Sunteți pe pagina 1din 38

Cobalt Phosphosulphide:

An Advanced Bifunctional Material for Superior


Battery Anodes and Electrocatalytic Oxygen Evolution

<Chia Ching Keat>


<U1422077G>

School of Materials Science & Engineering

A final year project submitted to the Nanyang Technological University in


fulfilment of the requirement for the degree of Bachelor of Engineering

1
Abstract

Lithium-ion batteries (LIBs) and hydrogen fuel are considered to be two of the most effective
energy storage systems in recent times. Many modern portable devices rely almost exclusively
on the former while hydrogen is the cleanest and most energy dense fuel known to man.
However, heavy reliance of such batteries is fast depleting the Earth’s lithium reserves and
clean hydrogen fuel production is currently too expensive to be commercially viable. Hence,
the development of materials that serve to improve the efficiency of lithium-ion batteries (so
as to reduce lithium demand) and reduce the cost of clean hydrogen fuel production is of great
importance.

This report reviews the use of LIBs in modern society and the feasibility of hydrogen
production via the water-splitting process. The mechanisms and recent developments for both
energy storage systems were explored. Subsequently, a new bi-functional anode material was
developed for both the afroementioned technologies.

Homogenous yolk – shell structured spheres of cobalt phosphosulfide (CoPS) were


synthesized via a novel partial-sulfurization/phosphorization hydrothermal synthesis method.
Cobalt precursor was first prepared, followed by the separate half-sulfurization and
phosphorization steps. As an anode for lithium-ion batteries, the CoPS material showed a high
specific energy capacity of 870 mAh/g at the current rate at 0.2 A/g and was highly stable with
a capacity of 475 mAh/g at 4 A/g for 5000 cycles. CoPS as an electro-catalyst also displayed
good performance and stability; the overpotential required for oxygen evolution reaction
(OER) was at a respectable 270 mV at a current density of 10 mA/cm2 and catalytic activity
was stable at 10 mA/cm2 for 15 hours.

The superior LIB performance and stability stemmed from the Co-P-S synergy and the
yolk – shell structure, respectively. Meanwhile, phosphorous doping in CoPS for water splitting
led to a greatly modulated electronic structure, adsorption Gibb’s free energy of OER
intermediates and overall electrical conductivity.

The CoPS spheres were synthesized with diameters ranging from 400 nm to 552 nm and
yet the LIB and OER performance are already comparable with the latest state-of-the-art
nanomaterials. Hence, there is potential for CoPS performance to be improved if the spheres
can be miniaturized or modified further.

2
Acknowledgment

The author would like to express his utmost gratitude and appreciation for the following people
for their guidance and support in the completion of this Final Year Project. He has learnt an
invaluable set of technical and transferable soft skills that will be applicable to his working life
upon graduation.

The author would like to thank his supervisor, Associate Professor Alex Yan Qingyu,
for offering him the opportunity to pursue his interest in the latest of battery and hydrogen
production technology. It was through his feedback that the author gained an insight to the
latest in emerging energy storage technologies. Secondly, the author would like to express his
deepest gratitude to his mentor, Dr Dai Zhengfei for his constant guidance, advice and input
for the completion of this study. He has imparted a positive mindset towards research in the
author.

3
Contents
Abstract ................................................................................................................................ 2

Acknowledgment .................................................................................................................. 3

List of Figures ....................................................................................................................... 5

1. Introduction....................................................................................................................... 6

1.1 Background ................................................................................................................. 6


1.2 Objectives ................................................................................................................... 7
1.3 Scope .......................................................................................................................... 7
2. Literature Review .............................................................................................................. 8

2.1 Lithium Ion Batteries ................................................................................................... 8


2.2 Electro-Catalytic Water Splitting ............................................................................... 10
2.3 General Electro-Catalyst Selection Criteria ................................................................ 12
2.3.1 HER Catalyst Selection ....................................................................................... 13
2.3.2 OER Catalyst Selection ....................................................................................... 14
2.4 Transition Metal Phosphides ...................................................................................... 16
2.5 Transition Metal Dichalcogenides.............................................................................. 17
2.6 Transition Metal Phosphosulfides .............................................................................. 18
3. Materials and Methods .................................................................................................... 19

3.1 Synthesis ................................................................................................................... 19


3.2 Characterization ........................................................................................................ 20
3.3 Electrochemical Measurement ................................................................................... 20
4. Results and Discussion .................................................................................................... 21

4.1 Characterization ........................................................................................................ 21


4.2 Performance in Lithium Ion Batteries ........................................................................ 24
4.3 Performance in Oxygen Evolution Reaction .............................................................. 27
5. Recommendations for Further Research .......................................................................... 31

6. Conclusion ...................................................................................................................... 33

References .......................................................................................................................... 34

4
List of Figures

Figure 1: Charging/Discharging of Li-ion batteries.6 ............................................................. 8


Figure 2: Projected demand for lithium-ion battery packs.10 .................................................. 9
Figure 3: Rising cost of lithium.11........................................................................................ 10
Figure 4: A simple water splitting device setup. 20 ................................................................ 10
Figure 5: Relationship between exchange current density and hydrogen absorption energy for
HER catalysts. 22 ................................................................................................................. 13
Figure 6: Proposed reaction mechanisms for OER.26 ........................................................... 14
Figure 7: Schematic representation of the scaling relations for ΔGMOH, ΔGMO and
ΔGMOOH as functions of ΔGMOH.18 ................................................................................ 15
Figure 8: Comparison OER overpotential trends for rutile, anatase, Co3O4 and MnxOy
oxides using the universal descriptor ΔG2.18 ........................................................................ 15
Figure 9: Capacity and cycle life of hollow and solid CoP nanoparticles. 34 .......................... 17
Figure 10: From the left: Trigonal prismatic (2H) and octahedral (1T) configurations of
MoS2, respectively42 ............................................................................................................ 18
Figure 11: Pyrite structure of CoPS. 43.................................................................................. 18
Figure 12: Schematic illustration of synthesizing CoPS yolk-shell nanoparticles ................. 19
Figure 13: Morphology and chemical analysis of CoPS samples. (a) to (c) SEM images of
uniform CoPS spheres (d) XRD spectra of CoPS vs CoS ..................................................... 21
Figure 14: Reaction steps to CoPS....................................................................................... 22
Figure 15: TEM images of (a) CoPS spheres with defects (b)/(c) crystalites present in CoPS
spheres. ............................................................................................................................... 22
Figure 16: (a) HAADF-STEM and (b) EDS-STEM images of CoPS spheres. ...................... 23
Figure 17: XPS spectra of CoPS and CoS. ........................................................................... 23
Figure 18: Electrochemical performance of CoPS in LIB/OER. (a) 1st and 2nd
charge/discharge at current density of 0.2 A/g (b) Stability performance at different current
densities from 0.2 to 8.0 A/g of CoPS, CoP and CoS. .......................................................... 24
Figure 19: Stability and specific capacity of CoPS anodes at a current density of (a) 0.2 A/g
(b) 1A/g and (c) 4 A/g ......................................................................................................... 26
Figure 20: Current vs. Applied Potential (with reference to AgCl electrode) of CoPS anodes
in three-electrode system. .................................................................................................... 27
Figure 21: Graph of Current Density vs. Potential of CoPS anodes in three-electrode system
........................................................................................................................................... 28
Figure 22: Graph of Overpotential vs. Log (Current)/OER activity in three-electrode system
........................................................................................................................................... 28
Figure 23: OER activity level of CoPS anode at a constant E Ag/AgCl of 0.5 V ....................... 30

5
1. Introduction

1.1 Background

The worldwide awareness of the benefits of a low-carbon economy empowers the ever-
growing demands for replacing fossil fuels with clean and sustainable energy sources, such as
solar, wind and tidal energy. Unfortunately, the reliance of such intermittent energy sources on
various environmental factors – most notably time, location and weather – limits their
implementation. In order to fully realize the potential of such renewable energy sources,
effective energy storage systems are needed so that the intermittent energy harvested from such
sources can be stored efficiently.1-3
Electrochemically rechargeable batteries and hydrogen fuel are among the most successful
energy storage systems over the past decade, of which lithium ion batteries (LIBs) and the
precious metal-catalyzed water – splitting process (which produces molecular hydrogen fuel)
are popular1-8. However, the scarcity and rising costs of lithium and precious metals (usually
platinum) hamper their sustainable implementation.9 Therefore, it is an intensively pursued
goal for one to access alternative materials that are cheap, abundant in supply and possess
desirable electrochemical properties for LIBs and water – splitting.
Transition metal phosphides (TMPs) are one of the promising materials considered for
replacing precious metal catalysts due to their high theoretical capacities. One acknowledged
drawback that limits their practical utilization is the structural instability during the
electrochemical process.20-21 On the other hand, transition metal dichalcogenides (TMDs)
possess the desirable electrochemical and mechanical stability, but have lower theoretical
capacities and catalytic activity. Recent studies have shown that combining the aforementioned
materials can produce synergistic effects: a MoS2-MoP hybrid system exhibits both excellent
stability and higher catalytic activity than its individual components.22-23 However, the
disordered interfaces distributed in the material architecture may impose an obstacle in the
precise understanding of the synergetic effects and design of advanced electrochemical
materials. A miscible M-P-S phase, or metal phosphosulphides, will hold the intriguing
promise in this regard where it might allow for precise control of the catalytic and other
physiochemical properties.23
Note that research on CoPS systems has so far been focused on HER performance; few
reports have included research in the counterpart OER in water splitting or in other
electrochemical applications, such as LIBs. Henceforth, the goal of this study is to improve
current Li-ion batteries and OER technologies by using cobalt phosphosulphides as electrodes.

6
1.2 Objectives

- To synthesize yolk – shell structured cobalt phosphosulfide spheres

- To characterize the synthesized spheres

- To evaluate the performance of the synthesized material as anodes in lithium-ion


batteries and in oxygen evolution reactions

1.3 Scope

Herein, we report a novel partial-sulfurization/phosphorization strategy to synthesize a


multifunctional yolk-shell structured cobalt phosphosulfide material for superior battery
anodes and efficient electrocatalysis. This strategy involves yolk-shell spheres with a diameter
of 400 nm and a shell thickness of 20 nm, as well as homogenous distribution of Co, P and S
throughout the entire sphere structure. By regulating the degree of phosphorization, the P/S
ratio in the cobalt phosphosulfide can be controlled and in turn, the electrochemical properties
can be optimized.

In addition to synthesis, the characterization and performance of cobalt phosphosulfides


(CoPS) as anodes in lithium-ion batteries and oxygen evolution reactions were evaluated in
this report.

7
2. Literature Review

2.1 Lithium Ion Batteries

First proposed in the 1970s, lithium-ion batteries (LIBs), or Li-ion batteries, are now one of the
most widely used energy storage systems. Since the 1990s, LIBs have trumped over the
previously preferred nickel-cadmium and nickel-metal-hydride systems as the most popular
portable secondary battery. This is due to the higher energy density, higher operating voltage,
less toxic components, lower maintenance requirements, limited self-discharging and
negligible memory effect of LIBs.2-3 LIBs are mainly used in portable consumer electronics
ranging from power systems and laptops to the ubiquitous smartphones. Recent developments
in transportation vehicles, most notably by the American company Tesla, include the push for
LIBs to replace the fuel engines of cars.4

Generally, several electrochemical cells make up a LIB. Each individual cell in turn consists
of an anode, a cathode and electrolyte. As opposed to other secondary batteries, lithium-ion
batteries store electrical energy in the two electrodes as Li-ion intercalated compounds. During
charging or discharging, the insertion (intercalation) or extraction (deintercalation) of Li+ ions
occurs without destroying the core electrode structure5.

Figure 1: Charging/Discharging of Li-ion batteries.6

As shown on Figure 1, the Li+ ions move from cathode to anode during charging, and vice
versa during discharging. LIBs typically use LiCoO2 cathodes and carbon anodes.7 The
chemical reactions are as follows:

8
Cathode: LiCoO2 ⇌ Li1-xCoO2 + xLi+ + xe-
Anode: C + xLi+ + xe- ⇌ LixC
Overall: LiCoO2 + C ⇌ LixC + Li1-xCoO2 (E0cell = 3.6V)

With different types of LIBs, the performance, cost and safety characteristics can vary. For
example, typical LIBs based on LiCoO2 offer high energy density but are less safe when
damaged or overcharged while lithium-ion manganese oxide batteries offer lower energy
densities but possess less safety risks and better thermal stability.8

Scarcity and Rising Costs of Lithium

Despite the large reserves of lithium available today (the U.S. Geological Survey concluded
that the world has enough reserves for another 365 years of current global production), the
demand for lithium is ever increasing due to our reliance on lithium-ion batteries and other
lithium-related industrial applications.9

With Tesla expecting to produce 500,000 LIB-dependant electric vehicles per year by 2020,
the additonal demand (Figure 2) may cause current lithium reserves to last a mere 17 years
instead of the projected 365 years.9 Furthermore, the cost of lithium has been rising steadily,
as shown on Figure 3. Hence, the Earth’s supply of lithium alone cannot shoulder the needs of
a technologically advanced world that heavily relies on portable and high-density energy
storage. It is simply not a sustainable option. Research on alternative energy storage methods
with similar or better energy densities, cost, reliability and efficiency is needed to reduce the
demand on the world’s lithium reserves and the cost of LIBs. Examples of alternative energy
storage methods are batteries that use non-precious metals electrodes and hydrogen fuel.

Figure 2: Projected demand for lithium-ion battery packs.10

9
Price of Lithium vs Years
8000

Lithium Price (USD/metric ton)


7000
6000
5000
4000
3000
2000
1000
0
2000 2002 2004 2006 2008 2010 2012 2014 2016 2018
Year

Figure 3: Rising cost of lithium.11

2.2 Electro-Catalytic Water Splitting


Molecular hydrogen is known as the most efficient and sustainable energy carrier; it is the most
energy dense of all fuels and it produces zero carbon emissions. Additionally, hydrogen fuel
theoretically produces only water as by-products. Not only is hydrogen a clean and sustainable
fuel, hydrogen is also ideal for storing energy from intermittent renewable energy sources. This
can be done by coupling large-scale renewable energy harvesting technologies with
electrochemical water splitting devices, allowing renewable energy to be converted into
hydrogen fuel for later use.17-19 The figure below shows sunlight being harnessed by a water
splitting setup to produce hydrogen.

Figure 4: A simple water splitting device setup. 20

The problem with hydrogen fuel is that current large-scale production of the substance is
unsustainable; 96% of all hydrogen produced comes from steam reformation of fossil fuels,
resulting in significant greenhouse emissions. Clean production of hydrogen – the splitting of
water into hydrogen and oxygen using various methods – is costlier and still in the research

10
stage.18-19 The water splitting process is represented as follows, with molecular hydrogen being
generated at the cathode and oxygen generated at the anode:

2 H2O + Energy → 2H2 + O2


This method of clean hydrogen production is dependent on the electrochemistry of the
above reaction; the theoretical equilibrium potential of water is 1.23 V. However, in reality,
more than 1.23 V must be supplied to the system to drive the above reaction forward and
produce hydrogen. This additional potential (also known as overpotential) required is due to
the kinetic barriers of the oxygen evolution reaction (OER) at the anode and hydrogen
evolution reaction (HER) at the cathode.18 The OER, HER and overpotential equations are as
follows:

OER in alkaline solution: 4OH- → O2 + 2H2O + 4e-


HER in alkaline solution: 4H2O + 4e- → 2H2 + 4OH-
or
OER in acidic solution: 2H2O → O2 + 4H+ + 4e-
HER in acidic solution: 4H+ + 4e- → 2H2

Operational Voltage Required = Veq + ɳA + | ɳC | + ɳΩ - (E1)


Overpotential due to kinetic barrier = ɳA + | ɳC | + ɳΩ - (E2)
Veq = equilibrium potential of water

ɳA = overpotential required to overcome kinetic barrier of OER (anode)

ɳC = overpotential required to overcome kinetic barrier of HER (cathode)

ɳΩ = overpotential required to compensate resistance loss of device

Problems Facing Large-Scale Water Splitting


Three fundamental issues with water splitting prevent it from being used for large-scale
hydrogen production or renewable energy storage21:

(1) Due to the kinetic barrier imposed by HER/OER reactions, the real-world efficiency
of the process is significantly lower than theoretical thermodynamic limits as the voltage
needed is too high.

(2) The lifetime of the electrode materials is insufficient.

11
(3) Platinum, currently the best material for catalysing water splitting, is expensive and
scarce in supply.

Over the last decade, much research has been done to address the above issues by
investigating hydrogen and oxygen evolution reactions with various classes of materials. Since
water splitting is dependent on OER and HER, academic focus has been on investigating
electrode materials that allow high HER/OER activity with minimal overpotential and excellent
stability.21-22

2.3 General Electro-Catalyst Selection Criteria

Two important factors to consider when selecting a catalyst material (either as a cathode or as
an anode) are the overpotential and Tafel slope. As mentioned earlier, the kinetic barriers
imposed by HER/OER is currently an obstacle to efficient water splitting as the activation
energy needed, or the overpotential, is too high. The Butler–Volmer equations that govern
HER/OER activity is as follows21:

ɳA = EH + iR + ɳH - (E3)
ɳC = EO + iR + ɳO - (E4)
Where ɳA = overpotential at cathode ɳC =overpotential at anode

ɳH = HER overpotential ɳO = OER overpotential

EHER = Nernst potential of HER EOER = Nernst potential of OER

iR = ohmic potential drop caused by flow of current in electrolyte

𝑹𝑻
When ɳH/O < 0.005 V, ɳH/O = (ɳ𝐅𝐣o) j

−𝟐.𝟑𝑹𝑻 𝟐.𝟑𝑹𝑻
When ɳH/O ≥ 0.005 V, ɳH/O = a + b log (j) = log jo + log j - (E5)
𝒏𝑭𝒋𝒐 𝒏𝑭𝒋𝒐

Where
2.3𝑅𝑇
Where j = current density b= = Tafel slope
𝑛𝐹𝑗𝑜

jo = current exchange density n = number of electrons transferred

R = ideal gas constant F = Faraday constant

T = temperature

With equations (E3) to (E5), it can be deduced that low Tafel slopes and large exchange
current densities result in lower overpotential at the electrodes. This in turn lowers the
operational voltage needed for hydrogen to be produced, as shown by equation (E1). Hence,
electrochemically speaking, ideal catalysts for HERs or OERs should possess low Tafel slopes,

12
large exchange current densities and small overpotentials. In cases where catalysts have low
Tafel slopes and small exchange current densities, the better catalyst is the one with a smaller
overpotential at a given operational current density. With a smaller overpotential, less voltage
would be needed to drive the water splitting reaction, resulting in more efficient hydrogen
generation.21-23

2.3.1 HER Catalyst Selection

Hydrogen evolution reactions may involve two out of three possible reaction steps:

In a HER, either the Volmer-Heyrovsky (Steps 1+2) or Volmer-Tafel (Steps 1+3)


mechanism takes place at the cathode. Both reactions involve adsorbed hydrogen (Had) on the
catalyst (or electrode, since the electrode surface acts as a catalyst as well) surface and thus the
free energy of hydrogen adsorption (ΔGH) of the catalyst is critical. For HER, the hydrogen to
surface bonding cannot be too strong or too weak; strong hydrogen adsorption limits the
Volmer (or absorption) step while a weak ΔGH leads to a slow desorption (Heyrovsky or Tafel)
step. This is known as the Sabatier principle, which results in a “volcano plot” relationship
between the HER rate and ΔGH (Figure 5). According to literature, the ΔGH for optimal HER
activity is zero.21-23

22
Figure 5: Relationship between exchange current density and hydrogen absorption energy for HER catalysts.

13
2.3.2 OER Catalyst Selection

Compared to HER, the working principles of oxygen evolution reaction are not as well
understood. This is because thus far, there has been no conclusive evidence that are able to
validate any of the various proposed mechanisms of OER, some of which can be seen on Figure
6. Various issues plague the study of OER kinetics, such as the fact that it is the metal oxide –
not the metal – that catalyses OERs. Still, it is generally accepted that for OER catalysis, surface
metal ions (M) of the electro-catalyst are the active sites and OER is believed to proceed
through several intermediate reactions involving M-O, M-OOH and M-OO bonds.18, 25-27

Figure 6: Proposed reaction mechanisms for OER.26

Of the proposed mechanisms, most studies have chosen to use the Density Functional
Theory (DFT) calculated pathway for analysis of alkaline OERs, indicated by Path IV in Figure
6. This is because this pathway describes the reactivity trends on different catalysts fairly well
and thus appears the most reasonable.26

DFT-predicted OER Mechanism (in acidic conditions):

Reaction Step 1) H2 O + M M-OH + H+ + e-

Reaction Step 2) M-OH M-O + H+ + e-

Reaction Step 3) M-O + H2O M-OOH + H+ + e-

Reaction Step 4) M-OOH M + O2


where ΔGi is the Gibbs free reaction energy of step i

Unlike HERs, the adsorption energy of three types of bonding have to be considered in
OERs. This makes the selection of a catalyst more difficult, as the scaling relationship among
ΔGMO, ΔGMOH and ΔGMO across metal and metal oxide surfaces27 – as shown on Figure 7 –
means that it is not possible to fine-tune the ΔG of any single intermediate without affecting
the binding energies of the other two.

14
Figure 7: Schematic representation of the scaling relations for ΔGMOH, ΔGMO and ΔGMOOH as functions of ΔGMOH.18

Seeing that OER consists of multiple steps, the step with the largest change in Gibb’s Free
Energy determines the overpotential required of the overall OER:

ΔGOER = max {ΔG1, ΔG2, ΔG3, ΔG4}


ɳOER = (ΔGOER / q) – 1.23 V
Finally, since reaction step 2 of the theoretical OER mechanism contains information of
binding energies of all three intermediates (through the scaling relationship), catalysts can be
compared (Figure 8) by using a single universal descriptor to estimate the required OER
overpotential18:

ΔG2 = ΔGMOH – ΔGMO – qV

Figure 8: Comparison OER overpotential trends for rutile, anatase, Co3O4 and MnxOy oxides using the universal
descriptor ΔG2.18

Currently, platinum is the best electro-catalyst for HER and ruthenium oxide offers the
most optimal OER activity. Due to rising costs and limited supply of these precious metals,
much research has been focused on minimizing precious metal usage by using nanostructured
electrodes, alloying precious metals with other elements or completely replacing them as the
electro-catalyst.21-22

15
Of the two half reactions in water splitting, OER is the more demanding reaction. This is
because “it requires the distribution of four redox processes over a narrow potential range, the
coupling of multiple proton and electron transfers, and the formation of two oxygen-oxygen
bonds”.31 Hence, the efficiency of OER is a major factor in determining the feasibility of large-
scale water splitting processes and thus the focus from here on will be on OERs.

2.4 Transition Metal Phosphides

Currently, the mechanistic understanding of OERs with transition metal phosphide (TMP)
based catalysts is even more lacking than with traditional catalysts – the DFT-predicted OER
mechanism mentioned in the previous section may not even apply as TMPs undergo oxide
dissolution under acidic conditions. Several other factors, such as the method of preparation
and oxide layer thickness, strongly affect the current-voltage behaviour of TMPs; this results
in inconsistent kinetic data, thus causing much difficulty in ascertaining exact OER
mechanisms. Furthermore, most studies of TMPCs are focused on HER instead of the more
demanding oxygen evolution.31-33

Despite the lack of understanding of its catalytic mechanism in literature, TMPs have
shown much potential as an anode due to their high volumetric capacities.34-36 Additionally, it
is generally agreed that the ample amount of hydroxyl groups in the crystal structure of TMP
oxides grant them a high level of self-functionalization as the stable interactions between
oxygen and metal prevents particle agglomeration. This in turn helps retain the beneficial
effects of being a nano-sized material.35

The major drawback of TMPCs is their poor stability – in LIBs, Li+ ions tend to pulverize
the catalysts after a few cycles, leading to poor capacity retention and short cycling life. One
way to mitigate this problem is to construct the TMPCs into nanostructures where the
significantly higher interface surface area and allowance for volume expansion may improve
the stability of TMP-based catalysts.36 In particular, TMP hollow/solid nanoparticles, nanorods
and nanospheres have been reported to show high specific capacities and good stability34-36
(Figure 9). These improvements to TMPs may make them good contenders to be electro-
catalysts for Li/Na ion batteries and for OERs in water splitting.

16
Figure 9: Capacity and cycle life of hollow and solid CoP nanoparticles. 34

2.5 Transition Metal Dichalcogenides

Transition Metal Dichalcogenides (TMDs), with the general formula MX 2, are two-
dimensional layered materials that are currently being developed for catalysing water splitting.
With MX2 nanosheets in particular, each sheet is comprised of covalently bonded X-M-X
atomic layers and these layers are bound to each other by weak van der Waals forces.37

Since MX2 layers are only weakly bonded, TMDs are able to intercalate Li+/Na+ ions in
between the nanosheet layers. Indeed, TMD anodes in both LIBs and SIBs have been reported
to show reasonable capacity and excellent intercalation-deintercalation stability.38-40 For OERs,
the use of the ‘1T’ configuration (Figure 10) of molybdenum disulphide (MoS2) as a catalyst
requires an overpotential of 420mV, which is almost comparable to iridium oxides
(overpotential of 390mV) under standard conditions.38 For TMDs like 2H-MoS2 and WS2, the
superior catalytic properties are said to stem from the sulphur-terminated edges – that act as
catalytically active sites – and their nano-scale size.41 For 1T-MoS2, the basal plane is the
catalytically active site.38

Recently, a yolk – shell structured sphere of Co 9S8/MoS2 composite have been shown to
possess better charge rate and cycling stability than graphite anodes in LIBs.42 Thus, this report
attempts to achieve a superior catalytic material with such a yolk – shell structure, but by
combining TMPs with TMDs.

17
Figure 10: From the left: Trigonal prismatic (2H) and octahedral (1T) configurations of MoS2, respectively42

2.6 Transition Metal Phosphosulfides

Efforts to produce a catalyst with the advantages of both TMPs and TMDs have been
successful: cobalt phosphosulfides (CoPS) have been reported to require a HER overpotential
of 48 – 68mV when achieving a current density of 10mA/cm2 – better than any other non-
precious metal catalysts. The stability of CoPS is also excellent.24, 45-46 To our knowledge, there
has been no reports done on OERs with CoPS catalysts, but a recent study on pyrite-structured
NiCoPS system had reported with an overpotential of 230 mV47 (a respectable value for non-
precious metal catalysed OERs). The cause of these improvements over TMPs and TMDs is
the partial substitution of sulphur atoms by phosphorous atoms in the pyrite-structured CoS2
(Figure 11).

Figure 11: Pyrite structure of CoPS.43

Although the exact mechanics of CoPS catalysis is unclear, it is believed that the functions
of phosphorous are to induce the local charge density and accommodate the surface charge
state.48 It has also been confirmed that the improved HER activity is due to the fine-tuning of
hydrogen absorption energy by doping in the less electronegative P atoms.44 To our knowledge,
there are currently no reports on the performance of CoPS as anodes for lithium ion batteries.
Hence, encouraged by the promising HER activity of CoPS, this report aims to investigate the
material’s catalytic performance for the OER of water splitting, as well as look into the use of
CoPS in lithium batteries.

18
3. Materials and Methods

3.1 Synthesis

Figure 12: Schematic illustration of synthesizing CoPS yolk-shell nanoparticles

The procedure for synthesizing CoPS yolk-shell spheres via a three-step process is shown on
Figure 12.

Firstly, cobalt-precursor was synthesized by reacting cobalt powder with isopropyl alcohol
(IPA) and glycerol. Uniform Co-precursor spheres were synthesized through a facile modified
hydrothermal reaction. 0.2 mol of Co(NO3)2·6H2O (Sigma-Aldrich) and 4 ml of glycerol
(Sigma-Aldrich) were first dissolved into 20 ml of isopropanol (Sigma-Aldrich) under vigorous
magnetic stirring for 30 minutes. Then, the transparent solution was transferred to a Teflon-
lined stainless steel autoclave with a capacity of 45 mL and set aside at 180 °C for 6 hours.
After cooling down, the product was washed with ethanol for a few times and dried in an oven
at 60 °C for 12 hours.

Secondly, 20 mg of the as-prepared Co-precursor spheres was added it into 30 mL ethanol


solution containing 40 mg of thioacetamide (Sigma-Aldrich). This mixture was then shifted to
a Teflon-lined stainless steel autoclave with a size of 45 mL and set aside at 180 °C for 1-3
hours. The next step was to collect the black precipitate by repetitive washing with ethanol for
at least six times before placing it to vacuum dry at room temperature overnight.

The last step is the phosphorization of Co-S. The black precipitate and NaH2PO2 (Sigma-
Aldrich) were put at two separate positions in a porcelain boat with NaH 2PO2 at the upstream
side of the furnace. The weight ratio of precipitate to NaH2PO2 is 1:5. After being flushed with
argon gas, the centre of the furnace was elevated to 350 °C with a ramping rate of 2 °C min -1
and held at this temperature for 2 hours, and then naturally cooled to ambient temperature under

19
Ar atmosphere. The sulfurization and phosphorization time were varied to synthesize samples
with different S: P ratios.

The pure cobalt phosphide (Co-P) and cobalt sulphide (Co-S) yolk-shell structures were
obtained by one-step phosphorization in Ar funace and hydrothermal sulfurization (at 180 °C
for 6h), respectively.

3.2 Characterization

 X-Ray Diffraction (XRD) analysis was done with a Shimazu XRD-6000 diffractometer
(Cu Kα=1.5406Å at 40kV, 30mA) at a scan rate of 20C/min

 Morphology of the samples was characterized with a scanning electron microscope


(JEOL-2100F, 5.0kV)

 Elemental composition of the samples was done with Energy Dispersive Spectroscopy
(XPS) using JEOL-JSM5410LV

 Transmission Electron Microscope (TEM) images were obtained with a JEOL JEM-
2010 microscope with an accelerating voltage of 200 kV

 Galvanostatic charge and discharge was tested on a NEWARE multichannel battery


testing system in the voltage range of 0.01 – 3.0 V versus Li+/Li for LIB testing and
versus Ag+/AgCl for OER testing

3.3 Electrochemical Measurement

For LIB testing, the anode was prepared by mixing 70% active materials, 20% graphene, and
10% polyvinylidene difluoride. For the electrolyte, 1 M LiPF6 of solution was poured in a 1:1
(v/v) ratio mixture of ethylene carbonate and dimethyl carbonate. Lithium foil was used as both
the counter and reference electrode. Coin-type half-cells were assembled in an argon-filled
glove box and then tested on a NEWARE multichannel battery test system with galvanostatic
discharge and charge in the voltage range of 0.01–3.0 V versus Li+/Li.

For OER testing, the anode slurry was made with 5 mg cobalt phosphosulfide in 1 ml
isopropyl alcohol and 20 µL Nafion (5% wt. in lower aliphatic alcohols and water). Platinum
was used as the counter electrode and Ag/AgCl (in saturated KCl solution) was used as the
reference electrode in this three-electrode system, together with 1 M KOH as the electrolyte.
In all measurements, Ag/AgCl reference electrode was calibrated with respect to reversible
hydrogen electrode (RHE):

E (RHE) = E (Ag/AgCl) + 0.059 pH + 0.197.

20
4. Results and Discussion

4.1 Characterization

As shown on the following Figures 13(a) to 13(c), Co-P-S spheres were successfully
synthesized. The spheres were uniformly sized (400 – 552 nm) and uniformly distributed.
Additionally, the darker contrast seen on the middle of spheres compared to their edges suggest
that the spheres are hollow.

Figure 13: Morphology and chemical analysis of CoPS samples. (a) to (c) SEM images of uniform CoPS spheres (d) XRD
spectra of CoPS vs CoS

Crystal structure wise, XRD analysis of the samples (Figure 13d) shows that the pyrite
structure from Co-S remained intact after the final step of synthesis from Co-S to Co-P-S, as
the XPS spectrographs of the two compounds are the same. It can thus be inferred that the
phosphorization step simply substituted some sulphur atoms with phosphorus atoms.44 The
other possible phosphorization step is the substitution of R-complex groups by phosphorous
atoms, as shown below on Figure 14.

21
Figure 14: Reaction steps to CoPS.

On the following Figure 15(a), a small number of spheres with defects can be seen, where
the shells are broken and the ‘yolks’ are exposed. These defects were possibly caused by
excessive sonication during the mixing of reagents in both the sulfurization and
phosphorization steps. Excessive physical agitation from sonication may have damaged a small
percentage of the spheres. In Figures 15(b) and 15(c), the crystallites of the CoPS spheres can
be seen. The HRTEM clearly shows the fringe spacing of 0.54 nm and 0.25 nm, which
corresponds to the (111) and (220) planes of the phase (Figure 13d).

Figure 15: TEM images of (a) CoPS spheres with defects (b)/(c) crystalites present in CoPS spheres.

The morphology of the CoPS composite is shown on the following figures. The HAADF-
STEM image on 16(a) shows a distinct void space between the shell and the ‘yolk’ of the sphere,
confirming the yolk-shell structure of the cobalt phosphosulfide spheres that were synthesized.
With the EDS-STEM elemental mapping image shown on Figure 16(b), it can be seen that the
Co, P and S elements were uniformly throughout the entire structure of the yolk-shell sphere.

22
Figure 16: (a) HAADF-STEM and (b) EDS-STEM images of CoPS spheres.

XPS analysis of CoS after the sulfurization step and CoSP after the phosphorization step
are as shown below on Figure 17.

Figure 17: XPS spectra of CoPS and CoS.

On Figure 17(a) – at the second synthesis step where CoS is phosphorized to CoPS – XPS
analysis show that the concentration of Co (III) increased while Co (II) concentration decreased
significantly. From Figures 17(b) and 17(c), the XPS results also confirm the presence of P (II),
S (II) and phosphate/sulphate side products.

23
4.2 Performance in Lithium Ion Batteries

The galvanostatic discharge and charge behavior of the CoPS yolk-shell spheres in LIBs can
be seen on Figure 18. From 18(a), the discharge/charge capacities (in LIBs) of the first and
second cycles are 1650 mAhg -1/940 mAhg-1 and 900 mAhg-1/870 mAhg-1, respectively. This
data indicates a coloumbic efficiency (CE) of 56.9% for the first cycle and a high 96.7% for
the second cycle. The low coloumbic efficiency of the first cycle is attributed to initial side
reactions with the electrolyte, such as the formation of the solid-eletrolyte interphase (SEI)
layer.42

Figure 18: Electrochemical performance of CoPS in LIB/OER. (a) 1st and 2nd charge/discharge at current density of 0.2
A/g (b) Stability performance at different current densities from 0.2 to 8.0 A/g of CoPS, CoP and CoS.

The specific capacities of yolk-shell CoPS spheres as anodes in LIB (Figure 18b) are 870,
720, 600, 450, 390, 360 and 330 mAhg-1 at current densities of 0.2, 0.5, 1.0, 2.0, 4.0, 6.0 and
8.0 Ag-1, respectively. This is comparable to the capacity of yolk-shell structured Co9S8/MoS2
anodes in a previous study42. Considering that the theoretical capacity of graphite at 372 mAh/g,
the capacity of CoPS as LIB anodes is high. From Figure 18(b), it is clear that the CoPS samples
exhibit better performance than those of CoS and CoP samples.

To understand why the capacity of CoPS is higher, we can look at its intercalation
mechanisms. In TMPs, lithium storage can occur either via intercalation or conversion reaction
mechanisms. During intercalation, the P-P bonds are cleaved to allow the insertion of Li ions49.
During discharge, the P-P bonds are once again formed, as shown below in a cobalt phosphide
anode:

CoPn + xLi+ + xe− ↔ LixCoPn

24
During and after the intercalation mechanism, the Co-P bonds remain intact. Conversely,
Co-P bonds are broken when lithium is stored via the conversion reaction mechanism. This
mechanism results in nano-sized metal particles and lithium phosphides being formed49, as the
following reaction shows:

CoPn + 3nLi+ + 3ne− ↔ nLi3P + Co


Since CoPS has the same pyrite structure and the same presence of Co-P bonds as TMPs,
we postulate that CoPS stores lithium the same way as TMPs via the two mechanisms
mentioned above:

(E6) CoPS + xLi+ + xe− ↔ LixCoPS

(E7) CoPS + 4Li+ + 4e− ↔ Li2P + Li2S + Co

Reaction (E7) has a defined number of lithium ions, as the electronic states of P and S was
determined earlier in 4.1. For comparison purposes, the half reaction at a graphite anode with
a LiCoO2 cathode would be:

(E8) C6 + 6Li+ + 6e- ↔ LiC6

The high specific capacity recorded in this report can thus be attributed to the molecular
makeup of CoPS. The half-reaction of a graphite anode (E8) in an LIB theoretically requires 6
carbon atoms – a total molecular weight of 72 – to intercalate one lithium ion. In contrast, a
122
CoPS anode (E7) requires a total molecular weight of = 30.5 to contain one lithium ion.
4
72
This means that it takes theoretically 30.5 = 2.36 times less active materials for a CoPS anode

than a graphite anode to intercalate the same amount of charges. This agrees with the recorded
870 𝑚𝐴ℎ/𝑔
specific capacity of CoPS which is approximately = 2.33 times more than the
372 𝑚𝐴ℎ/𝑔

theoretical specific capacity of graphite.

The small discrepancy between the calculated and recorded CoPS capacity may be due to
side reactions that reduce the weight fraction of active CoPS material and the possible change
in electronic states of sulphur/phosphorous atoms during lithium intercalation/de-intercalation,
both of which were considered negligible and thus not examined in this study.

In terms of stability, CoPS shows much promise, as shown on Figure 19.

25
Figure 19: Stability and specific capacity of CoPS anodes at a current density of (a) 0.2 A/g (b) 1A/g and (c) 4 A/g

Figure 19(a) shows that CoPS remained stable for 50 cycles with a capacity of 850 mAh/g
at a current density of 0.2 A/g. Remarkably, CoPS was able to maintain a capacity of 475
mAh/g for 5000 charge-discharge cycles even at a high current density of 4A/g, as shown on
Figure 19(c).

Stability in CoPS is good due to the space in between the sphere and ‘yolk’, which is able
to accommodate the volumetric expansion of the anode during Li intercalation without
pulverizing the material. This yolk-shell configuration maintains the structural integrity of the
CoPS anode for a longer period of time than simple nanostructures, thus allowing for better
stability under the same electrochemical conditions.

26
4.3 Performance in Oxygen Evolution Reaction

Performance of CoPS anodes in catalysing OER was tested using a three-electrode system.
First, EAg+/AgCl (RHE) values of the system for the tested anodes were obtained, as shown on
the following figure.

Figure 20: Current vs. Applied Potential (with reference to AgCl electrode) of CoPS anodes in three-electrode system.

Then, the total operational potential in a water splitting reaction was calculated:

Operational Potential (vs. RHE) = EAg+/AgCl + AgCl Reference Electrode Potential + 0.059pH
EiR correction = I x R
Operational Potential = EAg+/AgCl + 0.197 + (0.059 x 14)
Operational Potential (iR corrected) = Operational Potential – EiR correction
Where

I = current

R = resistance between Luggin capillary and working electrode

A = surface area of working electrode

Using the above equations, Figure 21 (below) was obtained and it shows the current density
(signifying OER activity) vs. the operational potential of the water splitting reaction. It can be
seen that compared to CoP and CoS alone, CoSP spheres require less operational potential to
generate the same amount of current density, or OER activity.

27
Figure 21: Graph of Current Density vs. Potential of CoPS anodes in three-electrode system

Figure 22 below shows a clearer comparison of the overpotential required of each anode
tested in this study. The results show that cobalt phosphosulphide is the superior catalyst for
OER when compared to either CoS or CoP anodes. The presence of both sulphur and
phosphorous atoms in CoPS result in an obvious synergistic effect in catalyzing OER, with
CoS0.46P0.54 exhibiting the lowest overpotential slope of 48 mV/decade, or 270 mV of required
overpotential at a current density of 10 mA cm-2 . This is comparable with the most recent
advances in non-precious metal OER catalysts, such as the NiCoPS/carbon cloth anode46 which
requires 230 mV of overpotential at 10 mAcm-2. However, since that particular study used
NiCoPS nanowires that were many times smaller than the 400nm – 552nm yolk-shell spheres
in this report, the lower overpotential of NiCoPS anode can be attributed to its smaller size.

Figure 22: Graph of Overpotential vs. Log (Current)/OER activity in three-electrode system

28
As mentioned in section 2.2, an optimal catalyst for OER can be selected by using a
universal descriptor. However, this descriptor uses the binding energies of intermediates to the
catalyst surface, which are difficult values to obtain accurately. In light of this issue, authors J.
Suntivich et al. suggested that the filling of 3d eg-orbitals in a transition metal oxide affects the
binding strength of OER intermediates. The authors noted that a catalyst with an eg-orbital
filling of close to 1.25 (with a “volcano” plot as per Sabatier principle) would possess the
optimal catalytic ability for OERs50.

In section 4.1, we determined that Co3+ content increased when CoS was phosphorized to
CoSP. Thus, Tthe increase in Co3+ content may have modulated the eg-orbital filling towards
the optimal value. Additionally, the Co3+/Co2+ redox pair may have resulted in a combination
of cobalt (II) and cobalt (III) hydroxides/oxides being formed on the the surface, which in turn
are possibly more catalytically active than cobalt (II) hydroxide/oxide alone.

In addition to modifying the eg-orbital of cobalt, the heterodoping of phosphorus modified


the Gibb’s free energy of adsoption of the OER intermediates too. Since OER consists of three
intemdiates whose ΔG are all in a fixed relationship with each other, the exact modification of
ΔG of all the intermediates by phosphorus doping is difficult to decipher. However, a facile
deduction can be made; the universal OER activity indicator ΔG2 was possibly modified by
phosphorous doping and the optimal P content for the lowest ΔG2 possible in cobalt
phosphosulphides is close to being equal to the sulphur atom content.

Previous studies have shown that the doping of P into metals can result in the metals
becoming less conductive or even insulated. This is due to the electronegative P atoms
restricting electron delocalization in the metal. However, with appropriate M-P ratios, metal
phosphides have been known to exhibit excellent conductivity or even superconductivity51.
This implies that the Co-P ratio in CoS0.46P0.54 is close to the appropriate ratio for high
conductivity, thus allowing for better electrocatalytic performance.

Although there is currently no direct evidence that these sites are responsible for HER or
OER activity24, literature suggests that phosphorous and sulphur-terminated edges act as
catalytic active sites41; experiments have showed that the square-pyramidal surfaces of the
pyrite structure – metal centres bridged by dichacogenide dumbells – catalyzes OER24. With a
higher Co3+ content, the number of active phosphorous and sulphur edges may have increased,
further enhancing the OER ability of CoPS relative to CoP or CoS.

29
With regards to stability, the three-electrode system showed that CoPS maintained stable
OER activity for as long as 15 hours.

Figure 23: OER activity level of CoPS anode at a constant EAg/AgCl of 0.5 V

Figure 23 shows that OER activity started off at 12.6 mAcm-2 at the applied voltage of
EAg/AgCl = 0.5 V. Activity was reduced to 10 mAcm-2 at the 5th hour and remained stable for
another 10 hours. However, deviation of activity started to increase at the 4th hour.

The initial drop of activity from the start to the 5th hour may have been due to the build up
of side products over time, such as non-cobalt sulphates and phosphates. Contaminants in the
air may have entered the electrolyte and then reacted with the electrode surface to form inactive
compounds as well. This mix of catalytically inactive side products on the electrode surface
may have hindered the complex OER mechanics, thus resulting in a decrease in activity and
stability over time.

Regardless, CoPS is able to maintain OER activity without any large decrease in catalytic
ability.

30
5. Recommendations for Further Research

The results in section 4.1 have shown that the cycling stability and specific capacity of CoPS
anodes in LIBs are superior to graphite anodes. However, higher specific energy capacities are
always welcomed. We suggest that further research be performed on CoPS by changing the
synthesis method to produce hierarchical nanoflower-structured or urchin-like hollow spheres
instead, as these structures have been shown to increase charge transport, strain relaxation and
surface area.52-53 We recommend that a comparison study be done on the various nanostructures
that are possible, as rational nanostructure design can improve on the specific capacity while
retaining the stability of yolk – shell spheres.

Technically, the definition of a nanomaterial is a material with at least one dimension that
is between 1 and 100 nm. Therefore, the spheres reported in this study are not nano-sized in
the strictest sense. Yet, the spheres were able to compete with technically nano-sized materials
in terms of the overpotential required in OER46. We predict that if the size of CoPS can be
decreased (thus increasing the surface area) and/or porosity can be increased, the OER
performance will improve.

Although improving the capacity of LIBs may help reduce the global demand of lithium
metal, the rate at which new LIB-dependant application are being invented and used (such as
drones and electric cars) nevertheless imposes a heavy burden on the global lithium supply.
During this project, preliminary experiments carried out with CoPS anodes in sodium-ion
batteries (SIBs) had shown superior capacity and stability, especially when compared with
graphite anodes. Development of SIBs that can compete with LIBs (in terms of performance)
may prove to be a more impactful goal to pursue, as sodium is an abundant resource and thus
SIBs are cheaper to produce.

Finally, we also suggest studying on the different combinations of transition metals


possible in phosphosulphides. As mentioned in earlier sections, this is because NiCoPS coated
on carbon cloth was reported to require a low overpotential46 of 230 mv at 10 mAcm-2. Further
studies may include examining if the low overpotential is due to the Ni-Co-P-S synergy or the
improved conductivity from the carbon cloth. Previous studies have also shown that iron-nickel
oxides/hydroxides reduce the overpotential needed for OER54. Thus, readers may want to
investigate the synergy between Ni-Fe-Co transition metals and their oxide/hydroxides. These
studies may help improve transition metal phosphosulphides to the point of being commercially
viable.

31
In summary, future research on CoPS anodes may include the following:

1) The synthesis of CoPS in the form of nanoflowers or urchin-like hollow spheres

2) The reduction of CoPS yolk – shell spheres to less than 100 nanometers in diameter

3) Usage of transition metal phosphosulfides as anodes for sodium-ion batteries

4) Additional doping of nickel and iron into CoPS for improving catalytic efficiency

32
6. Conclusion

Homogenous yolk – shell structured spheres of cobalt phosphosulphide (diameter of 400 – 552
nm) were successfully synthesized. When tested as anodes for lithium-ion batteries, CoPS was
found to result in high specific capacity (870 mAh/g) and high rate capability due to the pyrite
structure and Co-P-S synergy. Due to the yolk – shell configuration that allows for rapid
volume expansion without pulverization, the cycling stability of CoPS was found to be good
as well.

With respect to OER, CoPS displayed an overpotential of 270 mV at 10 mA/cm2 and a low
Tafel slope of 48 mV dec-1. Although still inferior to platinum, this value is comparable to
many state-of-the-art non-precious metal catalysts studied in recent literature. Heteroatom
phosphorus doping was postulated to greatly modulate the electronic structure of cobalt, the
universal catalytic activity indicator ΔG2 and the electrical conductivity of CoPS. These three
main factors resulted in increased OER activity.

CoPS has proven to be a capable and stable anode for both LIBs and OERs. If
improvements can be made, such as decreasing the size and increasing the porosity of the
material, the performance of this bi-functional material may be enhanced even further. In
particular, OER performance needs to be improved; possibly by doping additional transition
metals like nickel and iron into CoPS.

33
References

[1]Y. Li, Y.Lu, C. Zhao, Y.Hu, M.Titirici, H. Li, X. Huang and L. Chen. “Recent advances of
electrode materials for low-cost sodium ion batteries towards practical application for grid
energy storage”, Energy Storage Materials, 2017, vol. 7, p. 130–151.
http://dx.doi.org/10.1016/j.ensm.2017.01.002
[2] X. Zuo, J. Zhu, P. Müller-Buschbaum and Y. Cheng. “Silicon based lithium-ion battery
anodes: A chronicle perspective review”, Nano Energy, 2017, vol. 31, p. 113-143.
http://dx.doi.org/10.1016/j.nanoen.2016.11.013
[3] Leonid Leiva, 2013. “Memory effect now also found in lithium ion batteries. Science X
Network. Retrieved from https://phys.org/news/2013-04-memory-effect-lithium-ion-
batteries.html (accessed on 14 Jan 2017)
[4] K. Fehrenbacher, 2016. “Why Tesla’s New Battery Pack Is Important”. Fortune. Retrieved
from http://fortune.com/2016/08/24/tesla-100kwh-battery-pack/ (accessed on 14 Jan 2017)
[5] S. Fang, L. Shen and X. Zhang. “Chapter 9 - Application of Carbon Nanotubes in Lithium-
ion Batteries”, In Micro and Nano Technologies. Industrial Applications of Carbon Nanotubes,
2016, p. 251-276. http://dx.doi.org/10.1016/B978-0-323-41481-4.00009-5 (How LIBs work,
nano)
[6] The Electric Energy, N.D. “Lithium and Lithium Ion Battery Technology”. Retrieved from:
http://theelectricenergy.com/lithium-and-lithium-ion-battery/ (accessed on 14 Jan 2017)
[7] Y. Wang, B. Liu, Q. Li, S. Cartmell, S. Ferrara, Z. D. Deng and J. Xiao. “Lithium and
lithium ion batteries for applications in microelectronic devices: A review”, Journal of Power
Sources, 2015, Volume 286, Pages 330-345. http://dx.doi.org/10.1016/j.jpowsour.2015.03.164.
[8] Doughty D. H., 2013. “Vehicle Battery Safety Roadmap Guidance” (PDF). National
Renewable Energy Laboratory. Retrieved from:
https://searchworks.stanford.edu/view/11256279 (accessed on 15 Jan 2017)
[9] Tam Hunt, 2015. “Is There Enough Lithium to Maintain the Growth of the Lithium ion
Battery Market?”. Greentech Media. Retrieved from:
https://www.greentechmedia.com/articles/read/Is-There-Enough-Lithium-to-Maintain-the-
Growth-of-the-Lithium-ion-Battery-M (accessed on 14 Jan 2017)
[10] J. Dessjardins, 2016. “Explaining the Surging Demand for Lithium ion Batteries (Part 3).
Visual Capitalist”. Retrieved from http://www.visualcapitalist.com/explaining-surging-
demand-lithium-ion-batteries/ (accessed on 14 Jan 2017)
[11] Metalary, 2016. “Lithium Price”. Metalary. Retrieved from:
https://www.metalary.com/lithium-price/ (accessed on 14 Jan 2017)
[12] S.T. Senthilkumar, M. Abirami, J. Kim, W. Go, S. M. Hwang and Y. Kim. “Sodium ion
hybrid electrolyte battery for sustainable energy storage applications”, Journal of Power
Sources, 2017, Volume 341, Pages 404-410. http://dx.doi.org/10.1016/j.jpowsour.2016.12.015
[13] Q. Pan, Z. Li, W. Zhang, D. Zeng, Y. Sun and H. Cheng. “Single ion conducting sodium
ion batteries enabled by a sodium ion exchanged poly(bis(4-carbonyl benzene sulfonyl)imide-

34
co-2,5-diamino benzesulfonic acid) polymer electrolyte”, Solid State Ionics, 2017, Volume 300,
Pages 60-66. http://dx.doi.org/10.1016/j.ssi.2016.12.001
[14] S. Peng, X. Han, L. Li, Z. Zhu, F. Cheng, M. Srinivansan , S.Adams and S. Ramakrishna.
“Unique Cobalt Sulfide/Reduced Graphene Oxide Composite as an Anode for Sodium ion
Batteries with Superior Rate Capability and Long Cycling Stability”. Small, 2016, Volume 12.
http://dx.doi.org/10.1002/smll.201502788
[15] H. Zhu, Z. Jia, Y. Chen, N. Weadock, J. Wan, O. Vaaland, X. Han, T. Li, and L. Hu. “Tin
Anode for Sodium ion Batteries Using Natural Wood Fiber as a Mechanical Buffer and
Electrolyte Reservoir”, Nanoletters, 2013, Volume 13. http://dx.doi.org/10.1021/nl400998t
[16] Y. Liu, N Zhang, L. Jiao, Z. Tao and J. Chen. “Ultra small Sn Nanoparticles Embedded in
Carbon as High-Performance Anode for Sodium ion Batteries”, Advanced Functional
Materials, 2015, Volume 25 (2), Pages 214-220. http://dx.doi.org/10.1002/adfm.201402943
[17] V. Chintala and K.A. Subramanian. “A comprehensive review on utilization of hydrogen
in a compression ignition engine under dual fuel mode”. Renewable and Sustainable Energy
Reviews, 2015, Volume 70, Pages 472-491. http://dx.doi.org/10.1016/j.rser.2016.11.247
[18] R. L. Doyle, & M. E. G. Lyons. The Oxygen Evolution Reaction: Mechanistic Concepts
and Catalyst Design, 2016, Photoelectrochemical Solar Fuel Production. Springer
International Publishing. Retrieved from: http://link.springer.com/chapter/10.1007%2F978-3-
319-29641-8_2#page-1 (accessed on 16 Jab 2017)
[19] S. Banerjee, Md. Nor. Musa and A. B. Jaafar. “Economic assessment and prospect of
hydrogen generated by OTEC as future fuel”, International Journal of Hydrogen Energy, 2016
Volume 42, Issue 1, 5, Pages 26–37. http://dx.doi.org/10.1016/j.ijhydene.2016.11.115
[20] M. D. Bhatt, and J. S. Lee. “Recent theoretical progress in the development of photoanode
materials for solar water splitting photoelectrochemical cells”, Journal of Materials Chemistry
A, 2015, Volume 3. http://dx.doi.org/10.1039/c5ta00257e
[21] D. Strmcnik, P. P. Lopes, B. Genorio, V. R. Stamenkovic, & N. M. Markovic. “Design
principles for hydrogen evolution reaction catalyst materials”, Nano Energy, 2016, Volume 29,
Pages 29-36. http://dx.doi.org/10.1016/j.nanoen.2016.04.017
[22] J. D. Benck, T. R. Hellstern, J. Kibsgaard, P. Chakthranont, and T. F. Jaramillo.
“Catalyzing the Hydrogen Evolution Reaction (HER) with Molybdenum Sulfide
Nanomaterials”, ACS Catalogue, 2014, Volume 4 (11), Pages 3957–3971.
http://dx.doi.org/10.1021/cs500923c
[23] M. Zeng and Y. Li. “Recent advances in heterogeneous electrocatalysts for the hydrogen
evolution reaction”, Journal of Materials Chemistry A, 2015, Volume 3.
http://dx.doi.org/10.1039/C5TA02974K
[24] M. Cabán-Acevedo, M. L. Stone, J. R. Schmidt, J. G. Thomas, Q. Ding, H. Chang, M.
Tsai, J. He, and S. Jin. “Efficient hydrogen evolution catalysis using ternary pyrite-type cobalt
phosphosulphide”, Nature Materials, 2015, Volume 14, Pages 1245–1251.
http://dx.doi.org/10.1038/NMAT4410

35
[25] M. S. Burke, L. J. Enman, A. S. Batchellor, S. Zou, and S. W. Boettcher. “Oxygen
Evolution Reaction Electrocatalysis on Transition Metal Oxides and (Oxy) hydroxides:
Activity Trends and Design Principles”, Chemistry of Materials, 2015, Volume 27 (22), Pages
7549 - 7558. http://dx.doi.org/10.1021/acs.chemmater.5b03148
[26] L. Trotochaud and S. W. Boettcher. “Precise oxygen evolution catalysts: Status and
opportunities”, Scripta Materialia, 2014, Volume 74, Pages 25-32.
http://dx.doi.org/10.1016/j.scriptamat.2013.07.019
[27] T. Reier, H. N. Nong, D. Teschner, R. Schlögl and P. Strasser. “Electrocatalytic Oxygen
Evolution Reaction in Acidic Environments – Reaction Mechanisms and Catalysts”, Advanced
Energy Materials, 2016, Volume 7 (1). http://dx.doi.org/10.1002/aenm.201601275
[28] I. C. Man, H. Su, F. Calle-Vallejo, H. A. Hansen, J. I. Martínez, N. G. Inoglu, J. Kitchin,
T. F. Jaramillo, J. K. Nørskov and J. Rossmeisl. “Universality in Oxygen Evolution
Electrocatalysis on Oxide Surfaces”, ChemCatChem, 2011, Volume 3 (7), Pages 1159 – 1165.
http://dx.doi.org/10.1002/cctc.201000397
[29] E. Fabbri, A. Habereder, K. Waltar, R. Kötz, & T. J. Schmidt. “Developments and
perspectives of oxide-based catalysts for the oxygen evolution reaction”, Catalysis Science &
Technology, 2014, Volume 4 (11). http://dx.doi.org/10.1039/c4cy00669k
[30] X. Xu, F. Song, & X. Hu. “A nickel iron diselenide-derived efficient oxygen-evolution
catalyst”, Nature Communications, 2016, Volume 7. http://dx.doi.org/10.1038/ncomms12324
[31] C. C. L. McCrory, S Jung, I. M. Ferrer, S. M. Chatman, J. C. Peters and T. F. Jaramillo.
"Benchmarking Hydrogen Evolving Reaction and Oxygen Evolving Reaction Electrocatalysts
for Solar Water Splitting Devices”, Journal of American Chemical Society, 2015Volume 137
(13), Pages 4347-4357. http://dx.doi.org/10.1021/ja510442p
[32] Y. Surendranath, M. W. Kanan and D. G. Nocera. “Mechanistic Studies of the Oxygen
Evolution Reaction by a Cobalt-Phosphate Catalyst at Neutral pH”, Journal of American
Chemical Society, 2010, Volume 132, Pages 16501–16509.
http://dx.doi.org/10.1021/ja106102b
[33] E. J. Popczun, C. G. Read, C. W. Roske, N. S. Lewis, & R. E. Schaak. “Highly Active
Electrocatalysis of the Hydrogen Evolution Reaction by Cobalt Phosphide Nanoparticles”,
Angewandte Chemie International Edition, 2014, Volume 53 (21), Pages 5427 – 5430.
http://dx.doi.org/10.1002/anie.201402646
[34] E. Aslan, I. Akin and I. H. Patir. “Highly Active Cobalt Sulfide/Carbon Nanotube Catalyst
for Hydrogen Evolution at Soft Interfaces”, Chemistry A European Journal, 2016, Volume 22
(15), Pages 5342 – 5349. http://dx.doi.org/10.1002/10.1002/chem.201505048
[35] D. Yang, J. Zhu, X. Rui, H. Tan, R. Cai, H. E. Hoster, D. Y. W. Yu, H. H. Hng and Q.
Yan. Synthesis of Cobalt Phosphides and Their Application as Anodes for Lithium Ion
Batteries”, Applied Materials and Interfaces, 2012, Volume 5 (3), Pages 1093–1099.
http://dx.doi.org/10.1021/am302877q
[36] M. Sun, H. Liu, J. Qu and J. Li. “Earth-Rich Transition Metal Phosphide for Energy
Conversion and Storage”, Advanced Energy Materials, 2016, Volume 13.
http://onlinelibrary.wiley.com/doi/10.1002/aenm.201600087/abstract
36
[37] S. Boyanov & K. Annou & C. Villevieille & M. Pelosi, D. Zitoun and L. Monconduit.
“Nanostructured transition metal phosphide as negative electrode for lithium-ion batteries”,
Ionics, 2007, Volume 14 (3), Pages 183-190. http://dx.doi.org/10.1007/s11581-007-0170-3
[38] F. Wang, T. A. Shifa, X. Zhan, Y. Huang, K. Liu, Z. Cheng, C. Jiang and J. He. “Recent
advances in transition-metal dichalcogenide based nanomaterials for water splitting”,
Nanoscale, 2015, Volume 7, Pages 19764-19789. http://dx.doi.org/10.1039/C5NR06718A
[39] L. Ji, Z. Lin, M. Alcoutlabi and X. Zhang. “Recent developments in nanostructured anode
materials for rechargeable lithium-ion batteries”, Energy & Environmental Science, Volume 4,
2011, Pages 2682-2699. http://dx.doi.org/10.1039/C0EE00699H
[40] S. Peng, X. Han, L. Li, Z. Zhu, F. Cheng, M. Srinivansan, S. Adams, and S. Ramakrishna.
“Unique Cobalt Sulfide/Reduced Graphene Oxide Composite as an Anode for Sodium-Ion
Batteries with Superior Rate Capability and Long Cycling Stability”, Small, 2016, Volume 13
(10), Pages s 1359–1368. http://dx.doi.org/10.1002/smll.201502788
[41] Jingjie Wu, M. Liu, K. Chatterjee, K. P. Hackenberg, J. Shen, X. Zou, Y. Yan, J. Gu, Y.
Yang, J. Lou, & P. M. Ajayan. “Exfoliated 2D Transition Metal Disulfides for Enhanced
Electrocatalysis of Oxygen Evolution Reaction in Acidic Medium”, Advanced Materials
Interfaces, 2016Volume 3 (9). http://dx.doi.org/10.1002/admi.201500669
[42] H. Geng, J. Yang, Z. Dai, Y. Zhang, Y. Zheng, H. Yu, H. Wang, Z. Luo, Y. Guo, Y. Zhang,
H. Fan, X. Wu, J. Zheng, Y. Yang, Q. Ya, & Hongwei Gu. “Co9S8/MoS2 Yolk–Shell Spheres
for Advanced Li/Na Storage”, Small, 2017. http://dx.doi.org/10.1002/smll.201603490
[43] Rajesh Kappera, D. Voiry, S. E. Yalcin, B. Branch, G. Gupta, A. D. Mohite and M.
Chhowalla. Phase-engineered low-resistance contacts for ultrathin MoS2 transistors”, Nature
Materials, 2014, Volume 13, Pages 1128–1134. http://dx.doi.org/10.1038/nmat4080
[44] A. Ambrosi, Z. Soferb and M. Pumera. “2H - 1T phase transition and hydrogen evolution
activity of MoS2, MoSe2, WS2 and WSe2 strongly depends on the MX2 composition”,
Chemical Communications, 2015, Volume 51, Pages 8450-8453.
http://dx.doi.org/10.1039/c5cc00803d
[45] J. Kibsgaard and T. F. Jaramillo. “Molybdenum Phosphosulfide: An Active, Acid-Stable,
Earth-Abundant Catalyst for the Hydrogen Evolution Reaction”, Angewandte Chemie
International Edition, 2014, Volume 53 (52), Pages 14433–14437.
http://dx.doi.org/10.1002/anie.201408222
[46] W. Liu, E. Hu, H. Jiang, Y. Xiang, Z. Weng1, M. Li, Qi Fan, X. Yu, E. I. Altman and H.
Wang. “A highly active and stable hydrogen evolution catalyst based on pyrite-structured
cobalt phosphosulfides”, Nature Communications Volume, 2016, 7.
http://dx.doi.org/10.1038/ncomms10771 10
[47] J. Li, Z. Xia, X. Zhou, Y. Qin, Y. Ma, Y. Qu. “Quaternary pyrite-structured nickel/cobalt
phosphosulfide nanowires on carbon cloth as efficient and robust electrodes for water
electrolysis”, Nano Research, 2017, Pages 1-12. http://dx.doi.org/10.1007/s12274-016-1335-z
[48] A. Wu, C. Tian, H. Yan, Y. Jiao, Q. Yan, G. Yang and H. Fu. “Hierarchical MoS2@MoP
core–shell heterojunction electrocatalysts for efficient hydrogen evolution reaction over a
broad pH range”, Nanoscale, 2016, Volume 8 (21). http://dx.doi.org/10.1039/c6nr02803a
37
[49] X. Wang, H. Kim, Y. Xiao and Y. Sun. “Nanostructured metal phosphide-based materials
for electrochemical energy storage”, Journal of Materials Chemistry A, 2016, Volume 4, Pages
14915-14931. http://dx.doi.org/10.1039/C6TA06705K
[50] Suntivich J, May KJ, Gasteiger HA, Goodenough JB, & Shao-Horn Y. “A perovskite
oxide optimized for oxygen evolution catalysis from molecular orbital principles”, Science,
2011, Volume 334 (6061) Pages 1383-1385 http://dx.doi.org/10.1126/science.1212858
[51] Y. Shi and B. Zhang. “Recent advances in transition metal phosphide nanomaterials:
synthesis and applications in hydrogen evolution reaction”, Royal Society of Chemistry, 2016,
Volume 45, Pages 1529-1541. http://dx.doi.org/10.1039/C5CS00434A
[52] X. Rui, H. Tan, D. Sim, W. Liu, Chen Xu, Huey Hoon Hng, R. Yazami, T. M. Lim, and
Q. Yan. “Template-free synthesis of urchin-like Co3O4 hollow spheres with good lithium
storage properties” Journal of Power Sources, 2103, Volume 222, 15 January 2013, Pages 97–
102. http://dx.doi.org.ezlibproxy1.ntu.edu.sg/10.1016/j.jpowsour.2012.08.094
[53] Y. Tang, X. Rui, Y. Zhanga, T. M. Lim, Z. Dong, H. H. Hnga, X. Chena, Q. Yan and Z.
Chen. “Vanadium pentoxide cathode materials for high-performance lithium-ion batteries
enabled by a hierarchical nanoflower structure via an electrochemical process”, J. Material
Chemistry A, 2013, Volume 1, Pages 82-88.
http://pubs.rsc.org.ezlibproxy1.ntu.edu.sg/en/content/http://dx.doi.org.ezlibproxy1.ntu.edu.sg/
10.1039/C2TA00351A/2013/ta/c2ta00351a
[54] Y. Cheng and S. P. Jiang. “Advances in electrocatalysts for oxygen evolution reaction of
water electrolysis-from metal oxides to carbon nanotubes”, Progress in Natural Science:
Materials International, 2015, Volume 25, Issue 6, Pages 545–553.
http://dx.doi.org.ezlibproxy1.ntu.edu.sg/10.1016/j.pnsc.2015.11.008

38

S-ar putea să vă placă și