphides is the phosphidation of metal The need for high-surface-area phosphide materials for
oxides, hydroxides, or other precursors by highly active catalytic and electrocatalytic applications has led to renewed
phosphorus species. The phosphidation can be obtained either interest in alternative synthetic strategies for the production of
through direct exposure to phosphine gas75 or by exposure to metal phosphides. For example, solvothermal reactions,87
related compounds generated in situ through the reduction of thermal decomposition of single-source organometallic pre-
phosphate salts by hydrogen48 or carbon.76 For instance, cursors,88 and the reaction of organometallic compounds or
several phosphide phases, such as Ni2P, CoP, and FeP, can be metallic nanoparticles with organophosphine reagents57,89 have
readily obtained by the temperature-programmed reduction been used to produce crystalline high-surface-area metal
(TPR) of the corresponding metal phosphate.48,49 In a phosphides under reaction conditions that are frequently
representative TPR synthesis of CoP,77 a stoichiometric milder than those found in direct reactions or flux approaches.
mixture of cobalt nitrate and ammonium hydrogen phosphate Furthermore, these methods typically produce metal phos-
is calcined in air at 500 °C for approximately 6 h to produce a phides in the form of dispersible nanocrystals that can be
cobalt phosphate precursor, which is subsequently reduced by directly applied by drop-casting or spin-coating onto the
heating to 1000 °C for 2 h in a H2-containing atmosphere. surfaces of electrodes. Highly reactive reagents such as white
Originally developed to produce metal-oxide-supported phosphorus (P4) or P(SiMe3)3 can be used as phosphorus
6021 DOI: 10.1021/acs.chemmater.6b02148
Chem. Mater. 2016, 28, 6017−6044
Chemistry of Materials Review

sources, but milder reagents such as tri-n-octylphosphine


(TOP) have been used as general phosphorus sources for the
low-temperature conversion of metals into metal phosphides.
Multiple phases such as Ni2P, Ni12P5, Ni5P4, Cu3P, Fe2P, FeP,
Co2P, CoP, InP, PtP2, PdP2, RhP2, Au2P3, Pd5P2, and
MnP,57,89,90 as well as mixed-metal solid solutions such as
(NixFe1−x)2P,64 (NixCo1−x)2P,91 and (CoxFe1−x)2P,92 have been
synthesized as nanoparticles through these methods (Figure 3).
However, due to the use of solvents for such reactions, the
temperature range is limited and generally cannot exceed 400
°C or the maximum reflux temperature of the highest-boiling
solvents.
Several thin-film growth techniques have also been applied to
the synthesis of transition metal phosphides, mainly for the
fabrication of semiconducting or optoelectronic devices. Figure 4. Polarization data for the HER in 0.5 M H2SO4 for a series of
Chemical vapor deposition (CVD) and metal−organic Pt electrodes, including a Pt mesh, Pt nanoparticles, Pt on carbon
chemical vapor deposition (MOCVD) (Figure 3) have been black, and a Pt disk.
widely used to produce high-quality crystalline and amorphous
thin films of several transition metal phosphides, including Linear sweep voltammetry (LSV) and cyclic voltammetry
InP,93 GaP,94 Zn3P2,95 Ni2P,96 and TiP,97 among others. (CV) are commonly used to evaluate the HER performance by
Typically in these processes, volatile gaseous precursors such as measurement of the catalytic current as a function of applied
metal alkyls, metal halides, or in the case of MOCVD, single- potential. The observed catalytic current is typically plotted as
source metal−organic compounds, are decomposed at high the experimentally observed current density, normalized to the
temperatures over an appropriate substrate. Similarly, physical geometric area of the electrode, to facilitate comparisons
vapor deposition (PVD) techniques, such as sputtering98,99 and among electrodes having different sizes. This approach however
pulsed laser deposition (PLD),100 have also been used to does not account for variations in catalyst loading or surface
produce metal phosphide thin films on various substrates. area. Methods to estimate the actual surface area of the catalyst
However, in most cases the films obtained through these include the use of bulk surface areas obtained through BET
techniques are poorly crystalline or amorphous. analysis or geometric estimates using mathematically derived
surface areas based on average particle sizes and shapes.50,101
4. CHARACTERIZATION OF TRANSITION METAL These estimates are limited by the realization that not all
PHOSPHIDE CATALYSTS FOR THE exposed surface sites are catalytically active. Electrochemical
HYDROGEN-EVOLUTION REACTION measurements of the surface area most closely relate to HER
In this section we discuss methods for thorough and rigorous operating conditions, but they have been predominately
characterization of metal phosphides. Such methods allow the developed for noble metal systems and therefore may not be
establishment of catalytic performance metrics using a common directly applicable to the metal phosphides. The exact
framework that permits both benchmarking and comparisons determination of the true electrochemically active surface area
and also enable accurate attribution of the observed catalytic may not be possible, so electrocatalytic testing on flat electrode
properties to the key material features that define the systems substrates is preferable to best obtain the inherent electro-
of interest. chemical performance of a material. The use of flat substrates
Characterization of Electrocatalytic Properties. Metal prevents artificial enhancement of electrochemical performance
phosphide systems have emerged as highly active HER through increased surface area effects, thereby facilitating
electrocatalysts, and they are being increasingly investigated comparisons with other catalysts. Although the development
for this and other catalytic reactions. Accordingly, several recent of highly efficient 3D and porous electrodes is an important
articles have outlined and reviewed best practices for testing, area of research for electrode design and optimization, such
reporting, and benchmarking such electrocatalytic materials as studies are most useful when the high surface area systems can
well as the related photocatalyst systems.5,30,101,102 Figure 4 be compared with the catalytic performance on a flat electrode.
shows typical data for various types of benchmark Pt catalysts, When reporting the results of electrocatalytic testing, the
which will be discussed in more detail below. Key overpotential required to reach a specific current density, which
considerations for appropriate electrocatalytic testing of the can be chosen depending on the target application, allows
metal phosphide systems are briefly outlined below as well. reliable comparison between catalysts tested under similar
Materials for HER electrocatalysis are typically evaluated experimental conditions. The “onset potential”, which is the
using a three-electrode setup, in which a reference electrode, a potential at which catalytic current first appears, is often
counter electrode, and a catalyst-modified working electrode reported in the literature. However, because the onset potential
are immersed in an aqueous electrolyte. The electrolyte must can be ascribed to the production of an arbitrarily defined
be continuously purged with high purity H2 gas to establish current density, the “onset potential” is not a well-defined
standard conditions. When Pt is used as a counter electrode, a electrochemical property and hence is not a suitable metric for
two-compartment cell, separated by a proton exchange analytically evaluating or comparing different HER catalysts.101
membrane, must be utilized to separate the working and Other relevant metrics that are often reported for HER catalysts
counter electrodes and thereby prevent cross-contamination by include Tafel slope, exchange-current density, and turnover
trace noble metal species, as well as undesired back reactions. frequency (TOF). These parameters in combination are also
When using a single compartment cell, only graphite rods or important for evaluating electrocatalytic performance and have
other inert materials should be used as counter electrodes. been reviewed in detail in recent articles.5,30,101,102
6022 DOI: 10.1021/acs.chemmater.6b02148
Chem. Mater. 2016, 28, 6017−6044
Chemistry of Materials Review

Figure 5. Representative, nonexhaustive summary of materials-characterization data that can be used to understand the key characteristics that
underpin the observed catalytic performance, including important aspects of surface and bulk structure, composition, and morphology. Color-coding
shows complementary types of information that are provided by different characterization techniques.

Galvanostatic measurements and cyclic voltammetry are electrolyte analysis studies to identify and understand slow
typically used to characterize the stability of a catalyst. dissolution processes, and impedance measurements to probe
Galvanostatic measurements maintain an application-relevant the electrical resistivity of the catalyst.
current density, such as 10 mA cm−2 for photoelectrochemical As mentioned above, catalyst benchmarking is important,
(PEC) cells, for a sufficiently long time to establish the desired and as a result, proper controls and reporting metrics are
degree of catalyst stability. Galvanostatic testing is particularly mandatory. For example, showing electrochemical data for
useful for evaluating the longevity of a catalyst under device- standard Pt electrodes under the same conditions used to
relevant operating conditions. For catalysts that dissolve slowly evaluate new catalytic materials provides a necessary baseline
under operating conditions, high catalyst mass loadings may for comparison. However, the availability of a wide range of
result in artificially prolonged stability. For this reason, different Pt standards having different surface areas and mass
galvanostatic stability measurements are most useful when loadings, including flat Pt disks, Pt mesh electrodes, and
performed at low mass loadings or when coupled with sensitive supported Pt nanoparticles, complicates matters, as such
elemental analysis of the electrolyte, such as atomic absorption systems exhibit very different catalytic performances (Figure
spectroscopy or inductively coupled plasma mass spectrometry 4). Pt meshes are particularly desirable because they are
(ICP-MS), to detect dissolution processes. An added benefit of commercially available, have high catalytic activity, and offer
galvanostatic testing is that the current density at which the highly reproducible performance. Ultimately, the most active Pt
experiment is conducted can be selected to investigate the HER standards should be used for comparison to new catalysts. The
stability for a wide range of applications and device designs, overpotentials for a clean Pt mesh are ∼−15 to −20 mV at a
ranging from PECs (−10 to −20 mA cm−2) to electrolyzers current density of −10 mA cm−2. When reporting the results of
(−1 to −2 A cm−2). Multiple cyclic voltammetry cycles over an electrocatalytic testing, the overpotential required to reach a
appropriate potential window, such as between 0 V vs RHE and specific current density (the operationally relevant benchmark
the potential required to reach or exceed a target operational current density) allows facile comparison between catalysts
current density, mimic the ramp-up and ramp-down cycles tested under similar experimental conditions. Reporting the
expected for solar-driven water-splitting systems. Additional electrode details is also important for characterizing and
useful studies include prolonged testing under the open-circuit comparing catalysts. For example, differences in loading density
potential (to simulate the system at rest), tests over longer and surface area can influence the reported metrics and are
periods of time (months to years), quartz-microbalance and important considerations.
6023 DOI: 10.1021/acs.chemmater.6b02148
Chem. Mater. 2016, 28, 6017−6044
Chemistry of Materials Review

Materials Characterization. Coupled with electrochemical catalytic materials to characterize their morphological features,
characterization, it is important to fully characterize the key TEM data corroborate bulk structural and compositional
aspects of catalytic materials that contribute to their perform- information. For example, carefully and accurately analyzed
ance, including techniques that probe the bulk crystal structure, lattice spacings, angles, and structural motifs observed by high-
morphology, and chemical composition, as well as key surface resolution TEM (HRTEM) can corroborate the assigned
chemistry details (Figure 5). This, coupled with the structure and provide knowledge about the facets that are
benchmarking efforts described in the preceding section, exposed. Comparing the average particle or grain sizes observed
facilitates meaningful comparisons among catalytic systems, by TEM to the grain sizes determined by Scherrer analysis of
establishes the relevant parameters, and sets the stage for powder XRD data confirms that the bulk of the sample is
elucidating structure−property relationships. Ultimately, one comprised of the same relative sizes as those observed
should make deliberate and informed choices about which microscopically. As with all microscopy techniques, only a
characterization tools will provide the necessary information for small fraction of the sample is interrogated using SEM or TEM
adequately evaluating a catalyst at each stage of discovery, analysis, so care must be taken when formulating conclusions
development, and detailed understanding. Figure 5 shows about the properties of a bulk sample from microscopic data
conceptually how various materials and surface characterization alone.
tools can be used together to study catalytically active materials. Electron-diffraction data, both for individual particles and for
For metal phosphide HER catalysts in nanoparticle, bulk- large ensembles of particles, should match the bulk XRD data
powder, or thin-film form, powder X-ray diffraction (XRD) and, consequently, further confirm the structural assignment.
data enables the identification of all crystalline phases present in Compositional analysis by energy-dispersive X-ray spectroscopy
a sample, as well as an estimation of the size of crystalline (EDS), again for both large ensembles of particles and also for
domains through Scherrer analysis. Because phase diagrams of individual particles as needed, can further validate the phase
metal phosphides contain multiple crystalline compounds of assignment. In conjunction with other types of materials-
different compositions and crystal structures, each of which can characterization data, such compositional analysis helps confirm
have different properties, high-quality XRD data is important that substantial amounts of amorphous or impurity phases are
for establishing phase formation and purity. Additionally, for not present in a sample. SEM is particularly helpful because of
particles that are found to be highly anisotropic or films that the relatively large sample size (compared to TEM) that can be
contain highly oriented crystallites, the observation of preferred interrogated. For amorphous materials, high-resolution EDS
orientation by powder XRD data can confirm that the element maps obtained using TEM are particularly helpful
morphology is characteristic of the bulk sample. Powder because such maps can identify the presence, amounts, and
XRD cannot, however, unambiguously confirm phase purity, distributions of the elements within the catalytic material, albeit
nor can it exclude the possibility that catalytically relevant for only a small region of the sample. For all materials, analysis
impurities are present or reveal the presence of amorphous before and after extended testing allows confirmation that the
components. catalyst is stable under operating conditions. Materials
Bulk elemental analysis can be used to compare the overall characterization after electrochemical testing can be challenging
composition with the crystalline phase and thus to determine because the catalyst is anchored directly to an electrode
whether significant amorphous phases not detectable by substrate, but surface analysis, as well as bulk analysis of the
powder XRD are present in a sample. For fully amorphous catalyst after physically detaching it from the electrode
samples that have no long-range crystalline order, character- substrate, can still provide useful insights. These materials-
ization can be more challenging. In these cases, rigorous characterization tools can provide valuable information when
analysis of composition, sample heterogeneity, and oxidation used separately but can be even more powerful when used in a
states can be especially important, along with any other complementary fashion. Specifically, structural information
microscopic or spectroscopic techniques that are appropriate from XRD should be in agreement with electron diffraction
and available. Because the structure and composition at the and with lattice spacings obtained from HRTEM. Similarly,
surface of a catalyst may be quite different from that in the bulk, grain sizes observed by TEM should be in agreement with
X-ray photoelectron spectroscopy (XPS) can be especially Scherrer analysis from XRD. Elemental analyses from multiple
powerful. XPS can be utilized for identification of the oxidation techniques should match one another and should also match
states and chemical composition near the surface, both for the stoichiometry expected for the assigned phase. Electro-
crystalline and amorphous materials. However, for some catalysts can be examined both before and after electrocatalytic
systems, such as CoP and Co2P, differentiating oxidation states testing because several techniques can be performed directly on
and quantifying their relative ratios is often not possible or electrode substrates, including SEM, XPS, and XRD. Such
straightforward. Synchrotron-based techniques such as X-ray analyses enable the identification of structural, compositional,
absorption spectroscopy (XAS) and pair distribution function and morphological changes during catalysis that provide
(PDF) analysis offer additional insights into the bulk and local important insights into both the true active form of the
structure of catalytic materials and can be performed in situ and material and its stability.
under operationally relevant conditions.103
Electron microscopy, particularly for nanoparticulate and 5. TRANSITION METAL PHOSPHIDES FOR THE HER
thin-film metal phosphide catalysts, complements the bulk and Despite being well-known catalysts for various hydrotreating
surface analyses highlighted above. Both scanning electron processes, such as hydrodesulfurization and hydrodenitrogena-
microscopy (SEM) and transmission electron microscopy tion,48,49 transition metal phosphides have only recently been
(TEM) can provide detailed information about the morpho- explored as catalysts for the HER. Early work by Paseka84,86
logical aspects of a catalyst sample, including the distribution of and Burchardt82 demonstrated that amorphous alloys of Ni,
shapes and sizes of the constituent particles or grains, as well as Co, and Fe with small amounts of phosphorus (1 to 27 wt %)
insights into sample heterogeneity. Beyond simply imaging the were able to catalyze the HER at relatively low overpotentials in
6024 DOI: 10.1021/acs.chemmater.6b02148
Chem. Mater. 2016, 28, 6017−6044
Chemistry of Materials Review

Table 1. Compilation of HER Performance Metrics for Various Nickel Phosphide Catalysts Synthesized under Different
Conditions and Evaluated in 0.5 M H2SO4
material η−10mAcm−2 η−20mAcm−2 Tafel slope (mV dec−1) exchange current density (A cm−2) loading density (mg cm−2) ref
−5
Ni2P NPs/Ti ∼−116 −130 ∼46 3.3 × 10 1.0 50
Ni2P/Ni −128 −153 66 - - 58
Ni2P NS/Ni foam ∼−115 ∼−140 68 - - 79
Polydispersed Ni2P/GCE ∼−125 −140 ∼87 - 0.38 104
nanoparticle films Ni2P/Ti ∼−130 −138 60 - 2.0 107
Ni2P/CNT −124 - 53 5.37 × 10−5 - 108
peapod-like Ni2P/C −87 −115 54 - 0.36 109
Ni2P−G@NF ∼−150 - ∼30 - 110
Ni5P4 MP pellet −23 - 33 - 177 111
Ni2P MP pellet −42 - 38 - 177 111
Ni12P5/Ti −107 −141 63 - 3 112
NiP2 NS/CC −75 - 51 2.60 × 10−4 4.3 113
Ni2P NPs −137 - 49 - 1.99 114
Ni5P4 NPs −118 - 42 - 1.99 114
Ni12P5 NPs −208 - 75 - 1.99 114
Ni5P4/Ni −140 - 40 - - 115
Ni12P5/CNT −129 - 56 7.10 × 10−5 0.75 116
Ni2P/GCE - - 84 2.90 × 10−6 0.15 117
Ni12P5/GCE - - 108 3.70 × 10−7 0.15 117
MOF-derived Ni2P ∼−200 - 62 7.10 × 10−5 0.35 118
MOF-derived Ni12P5 ∼−650 - 270 4.50 × 10−5 0.35 118
Ni2P/CNSs −92 −108 47 4.90 × 10−4 - 119
Ni2P NPs/Ni foam −136 - - - 0.14 120
Ni−P films −93 - 33 and 98 - 0.35 121
Ni2P-NRs/Ni −131 −163 106.1 8.62 × 10−5 - 122
Ni5P4 - −62 46.1 2.75 × 10−4 0.15 123
Ni2P - −228 83.3 2.10 × 10−4 - 124
Ni2P/NRGO −102 −122 59 4.90 × 10−5 - 125
Ni2P−G/NF −75 - 51 - - 126
(Ni2P)@graphitized carbon −45 - 46 - 0.38 127

alkaline electrolytes. In 2005, based on density functional different conditions and evaluated in 0.5 M H2SO4. The first
theory (DFT) calculations, Liu and Rodriguez predicted that Ni2P materials studied experimentally as HER catalysts were
Ni2P may be a potential alternative to Pt, suggesting that nanoparticles synthesized by reacting trioctylphosphine (TOP)
synergistic effects between exposed proton-acceptor and and nickel(II) acetylacetonate in 1-octadecene and oleylamine
hydride-acceptor centers on the (001) surface of Ni2P could at 320 °C for 2 h.50 Several synthetic routes to colloidal Ni2P
mimic features of the active sites of hydrogenase enzymes to nanoparticles have been reported. For example, work by the
facilitate efficient HER catalysis.45 In 2013, we experimentally groups of Brock,47 Chiang,62 Hyeon,105 Robinson,106 and
validated this prediction, showing that nanostructured Fe2P- Tracy,63 along with our group,57,89 demonstrated that high-
type Ni2P was indeed a highly active HER electrocatalyst in quality Ni2P nanocrystals could be readily obtained in solution
acidic aqueous solutions.50 Hu and co-workers similarly showed by the coreaction of organophosphine reagents such as TOP
that Ni2P nanopowders prepared through alternative solid-state with nickel complexes or premade Ni nanopartices. The Ni2P
approaches were also highly active HER catalysts in acidic particles initially evaluated as HER catalysts were synthesized
aqueous solutions.104 Since then, many groups worldwide have using the method reported by Tracy and co-workers because it
contributed extensively to the advancement of this field, produced a high yield of monodisperse, phase-pure Ni2P
including the discovery of other metal phosphide HER nanocrystals through a simple one-pot reaction.63 As shown in
catalysts, the development of new and improved methods for Figure 6, the as-synthesized Ni2P particles were monodisperse,
the synthesis and processing of catalytic metal phosphide hollow, multifaceted, and single-crystalline, with an average
materials, the interrogation of their electrocatalytic and diameter of 20 nm. The hollow morphology is the result of the
photocatalytic properties, investigations into the mechanisms nanoscale Kirkendall effect, which is caused by differences in
by which they function, and demonstrations of their the inward vs outward diffusion rates of the constituent
applicability in integrated systems and devices. While transition elements during the reaction. Kirkendall voids are commonly
metal phosphides have also been shown to be active HER observed in metal phosphide nanoparticles synthesized by the
catalysts in pH-neutral and alkaline aqueous solutions, in this decomposition of trioctylphosphine. Working electrodes of the
review we focus primarily on their behavior under acidic Ni2P material were prepared by applying the as-made
conditions that are relevant to proton-exchange-membrane nanoparticles to Ti foil substrates, followed by annealing at
electrolysis systems. 450 °C in H2(5%)/N2(95%) to remove the organic ligands that
Nickel Phosphides. Table 1 summarizes the performance capped the surface of the nanoparticles. The resulting
of various nickel phosphide HER catalysts synthesized under nanoparticulate Ni2P films required an overpotential of only
6025 DOI: 10.1021/acs.chemmater.6b02148
Chem. Mater. 2016, 28, 6017−6044
Chemistry of Materials Review

Figure 6. (A) TEM image of Ni2P nanoparticles. (B) HRTEM image


of a representative Ni2P nanoparticle highlighting the exposed
Ni2P(001) facet and the 5.2-Å lattice fringes that correspond to the
(010) planes. (C) Experimental powder XRD pattern for the Ni2P
nanoparticles, with the simulated pattern of Ni2P shown for
comparison. (D) Polarization data for three individual Ni2P electrodes
in 0.5 M H2SO4, along with glassy carbon, Ti foil, and Pt in 0.5 M
H2SO4, for comparison. Adapted with permission from ref 50. Figure 7. (A) SEM image of Ni2P particles on a Ti plate and (B)
Copyright 2013 American Chemical Society. corresponding linear sweep voltammograms in 0.5 M H2SO4. Adapted
with permission from ref 107. Copyright 2014 Royal Society of
Chemistry. (C) TEM image of Ni2P particles decorating multiwalled
−116 mV to produce an operationally relevant current density carbon nanotubes and (D) corresponding linear sweep voltammo-
of −10 mA cm−2 (η−10mAcm−2 = −116 mV) in a strongly acidic grams in 0.5 M H2SO4. Adapted with permission from ref 108.
electrolyte (0.50 M H2SO4), while also demonstrating good Copyright 2015 Royal Society of Chemistry. (E) TEM image of bulk
stability and quantitative Faradaic efficiencies over 2 h of Ni2P nanopowders and (F) corresponding linear sweep voltammo-
sustained hydrogen production.50 Hu and co-workers similarly grams in 0.5 M H2SO4. Adapted with permission from ref 104.
demonstrated that Ni2P nanoparticles made using a bulk-scale Copyright 2014 Royal Society of Chemistry.
reaction between NaH2PO2 and NiCl2·6H2O showed excellent
activity and stability in both acidic and alkaline solutions, Recently, our group also presented a general and scalable
requiring overpotentials of approximately η−10mAcm−2 = −125 strategy for the synthesis of metal phosphide electrodes,
mV and η−10mAcm−2 = −230 mV in acidic and alkaline including Ni2P, based on the phosphidation of commercially
conditions, respectively.104 The observed catalytic performance available metal foils through the vapor-phase decomposition of
placed Ni2P among the best non-noble-metal HER electro- various organophosphine reagents (tributylphosphine and
catalysts in acidic aqueous solutions reported up to that point, trioctylphosphine).58 The resulting films exhibited excellent
including MoS2,40−42 NiMoN,38 MoB,35 and Mo2C cata- activities, with the Ni2P electrodes requiring overpotentials for
lysts.35,36 the HER of approximately η−10mAcm−2 = −128 mV in 0.50 M
Many other groups have described similar activities and H2SO4 and η−10mAcm−2 = −183 mV in 1 M KOH. Additionally,
stabilities for a wide portfolio of Ni2P materials, including we demonstrated that the same phosphidation strategy could
various nanostructures, films, and bulk powders (Figure 7). For be applied to evaporated metal thin films to form conformal
example, Sun and co-workers reported Ni2P nanoparticle films metal phosphide coatings on a variety of substrates, including
prepared via the low-temperature phosphidation of electro- relevant photocathode materials such as highly doped Si.
deposited nickel hydroxide precursors.107 The resulting Ni2P Despite the low loadings and low surface areas of the samples,
films displayed a HER performance comparable to those the Ni2P thin films on Si exhibited moderate activities for the
reported previously and made by other methods, requiring an HER, requiring an overpotential of η−10mAcm−2 = −240 mV in
overpotential of approximately η−10mAcm−2 = 130 mV in 0.50 M 0.50 M H2SO4.
H2SO4 and exhibiting stable hydrogen production for at least Significant enhancements to the HER activity of Ni2P have
15 h. Likewise, Liu and co-workers observed similar activities been reported through the use of various phosphide−carbon
(η−10mAcm−2 = approximately −124 mV) in samples of Ni2P composites and 3D electrode geometries. For instance, Wang
nanoparticles decorated on multiwalled carbon nanotubes and co-workers presented carbon-encapsulated Ni2P nano-
(Ni2P/CNT).108 The Ni2P/CNT material was synthesized in particles (Ni2P/C) prepared by the reduction of glucose-coated
a one-pot reaction by the in situ thermal decomposition of NiNH4PO4·H2O nanorods with H2 at 700 °C.109 These Ni2P/
nickel acetylacetonate and TOP in an oleylamine solution of C nanocomposites showed enhanced HER performance,
acid-treated CNTs, followed by deposition onto glassy carbon requiring only η−10mAcm−2 = −87 mV in 0.50 M H2SO4. The
electrodes. improved electrocatalytic activity of the Ni2P/C composite was
6026 DOI: 10.1021/acs.chemmater.6b02148
Chem. Mater. 2016, 28, 6017−6044
Chemistry of Materials Review

attributed to enriched nanoporosity and a more efficient use of pellets and sealing all but one side with epoxy. The reported
the available active sites. Similarly, Du and co-workers reported overpotential required by the Ni5P4 electrodes in 0.50 M
a three-dimensional few-layer graphene/nickel foam (G@NF) H2SO4 was an exceptionally small value of η−10mAcm−2 = −23
electrode coated with nanostructured Ni2P that displayed mV, which is almost identical to that of Pt. It is unclear why this
exceptional HER activity, requiring an overpotential of Ni5P4 catalyst has such low overpotentials relative to all other
η−10mAcm−2 = −55 mV in 0.50 M H2SO4.110 Such high catalytic reported metal phosphide HER catalysts. Ni5P4 nanoparticles
performance was attributed to the presence of more catalyti- prepared through similar methods and of comparable surface
cally active sites provided by the larger surface area of the areas were reported by Liu and co-workers under similar testing
porous electrode and to enhanced ion and electron transfer. conditions to require η−10mAcm−2 = −118 mV.114 Likewise,
However, the activity of the 3D Ni2P−G@NF electrode was Shalom and co-workers reported the growth of Ni5 P 4
normalized to a flat geometric surface area despite being highly nanoarchitectures directly on Ni foils by heating the metal
porous. Despite the wide diversity of synthetic preparations, with red phosphorus at 550 °C for 1 h under an inert
sizes, morphologies, and supports that have been reported for atmosphere.115 In this case, the reported overpotential for the
Ni2P-based HER catalysts, most results are in agreement, with Ni5P4 nanoarchitectures was η−10mAcm−2 = −140 mV, in close
an average reported overpotential for Ni2P of η−10mAcm−2 = agreement with other reports.
−125 mV in 0.50 M H2SO4. The Ni12P5 phase has also been identified as an active HER
Other nickel phosphide phases with different compositions electrocatalyst (Figure 8). Ni12P5 nanoparticles on a titanium
and structures have also been explored as HER electrocatalysts substrate112 and Ni12P5/CNT nanohybrids116 were reported to
(Figure 8). Dismukes and co-workers reported micrometer- require overpotentials of approximately η−10mAcm−2 = −105 and
sized Ni5P4 particles prepared by the decomposition of nickel η−10mAcm−2 = −129 mV in acid, respectively. In addition, the
acetylacetonate, trioctylphosphine, and trioctylphosphine oxide HER activity of NiP2 nanosheet arrays supported on carbon
at ∼330 °C.111 Electrodes of the material were fabricated by cloth (NiP2 NS/CC) has also been reported by Sun and co-
pressing 50 mg of dried Ni5P4 powders into 6 mm diameter workers.113 The NiP2 NS/CC material was obtained through a
two-step synthetic strategy. First, Ni(OH)2 nanosheets were
grown on carbon cloth through hydrothermal methods,
followed by phosphidation with NaH2PO2 at 300 °C for 2 h
in an inert atmosphere. The NiP2 NS/CC composites were
highly active for the HER in acidic solutions and required an
overpotential of η−10mAcm−2 = −75 mV. The NiP2 NS/CC
electrodes also maintained their catalytic activity for at least 57
h. However, it is worth noting again that 3D electrode
geometries and other porous architectures can produce
artificially enhanced performance if the reported activity is
not normalized for exposed and/or active surface areas.
Comparisons among the different nickel phosphides are
particularly interesting and instructive (Figure 9). In face-
centered cubic (fcc) Ni, each Ni atom is surrounded by and
coordinated to 12 other Ni atoms. However, as phosphorus is
incorporated and the P:Ni ratio in metal phosphides increases,
the number of direct Ni−Ni interactions progressively
decreases while the Ni−P coordination increases. In the case
of pyrite-type NiP2, no direct Ni−Ni interactions remain and
phosphorus−phosphorus bonding (P−P dimers, as mentioned
previously) is observed. Additionally, the introduction of P into
the structure significantly changes the geometry and arrange-
ment of the Ni sites and leads to a gradual increase in the Ni−
Ni bond distance from around 2.49 Å on the Ni(111) surface to
3.85 Å on the NiP2 (001) surface (Figure 9). These differences
in crystal structure and bonding may have a direct impact on
the catalytic properties of the various metal phosphides. Liu and
co-workers prepared different nanostructured nickel phosphide
phases (Ni12P5, Ni2P, Ni5P4) and compared their activities for
the HER under similar conditions (Figure 10).114 The Ni5P4
phase exhibited superior electrocatalytic performance relative to
Ni12P5 and Ni2P, with the behavior attributed to a higher
Figure 8. (A) TEM image of Ni5P4 particles and (B) corresponding positive charge on the Ni and a stronger ensemble effect from P
linear sweep voltammograms in 0.5 M H2SO4. Adapted with in Ni5P4. However, as can be seen in Figure 10, variations in the
permission from ref 111. Copyright 2015 Royal Society of Chemistry.
particle sizes, morphologies, and surface areas of the samples
(C) TEM image of Ni12P5 nanoparticles and (D) corresponding linear
sweep voltammograms in 0.5 M H2SO4. Adapted with permission could account for some of the observed differences in HER
from ref 112. Copyright 2014 American Chemical Society. (E) TEM activity. Along with similar studies by Kucernak and Sundaram
of NiP2 nanosheets on carbon cloth and (F) corresponding linear comparing Ni12P5 and Ni2P117 and by Dismukes and co-
sweep voltammogram in 0.5 M H2SO4. Adapted with permission from workers comparing Ni5P4 and Ni2P,111 these results suggest
ref 113. Copyright 2014 Royal Society of Chemistry. that the metal-to-phosphorus ratio in metal phosphides might
6027 DOI: 10.1021/acs.chemmater.6b02148
Chem. Mater. 2016, 28, 6017−6044
Chemistry of Materials Review

Figure 10. TEM images of Ni12P5, Ni2P, and Ni5P4 nanoparticles and
their respective linear sweep voltammograms in 0.5 M H2SO4.
Adapted with permission from ref 114. Copyright 2015 Royal Society
of Chemistry.

active HDS catalyst. CoP nanoparticles, which were prepared


by reacting Co nanoparticles with TOP at ∼320 °C, were quasi-
spherical, multifaceted, highly uniform, and hollow, with an
average diameter of 13 ± 2 nm (Figure 11), similar to those of
Ni2P mentioned previously. Electrodes comprised of CoP
nanoparticles on a Ti support outperformed Ni2P and other
nickel-containing phosphides in both activity and stability in
0.50 M H2SO4, requiring an overpotential of η−10mAcm−2 = −75
mV and remaining stable for over 24 h while exhibiting 100%
Faradaic efficiency.
Several groups have since corroborated these results for a
wide variety of CoP morphologies and supports (Figure 12).
For example, Sun and co-workers have reported similar HER
activities in 0.50 M H2SO4 for a number of different CoP-based
Figure 9. Crystal structures of various nickel phosphides spanning a
range of Ni:P ratios. The (P2)2− dimer, which appears in the NiP2 and
electrodes obtained via the low-temperature phosphidation of
NiP3 polyphosphides, is also shown. various cobalt precursors with NaH2PO2. These reports include
carbon nanotubes decorated with CoP nanocrystals (η−10mAcm−2
= −122 mV),129 CoP nanosheet arrays supported on Ti plates
play an important role in affecting the HER performance, with (η−10mAcm−2 = −90 mV,128 self-supported nanoporous cobalt
more phosphorus-rich phases tending to exhibit higher HER phosphide nanowire arrays on carbon cloth (η−10mAcm−2 = −67
activities. While comparing the intrinsic activities of HER mV),127 CoP nanotubes (η−10mAcm−2 = −72 mV),128 three-
catalysts tested using different electrode fabrication methods dimensional interconnected networks of porous CoP nanowires
and loadings poses significant challenges, reports on the HER (η−10mAcm−2 = −100 mV),131 and CoP nanoparticle films on
activity of NiP2 nanosheets (η−10mAcm−2 = −75 mV) also seems carbon cloth (η−10mAcm−2 = −48 mV).132 It is worth noting that
to support these conclusions.113 The observed trends are the increased activity observed in a few of these instances could
consistent with the putative mechanism for the HER on be attributed to the use of highly porous 3D electrodes without
Ni2P(001), which invokes an ensemble effect involving normalization to exposed or active surface areas. Many other
cooperativity of the P and Ni atoms that implies a dependence reports, too numerous to include as an exhaustive list, describe
on the P:Ni ratio. related iterations and comparable HER activities for various
Cobalt Phosphides. Following the initial studies of Ni2P as CoP materials.
an Earth-abundant HER catalyst, cobalt phosphide (CoP) was Interestingly, Lewis, Soriaga, and co-workers have demon-
also identified as an active and acid-stable HER catalyst.128 Like strated that electrodeposited amorphous Co−P films exhibit
Ni2P, CoP is a structurally and compositionally distinct but HER activities comparable to those of crystalline CoP phases
6028 DOI: 10.1021/acs.chemmater.6b02148
Chem. Mater. 2016, 28, 6017−6044
Chemistry of Materials Review

(Figure 13).81 The Co−P films were synthesized by cathodic


deposition from a boric acid solution of Co2+ and H2PO2− on
Cu foils, followed by operando purification to produce an
electrocatalyst with a Co:P atomic ratio of 1:1. The
electrodeposited CoP catalysts showed high activities with an
overpotential of η−10mAcm−2 = −85 mV needed in highly acidic
solutions (0.50 M H2SO4). In agreement with this report, Sun
and co-workers later observed comparable activities using
electrodeposited amorphous Co−P films prepared under
similar experimental conditions (η−10mAcm−2 = −98 mV).
As in the case of Ni2P, despite the wide diversity of synthetic
preparations, sizes, morphologies, and supports that have been
reported, most results are in agreement, reporting an average
overpotential for CoP of η−10mAcm−2 = −80 mV in 0.50 M
H2SO4. Table 2 summarizes the performance of various cobalt
phosphide HER catalysts synthesized under various conditions
and evaluated in 0.5 M H2SO4.
Comparisons among electrocatalysts with different struc-
tures, but identical constituent elements and morphologies, are
important for identifying key structural characteristics that
could lead to higher HER activities. With this in mind, we
recently studied nanostructures of Co2P, a cobalt phosphide
Figure 11. (A) TEM image, (B) high-angle annular dark field phase that is compositionally and structurally distinct from
(HAADF) scanning transmission electron microscopy (STEM) image,
and (C) powder XRD data for CoP nanoparticles. (D) Polarization CoP, by synthesizing by morphologically equivalent Co2P and
data for CoP/Ti electrodes in 0.5 M H2SO4, along with Ti foil and Pt CoP nanoparticles and evaluating their catalytic activity for the
mesh for comparison. Panels (A), (C), and (D) adapted with HER in 0.50 M H2SO4 (Figure 14).135 The Co2P phase
permission from ref 128. Copyright 2014 WILEY-VCH Verlag GmbH displayed slightly higher overpotentials (η−10mAcm−2 = −95 mV)
& Co. KGaA, Weinheim. than CoP (η−10mAcm−2 = −75 mV). This behavior correlates with

Figure 12. Polarization data in 0.5 M H2SO4 and corresponding TEM or SEM images (in the insets) for various nanostructured CoP HER catalysts.
(A) CoP nanocrystals on carbon nanotubes. Adapted with permission from ref 133. Copyright 2014 WILEY-VCH Verlag GmbH & Co. KGaA,
Weinheim. (B) CoP nanosheets on a Ti plate. Adapted with permission from ref 134. Copyright 2014 American Chemical Society. (C) Self-
supported CoP nanowires on carbon cloth. Adapted with permission from ref 129. Copyright 2014 American Chemical Society. (D) CoP
nanosheets on a Ti plate. Adapted with permission from ref 131. Copyright 2014 Royal Society of Chemistry.

6029 DOI: 10.1021/acs.chemmater.6b02148


Chem. Mater. 2016, 28, 6017−6044
Chemistry of Materials Review

Both our group and the Sun group independently identified


MnP-type FeP as an exceptionally active HER electrocatalyst in
acidic, basic, and neutral-pH conditions (Figure 15),152,153 that
outperforms other comparable metal phosphide systems
including CoP and Ni2P. Table 3 summarizes the performance
of various iron phosphide HER catalysts synthesized under
various conditions and evaluated in 0.5 M H2SO4.
Similar to our CoP system, we synthesized colloidal FeP
nanoparticles by reacting premade Fe nanocrystals with TOP at
elevated temperatures.152 The resulting FeP nanoparticles were
spherical and hollow, with an average diameter of 13 ± 2 nm
(Figure 15), and were morphologically comparable to the Ni2P
and CoP nanoparticles that had been synthesized previously.
Working electrodes (FePNP/Ti) with FeP nanoparticles were
prepared by drop-casting appropriate amounts to obtain mass
loadings of 1 mg cm−2 of an FeP nanoparticle suspension onto
Ti substrates, followed by treatment at 450 °C in a reducing
atmosphere to remove any remaining surface ligands. Linear
sweep voltammetry measurements for the FeP NP/Ti electro-
des in acidic solutions (0.5 M H2SO4) demonstrated excep-
tionally high HER activities, with overpotentials of only
Figure 13. (A) Low-magnification and (B) high-magnification SEM
images of electrodeposited amorphous Co−P films. (C) XPS spectrum
η−10mAcm−2 = −50 mV required. Sun and co-workers studied
of Co−P films before and after voltammetry. (D) Linear sweep FeP nanowire arrays supported on conductive Ti plates
voltammograms in 0.5 M H2SO4, along with Pt, Co, and Cu controls. (FePNA/Ti) (Figure 15).133 These arrays were synthesized
Adapted with permission from ref 81. Copyright 2014 American chemically by converting FeOOH nanowire arrays, which were
Chemical Society. hydrothermally grown directly onto Ti plates, into FePNA/Ti by
a low-temperature phosphidation reaction with NaH2PO2. The
the different Co/P ratio of the phases and suggests that the resulting FePNA/Ti system required an overpotential of
increased Co−P character of CoP may provide a higher density η−10mAcm−2 = −55 mV in 0.5 M H2SO4, comparable to the
of proximal cobalt and phosphorus surface atoms, which are behavior of the FePNP/Ti system. While both FeP electrodes
hypothesized to be active sites for the HER. Liu and co-workers exhibited HER activities that were significantly higher than
further explored the influence of phase, structure, and support comparable electrodes of Ni2P and CoP, long-term stability
effects on the HER activity by synthesizing a series of cobalt measurements indicated decreases in activity of the Fe-based
phosphide-based electrocatalysts, including Co2P, CoP, Co2P/ systems after more than 15 h of sustained hydrogen production.
CNTs, CoP/CNTs, Co2P/nitrogen-doped carbon nanotubes In the case of the FePNP/Ti sample, the required overpotential
(NCNTs), and CoP/NCNTs, through the solution-based increased by approximately 52 mV, compared to an increase of
decomposition of various organophosphine reagents.136 Their 11 mV for comparable CoP nanoparticles after 24 h of
results indicated that catalytic activity followed the order CoP/ sustained operation at −20 mA cm−2. While activity is an
NCNTs > Co2P/NCNTs > CoP/CNTs > Co2P/CNTs > CoP important metric for HER performance, a practical catalyst
> Co2P, with the more phosphorus-rich CoP phase out- clearly requires a balance between activity and long-term
performing the corresponding Co2P counterparts in each case. stability. Hence, CoP might be a more robust candidate for
The highest activity of the series was obtained by the CoP/ further development and testing efforts, such as integration
NCNTs catalysts, which required η−10mAcm−2 = −85 mV. While with light absorbers and use in solar-driven water-splitting
preliminary, these reports are in agreement with observations devices.
on the nickel phosphide system, indicating that more Other phases in the iron phosphide system have also been
phosphorus-rich phases exhibit higher HER activities. Fur- studied as HER catalysts. For example, Yang and co-workers
thermore, direct comparisons by our group of multifaceted CoP reported phosphorus-rich FeP2/C nanohybrids prepared by the
nanoparticles with highly branched CoP nanostructures that pyrolysis of ferrocene and red phosphorus in an evacuated and
exposed a high density of (111) facets suggested that the high sealed quartz tube at 500 °C.154 The resulting material was
HER activity of CoP is intrinsic to the system and that shape investigated for the HER in 0.50 M H2SO4 and showed very
control may not play a significant role in defining the low HER activities, requiring an overpotential of −500 mV to
magnitude of the overpotentials required to produce opera- achieve a current density of −5 mA cm−2. In addition, Fe2P
tionally relevant cathodic current densities.137 nanoparticles encapsulated in a sandwichlike graphited carbon
Iron Phosphides. As the cheapest and most terrestrially envelope were reported by Wang and co-workers, with an
abundant transition metal, iron is a particularly interesting observed overpotential for the HER of η−10mAcm−2 = −88 mV.155
target for incorporation into metal phosphide HER catalysts. While still preliminary, the lower HER activity of Fe2P relative
Iron-containing clusters have been found in the active sites of to FeP is in agreement with previously established general
various enzymes including [FeFe] and [Fe]-only hydrogenases, trends that phosphide phases having a higher phosphorus
which have been demonstrated to be highly active and efficient content show improved activities for the HER. However, it is
HER catalysts in biological systems. Interestingly, iron possible that this activity trend may not extend to phosphorus-
phosphides such as FeP and Fe2P are also known hydro- rich phases, such as FeP2, because their different structures and
desulfurization catalysts, yet with significantly lower activities bonding (e.g., P−P bonds) may impact their relative activities
than other transition metal phosphides such as Ni2P.48,49,136,151 and stabilities.
6030 DOI: 10.1021/acs.chemmater.6b02148
Chem. Mater. 2016, 28, 6017−6044
Chemistry of Materials Review

Table 2. Compilation of HER Performance Metrics for Various Cobalt Phosphide Catalysts Synthesized under Different
Conditions and Evaluated in 0.5 M H2SO4
material η−10mAcm−2 η−20mAcm−2 Tafel slope (mV dec−1) exchange current density (A cm−2) loading density (mg cm−2) ref
Co2P/Co −174 - 60 - 58
CoP/CC −49 −59 30.1 - 13.6 78
Co−P films −85 - 50 2.00 × 10−4 - 81
CoP NPs/Ti −75 −85 50 1.40 × 10−4 2.0 128
CoP/CC −67 −100 51 2.88 × 10−4 0.92 129
CoP NTs - −129 60 - - 130
CoP NPs - −297 82 - - 130
CoP NWs/Ti - −95 65 - 0.8 131
CoP NPs/CC −48 - 70 - 4.0 132
CoP/CNT −122 - 54 1.30 × 10−4 0.285 133
CoP/Ti −90 - 43 - 2.0 134
Co2P NPs/Ti −95 −109 45 - 1.0 135
Co2P −406 101 3.20 × 10−3 - 136
CoP −383 90 1.70 × 10−2 - 136
Co2P/CNTs −195 −219 74 3.90 × 10−2 - 136
CoP/CNTs −165 −198 68 6.80 × 10−2 - 136
Co2P/NCNTs - −171 62 1.02 × 10−1 - 136
CoP/NCNTs −79 −99 49 3.20 × 10−1 - 136
branched CoP - −117 48 - 1 137
CoP NWs −110 −142 54 1.60 × 10−4 0.35 138
CoP NPs −221 - 61 5.40 × 10−5 0.35 138
CoP NSs −164 - 87 3.20 × 10−5 0.35 138
Co2P NRs/Ti −134 −167 51.7 - 1 139
Co−P films −94 - 42 - 2.6 140
CoP/RGO ∼−250 - 104.8 4.00 × 10−5 0.29 141
CoP/C −130 - - - - 142
Urchin-like CoP NCs ∼−100 - 46 - 0.28 143
CoP NBAs/Ti −203 - 40 - 1.96 144
CoP@C Nanocables −170 - 61 - 0.35 145
Co2P NW ∼−100 - 45 - 2 146
CoP NW ∼−100 - 41 - 2 146
CoP2/RGO −88 −106 50 - 0.285 147
CoP hollow polyhedron −159 - 59 3.70 × 10−2 0.102 148
CoP particles −355 - 77 5.00 × 10−3 0.102 148
CoP NPAs −393 105 - - 149
CoP NRAs −181 69 - - 149
Co2P/GCE - −160 53 2.10 × 10−4 - 150

Molybdenum and Tungsten Phosphides. Molybdenum overpotentials of η−10mAcm−2 = −90 mV in 0.50 M H2SO4. These
and tungsten phosphides are among some of the most active potentials remained constant after 18 h of galvanostatic testing
HDS catalysts reported to date, making them attractive targets and after over 500 cyclic voltammetric sweeps, indicating
as catalysts for the HER.48,49 Table 4 summarizes the substantial stability under operating conditions. Amorphous
performance of various molybdenum phosphide and tungsten tungsten phosphide (WP) nanoparticles of comparable
phosphide HER catalysts synthesized and evaluated under morphology and size were also obtained using similar synthetic
various conditions. Wang and co-workers reported the HER methods (Figure 17).167 When tested for the HER, WP/Ti
activity of crystalline Mo3P and MoP prepared through bulk electrodes displayed slightly lower activities than MoP,
solid-state approaches.166 The metal-rich Mo3P phase displayed producing a current density of −10 mA cm−2 at an
low HER activity, requiring an overpotential of η−10mAcm−2 = overpotential of −120 mV in 0.50 M H2SO4. Crystalline WP
−500 mV in 0.50 M H2SO4. However, the stoichiometric MoP nanorod arrays on carbon cloth (WP NAs/CC) reported by
phase exhibited much improved performance even in bulk Sun and co-workers displayed similar activity, requiring an
form, exhibiting an overpotential of approximately η−10mAcm−2 = overpotential of η−10mAcm−2 = −130 mV in 0.50 M H2SO4
−125 mV under the same conditions (Figure 16). Our group (Figure 17).168 Under the same conditions, crystalline,
reported on the synthesis and HER performance of amorphous submicrometer WP2 particles reported by the same group
molybdenum phosphide (MoP) nanoparticles having diameters required a slightly higher overpotential of η−10mAcm−2 = −161
of ∼3 nm prepared through the decomposition of Mo(CO)6 mV.169
and TOP in squalane at 320 °C (Figure 16).165 The MoP Copper Phosphides. Compared to the phosphides of the
nanoparticles remained amorphous even after annealing to 450 iron group elements (Fe, Co, Ni), copper phosphides have
°C to remove the organic surface ligands. Working electrodes attracted significantly less attention as HER electrocatalysts,
of the MoP nanoparticles on Ti foil (MoP/Ti) exhibited and only a few studies on Cu3P have been reported to date. An
6031 DOI: 10.1021/acs.chemmater.6b02148
Chem. Mater. 2016, 28, 6017−6044
Chemistry of Materials Review

Figure 15. (A) TEM image of FeP nanoparticles and (B) polarization
data for FeP/Ti electrodes. Adapted with permission from ref 152.
Copyright 2014 American Chemical Society. (C) SEM image of self-
supported FeP nanowires on Ti and (D) corresponding polarization
data in 0.5 M H2SO4. Adapted with permission from ref 153.
Copyright 2014 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.

Cation and Anion Substituted Metal Phosphides.


Mixed-metal transition metal phosphides have recently been
explored as electrocatalysts for the HER through a combined
experimental and theoretical approach. Jaramillo and co-
workers utilized density functional theory calculations to
Figure 14. (A) TEM image of Co2P nanoparticles and (B) predict the hydrogen adsorption free energies, ΔGH, for a
corresponding XRD and SAED patterns. (C) TEM image of CoP series of mixed-metal transition metal phosphides, and
nanoparticles and (D) corresponding XRD and SAED patterns. (E) compared these results to experimentally determined HER
Polarization data of CoP/Ti and Co2P/Ti electrodes in 0.5 M H2SO4 activities to identify optimized metal phosphide catalysts for the
along with Pt mesh and bare Ti foil for comparison. Adapted with HER (Figure 19).149 Applying this methodology, a Fe0.5Co0.5P
permission from ref 135. Copyright 2015 American Chemical Society. alloy was found to be the most active of all the studied systems,
with a near-zero theoretical ΔGH of 0.004 eV and an
initial study by Sun and co-workers on self-supported Cu3P experimental TOFavg of 0.19 ± 0.01 H2 s−1. Nevertheless, the
nanowire arrays grown on commercial porous copper foams study also indicated that monometallic CoP exhibits a
(Cu3P NW/CF) reported catalytic current densities at an comparable HER activity, with a theoretical ΔGH = −0.09 eV
overpotential of η−10mAcm−2 = −143 mV, with only minor and an experimental TOFavg of ∼0.18 H2 s−1, suggesting that
degradation after continuous hydrogen production for 25 h strategies based exclusively on cation solid solutions might not
(Figure 18).177 Likewise, Kong and co-workers described the lead to significant improvements in HER activity. A different
phosphidation of Cu(OH)2 precursors to form Cu3P nano- study by Wang and co-workers also explored the use of mixed-
cubes, which exhibited moderate activities for the HER metal phosphides for the HER, in particular Fe-substituted
(η−10mAcm−2 = −320 mV) when tested on glassy carbon Ni2P sandwich nanocomposites.179 These (Fe,Ni)2P nano-
electrodes (GCE) (Figure 18).178 Recently, our group reported composites were prepared by soaking NiNH4PO4·H2O nano-
the synthesis of phase-pure Cu3P films grown directly onto Cu sheets synthesized by a hydrothermal approach in an aqueous
foils through the vapor-phase phosphidation of commercially FeCl2 solution, followed by a second hydrothermal process to
available metal foils with organophosphine reagents.58 The carbon-coat the nanostructures, followed by reduction in a H2
Cu3P films appeared to be very unstable under the operating atmosphere at 680 °C. The polarization data for the Fe-
conditions, perhaps due to rapid degradation at the Cu3P/Cu substituted Ni2P composites and Ni2P controls indicated
interface. The previous report on Cu3P nanowire arrays enhanced activity for the (Fe,Ni)2P system, which required
presumably was performed on a more robust catalyst/substrate an overpotential of approximately η−10mAcm−2 = −75 mV
interface. Cu3P therefore shows some evidence of moderate compared to the η−10mAcm−2 = −150 mV required by Ni2P to
HER activity but overall is not as stable at this point as the Fe, reach the same current density. However, iron phosphide
Co, and Ni phosphide systems. controls were not included for comparison.
6032 DOI: 10.1021/acs.chemmater.6b02148
Chem. Mater. 2016, 28, 6017−6044
Chemistry of Materials Review

Table 3. Compilation of HER Performance Metrics for Various Iron Phosphide Catalysts Synthesized under Different
Conditions and Evaluated in 0.5 M H2SO4
material η−10mAcm−2 η−20mAcm−2 Tafel slope (mV dec−1) exchange current density (A cm−2) loading density (mg cm−2) ref
Fe2P/Fe −191 - 55 - - 58
FeP NPs/Ti −50 - 37 4.30 × 10−4 1.0 152
FeP NA/Ti −55 −72 38 4.20 × 10−4 0.285 153
FeP2/C - - 66 1.75 × 10−6 0.425 154
Fe2P/GCS ∼−100 - 49 - 0.36 155
FeP/CC −34 −43 29.2 6.80 × 10−4 4.9 156
HMFeP@C −115 - 56 1.91 × 10−4 - 157
FeP NWs −96 - 39 1.70 × 10−4 - 158
FeP2 NWs −61 - 37 5.50 × 10−4 - 158
FeP/GS −123 - 50 1.20 × 10−4 0.28 159
FeP NAs/CC −58 - 45 5.00 × 10−4 1.5 160
FeP NAs/Ti −85 - 60 - - 161
FeP NSs ∼−240 - 67 - - 162
FeP/CC - −54 32 5.90 × 10−4 4.2 163
FeP-CS −112 - 58 2.20 × 10−4 0.28 164

Table 4. Compilation of HER Performance Metrics for Various Molybdenum and Tungsten Phosphide Catalysts Synthesized
under Different Conditions and Evaluated in 0.5 M H2SO4
material η−10mAcm−2 η−20mAcm−2 Tafel slope (mV dec−1) exchange current density (A cm−2) loading density (mg cm−2) ref
−4
MoP/Ti −90 −105 45 1.20 × 10 1.0 165
MoP ∼−130 - 54 3.40 × 10−5 - 166
Mo3P ∼−500 - 147 - - 166
WP/Ti −120 −140 54 4.50 × 10−5 1.0 167
WP NAs/CC −130 - 69 2.90 × 10−4 2.0 168
WP2 SMPs −161 - 65 1.70 × 10−5 0.50 169
MoP NPs −125 - 54 8.60 × 10−5 0.36 170
MoP/GCE −246 - 60 4.15 × 10−6 0.071 171
MoP/CF −200 - 56.4 - 0.36 172
MoP MPs −150 - 50 1 × 10−5 0.1 173
MoP flakes - −155 71.77 - 1.425 174
MoP2 NPs/Mo −143 - 57 6.00 × 10−5 - 175
WP2 nanorods −148 - 52 1.30 × 10−5 - 176

Anion substitution has also been explored as an approach for = −48 mV in 0.50 M H2SO4. This activity places CoPS among
improving the HER activity of metal phosphides. Based on the the best Earth-abundant HER electrocatalysts reported to date
mechanistic commonalities shared by many HER and HDS and perhaps more importantly demonstrates that tuning the
catalysts, Jaramillo and co-workers hypothesized that the electronic structure and reactivity of catalysts by substituting
introduction of sulfur onto metal phosphides might lead to nonmetallic atoms can serve as a powerful strategy to enhance
enhancements in their catalytic activity and stability for the the activity of transition metal compounds.
HER, as surface phosphosulfides are considered to be the sites Mechanistic and Surface Studies of Metal Phos-
of enhanced activity during HDS catalysis by both Ni2P and phides. Despite great interest in the use of metal phosphides
MoP.180 To explore this possibility, sulfur was introduced into as catalysts for the HER, to date only a few mechanistic and
the surface region of MoP films by postsulfidation in a quartz surface studies have been reported. The surface structure of the
tube furnace at 400 °C under a mixture of 10% H2S in H2. The catalysts under operating conditions, the identity of the
introduction of sulfur led to significant improvements in the catalytically active sites, and key details of the HER mechanism
electrocatalytic HER activity of the MoP films, with the MoP|S on metal phosphide surfaces remain largely unknown. The
catalyst requiring an overpotential of only η−10mAcm−2 = −64 existing studies have focused mainly on density functional
compared to the required η−10mAcm−2 = −117 mV of the theory (DFT) calculations of the hydrogen absorption energies
unmodified MoP films (Figure 20). Similarly, Jin and co- on various metal phosphide surfaces. Based on such
workers developed a nanostructured pyrite-type cobalt calculations, Liu and Rodriguez associated the HER behavior
phosphosulfide (CoPS) catalyst for the HER (Figure 21).181 of Ni2P to an ensemble effect, in which the presence of P atoms
The CoPS electrodes were prepared by reacting a series of dilutes the number of highly active Ni sites on the surface,
cobalt-based precursors at 500 °C with sulfur and phosphorus potentially leading to more moderate binding of the products
vapors produced by the evaporation of the elemental powders and intermediates (Figure 22).45 Importantly, they also
in an inert atmosphere. High-surface-area CoPS nanoplates identified the P sites as active, suggesting that the presence of
grown on carbon paper displayed the highest HER perform- both proton-acceptor and hydride-acceptor centers on the
ance of the studied samples, achieving a geometric current surface could be playing a role in facilitating the HER. Based on
density of −10 mA cm−2 at an overpotential of only η−10mAcm−2 DFT calculations, Hu and co-workers suggested that
6033 DOI: 10.1021/acs.chemmater.6b02148
Chem. Mater. 2016, 28, 6017−6044
Chemistry of Materials Review

Figure 16. (A) TEM image of amorphous MoP nanoparticles, (B) corresponding XRD patterns at different annealing temperatures, and (C)
polarization data of bulk and nano-MoP in 0.5 M H2SO4, along with a Pt control. Adapted with permission from ref 165. Copyright 2014 American
Chemical Society. (D) SEM image of crystalline MoP particles, (E) corresponding XRD patterns, and (F) polarization data of MoP and Mo3P in 0.5
M H2SO4, along with a Pt control. Adapted from ref 166. Copyright 2014 Royal Society of Chemistry.

Figure 17. (A) TEM image of amorphous WP nanoparticles and (B)


corresponding polarization data in 0.5 M H2SO4, along with a Pt Figure 18. (A) SEM image of Cu3P nanocubes and (B) corresponding
control. Adapted with permission from ref 167. Copyright 2014 Royal polarization data in 0.5 M H2SO4, along with a Pt control. Adapted
Society of Chemistry. (C) TEM image of crystalline WP2 nano- with permission from ref 177. Copyright 2014 WILEY-VCH Verlag
particles and (D) corresponding polarization data in in 0.5 M H2SO4, GmbH & Co. KGaA, Weinheim. (C) SEM image of self-supported
along with a Pt control. Adapted with permission from ref 169. Cu3P nanowires on carbon cloth and (D) corresponding polarization
Copyright 2015 American Chemical Society. data in 0.5 M H2SO4. Adapted with permission from ref 178.
Copyright 2016 Royal Society of Chemistry.
phosphorus-stabilized Ni-bridge sites on Ni2P could also serve
as active sites and provide moderate binding to hydrogen nickel phosphide crystals, including low-energy electron
atoms.182 Further studies by Jaramillo28 and Wang166 indicated diffraction (LEED), scanning tunneling microscopy (STM),
that several metal phosphide surfaces have hydrogen absorption and photoemission electron microscopy (PEEM, indicate the
energies that are close to thermo-neutral, but specific structural presence of reconstructed surface structures, implying that the
or mechanistic characteristics that could lead to better catalyst bulk terminated structures of metal phosphides may not be
design have not been explicitly identified. Additionally, several stable.183,184 However, these and similar experiments are
experimental surface studies by Asakura and co-workers on carried out in environments that are very different from those
6034 DOI: 10.1021/acs.chemmater.6b02148
Chem. Mater. 2016, 28, 6017−6044
Chemistry of Materials Review

Figure 21. Polarization data in 0.5 M H2SO4 for CoPS films,


nanowires, and nanoplates, along with a Pt control. Adapted with
permission from ref 181. Copyright 2015 Macmillan Publishers
Limited.

Figure 19. (A) Polarization data for several transition metal


phosphides in 0.5 M H2SO4, along with a Pt control. (B) Volcano
plot showing the TOF of several transition metal phosphides as a
function of hydrogen adsorption free energies. The color-coded key
for both panels is shown in (B). Adapted with permission from ref
149. Copyright 2015 Royal Society of Chemistry.
Figure 22. Optimized structures for the (001) surface of Ni2P, during
different steps in hydrogen-evolution reaction. Hydrogen atoms shown
in white, nickel in blue, and phosphorus in purple. Reproduced with
permission from ref 45. Copyright 2005 American Chemical Society.

used during actual HER operating conditions. In situ and ex


situ characterization studies of transition metal phosphide
materials under HER conditions could thus provide insights
into the origin of the high catalytic activity and ultimately the
mechanism by which the HER occurs.

6. INTEGRATION WITH LIGHT ABSORBERS


Integration of kinetically competent catalysts with light
absorbers remains challenging because of the delicate interplay
between light absorption, carrier collection, charge transfer at
interfaces, and catalysis. For the HER, most studies involving
the integration of catalysts with light-absorbing materials have
focused on Pt and MoS2 but recently have expanded to metal
phosphide systems. In this section we discuss key metrics,
experimental conditions, specific challenges, and recent high-
lights involving the coupling of metal phosphide electrocatalysts
Figure 20. (A) Low-magnification and (B) high-magnification SEM to photocathodes.
images of MoP|S. (C) Polarization data for MoP and MoP|S in 0.5 M Background, Considerations, and Metrics. Solar-
H2SO4, along with a Pt control. Adapted with permission from ref 180. powered hydrogen production depends critically on the
Copyright 2014 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim. judicious choice of materials to maximize the overall system
solar-to-hydrogen efficiency, ηSTH. This quantity represents the
ratio of useful power out (H2 and O2) to the total power of
6035 DOI: 10.1021/acs.chemmater.6b02148
Chem. Mater. 2016, 28, 6017−6044
Chemistry of Materials Review

sunlight illuminating the system. A single light absorber can mass transport (i.e., should be performed with rapid stirring of
potentially reach a maximum efficiency of 11% while a stacked the solution), uncompensated resistance overpotentials (by
configuration consisting of two light absorbers may yield in correcting for the iR drop whenever possible), cell geometry,
excess of 22% (Figure 23).185 Hence, on the path toward the nature of the counter electrode (by avoiding precious
metals which could contaminate the cell) or the reference
electrode (by using a well-defined Nernstian potential with a
known pH and maintaining 1 atm of H2), or the nature of the
illumination source (by using an artificial light that matches well
with the solar spectrum). For example, for metal working
electrodes, the solution resistance can be electronically
compensated using a current-interrupt method. In contrast,
for a photoelectrode the uncompensated resistance of the setup
should be determined with a glassy carbon electrode in place of
the semiconductor. Instead of electronic compensation, a
Luggin capillary reference electrode can be used to minimize
solution resistance for either metal or semiconducting electro-
des. Pt contamination is a common pitfall in the evaluation of
HER catalysts, because even submonolayer deposits of Pt can
result in high HER activities and therefore produce apparent
enhancements in activity and/or stability.188,189 When possible,
it is best to avoid the use of Pt counter electrodes, by using
high-purity carbon rods or other Pt-free electrodes instead. The
catalyst and synthesis preparation should additionally be
performed with precautions taken to minimize the possibility
of noble metal contamination, e.g., all glassware should be
carefully cleaned using aqua regia. The possibility of trace
surface contamination underscores the importance of utilizing
XPS to check for trace metal impurities. Together the open-
circuit voltage (Voc), short-circuit photocurrent density (JSC),
and fill factor ( f f) form the basis set of metrics for evaluating
the characteristics of a given photoelectrode. The open-circuit
potential is the potential measured at zero current with a high-
quality multimeter. The magnitude of the photovoltage is
Figure 23. Solar-to-hydrogen efficiency for (A) single and (B) double governed by the band gap of the material, practically reaching
light-absorber water splitting systems. Adapted with permission from approximately two-thirds of the band gap in well-engineered
ref 185. Copyright 2013 Royal Society of Chemistry. systems. Photovoltages are lower in practice due to
recombination.
economic viability, considerable efforts have been invested in In contrast to the short-circuit current density, which is the
achieving high efficiencies by developing photoelectrosynthetic current measured with a high-resistivity multimeter at zero
cells driven by tandem photoabsorbers, with the highest voltage and generally depends linearly on the photon flux
achieved ηSTH of 18% reported by Licht and co-workers in reaching the semiconductor, the open-circuit voltage depends
2000 using Pt black and RuO2 as electrocatalysts for the logarithmically on the illumination intensity. The actual
hydrogen- and oxygen-evolution reactions, respectively.186 photocurrent depends on the rate at which the current
While ηSTH remains the key figure of merit for an entire cell, approaches the limiting photocurrent, with the potential
optimization occurs by understanding the behavior of the dependence of the current called the fill factor, reflecting the
individual components comprising the whole system. An “squareness” of the J−V characteristic of the photoelectrode.
important metric for a single component is the ideal Hence J−V measurements obtained with a potentiostat yield
regenerative cell efficiency, ηIRC, which represents the ratio of information about the light absorption and photocarrier
the total electrical power output relative to the input solar collection. The Faradaic yield for H2 production should also
power for each half reaction.187 This differs from ηSTH because be determined. The ultimate test is to measure all of these
only electrical power is generated while measuring ηIRC instead parameters after prolonged operation, which ultimately must be
of a net chemical reaction only driven by solar energy for ηSTH. stable on the order of 10 years for a commercially viable
Ideally, the photoelectrode performance should be determined electrode with ηSTH > 10%, but more tractably on the bench by
without a coupled catalyst, with an electrochemically reversible measuring ηIRC after accelerated degradation with an appro-
one-electron redox couple such as methyl viologen (MV2+/+), priate benchmarking protocol.190
to establish the maximum performance obtainable from the Structuring of Photoelectrodes. One guiding principle to
absorber. This measurement aids in distinguishing issues produce high ηIRC for a photocathode relies on maximizing
intrinsic to the semiconductor from those due to the coupling simultaneously the photovoltage and short-circuit current
of a catalyst with the electrode surface. density, while accommodating thick catalyst overlayers. In a
Fundamental electrode characteristics can be determined semiconducting device, both the voltage and current intimately
using three-electrode measurements to probe the current depend on the generation rate of electron−hole pairs. This rate,
density as a function of applied potential on each photo- in turn, depends on the flux of photons with energies greater
electrode. Additionally, experiments should not be limited by than the bandgap. Illumination can occur on either the backside
6036 DOI: 10.1021/acs.chemmater.6b02148
Chem. Mater. 2016, 28, 6017−6044
Chemistry of Materials Review

or front side of a solar cell, so the absorbance and reflectivity of


a catalyst are important properties. For frontside exposure to
the sun, it is critical for the catalyst film to have low light
absorption, but for backside illumination, light absorption by
the catalyst is not crucial.191 Micro- and nanostructured
photoelectrodes may offer an alternative approach for
decoupling light absorption of a catalyst film from the light
absorption necessary to produce photogenerated carriers within
the semiconductor. High-aspect-ratio microwire or nanowire
devices, for example, can utilize catalysts with lower TOFs by
enabling higher mass loadings of the catalytic materials without
compromising necessary light absorption by the semiconductor.
For example, when CoP nanoparticle HER catalysts were
coupled with n+p-Si microwires, high catalyst mass loadings
were easily accommodated on the microwires (Figure 24). The

Figure 24. (A) SEM image of CoP nanoparticles on n+p-Si microwire


(MW) arrays. (B) Effect of catalyst loading on the current density vs
potential behavior of n+p-Si planar photocathodes in contact with
H2(g)-saturated 0.50 M H2SO4(aq) and under 100 mW cm−2 of
AM1.5G simulated solar illumination. Adapted with permission from
ref 192. Copyright 2015 American Chemical Society.

resulting performance (ηIRC of 2%) was limited by parasitic


resistance, presumably due to poor contact,192 rather than by Figure 25. (A) SEM images of micropyramid-structured silicon
light absorption, as would have occurred with a planar photocathodes coated with 10 nm CoPS showing top and cross-
geometry. Likewise, Jin and co-workers used a structured section views. (B) J−V curves in 0.5 M H2SO4 and under one sun
photoelectrode with a catalytically active metal phosphide film illumination of microstructured n+p-Si electrodes coated with 3, 5, 7.5,
that had intimate contact between the catalyst and the light 10, and 15 nm of CoPS and 5 nm of Pt. A planar CoPS/p-Si electrode
absorber. Integration of CoPS with micropyramid n+-p silicon is shown for comparison. Adapted with permission from ref 181.
substrates was achieved first by electron-beam deposition of Copyright 2015 Macmillan Publishers Limited.
cobalt, followed by heating in a furnace in the presence of 1:1
sulfur:phosphorus powder. Using this approach, ηIRC was nearly over the course of an hour. However, a loss in activity occurred
5% (Figure 25).181 shortly thereafter, and this decline in performance was
Maintaining Charge Separation. Various strategies have attributed to passivation by silicon oxide. Zhang obtained
been developed for enhancing the separation of charge carriers, ηIRC ∼ 3% without using a p−n junction by dipping p-type
including judicious selection of materials with optimal proper- silicon nanowires into a solution of iron nitrate and then
ties and careful engineering of the junction between the catalyst heating in the presence of NaH2PO2, to yield FeP. The
and light absorber. Considering the coupling of metal photoelectrode remained active for over 1 h of operation.195
phosphide HER catalysts to light-absorbing semiconductors, a Another approach for enhancing the separation of charge
large barrier height and a correspondingly high photovoltage carriers relies on forming a metallurgical junction created by
in an ideal Schottky contactcould be obtainable with a low diffusing dopants that create an appropriate built-in electric
work function catalyst contacting the p-type semiconductor. field necessary for photocarrier collection. The built-in potential
The work functions of metal phosphide HER catalysts have not in this strategy does not depend on the work function of the
yet been determined. However, Liu and co-workers reported catalyst; however, the formation of well-behaved buried
that a heterostructure of CoSe and p-type Si exhibited ηIRC of junctions remains an unsolved challenge for many semi-
0.6%.193 Zhang and co-workers anchored Ni12P5 nanoparticles conductor systems. Both approaches potentially face challenges
onto p-type silicon nanowires and illuminated for 2.8 h with associated with colloidially synthesized nanoparticle catalysts
ηIRC = 3%.112 Liu observed an initial ηIRC of 2% upon coupling capped by organic ligands, which require high temperatures to
CoP nanoparticles with p-type silicon nanowire-array photo- expose the catalytically active sites. Interstitial defects
cathodes.194 This system was fabricated by first electro- introduced by diffusion of a foreign material into the bulk of
depositing Co and then using phosphidation at 500 °C to the semiconductor can lead to trap-assisted recombination,
produce crystalline CoP particles. Significant enhancement of thereby lowering the performance of the photoelectrode. Low-
catalytic activity (compared with a bare surface) was observed temperature processing is therefore desirable for integrating
6037 DOI: 10.1021/acs.chemmater.6b02148
Chem. Mater. 2016, 28, 6017−6044
Chemistry of Materials Review

materials with many photocathodes. For example, preliminarily produced H2 for over 10 h, obtaining a maximum H2
results suggest that extension of an electrodeposition production rate of 251 μmol h−1 with CoP.198
procedure81 for growing amorphous CoP onto planar p-WSe2 Outlook. Metal phosphides demonstrate promising photo-
or n+p-Si appears tenable. catalytic activity when coupled to light absorbers. Future efforts
Charge Transfer Between Catalyst and Light Absorb- should also include longer stability windows, which will be
er. The charge transfer of electrons from the absorber to the important for evaluating commercially viable photocathode
catalyst depends on the electrical and mechanical contact. Poor systems. The design of an efficient water-splitting system will be
contact could result in parasitic series resistances, thereby aided by reporting the intrinsic resistivities of these catalyst
detrimentally affecting the device performance. Charge transfer films and including absorption and reflection spectra of the
within the bulk of the catalyst could also result in a parasitic films. An expanded suite of deposition techniques that ensures
resistance. Hence, a thin layer or low-resistivity catalyst film is intimate contact between the semiconductor and catalyst film,
desirable. The resistivities of metal phosphides have not yet in addition to more complete characterization, will be useful for
been reported. Ensuring intimate electrical and mechanical integrating these materials into devices.
contact between nanoparticles and substrates remains an open A challenge facing the construction of a functional,
challenge. Appropriate binding agents could result in improved inexpensive water-splitting device operating in acidic regimes
adhesion and electron transport. An alternative strategy is to is the lack of Earth-abundant oxygen-evolving catalysts. As the
deposit the materials directly by vacuum deposition, potentially search continues for compatible families of anode materials,
resulting in thin, conformal, highly active electrodes. Jaramillo another approach relies on the incorporation of a bipolar
and co-workers demonstrated that Co metal deposited by membrane situated in between an acidic catholyte and an
electron-beam evaporation onto n+p-Si, followed by phosphi- alkaline anolyte. The membrane simultaneously maintains a
dation, yielded highly active CoP on planar silicon (Figure steady-state pH difference between each compartment and
26).196 The TOF values were nearly 10 times those reported prevents gaseous product crossover, thereby allowing the use of
an optimized acid-compatible photocathode and base-compat-
ible photoanode. Sun demonstrated solar-driven water-spitting
with an ηSTH of 10% for >100 h using a GaAs/InGaP tandem
stack immersed in an anolyte compartment (pH 9.3) driving
the OER with a NiOx coating and a back contact wired to a Ti
mesh coated with CoP carrying out the HER in the catholyte
compartment.199 In addition to the overpotential losses from
catalysis, the bipolar membrane operating at a current density
of 10 mA cm−2 results in an additional 400−500 mV resistive
loss. Luo similarly reported a device using a pervoskite light
harvester located outside of the solutions wired to a NiFeOx
electrode in the anolyte (pH 13.9) and a CoP on Ti foil
electrode in the catholyte (pH 0.4) giving ηSTH of 12% for
nearly 100 h.200 Future work may result in completely wireless
devices.
Figure 26. Linear sweep voltammograms of a CoP thin film on n+p-Si
photocathodes under one sun illumination before and after 24 h of 7. CONCLUSIONS
operation in 0.5 M H2SO4. Adapted with permission from ref 196. Transition metal phosphides have emerged as a robust family of
Copyright 2012 Macmillan Publishers Limited.
Earth-abundant catalysts for the HER. The high catalytic
activity and stability displayed by these materials, including in
strongly acidic aqueous electrolytes, has motivated extensive
previously, allowing for thinner films and suggesting a more efforts in synthesis, characterization, catalytic testing, and device
intrinsically active material than the nanoparticles and/or better integration. Currently, several phosphides of Ni, Co, Fe, Mo,
interfacial contact between the substrate and catalyst film. The and W are considered to be highly promising HER catalysts,
resulting photocathode exhibited ηIRC = 5% while showing and anion substitution appears to be a particularly useful
stability for 24 h of operation. approach for further improving their catalytic performance.
Particle-Based Absorbers. One technoeconomic analysis Despite these discoveries and ongoing improvements as more
suggests that H2 generation systems using multijunction systems are being interrogated, mechanistic studies remain
semiconductors have an undesirably high calculated H2 cost scarce. In situ studies, coupled with computational inves-
($6 per kg of H2), which is driven primarily by the expected tigations, will be especially useful for furthering our under-
cost of cell packaging. In contrast, water splitting using colloidal standing of how the HER proceeds on transition metal
systems that function in inexpensive plastic bags could phosphide surfaces under operational conditions. Such studies
potentially result in a system that has a very low cost of H2 will also help to reveal design guidelines for producing catalytic
($1.49 per kg of H2).197 Thus, a particle-bed PEC device in materials that expose the highest possible density of surface
bags is an attractive option for an economically scalable active sites. Finally, the integration of metal phosphides with
hydrogen generation platform, provided that an intrinsically light absorbers is still at an early stage. Nonetheless, the
safe system can be designed such that H2 and O2 are not development of methods capable of producing favorable
evolved in the same bag at the same time. Along these lines, interfaces between catalysts and light absorbers, as well as
FeP nanoparticles anchored onto TiO2 under UV illumination novel photocathode architectures, may be key steps toward
produced H2 for 16 h at nearly the same rate as Pt.152 Fu and realizing integrated water-splitting systems comprised entirely
co-workers coupled CoP, Ni2P, and Cu3P to CdS particles and of inexpensive and Earth-abundant materials.
6038 DOI: 10.1021/acs.chemmater.6b02148
Chem. Mater. 2016, 28, 6017−6044
Chemistry of Materials Review

■ AUTHOR INFORMATION
Corresponding Authors
Structure of [Fe]-Hydrogenase Reveals the Geometry of the Active
Site. Science 2008, 321, 572−575.
(16) Burgess, B. K.; Lowe, D. J. Mechanism of Molybdenum
*(R.E.S.) E-mail: res20@psu.edu. Nitrogenase. Chem. Rev. 1996, 96, 2983−3012.
*(N.S.L.) E-mail: nslewis@caltech.edu. (17) Le Gall, T.; Ibrahim, S. K.; Gormal, C. A.; Smith, B. E.; Pickett,
Author Contributions C. J. The Isolated Iron-Molybdenum Cofactor of Nitrogenase
§ Catalyses Hydrogen Evolution at High Potential. Chem. Commun.
(J.F.C. and C.G.R.) These authors contributed equally.
1999, 9, 773−774.
Notes
(18) Darensbourg, M. Y.; Bethel, R. D. Biomimetic Chemistry:
The authors declare no competing financial interest.


Merging the old with the New. Nat. Chem. 2012, 4, 11−13.
(19) Mejia-Rodriguez, R.; Chong, D.; Reibenspies, J. H.; Soriaga, M.
ACKNOWLEDGMENTS P.; Darensbourg, M. Y. The Hydrophilic Phosphatriazaadamantane
This work was supported by the National Science Foundation Ligand in the Development of H2 Production Electrocatalysts: Iron
(NSF) Center for Chemical Innovation in Solar Fuels (CHE- Hydrogenase Model Complexes. J. Am. Chem. Soc. 2004, 126, 12004−
1305124). C.W.R. thanks the Link Energy Foundation for a 12014.
graduate research fellowship. (20) Le Goff, A.; Artero, V.; Jousselme, B.; Tran, P. D.; Guillet, N.;


Métayé, R.; Fihri, A.; Palacin, S.; Fontecave, M. From Hydrogenases to
Noble Metal−Free Catalytic Nanomaterials for H2 Production and
REFERENCES Uptake. Science 2009, 326, 1384−1387.
(1) Lewis, N. S.; Nocera, D. G. Powering The Planet: Chemical (21) Helm, M. L.; Stewart, M. P.; Bullock, R. M.; Dubois, M. R.;
Challenges N Solar Energy Utilization. Proc. Natl. Acad. Sci. U. S. A. Dubois, D. L. A Synthetic Nickel Electrocatalyst with a Turnover
2006, 103, 15729−15735. Frequency Above 100,000 S−1 for H2 Production. Science 2011, 333,
(2) Hunter, B. M.; Blakemore, J. D.; Deimund, M.; Gray, H. B.; 863−866.
Winkler, J. R.; Müller, A. M. Highly Active Mixed-Metal Nanosheet (22) Liu, T.; Dubois, D. L.; Bullock, R. M. An Iron Complex with
Water Oxidation Catalysts Made by Pulsed-Laser Ablation in Liquids. Pendent Amines as a Molecular Electrocatalyst for Oxidation of
J. Am. Chem. Soc. 2014, 136, 13118−13121. Hydrogen. Nat. Chem. 2013, 5, 228−233.
(3) A National Vision of America’s Transition to a Hydrogen (23) Liu, T.; Darensbourg, M. Y. A Mixed-Valent, Fe(II)Fe(I),
EconomyTo 2030 and Beyond; United States Department of Energy:
Diiron Complex Reproduces the Unique Rotated State of the [FeFe]
Washington, DC, 2002.
Hydrogenase Active Site. J. Am. Chem. Soc. 2007, 129, 7008−7009.
(4) Cheng, X.; Shi, Z.; Glass, N.; Zhang, L.; Zhang, J.; Song, D.; Liu,
(24) Rose, M. J.; Gray, H. B.; Winkler, J. R. Hydrogen Generation
Z. S.; Wang, H.; Shen, J. A Review of PEM Hydrogen Fuel Cell
Contamination: Impacts, Mechanisms, and Mitigation. J. Power Sources Catalyzed by Fluorinated Diglyoxime−Iron Complexes at Low
2007, 165, 739−756. Overpotentials. J. Am. Chem. Soc. 2012, 134 (20), 8310−8313.
(5) McKone, J. R.; Lewis, N. S.; Gray, H. B. Will Solar-Driven Water- (25) Hu, X.; Brunschwig, B. S.; Peters, J. C. Electrocatalytic
Splitting Devices See The Light of Day? Chem. Mater. 2014, 26, 407− Hydrogen Evolution at Low Overpotentials by Cobalt Macrocyclic
414. Glyoxime and Tetraimine Complexes. J. Am. Chem. Soc. 2007, 129
(6) Xiang, C.; Papadantonakis, K. M.; Lewis, N. S. Principles and (29), 8988−8998.
Implementations of Electrolysis Systems for Water Splitting. Mater. (26) Kristensen, J.; Zhang, J.; Chorkendorff, I.; Ulstrup, J.; Ooi, B. L.
Horiz. 2016, 3, 169−173. Assembled Monolayers Of Mo3S44+ Clusters on Well-Defined
(7) Hernandez-Pagan, E. A.; Vargas-Barbosa, N. M.; Wang, T.; Zhao, Surfaces. Dalton Trans. 2006, 3985−3990.
Y.; Smotkin, E. S.; Mallouk, T. E. Resistance and Polarization Losses in (27) Karunadasa, H. I.; Montalvo, E.; Sun, Y.; Majda, M.; Long, J. R.;
Aqueous Buffer-Membrane Electrolytes for Water-Splitting Photo- Chang, C. J. A Molecular MoS2 Edge Site Mimic for Catalytic
electrochemical Cells. Energy Environ. Sci. 2012, 5, 7582−7589. Hydrogen Generation. Science 2012, 335, 698−702.
(8) Lewis, N. S. Research Opportunities to Advance Solar Energy (28) Kibsgaard, J.; Jaramillo, T. F.; Besenbacher, F. Building an
Utilization. Science 2016, 351, aad1920. Appropriate Active-Site Motif Into a Hydrogen-Evolution Catalyst
(9) Gray, H. B. Powering The Planet with Solar Fuel. Nat. Chem. with Thiomolybdate [Mo3S13]2− Clusters. Nat. Chem. 2014, 6, 248−
2009, 1, 7. 253.
(10) Parkin, A.; Cavazza, C.; Fontecilla-Camps, J. C.; Armstrong, F. (29) Bethel, R. D.; Darensbourg, M. Y. Bioinorganic Chemistry:
A. Electrochemical Investigations of the Interconversions Between Enzymes Activated by Synthetic Components. Nature 2013, 499, 40−
Catalytic and Inhibited States of the [FeFe]-Hydrogenase From 41.
Desulfovibrio Desulfuricans. J. Am. Chem. Soc. 2006, 128, 16808− (30) McKone, J. R.; Marinescu, S. C.; Brunschwig, B. S.; Winkler, J.
16815. R.; Gray, H. B. Earth-Abundant Hydrogen Evolution Electrocatalysts.
(11) Madden, C.; Vaughn, M. D.; Díez-Pérez, I.; Brown, K. A.; King, Chem. Sci. 2014, 5, 865−878.
P. W.; Gust, D.; Moore, A. L.; Moore, T. A. Catalytic Turnover of (31) Walter, M. G.; Warren, E. L.; McKone, J. R.; Boettcher, S. W.;
[FeFe]-Hydrogenase Based On Single-Molecule Imaging. J. Am. Chem.
Mi, Q.; Santori, E. A.; Lewis, N. S. Solar Water Splitting Cells. Chem.
Soc. 2012, 134, 1577−1582.
Rev. 2010, 110 (11), 6446−6473.
(12) Volbeda, A.; Charon, M.-H.; Piras, C.; Hatchikian, E. C.; Frey,
M.; Fontecilla-Camps, J. C. Crystal Structure of the Nickel-Iron (32) Brown, D. E.; Mahmood, M. N.; Turner, A. K.; Hall, S. M.;
Hydrogenase from Desulfovibrio Gigas. Nature 1995, 373, 580−587. Fogarty, P. O. Low Overvoltage Electrocatalysts for Hydrogen
(13) Mcintosh, C. L.; Germer, F.; Schulz, R.; Appel, J.; Jones, A. K. Evolving Electrodes. Int. J. Hydrogen Energy 1982, 7, 405−410.
The [NiFe]-Hydrogenase of the Cyanobacterium Synechocystis Sp. (33) Brown, D. E.; Mahmood, M. N.; Man, M. C. M.; Turner, A. K.
PCC 6803 Works Bidirectionally With a Bias to H2 Production. J. Am. Preparation and Characterization of Low Overvoltage Transition
Chem. Soc. 2011, 133, 11308−11319. Metal Alloy Electrocatalysts for Hydrogen Evolution In Alkaline
(14) Tard, C.; Liu, X.; Ibrahim, S. K.; Bruschi, M.; Gioia, L. D.; Solutions. Electrochim. Acta 1984, 29, 1551−1556.
Davies, S. C.; Yang, X.; Wang, L.-S.; Sawers, G.; Pickett, C. J. Synthesis (34) Nocera, D. G. The Artificial Leaf. Acc. Chem. Res. 2012, 45,
of the H-Cluster Framework of Iron-Only Hydrogenase. Nature 2005, 767−776.
433, 610−613. (35) McKone, J. R.; Sadtler, B. F.; Werlang, C. A.; Lewis, N. S.; Gray,
(15) Shima, S.; Pilak, O.; Vogt, S.; Schick, M.; Stagni, M. S.; Meyer- H. B. Ni−Mo Nanopowders for Efficient Electrochemical Hydrogen
Klaucke, W.; Warkentin, E.; Thauer, R. K.; Ermler, U. The Crystal Evolution. ACS Catal. 2013, 3, 166−169.

6039 DOI: 10.1021/acs.chemmater.6b02148


Chem. Mater. 2016, 28, 6017−6044
Chemistry of Materials Review

(36) Vrubel, H.; Hu, X. Molybdenum Boride and Carbide Catalyze (56) Otani, S.; Ohashi, N. Preparation of Ni2P and Fe2P Single
Hydrogen Evolution in Both Acidic and Basic Solutions. Angew. Chem., Crystals by the Floating-Zone Method. J. Ceram. Soc. Jpn. 2013, 121,
Int. Ed. 2012, 51, 12703−12878. 331−332.
(37) Chen, W. F.; Wang, C. H.; Sasaki, K.; Marinkovic, N.; Xu, W.; (57) Henkes, A. E.; Schaak, R. E. Trioctylphosphine: A General
Muckerman, J. T.; Zhu, Y.; Adzic, R. R. Highly Active and Durable Phosphorus Source for the Low-Temperature Conversion of Metals
Nanostructured Molybdenum Carbide Electrocatalysts for Hydrogen Into Metal Phosphides. Chem. Mater. 2007, 19, 4234−4242.
Production. Energy Environ. Sci. 2013, 6, 943−951. (58) Read, C. G.; Callejas, J. F.; Holder, C. F.; Schaak, R. E. General
(38) Cao, B.; Veith, G. M.; Neuefeind, J. C.; Adzic, R. R.; Khalifah, P. Strategy for fhe Synthesis of Transition Metal Phosphide Films for
G. Mixed Close-Packed Cobalt Molybdenum Nitrides as Non-Noble Electrocatalytic Hydrogen and Oxygen Evolution. ACS Appl. Mater.
Metal Electrocatalysts for the Hydrogen Evolution Reaction. J. Am. Interfaces 2016, 8, 12798−12803.
Chem. Soc. 2013, 135, 19186−19192. (59) Panneerselvam, A.; Malik, M. A.; Afzaal, M.; O’Brien, P.;
(39) Chen, W.-F.; Sasaki, K.; Ma, C.; Frenkel, A. I.; Marinkovic, N.; Helliwell, M. The Chemical Vapor Deposition of Nickel Phosphide or
Muckerman, J. T.; Zhu, Y.; Adzic, R. R. Hydrogen Evolution Catalysts Selenide Thin Films from a Single Precursor. J. Am. Chem. Soc. 2008,
Based on Non-Noble Metal Nickel−Molybdenum Nitride Nano- 130, 2420−2421.
sheets. Angew. Chem., Int. Ed. 2012, 51, 6131−6135. (60) Infantes-Molina, A.; Moreno-León, C.; Pawelec, B.; Fierro, J. L.
(40) Hinnemann, B.; Moses, P. G.; Bonde, J.; Jørgensen, K. P.; G.; Rodríguez-Castellón, E.; Jiménez-López, A. Simultaneous Hydro-
Nielsen, J. H.; Horch, S.; Chorkendorff, I.; Nørskov, J. K. Biomimetic desulfurization and Hydrodenitrogenation on MoP/Sio2 Catalysts:
Hydrogen Evolution: MoS2 Nanoparticles as Catalyst for Hydrogen Effect of Catalyst Preparation Method. Appl. Catal., B 2012, 113−114,
Evolution. J. Am. Chem. Soc. 2005, 127, 5308−5309. 87−99.
(41) Jaramillo, T. F.; Jørgensen, K. P.; Bonde, J.; Nielsen, J. H.; (61) Mobarok, M. H.; Luber, E. J.; Bernard, G. M.; Peng, L.;
Horch, S.; Chorkendorff, I. Identification of Active Edge Sites for Wasylishen, R. E.; Buriak, J. M. Phase-Pure Crystalline Zinc Phosphide
Electrochemical H2 Evolution from MoS2 Nanocatalysts. Science 2007, Nanoparticles: Synthetic Approaches And Characterization. Chem.
317, 100−102. Mater. 2014, 26, 1925−1935.
(42) Li, Y.; Wang, H.; Xie, L.; Liang, Y.; Hong, G.; Dai, H. Mos2 (62) Chiang, R.-K.; Chiang, R. T. Formation of Hollow Ni2P
Nanoparticles Grown on Graphene: An Advanced Catalyst for the Nanoparticles Based on the Nanoscale Kirkendall Effect. Inorg. Chem.
Hydrogen Evolution Reaction. J. Am. Chem. Soc. 2011, 133, 7296− 2007, 46, 369−371.
7299. (63) Wang, J.; Johnston-Peck, A. C.; Tracy, J. B. Nickel Phosphide
(43) Kibsgaard, J.; Chen, Z.; Reinecke, B. N.; Jaramillo, T. F. Nanoparticles with Hollow, Solid, and Amorphous Structures. Chem.
Engineering the Surface Structure of Mos2 to Preferentially Expose
Mater. 2009, 21, 4462−4467.
Active Edge Sites for Electrocatalysis. Nat. Mater. 2012, 11, 963−969. (64) Hitihami-Mudiyanselage, A.; Arachchige, M. P.; Seda, T.; Lawes,
(44) Morales-Guio, C. G.; Stern, L. A.; Hu, X. Nanostructured
G.; Brock, S. L. Synthesis and Characterization of Discrete FexNi2−XP
Hydrotreating Catalysts for Electrochemical Hydrogen Evolution.
Nanocrystals (0 < X < 2): Compositional Effects on Magnetic
Chem. Soc. Rev. 2014, 43, 6555−6569.
Properties. Chem. Mater. 2015, 27, 6592−6600.
(45) Liu, P.; Rodriguez, J. A. Catalysts for Hydrogen Evolution from
(65) Habas, S. E.; Baddour, F. G.; Ruddy, D. A.; Nash, C. P.; Wang,
the [NiFe] Hydrogenase to the Ni2P(001) Surface: The Importance of
J.; Pan, M.; Hensley, J. E.; Schaidle, J. A. A Facile Molecular Precursor
Ensemble Effect. J. Am. Chem. Soc. 2005, 127, 14871−14878.
(46) Oyama, S. T.; Wang, X.; Lee, Y. K.; Chun, W. J. Active Phase of Route to Metal Phosphide Nanoparticles and their Evaluation as
Ni2P/SiO2 in Hydroprocessing Reactions. J. Catal. 2004, 221, 263− Hydrodeoxygenation Catalysts. Chem. Mater. 2015, 27, 7580−7592.
273. (66) Zhang, H.; Ha, D.-H.; Hovden, R.; Kourkoutis, L. F.; Robinson,
(47) Senevirathne, K.; Burns, A. W.; Bussell, M. E.; Brock, S. L. R. D. Controlled Synthesis of Uniform Cobalt Phosphide Hyper-
Synthesis and Characterization of Discrete Nickel Phosphide Nano- branched Nanocrystals Using Tri-N-Octylphosphine Oxide as a
particles: Effect of Surface Ligation Chemistry on Catalytic Hydro- Phosphorus Source. Nano Lett. 2011, 11, 188−197.
desulfurization of Thiophene. Adv. Funct. Mater. 2007, 17, 3933−3939. (67) Stein, B. F.; Walmsley, R. H. Magnetic Susceptibility and
(48) Oyama, S. T. Novel Catalysts for Advanced Hydroprocessing: Nuclear Magnetic Resonance in Transition-Metal Monophosphides.
Transition Metal Phosphides. J. Catal. 2003, 216, 343−352. Phys. Rev. 1966, 148, 933−939.
(49) Oyama, S. T.; Gott, T.; Zhao, H.; Lee, Y.-K. Transition Metal (68) Arrigo, A. Preparation of Metal Phosphides. U.S. Patent
Phosphide Hydroprocessing Catalysts: A Review. Catal. Today 2009, 3,008,805, November 14, 1961.
143, 94. (69) Ripley, R. L. The Preparation and Properties Of Some
(50) Popczun, E. J.; McKone, J. R.; Read, C. G.; Biacchi, A. J.; Transition Phosphides. J. Less-Common Met. 1962, 4, 496−503.
Wiltrout, A. M.; Lewis, N. S.; Schaak, R. E. Nanostructured Nickel (70) Kaner, R.; Castro, C. A.; Gruska, R. P.; Wold, A. Preparation
Phosphide as an Electrocatalyst for the Hydrogen Evolution Reaction. and Characterization of the Platinum Metal Phosphides RuP2 and IrP2.
J. Am. Chem. Soc. 2013, 135, 9267−92270. Mater. Res. Bull. 1977, 12, 1143−1147.
(51) Blaugher, R. D.; Hulm, J. K.; Yocom, P. N. Superconducting (71) Braun, D. J.; Jeitschko, W. Ü ber Polyphosphide Von Chrom,
Phosphides of the Transition Metals. J. Phys. Chem. Solids 1965, 26, Mangan, Ruthenium und Osmium. Synthese und Kristallstruktur von
2037−2039. Rup4 und Osp4. Z. Anorg. Allg. Chem. 1978, 445, 157−166.
(52) Jeitschko, W.; Glaum, R.; Boonk, L. Superconducting LaRu2P2 (72) Jeitschko, W.; Rühl, R.; Krieger, U.; Heiden, C. Preparation,
and Other Alkaline Earth and Rare Earth Metal Ruthenium and Properties, and Structure Refinement of the Stacking Variant 2-MnP4.
Osmium Phosphides and Arsenides with Thcr2Si2 Structure. J. Solid Mater. Res. Bull. 1980, 15, 1755−1762.
State Chem. 1987, 69, 93−100. (73) Ruhl, R.; Jeitschko, W. Stacking Variants of Mnp4: Preparation
(53) Witte, J. H.; Wilhelm, H. A. Oxidation studies of some refractory and Structure of 6-Mnp4. Acta Crystallogr., Sect. B: Struct. Crystallogr.
metals; Ames Laboratory Technical Reports, Paper 66; Ames Cryst. Chem. 1981, 37, 39−44.
Laboratory: 1964. (74) Ruehl, R.; Jeitschko, W. Preparation and Crystal Structure of
(54) Motojima, S.; Wakamatsu, T.; Sugiyama, K. Corrosion Stability Dirhenium Pentaphosphide, Re2P5, a Diamagnetic Semiconducting
of Vapour-Deposited Transition Metal Phosphides at High Temper- Polyphosphide with Rhomboidal Re4 Clusters. Inorg. Chem. 1982, 21,
ature. J. Less-Common Met. 1981, 82, 379−383. 1886−1891.
(55) Kanatzidis, M. G.; Pöttgen, R.; Jeitschko, W. The Metal Flux: A (75) Yang, S.; Liang, C.; Prins, R. A Novel Approach to Synthesizing
Preparative Tool for the Exploration of Intermetallic Compounds. Highly Active Ni2P/SiO2 Hydrotreating Catalysts. J. Catal. 2006, 237,
Angew. Chem., Int. Ed. 2005, 44, 6996−7023. 118−130.

6040 DOI: 10.1021/acs.chemmater.6b02148


Chem. Mater. 2016, 28, 6017−6044
Chemistry of Materials Review

(76) Yao, Z. W. Exploration on Synthesis of Activated Carbon (97) Lewkebandara, T. S.; Proscia, J. W.; Winter, C. H. Precursor for
Supported Molybdenum Carbide, Nitride and Phosphide via the Low-Temperature Deposition of Titanium Phosphide Films.
Carbothermal Reduction Route. J. Alloys Compd. 2009, 475, L38−L41. Chem. Mater. 1995, 7, 1053−1054.
(77) Abu, I. I.; Smith, K. J. The Effect of Cobalt Addition to Bulk (98) Schrey, F.; Boone, T.; Nakahara, S.; Robbins, M.; Appelbaum, A.
MoP and Ni2P Catalysts for the Hydrodesulfurization of 4,6- Structure of Sputtered Ni2P Films. Thin Solid Films 1987, 149, 303−
Dimethyldibenzothiophene. J. Catal. 2006, 241, 356−366. 311.
(78) Yang, X.; Lu, A. Y.; Zhu, Y.; Hedhili, M. N.; Min, S.; Huang, K. (99) Yoon, J. S.; Doerr, H. J.; Deshpandey, C. V.; Bunshah, R. F.
W.; Han, Y.; Li, L. J. CoP Nanosheet Assembly Grown on Carbon Amorphous Nickel Phosphide Alloy Coatings Obtained by Magnetron
Cloth: A Highly Efficient Electrocatalyst for Hydrogen Generation. Sputtering Methods for Magnetic Recording Disk. J. Electrochem. Soc.
Nano Energy 2015, 15, 634−641. 1989, 136, 3513−3517.
(79) Shi, Y.; Xu, Y.; Zhuo, S.; Zhang, J.; Zhang, B. Ni2P Nanosheets/ (100) Iribarren, A.; Castro-Rodríguez, R.; Ponce-Cabrera, L.; Peña, J.
Ni Foam Composite Electrode for Long-Lived and pH-Tolerable L. Growth of Polycrystalline InP Thin Films by the Pulsed Laser
Electrochemical Hydrogen Generation. ACS Appl. Mater. Interfaces Deposition Technique. Thin Solid Films 2006, 510, 134−137.
2015, 7, 2376−2384. (101) Benck, J. D.; Hellstern, T. R.; Kibsgaard, J.; Chakthranont, P.;
(80) Yu, Y.; Song, Z.; Ge, H.; Wei, G. Preparation of CoP Films by Jaramillo, T. F. Catalyzing the Hydrogen Evolution Reaction (HER)
Ultrasonic Electroless Deposition at Low Initial Temperature. Prog. with Molybdenum Sulfide Nanomaterials. ACS Catal. 2014, 4, 3957−
Nat. Sci. 2014, 24, 232−238. 3971.
(81) Saadi, F. H.; Carim, A. I.; Verlage, E.; Hemminger, J. C.; Lewis, (102) McCrory, C. C. L.; Jung, S.; Ferrer, I. M.; Chatman, S. M.;
N. S.; Soriaga, M. P. CoP as an Acid-Stable Active Electrocatalyst for Peters, J. C.; Jaramillo, T. F. Benchmarking Hydrogen Evolving
the Hydrogen-Evolution Reaction: Electrochemical Synthesis, Inter- Reaction and Oxygen Evolving Reaction Electrocatalysts for Solar
facial Characterization and Performance Evaluation. J. Phys. Chem. C Water Splitting Devices. J. Am. Chem. Soc. 2015, 137, 4347−4357.
2014, 118, 29294−29300. (103) Ha, D. H.; Moreau, L. M.; Bealing, C. R.; Zhang, H.; Hennig,
(82) Burchardt, T. The Hydrogen Evolution Reaction on NiPx R. G.; Robinson, R. D. The Structural Evolution and Diffusion During
Alloys. Int. J. Hydrogen Energy 2000, 25, 627−634. the Chemical Transformation from Cobalt To Cobalt Phosphide
(83) Rudnik, E.; Kokoszka, K.; Łapsa, J. Comparative Studies on the Nanoparticles. J. Mater. Chem. 2011, 21, 11498−11510.
Electroless Deposition of Ni−P, Co−P and their Composites with SiC (104) Feng, L.; Vrubel, H.; Bensimon, M.; Hu, X. Easily-Prepared
Particles. Surf. Coat. Technol. 2008, 202, 2584−2590. Dinickel Phosphide (Ni2P) Nanoparticles as an Efficient and Robust
(84) Paseka, I.; Velicka, J. Hydrogen Evolution And Hydrogen Electrocatalyst for Hydrogen Evolution. Phys. Chem. Chem. Phys. 2014,
Sorption on Amorphous Smooth Me-P(X) (Me = Ni, Co And Fe-Ni) 16, 5917−5921.
Electrodes. Electrochim. Acta 1997, 42, 237−242. (105) Park, J.; Koo, B.; Yoon, K. Y.; Hwang, Y.; Kang, M.; Park, J. G.;
(85) Morikawa, T.; Nakade, T.; Yokoi, M.; Fukumoto, Y.; Iwakura, Hyeon, T. Generalized Synthesis of Metal Phosphide Nanorods via
C. Electrodeposition of Ni-P Alloys from Ni-Citrate Bath. Electrochim. Thermal Decomposition of Continuously Delivered Metal−Phosphine
Acta 1997, 42, 115−118. Complexes Using a Syringe Pump. J. Am. Chem. Soc. 2005, 127, 8433−
(86) Paseka, I. Evolution of Hydrogen and its Sorption on 8440.
Remarkable Active Amorphous Smooth Ni-P(X) Electrodes. Electro- (106) Moreau, L. M.; Ha, D.-H.; Zhang, H.; Hovden, R.; Muller, D.
chim. Acta 1995, 40, 1633−1640. A.; Robinson, R. D. Defining Crystalline/Amorphous Phases of
(87) Xie, Y.; Su, H. L.; Qian, X. F.; Liu, X. M.; Qian, Y. T. A Mild Nanoparticles Through X-ray Absorption Spectroscopy and X-Ray
One-Step Solvothermal Route to Metal Phosphides (Metal = Co, Ni, Diffraction: The Case of Nickel Phosphide. Chem. Mater. 2013, 25,
Cu). J. Solid State Chem. 2000, 149, 88−91. 2394−2403.
(88) Lukehart, C. M.; Milne, S. B.; Stock, S. R. Formation of (107) Pu, Z.; Liu, Q.; Tang, C.; Asiri, A. M.; Sun, X. Ni2P
Crystalline Nanoclusters of Fe2P, RuP, Co2P, Rh2P, Ni2P, Pd5P2, Or Nanoparticle Films Supported on a Ti Plate as an Efficient Hydrogen
PtP2 In a Silica Xerogel Matrix from Single-Source Molecular Evolution Cathode. Nanoscale 2014, 6, 11031−11034.
Precursors. Chem. Mater. 1998, 10, 903−908. (108) Pan, Y.; Hu, W.; Liu, D.; Liu, Y.; Liu, C. Carbon Nanotubes
(89) Henkes, A. E.; Vasquez, Y.; Schaak, R. E. Converting Metals Decorated with Nickel Phosphide Nanoparticles as Efficient Nano-
Into Phosphides: A General Strategy for the Synthesis of Metal hybrid Electrocatalysts for the Hydrogen Evolution Reaction. J. Mater.
Phosphide Nanocrystals. J. Am. Chem. Soc. 2007, 129, 1896−1897. Chem. A 2015, 3, 13087−13094.
(90) Perera, S. C.; Tsoi, G.; Wenger, L. E.; Brock, S. L. Synthesis of (109) Bai, Y.; Zhang, H.; Li, X.; Liu, L.; Xu, H.; Qiu, H.; Wang, Y.
MnP Nanocrystals by Treatment of Metal Carbonyl Complexes with Novel Peapod-Like Ni2P Nanoparticles with Improved Electro-
Phosphines: A New, Versatile Route To Nanoscale Transition Metal chemical Properties for Hydrogen Evolution and Lithium Storage.
Phosphides. J. Am. Chem. Soc. 2003, 125, 13960−13961. Nanoscale 2015, 7, 1446−1453.
(91) Liyanage, D. R.; Danforth, S. J.; Liu, Y.; Bussell, M. E.; Brock, S. (110) Han, A.; Jin, S.; Chen, H.; Ji, H.; Sun, Z.; Du, P. A Robust
L. Simultaneous Control of Composition, Size, and Morphology in Hydrogen Evolution Catalyst Based on Crystalline Nickel Phosphide
Discrete Ni2−XCoxP Nanoparticles. Chem. Mater. 2015, 27, 4349− Nanoflakes on Three-Dimensional Graphene/Nickel Foam: High
4357. Performance for Electrocatalytic Hydrogen Production from pH 0−14.
(92) Li, D.; Arachchige, M. P.; Kulikowski, B.; Lawes, G.; Seda, T.; J. Mater. Chem. A 2015, 3, 1941−1946.
Brock, S. L. Control of Composition and Size in Discrete CoxFe2‑XP (111) Laursen, A. B.; Patraju, K. R.; Whitaker, M. J.; Retuerto, M.;
Nanoparticles: Consequences for Magnetic Properties. Chem. Mater. Sarkar, T.; Yao, N.; Ramanujachary, K. V.; Greenblatt, M.; Dismukes,
2016, 28, 3920−3927. G. C. Nanocrystalline Ni5P4: A Hydrogen Evolution Electrocatalyst of
(93) Razeghi, M.; Poisson, M. A.; Larivain, J. P.; Duchemin, J. P. Low Exceptional Efficiency in Both Alkaline and Acidic Media. Energy
Pressure Metalorganic Chemical Vapor Deposition of InP and Related Environ. Sci. 2015, 8, 1027−1034.
Compounds. J. Electron. Mater. 1983, 12, 371−395. (112) Huang, Z.; Chen, Z.; Chen, Z.; Lv, C.; Meng, H.; Zhang, C.
(94) Biefeld, R. M. The Preparation of Device Quality Gallium Ni12P5 Nanoparticles as an Efficient Catalyst for Hydrogen Generation
Phosphide by Metal Organic Chemical Vapor Deposition. J. Cryst. via Electrolysis and Photoelectrolysis. ACS Nano 2014, 8, 8121−8126.
Growth 1982, 56, 382−388. (113) Jiang, P.; Liu, Q.; Sun, X. NiP2 Nanosheet Arrays Supported on
(95) Long, J. The Growth Of Zn3P2 by Metalorganic Chemical Vapor Carbon Cloth: An Efficient 3D Hydrogen Evolution Cathode in Both
Deposition. J. Electrochem. Soc. 1983, 130, 725−728. Acidic and Alkaline Solutions. Nanoscale 2014, 6, 13440−13445.
(96) Motojima, S.; Haguri, K.; Takahashi, Y.; Sugiyama, K. Chemical (114) Pan, Y.; Liu, Y.; Zhao, J.; Yang, K.; Liang, J.; Liu, D.; Hu, W.;
Vapor Deposition of Nickel Phosphide Ni2P. J. Less-Common Met. Liu, D.; Liu, Y.; Liu, C. Monodispersed Nickel Phosphide Nanocrystals
1979, 64, 101−106. with Different Phases: Synthesis, Characterization and Electrocatalytic

6041 DOI: 10.1021/acs.chemmater.6b02148


Chem. Mater. 2016, 28, 6017−6044
Chemistry of Materials Review

Properties for Hydrogen Evolution. J. Mater. Chem. A 2015, 3, 1656− (132) Li, Q.; Xing, Z.; Asiri, A. M.; Jiang, P.; Sun, X. Cobalt
1665. Phosphide Nanoparticles Film Growth on Carbon Cloth: A High-
(115) Ledendecker, M.; Krick Calderón, S.; Papp, C.; Steinrück, H. Performance Cathode for Electrochemical Hydrogen Evolution. Int. J.
P.; Antonietti, M.; Shalom, M. Shalom, M. The Synthesis of Hydrogen Energy 2014, 39, 16806−16811.
Nanostructured Ni5P4 Films and their Use as a Non-Noble (133) Liu, Q.; Tian, J.; Cui, W.; Jiang, P.; Cheng, N.; Asiri, A. M.;
Bifunctional Electrocatalyst for Full Water Splitting. Angew. Chem., Sun, X. Carbon Nanotubes Decorated with CoP Nanocrystals: A
Int. Ed. 2015, 54, 12361−12365. Highly Active Non-Noble-Metal Nanohybrid Electrocatalyst for
(116) Wang, C.; Ding, T.; Sun, Y.; Zhou, X.; Liu, Y.; Yang, Q. Ni12P5 Hydrogen Evolution. Angew. Chem., Int. Ed. 2014, 53, 6710−6714.
Nanoparticles Decorated on Carbon Nanotubes with Enhanced (134) Pu, Z.; Liu, Q.; Jiang, P.; Asiri, A. M.; Obaid, A. Y.; Sun, X.
Electrocatalytic and Lithium Storage Properties. Nanoscale 2015, 7, CoP Nanosheet Arrays Supported on a Ti Plate: An Efficient Cathode
19241−19242. for Electrochemical Hydrogen Evolution. Chem. Mater. 2014, 26,
(117) Kucernak, A. R. J.; Naranammalpuram Sundaram, V. N. Nickel 4326−4329.
Phosphide: The Effect of Phosphorus Content on Hydrogen (135) Callejas, J. F.; Read, C. G.; Popczun, E. J.; McEnaney, J. M.;
Evolution Activity and Corrosion Resistance in Acidic Medium. J. Schaak, R. E. Nanostructured Co2P Electrocatalyst for the Hydrogen
Mater. Chem. A 2014, 2, 17435−17445. Evolution Reaction and Direct Comparison with Morphologically
(118) Tian, T.; Ai, L.; Jiang, J. Metal-Organic Framework-Derived Equivalent CoP. Chem. Mater. 2015, 27, 3769−3774.
Nickel Phosphides as Efficient Electrocatalysts Toward Sustainable (136) Pan, Y.; Lin, Y.; Chen, Y.; Liu, Y.; Liu, C. Cobalt Phosphide-
Hydrogen Generation from Water Splitting. RSC Adv. 2015, 5, Based Electrocatalysts: Synthesis and Phase Catalytic Activity
10290−10295. Comparison for Hydrogen Evolution. J. Mater. Chem. A 2016, 4,
(119) Pan, Y.; Liu, Y.; Liu, C. Nanostructured Nickel Phosphide 4745−4754.
Supported on Carbon Nanospheres: Synthesis and Application as an (137) Popczun, E. J.; Roske, C. W.; Read, C. G.; Crompton, J. C.;
Efficient Electrocatalyst for Hydrogen Evolution. J. Power Sources McEnaney, J. M.; Callejas, J. F.; Lewis, N. S.; Schaak, R. E. Highly
2015, 285, 169−177. Branched Cobalt Phosphide Nanostructures for Hydrogen-Evolution
(120) Stern, L. A.; Feng, L.; Song, F.; Hu, X. Ni2P As a Janus Catalyst Electrocatalysis. J. Mater. Chem. A 2015, 3, 5420−5425.
for Water Splitting: The Oxygen Evolution Activity of Ni2P (138) Jiang, P.; Liu, Q.; Ge, C.; Cui, W.; Pu, Z.; Asiri, A. M.; Sun, X.
Nanoparticles. Energy Environ. Sci. 2015, 8, 2347−2351. CoP Nanostructures with Different Morphologies: Synthesis, Charac-
(121) Jiang, N.; You, B.; Sheng, M.; Sun, Y. Bifunctionality and terization and a Study of Their Electrocatalytic Performance Toward
Mechanism of Electrodeposited Nickel−Phosphorous Films for The Hydrogen Evolution Reaction. J. Mater. Chem. A 2014, 2, 14634−
Efficient Overall Water Splitting. ChemCatChem 2016, 8, 106−112.
14640.
(122) Wang, X.; Kolen’ko, Y. V.; Liu, L. Direct Solvothermal
(139) Huang, Z.; Chen, Z.; Chen, Z.; Lv, C.; Humphrey, M. G.;
Phosphorization of Nickel Foam to Fabricate Integrated Ni2P-
Zhang, C. Cobalt Phosphide Nanorods as an Efficient Electrocatalyst
Nanorods/Ni Electrodes for Efficient Electrocatalytic Hydrogen
for the Hydrogen Evolution Reaction. Nano Energy 2014, 9, 373−382.
Evolution. Chem. Commun. 2015, 51, 6738−6741.
(140) Jiang, N.; You, B.; Sheng, M.; Sun, Y. Electrodeposited Cobalt-
(123) Li, J.; Li, J.; Zhou, X.; Xia, Z.; Gao, W.; Ma, Y.; Qu, Y. Highly
Phosphorous-Derived Films as Competent Bifunctional Catalysts for
Efficient and Robust Nickel Phosphides as Bifunctional Electro-
Overall Water Splitting. Angew. Chem., Int. Ed. 2015, 54, 6251−6254.
catalysts for Overall Water-Splitting. ACS Appl. Mater. Interfaces 2016,
(141) Ma, L.; Shen, X.; Zhou, H.; Zhu, G.; Ji, Z.; Chen, K. CoP
8, 10826−10834.
Nanoparticles Deposited on Reduced Graphene Oxide Sheets as an
(124) Pan, Y.; Lin, Y.; Liu, Y.; Liu, C. Size-Dependent Magnetic and
Electrocatalytic Properties of Nickel Phosphide Nanoparticles. Appl. Active Electrocatalyst for the Hydrogen Evolution Reaction. J. Mater.
Surf. Sci. 2016, 366, 439−447. Chem. A 2015, 3, 5337−5343.
(125) Pan, Y.; Yang, N.; Chen, Y.; Lin, Y.; Li, Y.; Liu, Y.; Liu, C. (142) Ryu, J.; Jung, N.; Jang, J. H.; Kim, H. J.; Yoo, S. J. In Situ
Nickel Phosphide Nanoparticles-Nitrogen-Doped Graphene Hybrid as Transformation of Hydrogen-Evolving CoP Nanoparticles: Toward
an Efficient Catalyst for Enhanced Hydrogen Evolution Activity. J. Efficient Oxygen Evolution Catalysts Bearing Dispersed Morphologies
Power Sources 2015, 297, 45−52. with Co-Oxo/Hydroxo Molecular Units. ACS Catal. 2015, 5, 4066−
(126) Cai, Z. X.; Song, X. H.; Wang, Y. R.; Chen, X. Electro- 4074.
deposition-Assisted Synthesis of Ni2P Nanosheets on 3D Graphene/ (143) Yang, H.; Zhang, Y.; Hu, F.; Wang, Q. Urchin-Like CoP
Ni Foam Electrode and its Performance for Electrocatalytic Hydrogen Nanocrystals as Hydrogen Evolution Reaction and Oxygen Reduction
Production. ChemElectroChem 2015, 2, 1665−1671. Reaction Dual-Electrocatalyst with Superior Stability. Nano Lett. 2015,
(127) Bai, Y.; Zhang, H.; Fang, L.; Liu, L.; Qiu, H.; Wang, Y. Novel 15, 7616−7620.
Peapod Array of Ni2P@Graphitized Carbon Fiber Composites (144) Niu, Z.; Jiang, J.; Ai, L. Porous Cobalt Phosphide Nanorod
Growing on Ti Substrate: A Superior Material for Li-Ion Batteries Bundle Arrays as Hydrogen-Evolving Cathodes for Electrochemical
and the Hydrogen Evolution Reaction. J. Mater. Chem. A 2015, 3, Water Splitting. Electrochem. Commun. 2015, 56, 56−60.
5434−5441. (145) Wang, C.; Jiang, J.; Zhou, X.; Wang, W.; Zuo, J.; Yang, Q.
(128) Popczun, E. J.; Read, C. G.; Roske, C. W.; Lewis, N. S.; Schaak, Alternative Synthesis of Cobalt Monophosphide@C Core−Shell
R. E. Highly Active Electrocatalysis of the Hydrogen Evolution Nanocables for Electrochemical Hydrogen Production. J. Power
Reaction by Cobalt Phosphide Nanoparticles. Angew. Chem., Int. Ed. Sources 2015, 286, 464−469.
2014, 53, 5427−5430. (146) Jin, Z.; Li, P.; Xiao, D. Metallic Co2P Ultrathin Nanowires
(129) Tian, J.; Liu, Q.; Asiri, A. M.; Sun, X. Self-Supported Distinguished from CoP as Robust Electrocatalysts for Overall Water-
Nanoporous Cobalt Phosphide Nanowire Arrays: An Efficient 3D Splitting. Green Chem. 2016, 18, 1459−1464.
Hydrogen-Evolving Cathode over the Wide Range of pH 0−14. J. Am. (147) Wang, J.; Yang, W.; Liu, J. CoP2 Nanoparticles on Reduced
Chem. Soc. 2014, 136, 7587−7590. Graphene Oxide Sheets as a Super-Efficient Bifunctional Electro-
(130) Du, H.; Liu, Q.; Cheng, N.; Asiri, A. M.; Sun, X.; Li, C. M. catalyst for Full Water Splitting. J. Mater. Chem. A 2016, 4, 4686−
Template-Assisted Synthesis of CoP Nanotubes to Efficiently Catalyze 4690.
Hydrogen-Evolving Reaction. J. Mater. Chem. A 2014, 2, 14812− (148) Liu, M.; Li, J. Cobalt Phosphide Hollow Polyhedron as
14816. Efficient Bifunctional Electrocatalysts for the Evolution Reaction of
(131) Gu, S.; Du, H.; Asiri, A. M.; Sun, X.; Li, C. M. Three- Hydrogen and Oxygen. ACS Appl. Mater. Interfaces 2016, 8, 2158−
Dimensional Interconnected Network of Nanoporous CoP Nanowires 2165.
as an Efficient Hydrogen Evolution Cathode. Phys. Chem. Chem. Phys. (149) Kibsgaard, J.; Tsai, C.; Chan, K.; Benck, J. D.; Norskov, J. K.;
2014, 16, 16909−16913. Abild-Pedersen, F.; Jaramillo, T. F. Designing an Improved Transition

6042 DOI: 10.1021/acs.chemmater.6b02148


Chem. Mater. 2016, 28, 6017−6044
Chemistry of Materials Review

Metal Phosphide Catalyst for Hydrogen Evolution Using Experimental Hydrogen Evolution Using Amorphous Tungsten Phosphide Nano-
and Theoretical Trends. Energy Environ. Sci. 2015, 8, 3022−3029. particles. Chem. Commun. 2014, 50, 11026−11028.
(150) Lu, A.; Chen, Y.; Li, H.; Dowd, A.; Cortie, M. B.; Xie, Q.; Guo, (168) Pu, Z.; Liu, Q.; Asiri, A. M.; Sun, X. Tungsten Phosphide
H.; Qi, Q.; Peng, D. L. Magnetic Metal Phosphide Nanorods as Nanorod Arrays Directly Grown on Carbon Cloth: A Highly Efficient
Effective Hydrogen-Evolution Electrocatalysts. Int. J. Hydrogen Energy and Stable Hydrogen Evolution Cathode at All pH Values. ACS Appl.
2014, 39, 18919−18928. Mater. Interfaces 2014, 6, 21874−21879.
(151) Wang, X.; Clark, P.; Oyama, S. T. Synthesis, Characterization, (169) Xing, Z.; Liu, Q.; Asiri, A. M.; Sun, X. High-Efficiency
and Hydrotreating Activity of Several Iron Group Transition Metal Electrochemical Hydrogen Evolution Catalyzed by Tungsten
Phosphides. J. Catal. 2002, 208, 321−331. Phosphide Submicroparticles. ACS Catal. 2015, 5, 145−149.
(152) Callejas, J. F.; McEnaney, J. M.; Read, C. G.; Crompton, J. C.; (170) Xing, Z.; Liu, Q.; Asiri, A. M.; Sun, X. Closely Interconnected
Biacchi, A. J.; Popczun, E. J.; Gordon, T. R.; Lewis, N. S.; Schaak, R. E. Network of Molybdenum Phosphide Nanoparticles: A Highly Efficient
Electrocatalytic and Photocatalytic Hydrogen Production from Acidic Electrocatalyst for Generating Hydrogen from Water. Adv. Mater.
and Neutral-pH Aqueous Solutions Using Iron Phosphide Nano- 2014, 26, 5702−5707.
particles. ACS Nano 2014, 8, 11101−11107. (171) Chen, X.; Wang, D.; Wang, Z.; Zhou, P.; Wu, Z.; Jiang, F.
(153) Jiang, P.; Liu, Q.; Liang, Y.; Tian, J.; Asiri, A. M.; Sun, X. A Molybdenum Phosphide: A New Highly Efficient Catalyst for the
Cost-Effective 3D Hydrogen Evolution Cathode with High Catalytic Electrochemical Hydrogen Evolution Reaction. Chem. Commun. 2014,
Activity: FeP Nanowire Array as the Active Phase. Angew. Chem. 2014, 50, 11683−11685.
126, 13069−13073. (172) Cui, W.; Liu, Q.; Xing, Z.; Asiri, A. M.; Alamry, K. A.; Sun, X.
(154) Jiang, J.; Wang, C.; Zhang, J.; Wang, W.; Zhou, X.; Pan, B.; MoP Nanosheets Supported on Biomass-Derived Carbon Flake: One-
Tang, K.; Zuo, J.; Yang, Q. Synthesis of FeP2/C Nanohybrids and their Step Facile Preparation and Application as a Novel High-Active
Performance for Hydrogen Evolution Reaction. J. Mater. Chem. A Electrocatalyst Toward Hydrogen Evolution Reaction. Appl. Catal., B
2015, 3, 499−503. 2015, 164, 144−150.
(155) Zhang, Y.; Zhang, H.; Feng, Y.; Liu, L.; Wang, Y. Unique Fe2P (173) Wang, T.; Du, K.; Liu, W.; Zhu, Z.; Shao, Y.; Li, M. Enhanced
Nanoparticles Enveloped in Sandwichlike Graphited Carbon Sheets as Electrocatalytic Activity of MoP Microparticles for Hydrogen
Excellent Hydrogen Evolution Reaction Catalyst and Lithium-Ion Evolution by Grinding and Electrochemical Activation. J. Mater.
Battery Anode. ACS Appl. Mater. Interfaces 2015, 7, 26684−26690. Chem. A 2015, 3, 4368−4373.
(156) Yang, X.; Lu, A. Y.; Zhu, Y.; Min, S.; Hedhili, M. N.; Han, Y.; (174) Chen, Z.; Lv, C.; Chen, Z.; Jin, L.; Wang, J.; Huang, Z.
Huang, K. W.; Li, L. J. Rugae-Like FeP Nanocrystal Assembly on a Molybdenum Phosphide Flakes Catalyze Hydrogen Generation in
Carbon Cloth: An Exceptionally Efficient and Stable Cathode for Acidic and Basic Solutions. Am. J. Anal. Chem. 2014, 5, 1200−1213.
Hydrogen Evolution. Nanoscale 2015, 7, 10974−10981. (175) Pu, Z.; Saana Amiinu, I.; Wang, M.; Yang, Y.; Mu, S.
(157) Zhu, X.; Liu, M.; Liu, Y.; Chen, R.; Nie, Z.; Li, J.; Yao, S. Semimetallic MoP2: An Active and Stable Hydrogen Evolution
Carbon-Coated Hollow Mesoporous FeP Microcubes: An Efficient Electrocatalyst Over the Whole pH Range. Nanoscale 2016, 8,
and Stable Electrocatalyst for Hydrogen Evolution. J. Mater. Chem. A 8500−8504.
2016, 4, 8974−8977. (176) Du, H.; Gu, S.; Liu, R.; Li, C. M. Tungsten Diphosphide
(158) Son, C. Y.; Kwak, I. H.; Lim, Y. R.; Park, J. FeP And FeP2 Nanorods as an Efficient Catalyst for Electrochemical Hydrogen
Nanowires for Efficient Electrocatalytic Hydrogen Evolution Reaction. Evolution. J. Power Sources 2015, 278, 540−545.
Chem. Commun. 2016, 52, 2819−2822. (177) Tian, J.; Liu, Q.; Cheng, N.; Asiri, A. M.; Sun, X. Self-
(159) Zhang, Z.; Lu, B.; Hao, J.; Yang, W.; Tang, J. FeP Supported Cu3P Nanowire Arrays as an Integrated High-Performance
Nanoparticles Grown on Graphene Sheets as Highly Active Non- Three-Dimensional Cathode for Generating Hydrogen from Water.
Precious-Metal Electrocatalysts for Hydrogen Evolution Reaction. Angew. Chem., Int. Ed. 2014, 53, 9577−9581.
Chem. Commun. 2014, 50, 11554−11557. (178) Ma, L.; Shen, X.; Zhou, H.; Zhu, J.; Xi, C.; Ji, Z.; Kong, L.
(160) Liang, Y.; Liu, Q.; Asiri, A. M.; Sun, X.; Luo, Y. Self-Supported Synthesis of Cu3P Nanocubes and their Excellent Electrocatalytic
FeP Nanorod Arrays: A Cost-Effective 3D Hydrogen Evolution Efficiency for the Hydrogen Evolution Reaction in Acidic Solution.
Cathode with High Catalytic Activity. ACS Catal. 2014, 4, 4065−4069. RSC Adv. 2016, 6, 9672−9677.
(161) Liu, R.; Gu, S.; Du, H.; Li, C. M. Controlled Synthesis of FeP (179) Feng, Y.; Ouyang, Y.; Peng, L.; Qiu, H.; Wang, H.; Wang, Y.
Nanorod Arrays as Highly Efficient Hydrogen Evolution Cathode. J. Quasi-Graphene-Envelope Fe-Doped Ni2P Sandwiched Nanocompo-
Mater. Chem. A 2014, 2, 17263−17267. sites for Enhanced Water Splitting and Lithium Storage Performance.
(162) Xu, Y.; Wu, R.; Zhang, J.; Shi, Y.; Zhang, B. Anion-Exchange J. Mater. Chem. A 2015, 3, 9587−9594.
Synthesis of Nanoporous FeP Nanosheets as Electrocatalysts for (180) Kibsgaard, J.; Jaramillo, T. F. Molybdenum Phosphosulfide: An
Hydrogen Evolution Reaction. Chem. Commun. 2013, 49, 6656−6658. Active, Acid-Stable, Earth-Abundant Catalyst for the Hydrogen
(163) Tian, J.; Liu, Q.; Liang, Y.; Xing, Z.; Asiri, A. M.; Sun, X. FeP Evolution Reaction. Angew. Chem., Int. Ed. 2014, 53, 14433−14437.
Nanoparticles Film Grown on Carbon Cloth: An Ultrahighly Active (181) Caban-Acevedo, M.; Stone, M. L.; Schmidt, J. R.; Thomas, J.
3D Hydrogen Evolution Cathode in Both Acidic and Neutral G.; Ding, Q.; Chang, H. C.; Tsai, M. L.; He, J. H.; Jin, S. Efficient
Solutions. ACS Appl. Mater. Interfaces 2014, 6, 20579−20584. Hydrogen Evolution Catalysis Using Ternary Pyrite-Type Cobalt
(164) Zhang, Z.; Hao, J.; Yang, W.; Lu, B.; Tang, J. Modifying Candle Phosphosulphide. Nat. Mater. 2015, 14, 1245−1251.
Soot with FeP Nanoparticles into High-Performance and Cost- (182) Hansen, M. H.; Stern, L. A.; Feng, L.; Rossmeisl, J.; Hu, H.
Effective Catalysts for the Electrocatalytic Hydrogen Evolution Widely Available Active Sites on Ni2P for Electrochemical Hydrogen
Reaction. Nanoscale 2015, 7, 4400−4405. Evolution − Insights from First Principles Calculations. Phys. Chem.
(165) McEnaney, J. M.; Crompton, J. C.; Callejas, J. F.; Popczun, E. Chem. Phys. 2015, 17, 10823−10829.
J.; Biacchi, A. J.; Lewis, N. S.; Schaak, R. E. Amorphous Molybdenum (183) Suzuki, S.; Moula, G. M.; Miyamoto, T.; Nakagawa, Y.;
Phosphide Nanoparticles for Electrocatalytic Hydrogen Evolution. Kinosthita, K.; Asakura, K.; Oyama, S. T.; Otani, S. Scanning
Chem. Mater. 2014, 26, 4826−4831. Tunneling Microscopy and Photoemission Electron Microscopy
(166) Xiao, P.; Sk, M. A.; Thia, L.; Ge, X.; Lim, R. J.; Wang, J. Y.; Studies on Single Crystal Ni2P Surfaces. J. Nanosci. Nanotechnol.
Lim, K. H.; Wang, X. Molybdenum Phosphide as an Efficient 2009, 9, 195−201.
Electrocatalyst for the Hydrogen Evolution Reaction. Energy Environ. (184) Hernandez, A. B.; Ariga, H.; Takakusagi, S.; Kinoshita, K.;
Sci. 2014, 7, 2624−2629. Suzuki, S.; Otani, S.; Oyama, S. T.; Asakura, K. Dynamical LEED
(167) McEnaney, J. M.; Chance Crompton, J.; Callejas, J. F.; Analysis of Ni2P (0 0 0 1)-1 × 1: Evidence for P-Covered Surface
Popczun, E. J.; Read, C. G.; Lewis, N. S.; Schaak, R. E. Electrocatalytic Structure. Chem. Phys. Lett. 2011, 513, 48−52.

6043 DOI: 10.1021/acs.chemmater.6b02148


Chem. Mater. 2016, 28, 6017−6044
Chemistry of Materials Review

(185) Pinaud, B. A.; Benck, J. D.; Seitz, L. C.; Forman, A. J.; Chen,
Z.; Deutsch, T. G.; James, B. D.; Baum, K. N.; Baum, G. N.; Ardo, S.;
Wang, H.; Miller, E.; Jaramillo, T. F. Technical and Economic
Feasibility of Centralized Facilities for Solar Hydrogen Production via
Photocatalysis and Photoelectrochemistry. Energy Environ. Sci. 2013, 6,
1983−2002.
(186) Licht, S.; Wang, B.; Mukerji, S.; Soga, T.; Umeno, M.;
Tributsch, H. Efficient Solar Water Splitting, Exemplified by RuO2-
Catalyzed AlGaAs/Si Photoelectrolysis. J. Phys. Chem. B 2000, 104,
8920−8924.
(187) Coridan, R. H.; Nielander, A. C.; Francis, S. A.; Mcdowell, M.
T.; Dix, V.; Chatman, S. M.; Lewis, N. Methods for Comparing the
Performance of Energy-Conversion Systems for Use in Solar Fuels and
Solar Electricity Generation. Energy Environ. Sci. 2015, 8, 2886−2901.
(188) Esposito, D. V.; Hunt, S. T.; Stottlemyer, A. L.; Dobson, K. D.;
Mccandless, B. E.; Birkmire, R. W.; Chen, J. G. Low-Cost Hydrogen-
Evolution Catalysts Based on Monolayer Platinum on Tungsten
Monocarbide Substrates. Angew. Chem. 2010, 122, 10055−10058.
(189) Stephens, I. E. L.; Chorkendorff, I. Minimizing the Use of
Platinum in Hydrogen-Evolving Electrodes. Angew. Chem., Int. Ed.
2011, 50, 1476−1477.
(190) Bard, A. J.; Fox, M. A. Artificial Photosynthesis: Solar Splitting
of Water to Hydrogen and Oxygen. Acc. Chem. Res. 1995, 28, 141−
141.
(191) Bae, D.; Pedersen, T.; Seger, B.; Malizia, M.; Kuznetsov, A.;
Hansen, O.; Chorkendorff, I.; Vesborg, P. C. K. Back-Illuminated Si
Photocathode: A Combined Experimental and Theoretical Study for
Photocatalytic Hydrogen Evolution. Energy Environ. Sci. 2015, 8, 650−
660.
(192) Roske, C. W.; Popczun, E. J.; Seger, B.; Read, C. G.; Pedersen,
T.; Hansen, O.; Vesborg, P. C. K.; Brunschwig, B. S.; Schaak, R. E.;
Chorkendorff, I.; Gray, H. B.; Lewis, N. S. Comparison of the
Performance of CoP-Coated and Pt-Coated Radial Junction
N+P−Silicon Microwire-Array Photocathodes for the Sunlight-Driven
Reduction of Water To H2(G). J. Phys. Chem. Lett. 2015, 6, 1679−1683.
(193) Basu, M.; Zhang, Z.-W.; Chen, C. J.; Chen, P. T.; Yang, K. C.;
Ma, C. G.; Lin, C. C.; Hu, S. F.; Liu, R. S. Heterostructure of Si and
CoSe2: A Promising Photocathode Based on a Non-Noble Metal
Catalyst for Photoelectrochemical Hydrogen Evolution. Angew. Chem.,
Int. Ed. 2015, 54, 6211−6216.
(194) Bao, X.-Q.; Fatima Cerqueira, M.; Alpuim, P.; Liu, L. Silicon
Nanowire Arrays Coupled with Cobalt Phosphide Spheres as Low-
Cost Photocathodes for Efficient Solar Hydrogen Evolution. Chem.
Commun. 2015, 51, 10742−10745.
(195) Lv, C.; Chen, Z.; Chen, Z.; Zhang, B.; Qin, Y.; Huang, Z.;
Zhang, C. Silicon Nanowires Loaded with Iron Phosphide for Effective
Solar-Driven Hydrogen Production. J. Mater. Chem. A 2015, 3,
17669−17675.
(196) Hellstern, R. T.; Benck, J. D.; Kibsgaard, J.; Jaramillo, J. F.;
Hahn, C. Engineering Cobalt Phosphide (CoP) Thin Film Catalysts
for Enhanced Hydrogen Evolution Activity on Silicon Photocathodes.
Adv. Energy Mater. 2016, 6, 1501758.
(197) James, B. D.; Baum, G. N.; Perez, J.; Baum, K. N.
Technoeconomic Analysis of Photoelectrochemical (PEC) Hydrogen
Production; U.S. Department of Energy: 2009; Final Report.
(198) Cao, S.; Chen, Y.; Wang, C. J.; Lv, X. J.; Fu, W. F. Spectacular
Photocatalytic Hydrogen Evolution Using Metal-Phosphide/CdS
Hybrid Catalysts Under Sunlight Irradiation. Chem. Commun. 2015,
51, 8708−8711.
(199) Sun, K.; Chen, Y.; Verlage, E.; Lewis, N. S.; Xiang, C.; Liu, R. A
Stabilized, Intrinsically Safe, 10% Efficient, Solar-Driven Water-
Splitting Cell Incorporating Earth-Abundant Electrocatalysts with
Steady-State pH Gradients and Product Separation Enabled by a
Bipolar Membrane. Adv. Energy Mater. 2016, 6, 1600379.
(200) Luo, J.; Im, J. H.; Mayer, M. T.; Schreier, M.; Nazeeruddin, M.
K.; Park, N. G.; Fan, H. J.; Grätzel, M.; Tilley, S. D. Water Photolysis
at 12.3% Efficiency via Perovskite Photovoltaics and Earth-Abundant
Catalysts. Science 2014, 345, 1593−1596.

6044 DOI: 10.1021/acs.chemmater.6b02148


Chem. Mater. 2016, 28, 6017−6044

You might also like