You are on page 1of 15

REVIEWS

Achieving high energy density and high


power density with pseudocapacitive
materials
Christopher Choi 1,4, David S. Ashby 1,2,4, Danielle M. Butts 1,4
, Ryan H. DeBlock 1
,
Qiulong Wei 1, Jonathan Lau 1,3 and Bruce Dunn 1*
Abstract | Batteries and supercapacitors serve as the basis for electrochemical energy-​storage
devices. Although both rely on electrochemical processes, their charge-​storage mechanisms
are dissimilar, giving rise to different energy and power densities. Pseudocapacitive materials
store charge through battery-​like redox reactions but at fast rates comparable to those of
electrochemical double-​layer capacitors; these materials, therefore, offer a pathway for
achieving both high energy and high power densities. Materials that combine these properties
are in demand for the realization of fast-​charging electrochemical energy-​storage devices
capable of delivering high power for long periods of time. In this Review, we describe the
fundamental electrochemical properties of pseudocapacitive materials, with emphasis on
kinetic processes and distinctions between battery and pseudocapacitive materials. In addition,
we discuss the various types of pseudocapacitive materials, highlighting the differences
between intrinsic and extrinsic materials; assess device applications; and consider the future
prospects for the field.

Electrochemical energy-​storage (EES) technologies power power density of a supercapacitor. There is a clear need
the portable, electronic devices that are an indispensable for such materials to enable fast charging and to deliver
part of our daily lives. All evidence indicates that the high power for long periods of time to enhance the capa-
growth of EES technologies will continue almost unabated bilities of mobile power sources. Pseudocapacitance pro-
in the near future, driven by expanding markets for mobile vides an opportunity to achieve high energy density at
consumer electronics, the electrification of transportation high power density, and an increasing number of studies
and the need for wireless communication1,2. In addition, show that pseudocapacitive materials can fulfil this goal.
substantial growth for stationary EES is projected, owing Pseudocapacitive materials exhibit battery-​like redox
1
Department of Materials
to the increasing integration of renewable technologies reactions that occur at rates comparable to those of
Science and Engineering, into the energy grid3,4. Although batteries are the fore- double-​layer formation in a capacitive material7 (Box 1).
University of California, most devices used for EES, electrochemical capacitors, As a consequence, pseudocapacitive materials gener-
Los Angeles, Los Angeles, known as supercapacitors, generally have charge-​storage ally charge much faster than battery materials, on the
CA, USA.
properties that complement those of batteries, and their timescale of a few minutes. Although pseudocapaci-
2
Sandia National Laboratories,
usage has increased substantially over the past 10 years5. tive behaviour in materials such as hydrated RuO2 and
Livermore, CA, USA.
The energy-​storage processes within the two types of MnO2 has been known since the 1970s (Box 2), the field
3
Energy Storage and
Distributed Resources
devices derive from fundamentally different mechanisms, has received renewed interest, owing to the increased
Division, Energy Technologies leading to the different charge-​storage properties. Battery demand for EES and the greater understanding of the
Area, Lawrence Berkeley materials store large amounts of energy (~200 Wh kg–1) electrochemical properties of nanoscale materials.
National Laboratory, through diffusion-​limited redox reactions, which results Consequently, the diversity of pseudocapacitive mate-
Berkeley, CA, USA.
in slow charging (on the order of hours)3. By contrast, rials has expanded to include perovskite, non-​oxide,
4
These authors contributed capacitive materials store much smaller amounts of polycationic and 2D materials.
equally: Christopher Choi,
David S. Ashby, Danielle
energy (~5 Wh kg–1) very rapidly (on the order of seconds) This Review begins with a discussion of the kinetic
M. Butts through the formation of electrical double layers6. signatures and charge-​storage mechanisms of pseudo-
*e-​mail: bdunn@ucla.edu This Review addresses the question of whether there capacitance, with a focus on establishing the distinc-
https://doi.org/10.1038/ are energy-​storage materials that can simultaneously tions between pseudocapacitive and battery materials.
s41578-019-0142-z achieve the high energy density of a battery and the high The characterization techniques that can be used for

Nature Reviews | Materials


Reviews

Box 1 | electrochemical charge-​storage mechanisms In this section, we discuss the fundamental aspects of
the electrochemical, chemical and structural signa-
a Li-​ion battery material stores charge through diffusion-​limited, faradaic reactions tures of pseudocapacitive materials and the various
throughout the bulk of the active material. Diffusion-​limited redox is often slow characterization techniques used in their analysis.
and, although the redox reactions lead to high energy density, the electrochemical
performance of these devices is characterized by low power density and charging
Electrochemical signatures
requires tens of minutes to hours. supercapacitor devices, also known as electrical
double-​layer capacitors (eDLCs), do not involve faradaic charge storage and only The charge-​storage mechanisms of pseudocapacitive
store charge through surface-​controlled ion adsorption (see panel a of the figure). materials are based on battery-​like redox reactions,
this capacitive charge-​storage mechanism leads to high power and rapid charging which occur at rates comparable to that of electrical
on the order of minutes. a pseudocapacitive material exhibits a second mechanism double-​layer charge storage in capacitive materials, and
for capacitive charge storage. these materials undergo surface-​controlled, faradaic display an electrochemical response similar to that of a
reactions, which, because they are confined to the surface (or near the surface), capacitor. The capacitance, C (F g–1), can be defined as
occur at rates comparable to those of double-​layer formation in eDLCs (see panels b a function of the potential, V (ref.6):
and c of the figure). thus, pseudocapacitive materials differ from eDLCs because
they undergo faradaic redox reactions. However, pseudocapacitive materials also Q  nF  X
differ from traditional Li-​ion battery materials because the kinetics of the redox C[F g −1] = =  (1)
reactions are dramatically faster and are not limited by semi-​infinite diffusion.
V  m  V
within pseudocapacitive materials, there are two main mechanisms of charge
storage: surface redox pseudocapacitance and intercalation pseudocapacitance. where X is the extent of fractional coverage on the sur-
with surface redox materials, the charge-​storage mechanism is attributed to charge face or inner surface of the active material, m is the
transfer occurring at or near the surface of the material (see panel b of the figure, molecular weight of the active material, n is the number
where O and r are the oxidized and reduced forms of a redox couple, respectively), of electrons and F is the Faraday constant. For a pseudo­
where ions are electrochemically adsorbed, creating short diffusion distances and, capacitive material, the stored charge, Q, is expected
subsequently, short diffusion times. By contrast, intercalation pseudocapacitance (see to increase linearly with an increase in potential. Thus,
panel c of the figure) involves the rapid diffusion of ions in and out of ion-​conduction a nearly linear dependence of charge storage within
channels or layers within a host material, without the material undergoing a phase the potential window is a defining feature of a pseudo­
change. these two mechanisms may occur in intrinsic pseudocapacitive materials or
capacitive material and indicates that the faradaic
can be achieved by engineering materials to elicit a pseudocapacitive response and
are, thus, termed extrinsic pseudocapacitive materials. charge storage is not limited by diffusion (Fig. 1a). The
energy stored in a pseudocapacitive material is there-
a EDLC b Redox c Intercalation fore given by E = ½CV2 = ½QV (ref.8), with E continu-
pseudocapacitance pseudocapacitance ously increasing as charge is added incrementally with
– – + O + ze ↔ R
– a changing potential. By contrast, a battery material dis-
– – Li+ Li+ Li+
+ + + plays a defined voltage plateau (Fig. 1a) and, therefore,
– –
– Li+ Li+ Li+ the majority of the charge is stored at a constant poten-
– + tial. The energy stored in a battery material is given by
+ + + + Li+ Li+ Li+
– E = QV when incremental charge is added at a constant

– – + Li+ Li+ potential7. Characterization of the charge-​storage prop-
+ – –

erties of a pseudocapacitive material should establish
Carbon Anion Cation Redox-active material Redox host material the possible processes (for example, surface adsorp-
Figure adapted from ref.55, courtesy of Kristy Jost, 3M, Minnesota, usa. tion or ion intercalation) that occur over the operating
potential range. For this reason, it is more appropriate
the analysis and identification of pseudocapacitive to describe pseudocapacitive materials using gravimet-
materials are also discussed. Although the fundamen- ric capacity (C g–1 or mAh g–1), rather than gravimetric
tals of pseudocapacitive behaviour have been well stud- capacitance (F g –1), which can lead to higher than
ied, understanding continues to evolve as the range of theoretical values9.
pseudocapacitive materials expands. The second part Experimentally, the cyclic voltammetry (CV) res­
of this Review illustrates the current inventory of pseudo­ ponse for a pseudocapacitive material resembles that
capacitive materials. In particular, we highlight the dif- of a capacitive material. Pseudocapacitive voltammo-
ferences between intrinsic pseudocapacitive materials, grams can be quasi-​rectangular (as for hydrated RuO2,
which exhibit inherent capacitor-​like charge storage, and top-​left part of Fig. 1b) but will often have broad fara-
extrinsic pseudocapacitive materials, which need to be daic charge-​transfer peaks superimposed over a ‘box-​
nanostructured to exhibit pseudocapacitive behaviour. like’ profile10 (as for orthorhombic Nb2O5 (T-​Nb2O5)
In addition, we briefly consider the nascent integration or extrinsic MoS2, top-​right and bottom-​left parts of
of pseudocapacitive materials into EES devices and the Fig. 1b, respectively). The separation between the broad
direction of this research. Finally, we provide our per- anodic and cathodic peaks is small, indicating that there
spective on the role of pseudocapacitive materials in the is only a small amount of polarization in the charge-​
future energy landscape. storage process and that the charged and discharged
states have a similar free energy. The parameters that
Pseudocapacitive charge storage lead to broad faradaic peaks and quasi-rectangular
Pseudocapacitive materials exhibit a combination shapes in CV have been investigated computationally
of characteristics that distinguish them from battery using numerical methods and multiphysics models11.
materials. The mechanisms that underpin pseudo­ It has been suggested that peak broadening is due to
capacitive processes in materials are discussed in Box 1. concentration-​dependent activity coefficients and the

www.nature.com/natrevmats
Reviews

interactions between reactant molecules according to (Box 1). Electrochemical methods, such as CV and EIS,
Eq. 2 (ref.12): have been widely used in an attempt to deconvolute the
kinetic responses and identify charge-​storage mecha­
nisms in materials. In a CV experiment, the current
 RT   2Γ 
o
E * = E app +  (a R + a O − 2a RO)1 − R  (2) response i upon varying the sweep rate (υ) is dependent
 F   Γm  on the charge-​storage process. Specifically, a power-​
law relationship can be used to correlate the current
where E* is the standard potential for an ideal couple; aR, dependence17:
aO and aRO are interaction coefficients (where R and O
indicate the reduced and oxidized forms of a redox couple, i(υ) = aυ b (3)
respectively); ΓR is the surface concentration of the redu­
ced species; Γm is the maximum surface concentration; where a is a constant and b is the power-​law exponent.
o
and E app is the apparent standard potential, which The b value can be obtained from the slope of the plot of
o RT
is given by E app = E o + ( F )(a O − a R ). Peak broadening log(i) versus log(υ) for the cathodic and anodic peaks.
occurs when the interaction parameter (aR + aO – 2aRO) is On the basis of the experimental b values, the kinetics
positive12. In addition to voltammetry, electrochemical of the charge-​storage mechanism can be determined
impedance spectroscopy (EIS) provides a specific pseudo­ qualitatively.
capacitive electrochemical signature. Namely, 3D Bode For a redox reaction limited by semi-​infinite diffu-
plots of frequency, potential and either real/imaginary sion, the relationship between the peak current, ip, and
impedance (Zʹ/Zʺ), real/imaginary capacitance (Cʹ/Cʺ) or the sweep rate follows the Randles–Sevcik equation18,19:
phase angle (ϕ) as the third axis can distinguish between
1∕2
capacitive, pseudocapacitive or battery-​like materi-  αnF 
i p = 0.4463nFAcD 1∕2υ1∕2 (4)
als (Fig. 1c–e). 3D Bode representations offer a means  RT 
of deconvoluting the faradaic contribution to the total
capacitance on top of double-​layer capacitance across where c is the surface concentration of the electrode
the operating potential window and calculating the material, α is the transfer coefficient, D is the chemical
corresponding relaxation time constants13–16. diffusion coefficient, A is the surface area of the elec-
trode material, R is the molar gas constant and T is the
Analysis of electrochemical kinetics temperature. The redox peak current for a diffusion-​
The kinetics of charge storage provide insight into the controlled process is, thus, proportional to the square
charge-​storage mechanisms as well as information root of the sweep rate, hence b = 0.5. This value indicates
about device operation. In general, EES processes can a faradaic charge-​storage process, as occurs in battery
be divided into diffusion-​controlled (that is, battery-​like) materials. By contrast, for a non-​faradaic, capacitive
and surface-​controlled (that is, capacitor-​like) responses charge-​storage mechanism, the peak current has a linear
relationship with the scan rate υ (ref.7), resulting in b = 1:

Box 2 | ruO2 as a pseudocapacitive material i p = υCA (5)


the idea of pseudocapacitive charge storage was addressed as early as 1962 by
Conway and Gileadi to explain the underpotential deposition of hydrogen adatoms on Although the charge-​storage mechanisms in pseudo­
noble-​metal surfaces174. the prototypical pseudocapacitive charge-​storage material, capacitive materials are faradaic in nature, these mate-
namely ruO2, was shown to be “a new interesting electrode material” in a preliminary rials also display b = 1, owing to their capacitor-​like
report in 1971 (ref.63), with a follow-​up study in 1974 (ref.64); subsequently, ruO2 (surface-​c ontrolled) electrochemical response. On
has become the benchmark for an ultrafast, high-​capacity charge-​storage material. the basis of the different current dependencies, capac-
although redox reactions were clearly observed, the exact nature was elusive itive processes, rather than faradaic processes, dom-
until trasatti and Buzzanca accurately theorized that ruO2 is a hydrous oxide that inate the peak current response at high sweep rates,
undergoes protonation in mildly acidic aqueous solutions63,66. the mechanism of owing to the linear relationship. Experimentally, the
electron charge transfer in hydrated ruO2 (see the figure) is mediated by facile
well-​known battery material LiFePO4 has a b value of
transport of H+ through the hydrous boundaries of ruO2 as the ru valence changes
from ru4+ to ru3+ to ru2+. without abundant hydrous boundaries throughout 0.5, whereas orthorhombic Nb2O5, a pseudocapacitive
the nanoporous ruO2 network, proton diffusion becomes sluggish and redox is material, shows a b value of 1.0 over a wide range of
restricted to the outermost surface, resulting in low capacities. Hydrated ruO2 sweep rates20,21. For several other battery materials, b
exhibits electrochemical signatures values between 0.5 and 1.0 have been observed, which
similar to those of a capacitor, such Aqueous electrolyte (e.g. H2SO4) suggests a mixture of diffusion-​controlled and capacitor-​
as multicycle repeatability and like responses22–24. For example, cyclic voltammograms
box-​like voltammograms, despite Ru 2+
↔ Ru3+ ↔ Ru4+ Hydrous
boundary of nanostructured (extrinsic) MoS2 (bottom-​left part of
the faradaic charge-​storage H +
Fig. 1b) exhibit broad redox peaks superimposed on a
mechanism, making ruO2·xH2O box-​like profile, which suggests that nanostructured
a unique redox pseudocapacitive
e– MoS2 is pseudocapacitive. In addition, the b values for
material65,67. although there were
efforts in the mid-1970s to early 1980s these redox peaks are ~1 (ref.25). By contrast, the cyclic
to commercialize ruO2 capacitors, voltammograms of bulk MoS2 present well-​defined
the prohibitively high cost of ru has redox peaks (bottom-​right part of Fig. 1b). Although
largely restricted the use of these Nanoporous RuO2·xH2O Eq. 4 was derived for redox molecules in solution with
devices to military applications7. a fixed, single-​redox peak potential, the extension to

Nature Reviews | Materials


Reviews

a 1.0 b Redox pseudocapacitance Intercalation pseudocapacitance


(RuO2·xH2O) (T-Nb2O5)
0.9 4
30
3

Current (mA cm–2)


0.8 20

Current (A g–1)
2
Normalized potential (V/V0)

0.7 10 1
0
0
0.6 –1
–10 –2
0.5 –3
–20
–4
0.4 –30 –5
0.0 0.2 0.4 0.6 0.8 1.0 0.0 1.2 1.5 1.8 2.1 2.4 2.7 3.0
0.3 Potential (V vs SCE) Potential (V vs Li/Li+)
0.2
Extrinsic pseudocapacitance
0.1 (nano-MoS2) Battery material (bulk MoS2)
0.6 0.6
0.0
0.4 0.4

Current (mA)

Current (mA)
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
Normalized capacity (Q/Q0) 0.2 0.2
0.0 0.0
Battery material (bulk LCO)
Extrinsic pseudocapacitance (nano-LCO) –0.2 –0.2
Redox pseudocapacitance (RuO2·xH2O)
–0.4 –0.4
Intercalation pseudocapacitance (T-Nb2O5) 0.0 1.0 1.5 2.0 2.5 3.0 0.0 1.0 1.5 2.0 2.5 3.0
Potential (V vs Li/Li+) Potential (V vs Li/Li+)

c Capacitive d Pseudocapacitive e Battery-like


2.5 1.4 1.4
1.2 0.01 Hz 1.2
2.0 0.01 Hz
1.0 1.0
C′ (F cm–2)
C′ (F cm–2)

C′ (F cm–2)
1.5
0.8 0.8 0.03 Hz
1.0 0.6 0.03 Hz 0.6
0.4 0.4
0.5 0.1 Hz
0.01 Hz 0.2
0.03 Hz 0.1 Hz 0.2
0.0 0.1 Hz 0.0 0.0
0.01 0.01 1 Hz 0.01 1 Hz
Fr

1 Hz Fre Fre
eq

0.1 q ue 0.1 1.8 2.0 q ue 0.1 1.8 2.0


ue

1.2 nc 1.4 1.6 nc 1.4 1.6


nc

0.8 1.0 y (H Zn )
1 2+ )
1.0 1.2 y (H
2+
0.6 1
0.4 n / 1 1.0 1.2 s /Zn
n
y(

0.2 s Z
l (V v Z
Potential (V vs
Ag/AgCl) z) ia z ) ial (V v
Hz

Potent Potent
)

Fig. 1 | representative characteristics of electrochemical energy-​storage materials. a | Illustrative galvanostatic


discharge curves for various kinds of pseudocapacitive materials (surface redox, intercalation pseudocapacitance
and a nanostructured extrinsic pseudocapacitive material) and, for comparison, a battery material, bulk LiCoO2 (LCO).
The pseudocapacitive materials have capacitor-​like electrochemical features, showing linear potential versus capacity
responses during galvanostatic discharging, whereas the battery material shows a potential plateau. Data from refs26,70,108.
b | Cyclic voltammograms for various pseudocapacitive materials: surface redox, RuO2 (500 mV s–1 at a mass loading of
1 mg cm–2)68; intercalation pseudocapacitance, Nb2O5 (10 mV s–1, 1 M LiClO4 in propylene carbonate, 40 µg cm–2)95; and
an extrinsic pseudocapacitive material, nano-​MoS2 (1 mV s–1, 1 M LiPF6 in ethylene carbonate:dimethyl carbonate 1:1,
5.8 mg cm–2)25; as well as a battery material, bulk MoS2 (1 mV s–1, 1 M LiPF6 ethylene carbonate:dimethyl carbonate 1:1,
3.4 mg cm–2)25, highlighting the variation in peak broadening. c–e | Examples of 3D Bode plots showing the normalized
capacitance (C′) versus frequency and potential for capacitive (carbon nanofoam (CNF)15; panel c), pseudocapacitive
(MnOx@CNF in a NaSO4 electrolyte; panel d) and battery (MnOx@CNF in a 3:1 NaSO4:ZnSO4 electrolyte; panel e)
materials16. MnOx@CNF electrodes can display either battery-​like (Zn-​ion insertion) or capacitor-​like (pseudocapacitive
from Na-​ion insertion) character, depending on the cations in the electrolyte. The capacitive and pseudocapacitive
materials exhibit near-​constant capacitance values across the potential range, whereas the battery material shows peak-​
like features near the redox potential at low frequency. SCE, standard calomel electrode. Panel a curve for bulk LCO is
adapted with permission from ref.108, ACS. Panel b top-​left part is adapted with permission from ref.68, ACS. Panel b
top-​right part is adapted with permission from ref.95, Wiley-​VCH. Panel b bottom parts are adapted with permission from
ref.25, ACS. Panel c is adapted with permission from ref.15, Elsevier. Panels d and e are adapted with permission from ref.16, RSC.

solid-​phase pseudocapacitive systems, in which both way, the variation in b value can be determined across
faradaic and non-​faradaic phenomena occur over a the entire operating potential range. Thus, it is possi-
range of potentials due to peak broadening, suggests ble to ascertain whether materials, pseudocapacitive
that b values can also be obtained at potentials other or battery-​like, show a transition between capacitor-​
than that at which the peak current is observed. In this like and diffusion-​controlled processes, depending on

www.nature.com/natrevmats
Reviews

potential26. For pseudocapacitive materials at potentials current iC decays exponentially in response to a potential
far from the redox regime, it is also possible to observe step of magnitude, ΔV (ref.34):
b = 1, which is representative of electrical double-​layer
processes. For example, multiphysics modelling studies of ΔV −t ∕ R sC
iC = e (8)
Nb2O5 have shown that b = 1 above 2.4 V (versus Li/Li+) Rs
and arises from non-​faradaic electrical double-​layer
formation11. Indeed, this analysis shows that, although where Rs is the series resistance. A few studies using
the charge-​storage mechanism in Nb2O5 is surface con- activated carbon and MnO2 have demonstrated practi-
trolled throughout the potential range, there is a change cal applications of the SPECS model35,36. The MUSCA
in the dominant charge-​storage mechanism between method has been used to examine the energy-​storage
pseudocapacitive faradaic (b = 1) and electrical double-​ mechanism of Ti3C2Tx (where T is the surface termi-
layer capacitive (b = 1). The transition between regimes nation functional group, such as OH, O or F) MXene
is indicated by an abrupt dip in the b value2,11, which materials30. With the MUSCA technique, the influence
arises from Li-​ion depletion at the electrode–electrolyte of polarization from the ohmic drop is minimized, as the
interface11,26. mean current during the titration experiment is calcu-
In a related analysis approach, the current response at lated by integrating transient currents at each incremental
a fixed potential is expressed as the combination of two potential step. The voltammogram is then reconstructed
separate contributions: surface capacitive and diffusion using the mean current to deconvolute the surface and
controlled27,28: bulk diffusion processes with a minimal ohmic contribu-
tion. This analysis will be particularly useful for material
i(υ) = k 1υ + k 2υ1∕2 (6) systems with limited electronic conductivities, for which
the ohmic drop can be substantial.
where k1 and k2 are the proportionality constants that Another electrochemical signature that might pro-
describe the surface-​controlled and diffusion-​controlled vide insight into charge-​storage mechanisms is the rate
processes, respectively. This method is qualitative, giving of reaction. In terms of electrochemical redox reaction
an indication only as to the fraction of the current car- kinetics, pseudocapacitive charge storage is an electro-
ried by the two processes. Moreover, this approach is not chemically reversible or quasi-​reversible process (that
applicable if there are other charge-​storage mechanisms is, has fast reaction rates), owing to a small separation
(for example, conversion reactions). In order to more between the anodic and cathodic peaks, stemming from
accurately evaluate the surface-​controlled component, the minimal polarization in the charge-​storage process.
this analysis is often carried out using slow sweep rates The effect of reaction kinetics on the pseudocapacitive
(<1 mV s–1) to enhance the contribution from the slower charge-​storage mechanisms for Nb2O5 has been studied
diffusion-​controlled process. computationally using numerical methods37. Results
Electrochemical analytical methods based on the from this study suggest that the rate constant for hetero­
potentiostatic intermittent titration technique (PITT) geneous electron transfer, ko, has to be sufficiently
have recently been developed to elucidate the behav- high to maintain a faradaic process and to avoid the
iour of pseudocapacitive materials. These analytical non-​faradaic electrical double-​layer process from dom-
methods include step potential electrochemical spec- inating, especially at a high scan rate (that is, at 1 V s–1).
troscopy (SPECS)29 and multiple-​step chronoampero­ When the electrochemical storage capabilities of Nb2O5
metry (MUSCA)30. PITT was initially developed for are not limited by the rate of charge transfer, ko had a
determining the diffusion coefficients of intercalation minimal effect on the charge-​storage process.
species and to identify whether ion diffusion is confined A well-​known kinetic analysis based on Nicholson’s
by finite space or is governed by semi-​infinite diffusion, parameter (Ψ)38 has also been applied to experimentally
as in classical electrochemical cells31,32. In PITT experi- determine the reaction kinetics of faradaic charge trans-
ments, small potential pulses of specified increments are fer for pseudocapacitive materials39. Nicholson’s param-
applied, each followed by a relaxation period intended eter was initially derived for redox molecules in solution
to allow the system to reach pseudo-​equilibrium at each in the quasi-​reversible regime to qualitatively describe
given state of charge. The pertinent diffusion properties how a system evolves from displaying a reversible to
can then be obtained from current transients monitored an irreversible reaction. This parameter can be deter-
during the incremental titration. mined by correlating changes in peak-​to-peak separation
The SPECS and MUSCA analytical methods enable with respect to varying CV sweep rates; ko can then be
further interpretation of PITT data but have only been calculated using Eq. 9 (ref.38):
demonstrated for a few pseudocapacitive materials thus far.
With the SPECS method, the transient current response 1∕2
iD of the diffusion-​controlled process is defined by the  RT  −1∕2
Ψ = ko  υ (9)
Cottrell equation for semi-​infinite planar diffusion33:  πnFD 

 D 1∕2 A higher ko value indicates that the system is able to


iD = nFAc  (7)
 πt  reach equilibrium more rapidly through faster redox
reactions. In a study of the extrinsic pseudocapacitive
where t is the time elapsed at the applied potential. By behaviour of MoS2, the ko value of nanoscale MoS2 on
contrast, for the capacitive-​like process, the transient reduced graphene oxide (rGO) was roughly two orders

Nature Reviews | Materials


Reviews

of magnitude greater than that of bulk MoS2 (ref.39). This requirement can be satisfied by having either a very
Although there remain questions regarding the appli- short diffusion length or a large diffusion coefficient.
cation of Nicholson’s parameter to solid-​state systems, It is difficult to establish the materials that have a large
it may provide a qualitative method for determining mobile-​ion (typically Li+) diffusion coefficient, owing
the difference in the rate of reaction between bulk and to the dependence of this value on the measurement
nanostructured materials in extrinsic pseudocapacitive method and the mobile-​ion concentration. By contrast,
systems. there has been considerable success in creating nanoscale
materials that exhibit pseudocapacitive properties. One
Chemical and structural signatures reason for this success is that, when the electrode film
The ability to track changes in the valence state of a is very thin — that is, less than the characteristic diffu-
metal ion can be useful in characterizing pseudocapaci- sion length — diffusion of the species involved in the
tive materials. Monitoring changes in the oxidation state redox reactions is governed not by conventional electro­
of the materials using X-​ray techniques, such as X-​ray chemistry but by thin-​layer electrochemistry44,45. In
absorption near-​edge structure40 and X-​ray photo­electron thin-​layer electrochemistry, the faradaic reactions occur
spectroscopy41, combined with the electrochemical within a finite diffusion space within which the concen-
kinetic analyses described above further enables decon- tration gradients that occur in the bulk electrode are neg-
volution of the charge-​storage mechanisms. In pseudo- ligible and the redox kinetics begin to resemble capacitive
capacitive materials, the total capacitance (Ctotal) has a processes (b = 1). This behaviour is shown by the propor-
non-​faradaic electrical double-​layer contribution (CNF) tional relation between i and υ in the following equation,
and/or a faradaic pseudocapacitive contribution (CF). which considers no mass transport limitations34:
The faradaic pseudocapacitive contribution is directly
proportional to the change in oxidation state, [Ox], and
the theoretical capacity per valence state change, ξ:
 2 2 1
 n F υVAcD 2
i = 



( )
exp RT (E − E 0′)

nF

(11)
{ }
 2
 RT
C total = C NF + C F = C NF + ξ[Ox] (10) ( )
 nF 0′ 
1 + exp RT (E − E )
 

Recently, this approach was used to determine the where E is the potential of an electrode and E0′ is the
capacitive charge-​storage contributions from fara- formal potential of an electrode.
daic and non-​faradaic processes in bronze-​phase TiO2 Early studies on finite diffusion primarily focused
(TiO2(B)) nanotubes42. On the basis of the quantitative on the mechanism of redox processes in conducting
charge-​storage analysis using in situ, dynamic valence-​ polymer films46–48. The electrochemical features of thin
state monitoring, the decrease in the specific capacity conducting polymers resemble those of capacitive redox
with increasing sweep rate was attributed to a decrease systems, with both demonstrating changes in oxidation
in battery-​like, diffusion-​limited contributions. Thus, in state with increasing electrode potential and reversible
a material with multiple charge-​storage processes, such charge-​transfer redox processes. Consequently, these
as TiO2(B), charge-​storage contributions from diffusion-​ polymers exhibit nearly mirror-​image voltammograms
limited, pseudocapacitive and non-​faradaic reactions can and a linear relationship between the current and the
be precisely evaluated using the valence-​state monitor- scan rate. Electron propagation through a polymer film
ing technique in conjunction with the aforementioned occurs between immobile redox centres during the
kinetic analyses. faradaic process. Furthermore, the sweep-​rate depend-
A structural feature that seems to correlate with the ence of the peak current has been explained in terms
electrochemical characteristics discussed above is that of 1D diffusion and surface-​confined redox processes46.
faradaic charge transfer occurs without a phase tran- Thus, with an increase in film thickness, mass transport
sition in pseudocapacitive materials. For example, the within the polymer film becomes limiting and the cyclic
pseudocapacitive response of orthorhombic Nb2O5 voltammetric response deviates from a mirror image
arises from fast 2D Li-​ion transport through empty to the classic, asymmetrical shape, with the current
octahedral sites between (001) planes. This structural proportional to the square root of the sweep rate49.
arrangement provides tunnels in the a–b crystallo- The reaction kinetics of thin-​film electrodes have
graphic plane to accommodate the ions; thus, the Li+ been evaluated using different electrochemical tech-
insertion–deinsertion process does not cause phase tran- niques, such as chronopotentiometry39, cyclic voltam-
sitions in LixNb2O5 when x ≤ 2 (refs10,43). When structural metry49,50, staircase voltammetry51 and square-​wave
changes need to be considered, X-​ray diffraction (XRD) voltammetry52–54. These studies show that, in thin-​film
should be used to confirm that a phase change does not electrochemistry, the mass-​t ransport dynamics are
occur upon varying the Li+ content. If no changes in determined primarily by a few key kinetic parameters.
the lattice parameters are observed, redox reactions are For example, a finite diffusion parameter, w, was pro-
likely confined to the surface of the material. posed to describe the reaction kinetics of thin-​film elec-
trode materials and is dependent on variables such as
Thin-​film electrochemistry the electrode thickness, the diffusion coefficient and the
The charge-​storage kinetics in pseudocapacitive materi- potential sweep rate49:
als are not characterized by semi-​infinite diffusion; that
is, the ion diffusion length l and the diffusion coefficient nFl 2υ
w= (12)
D are related by l « (Dtd)1/2, where td is the diffusion time. RTD

www.nature.com/natrevmats
Reviews

Depending on the value of the finite diffusion more than one transition-​metal ion, an area of increas-
parameter, the voltammetric response of the reaction ing interest60. More extensive lists of pseudocapacitive
is characteristic of a surface redox process (w < 0.2), materials have been reported elsewhere59,61,62.
an electrode reaction controlled by restricted diffusion
(0.2 < w < 10) or a faradaic process occurring under Intrinsic pseudocapacitive materials
semi-​infinite diffusion (w > 10). Thus, it is evident that RuO2·xH2O. RuO2 is considered the archetypal pseu-
thin-​film electrochemistry, as described by surface-​ docapacitive material, owing to its well-​understood
confined redox processes and finite diffusion, will charge-​storage mechanism, historical significance and
produce a pseudocapacitive electrochemical response. high theoretical capacity7,63–70 (Box 2). In early investi-
However, the interpretation of such a response at the gations, the redox reaction was found to be localized
thin-​film length scale is not straightforward. It is diffi- on the surface and, thus, pseudocapacitive charge stor-
cult to discern whether the finite diffusion response in age was thought to be dependent on surface area71,72.
a thin-​film electrochemical measurement comes from Further analysis revealed the importance of the elec-
the intrinsic or extrinsic material properties because trolyte and the surface chemistry, as only hydrated
both factors contribute to the emergence of pseudo­ RuO2 (RuO2·0.5H2O, which has a theoretical capacity of
capacitance in materials as the electrode dimensions 375 mAh g–1 at 1 V) displayed high-​rate pseudocapaci-
reach the thin-​film regime. tance73,74 (Table 1). Anhydrous RuO2 does demonstrate a
The occurrence of extrinsic pseudocapacitive behav- pseudocapacitive response but the capacitance and rate
iour in materials is influenced significantly by thin-​film capability are severely limited3. The discrepancy between
electrochemistry55. Nanostructuring materials can sup- the hydrated and anhydrous forms is due to the hydrated
press crystallographic phase transitions by decreasing state acting as a proton lubricator and, therefore, facil-
intercalation stress, leading to fast kinetics that are not itating fast proton conduction along the surface and
limited by the reaction rates of phase transitions. This through grain boundaries3. This storage of protons in
behaviour has been observed experimentally in sev- hydrated RuO2 can be expressed as:
eral materials, including MoO2 (ref.56), LiMn2O4 (ref.57)
and MoS2 (ref.25), which exhibit battery-​like behaviour RuOx (OH)z + δ H + + δ e − ↔ RuOx − δ (OH)z + δ
in bulk form. The fast kinetics and pseudocapacitive-​
like responses of these materials arise primarily from Density functional theory analysis of the charge-​
thin-​layer electrochemistry, as the condition l « (Dtd)1/2 storage mechanism supports that proton migration
is satisfied for nanoscale dimensions. The experi­mental through the bulk is unfavourable (with an energy
phenomenon of extrinsic pseudocapacitive nano­scale barrier of 1.62 eV for proton intercalation into bulk
materials has also been observed using numerical RuO2), confirming that redox is confined to the sur-
model­ling simulations58. Thus, nanostructuring of the face of RuO 2 (ref. 75) . The rutile phase of RuO 2 is a
electrode materials to the thin-​layer electrochemistry metallic conductor, enabling fast electronic transport
regime is one important consideration for obtaining through the bulk, while chemisorption or physisorp-
extrinsic pseudocapacitive behaviour. tion of water on the RuO2 surface is responsible for
facile proton insertion. Thus, a hydrated RuO2 material
Classifying pseudocapacitive materials can accommodate rapid charge transfer at the inter-
For many years, RuO2 and MnO2 dominated the field of face between the RuO2 nanocrystal and the boundary
pseudocapacitive materials. Over the past 10 years, the water76. Moreover, there is a key relationship between
field has broadened considerably following the discovery the electrolyte and available capacitance. Protons have
that several battery materials exhibit pseudo­capacitive been shown to be the dominant charge-​carrying ions
responses when nanostructured or engineered into a over hydroxides. Similarly, a direct relationship has been
nanoscale form. To account for this ability to modify shown between cation concentration in the electrolyte
charge-​storage behaviour, it has been proposed that mate- and capacitance77. In many respects, the electrochemical
rials should be classified as either intrinsic or extrinsic, signatures of RuO2 provide a model system for pseudo-
depending on whether pseudocapacitance is inherent to capacitive materials in aqueous electrolytes. The high
the material in its bulk form or arises upon nanostructur- electronic conductivity (10 4 S cm–1 in single-​crystal
ing, respectively59. This distinction is used in the following RuO 2) 7 coupled with the rapid proton migration
as we review model intrinsic pseudocapacitive materials in hydrated systems provides a basis for understanding
and the design of extrinsic pseudocapacitive materials. pseudocapacitive materials when the redox processes
We also discuss the emergence of pseudocapacitive are confined to the surface.
non-​oxide systems.
Selected properties of model intrinsic and extrinsic MnO2. Since the discovery in 1999 that amorphous
pseudocapacitive materials are provided for reference in MnO2 displays a faradaic, rectangular voltammogram
Table 1 and compared in Fig. 2. Table 1 highlights the in aqueous electrolytes, MnO2 has been explored as
representative electrochemical performance expected an inexpensive alternative to RuO2 (ref.78). In general,
for several common pseudocapacitive materials and charge storage in MnO2 occurs through a pseudo­
their inherent advantages and disadvantages. It should capacitive faradaic reaction between Mn4+ and Mn3+ in
also be noted that the group of materials identified here aqueous electrolytes:
is not intended to cover all pseudocapacitive materials.
In particular, we have not reviewed systems that have MnO2 + δ M + + δ e − ↔ M δ MnO2

Nature Reviews | Materials


Reviews

Table 1 | summary of representative intrinsic, extrinsic and non-​oxide pseudocapacitive materials


Classification Material electrolyte Capacity (mah g–1) Cycle life Cell advantages Disadvantages refs
(C-​rates) (no. of cycles) potentiala
Intrinsic pseudocapacitive
Surface redox RuO2•xH2O Aqueous 172 (C/3), 139 (80C) 5,000 (82% 1-V High σe; fast H+ High cost; 69

retention) window diffusion; high performance


theoretical dependent on extent
capacity of hydration
Surface redox MnO2 Aqueous 50 (20C), 40 (720C) 10,000 (100% 0.9-V High theoretical Low σe; non-​aqueous 91

(thick film) retention) window capacity; low cost instability


MnO2 263 (15C), 124 (80C) – 1.2-V 81

(thin film) window


Intercalation T-​Nb2O5 Non-​aqueous 140 (1C), 110 (100C) 1,000 (100% 1.2–3.0 V Fast bulk diffusion Low σe; high anodic 21

pseudocapacitance retention) vs Li/Li+ voltage


Intercalation TiO2(B)b Non-​aqueous 230 (1C), 160 (30C) 1,000 (100% 1–3 V vs Fast bulk diffusion; Low σe; high anodic 99

pseudocapacitance retention) Li/Li+ high theoretical voltage


capacity
Extrinsic pseudocapacitive
Size dependence Nano-​LiCoO2 Non-​aqueous 120 (1C), 75 (100C) 20 (75% 3.0–4.3 V High σi and σe; Inactive surface 108

retention) vs Li/Li+ 2D diffusion sites; poor cyclability


Bulk LiCoO2 145 (1C), 40 (100C)
Dopant chemistry MoO3–x Non-​aqueous 270 (1 mV s–1), 50 (97% 1.5–3.5 V 2D diffusion; Initial phase change; 118

150 (100 mV s–1) retention) vs Li/Li+ high σe; fast bulk poor cyclability
diffusion
MoO3 250 (1 mV s–1),
78 (100 mV s–1)
Anion intercalation LaMnO3–x Aqueous 200 (1 mV s–1), – 1.27-V Fast O2– diffusion; High cost; cyclability 122

127 (100 mV s–1) window high volumetric dependent on


capacity O2– stability
Non-​oxide
Nitride VN Aqueous 100 (10 mV s–1), 10,000 (95% 1.2-V 2D diffusion; Capacity dependent 125

73 (100 mV s–1) retention) window high σe on the degree of


surface oxidation
Sulfide Nano-​MoS2 Non-​aqueous 180 (1C), 90 (100C) 3,000 (80% 1–3 V vs High σe; Li+ and Initial phase change; 129

retention Li/Li+ Na+ insertion high anodic voltage


at 20C)
Carbide Ti3C2Tx Aqueous 110  (2 mV s–1), 10,000 (98% 0.9-V High σe; easy Performance 138

63  (1,000 mV s–1) retention) window fabrication dependent on


surface functional
group
Summary of useful parameters and performance metrics for intrinsic, extrinsic and non-​oxide pseudocapacitive materials, as well as the main advantages and
disadvantages of the materials. σe, electronic conductivity; σi, ionic conductivity; TiO2(B), bronze phase of TiO2; T-​Nb2O5, orthorhombic phase of Nb2O5; Tx, surface
termination functional group. aThe cell potential is the operating potential range in which redox occurs versus Li/Li+ for non-​aqueous electrolytes or the voltage
stability window for aqueous electrolytes. bElectrochemical data are for TiO2(B) nanosheets.

where M+ is an alkali metal cation or H+. Assuming a polymorph79,80. Increasing the number of diffusion path-
one-​electron redox reaction (δ = 1), a theoretical capacity ways and the size of the intercalation channels, which
of 308 mAh g–1 (over 0–0.9 V) can be obtained. However, depend on the crystal structure, increases the ionic
several studies have shown that the electrochemical conductivity and thus the obtainable capacity. Owing to
behaviour of bulk MnO2 is dependent on the phase as the variations in testing conditions, a verdict has not yet
well as the charge-​carrying cation in the electrolyte79,80. been reached on which phase of MnO2 has the high-
Unlike RuO2, MnO2 shows high ion-​diffusion rates, est capacity (λ-​MnO2 achieves a capacity of 68 mAh g–1
enabling bulk redox, but the low electronic conductiv- after 2.7 min charging in 0.5 M K2SO4 (ref.84), whereas α-​
ity (on the order of 10–7 S cm–1) limits redox reactions to MnO2 achieves 83 mAh g–1 after 5 min charging in 0.1 M
the near surface. Thin films have achieved high specific Na2SO4 (ref.79)). In contrast to the effects of the crystal
capacities (263 mAh g–1) but increasing the electrode structure on MnO2 properties, the surface area does not
thickness limits the obtainable capacity, typically to influence the performance. The absence of a surface-​
40–90 mAh g–1 (refs81–84) (Table 1). Through careful con- area dependence is an indication that bulk redox con-
trol of the synthesis conditions, it is possible to fabricate tributes significantly to the capacity, and this has been
MnO2 with a specific morphology and crystal struc- further confirmed using in situ Raman spectroscopy87
ture85,86. The crystal structure influences key material and in situ XRD88. In neutral aqueous electrolytes that
properties, such as conductivity, diffusivity and capacity, contain alkali metal salts, charge storage is dominated
and, thus, the performance of MnO2 is dependent on the by the proton contribution (>60%), stemming from

www.nature.com/natrevmats
Reviews

the high proton mobility and ease of proton insertion linear dQ/dV response10,43. Although T-​Nb2O5 retains
into the bulk88. With aprotic solvents, the capacitance a high energy density at high rates of charge and dis-
is lower and there is a capacity loss upon cycling com- charge, poor electronic conductivity necessitates nano-
pared with that in aqueous electro­lytes89–91. To increase structuring and a conductive network to achieve fast
the potential window and, thus, the energy density, the charge-​transfer reactions for thick electrodes26.
‘water-​in-salt’ approach has been explored, leading to an
increased capacity (106 mAh g–1 at 1.7 V in 5 M LiNO3)92. TiO2(B). Another promising intercalation material is
However, the low ion mobility of the electrolyte limited the bronze phase of titanium dioxide, TiO2(B)97. This
the achievable capacity at higher scan rates. phase of TiO2 can intercalate up to one Li ion into its
structure, giving a theoretical capacity of 335 mAh g–1
T-​Nb2O5. Nb2O5 has been of interest as an electro­ between 1.2 V and 3.0 V. Owing to its low structural
chemical energy-​storage material since the 1980s, when density and layered structure, fast Li+ diffusion is possi-
Li-​ion solid-​s olution intercalation was observed in ble in bulk TiO2(B) through channels between edge and
Nb2O5 at potentials <2 V versus Li/Li+ (refs93,94). This corner-​sharing [TiO6] octahedra parallel to the b axis.
intercalation behaviour produces broad faradaic charge-​ CV analysis of bulk TiO2(B) particles indicated a nearly
transfer peaks in the cyclic voltammogram, which dis- linear dependence of the peak current on the sweep rate
plays a notable surface-​charge-storage contribution (b = 1), signifying a surface-​controlled redox reaction98.
(>80%) with ion-​diffusion kinetics that are not limited The morphology of TiO2(B) also has a strong effect on
by semi-​infinite diffusion (b = 1)10,95. There are several the Li+ insertion mechanism and cycling kinetics. For
polytypes of crystalline Nb2O5, with the orthorhombic example, nanosheets99 (with a capacity of 230 mAh g–1
phase (T-​Nb2O5) displaying the highest capacity, cycling at 1C; Table  1 ) display improved Li-​i on insertion
stability and power density59 (140 mAh g–1 at 1C over kinetics over nanoparticles or nanowires (150 mAh g–1
1.2–3 V versus Li/Li+; Table 1). The amorphous phase at 1C for nanowires)100. This improvement originates
has also displayed pseudocapacitive behaviour, albeit from the decreased confinement in the c axis, which
at much lower capacities (69 mAh g–1 after 4 min charg- reduces the repulsion between neighbouring axial oxy-
ing), owing to the lack of intercalation channels in non-​ gen sites and channel sites100–103. Interestingly, NMR
crystalline Nb2O5 (ref.95). Li-​ion-based charge storage analyses of Li+ diffusion in these various morphologies
occurs in Nb2O5 through an intercalation reaction: do not corro­borate the experimental results104,105; how-
ever, the inclusion of anatase impurities could mask
Nb 2O5 + x Li+ + x e − ↔ Li x Nb 2O5 the true kinetics and capacity of the different morpho­
logies106,107. Similar to Nb2O5, the fast Li-​ion intercala-
where x ≤ 2. Fast intercalation was calculated to occur tion within TiO2(B) makes this material promising for
in the (001) plane because the large oxygen–oxygen dis- high-​rate devices, but the poor electronic conductivity
tance (3.9 Å) results in a low-​energy diffusion barrier restricts its incorporation into EES designs.
for hopping between adjacent NbO6 octahedra96. In situ
X-​ray absorption fine structure and XRD analysis of Extrinsic pseudocapacitive materials
T-​Nb2O5 during intercalation confirm that Li+ interca- Size dependence. Decreasing the size of the active mate-
lates as a solid solution with a linear change in its oxida- rial is the most commonly investigated route to engineer
tion state with voltage, reducing Nb5+ to Nb4+ and giving a extrinsic pseudocapacitive materials. This approach has
been widely explored with battery materials as a route
180 for increasing power density. With some battery materi-
160 TiO2(B) als, however, the electrochemical characteristics change
significantly and the materials exhibit a pseudocapac-
140 RuO2·xH2O
itive response. One of the more dramatic examples is
Capacity (mAh g–1)

LaMnO2.91
120 LiCoO2, the well-​known cathode material for Li-​ion bat-
T-Nb2O5 Charge time teries. When the nanoscale material (with a particle size
100
MoS2 Aqueous of <10 nm) is subjected to constant-​current (galvano­
80
VN static) conditions, the voltage profile changes from that of
60 Ti3C2 1s 120 s a typical battery, for which the majority of the charge is
40
MnO2 Non-aqueous stored at a constant potential, to a capacitive profile with
20
a linear dependence of charge storage within the poten-
1s 120 s
EDLC tial window108 (Fig. 1a). The change in electrochemical
0
–2.5 –2.0 –1.5 –1.0 –0.5 0.0 0.5 1.0 1.5 2.0 characteristics leads to a notable increase in the retention
Potential (V vs SHE) of capacity as the charging rates increase (Table 1). This
capacitive electrochemical signature has been associated
Fig. 2 | electrochemical performance of selected pseudocapacitive materials. with a decrease in the ionic and electronic diffusion dis-
The capacity of various aqueous (blue) and non-​aqueous (green) pseudocapacitive
tances (as the system enters the thin-​film electrochem-
material thick films with the corresponding operating potential range versus the standard
hydrogen electrode (SHE). The properties of these materials are summarized in Table 1. istry regime), suppression of phase transitions and an
The colour intensity indicates the time needed to charge, with a darker colour indicating increase in the number of available surface sites for Li+.
a faster charging time. The benchmark electrical double-​layer capacitor (EDLC) capacity The size dependence for pseudocapacitive materials
(20 mAh g–1 in ~1 s) is displayed in red. All materials have mass loadings above 0.5 mg cm–2. has been demonstrated for several systems25,28,57,109. For
TiO2(B), bronze phase of TiO2; T-​Nb2O5, orthorhombic phase of Nb2O5. example, decreasing the particle size of MoO2 to 20 nm

Nature Reviews | Materials


Reviews

suppresses the monoclinic–orthorhombic phase tran- The intercalated charge is compensated by the oxida-
sition that occurs upon lithiation. The capacity of the tion of Mn3+ to Mn4+, which diffuses towards the sur-
nanomaterial at 10C (120 mAh g–1) is three times greater face, maintaining a stable system122,123. The intercalation
than that of the micro-​sized powder (40 mAh g–1)56. of excess oxygen ions promotes high capacities that are
Moreover, by anchoring the MoO2 particles on a rGO retained at high rates122 (Table 1). However, the capac-
scaffold, the charge-​transfer properties improve greatly ity is highly dependent on the electrolyte composition,
and the MoO2–rGO material achieves a high capacity with hydroxide-​rich solutions (high pH) increasing
(150 mAh g–1) at a charge–discharge rate of 50C. the capacitance in comparison with acidic electrolytes
as the Mn2+/Mn3+ redox couple becomes dominant.
Dopant chemistry. Among energy-​storage materials, SrCo0.9Nb0.1O3–x and PrBaMn2O6–x also possess the per-
doping is routinely used to increase the electronic or ovskite structure and demonstrate excellent capacitive
ionic conductivity of a material and improve its electro- storage and stability at high specific currents (147 mAh g–1
chemical kinetics. One approach developed for layered at 10 A g–1 for SrCo0.9Nb0.1O3–x and 132 mAh g–1 at 10 A g–1
materials is to increase the interlayer spacing by doping for PrBaMn2O6–x), demonstrating a b value of 0.9, which
large moieties, such as polyoxometalates110, carbon111, indicates a surface-​controlled reaction123,124.
polymer chains112 or metal cations113,114, into interca-
lation channels. This approach has been examined for Non-​oxides. In addition to the oxide systems, pseudo-
materials such as MnO2, VS2 and Ti3C2, for which the capacitive responses have also been reported for vari­
increase in interlayer spacing can drastically increase ous nitrides, sulfides and carbides. One interesting
2D conduction115,116. Note that the choice of insertion pseudocapacitive nitride is thin-​film vanadium nitride
species is important, as it must not impede diffusion (VN), which has an electronically conductive rock salt
through the channel117. Another route to improving the structure that develops an oxy-​VN surface in aqueous
electrochemical kinetics is to introduce oxygen vacan- electrolytes. The charge-​storage mechanism is pro-
cies. For example, by introducing vacancies in MoO3, posed to involve the simultaneous redox of V3+ to V2+
the b-​axis spacing increased, allowing for faster Li-​ion with that of an oxy/oxyhydrate couple at the surface. For
insertion kinetics without a phase change and, thus, sub- thin films, this reaction gives a broad CV response with
stantially higher charge-​storage contributions and rate capacities of up to 446 mAh g–1 over a 1.2-V window125.
capabilities118 (Table 1). The introduction of vacancies In general, the charge-​storage properties of nitrides are
has the additional benefit that these sites act as shal- strongly dependent on the electrolyte; for example, the
low donors, forming localized midgap states that can highest pseudocapacitive contributions occur with basic
increase the electronic conductivity, but they can also electrolytes126,127.
lead to instability of the structure during cycling. A simi­ Another class of non-​oxide pseudocapacitive materi-
lar improvement to the kinetics has been observed with als is the transition-​metal dichalcogenides (MX2). These
anatase, for which partial reduction increased the electro­ materials, which typically have trigonal prismatic (H)
nic and ionic conductivity, leading to an enhanced rate or octahedral (T) phases, are of considerable interest for
capability119. energy storage, as their layered structure allows for facile
ion access and high electronic conductivity128. One of
Anion-​intercalation pseudocapacitance. Another route the most well-​studied systems is MoS2. During electro­
to increase pseudocapacitive-​b ased charge storage chemical cycling, Li+ inserts within the van der Waals
is through redox reactions based on anion intercala- gap, enabling MoS2 to intercalate 1 mol of Li+ (corre-
tion. In this approach, an anion (for example, oxygen), sponding to a capacity of 167 mAh g–1) between 0.8 V
is intercalated into a host material, causing cationic and 3.0 V versus Li/Li+. At lower voltages (<0.8 V), MoS2
charge transfer. Anion-​intercalation pseudocapacitive is converted into metallic Mo, which limits the electro-
materials should not be confused with anionic redox chemical stability of the cell129. Li+ insertion into the lay-
materials, such as Li-​rich oxides, which undergo anionic ered structure has been demonstrated computationally
charge transfer120,121. Non-​traditional systems such as and experimentally, confirming facile 2D ion conduc-
oxygen-​d eficient perovskites have been shown to tion128,129. Upon Li+ intercalation in bulk MoS2, the metal-
display pseudo­c apacitive behaviour through anion lic 1T phase undergoes a reversible structural change to
intercalation in aqueous solutions. For example, for the triclinic phase (LixMoS2), owing to the low solubility
pseudocapacitive LaMnO3 (ref.122), anion intercalation of Li+ in the 1T phase. Recent operando XRD studies
was proposed to occur through the insertion of O2– from have indicated that the intercalation mechanism depends
electrolyte OH– ions; the protons released during this on the size of the MoS2 particles. When the particle size
process combine with OH– ions in the electrolyte to is reduced to tens of nanometres, intercalation of Li+
form water: occurs by a solid-​solution mechanism without a phase
change. This process enables increased energy storage at
high charge–discharge rates (the capacities of nanostruc-
A(B 22+x ,B1−2
3+
x )O3−x +2x OH

tured and bulk MoS2 are 115 mAh g–1 and 40 mAh g–1,
↔ AB3+ O3+2x e − + x H 2O respectively, at 60C)25. Selected data for a nano-​MoS2
are provided in Table 1. Similar electrochemical behavi­
where A is La and B is Mn. The oxygen vacancies in our has also been reported for other layered dichalco-
perovskite structures enable oxygen-​ion intercalation genides, such as WS2 and TiS2; both materials show high
and hopping of O2– ions between adjacent vacancies. surface capacity contributions of 85% at slow rates130,131.

www.nature.com/natrevmats
Reviews

The intercalation kinetics of layered dichalcogenides MoO2–rGO) and a stable phase that allows for solid-​
are greatly affected by the interlayer spacing and multi- solution intercalation (for example, T-​Nb2O5). Intrinsic
ple methods are being explored in an effort to increase pseudocapacitive materials inherently display these
the spacing to ultimately improve solid-​state ion diffu- characteristics. For extrinsic pseudocapacitive materi-
sion and rate capability132,133. Although the ion diffusion als, one or more of these criteria are addressed through
coefficient can be tuned, there is still a need for short ion materials engineering strategies, such as nanostructuring
and electron conduction pathways to obtain apprecia- and dopant chemistry. Although a material exhibiting
ble capacities at high charge–discharge rates, as demon- these properties does not guarantee pseudocapacitive
strated for nano-​MoS2 versus bulk MoS2 (refs25,129). The electrochemical signatures, these criteria can provide
prospect of aligning the intercalation pathways of layered some guidance as to whether a new material should
materials may enable a significant improvement in the be evaluated for a pseudocapacitive response.
performance and use in commercial-​sized devices.
Carbides constitute a third class of non-​oxide pseu- Integration into hybrid EES devices
docapacitive materials. Over the past decade, pseudo- Early efforts in integrating pseudocapacitive materials
capacitive carbides, based on the 2D layered MXene (such as MnO2 and RuO2) into EES devices centred
family of materials (for example, Ti3C2, Ti2C, V2C, Nb2C on the development of hybrid devices with aqueous
and Ti3CN), have been the subject of intense research134. electro­lytes141–144. A hybrid device comprises two dif-
MXenes consist of 2D hexagonal sheets with an electron- ferent electrodes; one electrode is a redox material and
ically conductive carbide core and a transition-​metal (M) the other is a carbon-​based electrical double-​layer elec-
surface produced from etching the A atoms from MAX trode, typically activated carbon (Fig. 3a). By contrast,
(where A is typically a group 13 or 14 element and X is an EDLC comprises two activated-​carbon electrodes
carbon and/or nitrogen) phase materials. MXenes have (Fig. 3b). Hybrid devices were designed with the intent of
high surface areas and can undergo faradaic reactions, operating at power levels comparable to those of EDLC
leading to theoretical capacities of up to 320 mAh g–1 devices but with much higher energy densities through
between 0.05 V and 2.5 V versus Li/Li+ for Li2Ti3C2. the incorporation of redox-​active materials145. Although
This family of materials is amenable to the intercalation these early aqueous devices did not significantly sur-
of various ions between the 2D layers135, including Li+ pass EDLC performance (5 Wh kg–1 at 10 kW kg–1 (note
(in Ti3CNTx; 190 mAh g–1 at 50 mA g–1)136, Na+ (Ti3C2Tx; that the specific energies and powers listed here are
140 mAh g–1 at 50 mA g–1)137, K+ (Ti3C2Tx; 100 mAh g–1 at based on the total mass of the device unless otherwise
50 mA g–1)137 and H+ (Ti3C2Tx; 110 mAh g–1 at 1 mV s–1 indicated))146,147, they laid the groundwork for further
Table 1)138; these intercalated materials all display broad development of hybrid devices.
faradaic peaks in the CV response. MXenes are typically Hybrid devices based on non-​aqueous electrolytes
cycled in aqueous environments, in which the charging can achieve much higher capacities and operating volt-
mechanism can be controlled by changing the pH of the ages (~4 V) and, therefore, higher energy densities, than
electrolyte. In acidic electrolytes, redox is centred on those based on aqueous electrolytes. Indeed, hybrid
the M=O/M–OH couple; the redox couple is induced supercapacitors with non-​aqueous electrolytes have
by the adsorption of H3O+ from the electrolyte onto the achieved higher energy densities than symmetric EDLC
surface of the MXene, shifting the valence state of devices (Fig. 3c). Li-​ion capacitors (LICs) are prominent
the transition metal139. An important consideration with hybrid devices based on Li-​ion battery electrode mate-
MXenes is the correlation between the charge-​storage rials paired with activated carbon. High-​rate LICs based
mechanism and etching processes, which determines the on Li4Ti5O12 were the first established LIC anodes and
surface termination groups. Etching using hydrofluoric undergo redox reactions at 1.55 V versus Li/Li+, which
acid leaves a surface terminated with either –F or –OH is within the electrochemical stability window of most
groups that hinder ion access to the transition metal, common non-​aqueous, organic electrolytes148. LICs
while also increasing the charge-​transfer barrier. For based on Li 4Ti5O12 have demonstrated gravimetric
Ti3C2, –OH termination was found to be more detrimen- energy densities 3–4 times higher than those of conven-
tal than –F termination, resulting in diffusion barriers tional EDLCs, up to 40 Wh kg–1 at a specific power of
that are several times larger than those for F-​terminated 5 kW kg–1 (refs149,150). Li4Ti5O12 remains one of the few
and bare Ti3C2 (0.9 eV (–OH) > 0.36 eV (–F) > 0.07 eV materials that can cycle more than 10,000 times without
(bare))140. Although MXenes have promising electronic considerable degradation, as it undergoes only a small
and ionic kinetics, further study is needed to understand volume change and has a limited potential window.
how the surface functionalization and interlayer spacing Modern devices commonly use graphite as the faradaic
affects the charge mechanism and kinetics. anode (owing to its higher specific capacity and lower
The selected intrinsic and extrinsic pseudocapacitive potential compared with Li4Ti5O12), which results in
materials discussed above display the various electro- an increased working voltage and energy densities of
chemical signatures of pseudocapacitance. It is apparent 40–60 Wh kg−1 (based on the total mass of the electrode
that there are underlying structural characteristics that materials)151. Unfortunately, graphite-​based LICs are
are typically displayed in a pseudocapacitive material: limited by the sluggish diffusion of Li+ in graphite, the
an open crystal structure with multiple diffusion path- formation of a solid-​electrolyte interphase and the need
ways that enable fast ion conduction (as in α-​MnO2 for pre-​lithiation of the graphite electrode152,153.
or TiO2(B)), high electronic conductivity (as in RuO2) or The next step in the evolution of hybrid devices is
a means of enhancing electron transport (for example, the incorporation of non-​aqueous pseudocapacitive

Nature Reviews | Materials


Reviews

materials as a route for maintaining high power den- achievable energy density and greatly expand device
sities while simultaneously increasing energy densities. applications for mobile power.
Prototypes of hybrid devices based on the intrinsic Na-​ion capacitors are a future area of development
pseudocapacitive materials T-​Nb2O5 and TiO2(B) have that can take advantage of high-​rate pseudocapacitive
demonstrated energy densities at high-​p ower lev- materials. Despite the obvious size disadvantage of Na+
els comparable to those of the LICs that use Li4Ti5O12 compared with Li+, the lower desolvation energy of
(refs21,154–156) (Fig. 3d). Moreover, thick Nb2O5 electrodes Na+ may enable the development of high-​rate devices159,160.
(with mass loadings >10 mg cm–2) have demonstrated Although there are few reports on pseudocapacitive
that intrinsic pseudocapacitive materials can deliver materials for Na-​ion capacitors, MXenes are promising
high-​rate performance at practical mass loadings for candidates, and Na-​ion capacitors may yet turn out to be
commercialization157. In the progression towards high-​ an area in which significant progress can be made161,162.
energy-density devices, positive-​electrode pseudoca-
pacitive materials with high voltages versus Li/Li+ still Future perspectives
present a challenge for hybrid devices, owing to the lack Pseudocapacitive materials exhibit a unique combi-
of identified materials and the difficulty of nanostruc- nation of properties: redox reactions that occur with
turing158. Asymmetric pseudocapacitor devices with kinetics similar to those of charge storage in EDLC
two pseudocapacitive electrodes would increase the materials. These materials fill an important gap in the
energy-​storage field, namely the lack of materials that
have the energy density of battery materials and the
a + – c 5 power density of capacitive materials. The interest in
Anion Oxidative decomposition of electrolyte
pseudocapacitive materials spans from fundamental
– Li+ AC AC
studies to EES devices. It is certain that the allure of

Electrolyte stability window


4
– Li+ these materials will become more widespread because
Potential (V vs Li+/Li)

their unique combination of properties is suitable for


– Li+ 3 several of the expected growth areas for energy storage.
AC MOx Looking ahead, there are key questions regarding new
AC materials, energy densities and device performance that
b + – 2
Anion Cation have to be addressed before practical applications of
Nb2O5
– + pseudocapacitive materials are realized.
1 An apparent question to ask is whether any poten-
– + Reductive decomposition of electrolyte and tial classes of materials have been overlooked during
SEI formation
– + the exploration of pseudocapacitive materials. One
0
0 25 50 75 100 125 150 175 200 class of materials that seems to have gone unnoticed in
AC AC Capacity (C g–1) the context of pseudocapacitive materials is those that
d Pseudocapacitive hybrid devices exhibit anionic redox, an approach that has already led to
EDLC LICs Aqueous Non-aqueous increased capacity with battery materials163. In this regard,
80
sulfides provide an interesting opportunity. Owing to
their covalent nature, sulfides are better understood than
Specific energy (Wh kg–1)

70
60 oxides with regard to anionic redox behaviour164–166. It is
50 also interesting to note that amorphous pseudocapaci-
40 tive materials, which may have differing ionic and elec-
30 tronic properties from the bulk, have received limited
20 study. Although there has been some work with aqueous
10 systems167,168, to the best of our knowledge, there have
0 yet to be reports of amorphous pseudocapacitive mate-
AC Graphite Li4Ti5O12 MnO2 RuO2 TiO2(B) T-Nb2O5 Ti3C2
rials for non-​aqueous EES. Extrinsic pseudocapacitive
Maximum Specific energy Specific energy materials can be synthesized by numerous modifica-
specific energy at 1 kW kg–1 at 5 kW kg–1 tions and nanostructuring routes, and it is likely that
this area will continue to expand as researchers develop
Fig. 3 | Comparison of hybrid device architectures and performance. a, b | Architecture new approaches for synthesizing nanoscale materi-
of a hybrid device with one activated carbon (AC) electrode and one metal oxide (MOx) als. In addition, new scientific directions may begin
electrode (panel a) and an electrical double-​layer capacitor (EDLC) with two AC electrodes to emerge, as there are fundamental questions of ther-
(panel b). c | The electrochemical voltage profiles of a typical EDLC and a hybrid device modynamic metastability and the possibility to achieve
with a Nb2O5 electrode. The area between the curves of the hybrid device (yellow) is greater than one-​electron redox with materials on the
much larger than that between the curves of the EDLC (blue), indicating that the hybrid nanoscale. Colloidal chemistry, for example, offers new
device has a higher available energy density within the electrolyte stability window. synthetic routes for extrinsic pseudocapacitive materi-
d | Comparison of the maximum specific energy, specific energy at a specific power of
als and the ability to form compositional gradients to
1 kW kg–1 and specific energy at a specific power of 5 kW kg–1 for EDLCs147, Li-​ion capacitors
(LICs)149,151 and pseudocapacitive hybrid capacitors142,143,155,156,162, all versus an AC electrode improve the electrochemical performance169–171.
with mass loadings above 1 mg cm–2. Pseudocapacitive hybrid devices show promise for The second question to consider is how to close the
increasing the maximum specific energy of mobile power devices145. SEI, solid electrolyte energy-​density gap between batteries and pseudocapac-
interphase; TiO2(B), bronze phase of TiO2; T-​Nb2O5, orthorhombic phase of Nb2O5. itors. Although techniques such as nanostructuring can
Panels a–d are adapted with permission from ref.150, RSC. produce capacitor-​like kinetics, the energy density of

www.nature.com/natrevmats
Reviews

nanomaterials is limited by the inherent properties or high-​voltage pseudocapacitive materials. There is


of the bulk material. Although there are several pseudo- every reason to believe that the hybrid device field will be
capacitive materials that operate between 2.5 V and 1.0 V a robust one that achieves an energy density equal to half
(versus Li/Li+) that can achieve high energy densities at that of a Li-​ion battery at power densities on par with
high power densities, there are very few that operate at carbon-​based supercapacitors. Further progress could
potentials >3.5 V (Table 1). High-​voltage pseudocapac- lead to a Li-​ion battery based on two intercalation-​type
itive materials would enable the replacement of carbon electrodes, both of which are pseudocapacitive materials
in hybrid devices, thus producing asymmetric pseudo- (an example of an asymmetric pseudocapacitor), with
capacitors with increased energy densities159. Layered each electrode capable of achieving capacities well above
transition-​metal oxides, such as LiNi0.8Co0.15Al0.05O2 and 100 mAh g–1 and being charged in minutes.
LiNixMnyCo1–x–yO2, have high achievable capacities that The development of EES devices that exhibit the
arise from the substitution of nickel and present a prom- energy density of a battery at the high power density of
ising research direction, as these materials exhibit solid-​ a supercapacitor has yet to be realized. Using existing
solution intercalation above 3.8 V after the first charge, materials and architectures as a starting point, however,
as verified by operando XRD172,173. we now have various pathways to design novel materi-
The last question we consider is the practical advan- als that can help reach the desired performance of high-​
tages of pseudocapacitive materials and whether their energy, high-​power devices. Although the best materials
properties will lead to tangible improvements in EES and devices have yet to be achieved, pseudocapacitive
device performance. Hybrid pseudocapacitive devices materials offer glimpses of a new generation of energy-​
are currently comparable to Li4Ti5O12-based LICs in storage materials that will lead to unmatched energy for
terms of retaining high energy density at high power high-​power devices.
density. The ability to surpass present performance levels
will depend on advances in designing new high-​capacity Published online xx xx xxxx

1. IEA. Global EV Outlook 2017: Two Million and Counting. 15. Ko, J. S., Sassin, M. B., Rolison, D. R. & Long, J. W. energy storage in TiO2 (anatase) nanoparticles. J. Phys.
International Energy Agency https://www.iea.org/ Deconvolving double-​layer, pseudocapacitance, and Chem. C 111, 14925–14931 (2007).
publications/freepublications/publication/ battery-​like charge-​storage mechanisms in nanoscale 29. Forghani, M. & Donne, S. W. Method comparison
GlobalEVOutlook2017.pdf (2017). LiMn2O4 at 3D carbon architectures. Electrochim. Acta for deconvoluting capacitive and pseudo-​capacitive
2. Liang, Y. et al. A review of rechargeable batteries for 275, 225–235 (2018). contributions to electrochemical capacitor electrode
portable electronic devices. InfoMat 1, 6–32 (2019). 16. Ko, J. S., Sassin, M. B., Parker, J. F., Rolison, D. R. & behavior. J. Electrochem. Soc. 165, A664–A673
3. Gür, T. M. Review of electrical energy storage Long, J. W. Combining battery-​like and pseudocapacitive (2018).
technologies, materials and systems: challenges and charge storage in 3D MnOx@carbon electrode 30. Shao, H., Lin, Z., Xu, K., Taberna, P.-L. & Simon, P.
prospects for large-​scale grid storage. Energy Environ. architectures for zinc-​ion cells. Sustain. Energy Fuels 2, Electrochemical study of pseudocapacitive behavior
Sci. 11, 2696–2767 (2018). 626–636 (2018). of Ti3C2Tx MXene material in aqueous electrolytes.
4. Davies, D. M. et al. Combined economic and 17. Lindström, H. et al. Li+ ion insertion in TiO2 (anatase). Energy Storage Mater. 18, 456–461 (2019).
technological evaluation of battery energy storage 2. Voltammetry on nanoporous films. J. Phys. Chem. B 31. Weppner, W. & Huggins, R. A. Determination of the
for grid applications. Nat. Energy 4, 42–50 (2019). 101, 7717–7722 (1997). kinetic parameters of mixed-​conducting electrodes
5. Salanne, M. et al. Efficient storage mechanisms This paper set the foundation for providing kinetic and application to the system Li3Sb. J. Electrochem.
for building better supercapacitors. Nat. Energy 1, analysis that distinguishes between semi-​infinite Soc. 124, 1569–1578 (1977).
16070 (2016). and capacitor-​like behaviour. 32. Takami, N. Structural and kinetic characterization of
6. Conway, B. E. Transition from “supercapacitor” to 18. Randles, J. E. B. A cathode ray polarograph. Part II.— lithium intercalation into carbon anodes for secondary
“battery” behavior in electrochemical energy storage. The current-​voltage curves. Trans Faraday Soc. 44, lithium batteries. J. Electrochem. Soc. 142, 371–379
J. Electrochem. Soc. 138, 1539–1548 (1991). 327–338 (1948). (1995).
A seminal paper that distinguishes supercapacitor 19. Ševčík, A. Oscillographic polarography with periodical 33. Cottrell, F. G. Der Reststrom bei galvanischer
behaviour from that of batteries. triangular voltage. Collect. Czech. Chem. Commun. 13, Polarisation, betrachtet als ein Diffusionsproblem
7. Conway, B. E. Electrochemical Supercapacitors: 349–377 (1948). [German]. Z. Phys. Chem. 42U, 385–431 (1903).
Scientific Fundamentals and Technological Applications. 20. Come, J., Taberna, P.-L., Hamelet, S., Masquelier, C. 34. Bard, A. J. & Faulkner, L. R. Electrochemical Methods:
Ch. 2 & Ch. 3 (Springer US, 1999). & Simon, P. Electrochemical kinetic study of LiFePO4 Fundamentals and Applications. 2nd edn Ch. 11
8. Abruña, H. D., Kiya, Y. & Henderson, J. C. Batteries using cavity microelectrode. J. Electrochem. Soc. 158, & Ch. 15 (Wiley, 2001).
and electrochemical capacitors. Phys. Today 61, A1090–A1093 (2011). 35. Gibson, A. J. & Donne, S. W. A step potential
43–47 (2008). 21. Augustyn, V. et al. High-​rate electrochemical energy electrochemical spectroscopy (SPECS) investigation
9. Brousse, T., Bélanger, D. & Long, J. W. To be or not storage through Li+ intercalation pseudocapacitance. of anodically electrodeposited thin films of manganese
to be pseudocapacitive? J. Electrochem. Soc. 162, Nat. Mater. 12, 518–522 (2013). dioxide. J. Power Sources 359, 520–528 (2017).
A5185–A5189 (2015). This paper provides the electrochemical and 36. Dupont, M. F. & Donne, S. W. A step potential
This paper presents the defining features of structural characteristics that helped to define electrochemical spectroscopy analysis of electrochemical
pseudocapacitive materials and dispels some pseudocapacitive materials. capacitor electrode performance. Electrochim. Acta
of the misimpressions in the field. 22. Xiao, X. et al. Freestanding MoO3–x nanobelt/carbon 167, 268–277 (2015).
10. Come, J. et al. Electrochemical kinetics of nanotube films for Li-​ion intercalation pseudocapacitors. 37. Girard, H.-L., Wang, H., d’Entremont, A. L. & Pilon, L.
nanostructured Nb2O5 electrodes. J. Electrochem. Nano Energy 9, 355–363 (2014). Enhancing faradaic charge storage contribution in
Soc. 161, A718–A725 (2014). 23. Lai, C.-H. et al. Application of poly(3-hexylthiophene- hybrid pseudocapacitors. Electrochim. Acta 182,
11. Girard, H.-L., Dunn, B. & Pilon, L. Simulations 2,5-diyl) as a protective coating for high rate 639–651 (2015).
and interpretation of three-​electrode cyclic cathode materials. Chem. Mater. 30, 2589–2599 38. Nicholson, R. S. Theory and application of cyclic
voltammograms of pseudocapacitive electrodes. (2018). voltammetry for measurement of electrode reaction
Electrochim. Acta 211, 420–429 (2016). 24. Ko, J. S. et al. High-​rate capability of Na2FePO4F kinetics. Anal. Chem. 37, 1351–1355 (1965).
12. Costentin, C., Porter, T. R. & Savéant, J.-M. How do nanoparticles by enhancing surface carbon 39. Mahmood, Q. et al. Transition from diffusion-​controlled
pseudocapacitors store energy? Theoretical analysis functionality for Na-​ion batteries. J. Mater. Chem. A 5, intercalation into extrinsically pseudocapacitive charge
and experimental illustration. ACS Appl. Mater. 18707–18715 (2017). storage of MoS2 by nanoscale heterostructuring.
Interfaces 9, 8649–8658 (2017). 25. Cook, J. B. et al. Suppression of electrochemically Adv. Energy Mater. 6, 1501115 (2016).
13. Bai, L. & Conway, B. E. AC impedance of faradaic driven phase transitions in nanostructured MoS2 40. Wu, X. et al. Diffusion-​free Grotthuss topochemistry for
reactions involving electrosorbed intermediates: pseudocapacitors probed using operando X-​ray high-​rate and long-​life proton batteries. Nat. Energy 4,
examination of conditions leading to pseudoinductive diffraction. ACS Nano 13, 1223–1231 (2019). 123–130 (2019).
behavior represented in three-​dimensional impedance 26. Lai, C.-H. et al. Designing pseudocapacitance for 41. Toupin, M., Brousse, T. & Bélanger, D. Influence
spectroscopy diagrams. J. Electrochem. Soc. 138, Nb2O5/carbide-​derived carbon electrodes and hybrid of microstucture on the charge storage properties
2897–2907 (1991). devices. Langmuir 33, 9407–9415 (2017). of chemically synthesized manganese dioxide.
14. Taberna, P. L., Simon, P. & Fauvarque, J. F. 27. Liu, T.-C. Behavior of molybdenum nitrides as materials Chem. Mater. 14, 3946–3952 (2002).
Electrochemical characteristics and impedance for electrochemical capacitors. J. Electrochem. Soc. 42. Tang, Y. et al. Identifying the origin and contribution
spectroscopy studies of carbon-​carbon 145, 1882–1888 (1998). of surface storage in TiO2(B) nanotube electrode by
supercapacitors. J. Electrochem. Soc. 150, 28. Wang, J., Polleux, J., Lim, J. & Dunn, B. in situ dynamic valence state monitoring. Adv. Mater.
A292–A300 (2003). Pseudocapacitive contributions to electrochemical 30, 1802200 (2018).

Nature Reviews | Materials


Reviews

43. Kodama, R., Terada, Y., Nakai, I., Komaba, S. & devices. J. Solid State Electrochem. 7, 637–644 91. Li, L. et al. Facile synthesis of MnO2/CNTs composite
Kumagai, N. Electrochemical and in situ XAFS-​XRD (2003). for supercapacitor electrodes with long cycle stability.
investigation of Nb2O5 for rechargeable lithium 68. Zhang, J. et al. Template synthesis of tubular ruthenium J. Phys. Chem. C 118, 22865–22872 (2014).
batteries. J. Electrochem. Soc. 153, A583–A588 oxides for supercapacitor applications. J. Phys. Chem. C 92. Gambou-​Bosca, A. & Bélanger, D. Electrochemical
(2006). 114, 13608–13613 (2010). characterization of MnO2-based composite in the
44. Christensen, C. R. & Anson, F. C. Chronopotentiometry 69. Ma, H. et al. Disassembly–reassembly approach to presence of salt-​in-water and water-​in-salt electrolytes
in thin layer of solution. Anal. Chem. 35, 205–209 RuO2/graphene composites for ultrahigh volumetric as electrode for electrochemical capacitors. J. Power
(1963). capacitance supercapacitor. Small 13, 1701026 (2017). Sources 326, 595–603 (2016).
45. Hubbard, A. T. Study of the kinetics of electrochemical 70. Mondal, S. K. & Munichandraiah, N. Anodic 93. Reichman, B. & Bard, A. J. The application of Nb2O5 as
reactions by thin-​layer voltammetry. J. Electroanal. deposition of porous RuO2 on stainless steel for a cathode in nonaqueous lithium cells. J. Electrochem.
Chem. Interfacial Electrochem. 22, 165–174 (1969). supercapacitor studies at high current densities. Soc. 128, 344–346 (1981).
46. Andrieux, C. P. & Savéant, J. M. Electron transfer J. Power Sources 175, 657–663 (2008). 94. Ohzuku, T., Sawai, K. & Hirai, T. Electrochemistry
through redox polymer films. J. Electroanal. Chem. 71. Sugimoto, W., Yokoshima, K., Murakami, Y. & Takasu, Y. of L-​niobium pentoxide a lithium/non-​aqueous cell.
Interfacial Electrochem. 111, 377–381 (1980). Charge storage mechanism of nanostructured J. Power Sources 19, 287–299 (1987).
47. Peerce, P. J. & Bard, A. J. Polymer films on electrodes. anhydrous and hydrous ruthenium-​based oxides. 95. Kim, J. W., Augustyn, V. & Dunn, B. The effect of
J. Electroanal. Chem. Interfacial Electrochem. 114, Electrochim. Acta 52, 1742–1748 (2006). crystallinity on the rapid pseudocapacitive response
89–115 (1980). 72. Long, J. W., Swider, K. E., Merzbacher, C. I. & of Nb2O5. Adv. Energy Mater. 2, 141–148 (2012).
48. Laviron, E., Roullier, L. & Degrand, C. A multilayer Rolison, D. R. Voltammetric characterization of 96. Liu, C.-P., Zhou, F. & Ozolins, V. First principles study
model for the study of space distributed redox ruthenium oxide-​based aerogels and other RuO2 for lithium intercalation and diffusion behavior in
modified electrodes. J. Electroanal. Chem. Interfacial solids: the nature of capacitance in nanostructured orthorhombic Nb2O5 electrochemical supercapacitor.
Electrochem. 112, 11–23 (1980). materials. Langmuir 15, 780–785 (1999). APS Meet. Abstr. B26. 003 (2012).
49. Aoki, K., Tokuda, K. & Matsuda, H. Theory of linear 73. Hu, C.-C., Chang, K.-H., Lin, M.-C. & Wu, Y.-T. Design 97. Marchand, R., Brohan, L. & Tournoux, M. TiO2(B)
sweep voltammetry with finite diffusion space: and tailoring of the nanotubular arrayed architecture a new form of titanium dioxide and the potassium
Part II. Totally irreversible and quasi-​reversible cases. of hydrous RuO2 for next generation supercapacitors. octatitanate K2Ti8O17. Mater. Res. Bull. 15,
J. Electroanal. Chem. Interfacial Electrochem. 160, Nano Lett. 6, 2690–2695 (2006). 1129–1133 (1980).
33–45 (1984). 74. Bi, R.-R. et al. Highly dispersed RuO2 nanoparticles 98. Zukalová, M., KalbáČ, M., Kavan, L., Exnar, I. &
This paper outlines the development of the on carbon nanotubes: facile synthesis and enhanced Graetzel, M. Pseudocapacitive lithium storage in
parameters that helped to define finite supercapacitance performance. J. Phys. Chem. C 114, TiO2(B). Chem. Mater. 17, 1248–1255 (2005).
diffusion and its relationship with thin-​layer 2448–2451 (2010). 99. Li, X., Wu, G., Liu, X., Li, W. & Li, M. Orderly integration
electrochemistry. 75. Liu, Y., Zhou, F. & Ozolins, V. Ab initio study of of porous TiO2(B) nanosheets into bunchy hierarchical
50. Mirčeski, V. & Tomovski, Ž. Modeling of a voltammetric the charge-​storage mechanisms in RuO2-based structure for high-​rate and ultralong-​lifespan lithium-​ion
experiment in a limiting diffusion space. J. Solid State electrochemical ultracapacitors. J. Phys. Chem. C 116, batteries. Nano Energy 31, 1–8 (2017).
Electrochem. 15, 197–204 (2011). 1450–1457 (2012). 100. Liu, S. et al. Nanosheet-​constructed porous TiO2–B
51. Lovrić, M., Komorsky-Lovrić, Š. & Scholz, F. Staircase 76. Dmowski, W., Egami, T., Swider-​Lyons, K. E., Love, C. T. for advanced lithium ion batteries. Adv. Mater. 24,
voltammetry with finite diffusion space. Electroanalysis & Rolison, D. R. Local atomic structure and conduction 3201–3204 (2012).
9, 575–577 (1997). mechanism of nanocrystalline hydrous RuO2 from 101. Dylla, A. G., Xiao, P., Henkelman, G. & Stevenson, K. J.
52. Aoki, K. & Osteryoung, J. Square wave voltammetry X-​ray scattering. J. Phys. Chem. B 106, 12677–12683 Morphological dependence of lithium insertion in
in a thin-layer cell. J. Electroanal. Chem. Interfacial (2002). nanocrystalline TiO2(B) nanoparticles and nanosheets.
Electrochem. 240, 45–51 (1988). 77. Wen, S., Lee, J.-W., Yeo, I.-H., Park, J. & Mho, S. The role J. Phys. Chem. Lett. 3, 2015–2019 (2012).
53. Komorsky-​Lovrić, Š. & Lovrić, M. Kinetic of cations of the electrolyte for the pseudocapacitive 102. Hua, X. et al. Lithiation thermodynamics and kinetics
measurements of a surface confined redox reaction. behavior of metal oxide electrodes, MnO2 and RuO2. of the TiO2 (B) nanoparticles. J. Am. Chem. Soc. 139,
Anal. Chim. Acta 305, 248–255 (1995). Electrochim. Acta 50, 849–855 (2004). 13330–13341 (2017).
54. Mirčeski, V. Charge transfer kinetics in thin-​film 78. Lee, H. Y. & Goodenough, J. B. Supercapacitor 103. Ren, Y. et al. Nanoparticulate TiO2(B): an anode
voltammetry. Theoretical study under conditions behavior with KCl electrolyte. J. Solid State Chem. for lithium-​ion batteries. Angew. Chem. Int. Ed. 51,
of square-​wave voltammetry. J. Phys. Chem. B 108, 144, 220–223 (1999). 2164–2167 (2012).
13719–13725 (2004). 79. Ghodbane, O., Pascal, J.-L. & Favier, F. Microstructural 104. Wilkening, M., Lyness, C., Armstrong, A. R. &
55. Simon, P., Gogotsi, Y. & Dunn, B. Where do batteries effects on charge-​storage properties in MnO2-based Bruce, P. G. Diffusion in confined dimensions:
end and supercapacitors begin? Science 343, electrochemical supercapacitors. ACS Appl. Mater. Li+ transport in mixed conducting TiO2–B nanowires.
1210–1211 (2014). Interfaces 1, 1130–1139 (2009). J. Phys. Chem. C 113, 4741–4744 (2009).
56. Kim, H.-S., Cook, J. B., Tolbert, S. H. & Dunn, B. 80. Devaraj, S. & Munichandraiah, N. Effect of 105. Hoshina, K., Harada, Y., Inagaki, H. & Takami, N.
The development of pseudocapacitive properties crystallographic structure of MnO2 on its Characterization of lithium storage in TiO2(B) by
in nanosized-​MoO2. J. Electrochem. Soc. 162, electrochemical capacitance properties. 6
Li-​NMR and X-​ray diffraction analysis. J. Electrochem.
A5083–A5090 (2015). J. Phys. Chem. C 112, 4406–4417 (2008). Soc. 161, A348–A354 (2014).
57. Okubo, M. et al. Fast Li-​ion insertion into nanosized 81. Yan, W. et al. Lithographically patterned gold/ 106. Laskova, B., Zukalova, M., Zukal, A., Bousa, M.
LiMn2O4 without domain boundaries. ACS Nano 4, manganese dioxide core/shell nanowires for high & Kavan, L. Capacitive contribution to Li-​storage in
741–752 (2010). capacity, high rate, and high cyclability hybrid electrical TiO2 (B) and TiO2 (anatase). J. Power Sources 246,
58. Mei, B.-A., Li, B., Lin, J. & Pilon, L. Multidimensional energy storage. Chem. Mater. 24, 2382–2390 103–109 (2014).
cyclic voltammetry simulations of pseudocapacitive (2012). 107. Giannuzzi, R. et al. Ultrathin TiO2(B) nanorods with
electrodes with a conducting nanorod scaffold. 82. Zhao, Y., Li, M. P., Liu, S. & Islam, M. F. Superelastic superior lithium-​ion storage performance. ACS Appl.
J. Electrochem. Soc. 164, A3237–A3252 (2017). pseudocapacitors from freestanding MnO2-decorated Mater. Interfaces 6, 1933–1943 (2014).
59. Augustyn, V., Simon, P. & Dunn, B. Pseudocapacitive graphene-​coated carbon nanotube aerogels. ACS Appl. 108. Okubo, M. et al. Nanosize effect on high-​rate Li-​ion
oxide materials for high-​rate electrochemical energy Mater. Interfaces 9, 23810–23819 (2017). intercalation in LiCoO2 electrode. J. Am. Chem. Soc.
storage. Energy Environ. Sci. 7, 1597–1614 (2014). 83. Peng, L. et al. Ultrathin two-​dimensional 129, 7444–7452 (2007).
60. Ghodbane, O., Ataherian, F., Wu, N.-L. & Favier, F. MnO2/graphene hybrid nanostructures for high-​ 109. Rauda, I. E. et al. Nanostructured pseudocapacitors
In situ crystallographic investigations of charge performance, flexible planar supercapacitors. based on atomic layer deposition of V2O5 onto
storage mechanisms in MnO2-based electrochemical Nano Lett. 13, 2151–2157 (2013). conductive nanocrystal-​based mesoporous ITO
capacitors. J. Power Sources 206, 454–462 (2012). 84. Jabeen, N. et al. Enhanced pseudocapacitive scaffolds. Adv. Funct. Mater. 24, 6717–6728 (2014).
61. Crosnier, O. et al. Polycationic oxides as potential performance of α-​MnO2 by cation preinsertion. 110. Tang, Y.-J. et al. Molybdenum disulfide/nitrogen-​doped
electrode materials for aqueous-​based electrochemical ACS Appl. Mater. Interfaces 8, 33732–33740 reduced graphene oxide nanocomposite with enlarged
capacitors. Curr. Opin. Electrochem. 9, 87–94 (2018). (2016). interlayer spacing for electrocatalytic hydrogen
62. Yu, X. et al. Emergent pseudocapacitance of 2D 85. Cheng, F. et al. Facile controlled synthesis of MnO2 evolution. Adv. Energy Mater. 6, 1600116 (2016).
nanomaterials. Adv. Energy Mater. 8, 1702930 nanostructures of novel shapes and their application 111. Wang, R. et al. Elucidating the intercalation
(2018). in batteries. Inorg. Chem. 45, 2038–2044 (2006). pseudocapacitance mechanism of MoS2–carbon
63. Trasatti, S. & Buzzanca, G. Ruthenium dioxide: a new 86. Wang, X. & Li, Y. Selected-​control hydrothermal monolayer interoverlapped superstructure: toward
interesting electrode material. Solid state structure synthesis of α- and β-​MnO2 single crystal nanowires. high-​performance sodium-​ion-based hybrid
and electrochemical behaviour. J. Electroanal. Chem. J. Am. Chem. Soc. 124, 2880–2881 (2002). supercapacitor. ACS Appl. Mater. Interfaces 9,
Interfacial Electrochem. 29, A1–A5 (1971). 87. Chen, D. et al. Probing the charge storage mechanism 32745–32755 (2017).
This is the original report of the characterization of a pseudocapacitive MnO2 electrode using in 112. Liang, Y. et al. Interlayer-​expanded molybdenum
of RuO2 pseudocapacitive behaviour. operando Raman spectroscopy. Chem. Mater. 27, disulfide nanocomposites for electrochemical
64. Galizzioli, D., Tantardini, F. & Trasatti, S. Ruthenium 6608–6619 (2015). magnesium storage. Nano Lett. 15, 2194–2202
dioxide: a new electrode material. I. Behaviour in acid 88. Kuo, S.-L. & Wu, N.-L. Investigation of pseudocapacitive (2015).
solutions of inert electrolytes. J. Appl. Electrochem. 4, charge-​storage reaction of MnO2⋅nH2O supercapacitors 113. Chen, K., Pan, W. & Xue, D. Phase transformation
57–67 (1974). in aqueous electrolytes. J. Electrochem. Soc. 153, of Ce3+-doped MnO2 for pseudocapacitive electrode
65. Hadzi-Jordanov, S. Reversibility and growth behavior A1317–A1324 (2006). materials. J. Phys. Chem. C 120, 20077–20081
of surface oxide films at ruthenium electrodes. 89. Wang, G.-X., Zhang, B.-L., Yu, Z.-L. & Qu, M.-Z. (2016).
J. Electrochem. Soc. 125, 1471–1480 (1978). Manganese oxide/MWNTs composite electrodes for 114. Hu, Z. et al. Al-​doped α-​MnO2 for high mass-​loading
66. Zheng, J. P. Hydrous ruthenium oxide as an electrode supercapacitors. Solid State Ion. 176, 1169–1174 pseudocapacitor with excellent cycling stability.
material for electrochemical capacitors. J. Electrochem. (2005). Nano Energy 11, 226–234 (2015).
Soc. 142, 2699–2703 (1995). 90. Yang, Y., Xiao, L., Zhao, Y. & Wang, F. Hydrothermal 115. Sun, R. et al. Novel layer-​by-layer stacked VS2
67. Conway, B. E. & Pell, W. G. Double-​layer and synthesis and electrochemical characterization of nanosheets with intercalation pseudocapacitance for
pseudocapacitance types of electrochemical capacitors α-​MnO2 nanorods as cathode material for lithium high-​rate sodium ion charge storage. Nano Energy 35,
and their applications to the development of hybrid batteries. Int. J. Electrochem. Sci. 3, 67–74 (2008). 396–404 (2017).

www.nature.com/natrevmats
Reviews

116. Hu, Z. et al. 2D vanadium doped manganese dioxides 138. Kayali, E., VahidMohammadi, A., Orangi, J. & 159. Cresce, A. V. et al. Solvation behavior of carbonate-​
nanosheets for pseudocapacitive energy storage. Beidaghi, M. Controlling the dimensions of 2D MXenes based electrolytes in sodium ion batteries. Phys. Chem.
Nanoscale 7, 16094–16099 (2015). for ultrahigh-​rate pseudocapacitive energy storage. Chem. Phys. 19, 574–586 (2017).
117. Zhai, D. et al. A study on charge storage mechanism ACS Appl. Mater. Interfaces 10, 25949–25954 160. Okoshi, M., Yamada, Y., Yamada, A. & Nakai, H.
of α-​MnO2 by occupying tunnels with metal cations (2018). Theoretical analysis on de-​solvation of lithium, sodium,
(Ba2+, K+). J. Power Sources 196, 7860–7867 (2011). 139. Hu, M. et al. High-​capacitance mechanism for Ti3C2Tx and magnesium cations to organic electrolyte solvents.
118. Kim, H.-S. et al. Oxygen vacancies enhance MXene by in situ electrochemical Raman spectroscopy J. Electrochem. Soc. 160, A2160–A2165 (2013).
pseudocapacitive charge storage properties of MoO3−x. investigation. ACS Nano 10, 11344–11350 (2016). 161. Xie, X. et al. Porous heterostructured MXene/carbon
Nat. Mater. 16, 454–460 (2017). 140. Tang, Q., Zhou, Z. & Shen, P. Are MXenes promising nanotube composite paper with high volumetric
119. Shin, J.-Y., Joo, J. H., Samuelis, D. & Maier, J. anode materials for Li ion batteries? Computational capacity for sodium-​based energy storage devices.
Oxygen-​deficient TiO2−δ nanoparticles via hydrogen studies on electronic properties and Li storage Nano Energy 26, 513–523 (2016).
reduction for high rate capability lithium batteries. capability of Ti3C2 and Ti3C2X2 (X=F, OH) monolayer. 162. Wang, X. et al. Influences from solvents on charge
Chem. Mater. 24, 543–551 (2012). J. Am. Chem. Soc. 134, 16909–16916 (2012). storage in titanium carbide MXenes. Nat. Energy 4,
120. Li, B. & Xia, D. Anionic redox in rechargeable lithium 141. Hong, M. S., Lee, S. H. & Kim, S. W. Use of KCl aqueous 241–248 (2019).
batteries. Adv. Mater. 29, 1701054 (2017). electrolyte for 2 V manganese oxide/activated carbon This paper reports the performance of a state-​
121. Grimaud, A., Hong, W. T., Shao-​Horn, Y. & hybrid capacitor. Electrochem. Solid-​State Lett. 5, of-the-​art, high-​power hybrid supercapacitor device
Tarascon, J.-M. Anionic redox processes for A227–A230 (2002). based on a MXene as the redox electrode.
electrochemical devices. Nat. Mater. 15, 121–126 142. Hwang, J. Y. et al. Direct preparation and processing 163. Assat, G. & Tarascon, J.-M. Fundamental
(2016). of graphene/RuO2 nanocomposite electrodes for high-​ understanding and practical challenges of anionic
122. Mefford, J. T., Hardin, W. G., Dai, S., Johnston, K. P. & performance capacitive energy storage. Nano Energy redox activity in Li-​ion batteries. Nat. Energy 3,
Stevenson, K. J. Anion charge storage through oxygen 18, 57–70 (2015). 373–386 (2018).
intercalation in LaMnO3 perovskite pseudocapacitor 143. Wu, Z.-S. et al. High-​energy MnO2 nanowire/graphene 164. Grayfer, E. D., Pazhetnov, E. M., Kozlova, M. N.,
electrodes. Nat. Mater. 13, 726–732 (2014). and graphene asymmetric electrochemical capacitors. Artemkina, S. B. & Fedorov, V. E. Anionic redox
123. Liu, Y. et al. Highly defective layered double perovskite ACS Nano 4, 5835–5842 (2010). chemistry in polysulfide electrode materials for
oxide for efficient energy storage via reversible 144. Wang, Y.-G., Wang, Z.-D. & Xia, Y.-Y. An asymmetric rechargeable batteries. ChemSusChem 10,
pseudocapacitive oxygen-​anion intercalation. supercapacitor using RuO2/TiO2 nanotube composite 4805–4811 (2017).
Adv. Energy Mater. 8, 1702604 (2018). and activated carbon electrodes. Electrochim. Acta 50, 165. Goodenough, J. B. & Kim, Y. Locating redox couples
124. Zhu, L. et al. Perovskite SrCo0.9Nb0.1O3−δ as an anion-​ 5641–5646 (2005). in the layered sulfides with application to Cu[Cr2]S4.
intercalated electrode material for supercapacitors 145. Lukatskaya, M. R., Dunn, B. & Gogotsi, Y. J. Solid State Chem. 182, 2904–2911 (2009).
with ultrahigh volumetric energy density. Angew. Chem. Multidimensional materials and device architectures 166. Moreau, P., Ouvrard, G., Gressier, P., Ganal, P. &
Int. Ed. 55, 9576–9579 (2016). for future hybrid energy storage. Nat. Commun. 7, Rouxel, J. Electronic structures and charge transfer
125. Choi, D., Blomgren, G. E. & Kumta, P. N. Fast and 12647 (2016). in lithium and mercury intercalated titanium disulfides.
reversible surface redox reaction in nanocrystalline 146. Long, J. W. et al. Asymmetric electrochemical J. Phys. Chem. Solids 57, 1117–1122 (1996).
vanadium nitride supercapacitors. Adv. Mater. 18, capacitors—stretching the limits of aqueous electrolytes. 167. Subramanian, V., Zhu, H. & Wei, B. Synthesis and
1178–1182 (2006). MRS Bull. 36, 513–522 (2011). electrochemical characterizations of amorphous
126. Wang, B., Chen, Z., Lu, G., Wang, T. & Ge, Y. 147. Ra, E. J., Raymundo-​Piñero, E., Lee, Y. H. & Béguin, F. manganese oxide and single walled carbon nanotube
Exploring electrolyte preference of vanadium nitride High power supercapacitors using polyacrylonitrile-​ composites as supercapacitor electrode materials.
supercapacitor electrodes. Mater. Res. Bull. 76, based carbon nanofiber paper. Carbon 47, 2984–2992 Electrochem. Commun. 8, 827–832 (2006).
37–40 (2016). (2009). 168. Shi, P. et al. Design of amorphous manganese oxide@
127. Pande, P., Rasmussen, P. G. & Thompson, L. T. 148. Amatucci, G. G., Badway, F., Du Pasquier, A. & multiwalled carbon nanotube fiber for robust solid-​state
Charge storage on nanostructured early transition Zheng, T. An asymmetric hybrid nonaqueous energy supercapacitor. ACS Nano 11, 444–452 (2017).
metal nitrides and carbides. J. Power Sources 207, storage cell. J. Electrochem. Soc. 148, A930–A939 169. Kwon, B. J. et al. Effect of passivating shells on
212–215 (2012). (2001). the chemistry and electrode properties of LiMn2O4
128. Cong, X. et al. Intrinsic charge storage capability of 149. Naoi, K., Ishimoto, S., Isobe, Y. & Aoyagi, S. nanocrystal heterostructures. ACS Appl. Mater.
transition metal dichalcogenides as pseudocapacitor High-​rate nano-​crystalline Li4Ti5O12 attached on Interfaces 11, 3823–3833 (2019).
electrodes. J. Phys. Chem. C 119, 20864–20870 carbon nano-​fibers for hybrid supercapacitors. 170. Kim, U-​H., Myung, S-​T., Yoon, C. S. & Sun, Y-​K.
(2015). J. Power Sources 195, 6250–6254 (2010). Extending the battery life using an Al-​doped
129. Cook, J. B. et al. Pseudocapacitive charge storage in 150. Naoi, K., Ishimoto, S., Miyamoto, J. & Naoi, W. Li[Ni0.76Co0.09Mn0.15]O2 cathode with concentration
thick composite MoS2 nanocrystal-​based electrodes. Second generation ‘nanohybrid supercapacitor’: gradients for lithium ion batteries. ACS Energy Lett. 2,
Adv. Energy Mater. 7, 1601283 (2017). evolution of capacitive energy storage devices. 1848–1854 (2017).
130. Mahmood, Q. et al. Unveiling surface redox Energy Environ. Sci. 5, 9363–9373 (2012). 171. Ding, Z. et al. Understanding the enhanced kinetics of
charge storage of interacting two-​dimensional 151. Jeżowski, P. et al. Safe and recyclable lithium-​ion gradient-​chemical-doped lithium-​rich cathode material.
heteronanosheets in hierarchical architectures. capacitors using sacrificial organic lithium salt. ACS Appl. Mater. Interfaces 9, 20519–20526 (2017).
Nano Lett. 15, 2269–2277 (2015). Nat. Mater. 17, 167–173 (2017). 172. Weber, R., Fell, C. R., Dahn, J. R. & Hy, S. Operando
131. Muller, G. A., Cook, J. B., Kim, H.-S., Tolbert, S. H. 152. Tran, T. D., Feikert, J. H., Pekala, R. W. & Kinoshita, K. X-​ray diffraction study of polycrystalline and single-​
& Dunn, B. High performance pseudocapacitor Rate effect on lithium-​ion graphite electrode crystal LixNi0.5Mn0.3Co0.2O2. J. Electrochem. Soc. 164,
based on 2D layered metal chalcogenide nanocrystals. performance. J. Appl. Electrochem. 26, 1161–1167 A2992–A2999 (2017).
Nano Lett. 15, 1911–1917 (2015). (1996). 173. Grenier, A. et al. Reaction heterogeneity in
132. Shuai, J. et al. Density functional theory study of 153. Sivakkumar, S. R. & Pandolfo, A. G. Evaluation of LiNi0.8Co0.15Al0.05O2 induced by surface layer. Chem.
Li, Na, and Mg intercalation and diffusion in MoS2 lithium-​ion capacitors assembled with pre-​lithiated Mater. 29, 7345–7352 (2017).
with controlled interlayer spacing. Mater. Res. Express graphite anode and activated carbon cathode. 174. Conway, B. E. & Gileadi, E. Kinetic theory of pseudo-​
3, 064001 (2016). Electrochim. Acta 65, 280–287 (2012). capacitance and electrode reactions at appreciable
133. Xu, X. et al. Controllable design of MoS2 nanosheets 154. Naoi, K. et al. Ultrafast nanocrystalline-​TiO2(B)/carbon surface coverage. Trans. Faraday Soc. 58, 2493–2509
anchored on nitrogen-​doped graphene: toward fast nanotube hyperdispersion prepared via combined (1962).
sodium storage by tunable pseudocapacitance. ultracentrifugation and hydrothermal treatments for
Adv. Mater. 30, 1800658 (2018). hybrid supercapacitors. Adv. Mater. 28, 6751–6757 Acknowledgements
134. Anasori, B., Lukatskaya, M. R. & Gogotsi, Y. 2D metal (2016). The authors are grateful to the US Office of Naval Research
carbides and nitrides (MXenes) for energy storage. 155. Zhao, Y. et al. Fabrication of nanoarchitectured (grants nos. N00014-17-1-2244 and N00014-16-1-2164)
Nat. Rev. Mater. 2, 16098 (2017). TiO2(B)@C/rGO electrode for 4 V quasi-​solid-state for supporting their research.
135. Mashtalir, O. et al. Intercalation and delamination of nanohybrid supercapacitors. Electrochim. Acta 258,
layered carbides and carbonitrides. Nat. Commun. 4, 343–352 (2017). Author contributions
1716 (2013). 156. Kong, L. et al. High-​power and high-​energy asymmetric All authors researched data for the article. C.C., D.S.A., D.M.B.,
136. Du, F. et al. Environmental friendly scalable production supercapacitors based on Li+-intercalation into R.H.D. and B.D. contributed to the discussion of content and
of colloidal 2D titanium carbonitride MXene with a T-​Nb2O5/graphene pseudocapacitive electrode. writing and editing of the article prior to submission.
minimized nanosheets restacking for excellent cycle J. Mater Chem. A. 2, 17962–17970 (2014).
life lithium-​ion batteries. Electrochim. Acta 235, 157. Sun, H. et al. Three-​dimensional holey-​graphene/ Competing interests
690–699 (2017). niobia composite architectures for ultrahigh-​rate The authors declare no competing interests.
137. Lian, P. et al. Alkalized Ti3C2 MXene nanoribbons with energy storage. Science 356, 599–604 (2017).
expanded interlayer spacing for high-​capacity sodium 158. Myung, S.-T., Amine, K. & Sun, Y.-K. Nanostructured Publisher’s note
and potassium ion batteries. Nano Energy 40, 1–8 cathode materials for rechargeable lithium batteries. Springer Nature remains neutral with regard to jurisdictional
(2017). J. Power Sources 283, 219–236 (2015). claims in published maps and institutional affiliations.

Nature Reviews | Materials

You might also like