You are on page 1of 12

Journal of Electroanalytical Chemistry 717–718 (2014) 177–188

Contents lists available at ScienceDirect

Journal of Electroanalytical Chemistry


journal homepage: www.elsevier.com/locate/jelechem

Determination of the ‘‘NiOOH’’ charge and discharge mechanisms


at ideal activity
Matthew Merrill ⇑, Marcus Worsley, Arne Wittstock, Juergen Biener, Michael Stadermann
Chemical Sciences Division, Lawrence Livermore National Laboratory, Livermore, CA 94550, United States

a r t i c l e i n f o a b s t r a c t

Article history: Optimization of electrodeposition conditions produced Ni(OH)2 deposits chargeable up to 1.84 ± 0.02 e
Received 26 March 2013 per Ni on and the resulting nickel oxide/hydroxide active material could subsequently deliver 1.58 ± 0.02
Received in revised form 11 January 2014 e per Ni ion (462 mA h/g) over a potential range <0.2 V. The ability of the ‘‘NiOOH’’ active material to
Accepted 16 January 2014
deliver an approximately ideal charge and discharge facilitated a coulometric and thermodynamic anal-
Available online 24 January 2014
ysis through which the charge/discharge mechanisms were determined from known enthalpies of forma-
tion. The (dis)charge states were confirmed with in situ Raman spectroscopy. The mechanisms were
Keywords:
additionally evaluated with respect to pH and potential dependence, charge quantities, hysteresis, and
NiOOH
Ni(OH)2
fluoride ion partial inhibition of the charge mechanism. The results indicate that the ‘‘NiOOH’’ (dis)-
Nickel electrode charges as a solid-state system with mechanisms consistent with known nickel and oxygen redox reac-
Nickel oxy-hydroxide tions. A defect chemistry mechanism known for the LiNiO2 system also occurs for ‘‘NiOOH’’ to cause both
Active oxygen high activity and hysteresis. Similar to other cation insertion nickel oxides, the activity of the ‘‘NiOOH’’
mechanism is predominantly due to oxygen redox activity and does not involve the Ni4+ oxidation state.
The ‘‘NiOOH’’ was produced from cathodic electrodeposition of Ni(OH)2 from nickel nitrate solutions onto
highly oriented pyrolytic graphite at ideal electrodeposition current efficiencies and the deposition mech-
anism was also characterized.
Published by Elsevier B.V.

1. Introduction Recent developments in spectroscopic analysis, theoretical


modeling, and the analysis of nickel oxide materials in Li-ion bat-
Nickel oxides have been the dominant cathode material for sec- teries challenge the conventionally assumed nickel oxide mecha-
ondary batteries over many years and Li-ion batteries have only re- nism in aqueous conditions and indicate that the mechanism
cently emerged as a superior alternative due to their higher energy needs revision. The conventionally assumed mechanism for the en-
densities. Still, batteries with nickel oxide cathodes have several ergy producing reaction in nickel oxide/hydroxide batteries is
features that are superior to Li-ion batteries with regards to higher based upon the oxidation of nickel ions from Ni2+ to Ni4+ [3,5].
power densities, non-flammable electrolytes, and less costs [1,2]. The only evidence specifically supporting the Ni4+ state is the sim-
In practice, the nickel oxide material is the limiting electrode for ilarity in the first Ni–O bond length or Ni K-edge shift energy ob-
energy density in the battery with an activity of 0.4 e per Ni ion served though X-ray absorption studies [6–10]. The actual Ni4+
or 115 mA h/g Ni(OH)2 [2,3]. Nickel oxide materials have been re- valence state has been conventionally thought to occur in the fully
ported in the literature to reach high activities of up to 1.65 e per oxidized c-NiOOH state because of the X-ray adsorption experi-
Ni ion or 475 mA h/g Ni(OH)2 yet this higher performance has ments in which BaNiO3 or KNiIO6 reference materials were as-
never been realized in practical batteries [2–4]. An improvement sumed to represent an actual Ni4+ state. These materials were
in the energy density of nickel oxides by a factor of 3–4 could re- originally assumed to contain nickel in the Ni4+ valence state be-
store interest in this battery material. Both the highly active nickel cause there was reasonable confusion between the formal and ac-
oxide mechanisms and the mechanisms for producing the highly tual valence states at the time [11]. A formal oxidation state of Ni4+
active nickel oxide materials are further investigated here in order includes the possibility that the actual valence state is Ni3+ because
to facilitate the practical implementation of more active nickel oxidation of oxide has occurred. It has now been well-established
oxide materials. (even by some of the same authors, O’Grady and Mansour) from
Mössbauer, magnetic susceptibility, and X-ray photoelectron spec-
⇑ Corresponding author. Tel.: +1 925 422 9262. troscopy (XPS) experiments that the BaNiO3 or KNiIO6 reference
E-mail address: merrill10@llnl.gov (M. Merrill). materials (as well as Li(1x)NiO2 and c-NiOOH) all have the same

http://dx.doi.org/10.1016/j.jelechem.2014.01.022
1572-6657/Published by Elsevier B.V.
178 M. Merrill et al. / Journal of Electroanalytical Chemistry 717–718 (2014) 177–188

D53d space group symmetry involving a Ni3+ which has been some- (iii) and (iv) [4]. The characterization of the cathodic electrodepos-
what distorted due to the oxidation of some O2 [12–18]. XPS ition mechanism is relevant for explaining why deposits of Ni(OH)2
experiments specifically indicate that nickel is Ni3+ in c-NiOOH with optimum activity as ‘‘NiOOH’’ can only be produced in a rel-
and not Ni4+ [19–22]. The uncertainty involved with Jahn-Teller atively narrow range of conditions as well as why the limited con-
distortions causes the modeling of nickel oxide structures and ditions may be challenging for practical implementation in
bond lengths alone to be insufficient for determining nickel oxida- batteries.
tion states despite the numerous attempts to do so [14,15,18].
Recent experiment and theoretical modeling of the LiXNiO2 i. The OH gradient generated by the electrochemical reduc-
2+
electrode in lithium batteries has also indicated that oxygen redox tion of NO 3 through (1) chemically precipitates Ni as
chemistry is primarily responsible for the charge storage instead of Ni(OH)2 according to (2) in an 1 Ni(OH)2 per 2 e ratio
the oxidation of nickel ions to the 4+ state [16,18,23–26]. The oxy- [4,31–33]. The further reduction of NO 2 to either N2 or
gen redox chemistry occurs when electron vacancy forms on O2 NH4+ can also electrochemically generate a pH gradient,
instead of on Ni3+ and so the nickel ions are not being oxidized both with an additional 2 Ni(OH)2 per 3 e ratio [34].
to an actual Ni4+ valence state. Similarly, the electron vacancies ii. The reduction of water to form molecular hydrogen, H2,
in LiXNi(1X)O also occur on oxygen species to form O plus Ni2+ according to reaction (3) is another mechanism which can
when X < 0.4 and the electron vacancy only shifts to nickel to form electrochemically an OH gradient for chemically precipitat-
O2 plus Ni3+ as X approaches the 0.5 of LiNiO2 [17,27–30]. It ing Ni(OH)2 through reaction (2) in a 1 Ni(OH)2 per 2 e ratio
should therefore be expected that the aqueous ‘‘NiOOH’’ mecha- [5]. This mechanism is a likely mechanism when NO 3 is
nism involves oxygen redox activity instead of the actual Ni4+ state. absent from the electrodeposition solution.
A more thorough and comprehensive review and analysis of the iii. The reduction of Ni2+ to the metallic state Ni0 according to
‘‘NiOOH’’ mechanism is available online in the Supporting Informa- (4) occurs at relevantly cathodic potentials and leads to
tion section. We develop a new mechanism for the charging mech- two more possible cathodic deposition mechanisms. Water
anism of nickel oxide materials which is consistent with both our is not electrochemically stable on Ni at moderate potentials
experimental results and the literature. This mechanism incorpo- and will oxidize Ni0 to Ni2+ through (4) or to Ni(OH)2
rates the oxygen redox activity caused by electron vacancies on through (5), depending upon pH [34,35].
oxide species. A better understanding of oxygen redox center for- iv. The nitrate ion is such a strong oxidizing agent that it can
mation and stabilization could lead to better battery active materi- oxidize Ni0 even under a large, negative potential bias. The
als regardless of whether it is a H+-insertion system of an aqueous direct coupling of nitrate reduction (1) to Ni0 oxidation (4)
battery or a Li+-insertion systems of a Li+ ion battery. is expected to cause the immediate precipitation of Ni(OH)2
‘‘NiOOH’’ in its most active form is expected to constitute the before either the Ni2+ or OH products could diffuse back
optimal condition for mechanistic analysis. Minimizing the pres- into the bulk electrolyte phase and should therefore be
ence of side reactions will improve the resolution in the thermody- expected when ideal electrodeposition current efficiency is
namic analysis and minimizing the presence of inactive states will occurring.
improve the clarity of spectroscopic analysis. The oxidation of
Ni(OH)2 produced from cathodic electrodeposition from nickel ni- 2. Methods
trate solutions has resulted in the highest ‘‘NiOOH’’ activity (1.65
e per Ni ion) reported in the literature and will therefore be General electrochemical experiments were performed in a Tef-
implemented here [4]. The cathodic electrodeposition of Ni(OH)2 lon electrochemical cell. A Teflon face plate and corrosion-resistant
from nickel nitrate and other nickel solutions is likely more com- silicone gasket (McMaster-Carr) exposed a circular, 1 cm diameter
plex than the conventional mechanism typically presupposed in (0.785 cm2) working electrode surface of highly oriented pyrolytic
the literature because of the multiplicity of possible reaction path- graphite (HOPG, ZYH grade, Bruker AFM Probes) on the bottom of
ways. The five reactions of Table 1 collectively cause the four dif- the chamber. A 2 cm2 HOPG (ZYH grade, K-Tek Nanotechnology)
ferent mechanisms (i–iv) below through which Ni(OH)2 can be was used in conjunction with a 1.75 cm diameter Teflon face plate
cathodically deposited. Ni(OH)2 electrodeposition mechanisms (i) for in situ Raman experiments. Used HOPG surfaces were removed
and (ii) have been previously proposed in the literature. It is not with (Scotch) tape to expose a new, clean surface. A 30 ga platinum
likely that electrodeposition mechanisms (i) and (ii) can achieve wire (Hauser & Miller) counter electrode was coiled around the tip
ideal electrodeposition current efficiency because they depend of a double junction SCE reference electrode (Radiometer Analyti-
upon the chemical precipitation of Ni(OH)2 from OH concentra- cal) to be jointly inserted into the cell chamber from the top. The
tion gradients. The observation that the most highly active nickel general electrochemical experiments were performed with a Bio-
oxide materials are produced under the condition of ideal electro- logic VSP electrochemical workstation while the in situ Raman
deposition current efficiency motivates the re-evaluation of the spectroscopy experiments were performed with a Biologic SP200.
electrodeposition mechanism and the introduction of mechanisms Raman spectra were collected with Nicolet Almega XR dispersive
micro-Raman spectrometer, using 633 nm excitation length and a
water immersible Olympus LumPlanFl 60x objective for collecting
Table 1 the Raman scattered radiation. The in situ Raman samples were
Eqs. (1)–(5): the reactions of the Ni(OH)2 deposition mechanisms [32]. prepared by electrodeposition onto HOPG from 0.08 M Ni(NO3)2
NO    solutions at 1.0 mA/cm2 for 10 min. The spectra were collected
3 þ 2e þ H2 O $ NO2 þ 2OH (1)
E0 ð1Þ ¼ 0:835  0:0592  pH þ 0:0296  logð½NO 
3 =½NO2 Þ
for the first discharge and the second charge in 3 M KOH. The ‘‘NiO-
Ni

þ 2OH $ NiðOHÞ2 (2) OH’’ was initially charged at 20 C, the potential was held constant

logðNi Þ ¼ 12:18  2  pH for approximately 15 min during spectra collection, and the poten-
2H2 O þ 2e $ H2 þ 2OH (3) tial was changed between spectra at 1 mV/s. Metallic nickel layers
E0 ð3Þ ¼ 0:000  0:0592  pH of about 20 nm in thickness were deposited onto HOPG substrates
Ni2þ þ 2e $ Ni (4)
from Techni Nickel HT-2 solutions (Technic Inc.) at 10 mA/cm2 cur-
E0 ð4Þ ¼ 0:250 þ 0:0296  logð½Ni2þ Þ rent density for 5 s.
Ni þ 2H2 O $ NiðOHÞ2 þ 2e þ 2Hþ (5)
Masses of the Ni(OH)2 active material were determined by dis-
E0 ð5Þ ¼ 0:110  0:0592  pH
solving samples in 2 ml of 10% weight/volume nitric acid under
M. Merrill et al. / Journal of Electroanalytical Chemistry 717–718 (2014) 177–188 179

sonication for 15 min and then analyzing the Ni through induc- 0


tively coupled plasma mass spectrometry (ICP–MS) (Thermo_Sci-
entific X series 2). ICP–MS samples were prepared with a 500
dilution. The diluent contains an internal standard suite in 2% nitric

j / mA cm-2
acid. A fully quantitative analysis using a linear calibration curve
based on known standards was performed. The internal standard 0.16 M KCl on HOPG
corrects for instrument drift. NIST traceable certified reference 0.16 M KNO3 on HOPG
materials and serial dilutions were analyzed for accuracy and
duplicate samples were analyzed for precision. Scanning electron 0.16 M KCl on Ni/HOPG

microscopy (SEM) and energy-dispersive X-ray spectroscopy 0.16 M KNO3 on


(EDX) characterization were performed on a JEOL 7401-F at Ni/HOPG
-5
1-5 keV (20 mA) in lower secondary electron image mode with a -1.75 -1.25 -0.75 -0.25
working distance of 8 mm. X-ray diffraction (XRD) measurements E vs SHE / V
were performed on a AXS D8 ADVANCE X-ray diffractometer 0
(Bruker) equipped with a 1-dimensional linear Si strip detector
(LynxEye). The X-ray diffraction samples were initially prepared
by cathodic electrodeposition from 0.08 M Ni(NO3)2 solutions at
1 mA/cm2 for 30 min. The samples were scanned from 5° to 75°

j / mA cm-2
2h in 0.02° steps with a 2 s counting time per step, a 0.499° diver-
gence slit, and a 0.499° antiscatter slit. The X-ray source was Ni-fil- 0.08 M Ni(NO3)2
tered Cu radiation from a sealed tube operated at 40 kV and 40 mA. 0.08 M NiCl2
Goniometer alignment was ensured using an Al2O3 standard 0.08 M Ni(ClO4)2
(Bruker).
The problem with using Pourbaix’s equations for aqueous elec-
trochemical equilibria of metal oxides regards the thermodynamic -5
-1.75 -1.25 -0.75 -0.25
treatment of hydrating H2O in the enthalpy of formation values. E vs SHE / V
Pourbaix assumes that all of the H2O hydrating metal oxides have
entropies equivalent to that of the bulk liquid phase. Pourbaix’s Fig. 1. Linear sweep voltammograms of electrolytes and substrates relevant to the
treatment of H2O intercalated in hydrated metal oxides with liquid Ni(OH)2 electrodeposition mechanism(s). (a) Linear sweeps at 10 mV/s in KCl and
KNO3 solutions on HOPG and HOPG coated with about 20 nm of metallic nickel (Ni/
phase entropies was likely motivated by limited data and an inter-
HOPG). (b) Linear sweeps at 10 mV/s in Ni(NO3)2, NiCl2, and Ni(ClO4)2 electrolytes
est in the uniform treatment of intercalated H2O across all ele- on HOPG.
ments and oxidation states. This treatment causes inaccuracy in
the calculated equilibria because confining water in a solid phase
lowers the H2O’s entropy. For an example at 25 °C, liquid water than the nitrate reduction reaction near neutral pH despite the
(H2O(l)) has a standard entropy, DS0m , of 69.9 J/K/mol while the con- large difference in equilibrium potentials. The deposition of about
fined movement of water in the solid phase (H2O(s)) lowers the DS0m 20 nm of metallic nickel onto the HOPG from a commercial electro-
to 38.0 J/K/mol [36]. A comparison between NiO and NiO(H2O) free deposition solution (Nickel HT-2) lowered the H2 evolution over-
enthalpies of formation indicates that the hydrating water can potential by 0.4 V for the same KCl and KNO3 solutions. A low
have liquid-like DS0m values of 65.5 J/K/mol when associated with current shoulder in the KNO3 voltammogram on the nickel metal
Ni2+ ions and will be designated as H2O(l0 ) [34,37]. It is known that likely corresponds to NO 3 reduction at potentials similar to that
water associated with smaller, more highly charged ions will have of H2 evolution and is consistent the literature for Ni and Pt sub-
even lower entropies [38]. It will be assumed here that the water strates [32,39]. Low current voltammetric shoulders/peaks ap-
associated with Ni3+ has the solid phase entropy of ice and will peared for the Ni(NO3)2, NiCl2, and Ni(ClO4)2 solutions (Fig. 1b)
be designated as H2O(s). The states considered are therefore and likely corresponded with metallic nickel deposition (4) on
NiO(H2O(l0 )), Ni3O4(H2O(s))(H2O(l0 )) and Ni2O3(H2O(s)). The redox HOPG because the shoulders/peaks did not appear when Ni2+
potentials of Table 3 have been adjusted from their referenced val- was absent (Fig. 1a). Some Ni0 state nickel also must have depos-
ues to reflect this more accurate treatment of intercalated H2O. ited at the lower current/potentials because the hydrogen evolu-
tion kinetics which predominate >2 mA/cm2 in the nickelous
solutions (Fig. 1b) were the same for the non-nickelous solutions
3. Results and discussion only when a thin film of Ni0 was coated on the HOPG (Fig. 1a).
Electrochemically charging the Ni(OH)2 deposits formed at
3.1. Ni(OH)2 electrodeposition characterization and optimization 1 mA/cm2 from NiCl2 and Ni(ClO4)2 electrolytes indicated the
presence of metallic nickel, Ni0, in these deposits. Residual Ni0
The cathodic electrodeposition of Ni(OH)2 on HOPG likely in- was considered detected in the NiCl2 and Ni(ClO4)2 deposits be-
volves Ni0 as an intermediate state. HOPG has a large overpotential cause their first charges initiated near 0.4 to 0.6 V vs. SHE in
for evolving H2 from water molecules in KCl and KNO3 (Fig. 1a). Fig. 2 chronopotentiograms, which corresponds with the oxidation
The hydrogen evolution reaction (3) is considered the only plausi- potential of metallic nickel through (5). In the absence of an ap-
ble reduction reaction on HOPG in the KCl solution. The linear plied potential, anhydrous NiO will form through the oxidation
sweep voltammogram of KNO3 on HOPG was approximately the of Ni0 by the water molecule hydrating neighboring Ni(OH)2
same as that of KCl. The hydrogen evolution reaction is expected (which can be alternatively expressed as NiOH2O) through (5) if
to also be the primary reduction reaction on HOPH in the KNO3 the Ni0 is no longer exposed to bulk water. In contrast to the
solution due to the similarities in KCl and KNO3 voltammograms semi-transparent green Ni(OH)2 observed for uncharged deposits
in Fig. 1a. The absence of peaks or shoulders in HOPG’s KNO3 vol- produced from Ni(NO3)2 solutions at <1.25 mA/cm2, significant
tammogram indicates insignificant rates of NO 3 reduction (1) in quantities of the black, anhydrous NiO was visually detected in
comparison with the H2 evolution reaction. The H2 evolution reac- the uncharged NiCl2 and Ni(ClO4)2 deposits produced at compara-
tion experimentally occurs on HOPG at more positive potentials ble current densities [34]. The presence of Ni0 and NiO was also
180 M. Merrill et al. / Journal of Electroanalytical Chemistry 717–718 (2014) 177–188

1 seems to be less consistent for ClO4- than it was for NO 3 [34].
The extra 170 mV offset was therefore likely caused by the differ-
Potential vs SHE / V

ence between deposition onto Ni0 and Ni(OH)2. Alternatively, fac-


0.6
toring in reactant and product concentrations into the
electrochemical equilibria indicates that water will spontaneously
0.2
Ni(NO3)2 oxidize metallic nickel to Ni(OH)2 through reaction (5) even at a
potential bias of 0.7 V vs. SHE if the interfacial Ni2+ concentration
NiCl2
-0.2
is depleted low enough from deposition and interfacial pH reaches
Ni(ClO4)2 pH P 11 due to hydrogen evolution [34]. Bubbles indicative of H2
evolution were observed during deposition from Ni(ClO4)2 and
-0.6 NiCl2 solutions but not from Ni(NO3)2, indicating that the oxidation
0 50 100 150 200
Time / s
of Ni0 by NO 3 is favorable if nitrate is present.
The most active ‘‘NiOOH’’ materials could only be generated at
Fig. 2. Chronopotentiograms for the initial charge of freshly deposited Ni(OH)2. The low concentrations (0.08 M) of Ni(NO3)2 and relatively high cur-
Ni(OH)2 deposits were formed from 0.08 M nickel solutions at 1 mA/cm2 for rent densities, where the most efficient deposition current efficien-
30 min and were initially charged with 8 mA/cm2.
cies and deposit activities occurred around 1 mA/cm2 (Fig. 4). The
quantity of deposited nickel and the corresponding ideal electrode-
position current efficiency was determined by dissolving thor-
detected when deposits from Ni(NO3)2 solutions were generated
oughly-washed deposits in 10% nitric acid and then determining
with current densities >1.25 mA/cm2 (data not shown), which
the Ni2+ concentration with ICP–MS. The ideal current efficiency
would be consistent with a mechanism in which Ni0 is depositing
for the cathodic electrodeposition of Ni(OH)2 was sustainable for
through (4) faster than it can be oxidized by NO 3 through (1).
over 10 min (Fig. 5).
The role of Ni0 in the deposition mechanisms was confirmed by
the similarity of Fig. 3’s chronopotentiograms of Ni(OH)2 deposi-
tion onto HOPG from 0.08 M Ni(NO3)2 and Ni0 deposition onto 3.2. Ni(OH)2 electrodeposition mechanisms
HOPG from a commercial deposition solution, Nickel HT-2. These
chronopotentiograms are characteristically unusual because of Ni(OH)2 electrodeposition mechanisms (i) and (ii) both rely
the exponential decay of the applied potential to lower overpoten- upon the chemical precipitation of Ni(OH)2 due to the formation
tials (more positive potentials) with time. An increase in the over- of a OH concentration gradient through (2). In this commonly pre-
potential of Fig. 3’s chronopotentiograms would be expected due sumed process, OH generated at the interface can diffuse back
to a decrease in Ni2+ concentration with time as the interfacial into the bulk phase much faster than Ni2+ can diffuse to the inter-
reactant diffusion layer grows yet this effect was not observed. face because OH’s diffusion coefficient of 5.27  105 cm2/s is al-
An increase in the overpotential of the chronopotentiograms most an order of magnitude faster than ½Ni2+’s diffusion
would also be expected if the deposition mechanisms primarily re- coefficient of 0.66  105 cm2/s [37]. The application of a constant
lied upon building up a OH concentration gradient according to current electrodeposition technique to this system will cause a
reactions (1) or (3). Significant increases in overpotential would non-ideal electrodeposition current efficiency which will further
have additionally been expected due to self-limiting processes if decrease with time because the OH generated at the interface will
the Ni(OH)2 was behaving like an insulator due to its band gap of diffuse back into the bulk solution phase faster than Ni2+ can dif-
about 2.8 eV (the material becomes more electrically conductive fuse from the bulk phase to the interface. Mechanisms (i) and (ii)
upon oxidation) and only the current collector was electrochemi- are therefore incompatible with the electrodeposition condition
cally active [4,19,31]. Note that the Ni2+ concentration of the Ni of 0.08 M Ni(NO3)2 at 1 mA/cm2 because the electrodeposition
HT-2 solution (0.48 M) appropriately accounts for the 40 mV off- current efficiency remained constant at the ideal 1 Ni(OH)2 per 2
set observed at <300 s for the 0.08 M Ni(ClO4)2 deposition solution, e with respect to the duration of electrodeposition for over
meaning Ni0 was likely the only material being deposited from 10 min (Fig. 5). Mechanisms (i) and (ii) are only incompatible with
Ni(ClO4)2 during the first 300 s. The 0.08 M Ni(ClO4)2 deposition cathodic Ni(OH)2 electrodeposition when the well-known ideal
solution acquired an extra 170 mV offset at >300 s to become sim- electrodeposition current efficiency of 1 Ni(OH)2 per 2 e is
ilar to the Ni(NO3)2 deposition solution’s potential (Fig. 3). Like the
nitrate ion, the perchlorate ion can oxidize metallic nickel even un-
der the large, negative potential bias although this mechanism 0.5
e- discharge / e- deposit

0.4
-0.5
0.3
Potential vs SHE / V

0.2

0.1
-0.7

Ni HT-2 0
-1.5 -1.2 -0.9 -0.6 -0.3 0
0.08 M Ni(NO3)2
Deposition current density / mA cm-2
0.08 M Ni(ClO4)2
-0.9 Fig. 4. Optimum deposition current density. The electrodeposition current density
0 600 1200 1800 for a 0.08 M Ni(NO3)2 solution was varied while the quantity of e (1.44 C) used for
Time / s electrodeposition was held constant. The Ni(OH)2 deposits were charged and then
discharged with 2 mA/cm2. A greater ratio of the number of electrons discharged
Fig. 3. Electrodeposition chronopotentiograms. Metallic nickel and/or Ni(OH)2 from the Ni(OH)2 deposit per the number of electron used to form the deposit
were cathodically deposited onto HOPG at 1 mA/cm2 for 30 min. reflect greater electrodeposition current efficiencies and/or deposit activities.
M. Merrill et al. / Journal of Electroanalytical Chemistry 717–718 (2014) 177–188 181

occurring [40,41]. Mechanisms (i) and (ii) are otherwise likely discharge >1 e per deposited Ni2+ ion of Ni(OH)2 requires the ‘‘NiO-
mechanisms involved when conditions result in non-ideal electro- OH’’ mechanism to either charge nickel to oxidation states greater
deposition current efficiency. than Ni3+ or the occurrence of oxygen redox chemistry if Ni3+ is
Mechanism (iii) is likely only present when the black NiO prod- the highest, stable nickel oxidation state. The discharge occurred
uct is formed at low overpotentials. The oxidation of Ni0 by water in a narrow voltage window of <0.2 V and yielded values of >8000
through (4) is negative of the hydrogen evolution equilibrium po- F/g. The application of this capacitor-appropriate metric to battery
tential of (3) between pH 5–8, depending on the interfacial Ni2+ materials is however misleading because the material can only de-
concentration, and so the coupling of the two reactions can be liver a maximum of 1640 C/g, regardless of the potential range uti-
spontaneous in the absence of a sufficiently negative potential bias. lized. The maximum charge was experimentally limited by oxygen
This coupling of H2 evolution with metal corrosion has been more evolution from water and was similar to the reported value of 1.65
widely recognized in the case of iron [34]. NiO can also be formed e per Ni ion [4]. The ability to discharge 1.58 e per Ni represents a
by depositing Ni0 at relatively negative potentials and then shifting high activity for a material where reports of activities >1 e per Ni
the potential to a more neutral value for oxidation through (4) or are rare and 0.4 e per Ni are more typical [45,46].
(5) with cyclic voltammetry or pulse techniques [42,43]. It is The freshly deposited Ni(OH)2 required exceptionally fast rates
known that Ni(OH)2 can be converted to NiO during electrodepos- of initial charge to achieve the highest activities. Subsequent
ition through electrochemically driven H2 evolution at large over- charging and discharging was performed at more conventional
potentials and current densities (>50 mA/cm2) however this was rates such as 2C. The ability to discharge >1.4 e per Ni could only
likely not the case observed here for the deposits formed with be achieved if the freshly electrodeposited Ni(OH)2 material was
1 mA/cm2 from the 0.08 M NiCl2 and Ni(ClO4)2 solutions [44]. charged under extremely fast rates of >17 C (charged under
Mechanism (iv) is most consistent with the observed results for 200 s) (Fig. 6). The activity of the freshly electrodeposited Ni(OH)2
the Ni(NO3)2 electrolytes when the highest activities are occurring material decreased if the initial charge rate was slower. The freshly
because they correlate with ideal Ni(OH)2 electrodeposition cur- deposited Ni(OH)2 oxidized by >1.7 e per Ni for all initial charging
rent efficiency. The direct coupling of Ni0 oxidation through (4) rates. The greater discharge capacities caused by the higher initial
with the conjugate reduction of NO 3 through (1) to form Ni(OH)2 charging rates indicates that the higher initial charging rates in-
prevents the diffusion of the nitrate reduction’s product OH back creased the reversibility of the charge storage system. Charging
into the solution’s bulk phase to produce ideal electrodeposition the freshly electrodeposited Ni(OH)2 materials at rates >17 C
current efficiencies. The nitrate ion is such a strong oxidizing agent caused overpotentials of 200–300 mV and may have affected the
according to (1) that it could oxidize Ni0 at potential biases more initial oxidation reaction pathways or structure of the Ni(OH)2.
negative than 1.0 V vs. SHE. The characterization and comparison The initial discharge rate also had to be 62 C in order to obtain
of electrolyte compositions in Section 3.1 using ion substitutions the highest activities, where slower discharge rates did not release
and omissions applied in this work indicated that both a metallic more charge. Assuming a molecular weight of 92.7 g/mol and a
Ni intermediate state and NO 3 are involved with achieving ideal density of 4.15 g/cm3, a measured deposit mass of 60.15 mg of
electrodeposition current efficiency for Ni(OH)2 as well as the ab- Ni(OH)2 would correspond with a thickness of 60.22 lm.
sence of Ni and NiO in the final product. Mechanism (iv)’s con-
straint of matching the rate of Ni0 electrodeposition with the rate
of its oxidation by NO 3.4. Raman spectroscopy
3 would also support the dramatic depen-
dence of the ideal electrodeposition current efficiency (or greatest
activity) upon both Ni(NO3)2 concentration and deposition current The features of the (dis)charge cycle illustrated in Fig. 7 will be
density (Fig. 4). Similarly, the rate dependence of upon both reac- supported here with in situ Raman spectroscopy and with addi-
tions (4) and (1) would also explain why neither ideal electrode- tional support from the thermodynamic analysis of Section 3.5. A
position current efficiency nor high activity could be achieved greater resolution in spectral potential-dependence (25 mV) than
using galvanostatic pulse or potentiostatic techniques for Ni(OH)2 previous characterizations enabled the identification and charac-
electrodeposition. terization of less prominent and/or previously unreported spectral
features and behaviors [47,48]. The cathodically deposited Ni(OH)2
was operated in 3 M KOH electrolyte and was initially charged at
3.3. Coulometric analysis
20 C. The spectra illustrated in Fig. 8A–D represent the first
The Ni(OH)2 deposited at 1.0 mA/cm2 became fully charged
when oxidized by 1.84 ± 0.02 e per Ni ion and could subsequently 1.6
deliver 1.58 ± 0.02 e per Ni ion. The ability to both charge and
e- discharged / Ni atom

160 1.4
experimental
Mass Ni in deposit / µg

120 1 Ni per 2 e- 1.2

80
1

40
0.8
0.3 0.7 1.1 1.5
0 Rate of initial charge / log (C rate)
0 0.1 0.2 0.3 0.4 0.5
Deposition charge / C Fig. 6. Ni(OH)2 activity as a function of initial charge rate. The activity of 0.25 mg/
cm2 deposits of Ni(OH)2 produced under ideal electrodeposition current efficiency
Fig. 5. Electrodeposition current efficiency. The mass of Ni deposited as Ni(OH)2 per was characterized as a function of initial charge rate. The rate of initial charge is
coulomb of deposition charge with a 1.0 mA/cm2 deposition current density from described in terms of the log of the C rate, where a 1.0 C rate means the electrode
0.08 M Ni(NO3)2. The slope of 304 lg/C corresponds to an ideal deposition current was charged in 1.0 h. The number of e discharged per Ni atom in the deposit
efficiency of 1 Ni deposited per 2 electrons. represents the material’s activity.
182 M. Merrill et al. / Journal of Electroanalytical Chemistry 717–718 (2014) 177–188

discharge and second charge. The spectra did not change signifi- Table 2
cantly for the subsequent charge cycle. All of the in situ Raman Raman peak assignments.

spectral features have been previously reported except for the Peak shift(s) Species State(s) Ref(s).
peaks in the 800–1500 cm1 range. The Raman peak assignments (cm1)
are in Table 2 and made in accordance with the literature. 445 A1 symmetric Ni–OH Ni(OH)2 [4,10,45,46]
The in situ Raman spectra indicated that a lower potential (0.1– 470–490 Ni(2+)–O stretching b-Ni(OH)2 or NiO(H2O) [45–47]
0.5 V vs. SHE) and a higher potential (0.5–0.65 V vs. SHE) phase 470–490a ? Ni3O4(H2O)2 [47]
470–490b Eg Ni(3+) = O or Ni2O3(H2O) or [10,46,50]
change occur during cycling between charged and discharged Ni(3+)–O bending c-NiOOH
states. The two phase changes have been designated according to 547 – 490c A1g Ni(3+)–O Ni2O3(H2O) or [10,46,50]
their relative potential regions because distinctly different spectral stretching c-NiOOH
changes occur with respect to potential within the two potential 900–1200 Active oxygen O0
1332 –OH Intercalated H2O [46,47]
regions. Fig. 9 presents the potential dependent shifting of the
spectral peaks and Fig. 10 presents the potential dependent change a
With the absence of 547–555 cm1 peak.
b
in relative peak intensity to better distinguish the different spectral With presence of 547–555 cm1 peak.
c
With presence of 470–490 cm1 peak.
changes which occur between the lower and upper potential
ranges. Two phase changes occur for both the charge and the dis-
charge mechanisms because both mechanisms have an intermedi-
to a state referred to as c-NiOOH. A new peak emerges around
ate state (Fig. 7).
547 cm1 and rapidly shifts to 555 cm1 as the 490 cm1 peak
The lower potential phase change during charging was associ-
shifts to 470 cm1 and more rapidly increases in intensity. The
ated with the oxidation of b-Ni(OH)2 to Ni3O4(H2O)2 due to the
similarity of the shifts and relative 470/555 cm1 peak intensities
changes of the 445 cm1 and 490 cm1 peaks upon charging from
are indicative of the c-NiOOH state, which is similar to the Raman
0.250 V to 0.500 V (Figs. 9 and 10) [47,49–51]. Chemically precip-
spectra of discharged LiNiO2 [18,47,51,57,58]. The nomenclature of
itated Ni(OH)2 is typically a more or less ordered b-Ni(OH)2 while
the c-NiOOH state may be somewhat misleading as the lack of any
the more active a-Ni(OH)2 produced from cathodic electrodeposit-
other visible or Raman peaks corresponding to an OH group indi-
ion is typically characterized as the more disordered and hydrated
cate that the charged ‘‘NiOOH’’ is both dehydrated and has approx-
a-Ni(OH)2 [31,32,52–54]. Cathodically deposited a-Ni(OH)2 has
imately no (<0.2 H+ per Ni) protons for the most active battery
been shown to adopt b-Ni(OH)2 character upon charge/discharge
materials [4,5,47,57,59]. Similarly, the remaining 1/3 of the inter-
and this was consistent with our results [5,55]. The shift of the
calated water is lost upon the completion of charge as indicated
490 cm1 peak towards 470 cm1 as the potential surpasses
by the disappearance of the 1332 cm1 peak at the highest poten-
0.450 V while the 540–560 cm1 peak remains absent is indicative
tials. The absence of OH or intercalated H2O species as well as the
of Ni3O4(H2O)2 as the lower potential phase product during charge
similarity with discharged LiNiO2 Raman spectra supports both the
(Fig. 9) [51]. The fully discharged state has spectral characteristics
absence of protons and the presence of activated oxygen, respec-
of both b-Ni(OH)2 and Ni3O4(H2O)2 as expected because coulomet-
tively, in Fig. 7’s higher potential product state.
ric and spectroscopic results from both this work and the literature
The higher potential phase stopped changing with respect to
indicate that the ‘‘NiOOH’’ never fully discharges back to a Ni(OH)2
potential during charging at about 0.600 V (Figs. 9 and 10), which
state [51,55,56]. The lower potential phase change also includes
coincided with the onset of the oxygen evolution reaction. The Ra-
the loss of 2/3 of the 1332 cm1 peak’s intensity up to 0.500 V
man spectra did not change noticeably even at 0.650 V even
and the emergence of the broad, low intensity peak ranging from
though the oxygen evolution reaction was proceeding at about
900 to 1200 cm1. The 1332 cm1 peak has been attributed to
1 mA/cm2. The results suggest that the characteristics of the 470,
intercalated H2O, which can appear in the 1000–1400 cm1 range
555, and/or 900–1200 cm1 peaks correspond with the presence
instead of contaminating NO 3 because of its consistent disappear-
of oxygen atoms, O0 (½O2), in ‘‘NiOOH’’ for the following three rea-
ance/reappearance with charging/discharging (NO 3 redox reac- sons: (1) these peak intensities and positions only emerge above
tions are irreversible in this voltage range) [50,51]. The loss of 2/
the oxygen evolution equilibrium potential of 0.400 V vs. SHE at
3 of the 1332 cm1 peak’s intensity up to 0.500 V indicates that
pH 14, (2) these peak intensities and positions become constant
the less hydrated Ni3O4(H2O) is the lower phase change’s predom-
as oxygen evolution dominates the charging current, and (3) there
inant product instead of the more hydrated Ni3O4(H2O)2 state
are no other features or phase changes which could be correlated
(Fig. 7).
to an oxygen evolution intermediate state. The broad peak of
The upper potential phase change during charging initiated
900–1200 cm1 may correspond to the O0, because it is much more
near 0.500 V and was associated with the oxidation of Ni3O4(H2O)
prominent in c-NiOOH than b-NiOOH and because it is red-shifted
from the intercalated water peak [47]. A more definitive

Table 3
Eqs. (6)–(10): known nickel and oxygen redox reactionsa [32].

Ni3 O4 ðH2 OÞ2 þ 2Hþ þ 2e $ 3NiOðH2 OÞ (6)


E0 ð6Þ ¼ 0:837  0:0592  pH
Ni2 O3 ðH2 OÞ þ 2Hþ þ 2e $ 2NiOðH2 OÞ (7)
E0 ð7Þ ¼ 0:912  0:0592  pH
3Ni2 O3 ðH2 OÞ þ 2Hþ þ 2e $ 2Ni3 O4 ðH2 OÞ2 (8)
E0 ð8Þ ¼ 1:184  0:0592  pH
2NiO2 ðH2 OÞ2 þ 2Hþ þ 2e $ Ni2 O3 ðH2 OÞ þ 4H2 O (9)
E0 ð9Þ ¼ 1:434  0:0592  pH
1=2O2 þ 2Hþ þ e $ H2 O (10)
E0 ð10Þ ¼ 1:288  0:0592  pH
a
See Methods section for details of how these potentials were adjusted from
Fig. 7. The ideal ‘‘NiOOH’’ active cycle. original source.
M. Merrill et al. / Journal of Electroanalytical Chemistry 717–718 (2014) 177–188 183

Intensity / Arb. Units


Intensity / Arb. Units

350 550 750 800 1100 1400


Raman shift / cm-1 Raman shift / cm-1
Intensity / Arb. units

Intensity / Arb. units

350 550 750 800 1100 1400


Raman shift / cm-1 Raman shift / cm-1

Fig. 8. The low (A) and high (B) in situ Raman shifts during charge and the low (C) and high (D) shifts during discharge. The lines indicate the positions for the 445 cm1 (red),
470 M 490 cm1 (blue), 447 M 545 cm1 (green), and 1332 cm1 (orange) peaks. The transitions between lower and higher potential phase changes occurs at approximately
0.500 V during charge and 0.350 V during discharge. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this
article.)

20 1.0

0.8
Relative intensity
Shift in peak / cm-1

10
0.6

0.4
0
0.2

-10 0.0
0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0 0.1 0.2 0.3 0.4 0.5 0.6
Potential vs SHE / V
Potential vs SHE / V
Fig. 10. The change in relative Raman peak intensity with respect to potential.
Fig. 9. The change in Raman peak cm1 with respect to potential.

assignment of this peak as the active oxygen species is difficult be-


being oxidized and not the permanently anionic O2 responsible
cause the 900–1200 cm1 range is commonly ignored in in situ Ra-
for structurally holding the nickel cations together. The oxidation
man studies of relevant metal oxides. Thermodynamic evidence
of only hydrating/intercalated H2O and not the H2O of bulk water
presented in the next section indicates that only the O2 belonging
is also consistent with quartz crystal microbalance analysis [60].
specifically to the hydrating/intercalated H2O at 1330 cm1 are
184 M. Merrill et al. / Journal of Electroanalytical Chemistry 717–718 (2014) 177–188

The ‘‘NiOOH’’ discharge mechanism was different than the Table 4


charge mechanism. The complete emergence of the intercalated Eqs. (11)–(18): relevant charge and discharge mechanisms.

H2O (1332 cm1) peak and the complete disappearance of the Ni3 O4 O0 ðH2 Oðl0 Þ Þ þ 4Hþ þ 4e $ 3NiOðH2 Oðl0 Þ Þ (11)
broad 900–1200 cm1 peak directly coincided with the decrease E0 ð11Þ ¼ 2:125  0:1183  pH
of the c-NiOOH (470 and 555, cm1) peaks during the discharge’s NiOðH2 Oðl0 Þ Þ þ Ni2 O3 ðH2 OðsÞ Þ $ Ni3 O4 ðH2 OðsÞ ÞðH2 Oðl0 Þ Þ (12)
higher potential phase down to 0.350 V (Fig. 10). The shifting of E0 ð12Þ ¼ 0:347
the 470 and 555 cm1 peaks back to their original discharge posi- Ni2 O3 O0 þ 4Hþ þ 4e $ 2NiOðH2 Oðl0 Þ Þ (13)

tions of 490 and 547 cm1, respectively, plus the reappearance of E0 ð13Þ ¼ 2:200  0:1183  pH
3Ni2 O3 O0 þ 4Hþ þ 4e $ 2Ni3 O4 O0 ðH2 Oðl0 Þ Þ (14)
the intercalated water peak at 1332 cm1 indicates the formation
of NiO(H2O) at potentials 60.450 V [47,50]. This indicates that E0 ð14Þ ¼ 2:472  0:1183  pH
2NiOO0 $ Ni2 O3 O0 (15)
the Ni2O3 state in combination with the active oxygen character
0 (16)
of the broad 900–1200 cm1 peak was being maintained through- NiOO þ 2Hþ þ 2e $ NiOðH2 OðsÞ Þ
E0 ð16Þ ¼ 1:288  0:0592  pH
out the charge delivery potentials and discharged to the NiO(H2O)
Ni3 O3 F2 ðH2 OðsÞ ÞðH2 Oðl0 Þ Þ þ H2 O þ 2e $ 3NiOðH2 Oðl0 Þ Þ þ 2F (17)
state instead of Ni(OH)2. The 445 cm1 peak indicating the proton-
Ni3 O4 O0 FðH2 Oðl0 Þ Þ þ 4Hþ þ 3e þ F $ Ni3 O3 F2 ðH2 OðsÞ ÞðH2 Oðl0 Þ Þ (18)
ated Ni–OH species of Ni(OH)2 did not reappear during discharge
until 0.150 V. The 470–490 cm1 peak was always accompanied
by the 547–555 cm1 peak during discharge, which indicates that
a hydrated Ni3O4 state did not form during discharge. Charging 0.01 M KOH
through a state with Ni3O4 character and discharging to a state 0.75 0.03 M KOH
with NiO(H2O) character is consistent with the well-known hyster- 0.1 M KOH
0.3 M KOH
esis of the ‘‘NiOOH’’ and underscores the importance of reporting

E vs SHE / V
0.65 1 M KOH
whether potential-dependent in situ spectra have been obtained 3 M KOH
during charge or discharge. 0.55

3.5. Thermodynamic analysis


0.45

An active oxygen mechanism of ‘‘NiOOH’’ consistent with spec-


0.35
troscopic and voltammetric characterizations (Fig. 7) can be con- 0 1 2 3 4 5 6
structed from the thermodynamics of known nickel and oxygen time / ks
redox reactions. The original source of the electrochemical poten-
tials of reaction (6)–(10) in Table 3 considered the hydrated nickel Fig. 11. ‘‘NiOOH’’ (dis)charge chronopotentiograms. The potential dependence of
oxide’s intercalated H2O to have the entropy of bulk phase liquid the 2nd charge/discharge curves with respect to KOH concentration for 0.25 mg/
cm2 loadings deposited from 0.08 M Ni(NO3)2 at 1 mA/cm2. The markers indicate
water. The nickel and oxygen redox potentials in Table 3 have been
the inflection points evaluated in Fig. 12.
adjusted because the water intercalated in the nickel oxides do not
have the entropy of bulk phase liquid water. Water associated with
Ni2+ ions have been designated as H2O(l0 ) because they have an en- 1.0
tropy similar to that of liquid water and the water associated with
Ni3+ ions have been designated as H2O(s) because they have the en-
E vs SHE / V

tropy of ice (see Methods section). The ‘‘NiOOH’’ reaction mecha- 0.8
nisms of Table 4 have been constructed using the potentials of
Table 3 and will be supported by experimentally derived thermo-
dynamic values.
0.6
The thermodynamics of ‘‘NiOOH’’ active material produced un-
der ideal electrodeposition current efficiency were galvanostatical-
ly (dis)charged with respect to variable electrolyte KOH
concentrations (Fig. 11). The pH and potential dependence of the 0.4
11.5 12.5 13.5 14.5
(dis)charge curve inflection points are illustrated in Fig. 12. The pH
(dis)charge curve inflection points should correspond to rational
equilibria in an approximately ideal system and correlate with Fig. 12. Correlation of experimental ‘‘NiOOH’’ thermodynamic values with calcu-
the pH and potential dependence of the mechanisms in Table 4. lated equilibria. The experimental equilibrium values derived from inflection points
of Fig. 11’s chronopotentiograms are compared with the equilibria calculated for
The slight deviations of the experimentally determined inflection
mechanisms described in Table 4.
points in the chronopotentiometric curves from the calculated
equilibria at the highest KOH concentrations (Fig. 12) was likely
due to the significant deviation of the solvent water’s activity (or
osmotic) coefficient from the ideal conditions assumed in the ther- clear from thermodynamics alone why an active oxygen mecha-
modynamic calculations [34,61]. Note in Fig. 12 that the 120 mV/ nism would occur instead of the more energetically favorable oxi-
pH dependence of the charge curve inflection points and the dation of just the nickel ions yet equilibrium thermodynamics
60 mV/pH dependence of the discharge curve inflection points alone yield no insight into the active material’s coordination chem-
at steady state conditions further indicated that the hysteresis oc- istry. It is apparent from in situ X-ray adsorption experiments that
curs due to a permanent thermodynamic effect and that the charge the coordination of the nickel ions remains constant and only dis-
and discharge mechanisms were not simply the reverse of each tortion in the (a)symmetry occurs during (dis)charge [6–10,54,63].
other [62]. The differences between the experimentally determined inflection
The oxidation states derived from the thermodynamic analysis points and the calculated equilibria in Fig. 12 indicate that the acti-
corroborate the interpretation of the in situ Raman experiments. vation energies for the active oxygen mechanism are small. The
The inflection points of the galvanostatic (dis)charge curves match activation energies for changing the material’s coordination chem-
both the potential and pH dependence of mechanism (11). It is not istry to accommodate only nickel redox chemistry must therefore
M. Merrill et al. / Journal of Electroanalytical Chemistry 717–718 (2014) 177–188 185

be relatively large for the observed selectivity for the active oxygen The potential-dependent selectivity of viable nickel redox reac-
mechanism to occur. Thermodynamics indicate that O0 is stabi- tions during discharge was also supported by coulometric analysis.
lized in the H2O(s) position so that the product is Ni3O4O0(H2O(l0 )) The 7–9% charge loss due to the inability to completely charge the
(Fig. 7). The preference for specifically the Ni3O4O0(H2O(l0 )) state ‘‘NiOOH’’ because of oxygen evolution plus a 4 – 5% charge loss due
near 0.500 V in 3 M KOH was consistent with the Raman spectra to instability with cycling causes the ‘‘NiOOH’’ to only be able to
in that the Ni3O4 state appears during charge, a second water-cor- discharge by 1.58 ± 0.02 e per Ni ion to an average Ni valence of
related peak never actually appears for H2O(s), and that 1/3 of the +2.26 ± 0.02. The discharge potentials and inflection points
water species present in the discharged state notably persisted matched that expected for the discharge of NiOO0 to NiO(H2O(s))
up to 0.500 V (Fig. 10). through (16) (Fig. 12). This means that both the thermodynamics
The ‘‘NiOOH’’ oxidizes from the Ni3O4O0(H2O(l0 )) state towards and coulometrics indicate that Ni2O3O0 relaxes into 2 NiOO0
the Ni2O3O0 state while charging through the upper potential according to (15) before discharging to 2 NiO(H2O(s)) as H+ are in-
phase change (Fig. 7). The potential during charge surpasses serted into the local environment. Note that only the conversion of
the disproportionation reaction of (12) so that Ni3O4(H2O(s))(H2- Ni2O3O0 to 2 NiO(H2O(s)) was detected in the in situ Raman exper-
O(l0 )) can spontaneously transition into separate NiO(H2O(l0 )) and iments while only the conversion of NiOO0 to NiO(H2O(s)) was de-
Ni2O3(H2O(s)) states. The degeneracy of (12) provides a means tected through thermodynamics and so the latter conversion must
for the Ni2+ ion in Ni3O4O0(H2O(l0 )) to reach the Ni2O3O0 state have proceeded immediately after the former conversion (Fig. 7).
as the latter becomes stable through (13). The ‘‘NiOOH’’ can The proposed discharge mechanism is in agreement with other
therefore oxidize through (14) at the potential of (13) in a solid cation insertion electrode materials such as LiNiO2 as expected due
state model. Oxygen atoms, O0, seem to be able to combine to to the similarities of between H+ and Li+ (de-)intercalation. Struc-
form molecular oxygen, O2(g), in the Ni2O3O0 state and so the turally, the biggest difference between LiNiO2 and ‘‘NiOOH’’, which
experimental charging curve inflection point caused by oxygen can be alternatively written as HNiO2, is that LiNO2 is convention-
evolution (Fig. 12) corresponds with the potential of mechanism ally well-ordered in Li-ion electrode materials while HNiO2 is con-
(13) as well as the lack of further potential-dependent evolution ventionally highly defective structurally in electrochemically
in the Raman spectra above 0.6 V vs. SHE (Figs. 8A–D). The pro- deposited ‘‘NiOOH’’ materials. Both experimental and modeling
posed fully charged state, which is composed of an equilibrium techniques report that at defect sites of LiNiO2, it is thermodynam-
between the Ni3O4O0(H2O(l0 )) and Ni2O3O0 states due to the con- ically favorable for oxygen species to oxidize so that Ni3+ can re-
straints of O2(g) evolution, is approximately absent of protonated duce to Ni2+ as the Ni2+ moves into Li+ ion vacancies [15,16,18].
species and in accordance with Corrigan’s and our inability to This discharge mechanism is consistent with the report that hys-
detect OH interactions with Ni ions in the fully charged condi- teresis is thermodynamically governed [62]. The experimental re-
tion. The fully charged state is also in accordance with a pro- sults here indicate that the ‘‘NiOOH’’ material was so defective
posed mechanism in which optimal O2(g) evolution catalytic that 95–96% of the Ni3+ moved into H+ vacancies as Ni2+ to form
efficiency occurs when the M3O4 and M2O3 oxidation states NiOO0 during discharge and would explain why the residual Ni3+
interact in predominantly dehydrated first row transition metal does not discharge to lower oxidation states until the evolution
oxide compounds [44]. The intermediate state of Ni3O4O0(H2O(l0 )) of aging effects. The degree of structural and bonding changes
has released 1.67 e per Ni ion, the ideal charged state of Ni2O3- which must be occurring at defects for the active oxygen mecha-
O0 has released 2.00 e per Ni ion, and the equilibrium (average) nism unfortunately may indicate that instability in aqueous elec-
condition between the two states corresponded to the experi- trolytes may be inherent for such disordered and highly active
mental limit of oxygen evolution where the ‘‘NiOOH’’ was oxi- materials.
dized by 1.84 ± 0.02 e per Ni ion.
The proposed charging process indicates that the fully charged 3.6. Fluoride partial inhibition
condition was composed of predominantly Ni2O3O0 plus a fraction
of Ni3+ whose O0 is absent as a result of O2 evolution catalysis. The The H2O(l0 ) and H2O(s) was experimentally distinguishable be-
Ni3+ without O0 constitutes about 20–22% of the nickel in the fully cause of the exchange of H2O(l0 ) with the aqueous solvent. Water’s
charged condition and was inert during the discharge process be- self-diffusion is exceptionally fast because it involves the dissocia-
cause the ‘‘NiOOH’’ finished discharging before approaching the tion into H+ + OH in an intermediate state [64,65]. This solvent ex-
potential of (7). While the discharge potential of Ni2O3 to Ni3O4 change mechanism can be inhibited in a condition when an F ion
was similar to the experimental discharge asymptote according (133 pm) can kinetically and thermodynamically outcompete OH
to the conventional potential of (8), the adjustment of the redox (133 pm). Solution phase calculations supported by experimental
potential of (8) to account for the lower entropies of the hydrating results indicated that the electronegative F is competitive with
H2O mean that the ‘‘NiOOH’’ is fully discharged at potentials 50 mV OH for Ni2+ in a 1 M KF:0.01 M KOH solution [66]. The inhibition
more positive than the unadjusted value. The absence of Ni3O4 Ra- by F caused oxidation by 0.67 e per Ni ion to the Ni3O3F2(H2-
man spectroscopic character during discharge means that the dis- O(s))(H2O(l0 )) according to (17) and the coulometric charge indi-
charge of the Ni3+ unassociated with O0 to Ni2.67+ cannot be a cated in Fig. 13. Elemental analysis with EDX after the first
predominant mechanism if it is active at all. Similarly, ‘‘NiOOH’’ charge step at the point marked by h in Fig. 13 indicated that there
finished discharging at potentials 0.2–0.3 V positive of the unad- was 0.59 of the 0.67 F per Ni predicted in the product of (17). The
justed equilibria of (6) and (7) and 0.3–0.4 V positive of the ad- subsequent oxidation of Ni3O3F2(H2O(s))(H2O(l0 )) to Ni3O4O0F(H2-
justed equilibria of (6) and (7) (see Fig. 11) and so these nickel O(l0 )) by another e per Ni ion through (18) was limited by oxygen
reduction reactions by themselves could not have been significant evolution, where 0.31 of the expected 0.33 F per Ni measured at
pathways during discharge. This means that all of the Ni3+ unasso- the point marked by d in Fig. 13. The potential-dependence dem-
ciated with the active oxygen species in the fully charged state did onstrated by the effects of KF in the electrolyte suggests that the
not discharge to lower oxidation states during discharge and re- number of hydrating H2O which can exchange with the solution
mained as Ni3+ in the fully discharged state. This may indicate that phase could change with activity, cycling, or exchange with KOH
changes in coordination chemistry have occurred for the inert Ni3+. and may explain the potential shifts which occur with aging [3].
Charge and discharge through other reactions, which could include The potential shifts correlated with aging could alternatively be
active peroxides instead of oxygen, may have developed after sig- due to the oxygen redox chemistry switching to a peroxide product
nificant deactivation of the ‘‘NiOOH’’ through repetitive cycling. state instead of the O0 state.
186 M. Merrill et al. / Journal of Electroanalytical Chemistry 717–718 (2014) 177–188

1.2 3 M KF + 0.001 M KOH


1 M KOH
E vs SCE / V

0.8

0.6

0.4
0 1 2 3 4
time / ks

Fig. 13. Partial inhibition of ‘‘NiOOH’’ with F. 0.25 mg/cm2 loadings deposited
from 0.08 M Ni(NO3)2 at 1 mA/cm2 were galvanostatically characterized in both
1 M KOH and 3 M KF + 0.001 M KOH to determine if F ions could selectively inhibit
the active oxygen mechanism of H2O(l0 ) but not H2O(s). The symbols h and d Fig. 14B. SEM image of Ni(OH)2 deposited with pulsed current. The Ni(OH)2 was
indicate where the ‘‘NiOOH’’ charged in KF was sampled for elemental analysis. deposited from 0.08 M Ni(NO3)2 at 1 mA/cm2 in 0.3 s pulses and 2.1 s pauses.

100

charge retention / %
80

60

40

20

0
0 30 60 90 120
cycle #

Fig. 15. Cycle stability of ‘‘NiOOH’’ at optimum activity. A 0.25 mg/cm2 deposit of
‘‘NiOOH’’ at optimum activity was prepared by electrodeposition at ideal current
Fig. 14A. SEM image of Ni(OH)2 deposited under ideal electrodeposition current efficiency followed by an initial charge at 20 C in 1 M KOH. The ‘‘NiOOH’’ was then
efficiency. The cathodic deposition of Ni(OH)2 from 0.08 M Ni(NO3)2 at 1 mA/cm2 cycled between 0.4 and 0.6 V vs. SCE with a 2 C rate.
would have produced a 0.22 lm thick film if the deposit was uniformly flat and
fully dense at 4.15 g/cm3.

3.7. Prospects for practical implementation the charge/discharge mechanism, it is not yet clear whether highly
active ‘‘NiOOH’’ materials can have competitive stabilities or if the
The morphology of the Ni(OH)2 deposited onto HOPG was not deactivation by changes in the hydrated states are unavoidable.
ideal for practical implementation. The narrow range of optimum
electrodeposition current densities (Fig. 4) could be difficult to 4. Conclusion
extrapolate from flat to porous electrodes. The 300 mC charge
passed during deposition at 1 mA/cm2 was ideally expected to An optimal combination of Ni(NO3)2 concentration and current
form a uniform film 0.22 lm thick, assuming the density of density resulted in an ideal cathodic electrodeposition current effi-
4.15 g/cm3, because the electrodeposition current efficiency was ciency and the highest deposit activities. The cathodic electrode-
100%. The Ni(OH)2 instead formed clumps with heights of about position of nickel hydroxide involved the deposition of Ni2+ into
1.5 lm (Fig. 14A). The morphology of Fig. 14A suggests that the the metallic (Ni0) state. The Ni0 is oxidized back to Ni2+ by the
density of the Ni(OH)2 may be approximately half of the expected reduction of NO 
3 to NO2 and the coupling of these reactions and
4.15 g/cm3 due to porosity. A low density and high surface area de- their products causes Ni(OH)2 precipitation with an ideal deposi-
posit from cathodic electrodeposition was not unexpected due to tion current efficiency of 1 Ni2+ per 2 e. Ni0 can alternatively cat-
the resulting material’s proficiency for supercapacitor applications alyze NO 
3 reduction or hydrogen evolution to generate the OH
[67–70]. Switching to a 0.3 s pulse at 1.0 mA/cm2 and 2.1 s product gradient which can chemically precipitate Ni(OH)2 from
pauses resulted in a more evenly distributed film with a signifi- aqueous Ni2+ under conditions of non-ideal electrodeposition cur-
cantly lower electrodeposition current efficiency and/or activity rent efficiencies.
(Fig. 14B). Highly active ‘‘NiOOH’’ deposits could be initially oxidized to
The highly active nickel material was unstable throughout re- store up to 1.84 e per Ni ion and could subsequently deliver a
peated charge and discharge cycles (Fig. 15). Charge storage capa- charge of 1.58 e per Ni ion at optimal performance. The approxi-
bility degenerated with a 5% loss with each successive charge or mately ideal charge capacity of the ‘‘NiOOH’’ facilitated a thermo-
discharge. The film thickness, charge and discharge rates, depth of dynamic analysis through which the specific charge and discharge
discharge, overcharge, electrolyte KOH concentration, or the addi- mechanisms could be resolved from the enthalpies of formation for
tion of LiOH to the electrolyte did not improve the stability signif- known states. Significant changes in the galvanostatic charge and
icantly enough (data not shown) to be useful in practical discharge curves occurred as the ‘‘NiOOH’’ aged with cycling. A
rechargeable battery applications. Despite the new insight into quantitative thermodynamic or in situ Raman spectroscopic
M. Merrill et al. / Journal of Electroanalytical Chemistry 717–718 (2014) 177–188 187

analysis of the aged material to determine whether active oxygen [7] N.R.S. Farley, S.J. Gurman, A.R. Hillman, Electrochim. Acta 46 (2001) 3119–
3127.
species switched from O0 to O1 should be the subject of future
[8] A.N. Mansour, R.A. Brizzolara, C.A. Melendres, M. Pankuch, J. Electrochem. Soc.
work. Poor electrodeposition constraints, morphology and stability 141 (6) (1994) L69–L72 (United States).
with cycling currently prevent practical application of this highly [9] W.E. Ogrady, K.I. Pandya, K.E. Swider, D.A. Corrigan, J. Electrochem. Soc. 143
active material in secondary batteries. (1996) 1613–1616.
[10] Y.N. Hu, I.T. Bae, Y.B. Mo, M.R. Antonio, D.A. Scherson, Can. J. Chem. – Rev. Can.
The results of both the thermodynamic analysis and the in situ De Chim. 75 (1997) 1721–1729.
Raman spectroscopy characterization of the ‘‘NiOOH’’ charge and [11] X. Qian, H. Sambe, D.E. Ramaker, K.I. Pandya, W.E. O’Grady, J. Phys. Chem. B
discharge pathways at optimal performance were consistent with 101 (1997) 9441–9446.
[12] R. Gottschall, R. Schöllhorn, M. Muhler, N. Jansen, D. Walcher, P. Gütlich, Inorg.
an active oxygen mechanism. Successfully modeling the thermo- Chem. 37 (1998) 1513–1518.
dynamics of the ‘‘NiOOH’’ (dis)charge cycle represents a distinct [13] H. Sambe, T.M. Nabi, D.E. Ramaker, A.N. Mansour, W.E. O’grady, The Oxidation
achievement which cannot be obtained through other proposed State of Ni in the Nickel Oxide Electrode and Related Nickel Oxide Compounds:
I. Spectroscopic Evidence, Defense Technical Information Center, 1997.
mechanisms (see Supplementary Information). Because it is now [14] H. Sambe, T.M. Nabi, D.E. Ramaker, The Oxidation State of Ni in the Nickel
known that (1) bond lengths alone cannot distinguish between Oxide Electrode and Related Nickel Oxide Compounds: II. Geometric Evidence,
the Ni4+ state vs. the Ni3+ state plus activated oxygen, and (2) that Defense Technical Information Center, 1997.
[15] A. Rougier, C. Delmas, A.V. Chadwick, Solid State Commun. 94 (1995) 123–127.
the BaNiO3 and KNiIO6 X-ray adsorption references actually corre- [16] M.K. Aydinol, A.F. Kohan, G. Ceder, K. Cho, J. Joannopoulos, Phys. Rev. B 56
spond to the Ni3+ state plus activated oxygen, there is no longer (1997) 1354–1365.
any significant evidence substantiating the Ni4+ state in the ‘‘NiO- [17] J. van Elp, H. Eskes, P. Kuiper, G.A. Sawatzky, Phys. Rev. B 45 (1992) 1612–
1622.
OH’’ mechanism. Constructing the ‘‘NiOOH’’ mechanism from
[18] P. Kalyani, N. Kalaiselvi, Sci. Technol. Adv. Mater. 6 (2005) 689–703.
known redox reactions required adjustment of the nickel oxide [19] M.C. Biesinger, B.P. Payne, L.W.M. Lau, A. Gerson, R.S.C. Smart, Surf. Interface
reduction potentials because Pourbaix considered all hydrating Anal. 41 (2009) 324–332.
H2O species to have an entropy equivalent to the liquid phase, [20] A.P. Grosvenor, M.C. Biesinger, R.S.C. Smart, N.S. McIntyre, Surf. Sci. 600 (2006)
1771–1779.
which is not an accurate assumption. Modifying the entropies of [21] I.G. Casella, M.R. Guascito, M.G. Sannazzaro, J. Electroanal. Chem. 462 (1999)
H2O(l0 ) associated with Ni2+ to a known value slightly smaller than 202–210.
that of the liquid phase and the entropy of the H2O(s) associated [22] T. Dickinson, A.F. Povey, P.M.A. Sherwood, J. Chem. Soc., Faraday Trans. 1: Phys.
Chem. Condens. Phases 73 (1977) 327–343.
with Ni3+ to an ice-like entropy resulted in an excellent correlation [23] Y. Koyama, T. Mizoguchi, H. Ikeno, I. Tanaka, J. Phys. Chem. B 109 (2005)
with the experimentally derived equilibria. The selectivity for only 10749–10755.
H2O(s)’s in the oxygen redox activity was consistent with the [24] Y. Koyama, Y.-S. Kim, Isao Tanaka, H. Adachi, Jpn. J. Appl. Phys. 4 (38) (1999)
2024.
Li(1X)NiO2 system in which oxygen electron vacancies only occur [25] Y. Uchimoto, H. Sawada, T. Yao, J. Power Sources 97–98 (2001) 326–327.
in the presence of Ni3+ upon charging. The ‘‘NiOOH’’ discharge [26] J. Graetz, C.C. Ahn, H. Ouyang, P. Rez, B. Fultz, Phys. Rev. B 69 (2004) 235103.
mechanism through the formation of oxygen electron vacancies [27] P. Kuiper, G. Kruizinga, J. Ghijsen, G.A. Sawatzky, H. Verweij, Phys. Rev. Lett. 62
(1989) 221–224.
associated with Ni2+ was similarly consistent with the discharge [28] F. Reynaud, D. Mertz, F. Celestini, J.M. Debierre, A.M. Ghorayeb, P. Simon, A.
of a Li(1X)NiO2 system into a LiXNi(1X)O system at defect sites. To- Stepanov, J. Voiron, C. Delmas, Phys. Rev. Lett. 86 (2001) 3638–3641.
gether, these (dis)charge mechanisms are also consistent with a [29] A.W. Moses, H.G.G. Flores, J.-G. Kim, M.A. Langell, Appl. Surf. Sci. 253 (2007)
4782–4791.
thermodynamically governed hysteresis as well as proton interca-
[30] H. Chen, C.L. Freeman, J.H. Harding, Phys. Rev. B 84 (2011) 085108.
lation. The elucidation of these relatively subtle mechanistic [31] M. Wohlfahrt-Mehrens, R. Oesten, P. Wilde, R.A. Huggins, Solid State Ionics
behaviors was facilitated in part by generating Raman spectra with 86–88, Part 2 (1996) 841–847.
a greater potential-dependent resolution to observe a secondary, [32] R.S. Jayashree, P.V. Kamath, J. Power Sources 93 (2001) 273–278.
[33] C.C. Streinz, A.P. Hartman, S. Motupally, J.W. Weidner, J. Electrochem. Soc. 142
less prominent phase change and distinguishing between the (1995) 1084–1089.
charge and discharge processes. [34] M. Pourbaix, Atlas of Electrochemical Equilibria in Aqueous Solutions, National
Association of Corrosion Engineers, 1974.
[35] C.D. Taylor, S.A. Wasileski, J.-S. Filhol, M. Neurock, Phys. Rev. B 73 (2006)
Acknowledgements 165402.
[36] P. Atkins, J.D. Paula, Physical Chemistry, W. H. Freeman and Co., 2009.
[37] W.M. Haynes, CRC Handbook of Chemistry and Physics: A Ready-Reference
This work was performed under the auspices of the U.S. Depart- Book of Chemical and Physical Data, CRC Press, Boca Raton, FL, 2011.
ment of Energy by Lawrence Livermore National Laboratory under [38] Y. Marcus, Chem. Rev. 109 (2009) 1346–1370.
[39] K. Bouzek, M. Paidar, A. Sadílková, H. Bergmann, J. Appl. Electrochem. 31
Contract DE-AC52-07NA27344. Project 11-LW-037. The authors
(2001) 1185–1193.
would also like to acknowledge Rachel E. Lindvall’s contribution [40] Zhao, Zhao, Li, Chem. Mater. 19 (2007) 3882–3891.
for the ICP–MS analysis and Jonathan Lee’s discussions regarding [41] T. Subbaiah, S.C. Mallick, K.G. Mishra, K. Sanjay, R.P. Das, J. Power Sources 112
X-ray absorption spectroscopic techniques. (2002) 562–569.
[42] A.C. Sonavane, A.I. Inamdar, P.S. Shinde, H.P. Deshmukh, R.S. Patil, P.S. Patil, J.
Alloy. Compd. 489 (2010) 667–673.
[43] K.-W. Nam, E.-S. Lee, J.-H. Kim, Y.-H. Lee, K.-B. Kim, J. Electrochem. Soc. 152
Appendix A. Supplementary material (2005) A2123–A2129.
[44] M.D. Merrill, R.C. Dougherty, J. Phys. Chem. C 112 (2008) 3655–3666.
Supplementary data associated with this article can be found, in [45] H.G. Zhang, X.D. Yu, P.V. Braun, Nat. Nanotechnol. 6 (2011) 277–281.
[46] H. Wang, H.S. Casalongue, Y. Liang, H. Dai, J. Am. Chem. Soc. 132 (2010) 7472–
the online version, at http://dx.doi.org/10.1016/j.jelechem.20 7477.
14.01.022. [47] Y.L. Lo, B.J. Hwang, Langmuir 14 (1998) 944–950.
[48] D. Gosztola, M.J. Weaver, Langmuir 5 (1989) 776–782.
[49] S. Deabate, F. Fourgeot, F. Henn, J. Power Sources 87 (2000) 125–136.
References [50] C.A. Melendres, S. Xu, J. Electrochem. Soc. 131 (1984) 2239–2243.
[51] C. Johnston, P.R. Graves, Appl. Spectrosc. 44 (1990) 105–115.
[52] P. Oliva, J. Leonardi, J.F. Laurent, C. Delmas, J.J. Braconnier, M. Figlarz, F. Fievet,
[1] J. Phillips, S. Mohanta, M. Geng, J. Barton, B. McKinney, J. Wu, ECS Trans. 16
A.D. Guibert, J. Power Sources 8 (1982) 229–255.
(2009) 11–17.
[53] G. Nazri, D.A. Corrigan, S.P. Maheswari, Langmuir 5 (1989) 17–22.
[2] T. Reddy, D. Linden, Linden’s Handbook of Batteries, McGraw-Hill, 2010.
[54] K.I. Pandya, W.E. O’Grady, D.A. Corrigan, J. McBreen, R.W. Hoffman, J. Phys.
[3] A.H. Zimmerman, Nickel–Hydrogen Batteries: Principles and Practice,
Chem. 94 (1990) 21–26.
Aerospace Press, 2010.
[55] R. Kostecki, F. McLarnon, J. Electrochem. Soc. 144 (1997) 485–493.
[4] D.A. Corrigan, S.L. Knight, J. Electrochem. Soc. 136 (1989) 613–619.
[56] B.C. Cornilsen, P.J. Karjala, P.L. Loyselle, J. Power Sources 22 (1988) 351–
[5] J. Desilvestro, D.A. Corrigan, M.J. Weaver, J. Electrochem. Soc. 135 (1988) 885–
357.
892.
[57] S. Deabate, F. Fourgeot, F. Henn, Electrochim. Acta 51 (2006) 5430–5437.
[6] T.W. Capehart, D.A. Corrigan, R.S. Conell, K.I. Pandya, R.W. Hoffman, Appl. Phys.
[58] B.S. Yeo, A.T. Bell, J. Phys. Chem. C 116 (2012) 8394–8400.
Lett. 58 (1991) 865–867.
188 M. Merrill et al. / Journal of Electroanalytical Chemistry 717–718 (2014) 177–188

[59] M.K. Carpenter, D.A. Corrigan, J. Electrochem. Soc. 136 (1989) 1022–1026. [66] D.o.L.a.W.R. Engineering, Visual MINTEQ 2.54, in: K.R.I.o. Technology (Ed)
[60] G.T. Cheek, W.E. O’Grady, J. Electroanal. Chem. 421 (1997) 173–177. <http://www.lwr.kth.se/English/OurSoftware/vminteq/>, Stockholme, 2010.
[61] W.J. Hamer, Y.-C. Wu, J. Phys. Chem. Ref. Data 1 (1972) 1047–1100. [67] E.E. Kalu, T.T. Nwoga, V. Srinivasan, J.W. Weidner, J. Power Sources 92 (2001)
[62] K.P. Ta, J. Newman, J. Electrochem. Soc. 146 (1999) 2769–2779. 163–167.
[63] J. McBreen, W.E. O’Grady, G. Tourillon, E. Dartyge, A. Fontaine, K.I. Pandya, J. [68] D.-D. Zhao, S.-J. Bao, W.-J. Zhou, H.-L. Li, Electrochem. Commun. 9 (2007) 869–874.
Phys. Chem. 93 (1989) 6308–6311. [69] K.-W. Nam, K.-B. Kim, J. Electrochem. Soc. 149 (2002) A346–A354.
[64] M. Eigen, Discuss. Faraday Soc. 17 (1954). [70] X. Zhang, W. Shi, J. Zhu, W. Zhao, J. Ma, S. Mhaisalkar, T. Maria, Y. Yang, H.
[65] T. Dippel, K.D. Kreuer, Solid State Ionics 46 (1991) 3–9. Zhang, H. Hng, Q. Yan, Nano Res. 3 (2010) 643–652.

You might also like