You are on page 1of 59

Progress in Materials Science 63 (2014) 58–116

Contents lists available at ScienceDirect

Progress in Materials Science


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

Deformation and stress in electrode materials


for Li-ion batteries
Amartya Mukhopadhyay a,⇑, Brian W. Sheldon b
a
High Temperature and Energy Materials Laboratory, Department of Metallurgical Engineering and Materials Science, IIT Bombay,
Powai, Mumbai 400076, India
b
School of Engineering, Brown University, Providence, RI 02912, USA

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

Article history: Structural stability and mechanical integrity of electrode materials


Received 5 April 2013 during lithiation/delithiation influence the performance of Li-ion
Received in revised form 10 January 2014 batteries. Significant dimensional and volume changes are associ-
Accepted 28 January 2014
ated with variations in lattice parameters and transformations of
Available online 6 February 2014
crystalline/amorphous phases that occur during electrochemical
cycling. These phenomena, which occur during Li-intercalation/
deintercalation, Li-alloying/dealloying and conversion reactions,
result in deformations and stress generation in the active cathode
and anode materials. Such stresses can cause fragmentation, disin-
tegration, fracturing, and loss in contact between current collectors
and the active electrode materials, all of which can also expose
fresh surfaces to the electrolyte. These degradation processes ulti-
mately lead to capacity fade with electrochemical cycling for
nearly all electrode materials, and are some of the major causes
for the eventual failure of a Li-ion cell. Furthermore, severe stresses
have made it nearly impossible to use higher capacity anode mate-
rials (e.g., Si, Sn) in practical batteries and also limit the ‘usable’
capacity of the present cathode materials (e.g., LiCoO2, LiMn2O4)
to nearly half the theoretical capacity. Against this backdrop, this
review presents an overview of the causes and the relative magni-
tudes of stresses in the various electrode materials, highlights
some of the more recent discoveries concerning the causes (such
as stress development due to passivation layer formation), intro-
duces the recently developed techniques for in situ observations
of lithiation induced deformations and measurement of stresses,
analyses the strategies adopted for addressing the stress-related

⇑ Corresponding author. Tel.: +91 22 25767612; fax: +91 22 25726975.


E-mail address: amartya_mukhopadhyay@iitb.ac.in (A. Mukhopadhyay).

http://dx.doi.org/10.1016/j.pmatsci.2014.02.001
0079-6425/Ó 2014 Elsevier Ltd. All rights reserved.
A. Mukhopadhyay, B.W. Sheldon / Progress in Materials Science 63 (2014) 58–116 59

issues, and raises various issues that still need to be addressed to


overcome the stress related problems that are some of the major
bottlenecks towards the development of new high-capacity elec-
trode materials for Li-ion batteries.
Ó 2014 Elsevier Ltd. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
2. Li-ion battery – components, basic principles and operation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
3. Stress development in electrode materials – causes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
3.1. Stresses arising from lithiation/delithiation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
3.1.1. Anode materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
3.1.2. Cathode materials. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
3.2. Stress due to surface reactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
4. Effects of stress development in electrode materials. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
4.1. Impact on battery performance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
4.1.1. Effects of stress on mechanical integrity of electrodes . . . . . . . . . . . . . . . . . . . . . . . . . . 79
4.1.2. Effects of stress on thermodynamics of the electrochemical phenomena . . . . . . . . . . . 81
4.2. Impact on further progress in Li-ion battery technology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
5. Real time observations of electrode material deformations/dilations during lithiation/delithiation . . . 84
6. Experimental determination of stresses in electrode materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
6.1. In situ determination of stress development using multi-beam optical stress sensor (MOSS) . . . 89
6.2. Experimentally measured stresses in various electrode materials using MOSS . . . . . . . . . . . . . . 90
6.2.1. Si thin film electrodes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
6.2.2. Carbon based thin film electrodes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
6.3. Other techniques for experimental determination of stresses/strains in electrode materials . . . 92
7. Assessment of stresses in electrode materials via computational research . . . . . . . . . . . . . . . . . . . . . . . 93
8. Stress management in electrode materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98
8.1. Formation of alloys and composites. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98
8.2. Formation of nanostructures and thin films . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
8.3. Limiting the usable Li-capacity. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
9. Conclusion and outlook . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110

1. Introduction

Li-ion batteries possess the highest energy densities (capability to store energy per unit weight or
volume) among all rechargeable batteries (see Fig. 1) [1–3]. Additionally, they typically have relatively
low self-discharge rates of 5% per month (pm), compared to >30% pm and 20% pm for Ni-metal hy-
dride batteries and Ni–Cd batteries, respectively. They are devoid of ‘memory effects’, are more envi-
ronmentally friendly compared to most other batteries, and can also operate over a wider temperature
range (25 °C to +50 °C) [1–3]. Presently, Li-ion batteries are the dominant power source for most
portable electronic devices such as cell phones, digital cameras, laptops, tablet e-book readers etc. In
order to address issues associated with the depletion of non-renewable energy sources (in particular,
fossil fuels), researchers worldwide are also contemplating the potential usage of such high capacity
energy storage device in more advanced and heavy duty applications, including automobiles (hybrid
electric vehicles and total electric vehicle), and in efficient storage/utilization of intermittent renew-
able energies like solar and wind energies [1–6]. This has instigated extensive research activities on
Li-ion batteries worldwide in the last few years. Since the materials used for the electrodes and elec-
trolytes are at the core of functioning and performance of Li-ion batteries, recent research has been
60 A. Mukhopadhyay, B.W. Sheldon / Progress in Materials Science 63 (2014) 58–116

Fig. 1. (a) Plot comparing the energy densities (y-axis) and specific energies (x-axis) of the different battery systems [1]. Note
that Li-ion batteries lie near the top of the plot. (b) Ragone plot comparing the energy densities (x-axis) and power densities (y-
axis) of the various electrochemical energy storage systems [94]. Note that Li-ion battery suffers from relatively poor power
density, especially when compared to electrochemical capacitors.

devoted towards understanding and resolving issues associated with both; the currently used mate-
rials, and the wide range of potential higher performance materials.
In a broader sense, the primary requirements for rendering Li-ion batteries suitable for advanced
applications are further improvements in the energy densities (i.e. increase Li-storage capacities of
electrodes), in the cycle lives (i.e. minimize capacity fade with electrochemical cycling) and in the rate
capabilities (i.e. allow faster Li-insertion/removal into/from the electrodes). Higher energy densities
[7–52] can be obtained by replacing the presently used electrode materials with materials possessing
higher Li-capacities or by increasing the usable Li-capacities of the presently used electrode materials
(here ‘capacities’ denote the charge storage capabilities; Li+ in this case). Cycle lives can be improved
[10–12,14,15,17,19,24,29,30,33–38,46–53] by preventing/minimizing the degradation of the electrode
materials or build-up of impedance (which are often interrelated) with electrochemical cycling. Rate
A. Mukhopadhyay, B.W. Sheldon / Progress in Materials Science 63 (2014) 58–116 61

capabilities depend critically on the diffusivity of Li-ions and the conductivity of electrons in the
various phases (electrodes, electrolytes and also circuit connections), and also across key interfaces.
Attempts to improve rate capabilities include using of novel electrode materials that possess higher
Li-diffusivities, reducing the dimensions of the electrodes (to increase surface area and reduce
diffusion lengths), careful engineering of crystal structures to allow easier access of Li-ions to the
intercalation paths, and modifying electrode/electrolyte interfaces [49–59].
The logical avenues for improvements are based on each of the three aforementioned areas, which
also presently form the major thrust areas for research on Li-ion batteries. Interestingly, all these areas
are potentially related to the stress developments in electrode materials during electrochemical
cycling, when Li-ions are repeatedly inserted into and removed from the electrode materials
[9–15,17,19–26,29–40,46–54,60–102]. Section 2 provides more details on the operation principles
of Li-ion batteries. From the broader perspective of materials science and engineering, the electro-
chemical cycling can be visualized as repeated insertion and removal of foreign ions (Li+) into/from
the host materials (active electrodes). The actual mechanisms of such insertion/removal reactions
are dependent on the type of electrode material under consideration, and can be classified into
(i) intercalation/deintercalation (as in graphitic carbon anode and layered cathodes such as LiCoO2,
LiMn2O4); (ii) Li-alloying/de-alloying (as in potential metallic anodes such as Si, Sn, Al, Sb, Ge etc.;
which include solid solution formation, as well as formation of new intermetallics); (iii) conversion
reactions (as in potential anodes such as Fe3O4, Co3O4 etc.).
The stresses in the electrode materials arise primarily due to changes in lattice dimensions,
changes in crystal structures and phase transformations involving both crystalline and amorphous
phases, which accompany the insertion/removal reactions. These effects are associated with overall
volume changes that lead to deformation of the electrode materials against the constraining effects
of the inactive regions of the electrodes or other parts of the cell and mismatches between regions
with different Li-concentrations, possessing different molar volumes (see Section 3 for more detailed
illustrations of the mechanisms leading to stress development). Mechanical instabilities, including
plastic deformation, fragmentation, disintegration and fracturing, and concomitant loss in contact of
the electrode material with the current collector, result in severe capacity fade with electrochemical
cycling, which ultimately render the electrode material unsuitable for further cycling. Hence, stress
development in electrode materials is one of the primary causes for capacity fade and the eventual
failure of Li-ion batteries. Such stress development is often more pronounced at the faster electro-
chemical cycling rates [9,40,60–62,73], and thus contributes towards limiting the rate capabilities
of a battery. In the presently used cathode materials (LiCoO2, LiMn2O4), lattice instabilities and stres-
ses associated with phase transformations and lattice parameter changes during Li-deintercalation/
intercalation limit the usable capacities to nearly half of the theoretical capacities [20–26,29–39].
The deformation and stress development are also the major bottlenecks towards improving the capac-
ity of the anode side, since the potential metallic materials (such as Si, Sn, Al) with very high Li-capac-
ities (3–10 times higher than the presently used graphitic carbons) retain their capacities and
cyclability for only few cycles because of the large volume changes (up to 400%) during Li-alloy-
ing/de-alloying [8–15,17,46–53,65–68,73–88]. Such stress developments due to dimensional changes
are even less forgiving in batteries having solid electrolytes (all-solid-state batteries), where the solid
electrolyte provides additional constraining effect [5].
Because stress development has a major impact on the performance of Li-ion batteries, there have
been extensive research activities devoted towards modeling the stresses that get developed under
different conditions [40,52,60,63,65–89] and for devising strategies for minimization/buffering of
the stresses [9–12,15,23,24,33,34,47–53,98–118]. More recently, efforts have also been directed to-
wards in situ observations of lithiation/delithiation induced deformation/structure/volume changes
[64,92–99,119,120], using transmission electron microscopy, X-ray transmission microscopy, AFM
and NMR. Though leading to very useful understanding, none of these methods provide information
on the magnitude of stresses that arise in different electrode materials. It must be mentioned here that
experimental measurement of the stresses, accompanied with analysis and interpretation of the inter-
related chemical and stress induced changes in a controlled fashion is extremely challenging with
electrodes with complex porous microstructures. Against this backdrop, the authors and few other
researchers [19,53,54,117,121–126] have recently employed optical stress sensors [127,128] to make
62 A. Mukhopadhyay, B.W. Sheldon / Progress in Materials Science 63 (2014) 58–116

in situ measurements of stress generation during electrochemical cycling of thin film electrode
materials.
In spite of the importance of stress development in electrode materials, to-date there is no compre-
hensive review that provides an overview of the causes and effects of such stresses and the measures
adopted for alleviating the deleterious effects of such stress developments. Hence, after briefly intro-
ducing the composition/architecture/operational aspects of Li-ion batteries in Section 2, this review
presents a comprehensive survey and critical analysis of the causes for stress development during
electrochemical cycling of a variety of electrode materials in Section 3. Some of the more recent
and notable findings with respect to the causes for stress developments are also highlighted. Attempt
has been made to establish the correlation between the crystal structures, phase developments, mech-
anisms of lithiation/delithiation, deformations and stress development. Following this, the effects of
such stresses on battery performance and on the further development of Li-ion battery technology
are elucidated in Section 4. Section 5 summarizes some of the recent work that studies deformations
induced by lithiation/delithiation of electrode materials in situ using transmission electron micro-
scope, X-ray transmission microscope, AFM, as well as NMR. This work provides important insights
into the mechanisms associated with deformation and stress development, which are difficult to ob-
tain with ex situ methods. Section 6 then presents an overview of the development of stress measure-
ment techniques (for real-time measurements), along with some of the notable findings of this
research. Section 7 presents critical analysis of the magnitudes, the causes/mechanisms of stress
developments, and conditions that can be employed to minimize stresses, as well as avert crack for-
mation/propagation, as obtained via computational research. A major fraction of the stress-related re-
search on Li-ion batteries is being directed towards minimizing deleterious effects of such stress
developments, and accordingly, Section 8 surveys various strategies that have been explored/adopted
for buffering stresses, alleviating the magnitude of the stresses, and in turn minimizing the associated
negative impacts. Finally, the review concludes in Section 9 by summarizing the important aspects re-
lated to stress development in electrode materials, raising issues which remain unresolved to date and
suggesting some possible ways forward to resolve such issues.

2. Li-ion battery – components, basic principles and operation

Li-ion batteries employ the same basic operating principles used in other rechargeable batteries
(such as Pb–acid, Ni–Cd or Ni–MH), with the primary difference being the use of Li-ions as the charge
carrier that shuttles between the anode and the cathode via the electrolyte during the discharge/
charge cycles. The main components of a typical Li-ion cell are the two electrodes (anode and cath-
ode), an electrolyte that is an ionic conductor and an electronic insulator (typically Li-salts dissolved
in organic solvents [130]), and a separator that is permeable to the flow of Li-ions (usually porous
polymer films such as micro-porous polyolefin membrane [19,129]). The ‘active materials’ of the elec-
trodes are connected to an external circuit via metallic current collectors. The current then flows
through the external circuit to which the load is connected. The ‘active materials’ of the electrodes
are elements/compounds/composites that can reversibly intercalate/de-intercalate (layered oxides/
graphite) or alloy/‘de-alloy’ (metals) with Li in response to the voltage. During discharging of a cell,
the Li is released from the anode as Li+ ions (in oxidized state), that flow through the electrolyte via
the separator and are taken-up by the cathode, where they are reduced. Simultaneously the electronic
current flows through the external circuit from the anode to the cathode (via the respective current
collectors). The reverse happens during the charging of the cell. The overall cell reaction is shown
in Eq. (1) (anodic) and 2 (cathodic), where left to right denotes the discharge of cell and vice versa.
Schematic representations of the components and the basic working principle of a Li-ion cell are
shown in Fig. 2a.
þ
Anode-Li $ Anode þ Li þ e ð1Þ

þ
Cathode þ Li þ e $ Cathode-Li ð2Þ
A. Mukhopadhyay, B.W. Sheldon / Progress in Materials Science 63 (2014) 58–116 63

Fig. 2. Schematic representations of (a) Li-ion full cell, showing the components and the directions of flow of Li+ ions and e,
corresponding to the discharge of the cell [image partly adapted from PRLog (Press Release) - May 12, 2011]; (b) typical
electrode architectures for the more common ‘porous composite electrode’ and the thin film electrode.

The typical electrode architectures are twofold (see Fig. 2b): (a) Particles of the active materials
(Li-hosts) are mixed with binders (usually PVD) and conductive carbon additives (usually carbon
black) to form the electrode. Such electrodes are prevalent in commercial batteries and will be
referred to as ‘porous-composite electrodes’ throughout the text. The other type of electrodes consid-
ered is (b) thin film electrodes consisting of films of the active materials from nanometers to microns
in thickness. These are typically used for miniature batteries, and because they are also have a simpler
geometry and are devoid of inactive components (e.g., binders and carbon black), they are sometimes
better suited for scientific investigation and understanding various fundamental phenomena [19,123].
Different novel electrode architectures are also currently being explored to improve the performance
(such as nano-rods [49,105,107,118,119], islands [52,53,102,117], etc.). In all of these cases, the active
electrode components [as described in (a) and (b)] lie on metallic current collectors (usually Cu or Ni
for anodes and Al or Pt for cathodes), that are connected to the external circuit (as shown in Fig. 2).
The present generation of Li-ion batteries use graphitic carbon as negative electrodes (anode)
and layered transition metal oxides (LiCoO2) or phosphates (LiFePO4) as positive electrodes, which
result in a net energy density of between 100–150 W h/kg (cell capacity 70 mAh/g [9]), that is higher
than any other electrochemical energy storage system (see Fig. 1). The cycle lives of the present
generation of batteries typically lie between 500–1500 cycles [1–5,131]. For further improvements
in the energy densities of Li-ion cells, research is also underway to employ metallic anode materials
64 A. Mukhopadhyay, B.W. Sheldon / Progress in Materials Science 63 (2014) 58–116

(such as Si, Sn, Al) that possess Li-capacities which exceed that of graphitic carbon by factors of 3–10
[1,4,8–15,17,46–53,90–107]. However, for reasons that will be highlighted and critically analyzed in the
following sections, the development of these metallic anodes have been hampered by severe degradation
(capacity fade) that usually limits the cycle life to less than 50 cycles [1,4,8–15,17,46–53,90–107].
Another property, with respect to which the present generation Li-ion batteries are inferior compared
to some other electrochemical energy storage technologies (in particular, supercapacitors) is
power density (i.e., the ability to charge/discharge at high rates, as shown in Fig. 1b). Also high
charge/discharge rates typically reduce the cycle lives of Li-ion batteries [56–63], which hints at some
correlation between the mechanisms that are responsible for capacity fade and poor rate capabilities.
More thorough understanding of these issues will promote further improvements in the performance
of Li-ion batteries in general and render them suitable for more demanding applications, such as
electric vehicles. The electrode materials have a dominant impact on the overall performance of a
Li-ion cell, and thus the last decade has witnessed extensive research efforts to understand and
develop new electrode materials/compositions to address the aforementioned issues.

3. Stress development in electrode materials – causes

Stresses in electrode materials can occur due to the internal chemical processes taking place during
electrochemical cycling [9–15,17,19–26,29–40,46–54,60–102], and also due to external mechanical
loads acting on the battery pack during manufacturing [71,72]. The present review is concerned with
the stresses that are created by the electrochemical processes, that are fundamental towards
operation of a battery (internal stresses). As outlined in the previous section, the working principle
of Li-ion batteries involves repeated Li-ions insertion and removal from the electrode materials during
electrochemical cycling. The insertion/removal can result from different mechanisms depending on
the types of active electrode materials. For instance, in graphite or layered oxides such as LiCoO2,
the insertion and removal of the Li-ions occurs by intercalation and de-intercalation. The term
intercalation, as typically used in materials science or chemistry, can loosely be defined as reversible
insertion of guest species in-between crystallographic planes of the host crystal lattice, without
disturbing the arrangements of these planes [1,16,17,19,132,133]. On the other hand, metallic anode
materials such as Si, Sn, Al etc. allow intake and release of Li via alloying and de-alloying, forming
intermetallic compounds or solid solutions with Li [9–14]. In either case, the foreign Li ions require
temporary accommodation within the host lattice as they are repeatedly inserted and removed.
Accommodating these guest species during electrochemical cycling changes the spacing between
the crystallographic planes of the host structure, leading to concomitant dimensional changes. Such
insertion/removal also leads to phase transformations and changes in crystal symmetries in many
materials. The magnitudes and more specific causes of such stress development related to specific
materials are reviewed in Section 3.1. Additionally, several recent reports [54,125] have documented
other causes of stress that are associated with electrochemical cycling, but are not directly related to
the lithiation/delithiation phenomena. These phenomena are reviewed in Section 3.2. Section 3 will
thus shed light on the various causes of the stress development. The effects of these stress develop-
ments on the electrode materials and battery performance, at microscopic as well as macroscopic
levels, will then be described in Section 4.

3.1. Stresses arising from lithiation/delithiation

Insertion/removal of Li-ions into/from a host crystal lattice (electrode material) during electro-
chemical cycling of a Li-ion cell leads to a major fraction of the stresses in the electrode materials.
The stress magnitude varies over several orders of magnitude and depends on a number of factors
such as the particular electrode material, the electrode architecture, electrode composition, the mech-
anism of Li-insertion/removal, the electrochemical cycling rate and the potential range under consid-
eration. Furthermore, stresses are often not uniform throughout the active electrode material. The
major reasons for this stress generation can be broadly grouped into the mechanisms that are shown
schematically in Fig. 3 and summarized below:
A. Mukhopadhyay, B.W. Sheldon / Progress in Materials Science 63 (2014) 58–116 65

Fig. 3. Schematic illustrations of (a) constraining effects of neighboring active particles, inactive matrix and current collector/
substrate on the expanding active particles upon lithiation of ‘porous composite electrode’; (b) stress generated at inter-particle
contact between expanding electrode particles upon lithiation; (c) constraining of the in-plane dilation of thin film electrodes,
upon lithiation, by the inactive current collector/substrate; (d) Li-concentration gradient between lithiated and unlithiated
portions of a particle, resulting in the development of stress discontinuities. The same is true for adjacent regions possessing
different crystal structures/phases.

(a) Physical constraints on the dimensional changes in active materials: The major structural changes
that occur when ‘foreign’ Li-ions are inserted or removed from the host lattices of the active
electrode materials are lattice parameter variations [1,16,17,19,132–134]) or the formation of
new crystalline/amorphous phases [1,9–14,92–95,98,121–124]. More details about these mech-
anisms are discussed in the subsequent sub-sections. However, the major effects of these phe-
nomena are overall volume and/or morphologically changes. In most circumstances, such
dimensional/volume/morphology changes will be constrained by neighboring particles, by
the substrate/current collector, or by the space constraint/geometry of the cell (see Fig. 3a
and b). These constraining effects are the dominant sources of stress development in many elec-
trode materials. Table 1 lists unit cell volumes, as well as volume per host atom, of a variety of
important phases that form during lithiation.
Table 1

66
Phases, crystal structures, lattice parameters and crystal volumes (also volume changes) that form during progressive electrochemical lithiation (of anode materials) and de-lithiation (of
cathode materials) of the various electrode materials for Li-ion batteries [9,30,31,61,111,133–139,141,142].

Electrode Phase Space group (structure) Unit cell Unit cell % Increase in Volume per % Volume Theoretical
material parameters (Å) volume (Å3) unit cell volume host atom increase Li-capacity
(Å3) per host atom (mA h g1)
Graphite (C) For C and Li-GICs; el = edge length of the rhombohedral cell with saturated Li in-plane density; d[0 0 0 2] = average layer spacing of neighboring C planes and
V = average cell volume of the space defined by two neighboring C-planes (see Fig. 4c and Ref. [133])

A. Mukhopadhyay, B.W. Sheldon / Progress in Materials Science 63 (2014) 58–116


C P63/mmc (hexagonal) a = 2.464; el = 4.239 V = 51.38 0 (based on V) – – –
c = 6.711; d[0002] = 3.302
LiC18 (dilute stage II) P6/mmm (hexagonal) a = 2.47; el = 4.254 V = 54.37 5.8 (based on V) – – 71
c = 6.711; d[0002] = 3.469
LiC18 (stage III) P6/mmm (hexagonal) a = 4.288; el = 4.255 V = 53.22 3.6 (based on V) 105
c = 7.07; d[0002] = 3.395
LiC12 (stage II) P6/mmm (hexagonal) a = 4.288; el = 4.262 V = 53.76 4.6 (based on V) – – 185
c = 7.066; d[0002] = 3.417
p
LiC6 (stage I) P6/mmm (hexagonal) a = 4.305; 3a = 4.289 V = 56.51 10 (based on V) – – 372
c = 3.706; d[0002] = 3.547
Si Si  ðcubicÞ
Fd3m a = 5.472 163.84 0 19.6 0 –
LiSi I41/a (tetragonal) a = 9.357 503.071 31.4 60 954
b = 9.357
c = 5.746
Li12Si7 Pnma (orthorhombic) a = 8.532 2393.14 1360.66 43.5 122 1635
b = 19.612
c = 14.302
Li13Si4 Pbam (orthorhombic) a = 7.914 528.71 222.69 67.3 243 3100
b = 15.084
c = 4.429
Li15Si4  ðcubicÞ
4I3d a = 10,595 76.4 290 3590
Li22Si5 F23 (Cubic) a = 13.189 2294.22 1300.28 82.4 320 4200
Sn Sn I41/amd (tetragonal) a = 5.83 108.1 0 25.9 0 –
(for white Sn) c = 3.18
LiSn P2/m (monoclinic) a = 5.17 127.25 17.7 41.1 58.7 226
b = 7.74
c = 3.18
Li7Sn3 P21/m (monoclinic) a = 9.45 381.81 253.2 61.2 136.3 527
b = 8.56
c = 4.72
Li5Sn2  ðtrigonalÞ
R3m a = 4.74 445.53 312.1 64.3 148.3 564
c = 19.83
Li13Sn5 
P3m1 ðtrigonalÞ a = 4.70 379.94 251.5 65.5 152.9 587
c = 17.12
Li7Sn2 Cmmm (orthorhombic) a = 9.80 642.39 494.3 80.3 210 790
b = 13.80
c = 4.75
Li22Sn5 F23 (cubic) a = 19.78 7738.89 7059 96.7 273.4 993
Al Al Fm3m (cubic) a = 4.0495 66.4 0 16 0 –
LiAl Fd3m (cubic) a = 6.37 258.47 289.26 24.1 50.6 995
Li3Al2  ðtrigonalÞ
R3m a = 4.508 289.79 336.43 36.1 125.6 1490

A. Mukhopadhyay, B.W. Sheldon / Progress in Materials Science 63 (2014) 58–116


c = 14.26
Li9Al4 C2/m (monoclinic) a = 19.151 467.76 604.45 48.1 200.6 2235
b = 5.429
c = 4.499
LiCoO2 LiCoO2  ðtrigonalÞðHIÞ
R3m a = 2.817 96.75 0 – – 270
c = 14.058
Li0.9CoO2  ðtrigonalÞðHIÞ
R3m a = 2.817 96.8 0 240
c = 14.08
 ðtrigonalÞðH2Þ
R3m a = 2.814 97.25 0.5 (based on initial H1)
c = 14.19
Li0.78CoO2  ðtrigonalÞðH2Þ
R3m a = 2.812 97.75 1 (based on initial H1) 200
c = 14.25
Li0.51CoO2  ðtrigonalÞðH2Þ
R3m a = 2.812 98.0 1.3 (based on initial H1) 150
c = 14.3
Li0.5CoO2 C2/m (monoclinic) a = 2.813 98.6 1.9 (based on initial H1) 135
c = 14.42
Li0.45CoO2  ðtrigonalÞðH2Þ
R3m a = 2.809 98.3 0.3 (based on Li0.5CoO2) 120
c = 14.4
Li0.22CoO2  ðtrigonalÞðH2Þ
R3m a = 2.81 95.7 –2.9 (based on Li0.5CoO2) 65
c = 14
CoO2  ðtrigonalÞ
R3m a = 2.822 88.8 –9.9 (based on Li0.5CoO2) –
c = 12.879
LiMn2O4 Li2Mn2O4 I41/amd (tetragonal) a = 5.646 294.86 47.3 (based on LiMn2O4) – – 240
c = 9.25
LiMn2O4 Fd3m (cubic) a = 8.242 559.47 0 – – 120
Li0.5Mn2O4 Fd3m (cubic) a = 8.15 541.34 3.2 (based on LiMn2O4) 60
Li0.4Mn2O4 Phase I: Fd3m (cubic) a = 8.124 (fraction: 0.7) 536.18 4.2 (based on LiMn2O4) 48
Phase II: Fd3m (cubic) a = 8.103 (fraction: 0.3) 532.03 4.9 (based on LiMn2O4)
Li0.3Mn2O4 Phase I: Fd3m (cubic) a = 8.108 (fraction: 0.41) 533.02 4.7 (based on LiMn2O4) 36
Phase II: Fd3m (cubic) a = 8.08 (fraction: 0.59) 527.51 5.7 (based on LiMn2O4)
Li0.15Mn2O4 Phase II: Fd3m (cubic) a = 8.048 521.27 6.8 (based on LiMn2O4) 18
LiFePO4 LiFePO4 Pnma (orthorhombic) a = 10.33 291.2 0 – – 170
b = 6.01

(continued on next page)

67
68
Table 1 (continued)

Electrode Phase Space group (structure) Unit cell Unit cell % Increase in Volume per % Volume Theoretical
material parameters (Å) volume (Å3) unit cell volume host atom increase Li-capacity
(Å3) per host atom (mA h g1)
c = 4.69
FePO4 Pbnm (orthorhombic) a = 9.81 (5% shrinkage) 271.5 7.2 – – –
b = 5.79
c = 4.78 (1.9% elongation)

A. Mukhopadhyay, B.W. Sheldon / Progress in Materials Science 63 (2014) 58–116


A. Mukhopadhyay, B.W. Sheldon / Progress in Materials Science 63 (2014) 58–116 69

(b) Inter-particle contacts: As described in Section 2, the most prevalent electrode architectures cur-
rently consist of particles of active materials, along with binders and conducting additives (‘por-
ous composite electrodes’; Fig. 2b). Due to the increase in volume of the active particles during
lithiation, it is possible for them to come in contact with each other (see Fig. 3c), which can lead
to significant stresses at the points of contacts.
(c) Mismatch between crystalline phases and Li-concentration gradients: Phase transformations that
occur during Li-insertion/removal are associated with changes in the Li-content (i.e., generally
driven by changes in the electrochemical potential (usually with Li content)
[1,9,19,30,31,61,92–95,133,135–141]. Furthermore, since lithiation/delithiation occurs mainly
via diffusion of Li-ions through the bulk of the active material (from surface to core during
lithiation and vice versa), steep Li-concentration gradients can develop in the active material.
These differences depend on transport limitations that are associated with Li-diffusivity and
current density, and are thus more substantial at higher electrochemical cycling rates
[61–63,67,68,142,76,80,85]. These Li-concentration gradients lead to differential dimensional/
volume changes and phase transformations from the surface to the core, which implies that
at any instant, adjacent regions within the same active material may have different crystal
phases (structures) and different molar volumes. Furthermore, these different co-existing
phases can also possess different elastic properties [126,133,141]. Interaction/contact between
such regions within a continuum leads to mismatch induced stress development (see Fig. 3d).
Table 1 lists the crystal structures, as well as elastic moduli of various phases that form during
lithiation [30,31,61,133,135–143].

It must be noted that the lithiation/delithiation occurring during electrochemical cycling is dy-
namic and thus leads to repeated stress reversals. Such repeated reversals can further increase the
severity of damage accumulation due to these stresses. Analogies can be drawn to the more general
fatigue mechanisms, which describe the increasing severity of the effects of stresses due to cyclic load-
ing/unloading. The following sections provide further elucidation of the causes and issues associated
with lithiation/delithiation induced stresses, focusing on particular electrode materials.

3.1.1. Anode materials


Graphitic carbon has been the most widely used anode material since the initial development of Li-
ion batteries in 1991. These materials can be visualized as consisting of stacks of graphene sheets [or
graphite basal planes (0 0 0 2)] along the crystallographic c-axis, with 0.34 nm interatomic spacing
between the graphene sheets (i.e., 0 0 0 2 spacing). During electrochemical cycling, Li-ions are revers-
ibly intercalated in the spaces between the graphite basal planes, forming staged Li–graphite interca-
lation compounds (Li-GICs). The composition corresponding to the maximum possible Li-intake (LiC6;
stage I GIC) has one Li-ion in between every other basal plane of graphite. Hence, graphite can host a
maximum of 1 Li-ion per 6 carbon atoms, leading to a maximum theoretical gravimetric Li-capacity of
372 mAh/g. Schematic representations of the atomic arrangements of the different Li-GICs at the dif-
ferent stages of Li-intercalation, super-imposed on a typical potential (V) – capacity (mAh/g)/time (h)
obtained during Li-intercalation/de-intercalation half cycles in a cell consisting of Li foil as counter
electrode, are presented in Fig. 4. The lithium intercalation in graphite occurs at potentials below
0.3 V and up to 0.01 V vs Li/Li+ [1,16,17,19,125,132–134], as can be seen from the potential pla-
teaus (each of which corresponding to the co-existence of two phases). More details about the struc-
ture and electrochemical behavior of lithiated graphite can be found elsewhere [16,19,54,132,133].
The major structural change due to Li incorporation in graphite is the increased interlayer spacing
of the basal planes (0 0 0 2), typical of intercalation compounds. Theoretical studies indicate that upon
full lithiation (i.e. LiC6), the interlayer spacing increases by 10% [16,19,54,133]. Dimensional changes
along the basal planes (a-axis) is significantly lesser; 1% for full lithiation [16,19,54,133]. Reported
variations of the lattice dimensions along the c-axis and the a-axis as functions of the state of charge
are presented in Table. 1. These changes result in net changes in the volume of the graphitic carbon
based anodes, upon full lithiation, by 14% (as estimated theoretically) [133]. These overall changes
in lattice dimensions, along with the co-existence of different Li-GIC phases during lithiation/delithi-
ation (see Fig. 4) are the major sources of Li-intercalation/de-intercalation induced stresses in
70 A. Mukhopadhyay, B.W. Sheldon / Progress in Materials Science 63 (2014) 58–116

Fig. 4. Typical potential (against Li/Li+) vs. degree of lithiation/delithiation plot showing the various plateaus corresponding to
the co-existence of the different Li-GIC phases (staged compounds) [132] and schematics showing the arrangement of the Li-ion
intercalate layers with respect to the graphite (0 0 0 2) layers for the different stage compounds [133].

graphitic carbon electrodes. Furthermore, it has been estimated [133], as well as inferred from
experimental measurements [144], that lithiation progressively reduces the elastic modulus along
the basal planes (by 12% upon full lithiation) and increases the modulus perpendicular to the basal
planes (by 150% upon full lithiation). Hence, in addition to dimensional/strain mismatch among the
different co-existing Li-GIC phases, mismatch in elastic modulus also contributes to the stresses.
Even though graphitic carbon is still the most commonly used anode material, research efforts over
the last decade have been directed towards possibly replacing graphite with metallic anode materials
(such as Si, Sn, Al, Bi etc. [1,9–14,92–95,98,121–124]). These metallic anode materials have theoretical
Li-capacities that are superior to those of graphitic carbon by 3–10 times [1,9–14,92–95,98,
121–124]. Furthermore, since most lithiation occurs at potentials slightly above the lithium deposition
potentials (i.e. 0 V vs. Li/Li+), safety issues associated with Li-deposition and dendrite formation are
more easily circumvented with the metallic anode materials [1,9–14].
Lithiation/delithiation of the metallic anode materials typically occurs by alloying/dealloying,
with the Li-ions occupying the interstitial sites in the host crystalline or amorphous lattices
[1,9–14,92–95,98,119,121–124]. Among the various possible metallic electrode materials, Si, Sn and
to some extent Al have been in the forefront of the research investigations to date. Both Si and Sn
can theoretically host up to 4.4 Li atoms per host atom, progressively leading to a composition of
Li22S5 (S = Sn or Si). This leads to a theoretical gravimetric capacity of 4200 mAh/g upon full lithiation
of Si [1,9–14], which is more than an order of magnitude higher than graphitic carbon (372 mAh/g)
and even superior to Li-metal itself (3800 mAh/g) [1]. Due to higher atomic weight, Sn has a lower
theoretical gravimetric capacity of 1000 mAh/g [10–14], which still exceeds that of graphitic carbon
A. Mukhopadhyay, B.W. Sheldon / Progress in Materials Science 63 (2014) 58–116 71

(a)

(b) 350
% Change in volume per Si atom
350
% volume increase per Si atom

Elastic modulus (C11)


300 300

Elastic modulus (C11)


250 250

200 200

Li22Si5
Li13Si4

Li15Si4
Li12Si7

150 150
Li14Si6

100 100
LiSi
Si

50 50

0 0

0 1 2 3 4 5
x in LixSi (increase in degree of lithiation)

Fig. 5. (a) Volumetric and gravimetric capacities of the various anode materials, showing the metallic anode materials
possessing considerably superior capacities, compared to carbon [11]. (b) Variation of % volume change per Si atom and
longitudinal elastic modulus (C11) with degree of lithiation (data obtained from references [9,141]). Note the significant
increase in volume, along with softening, with progress of lithiation.

by a factor of 3. Similarly, Al can be alloyed with Li up to a maximum of 2.25 Li atom per Al atom,
leading to a composition of Al4Li9 upon full lithiation. Because of its lower atomic weight, its theoret-
ical gravimetric capacity is 2200 mAh/g [10–14,145]. A comparison of the Li-capacities of the various
anode materials is presented in Fig. 5a. It must be mentioned here that, Kasavajjula et al. [9] presented
an estimation that reveals that the cell capacity increases with anode capacity till it reaches
1200 mAh/g (for fixed cathode capacity), beyond which increase in anode capacity does not lead
to any further increase in net cell capacity.
While the capacities of such metallic anodes are higher than graphitic carbon by up to an order of
magnitude, the net volume expansion that occurs during the lithiation can be as high as 400%, espe-
cially for Si [9–14], which is greater than that of graphitic carbon by a factor 30. This tremendous
change in volume is the major factor that leads to stress development in the metallic anode materials.
Additionally, the intermetallic phases (such as LiS, Li12S7, Li14S6, Li13S4, Li15S4, Li22S5
[9,94,119,140,141]; S = Si/Sn; or AlLi, Al2Li3, Al4Li9 [142,145]) that progressively form and co-exist dur-
ing the course of lithiation have significant molar volume mismatches, which also leads to stress
development during cycling. The calculated unit cell volumes [9,94,140,141] for the various Li–Si
intermetallic phases (i.e. variation with degree of lithiation) are plotted in Fig. 5b. The longitudinal
elastic modulus (C11), as estimated via first principle calculations [133], for the respective phases
are also shown in Fig. 5b. Table 1 lists the properties corresponding to the different anode materials.
72 A. Mukhopadhyay, B.W. Sheldon / Progress in Materials Science 63 (2014) 58–116

(b)

(a)

(c)

(d)

micro-crack
edge (f)
(e) component screw
component

dissoctaion
into Shockley
partials edge
component
micro-crack

Fig. 6. (a–d) Variations of lattice parameters and unit cell volume, and crystalline phases present as functions of degree of
delithiation of LiCoO2 [111,137,146]. (e and f) TEM images of as-cycled LiCoO2 particles, showing dislocation networks [147],
severe lattice strains and micro-cracks [30].

It can be observed that lithiation leads to progressive decrease in the elastic modulus and hence, dur-
ing co-existence there is considerable mismatch of elastic modulus between the intermetallic phases.
A. Mukhopadhyay, B.W. Sheldon / Progress in Materials Science 63 (2014) 58–116 73

It has also been reported [9–14,92–95,98,119,121–124] that room temperature lithiation of crystalline
Si and Sn leads to amorphization that can be reversed during delithiation. This can result in co-exis-
tence of crystalline and amorphous phases during electrochemical cycling, which is another source of
mismatch induced stress development within the Si-based anodes.

3.1.2. Cathode materials


Transition metal oxides, such as LiCoO2 and LiMn2O4, are presently the most widely used cathode
materials. LiCoO2 possesses a net theoretical capacity of 270 mAh/g (assuming full delithiation up to
CoO2), only about half of which (140 mAh/g) is presently usable [1,20–30]. This corresponds to 50%
Li-ion extraction (i.e. up to Li0.5CoO2) and a corresponding cut-off voltage during Li deintercalation of
4.25 V against Li/Li+, such that LiCoO2 can be cycled between potential ranges of 3.5–4.25 V against
metallic Li [1,20–30].
High temperature LiCoO2 (HT-LiCoO2; formed by annealing above 550 °C) crystallizes in a trigonal
crystal lattice belonging to a space group of R3m (a-NaFeO2 structure), with the oxygen atoms forming
a closed packed hexagonal lattice in ABCABC stacking and the Li and Co atoms occupying the octahe-
dral sites between them in the form of alternating layers [1,20–30,138,146,147]. The lattice parame-
ters for fully lithiated (stoichiometric) HT-LixCoO2; x = 1, in terms of hexagonal setting are a = 2.82 Å
and c = 14.06 Å. The repulsive interactions of the negatively charged oxygen anions tend to de-stabi-
lize the layered structure, in particular the arrangements of the CoO2 layers along the c-axis. The pres-
ence of the positively charged Li-ions in-between the CoO2 slabs maintains the cohesiveness of the
layered structure [137,138,146]. During delithiation, when the Li-ions are removed from between
the CoO2 slabs, the shielding effect of the Li-cations decreases and the lattice parameter corresponding
to the c-axis increases monotonically, with concomitant decrease in the lattice parameter correspond-
ing to the a-axis (see Fig. 6a–c and Table 1) [29–31,137,138,146]. The single initial hexagonal phase
(H1) is maintained during Li-extraction till x = 0.9 (3.5–3.9 V; against Li/Li+), below which a new hex-
agonal phase (H2) is nucleated with lattice parameters different from the H1 phase (as shown in
Fig. 6a–c and Table 1). Both the phases co-exist till x = 0.78, leading to a potential plateau at 3.9 V
against Li/Li+ (see Fig. 6a–d), after which the initial phase H1 disappears. The phase H2 then exists
as a single phase, undergoing a progressive increase in c-axis lattice parameter and decrease in a-axis
lattice parameters, till x = 0.51. The unit cell volume decreases by 3% over this de-intercalation range.
It must be noted that a strain >0.1% is considered severe for the brittle ceramic (oxide) cathode mate-
rials [25,30], in contrast to the metallic anode materials. Further de-intercalation below x = 0.51 (i.e. at
voltages above 4.25 V against Li/Li+) leads to a severe distortion of the lattice as the hexagonal lattice
gets transformed to the monoclinic phase, along with ordering in the Li-layers [29–31,137,138,146].
The monoclinic phase converts back to a hexagonal phase at x = 0.45, which remains as a single phase
till x = 0.22, with drastic decrease in c-axis lattice parameter and increase in a-axis lattice parameter
progressively with delithiation (see Fig. 6a–d). With further de-intercalation, the hexagonal phase and
a new monoclinic phase co-exist until a single hexagonal phase forms at x = 0.04, which is structurally
similar to that of the end-member, CoO2, and with significantly decreased c-axis lattice parameter.
Few studies have confirmed that full de-lithiation (x = 0) is possible with the formation of CoO2, which
exists in a metastable hexagonal crystal structure with lattice parameters; a = 2.8222 Å and
c = 12.879 Å [138,146].
These drastic changes in the lattice parameters and crystal structures lead to changes in macro-
scopic dimensions (volume) during lithiation/delithiation. Furthermore, similar to the metallic anode
materials, the preceding discussion points towards co-existence of the various phases/crystal struc-
tures at the different voltage ranges for the cathode materials as well. These are the major causes
for stress generation in the LiCoO2-based cathode materials. It has also been reported that after exten-
sive cycling (after 300 electrochemical cycles), the trigonal HT-LiCoO2 partially transforms to the cu-
bic ‘Low Temperature’ spinel phase; LT-LiCoO2, the transformation being initiated near the surface
[29,30]. When this transformation, occurs severe mismatch stresses. Also, the LT-LiCoO2 phase, usually
otherwise obtained on annealing at lower temperatures of 400 °C, is not considered suitable for use
as Li-ion battery cathode material [1,20–30,138,146,147].
Significant dislocation activity, severe lattice strains and micro-fractures were observed during
Transmission electron microscopy (TEM) of as-cycled LiCoO2 particles (see Fig. 6e and f) [30,147].
74 A. Mukhopadhyay, B.W. Sheldon / Progress in Materials Science 63 (2014) 58–116

(a)

(b) 0.825
Phase I
Phase II
0.820
Lattice Parameter (nm)

0.815

0.810

0.805

0.800
0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
x in LixMn2O4 (degree of lithiation)

Fig. 7. (a) Variation of Potential (vs. Li/Li+) with respect to degree of lithiation (discharge) and delithiation (charge) for
LixMn2O4, even during lithiating beyond x = 1 [111]. (b) Variation of lattice parameter for both the cubic phases of LixMn2O4
with degree of lithiation (between x = 0.1 and 1) (data obtained from Refs. [31,135]).

Careful analysis of the TEM observations by Gabrish and co-workers [147] revealed that perfect dis-
locations possessing burgers vector of a/3h1 1 2 0i on the basal planes {0 0 0 1} also dissociated into
Shockley partials of the type a/3h1 0 1 0i. More importantly, the dislocations were observed in orienta-
tions that would allow them to be glissile. It was hypothesized that the transformation of the perfect
dislocation into partial dislocations and the simultaneous glide of the partial dislocations contribute
towards changes in the stacking of CoO2 slabs and hence the transformations between the various
crystal phases. Furthermore, the presence of glissile dislocations is evidence in itself of one of the pre-
valent mechanisms for deformation and concomitant stress development.
Since Mn is cheaper and more environmental friendly compared to Co, cathodes based on LiMn2O4
are also extensively used in Li-ion batteries [31–40,135]. LiMn2O4 possesses a normal cubic spinel
structure (space group: Fd3m) with cubic close packing of the oxygen ions, Mn-ions in half the octa-
hedral sites and the Li-ions in one-eighth of the tetrahedral sites. For the stoichiometric LixMn2O4
(x = 1), the average oxidation state of Mn-ions is 3.5 and the lattice parameter (a) of the cubic unit cell
is 8.24 Å. During delithiation (charge), Li-ions are extracted between x = 1 and x = 0 within the
A. Mukhopadhyay, B.W. Sheldon / Progress in Materials Science 63 (2014) 58–116 75

potential range of 3.0–4.5 V against Li/Li+, with the bulk of the Li-ions being extracted at potential of
4 V (see Fig. 7a). Electrochemical cycling within this range leads to contraction (and expansion) of
the cubic lattice isotropically and reversibly during delithiation (and lithiation), with lattice parameter
(a) decreasing uniformly to 8.03 Å at x = 0.33 [31,135]. This corresponds to a reversible lattice strain
of 2.5% and a volumetric strain (of unit cell volume) of 7.5%. Furthermore, co-existence of two cubic
phases (phase I and phase II) have also been observed between x = 0.5 and 0.13 (i.e. 4 V against Li/Li+)
for LixMn2O4, with the lattice parameters of both the phases decreasing isotropically during
delithiation (see Fig. 7a and b) [31,135].
Theoretically it is possible to insert Li in LixMn2O4 beyond x = 1 and up to x = 2, when Li-ions pro-
gressively fill-up the vacant octahedral sites. However, upon lithiation beyond x = 1, the average oxi-
dation state of Mn becomes less than +3.5 and progressively nears +3, which results in Jahn–Teller
(anisotropic) distortion of the cubic spinel lattice. As a result of this distortion, the c/a ratio of the unit
cell increases by 16%, resulting in the formation of a tetragonal phase [31–40,135]. Under conditions
of thermodynamic equilibrium, this excess lithiation (beyond x = 1) is believed to occur at potentials
below 3 V (2.96 V) against Li/Li+ (stage III of Fig. 7a). However, since electrochemical cycling occurs
under non-equilibrium conditions, the cubic to spinel transformation is initiated in the cathode mate-
rial (especially near the surface) towards the end of lithiation (discharge) half cycle, even if the voltage
does not drop below 3 V (against Li+) [31–40,135]. In fact, Gummow et al. [36] postulated that the
non-equilibrium and dynamic conditions prevailing during electrochemical cycling can lead to the for-
mation of the tetragonal LiMn2O4 at the spinel particle surfaces even at potentials of 4 V against Li/Li+,
especially at the higher electrochemical cycling rates. The existence of the tetragonal phase at the sur-
face, even on electrochemical cycling at potentials higher than 3 V, is also consistent with experimen-
tal observations by other groups [31,35,40,135]. This is mainly due to the fact that excess lithiation
occurs near the surface, as compared to the bulk during electrochemical cycling. Such a phenomenon
leads to the co-existence of two phases (spinel and tetragonal), which contributes towards the net
stress, in addition to that of the 16% lattice distortion (and volumetric strain of 50%; see Table 1).
Since the pioneering work of Padhi et al. [41] on the possibility of using phosphor-olivines (such as
LiFePO4) as cathode materials in Li-ion batteries, extensive research efforts have been concentrated on
exploring and understanding the electrochemical performance of LiFePO4 and its composites [41–45].
Presently such electrodes are replacing the LiCoO2 or LiMn2O4 based cathodes even in commercial bat-
teries, mainly due to the high Li-capacity (170 mAh/g), better environmental compatibility, im-
proved safety and superior performance under adverse conditions of LiFePO4 [41–45,61,62].
LiFePO4 (triphylite) has an orthorhombic unit cell (Pnma symmetry), formed by a distorted hexag-
onal close-packed stacking of oxygen ions with Fe and Li ions on octahedral interstitial sites and phos-
phorus in tetrahedral coordination [41–45,61,62,73]. The crystal structure consists of PO4 tetrahedra,
FeO6 octahedra and tunnels of Li-ions along the b-axis [41–45,61,62,73,143]. Upon delithiation, a new
phase, FePO4 (heterosite) forms which belongs to the same space group. There is negligible solid-sol-
ubility in the two end members (LiFePO4 and FePO4), which results in co-existence of the two phases
at all times during lithiation and delithiation, with the fraction of each phase varying with the net Li-
content. In fact TEM observations of Li0.5FePO4 (partially delithiated) particles have shown the pres-
ence of alternating stripes of the two end-members, with the stripes being parallel to the (1 0 0) planes
(see Fig. 8a and b) [143]. Theoretical calculations predict that the Li-diffusion takes place along tunnels
parallel to the b-axis or [0 1 0] directions, as schematically shown in Fig. 8c and d, and the phase
boundary progresses along the a-axis [73].
Even though the two phases (LiFePO4 and FePO4) possess the same crystal structure, they differ in
terms of their lattice parameters (LiFePO4: a = 10.33 Å, b = 6.01 Å and c = 4.69 Å; FePO4: a = 9.81 Å,
b = 5.79 Å and c = 4.78 Å [41–45,61,62,68,73,143]). Hence, lithiation leads to tri-axial state of strains
(ea = 5.03%; eb = 3.7%; ec = 1.9) [61,68], that are reversible upon delithiation. Furthermore, the volume
change associated with this phase transformation is 6.8% [41–45,61,62,73,143]. The constraining ef-
fects on such phase transformations and the misfit strains associated with the co-existence of the two
phases, with constantly changing volume fractions, are the major contributing factors towards the
stress development in LiFePO4 based cathodes. Edge dislocations, parallel to the c-axis, have been
observed in as-cycled LiFePO4 and it has been suggested that the greatest mismatch strain along
the a-axis (5%) leads to the generation of such dislocations (see Fig. 8e and f) [61,62,68,73].
76 A. Mukhopadhyay, B.W. Sheldon / Progress in Materials Science 63 (2014) 58–116

Fig. 8. (a and b) TEM images of as-cycled LiFePO4 particles, showing the presence of strips (domains) aligned along c-axis and
also microcrack along b–c plane [143]. (c and d) Schematic illustrations of Li-ion insertion path along the b-axis in the LiFePO4
olivine crystal [73]. (e) Dark field TEM image of as-cycled LiFePO4 particle, showing significant dislocation activity [62]. (f)
Schematic illustration of formation of edge dislocation, with (1 0 0) extra half-plane and (0 1 0) glide plane, and concomitantly
crack surface normal to (0 1 0) (under mode I loading), due to partial filling of one or two layers during lithiation along b-axis of
LiFePO4 crystal [73].

Furthermore, theoretical models have also been developed which indicate that partial filling/emptying
of Li-ion layers during repetitive cycling may also lead to the formation of extra-half planes, resulting
in the formation of dislocations (see Fig. 8f) and the associated dislocation stress fields [73].
A. Mukhopadhyay, B.W. Sheldon / Progress in Materials Science 63 (2014) 58–116 77

Fig. 9. Schematic illustrations of (a) solvated Li-ion intercalation between graphene layers, decomposition inside the layers and
concomitant significant expansion of the interlayer spacing [148]. (b) Proposed model for stress development due to SEI layer
formation, based on ‘deposition of supersaturated high-mobility ions’ [125]. (c) Proposed model for SEI formation, based on
‘initial amorphization of the surface region intercalated with solvated Li-ions and subsequent deposition of inorganic particles’,
which is capable of bearing the recorded magnitudes of irreversible compressive stresses [154].

3.2. Stress due to surface reactions

The previous sub-section describes the causes of stress development in the respective electrode
materials due to the actual process of lithiation/delithiation. A survey of the open literature
[148–154] and the authors’ very recent work [54,125,154] also point to some other factors that can
also lead to significant stress development. During electrochemical cycling the electrode materials
are in direct contact with the electrolyte, which contains Li-ions usually in the form of inorganic salts
dissolved in organic solvents.
In solution (electrolyte), Li-ions possess a solvation shell around them (solvated Li-ions), making
them significantly larger in size as compared to the unsolvated Li-ions. Hence incorporation of the sol-
vated Li-ions can lead to tremendous increase in the spacing between the graphite basal planes
(0 0 0 2), as shown schematically in Fig. 9a. In the case of graphitic carbon electrodes, Besenhard
et al. [148] proposed that in the first lithiation half cycle, at potentials (1.2 V against Li/Li+) much
78 A. Mukhopadhyay, B.W. Sheldon / Progress in Materials Science 63 (2014) 58–116

higher than the actual Li-intercalation potential (0.3 V against Li/Li+), solvated Li-ions are interca-
lated in the spaces between the basal planes of graphite, forming ternary Li(solv)yCn GICs. This has also
been confirmed by many other researchers [149–153]. Such ternary GIC formations are believed to be
thermodynamically more favorable compared to the binary Li-intercalation compounds at low Li-con-
tents (C/Li P 18) [148]. It has been reported that solvated ion co-intercalation can lead to 150%
expansion along the c-axis, which is more than an order of magnitude more than the c-axis expansion
induced by actual Li-intercalation [148–153]. Furthermore, the process of solvated ion co-intercala-
tion is irreversible, in contrast to Li-ion intercalation. In this regard, it was observed via in situ AFM
that after the first electrochemical cycle the spacing between the graphite basal planes can increase
irreversibly by 25%, which is more than the reversible change in spacing due to actual Li-intercala-
tion (17%) [153]. Furthermore, AFM investigations on highly oriented pyrolytic graphite (HOPG) have
shown that hills and blisters (swellings) up to 20 nm in height form at 1 V, and can be attributed to
solvated ion co-intercalation [149,150].
In certain electrolytes, particularly those containing propylene carbonate (PC) as solvent, the co-
intercalation continues in an uncontrolled manner even after the potential reaches the actual Li-inter-
calation potentials. Fortunately, with the more commonly used solvents, such as mixtures of ethylene
carbonate (EC) and dimethyl carbonate (DMC), solvated ion intercalation is suppressed once a protec-
tive layer/film composed of electrolyte decomposition (reduction) products forms on the surface of
the graphite electrode (starting slightly below 1.0 V against Li/Li+) [125,148–153]. This passivation
layer, known as the solid electrolyte interphase (SEI) [125,148–154], is comprised of various organic
(such as ROCO2Li) and inorganic (such as LiCO3, LiF) salts. Its formation on graphitic carbon electrodes
prevents further direct contact of the electrode with the electrolyte, while still allowing Li-ion diffu-
sion (but not electronic conduction). This passivation layer is beneficial as it prevents further contact
of the electrode with the electrolyte. However, it also leads to irreversible consumption of Li, which
cannot be further used for additional electrochemical cycling. Even though this layer prevents solvated
ion intercalation and the corresponding lattice distortion, it has recently been observed by the authors
[54,125,154] that the surface phenomena associated with SEI formation lead to the development of
(irreversible) compressive stresses over 18 electrochemical cycles (with the bulk being in the first
cycle), that are significantly greater than the actual lithiation induced stresses. Such stress develop-
ment was explained in terms of the development of compressive stresses during growth of films (in
this case, SEI) from high mobility atoms (in this case Li+) because of super-saturation [125,155,156]
(see Fig. 9b; [125]), along with disruption of the graphitic lattice very close to the surface (up to depths
of few nm; see Fig. 9c [154]).

4. Effects of stress development in electrode materials

As outlined in the previous section, various causes lead to stress development during electrochem-
ical cycling in Li-ion battery materials. Though the exact causes and magnitude (not experimentally
measured as yet for most materials/configurations) varies between different materials, such stress
development is expected to impact the performance of electrode materials and hence the overall bat-
tery performance. It is believed that the majority of the impact would come from the effects of stress
on the mechanical behavior/integrity of the electrode materials. However, stress in the electrodes can
also affect various phenomena associated with electrochemical behavior. The following sub-sections
will shed some light on such effects and their overall consequences.

4.1. Impact on battery performance

The performance of a battery is measured in terms of its energy density, power density and cycle
life. These metrics are respectively linked to the usable Li-capacities of the electrode materials, the
accessible lithiation/de-lithiation rates, and the ability of the electrodes to maintain integrity and
appreciable Li-capacity over a number of cycles, respectively. These three capabilities can be inter-re-
lated in many cases, and stress development can impact all of the capabilities to different extents.
A. Mukhopadhyay, B.W. Sheldon / Progress in Materials Science 63 (2014) 58–116 79

Fig. 10. Images obtained with as-cycled (a–c) anode and (d–f) cathode materials showing the presence of cracks that form due
to stress development during the discharge/charge cycles [13,52,61,62,70,157].

4.1.1. Effects of stress on mechanical integrity of electrodes


Stress can lead directly to mechanical degradation of the electrode materials. The extent of such
degradation depends on the electrode material and architecture, and is almost always determined
by the magnitude of volume changes upon lithiation/delithiation, the mechanical properties of the
material, and the mechanical constraints [9–136]. Numerous studies have shown that the rate of elec-
trochemical cycling and the degree of lithiation/delithiation also affect the degradation of the elec-
trode materials [61–63,67,68,142,76,80,85]. Mechanical degradation is often manifested in fracture/
disintegration of the active electrode materials. Electron microscope images showing fractured as-
cycled anode and cathode materials are presented in Fig. 10 [13,52,61,62,70,157]. Under most circum-
stances, the fractured portion loses contact with either the current collector or the rest of the active
material, which leads to electrical isolation. Hence, progressive fracturing reduces the amount of ac-
tive electrode material and decreases the Li-capacity of the cell. Furthermore, stress development can
also damage the binder and reduce the pore volume, which in turn affects the mechanical integrity of
the electrodes [10,66,75,158,159]. All of these effects lead to capacity fade and reduce the energy den-
sity of the cell as electrochemical cycling continues. Such stress related fracture/disintegration is be-
lieved to be one of the main reasons for the eventual failure of a Li-ion battery. Depending on the
severity of stress development for certain electrode materials and configurations, fracturing can even
render the electrode unsuitable for cycling beyond very few or sometimes even one electrochemical
cycle [1,9–14,52,65]. For example, severely cracked metallic anode materials after few electrochemical
cycles are shown in Fig. 10a and b.
80 A. Mukhopadhyay, B.W. Sheldon / Progress in Materials Science 63 (2014) 58–116

Fig. 11. Variation of specific capacities of different electrode materials with number of electrochemical cycles
[19,62,161,162,164].

The morphology of cracks in amorphous Si thin film electrodes can be used to obtain fundamental
insight into the fracture processes. For example, Dahn and co-workers [160] had observed that the
cracks formed in Li-alloy thin film electrodes look similar to those formed during drying of mud,
and was believed to having been formed due to lateral contraction of the films during the delithiation
half cycle. Li et al. [65] reported that cracks on thicker films (1000 nm) were relatively straight
with few sharp bends and mostly interconnected, whereas cracks were more jagged in thinner films
(500–200 nm). Furthermore, Xiao et al. [52] observed that for 100 nm thick continuous Si film
electrodes, the average crack spacing are around 7 lm (see Fig. 10a).
Accordingly, severe Li-capacity fade with electrochemical cycling occurs for metallic anode
materials that exhibit 300% volume change upon lithiation/delithiation (as shown in Fig. 11a–c)
A. Mukhopadhyay, B.W. Sheldon / Progress in Materials Science 63 (2014) 58–116 81

[9–14,161–163]. Those can be compared with significantly better capacity retention for graphitic car-
bon electrodes that exhibit lithiation induced volume changes of not more than 14% (see Fig. 11b and d)
[19,161]. In addition to the cracking due to volumetric strain, significant inelastic deformation (flow)
has also been observed during lithiation of amorphous Si (a-Si) after an initial linear elastic region
[53,77,117,121–124]. Considering that a-Si is brittle in its pure form, sustaining large flow in the lith-
iated form has been attributed to a phenomenon known as ‘reactive flow’, where the non-equilibrium
chemical processes involved in Li-insertion and the concomitant structural changes assists the
mechanical flow, supplementing the deviatoric stress components [53,77,117,121–124,164].
In addition to the capacity fade due to electrical insulation of fractured particles, fracture results in
the creation of fresh electrode surfaces. When such fresh surfaces come in contact with the electrolyte,
further electrolyte reduction takes place leading to formation of new SEI layers [10,165]. Hence, such
stress related fracture results in accrued irreversible consumption of Li, which in turn negatively af-
fects the net cell capacity. Dimensional changes of the active materials also lead to fracturing of the
as formed SEI layers (see Fig. 10c), which results in a dynamic process of SEI formation and reforma-
tion [157]. Additionally, it has now been realized that significant stress is also associated with the very
formation (deposition) of the SEI layers, which can further affect the mechanical integrity and the pas-
sivation effects of such layers [54,125,154]. Such stress and fracturing of the SEI layers result in con-
tinuous build-up of such layers, well past the initial cycles, leading to continuous irreversible
consumption of Li from the system. It has been observed by the authors [54,125,154] that considerable
irreversible capacity, and concomitantly irreversible stress, gets built-up up to 18 cycles in graphitic
carbon electrodes. The dynamic build-up – fracture and re-construction of SEI layers is expected to be
more severe in case of the metallic alloy materials, in which the SEI layers get additionally strained
due to the huge volume expansion (300%) of the active materials [9,10,166,167]. An interesting
observation by Inaba et al. [167] was that for Sn-thin film electrodes, irreversible capacity due to for-
mation of SEI layer loss was not observed in the first cycle, but was observed from second cycle on-
wards. Such behavior was attributed to the destruction of the inert oxide surface film due to stress
developments (dimensional changes) in the first cycle, which exposed the Sn surface to the electrolyte
from the second cycle onwards.
In addition to the capacity fade with electrochemical cycling, stress associated with phase transfor-
mations (changes in lattice dimensions) reduces the usable capacities of LiCoO2 (140 mAh/g) and
LiMn2O4 (120 mAh/g) to nearly half of the theoretical capacities (270–300 mAh/g) [1,21–40].
The severe lattice distortion on transformation from trigonal to monoclinic on extracting Li beyond
0.5 in LixCoO2 (on cycling above 4.2 V against Li/Li+) cannot be tolerated by the LiCoO2 based elec-
trodes [1,21–31]. On a similar note, the Jahn–Teller distortion induced transformation from cubic
(c/a = 1) to tetragonal (c/a = 16) symmetry on lithiation beyond x = 1 in LixMn2O4 (i.e. below 3 V
against Li/Li+) leads to immediate failure of the LiMn2O4-based electrodes [31–40]. It must be noted
here that even on cycling within the ‘allowable voltage ranges’ stress related capacity fade is observed
for both the cathode materials. Furthermore, as noted in the previous section, Jahn–Teller distortion
occurs near the surface of LiMn2O4 and results in fracture/disintegration even on cycling above 3 V
[33–39].
With respect to impacts of stress development on the rate capabilities, stress related fracture has
been observed in LiFePO4-based electrodes at the higher electrochemical cycling rates. Fractured as-
cycled LiFePO4 particles have been observed in both SEM [60] as well as TEM [62,143] (see Fig. 10e
and f). Literature reports [60,62,143] have confirmed that faster electrochemical cycling rates lead
to very rapid capacity fade for LiFePO4 (see Fig. 11e). In more general terms, such stress related prob-
lems get accrued at the higher cycling rates for most electrode materials mainly due to establishments
of steeper Li-concentration gradients. This in turn limits the electrochemical cycling rate that can be
used for practical purposes and hence the rate capability of the electrode.

4.1.2. Effects of stress on thermodynamics of the electrochemical phenomena


The effects of phase transformation, misfit strain and concomitant stress development will also
influence the thermodynamics of the electrochemical processes, particularly the potentials of the lith-
iation and delithiation steps. The effects of stress state on the open circuit voltage can be envisaged
82 A. Mukhopadhyay, B.W. Sheldon / Progress in Materials Science 63 (2014) 58–116

Fig. 12. Accommodation energies as functions of Li-content for LiFePO4 particles (a) of different sizes (40 nm and 100 nm); (b)
during different cycles (1st and 2nd) [67]. Potential hysteresis as functions of Li-content (c) for LiFePO4 particles of different
sizes (40 nm and 100 nm); (d) for LiFePO4 particles during different cycles (1st and 2nd) and (e) for amorphous Si thin films
[124]. (See above-mentioned references for further information.).

from modified relationships of chemical potential (C), based on basic thermodynamics, as below
[124,168]:
 
Ci ¼ ðdG=dNi ÞT;P ¼ l0i þ RTlnðci Ni =RNi Þ þ zi F u þ V rjj dejj =dNi T;P
ð3Þ

where G and V are the total Gibb’s free energy and volume of the material, Ni is the number of moles of
component i (in this case, Li-ions), loi and ci are the reference chemical potential and activity coeffi-
cient for i, zi is the charge of i (+1 for Li ions, etc.), u is the electric potential, F is the Faraday’s constant
A. Mukhopadhyay, B.W. Sheldon / Progress in Materials Science 63 (2014) 58–116 83

(96,500 C/mol), rjj and ejj are the elastic stress and strain tensors, respectively. The second last term
in Eq. (3) accounts for the contribution from the electric potential, while the last term accounts for
elastic strain energy in the material.
The strain accommodation and deformation consumes energy, which decreases the energy avail-
able for the phase transformation associated with lithiation/delithiation [122,124]. In other words,
the energy accommodation during elastic–plastic deformation of the electrode materials decreases
the lithiation equilibrium potential and increases the delithiation equilibrium potential (with respect
to the equilibrium potential under strain free conditions) [67,122,124]. The accommodation energies
during lithiation/delithiation, as estimated for LixFePO4 as a function of x, by Zhu and Wang [67] are
shown in Fig. 12a and b. Accordingly, the decrease in the lithiation equilibrium potential (DElithiation)
can be expressed as [67]:
DElithiation ¼ Eo  Ede ¼ ðDGelastic þ DGplastic Þ=nF ð4Þ
where Eo is the theoretical equilibrium potential under strain (accommodation) free conditions, Ede is
the actual discharge equilibrium potential under conditions of accommodation process during Li-ion
insertion, DGelastic and DGplastic are elastic and plastic accommodation energies during lithiation
process and n is the number of electrons utilized in the lithiation reaction per atom of host
material reacted. This leads to hysteresis between the lithiation and delithiation potentials
(DElithiation + DEdelithiation), which is known as open-circuit potential hysteresis. In a different work
related to studies on electrochemical cycling of amorphous Si thin film based anodes, Sethuraman
and co-workers [122] developed a relation for estimation of the change in potential (DE) in response
to the stress change (Dr) as below:
DE ¼ mgDr=3F ð5Þ
where m is the molar volume of the host (Si) and g is the rate of change of volumetric strain due to
lithiation. Based on this analysis, it was estimated that stress-potential coupling (DE/Dr) in Si-thin
film anodes is 62 mV/GPa [122]. Hence, lithiation/delithiation induced stress development by
2–3 GPa in such electrodes can lead to a substantial offset of the equilibrium potentials
(by 0.15 V for each half cycle). In a different work, Sheldon et al. [124] estimated that even though
the stresses can lead to significant alteration of the potentials for a-Si thin film electrodes, which is in
agreement with the prior work of Sethuraman et al. [122], stress variations across Si electrodes will
have only a modest impact on composition variations at a uniform potential because the of large
negative mixing enthalpy in this material. However, larger stress-induced composition variations
may occur in other materials [124].
Hysteresis in the lithiation/delithiation potentials, as observed for Si-based anodes and LiFePO4-
based cathodes, are shown in Fig. 12c, d and e. It can be seen from Fig. 12c that smaller particle size
(40 nm) leads to slightly lower hysteresis as compared to that for 100 nm sized particles [67]. Fur-
thermore, for LiFePO4 hysteresis loss during the second cycle is slightly lesser as compared to the first
cycle, which might be due to deformations, dislocation generations, cracking and particle rearrange-
ments already taken place during the first cycle that aid phase transformation and accommodation of
volume changes from second cycle onwards [67]. It must be noted that the area of the hysteresis loop
corresponds to the energy dissipated during a single lithiation–delithiation full cycle. Hence occur-
rence of such hysteresis due to stress development and concomitant larger area of hysteresis loop
leads to considerable energy loss during electrochemical cycling, that adds onto the energy losses
due to over potential and IR (resistance losses) in the electrochemical cell.

4.2. Impact on further progress in Li-ion battery technology

Present generation Li-ion batteries are primarily used in small consumer electronics. However,
they also offer potential solutions to the global energy crisis, particularly in electric vehicles. These
more demanding applications necessitate a calendar life up to 15 years, along with significant
improvements in energy and power density. The development of improved Li-ion batteries to meet
these requirements is currently limited by several important factors, one of which is the stress related
mechanical degradation.
84 A. Mukhopadhyay, B.W. Sheldon / Progress in Materials Science 63 (2014) 58–116

As discussed in the preceding sections, the higher capacity metallic alloy materials [1,9–14]
(deemed as replacement for graphitic carbon anodes) undergo huge volume changes during lithia-
tion/delithiation resulting in stress build-up and concomitant capacity fade (see Figs. 10 and 11). Such
stress related failures have hindered the use of these higher capacity anode materials and concomi-
tantly have limited the application of these materials. Furthermore, based on fracture criterion, stres-
ses limit the dimensions (particle sizes and film thicknesses) of the electrode materials that permit
stable electrochemical cycling [9,10,60,68,69]. As will be explained in more detail in Section 8.3, stress
development also limit the potential window for electrochemical cycling. These in turn limit the net
energy density of the cell. Due to the relative magnitudes of dimensional changes experienced upon
lithiation/delithiation this limitation is definitely more stringent for the metallic anode materials
[9,10]. At the same time, concerns have also been expressed over stress related fracture of the phos-
phate based cathode materials [60,62,143], and similarly for most electrode materials, when cycled at
the higher electrochemical cycling rates. This in turn negatively affects the rate capabilities of the
Li-ion batteries and, in particular, hinders the use of much safer and relatively less expensive
phosphate based batteries in high power applications.
In addition to the degradation of battery performance due to deformation and stress in electrode
materials, lithiation/delithiation induced deformations also result in swelling of the entire battery
pack [170]. Such swelling can result in the development of external stresses on the battery and also
the requirement of larger space for fitting the battery pack, which might limit the applications.

5. Real time observations of electrode material deformations/dilations during lithiation/


delithiation

It is now well documented that significant stress gets developed during electrochemical cycling of
electrode materials in Li-ion batteries and that such stress generation has numerous adverse effects.
The dimensional changes that are at the core of such stress development are also known for many
materials. However, from a more scientific point of view, detailed understandings of the actual mech-
anisms that are associated with stress development are still poorly understood. Such understanding
can provide vitally important information for designing and optimizing high performance electrodes
against adverse effects. An ideal way to understand the phenomena leading to deformation and stress
during electrochemical cycling is to observe the structural changes occurring in situ during lithiation/
delithiation. Such direct observation of the evolution of the internal structures of anode materials dur-
ing lithiation/delithiation have recently been made with in situ transmission electron microscopy
(TEM) [92–94] and in situ X-ray transmission microscopy (TXM) [95].
Huang and co-workers reported for the first time the fabrication of a nanoscale Li-ion battery for
in situ observation of structural evolution during electrochemical cycling of SnO2 based nanowire an-
odes [92]. A schematic of the battery consisting of LiCoO2 as cathode and ionic liquid as electrolyte just
touching the tip of the electrodes is shown in Fig. 13a. In this configuration, a reaction front propa-
gated along the length of the nanowire, that is along the h0 1 1i axial direction, leading to significant
deformation of the nanowire. On full lithiation during first charge, 90% elongation was observed
in the axial direction, with 50% swelling along the transverse direction, resulting in a net volume
change of 250%. The greater elongation along the axial h0 1 1i direction, as opposed to the expected
h0 0 1i direction, was believed to be due to the greater compliance along the axial direction in the
nanowire configuration. The spiraling of the nanowire indicates severe strains and also appreciable
ductility, as a result of the nanowire configuration. The fact that severe misfit stresses were generated
was proven by the presence of high density of mobile dislocations within a 100 nm wide zone at the
reaction front (termed as, ‘Medusa Zone’), separating the unreacted SnO2 ahead and the lithiated re-
gion consisting of crystalline LixSn, crystalline Sn nanoparticles and amorphous Li2O behind (see
Fig. 13b and c). The following reactions are expected to occur during lithiation of SnO2:
þ
4Li þ SnO2 þ 4e ! 2Li2 O þ Sn; irreversible forming reaction ð6Þ

þ
Sn þ xLi þ xe $ LixSnð0 6 x 6 4:4Þ; reversible lithiation reaction ð7Þ
A. Mukhopadhyay, B.W. Sheldon / Progress in Materials Science 63 (2014) 58–116 85

Fig. 13. In situ (real time) observations and schematic representations of the deformations and fracture taking place during
lithiation of anode materials [64,92–95,99].
86 A. Mukhopadhyay, B.W. Sheldon / Progress in Materials Science 63 (2014) 58–116

Fig. 13 (continued)
A. Mukhopadhyay, B.W. Sheldon / Progress in Materials Science 63 (2014) 58–116 87

The dislocations nucleate at the front-side (‘Medusa Zone’ – SnO2 interface) and annihilate at the
back-side (crystalline–amorphous interface) of the ‘Medusa Zone’. The dislocation densities were esti-
mated to be 1017/m2, which is more than two orders of magnitude greater than that observed in
heavily cold worked face centered cubic metals. Based on DFT calculations and the experimental
observations, it was postulated that a stress close to the ideal shear strength of Sn was present in
the ‘Medusa Zone’. It was also hypothesized that the exceptionally high dislocation density was
caused by the enormous lithiation-induced stress, which leads to structural disorder and subsequent
solid-state amorphization that results in the formation of amorphous Li2O behind the reaction front
(as seen in Fig. 13b). Amorphization following lithiation has also been observed in other materials,
particularly in crystalline Si anodes [9–14,92–95,98,119,121–124]. Surprisingly, despite the huge
stress development, no cracking was observed in the SnO2 nanowire anode after the full first lithiation
cycle, unlike that observed for bulk SnO2 electrodes. This, along with the observations of bending,
twisting and significant dislocation activity suggested considerable ductility for the otherwise brittle
SnO2, which was ascribed to the nanowire configuration.
In a similar work [93], where the cell geometry was modified to allow for flooding of the electrolyte
around the SnO2 nanowire electrode, improved understanding of the microstructural evolution during
lithiation of SnO2 electrodes was obtained. It was observed via the in situ TEM that upon lithiation
crystalline LixSn spherical nanoparticles were formed in the amorphous Li2O matrix (see Fig. 13c). Fur-
ther growth of the isolated LixSn particles took place via Li and Sn diffusion through the amorphous
Li2O matrix, which also helped maintain the structural integrity. In case of flooding with the electro-
lyte (as opposed to just contact), it was also observed that at the intermediate stage of lithiation, mul-
tiple strips were formed along the (0 2 0) planes (Fig. 13d) [94]. The strips were formed due to
preferential Li-insertion along those planes in the flooding electrolyte geometry. In this case, these
strips actually served as the reaction fronts that laterally expanded during the lithiation process
and resulted in the deformation and amorphization. Severe deformations during lithiation, that re-
sulted in zigzagging of the initially straight wire damaged the electrode to the extent that delithiation
step could not be performed using the same lithiated wire.
Similar study [94] on Si nanowires with h1 1 2i growth direction led to completely different obser-
vations. Unlike the reaction front advance mechanism for SnO2 [92], the lithiation of the Si nanowires
occurred via a core–shell mechanism, with the Li-ions entering the Si nanowire radially via a thin ionic
electrolyte layer on the surface and resulting in lateral expansion of the nanowire (see Fig. 13e), thus
leading to a net volume expansion of 300%. This is unlike the SnO2 nanowires, which tended to ex-
pand more along the axial direction rather than lateral direction [92–94]. Furthermore, for Si a region
with intense dislocation activity was not observed during lithiation. On the other hand, the one strik-
ing similarity between lithiation of Si and SnO2 was the lithiation induced solid-state amorphization.
It was also revealed by a recent in situ nuclear magnetic resonance (NMR) study [96] that during the
first lithiation cycle, isolated Si atoms and small Si–Si clusters are formed, which is also expected to
contribute to the amorphization. In the case of Si, the lithiated zone, that is the shell, was converted
to amorphous Si (a-Si), whereas the core remained single crystalline Si (c-Si) and formed a conical
shape. Though the amorphous shell thickened gradually during lithiation, the nanowire cross-section
was never completely converted into lithiated a-Si even after prolonged lithiation. A dumbbell-shaped
cross-section was observed and was attributed to plastic flow upon lithiation. Lithiation induced plas-
tic flow of amorphous Si has also been confirmed by other research groups [117,123,124]. On progres-
sive lithiation, ‘necking’ due to plastic low and concomitant stress concentration resulted in fracture of
the Si nanowires into smaller sections. The in situ TEM study also facilitated the characterization of the
structure of the a-Si, which was found to be similar to c-Li15Si4, which throws light on the possible
reasons behind the frequently observed transformation of the a-Si back to the c-Li15Si4 upon continued
lithiation [94]. Furthermore, it was confirmed that the maximum possible lithiation of Si at room tem-
peratures is 3.75 Li atoms per Si atom, which would limit the gravimetric specific capacity to
3600 mAh/g. In recent work by Liu et al. [171] careful observation and analysis also shows that
the velocity of the a-Si/c-Si reaction front decreases as it proceeds towards the core of the nanowire,
and that full litigation does not occur. This behavior was attributed to the large stress that develops
during lithiation. Stress-induced limitations on both diffusion and interface reactions were analyzed
88 A. Mukhopadhyay, B.W. Sheldon / Progress in Materials Science 63 (2014) 58–116

to explain these observations, and it was noted that more work is needed to determine which of these
kinetic effects is likely to be the primary factor that limits full lithiation.
Liu and Huang have also observed the core–shell mechanism of lithiation in Si nanoparticles with
their in situ TEM investigation [94]. During lithiation, a sharp interface between the lithiated shell
unlithiated core was clearly observed (see Fig. 13f). Also radial crack initiated at the particles surfaces
(red arrows in Fig. 13f) when the lithiated shell thickness exceeded 150 nm. This was attributed to
the development of tensile hoop stresses at the surface. The observation of surface cracking, along
with anisotropic expansion upon lithiation is also supported by Lee et al. [172], who observed that
cracking in crystalline Si nano-pillars occurs only in the amorphous lithiated shell. It was believed that
the anisotropic expansion was primarily responsible for the intensification of the surface tensile hoop
stress. These are in contrast to earlier models [60,79,86,87], based on single phase lithiation mecha-
nism, where it was proposed that cracking would initiate from tensile stresses at the particle centers
during lithiation, and compressive hoop stresses at the surfaces would then inhibit cracking. Here it
was presumed that cracking in Si occurs during delithiation step. Other important results were the
identification, via statistical analyses, of 150 nm as the critical diameter of Si particles [94] and
300 nm for Si nano-pillars [172], below which cracking during lithiation would be suppressed.
Such generation of tensile hoop stress during lithiation is also contradicted in another recent work
with in situ TEM observation of lithiation/delithiation of amorphous Si-coated carbon nanofibre
composite electrodes by Wang et al. [98,99]. In accordance to the more general belief, longitudinal
nano-cracks (Fig. 13g) propagate along the axial direction during the delithiation half cycle and were
attributed to the development of tensile hoop and axial stress during the delithiation half cycle. The
constraint imposed by the carbon fiber was believed to lead to the development of tensile radial stres-
ses, but compressive hoop and axial stresses during lithiation half cycle. Furthermore, in situ optical
microscopy and AFM, as used by Dahn and co-workers [97,160] to study the deformation in Li-alloy
thin film electrodes, revealed that cracks appear primarily due to the lateral contraction of the films
during the delithiation half cycle. Whereas, during lithiation the expansion occurs only in out-of-plane
directions and do not usually lead to cracking. It must be noted here that the observations made by Liu
et al. [94], Lee et al. [172], Wang et al. [93] and Dahn et al. [97,160] were for different geometries and
direct comparison may not be justified. However, these differences demonstrate that the geometries/
architecture of the electrode materials will have important ramifications on the deformation and
stress development.
Another interesting observation made with the amorphous Si-coated carbon nanofibre composite
electrodes was that lithiation proceeded from the Si surface, as well as from the Si/C interface, creating
a sandwich structure of LixSi/Si/LixSi (see Fig. 13h) [99]. Such lithiation from the Si/C interface also
indicates possible changes in the character of the interface, which might also influence adhesion be-
tween the Si and the C.
In situ TEM was also used to explore lithiation induced deformation in multi-walled carbon nano-
tubes (MWCNTs) [64]. The excellent mechanical and physical properties of carbon nanotubes (CNTs)
would appear to make them ideal anode materials, with significant potential for mitigating stress re-
lated problems. However, CNTs exhibit severe fracturing and drastic capacity fade compared to the
more usual graphitic carbon [64,173,174]. In situ TEM observations has revealed that lithiation results
in 5.9% expansion of the inter-tubular spacing, which leads to radial and circumferential expansions
and significant tensile hoop stress (50 GPa) along the outermost wall due to the confined cylindrical
geometry. Such stresses become negligible with the more open geometry of graphite. In fact, distorted
side walls of the lithiated MWCNTs were clearly observed in TEM (see Fig. 13i). Additionally, in situ
tensile and compressive tests clearly revealed the brittle nature of the lithiated MWCNTs (showing
brittle fracture), as opposed to pristine MWCNTs. The presence of Li in the intertubular spacing was
believed to act as a point force on the MWCNT walls. This, along with chemical weakening of the
C–C bonds upon charge transfer between the intercalated Li and the C atoms was attributed to the
MWCNT embrittlement, thus making them unsuitable for use as electrodes with good cycle lives.
The structural evolution of Sn particles during lithiation/delithiation was also studied in real time
using X-ray Transmission Microscopy (XTM) [95]. Unlike the TEM observations described earlier
[92–94], XTM revealed the formation of core–shell internal structure in the Sn particles, with a light
core consisting of the unlithiated portion and darker shell consisting of lithiated Sn during the
A. Mukhopadhyay, B.W. Sheldon / Progress in Materials Science 63 (2014) 58–116 89

first lithiation cycle (Fig. 13j). The particles expand radially during lithiation, with the smaller particles
expanding at much faster rates due to greater interfacial reaction kinetics in the smaller
particles (based on core–shell kinetic theory [175]). However, the net% expansion of the smaller
particles was comparatively much less than the coarser particles. For instance, a 10 lm sized particle
expands by 380%, which is expected based on theoretical estimations, whereas a 1 lm sized particle
expanded by only 180%. This highlights the advantages of finer sized electrode particles. It was
hypothesized that the faster lithiation rate of the smaller particles leads to pronounced structural dis-
ordering that prevents expansion. Along with radial expansions, radial cracks were observed in the
brittle shell region (LixSn), emanating from the periphery and extending towards the core (as indicated
by red arrows in Fig. 13j). The radial cracking suggests the presence of tangential tensile stresses in the
shell region. An interesting observation was the development of net-like porous structures, possibly
due to recrystallization induced local densification of pure Sn, during the idle period after complete
de-intercalation. Formation of this porous structure prevented pulverization of the Sn particles
during electrochemical cycling, which highlights the advantage of using a ductile material prone to
recrystallization (Sn), as opposed to a brittle material like Si.

6. Experimental determination of stresses in electrode materials

6.1. In situ determination of stress development using multi-beam optical stress sensor (MOSS)

Techniques that provide real time observations of the internal structures of electrode materials
during electrochemical cycling are leading to detailed understanding of deformation processes and
micro-mechanisms related to stress development. However, one limitation of the in situ microscopy
techniques reviewed in Section 5 is that they do not provide direct measurements/quantifications
of stresses. Direct experimental determination of the magnitude of lithiation/delithiation induced
stresses has only been recently accomplished [19,54,121,123], along with systematic study of stress
evolution as functions of potential, degree of lithiation/delithiation and composition/microstructure
of the electrode materials. The direct study of interrelated chemical and stress induced changes in a
controlled fashion is challenging, particularly with complex porous electrode microstructures consist-
ing of mixtures of active particles, fillers and binders. This challenge is exacerbated by the requirement
that lithiation be conducted in controlled atmospheres (low oxygen and moisture), because of the
inherent reactivity of the Li-containing compounds.
Thus to allow determination of the magnitude of stresses, researchers at Brown University (includ-
ing the authors) employed in situ stress measurements during electrochemical cycling of thin film
electrodes [19,121,123]. With a thin film configuration it is much easier to analyze the interrelated
changes in voltage, composition and stress that occur as Li is added and removed from these materials.
It is believed that this can then form the basis for studying more complex composite structures. Such
investigation was made possible by the use of multi-beam optical stress sensor (MOSS) [127], in con-
junction with the custom-made electrochemical cells [19,54,121–126]. MOSS can be used for real time
monitoring of stress development in thin film via measurement of the changes in the curvature of the
inactive and stiff substrate due to constraining of the relative in-plane dimensional changes of the film
(as happens during lithiation/delithiation). Parallel array of laser beams are reflected from the back
side of the substrate and the changes in the spacing of the reflected beams are monitored, which cor-
respond to the changes in the substrate curvature. The substrate curvature (R) is then converted to in-
plane thin film stress (rf) using Stoney’s equation [127,176], as follows:
2
1=R ¼ 6rf :hf =M s :hs ð8Þ

where hf is the film thickness, Ms is the bi-axial modulus of the substrate and hs is the substrate thick-
ness. More details about the assumptions involved in Stoney’s relation and the fundamental/mecha-
nistic aspects of MOSS can be found elsewhere [127,128]. The versatility of MOSS, as compared to
other techniques for measuring substrate curvature, lies in the extremely high resolution (curvature
20 km or thin film stress 1 GPa Å [127]), negligible sensitivity to vibration (due to presence of mul-
tiple laser beams) and adaptability to any thin film deposition technique/application (needing just
90 A. Mukhopadhyay, B.W. Sheldon / Progress in Materials Science 63 (2014) 58–116

Fig. 14. (a) Schematic representation of the set-up for in situ measurement of stress development in thin film electrodes [19].
Variations of (b) potential and stress [123] and (c) bi-axial modulus [126] with degree of lithiation for amorphous Si thin film
electrodes. Variations of potential and stress with time (d) in the 40th cycle [19] and (e) in the first 20 cycles [125], during the
electrochemical cycling of c-axis oriented graphitic carbon thin film electrodes (CVD C). (f) Comparison between the variations
of Li-capacity with C-rates for CVD C and VAGLA thin film electrodes and the variations of Potential and Stress with time in the
first 20 cycles for VAGLA [54].

optical access to the back side of substrate). Accordingly, the custom made cells designed for such
investigation have a transparent quartz window either on the top or on the side that allows optical
access to the back side of the substrate (usually quartz) of the thin film electrode material under con-
sideration. The counter electrode used for the investigations to date was Li-foil and electrolyte consist-
ing of LiPF6 in EC-DMC. A schematic representation of the experimental set-up is shown in Fig. 14a.

6.2. Experimentally measured stresses in various electrode materials using MOSS

6.2.1. Si thin film electrodes


Several of the initial MOSS studies have explored Si electrodes [53,77,117,121–124,126]. Fig. 14b
presents results obtained with simultaneous in situ potential and stress measurements during con-
stant current lithiation/delithiation of an amorphous Si (a-Si) thin film [123]. Interpretation of the
stress requires an inert thick elastic substrate (quartz here). This constrains the lithiation induced dila-
tion of the a-Si thin film parallel to the substrate/current collector, leading to the compressive stress
that is observed here. These compressive stresses are then relaxed during delithiation. The lithiation of
amorphous Si (a-Si) thin films of thicknesses between 50 and 100 nm produces a linear elastic stress
on the order of 1 GPa [77,121–124]. Beyond the initial elastic stresses, the Si-Li alloy shows
A. Mukhopadhyay, B.W. Sheldon / Progress in Materials Science 63 (2014) 58–116 91

evidence of plastic flow, which has been analyzed in terms of lithiation induced viscous reactive flow
[77,121–124]. Such plastic flow was observed to be suppressed with patterned a-Si thin films com-
prised of square islands of lateral dimensions 7 lm  7 lm. In these materials, the net compressive
stress development at the end of the lithiation half cycle was significantly lower (0.25 GPa)
[52,53,117]. However, evidence for plastic flow were found for a-Si thin film electrodes having pat-
terns of lateral dimensions greater than the critical dimension of 7 lm  7 lm. The implications of
such low stress developments and suppression of flow in patterned thin films, as compared to the
usual continuous thin films, will be discussed in more details in Section 8.
The in situ stress measurements also formed basis for analysis of the stress-potential coupling in Si-
based electrodes [122,124], as was discussed in Section 4.1.2. Furthermore, based on the results of
similar experiments on continuous a-Si thin films on stiff substrate, Sethuraman et al. [126] estimated
the variation of biaxial modulus of a-Si with degree of lithiation (see Fig. 14c). It was found that the
modulus decreases significantly with degree of lithiation. Such variation of elastic modulus with lith-
iation/delithiation is expected to have significant impact on the deformation and stress in the elec-
trode materials during electrochemical cycling. Despite the fact that this work brings into light
additional potential of MOSS system for extracting relevant material properties of the electrode mate-
rials, the possibility that the degree of lithiation and hence the elastic modulus results might be af-
fected by the relatively slow Li-ion diffusion in such films [77], which was not considered in the
referred work, warrants more rigorous analysis.

6.2.2. Carbon based thin film electrodes


Real-time stress measurements were made on 200 nm thick c-axis oriented graphitic carbon thin
film electrodes fabricated via CVD (CVD C; with graphite basal planes parallel to the substrate). Such
films showed compressive stress development parallel to the substrate/current collector of magni-
tudes as low as 0.25 GPa (Fig. 14d) due to actual lithiation/delithiation process, resulting in good
stability during continued cycling [19]. This is significantly less than that observed with a-Si elec-
trodes and is expected in light of the considerably lesser lithiation induced dilation of graphitic carbon
[16,19,54,133]. As expected there is no evidence of plastic flow in these films. Also, the stress-thick-
ness data exhibits different behaviors (i.e., slopes) that correspond to different stages of the graphite
lithiation process (i.e., different plateaus in the potential data (as seen in Fig. 14d). The relationships
between the stress-thickness and potential curves for these graphitic films have not yet been evalu-
ated. The lithiation induced compressive stresses were almost completely reversible beyond the
20th cycle. A critical subsequent discovery [125] was that a substantial contribution to the net stress
in these graphitic carbon films comes not from the actual Li-intercalation/deintercalation process, but
from the processes associated with the formation of the solid electrolyte interface (SEI), which is a pas-
sivation layer that forms on the surface of the electrode materials in contact with the electrolyte dur-
ing the initial cycling. Irreversible in-plane compressive stresses in this layer (Fig. 14e) up to 1 GPa
get developed during the initial cycles. Careful analysis indicated that such stresses were possibly
associated with the ‘electrochemical deposition’ of 100 nm thick SEI layer on the graphitic carbon
thin film electrodes. These in situ results are also supported by a very recent work [170] that maps
the stresses in particle based porous composite graphitic carbon electrodes using ex situ RAMAN
microscopy. As recorded with MOSS in our work [125], RAMAN mapping [170] also indicated the
development of compressive stresses mainly during the early stages of cycling.
In a subsequent work, Sethuraman et al. [134] measured stress development in situ during electro-
chemical cycling of porous-composite graphite electrode (50 lm thick on 20 lm thick Cu current
collector, attached using epoxy to 300 lm Si wafer coated with thermally grown oxide layer), using
similar technique. The lithiation induced reversible compressive stress was only 10 MPa for the
composite electrode, indicating the effective role of binder phase in buffering stress development.
An interesting observation was that swelling of the binder due to absorption of the electrolyte results
in the development of 2 MPa net compressive stresses even prior to electrochemical cycling.
In order to address the issue associated with poor rate capabilities of graphitic carbon electrodes,
the authors successfully fabricated thin film electrodes based on vertical alignment of graphenic layers
with respect to substrate (VAGLA; a-axis oriented carbon thin films) using chromonic liquid crystal
precursors [54]. For such crystallographic orientation, the (intercalating) spaces between the graphene
92 A. Mukhopadhyay, B.W. Sheldon / Progress in Materials Science 63 (2014) 58–116

planes would directly face the electrolyte, permitting faster Li-intercalation/deintercalation kinetics
and concomitantly higher rate capability. As expected, greater capacity retention at the faster electro-
chemical cycling rates occurred for the VAGLA films, in comparison with similar thickness CVD C films
with basal terminated surfaces (see Fig. 14f) [53]. However, for these VAGLA based electrodes, signif-
icantly higher reversible stresses due to lithiation/delithiation of 1.5 GPa were measured, with the
SEI layer formation contributing an additional irreversible stress of 2 GPa (see Fig. 14f). The higher
stresses during actual lithiation/delithiation in the VAGLA are due to the greater dilation along the
a-axis (by an order of magnitude), as compared to that along c-axis (see Section 3). However, the
relatively higher irreversible stresses for the VAGLA points towards differences in the composition/
character of the SEI layers formed on the surfaces of the two types of graphitic carbon films. It has been
reported in the literature that SEI films on the graphitic edge planes are thicker and contain a higher
fraction of inorganic phases (relevant to VAGLA), compared to the SEI on the basal planes (relevant to
CVD C). This might explain the higher irreversible stresses in case of the VAGLA.
From the preceding discussion, it is apparent that the irreversible stresses that develop in the initial
cycles were common to the different graphitic carbon electrodes that were used for the in situ stress
measurements. Very recently, we performed more detailed experiments using the c-axis oriented gra-
phitic carbon thin film electrode to obtain more insight into the origin of these irreversible stresses
[154]. These experiments show that large compressive stresses evolution on the order of – 1 GPa
evolve near the graphite surface, during constant voltage holds at potentials above the intercalation
threshold for the graphitic carbon electrodes (uncoated, as well as coated with Al2O3 via ALD). Com-
plimentary HRTEM observations and detailed analysis of the SEI composition (XPS and SIMS depth
profiling), indicated that the majority of the irreversible stress (starting at 0.6 V) is associated with
amorphization of the graphitic carbon structure near the surface (at depths on the order of 10 nm;
see Fig. 9c). This interpretation is based in part on the premise that the organic electrolyte decompo-
sition products that form as part of the SEI during the first few cycles at these potentials cannot sup-
port stresses of this magnitude.

6.3. Other techniques for experimental determination of stresses/strains in electrode materials

Mukaibo and co-workers [91] introduced a similar optical cantilever technique for in situ measure-
ment of stress development in thin film electrodes. This method uses a single laser beam (as opposed
to multiple laser beams of MOSS) to detect changes in the bending of a cantilever comprised of an ac-
tive thin film (Sn or Ni – 62 atm.% Sn) on an inactive thicker Cu substrate (also current collector) dur-
ing lithiation and delithiation. Compressive stresses were observed during the lithiation half cycle for
both the electrodes (Sn or Ni – 62 atm.% Sn), and these were then released during the delithiation half
cycle. The magnitudes of the net stresses in the active films, as estimated from the observed cantilever
deflections using Stoney’s relation [176], were 52 MPa and 490 MPa, for Sn and Ni – 62 atm.% Sn,
respectively. The difference in the stresses between the two electrodes were explained in terms of
greater elastic modulus for the composite electrode (Ni – 62 atm.% Sn). However, it must be noted that
the magnitudes of the stresses were significantly lower compared to those expected based on theoret-
ical calculations (by 2–3 orders of magnitude) and also compared to those recorded with Si thin films
using MOSS (by at least one order of magnitude) [77,121–124]. It must be mentioned here that the
substrate used in the work of Mukaibo et al. [91] was Cu and not a stiff substrate, as is required for
direct application of Stoney’s relation. At several places in the manuscript [91], the authors express
caution with respect to the possible influences of the deformation of the Cu substrate on the deflection
and stress results.
In an earlier work, Rosolen and Decker [151] also used similar cantilever bending with a single laser
beam to observe the deformation behavior of carbon films on flexible polyester substrates in situ when
subjected to lithiation/delithiation. The presence of irreversible compressive stresses was observed
during the initial cycles, and tentatively ascribed to the formation of SEI on the carbon electrodes.
As described in the previous section, our in situ study of stress development in graphitic carbon elec-
trodes using MOSS also detects the presence of irreversible compressive stress and careful analysis
confirmed that such stresses were indeed associated with formation of SEI layer [54,125,154]. While
the earlier study of Rosolen and Decker [151] provided useful qualitative observations, the magnitude
A. Mukhopadhyay, B.W. Sheldon / Progress in Materials Science 63 (2014) 58–116 93

of the stresses were not reported and possible irreversible stress due to lithiation of non-graphitic car-
bons was not addressed.
Qi and Harris [66] designed an innovative experimental set-up for in situ determination of strain
maps in composite graphitic carbon electrodes during electrochemical cycling. Side-by-side arrange-
ment of the electrodes (graphitic carbon composite electrode and Li-foil) and transparent quartz cov-
ering of the cell allowed optical access to the electrode under investigation. Since various GICs are
associated with characteristic colours, starting from black/grey for unlithiated graphite to golden
for fully lithiated graphite (LiC6) via blue (LiC18) and red (LiC12), the changes in color of the different
regions in the graphite electrode were monitored to understand the state of charge at different loca-
tions during lithiation/delithiation. Digital image correlation then provided in situ information about
the deformation and strain fields. This technique lacks the resolution to observe the strains in the indi-
vidual crystallite or particles, however, the deformation strains in two directions at various regions of
the composite electrode were evaluated and the results suggest that overall strain of not more than 1%
develops away from the cracks (i.e., significantly less than the expected average strains for fully lith-
iated graphite). Another interesting observation was that contraction also occurs during lithiation in
some locations. This was explained with lithiation induced stiffening along the graphitic c-axis, as esti-
mated using Density Functional Theory (DFT) in the previous work of Qi et al. [133]. Another impor-
tant result was the observation that the volume expansion of the active particles during lithiation is
accommodated by up to 25% reduction in porosity, which is enough to affect the mechanical integ-
rity and rate capability of the electrodes.
Among the cathode materials, stress in sputter-deposited LiCoO2 films has been monitored
in situ using an Electrochemical Quartz Crystal Microbalance (EQCM) technique [177]. The in-plane
stresses were determined by removing the contributions to the experimentally measured net
changes in resonant frequencies from the theoretically determined changes in resonant frequencies
associated with the lithiation/delithiation induced mass change. Even though this work was more
qualitative in nature, it provided the first experimental evidence of tensile and compressive stresses
during lithiation and delithiation, respectively (note that these stress directions are reversed
compared to the anode materials, because the unit cell volume of LiCoO2 decrease during lithiation,
as can be noted from Table 1). The more important observation was that the stress changes are
more significant in the voltage ranges that correspond to the co-existence of two phases (Li-rich
and Li-poor).
In another interesting study, Chung and Kim [178] qualitatively monitored strain variation in
LiMn2O4 thin film electrodes in situ using a laser probe beam deflection method (LPDM). The set-up
used a single laser beam to monitor the deflection of the film/substrate system and was somewhat
similar to that used by Rosolen and Decker [151] for monitoring the stresses in carbon electrodes.
Gradual tensile strain evolution during delithiation and vice versa was observed in the absence of
any Jahn–Teller distortion. Jahn–Teller distortions also led to sudden changes in the strain develop-
ment which was reversible. Superimposition of the differential strain curves over the cyclic voltam-
mograms allowed determination of the onset of Jahn–Teller distortion, which was found at
potentials as high as 3.90–3.95 V, depending on the scan rate. Dynamic conditions prevailing during
higher scan rates accentuated the Jahn–Teller distortion, and this also occurred at higher potentials
during the delithiation half cycle.

7. Assessment of stresses in electrode materials via computational research

The challenges associated with measuring many of the key properties and phenomena in electrode
materials have also led to extensive computational modeling efforts. This includes the effects of var-
ious factors that impact stress evolution, such as Li-diffusivity [70,74–77,80–89,71], electrochemical
cycling conditions [76,77,80,81], electrode architecture and geometry [70,74–77,80–89,71], properties
of the active electrode material [68–84,88] and also the current collector [82]. A comprehensive over-
view and detailed description of the specific methods and conditions/assumptions involved in such
computational research are beyond the scope of the present review. However, a concise overview
of some of the more significant findings will be presented in this section.
94 A. Mukhopadhyay, B.W. Sheldon / Progress in Materials Science 63 (2014) 58–116

Fig. 15. (a) Schematic representation of particle expansion and contraction during lithiation and delithiation indicating strain
differentials between surface and core [60]. (b) Variation of critical size for fracture with dilatational strain, for different values
of fracture toughness of the electrode materials [69]. (c) Schematic representation of the electrode configuration consisting of
active particle surrounded by inactive matrix, which is also used for the analysis in reference [78]. (d) ‘Stability index diagram’
for spherical Sn active particles embedded in glass matrix [70]. (e) Effects of aspect ratios on the lithiation induced stresses in
elliptical shaped electrode particles [80]. (f) ‘Electrochemical shock map’ for galvanostatic charging [76].
A. Mukhopadhyay, B.W. Sheldon / Progress in Materials Science 63 (2014) 58–116 95

Fig. 15 (continued)

One primary focus has been analysis of the diffusion induced stresses (DIS) in Li-ion battery elec-
trodes. These stresses usually arise from compositional differences within the electrode materials that
are created by kinetic limitations associated with solid-state diffusion of Li-ions. One result noted by
various researchers is that the maximum stress in particles is proportional to the product of the par-
ticle size and the concentration gradient [70,74–77,80,81,84–89,71]. Christensen and Newman’s [60]
analysis also shows that the maximum stress increases with current density, particle size of spherical
particles, and lower Li-diffusivity in the electrode material, all of which lead to steeper concentration
gradients between the surface and the core. Such analysis was performed using a continuum model for
spherical electrode particles, with the assumption that strain differentials occur between the surface
and the core during both lithiation and delithiation (see Fig. 15a). Using Finite Element Analysis to ad-
dress more complex microstructures, Garcia et al. [87] reported Li-concentration gradients and possi-
ble accumulation of Li-ions near the electrode particle surfaces, especially at the higher
electrochemical cycling rates and for the particles closer to the separator. This analysis showed that
during discharge of a full cell, the surfaces of the cathode particles are in compression and the interiors
are in tension, which implies that the particle interiors are the potential sites for crack developments.
Cheng and Verbrugge [74] analyzed the evolution of strain energy and DIS in spherical particles under
different cycling conditions, to obtain expressions relating the maximum radial stresses (rr) at the
96 A. Mukhopadhyay, B.W. Sheldon / Progress in Materials Science 63 (2014) 58–116

particle centers and the maximum hoop stresses (rh), as well as shear stresses (rshear), at the particle
surfaces (of radius R). Under potentiostatic and galvanostatic cycling conditions this led to:

rpotentiostatic;max
r CR  C0
galvanostatic;max
2 ð9Þ
rr IR=ðFDÞ

rpotentiostatic rpotentiostatic CR  C0
h
galvanostatic
¼ shear ¼5 ð10Þ
rh rgalvanostatic
shear
IR=ðFDÞ

where C0 = uniform initial concentration and CR = surface concentration, D is the Li diffusivity, I is the
current and F is the Faraday’s constant. This analysis shows the expected scaling for the maximum
stresses as functions of the cycling conditions and particle sizes. The same authors [86] proposed
0 0 0 0
an ‘Electrochemical Biot Number’; B [B = ðka þ kc ÞR=Dc; where ka and kc are the cathodic and anodic
rate constants, respectively, and c is the concentration of available sites for Li-insertion], which can
be used to correlate the evolution of Li-concentrations and stresses. This analysis also revealed that
during Li-insertion, maximum tensile stress occurs in the radial direction and at the particle center,
whereas during de-insertion the maximum tensile stress occurs at the particle surface and in the tan-
gential direction. This also tends to support the results obtained by Garcia et al. [87]. Furthermore, for
B  1 (constant surface solute concentration), the tendency for cracking increases during de-insertion
and accordingly becomes more likely at the particle surface. Slightly different results were obtained in
another work by Verbrugge and Cheng [165], where stress development and cracking were related to
cyclic fatigue behavior. Their analysis related the strain energy density to the surface tension and sur-
face modulus, which in turn depends on the Li accumulation (i.e., concentration gradient) at the sur-
face. This predicts that smaller particles should allow better cycling performance, which also supports
the results obtained for lower ‘Electrochemical Biot Number’ (B  1) by the same authors [86]. Fur-
thermore, the cracking was believed to initiate at the surfaces, where the maximum tensile stresses
occur during reversals in the electrochemical half cycles.
A more rigorous analysis by Christensen [85] for DIS in spherical electrode particle showed that
pressure diffusion has a significant influence on the stress response in nearly all types of electrode
materials (low expansion materials like graphite or high expansion materials like Li-alloys), whereas
it affects the voltage response for only the high expansion materials. On the other hand, variability in
the solid-state diffusion of Li-ions affects both the voltage and the stress behavior. It was also estab-
lished that porous electrodes lead to amplification of stresses and concomitantly non-uniform decrep-
itation or at least disordering of the entire electrode. However, fragmented particles near the
separators were found to bear the maximum stress. This result has also been supported in a very re-
cent publication by Purkayastha and McMeeking [84] where it was shown that cathode particles near
the separator experience the most severe conditions in terms of stress development. This model, based
on 2-D Li-ion battery and consideration of Li-transport equations through the electrolyte, in addition
to DIS, also revealed that the magnitude of stress generation varied directly with the expansion/con-
traction of the active particles and charging rate, whereas it varied inversely with the product of the
elastic modulus of the active particles and the Li-partial molar volume.
Deshpande et al. [89] correlated elastic strain energy, diffusion induced stresses, and stress discon-
tinuities at the phase boundaries of the different phases (say, a and b) that form during lithiation/
delithiation using a core–shell structural model. This work predicts that steep concentration gradient
at the phase boundaries lead to stress discontinuities, that varied directly with the equilibrium ratio of
Cb/Ca, where Ci denotes the molar concentration of Li in the respective phases (i = a and b). Hence, the
analysis suggests that avoiding co-existence of phases with large concentration difference should al-
low more stable electrochemical cycling. The mechanism of stress development due to co-existence of
different phases has already been discussed earlier and schematically shown in Fig. 3d. Furthermore, it
was determined that for a given combination of phases, the maximum stress discontinuity at the
interface occurs when XbCb/XaCa = 1, where Xi is the partial molar volume of Li in phase i.
The classic work of Huggins and Nix [69] evaluated the effects of electrode dimension and mechan-
ical properties on the lithiation induced decrepitation in terms of thin film mechanics, by assuming
the electrode strip to be a bi-layer containing pre-existing cracks (of flaw size c). Dimensional
A. Mukhopadhyay, B.W. Sheldon / Progress in Materials Science 63 (2014) 58–116 97

mismatch between the two layers as a result of dilation of the top layer was assumed to lead to stress
development, with fracture occurring on satisfaction of Griffith’s criterion. Based on such basic frac-
ture mechanics theory, a relationship for the critical thickness (hc) beyond which crack propagation
could occur was developed, which depended on the mechanical properties of the electrode material,
such as fracture toughness (KIc), elastic modulus (E), Poisson’s ratio (m) and a lateral dilation (transfor-
mation) strain parameter (eT):
 
23 3K IC ð1  mÞ
hc ¼ ð11Þ
p EeT
Fig. 15b presents the variation of the computationally determined critical size with the strain
parameter and elucidates that for the electrode materials used more commonly (fracture toughness
below 15 MPa m1/2) and for eT > 1, thicknesses in the nanosized regime are required for stable elec-
trochemical cycling. Using this relationship, a critical size of 200 nm was estimated for Si electrodes,
for cycling without fracturing. This result is consistent with recent direct observations on Si nanopar-
ticles and nanorods by Liu et al. [94] and Lee et al. [172], respectively. Extending this classical work,
Bhandakkar and Gao [83] developed a relationship for a critical dimension (Hft) for thin strip elec-
trodes not containing pre-existing flaws for prevention of decrepitation, under conditions of dynamic
evolution of DIS during galvanostatic cycling:
( )1=3
Cð1  mÞF 2 D2
Hft ¼ 13 ð12Þ
Eð1 þ mÞX2 I2

where C is the fracture energy of the electrode material, X is the partial molar volume of Li, D is the Li-
diffusivity and I is the surface current density. This analysis showed that, in addition to the material
properties (C, E and m), the critical dimension of the active electrode material also depends on the de-
gree of lithiation/delithiation (through X), the Li transport kinetics and the electrochemical cycling
rate (through I). Hence, the critical dimension would be less for high power applications (faster
charge/discharge rates) and greater Li-concentration gradients (materials with lower D). This supports
most experimental observations including that of Lee et al. [172], and generally agrees with previous
theoretical models.
Aifantis and Dempsey [70] analyzed crack propagation in composite electrode architectures with
active materials of radius a, embedded in an inactive shell of radius b, as shown in Fig. 15c for the case
of Tin Composite oxides (TCO). In this figure the region corresponding to the ring of radial thickness
equals to (q  a) and depicts the damaged zone containing radial cracks (and supporting only radial
stresses) while D depicts the possible maximum expansion of the active core in the absence of the
constraining matrix. The cracks are assumed to propagate from the active/inactive interface. For this
architecture, a stability index [k = (b/G)  (dg/dq)] was determined based on the balance between
strain energy release rate (G) and energy required for crack propagation, such that crack propagation
is stable for negative values of k (or, as k increases, crack propagation becomes unstable). An example
for the variation of k with q for spherical particles of active Sn embedded in inactive glassy matrix as-
sumed three different cases, that are shown in Fig. 15d: case 1, a battery system constrained by outer
casing; case 2, a traction free outer boundary; and case 3, external force exerted on the inactive glass
shell balanced by the force exerted by the expansion of the active core. Crack propagation is more sta-
ble for case 1, as compared to the other two cases. Similar stability index diagrams were also con-
structed for thin film and fiber like architectures in the same work and also for Si nanospheres
embedded in soda glass matrix in a later work [79]. Extending this approach further, Aifantis et al.
[78] analyzed different combinations of materials for active and inactive sites. A smaller volume frac-
tion of active sites led to more stable behavior and a shorter distance for crack arrest. Mechanical
properties (fracture toughness and strength) of the inactive matrix also influence the crack arrest dis-
tance and critical crack length for fracture. This work predicts that Si active sites should lead to more
stable cyclic performance as compared to Sn.
The effects of particle size and shape on stress development in cathode materials (mainly LiMn2O4)
were analyzed by Zhang et al. [80]. The overall conclusion was that finer and elliptical shaped particles
should suppress stress development. Increase in the aspect ratios of elliptical particles should further
98 A. Mukhopadhyay, B.W. Sheldon / Progress in Materials Science 63 (2014) 58–116

reduce the stress levels (see Fig. 15e). In a recent work, Zhang et al. [82] analyzed the influence of the
relative dimensions and properties of the current collector and active electrode material on the DIS,
assuming that the current collector acts as a substrate which constrains the deformations of the active
electrode material during lithiation/delithiation. It was concluded that thinner current collectors with
smaller elastic modulus (softer and more flexible) should reduce stresses in the active electrode mate-
rials, provided that current collector can bear the enhanced stress levels without fracturing.
Woodford et al. [76] developed an ‘Electrochemical Shock Map’ that can be used to determine
whether a brittle electrode material will fracture at a particular electrochemical cycling rate (C-Rate),
for a given combination of particle size and fracture toughness of the material (Fig. 15f). Not surpris-
ingly, it was predicted that for a given material, smaller particle sizes lead to higher allowable C-Rates
(without fracture). This analysis was based on linear elastic fracture mechanics, using the materials
properties of LiMn2O4 cathode material. Considering the gradients of DIS at the crack tip and pre-exist-
ing flaws, the analysis suggests that smaller pre-existing cracks might grow more unstably than the
larger cracks. As mentioned earlier, similar relationships between the C-Rate, electrode size and
fracture have also been recently reported by Lee et al. [172], based on observations made with Si
nano-pillar electrodes. Hu et al. [68] applied basic fracture mechanics principles, based on competition
between energy required to create fracture surfaces (c) and release of elastic strain energy (G) to cal-
culate the critical particle size needed to avert lithiation/delithiation induced fracture in LiFePO4.
Estimates of the elastic strain mismatch (em) between adjacent regions with different Li-concentrations
(LiFePO4 and FePO4) show that this is the primary contribution to the stored elastic strain energy. A
relationship for the critical particle size (dcritical) needed to avoid fracture was then estimated as:

2c
dcritical ¼ ð13Þ
Z max Ee2m

where Zmax is the maximum value of a function of the length of the crack (L) and particle diameter (d).
Values of dcritical for platelet shaped particles and equiaxed particles were estimated to be 600 nm
and 60 nm, respectively.
In a recent and innovative work, Huang and Wang [73] have modeled deformation and stress in
LiFePO4 via the generation of dislocations due to Li-ion diffusion along the b-axis, and the correspond-
ing interaction of multiple dislocations. The observed cracking along the c-axis, with the fracture sur-
face orientations along the (1 0 0) and (0 1 0) planes, was simulated based on the stresses associated
with edge and screw dislocations formed due to inhomogeneous lithiation (one such case is illustrated
schematically in Fig. 8f). Furthermore, the study also considered the effects of charging/discharging
rates on the build-up of elastic/plastic energy and, in accordance with many other studies, showed
that the stored energies increase with the electrochemical cycling rates.

8. Stress management in electrode materials

Although experimental investigations of stress development in Li-ion battery electrodes are in their
infancy, extensive research over the last decade has been directed towards addressing stress related
problems. The following discussion highlights some of the laboratory scale strategies that have suc-
cessfully minimized stress related damages and improved the cycle life. Note also that most of these
methods are associated with some compromise in battery performance.

8.1. Formation of alloys and composites

In order to incorporate a second phase or matrix that would act as a buffer towards the dilation/
contraction related stresses, several research groups have been investigating the effects of using alloys
based on the metallic anode materials (such as Si, Sn, Al, Sb, etc.). Logical choices are matrix materials
with a high yield strength and low elastic modulus, so that the elastic strains from the deformation of
the primary active material can be accommodated by elastic deformation of the matrix. Two types of
metallic alloys that accomplish this have been investigated: Inactive-matrix based alloys and Active-
matrix based alloys. For the former, materials such as Cu [179], Nb [180], Co [181], FeSi2 [182], SnFe3C
A. Mukhopadhyay, B.W. Sheldon / Progress in Materials Science 63 (2014) 58–116 99

Fig. 16. Effects of (a) alloy [100] and (a–e) composite [100–103] formation on the electrochemical behavior of the higher
capacity metallic and oxide anode materials, as compared to their (f) unreinforced/unalloyed counterparts [103].

[183] form in situ during ‘displacive-type’ lithiation [10,13,14,184,185] of the corresponding alloys and
compounds such as Co3N [184], Cu6Sn5 [186] and NbSb2 [180], CoSb3 [187], FeSi6 [182] Sn–Fe–C
[90,188]. The in situ formed Cu, Nb, Co, FeSi2, SnFe3C etc. act as buffers that do not participate in
the actual lithiation/delithiation reactions. In the second type of alloys, such as SnSb
[13,14,166,189], SbAl [13,190] and SnAg [13,191], both phases participate in the lithiation/delithiation
reaction. In general the onset of lithiation occurs at different potentials for the two phases (for exam-
ple; 0.9 V against Li/Li+ for Sb and 0.6 V against Li/Li+ for Sn [13,14,166,189]), allowing one phase to
act as buffer during the lithiation/delithiation of the other phase.
Even though some improvements in the cycle life have been observed with the unalloyed metals,
especially for the inactive-matrix materials, this approach reduces the net Li-capacity. In fact, reversible
100 A. Mukhopadhyay, B.W. Sheldon / Progress in Materials Science 63 (2014) 58–116

specific capacities recorded with most of these alloys are in the same range as graphite [1,18,19],
which negates the primary purpose for using metallic anode materials. Furthermore, as seen in
Figs. 11c and 16a, while some of these alloys exhibit better capacity retention than the unalloyed met-
als, they still exhibit capacity fade, unlike that of graphitic carbon electrodes (refer to Fig. 11c and d).
Furthermore, the decomposition of the inactive-matrix based alloys [91–94,101,102,113,192] con-
sumes some energy, thus reducing the overall energy density of the cell. Additionally, for some alloys,
the decomposition (‘conversion’) reaction can also lead to volume changes [91–94,113].
The concept of an inactive matrix as a stress buffer and inhibitor of particle aggregation during lith-
iation/delithiation led to the development of tin-based amorphous composite oxides (TCO) with com-
positions corresponding to SnMxOy (M = B, P, Al) [193,194]. This work by Idota et al. proposed that
SnO2 participated in the actual lithiation/delithiation, with the remaining glass forming elements cre-
ating an inactive matrix. The reversible capacity of 600–700 mAh/g (2200 mA h/cm3) obtained
with these materials is nearly double that of graphitic carbon. With minimal stress-related capacity
fade (nearly 100% columbic efficiency for 100 cycles), these materials seemed to be a major break-
through in the development of higher capacity anodes. However, an initial irreversible capacity of
400 mAh/g was also observed. Courtney and Dahn [195] subsequently associated the initial irrevers-
ible capacities of such oxides to the formation of metallic Sn from the irreversible ‘conversion’ reaction
of SnO2 with Li during the first lithiation half cycle, which leads to the formation of metallic Sn
particles within the Li2O matrix (see Eqs. (4) and (5)). Several other research groups
[91–94,101,102,113,192] have confirmed this mechanism, which implies that only the Sn particles
that participate in the reversible lithiation/delithiation, with the surrounding amorphous Li2O acting
as an inactive buffer in the in situ formed tin composite oxide (TCO). The first step associated with the
conversion of SnO2 to Sn and Li2O (Eq. (4)) is also expected to lead to some energy loss, and might
result in additional volume changes and concomitant stress development. These phenomena have
not yet been investigated. The other glass forming elements in the material developed by Idota
et al. [193,194] were referred to as merely ‘spectator atoms’ [195], although they helped minimize
the aggregation of Sn particles during continued lithiation/delithiation.
With respect to composite electrodes, Li et al. [101] recently reported that the net specific capacity
and capacity retention during continued electrochemical cycling can be drastically improved if SnO2–
graphene nanosheet layered composites are formed by depositing SnO2 on graphene nanosheets
(GNS) by Atomic Layer Deposition. The improvements are more notable in amorphous SnO2, since
in addition to the buffering effects of GNS, the isotropic nature of amorphous SnO2 further assists in
alleviating the stresses arising from volume changes. Amorphous SnO2–GNS nanocomposites exhibit
negligible capacity fade from second cycle onwards (in fact marginal increase in capacity was ob-
served), retaining a specific capacity of 800 mAh/g after 150 cycles (Fig. 16b). In a similar study,
Zhang et al. [102] showed that proper dispersion and attachment of SnO2 nanoparticles on graphene
sheets via wet chemical route (as schematically depicted in Fig. 16c) resulted in improved capacity
retention with electrochemical cycling, as compared to commercial SnO2 or MWCNT–SnO2 nanocom-
posites. In fact in an earlier work by Fan et al. [47], it was reported that incorporation of SnO2 into the
nanopores of mesoporous carbon (ONTC) also results in improved cycling behavior, compared to unre-
inforced SnO2.
On a somewhat similar note, Saint et al. [103] observed improved cycle performance for Si–C com-
posite electrodes where Si nanoparticles were embedded in a non-porous carbon matrix (Fig. 16d),
with the composite retaining a Li-capacity of 1000 mAh/g after 20 cycles (Fig. 16e). This is a substan-
tial improvement compared to isolated Si particles, which undergo nearly complete capacity fade after
15 cycles (Fig. 16f). Based on Raman spectroscopy, it was suggested that the carbon matrix exerts
compressive stress on the Si particles, which opposes the volume change during lithiation and sup-
presses stress development. Recently Magasinski et al. [196] also demonstrated that coating of an-
nealed carbon black with Si nanoparticles, followed by assembling into porous spherical granules
can lead to the fabrication of a composite anode material that exhibits good capacity retention of
1600 mAh/g for 100 cycles, along with columbic efficiencies of 100%. Similarly, carbon coatings
can also reduce the capacity fade for the metallic anode materials [10,197,198].
Composites of Si and carbon nanofibres, including carbon nanotubes (CNT), have also been inves-
tigated as possible Li-ion battery anodes [9,104–107]. For instance, Zhang et al. [104] prepared a
A. Mukhopadhyay, B.W. Sheldon / Progress in Materials Science 63 (2014) 58–116 101

Fig. 17. Engineering of nanocomposite architectures of active metallic anode materials and carbon to obtain significantly
improved capacity retention [104–107], as compared to the more simple mechanical mixtures of CNT and metal nanoparticles
[199] (see Fig. 17a).

composite electrode (termed as SGM) consisting of Si particles/graphite and multi-walled carbon


nanotubes (MWCNT) via simple ball milling. Improved capacity retention was observed (compared
to pure Si and also Si–graphite composites), and attributed to the improved resilience and formation
of better conducting network in the presence of MWCNT. However, the capacity fade was continuous
with electrochemical cycling (Fig. 17a). In subsequent work Shu et al. [199] revealed that simple mix-
tures of CNT and Si do not lead to improved electrochemical properties, with materials consisting of Si
coated on CNTs [105,200,201] or CNTs grown on Si particles [199]. Also, even for these composites the
stability of the C/Si interface is critical, particularly because of the large difference in volume changes
during lithiation/delithiation [107]. In this regard, Wang and Kumta [105] observed excellent cycling
performance with capacities of 2000 mAh/g for over 20 cycles with Si nanoparticles on vertically
aligned multi-walled carbon nanotube stacks (see Fig. 17b). An interfacial layer of amorphous carbon
between the CNTs and the Si nanoparticles was believed to enhance the interfacial adhesion. With an-
other innovative composite design, consisting of free standing Si film incorporating CNT network (see
Fig. 17c), Cui et al. [106] demonstrated capacity retention of 2000 mAh/g for over 50 cycles. The CNT
102 A. Mukhopadhyay, B.W. Sheldon / Progress in Materials Science 63 (2014) 58–116

network performed the dual functions of acting as a mechanical reinforcement and a percolating
conducting network (eliminating the need for a current collector). The CNT network also held frac-
tured pieces together, thus preventing loss of active material due to stress-related fragmentation
(see inset of Fig. 17c). In another recent work, Krishnan et al. [107] developed functionally graded
composite nanorods (‘nanoscoops’) based on C–Al–Si on stainless steel substrates. The presence of
an Al interlayer between C and Si rendered the composite strain-graded during lithiation/delithiation
since Al undergoes volumetric strain (100%) that is intermediate between carbon (10%) and Si
(300%) upon lithiation. Schematic representations of the architecture of such electrodes, the lithia-
tion induced strain gradation and cross-section SEM images obtained before and after lithiation, are
presented in Fig. 17d. Such gradual transition of strain from C to Si, via intermediate Al layer,
apparently improved the structural stability of the electrode even upon electrochemical cycling at
extremely high rates such as 40 C, where stable capacities of 400 mAh/g were recorded up to 100
cycles (90% capacity retention) (see Fig. 17d).
Binders also buffer stresses in conventional particle based porous composite electrodes. The stan-
dard PVDF based binders can only accommodate stresses of roughly 10%, and several researchers
[158,202] have tried to improve the stress induced failure via the use of elastomeric binders capable
of sustaining strains of nearly 100%. On a similar note, innovative heat treatment of PVDF binder (at
250 °C for 3 h in Ar), which leads to coating of anode active particles (Fe3O4) and current collector with
the binder phase, has also been very recently observed to successfully buffer the stress developments
[203]. It was hypothesized that such heat treatment of the binder beyond its melting point resulted in
the formation of a coating which improved the integrity of the active particles with the current col-
lector (Cu in this case). It must be mentioned here that even though the binders are beneficial in terms
of buffering the stress developments, they do not possess appreciable electronic and ionic conductiv-
ities to facilitate Li+ and e- transport, thus negatively affecting the rate capabilities and making it nec-
essary to use conducting additives.
In cathode materials, composite electrodes consisting of LiCoO2 and CNTs show significantly im-
proved cycle performance over LiCoO2 composites with other more conventional conducting additives
such as acetylene black (AB) and carbon fibers (Fig. 18a) [108]. Among the other techniques for
improving the electrochemical performance of LiCoO2 based cathodes, LiCoO2 particles coated with
oxides, in particular ZrO2 and Al2O3, led to excellent cyclic stability even when cycled above 4.2 V
against Li/Li+ (corresponding to delithiation beyond x = 0.5 in LixCoO2) and concomitantly resulted
in improved specific capacity (170 mAh/g) [22] (see Fig. 18b). It was hypothesized by Cho et al.
[22] that surface layers (10 nm) of superior fracture toughness, especially in the presence of ZrO2
sol–gel based coating, suppressed the delithiation induced expansion of the c-axis lattice parameter
of LiCoO2. Subsequent works by Kim et al. [26] and Chen and Dahn [110,204] also demonstrated
the beneficial effects of metal oxide coatings, irrespective of the composition of the coatings (see
Fig. 18c), but cited possible reasons that are not related to lattice expansions and stress developments.
In a different work, Wang et al. [29] noted that Al substitution for Co (LiAl0.25Co0.75O2) suppresses spi-
nel disordering during electrochemical cycling. Here, it is also possible that the surface solid-solutions
in the coated ZrO2 might exhibit greater stability and hence improved electrochemical performance.
However, it must be mentioned that despite the suppression of spinel disordering in the
LiAl0.25Co0.75O2 alloy in the work of Wang et al. [29], cycling induced strain and structural damage
were still observed.
Since strain induced Jahn–Teller distortion is believed to be a major reason for the capacity fade of
LiMn2O4 cathodes (see Section 3), attempts have been made to increase the average oxidation state of
Mn beyond +3.5 by substitution of Mn with other cations such as Ni2+, Co3+ or La3+ [33,37,38,112].
Fig. 18d shows improved cycling performance for LiLa0.01Mn1.99O4 as compared to the un-doped
LiMn2O4 spinel. Chung et al. [112] demonstrated that just coating of undoped LiMn2O4 film electrodes
with LiM0.05Mn1.95O4 (M = Ni or Co) is sufficient to minimize Jahn–Teller distortion related capacity
fade, without affecting the bulk properties of spinel cathode. This study, assisted by real time strain
monitoring using bending beam method [112] (see Section 6.3), further confirmed that the Jahn–Teller
distortion is a surface phenomenon and occurs due to the dynamic (non-equilibrium) nature of the
electrochemical cycling.
A. Mukhopadhyay, B.W. Sheldon / Progress in Materials Science 63 (2014) 58–116 103

Fig. 18. Effects of (a) carbon-based additives (multi-walled carbon nanotubes or CNTs, vapor deposited carbon fibers or VCF and
acetylene black or AB) [108] and (b and c) metal oxide coatings [22,110] on the cycle performance of LiCoO2. Effects of (d) La
doping [38] and (e) oxide coatings [112] on the cycle performance of LiMn2O4.

8.2. Formation of nanostructures and thin films

One of the avenues for addressing stress related issues and improving cycling stability of electrodes
is the development of nanostructured electrodes. Beneficial effects of nanoscale materials are appar-
ent from most of the computational research on deformation and stresses in Li-ion battery electrodes,
as already discussed in Section 7. The nanoscale refers to either the active material grain or particle
sizes [9,15,44,47,50,102,113,161,163,183,205,206]. Uniform thin films and innovative architecture
based films, often with the characteristic length scales in the nanosized regimes, have also been ex-
plored [48,49,92,93,99–108,114,115,118,119,161,173,178,199–201,205–207]. In fact thin film based
electrodes have direct applications in miniature batteries. With respect to basic materials science
and metallurgy, nanostructured materials can possess improved mechanical properties, as compared
to their coarser counterparts. For example, metals possessing nanosized grains exhibit better yield
strengths [208]. Nanosized brittle ceramic particles or ceramic nanocomposites may possess improved
fracture resistance due to lesser probability of finding critical sized flaws or due to reduced dislocation
104 A. Mukhopadhyay, B.W. Sheldon / Progress in Materials Science 63 (2014) 58–116

Fig. 19. Capacity retention with electrochemical cycling for nanosized/nanocrystalline (a and b) anode [113,114] and (c)
cathode materials [115].

pile-up/twin lengths [209]. On more specific terms with respect to applications of nanomaterials in Li-
ion batteries, smaller structures reduce the transport distances for charge carriers (both Li-ions and
electrons), which minimizes the composition or Li-concentration gradients, especially at higher elec-
trochemical cycling rates. This concomitantly reduces stresses associated with these gradients. Such
nanosize effects are relevant to both anode and cathode materials.
In fact, as also mentioned earlier, direct observations have revealed that Si particles tend not to
fracture during lithiation at sizes below 150 nm [9,161,207,210]. Overall, considerably improved cy-
cling stability for nanosized Si and Sn particles, as compared to the coarser (more standard) sized par-
ticles, were reported by Graetz et al. (Fig. 11b) [161] and Pereira et al. (Fig. 19a) [113], respectively. In a
very recent work [114], further advancement of nanofabrication to engineer spaces between Si nano-
particles encapsulated within a shell of hollow carbon tubes (schematically presented as inset in
Fig. 19b), to allow relatively unhindered expansion of the Si nanoparticles during lithiation, led to
excellent retention of specific capacity of 1000 mAh/g up to 200 cycles (see Fig. 19b). Another advan-
tage of the architecture, especially of the hollow carbon shell, was the prevention of dynamic forma-
tion and destruction of the SEI layer directly on the Si nanoparticles (schematically shown as inset in
Fig. 19b).
A. Mukhopadhyay, B.W. Sheldon / Progress in Materials Science 63 (2014) 58–116 105

On the cathode side, 70 nm sized LiCoO2 particles have been reported [115] to possess much better
capacity retention with respect to particles of sizes 0.3 lm (see Fig. 19c). However, Okubo et al. [211]
observed that particle size reduction below 15 nm can also be detrimental with respect to cyclabil-
ity, mainly due to enhancement of the deleterious reactions with the electrolyte, especially at the
higher voltages. With respect to LiFePO4, no cracking was observed in particles of sizes below
30 nm even on cycling at the higher electrochemical cycling rates [212]. It can be envisaged that
the critical sizes for resistance to cycling induced fracture is lower for the cathode materials by factors
of 2–5, as compared to the anode materials. We believe that this is due to the inherent brittleness of
the ceramic cathode materials, despite the dimensional changes being significantly lesser as compared
to the metallic anode materials.
Thin films have frequently been observed to result in better capacity retention as compared to par-
ticle based electrodes (see Fig. 11b) [9,10,34,161]. Coulombic Efficiencies near 100% have also been ob-
tained with thin film electrodes [9,10,34,161,213]. Such beneficial effects have been observed both for
anode materials, such as Si [9,10,161,213] that seem to suffer most from stress related capacity fading,
as well as cathode materials, such as LiMn2O4 [34] that experiences significant stress due to Jahn–Tell-
er distortion. However, it has been usually found that such beneficial effects of thin films are typical
for film thicknesses below 100 nm [52,123,161], which severely limit the energy density that can be
attained for practical purposes. In order to explain such behavior, Graetz et al. [161] calculated the
critical crack size (ac) for fracturing of polycrystalline Si films based on the following relation:

ac ¼ ð2=pÞðK Ic =rÞ2 ð14Þ


1/2
where KIc is the mode I fracture toughness (0.75 MPa m ) and r is the yield strength (1.1 GPa). A
value of 300 nm was obtained for the critical crack size, which could tentatively explain the minimi-
zation of stress induced cracking for films less than 100 nm. On a similar note, Xiao et al. [52] related
the accrued cracking behavior of thicker films to greater strain energy release rates (G), according to
the following relation:

pð1  t2f Þr2f h


G ¼ gða; bÞ ð15Þ
2Ef

where g(a, b) is a function that characterize the elastic mismatch between the film and the substrate, Ef
is the effective elastic modulus, rf is the average film stress normal to the cracks and h is the film
thickness.
In thin film electrodes, stiff substrates constrain dimensional changes in two dimensions (in plane)
directions. To better accommodate these volume changes, Yu et al. [116] recently explored the use of
softer substrates. Si thin films on poly(dimethylsiloxane) or PDMS substrates were reported to exhibit
excellent mechanical stability during continued electrochemical cycling up to film thicknesses of
400 nm, as opposed to the thickness limit of just 100 nm on stiffer substrates. In fact, the novel
Si-based anodes could be cycled through the entire potential range (i.e. below the potential corre-
sponding to onset of Li15Si4 formation) to attain near theoretical capacity (>4000 mAh/g) with 85%
capacity retention up to 500 cycles (see Fig. 20a). The stress relaxation in these films was attributed
to buckling (out-of-plane deformation) of the film-substrate system, as shown schematically in
Fig. 20a. The superior performance in the presence of softer and more flexible substrate also supports
the results of computational research by Zhang et al. [82] (see Section 7). The work Cui et al. [106] with
free-standing Si–CNT composite thin film electrodes are also believed to relieve the stresses via form-
ing ripples in these more complex structures (see Section 8.1).
Based on the observation [52] that average crack spacing in continuous Si films after lithiation was
7 lm (see Fig. 10a), Xiao et al. [52] and Soni et al. [117] used patterned Si thin films (100 nm thick)
with varying island sizes from 40 lm down to 7 lm (see Fig. 20b) for understanding the correla-
tion between the island dimensions, deformation, stress development and electrochemical behavior.
Islands below the same critical size of 7 lm suppressed cracking during lithiation/delithiation and
hence improved capacity retention, compared to the continuous films. It had also been earlier evi-
denced by Beaulieu et al. [160] that after cracks form during the first delithiation half cycle of thin film
electrodes, the separated islands (by virtue of cracks) tend not to undergo fracture during further
106 A. Mukhopadhyay, B.W. Sheldon / Progress in Materials Science 63 (2014) 58–116

Fig. 20. (a) Capacity retention with electrochemical cycling for Si thin film anode on flexible substrate (see inset) [116]. (b)
Variation of stress, as measured in situ using MOSS, with electrochemical cycling for patterned a-Si thin film anodes (see inset)
[52,117]. Capacity retention with electrochemical cycling for (c) Sn nanorods based anodes (see inset), showing better capacity
retention as compared to the corresponding planar thin film anode [49]; (d) LiMnO2 nanorods based cathodes (see inset),
showing better capacity retention compared to LiMnO2 nanoparticles [118].
A. Mukhopadhyay, B.W. Sheldon / Progress in Materials Science 63 (2014) 58–116 107

cycling. Additionally, Xiao et al. [52] and Soni et al. [117] observed that plastic flow and concomitant
deleterious effects were also suppressed for islands below the critical size (compare Fig. 20b with
Fig. 14b). Such behavior was explained in terms of shear lag model, with the assumption that the
stresses in the patterned islands were balanced by the shear resistance at the island/current collector
interface. It was proposed that the lateral size of the shear lag zone (Lp) during lithiation of the Si
islands is given by [52,117]:

rc h
Lp ffi ð16Þ
sint
where rc is (maximum) stress at the center of the island, h is the film thickness and sint is the inter-
facial shear strength between the island and the current collector (see Fig. 21). Hence, the islands will
deform elastically if the size (2L) is such that Lp reaches L before rc reaches ry (the flow stress of the
island material). This allows the estimation of the critical (half) island size (Lcr) below which no plastic
deformation, and concomitantly no plastic strain localization will occur in the Si islands, and is given
by [52,117]:

Fig. 21. Schematic illustrations of (a) side view of lithiated Si island with island size of 2L, shear lag zone distance 2Lp, and film
thickness h; (b) free body diagram of half of the lithiated Si island; (c) top view describing the shear lag model for 7 lm, 17 lm
and 40 lm Si islands sizes (2L) [117].
108 A. Mukhopadhyay, B.W. Sheldon / Progress in Materials Science 63 (2014) 58–116

ry h
Lcr ffi ð17Þ
sint
Since sint will be limited by either the shear flow stress of the current collector or the interfacial fric-
tion strength, Eq. (17) results in an estimated value of between 5 lm and 9 lm for 2Lcr. This value was
found to agree reasonably well with the observed island size of 7 lm which suppressed plastic
deformation and cracking during lithiation/delithiation. The analysis in Ref. [117] also indicated that
the interfacial sliding between the Ti current collector and Si islands, and not plastic deformation of
the current collector, accommodated the lithiation induced dilation of the Si islands within the shear
lag zone. Another observation was that while the degradation of continuous films are mostly related to
through-thickness cracking, the primary degradation mechanism of patterned films below the critical
size is interfacial delamination, which is in turn related to the interfacial shear strength between the
active Si film and current collector [52]. Such delamination induced failure was again suppressed be-
low the critical island size of 7 lm.
Based on more conventional belief, another plausible reason for the improved behavior of the pat-
terned films is that the gap between the islands in the patterned films acted as buffer for accommo-
dating the volume expansions and led to stress relaxation; a principle somewhat similar to the
engineering of empty spaces between Si nanoparticles within a hollow carbon shell, as described ear-
lier [114]. With respect to required separation between the islands on a patterned film, in a very re-
cent investigation He et al. [136] observed that individual islands (lateral dimension: 5 lm  5 lm;
thicknesss: 500 nm) of patterned Si thin films tend to merge together during lithiation, unless the dis-
tance of separation of the patterns are >1 lm.
Similar to the patterned thin films, nanowires/nanotubes of active electrode materials, preferably
on substrates acting as current collectors lead to better cycling stability [118,119,200,214]. Under
optimized processing condition, crystalline Si nanowires can suppress formation of the deleterious
crystalline Li15Si4 phase even on lithiation up to potentials as low as 30 mV against Li/Li+ [119]. This
would allow cycling to near theoretical capacity (4200 mAh/g), without capacity loss associated with
degradation due to the phase transformation. Similar behavior was also reported for a-Si films of
thicknesses below 50 nm [215]. In a different work, nanorods based on Sn-coated Cu [49]
(Fig. 20c), TiO2 [216] and SnO2 [217] nanofibres protruding from current collectors also show better
capacity retention compared to the corresponding thin film based electrodes. Nanorod morphologies
have also been extended to the cathodes, where improved capacity retention, along with higher spe-
cific capacities, have been observed [118,218,219]. For example, Fig. 20d shows that LiMnO2 nanorods
lead to significantly improved performance with respect to even nanoparticles of the same composi-
tion [118]. The specific capacities of 200 mAh/g obtained with such nanoparticles are among the
highest capacities reported for the layered cathode materials. Furthermore, such nanostructuring al-
lows better rate capabilities and reduces mechanical degradation at higher electrochemical cycling
rates. As with most nanomaterials, nanowire electrodes have also not yet been used in commercial
batteries because of the expensive and stringent processing conditions required for their successful
fabrication.
Allowing empty spaces for accommodation of the lithiation induced volume expansion have also
been accomplished by introducing mesopores in the active electrode materials, leading to improved
cycling stability [203,205,220,221]. Furthermore, wider pore size distributions (between 2 nm and
8 nm) lead to better electrochemical performance for mesoporous SnO2 based anodes [220]. Reduced
stiffness due to varying pore size distribution has also been tentatively associated with such improved
cycling stability. This is in accordance with the report by Mukaibo et al. [91], where real time obser-
vations of stress development in Sn and Ni – 62% Sn thin film electrodes revealed lower stress mag-
nitude in the pure Sn electrode owing to lower elastic modulus.
Even though nanostructuring is advantageous with respect to stability of electrochemical cycling,
the major problems with the nanostructured materials are the challenges/expenses incurred during
their processing and the greater specific surface areas that leads to accrued parasitic side reactions
with the electrolyte, leading to higher irreversible consumption of Li-ions (to form SEI layer). Thus,
for rendering practical applications of nanomaterials as Li-ion battery electrodes commercially feasi-
ble, progress in research should lead to the advantages in terms of increase in cycle life and the ability
A. Mukhopadhyay, B.W. Sheldon / Progress in Materials Science 63 (2014) 58–116 109

to cycle at higher electrochemical cycling rates that significantly outweigh the disadvantages, mainly
in terms of processing challenges and occurrences of side (surface) reactions. This warrants reducing
not only the particle sizes, but devising novel nano-architectures.

8.3. Limiting the usable Li-capacity

One of the strategies used to minimize the deleterious effects of stresses associated with the for-
mations of certain crystal phases is to cycle the electrode material within potential ranges that would
prevent formation of such phases [9,10,17,20–39,92–94,98,119,220]. This leads to improved capacity
retention for both anode and cathode materials, but at the cost of specific capacities and energy den-
sities. For instance, formation of the deleterious crystalline Li15Si4 phase during lithiation of Si based
anodes at potentials below 50 mV against Li/Li+ necessitates using a cut-off voltage that is higher
than the potential for the onset of formation of this phase [9,10,17,92–94,98,119,222]. Hence, the
maximum attainable specific capacity from Si drops down to 3600 mAh/g, as compared to the the-
oretical value of 4200 mAh/g in case of full lithiation up to Li22Si5. In a very recent work, Misra et al.
[119] showed that Si nanowires exhibit much improved cycling behavior when cycled within a fixed
potential range that would lead to a lower constant capacity of 1000 mAh/g (corresponding to x = 1.05
in LixSi), which is less than the theoretical capacity of Si by a factor of 4. The nanowires exhibited
more drastic capacity fade even when cycled within potential range corresponding to a capacity of just
2000 mAh/g, which is still half the theoretical capacity of Si. In a very recent work, Aravindan et al.
[223] reported excellent cyclability, i.e. essentially no capacity fade up to 500 cycles, for SnO2 thin film
electrodes (fabricated via atomic layer deposition) on cycling within a restricted potential range of
0.005 and 0.8 V against Li/Li+. This prevented the reversible conversion reactions of Sn getting oxidized
to SnO2 beyond 0.8 V and concomitant stresses arising from the volume changes. However, this lim-
ited the attainable reversible Li-capacity to 650 mAh/g (at current density of 5 lA/cm2), instead of
theoretically possible capacity of 1500 mAh/g on allowing the conversion oxidation reaction leading
to the formation of SnO2.
Due to similar reasons, the delithiation of LiCoO2 is historically limited to the composition of Li0.5-
CoO2, that is up to an upper cut-off potential of 4.25 V against Li/Li+. This allows obtaining a specific
capacity of 120 mAh/g, which is half the theoretical capacity upon full delithiation (240 mAh/g). In
order to suppress the Jahn–Teller distortion occurring during lithiation of LixMn2O4 beyond x = 1, the
lower cut-off voltage is set above 3 V, which also negates the possibility of obtaining the theoretically
possible maximum specific capacity of 280 mAh/g on progressive lithiation up to x = 2 (Li2Mn2O4)
[31–39]. Reduced potential ranges also directly reduce the energy density of the cell, in addition to
reducing the energy density via reduction in electrode capacity (since; energy density = net cell capac-
ity  operating potential range). Hence, such strategies should have been the last resort towards alle-
viating the stress related issues. It must also be mentioned here that for reasons primarily related to
safety and stability of electrolytes, batteries are cycled over reduced range for various commercial
applications. However, to the best of the authors’ knowledge, there is no report available yet in the
open literature, which has been able to convincingly address these issues, especially related to the sac-
rifice of capacities of the cathode materials.

9. Conclusion and outlook

Deformation and stresses in electrode materials during electrochemical cycling of Li-ion batteries
lead to fracture/disintegration of the active electrode materials, loss of contact with the current collec-
tors and exposure of fresh electrode surfaces, resulting in the continued uncontrolled formation of
passivation layer (SEI). Such phenomena contribute significantly towards the decrease in Li-capacity
of the electrodes and eventual failure of the electrode material, and hence the entire Li-ion battery.
The predominant cause for such stress developments can be summarized in terms of lithiation/delith-
iation induced dilation/contraction of the active materials, and the constraints imposed by the current
collector/substrate or neighboring particles/regions. Furthermore, the stress developments get ac-
crued at the higher electrochemical cycling rates.
110 A. Mukhopadhyay, B.W. Sheldon / Progress in Materials Science 63 (2014) 58–116

Though nearly all electrode materials suffer from stress related degradation, the stress related fail-
ures are extremely severe for high capacity Li-alloy anodes (such as Sn, Si, Al, Ge), since these elec-
trodes undergo volume expansions up to 300% or more upon full lithiation (compared to
maximum of 14% for the presently used graphitic carbon anodes). These issues have prevented
the use of the Li-alloy anodes in commercial batteries to date, and hence have been one of the bottle-
necks towards enhancement of the energy densities of the Li-ion batteries.
Since stress development in electrode materials is a major issue, extensive research efforts have
been devoted towards understanding the various causes, analyzing the effects of the different param-
eters related to electrochemical cycling and materials properties, modeling the stress developments
via computational research and devising strategies to minimize the negative impacts of such stresses.
Two important approaches for improving performance are reducing the dimensions of the active elec-
trode materials to nanosize regimes and the use of composites. These approaches are manifested in a
wide range of innovative nanostructures/architectures such as nanopillars, nanorods, patterned thin
films and free standing nanoelectrodes. Widespread fabrication of these materials has been conducted
at the laboratory scale, with numerous demonstrations of improved electrochemical performance (cy-
cle lives). However, the expenses and intricacies involved in such fabrication methods have prevented
commercial take-up.
Recent research on direct observations of deformations occurring during lithiation/delithiation of
the different electrode materials, preferably via in situ TEM, have provided some valuable information
on the deformations in Sn, Si and CNTs. However, this is still at its infancy and need to be extended to
the other electrode materials. Additionally, very few research reports are available that actually report
the experimentally determined magnitudes of the stresses developed during electrochemical cycling.
The authors were among the few researchers to monitor these stresses in thin film electrodes in situ
during electrochemical cycling. These types of measurements make it possible to quantify stresses
(e.g., 1 GPa flow stresses in lithiated amorphous Si). This approach has also been used to demon-
strate that substantial stresses arise during SEI layer formation, and has also been modified to measure
other properties. Further application of this technique is expected to provide important insight into a
wider range of materials. More results based on such in situ observations and stress measurements,
coupled with those obtained via computational research and with laboratory scale experiments on
composite or nanostructured electrodes, are necessary to allow development of higher capacity elec-
trodes possessing improved cycle lives and rate capabilities, which can be used in commercial Li-ion
batteries in the near future.

References

[1] Tarascon JM, Armand M. Issues and challenges facing rechargeable lithium batteries. Nature 2001;414:359.
[2] Dresselhaus MS, Thomas IL. Alternative energy technologies. Nature 2001;414:332.
[3] Besenhard JO. Handbook of battery materials. Weinheim: Wiley-VCH; 1999.
[4] Shukla AK, Kumar TP. Materials for next-generation lithium batteries. Curr Sci 2008;94:314.
[5] Kali R, Mukhopadhyay A. Spark plasma sintered/synthesized dense and nanostructured materials for solid-state Li-ion
batteries: overview and perspective. J Power Sources 2014;247:920.
[6] Broussely M, Archdale G. Li-ion batteries and portable power source prospects for the next 5–10 years. J Power Sources
2004;136:386.
[7] Fu LJ, Liu H, Li C, Wu YP, Rahm E, Holze R, et al. Electrode materials for lithium secondary batteries prepared by sol–gel
methods. Prog Mater Sci 2005;50:881.
[8] Besenhard J, Winter M. Advances in battery technology: rechargeable magnesium batteries and novel negative-electrode
materials for lithium ion batteries. ChemPhysChem 2002;3:155.
[9] Kasavajjula U, Wang C, Appleby AJ. Nano- and bulk-silicon-based insertion anodes for lithium-ion secondary cells. J Power
Sources 2007;163:1003.
[10] Zhang WJ. A review of the electrochemical performance of alloy anodes for lithium-ion batteries. J Power Sources
2011;196:13.
[11] Larcher D, Beattie S, Morcrette M, Edstrom K, Jumas J-C, Tarascon J-M. Recent findings and prospects in the field of pure
metals as negative electrodes for Li-ion batteries. J Mater Chem 2007;17:3759.
[12] Obrovac MN, Christensen L, Le DB, Dahn JR. Alloy design for lithium-ion battery anodes. J Electrochem Soc
2007;154:A849.
[13] Winter M, Besenhard JO. Electrochemical lithiation of tin and tin-based intermetallics and composites. Electrochim Acta
1999;45:31.
[14] Besenhard JO, Yang J, Winter M. Will advanced lithium-alloy anodes have a chance in lithium-ion batteries? J Power
Sources 1997;68:87.
A. Mukhopadhyay, B.W. Sheldon / Progress in Materials Science 63 (2014) 58–116 111

[15] Arico AS, Bruce P, Scrosati B, Tarascon J-M, Schalkwijk WV. Nanostructured materials for advanced energy conversion and
storage devices. Nature Mater 2005;4:366.
[16] Flandrois S, Simon B. Carbon materials for lithium-ion rechargeable batteries. Carbon 1999;37:165.
[17] Winter M, Besenhard JO, Spahr ME, Novak P. Insertion electrode materials for rechargeable lithium batteries. Adv Mater
1998;10:725.
[18] Endo M, Kim C, Nishimura K, Fujino T, Miyashita K. Recent development of carbon materials for Li ion batteries. Carbon
2000;38:183.
[19] Mukhopadhyay A, Tokranov A, Sena K, Xiao X, Sheldon BW. Thin film graphite electrodes with low stress generation
during Li-intercalation. Carbon 2011;49:2742.
[20] Liu H, Wu YP, Rahm E, Holze R, Wu HQ. Cathode materials for lithium ion batteries prepared by sol-gel methods. J Solid
State Eletrochem 2004;8:450.
[21] VanderVen A, Aydinol MK, Ceder G, Kresse G, Hafner J. First-principles investigation of phase stability in LixCoO2. Phys Rev
B 1998;58:2975.
[22] Venkatraman S, Subramanian V, Gopu Kumar S, Renganathan NG, Muniyandi N. Capacity of layered cathode materials for
lithium-ion batteries – a theoretical study and experimental evaluation. Electrochem Commun 2000;2:18.
[23] Cho J, Kim YJ, Park B. LiCoO2 cathode material that does not show a phase transition from hexagonal to monoclinic phase.
J Electrochem Soc 2001;148:A1110.
[24] Cho J, Kim YJ, Park B. Novel LiCoO2 cathode material with Al2O3 coating for a Li ion cell. Chem Mater 2000;12:3788.
[25] Ammatucci CC, Tarascon JM, Klein LC. Cobalt dissolution in LiCoO2-based non-aqueous rechargeable batteries. Solid State
Ionics 1996;83:167.
[26] Kim YJ, Cho J, Kim TJ, Park B. Suppression of cobalt dissolution from the LiCoO2 cathodes with various metal-oxide
coatings. J Electrochem Soc 2003;150:A1723.
[27] Sun YK. Cycling behaviour of LiCoO2 cathode materials prepared by PAA-assisted sol–gel method for rechargeable lithium
batteries. J Power Sources 1999;83:223.
[28] Zhou M, Yoshio M, Gopukumar S, Yamaki J. Synthesis of high-voltage (4.5 V) cycling doped LiCoO2 for use in lithium
rechargeable cells. Chem Mater 2003;15:4699.
[29] Wang H, Jang Y, Huang B, Sadoway DR, Chiang Y-M. Electron microscopic characterization of electrochemically cycled
LiCoO2 and Li(Al, Co)O2 battery cathodes. J Power Sources 1999;81–82:594.
[30] Wang H, Jang Y-I, Huang B, Sadoway DR, Chiang Y-M. TEM study of electrochemical cycling-induced damage and disorder
in LiCoO2 cathodes for rechargeable lithium batteries. J Electrochem Soc 1999;146:473.
[31] Thackeray MM. Structural considerations of layered and spinel lithiated oxides for lithium ion batteries. J Electrochem Soc
1995;142:2558.
[32] Park YJ, Kim JG, Kim MK, Chung HT, Um WS, Kim MH, et al. Fabrication of LiMn2O4 thin films by sol–gel method for
cathode materials of microbattery. J Power Sources 1998;76:41.
[33] Sun YK, Jeon YS, Lee HJ. Overcoming Jahn–Teller distortion for spinel Mn phase. Electrochem Solid-State Lett 2000;3:7.
[34] Park YJ, Kim JG, Kim MK, Kim HG, Chung HT, Park Y. Electrochemical properties of LiMn2O4 thin films: suggestion of
factors for excellent rechargeability. J Power Sources 2000;87:69.
[35] Thackeray MM, Horn YS, Kahaian AJ, Kepler KD, Skinner E, Vaughey JT, et al. Structural fatigue in spinel electrodes in high
voltage (4 V) Li/LixMn2O4 Cells. Electrochem Solid-State Lett 1998;1:7.
[36] Gummow RJ, Kock AD, Thackeray MM. Improved capacity retention in rechargeable 4 V lithium/lithium manganese oxide
(spinel) cells. Solid State Ionics 1994;69:59.
[37] Amatucci GG, Pereira N, Zheng T, Tarascon J-M. Failure mechanism and improvement of the elevated temperature cycling
of LiMn2O4 compounds through the use of the LiAlxMn2xO4zFz solid solution. J Electrochem Soc 2001;148:A171.
[38] Tu J, Zhao XB, Zhuang DG, Cao GS, Zhu TJ, Tu JP. Studies of cycleability of LiMn2O4 and LiLa0.01Mn1.99O4 as cathode
materials for Li-ion battery. Physica B 2006;382:129.
[39] Shin Y, Manthiram A. Factors influencing the capacity fade of spinel lithium manganese oxides. J Eectrochem Soc
2004;151:A204.
[40] Christensen J, Newman J. A mathematical model of stress generation and fracture in lithium manganese oxide. J
Electrochem Soc 2006;153:A1019.
[41] Padhi AK, Nanjundaswamy KS, Goodenough JB. Phospho-olivines as positive-electrode materials for rechargeable lithium
batteries. J Electrochem Soc 1997;144:1188.
[42] Lepage D, Michot C, Liang G, Gauthier M, Schougaard SB. A soft chemistry approach to coating of LiFePO4 with a
conducting polymer. Angew Chem Int Ed 2011;50:6884.
[43] Shi JY, Yi C-W, Kim K. An investigation of LiFePO4/poly(3,4-ethylenedioxythiophene) composite cathode materials for
lithium-ion batteries. Bull Korean Chem Soc 2010;31:2698.
[44] Dinh H-C, Mho S-I, Yeo I-H. Electrochemical analysis of conductive polymer-coated LiFePO4 nanocrystalline cathodes
with controlled morphology. Electroanalysis 2011;23:2079.
[45] Wang GX, Yang L, Chen Y, Wang JZ, Bewlay S, Liu HK. An investigation of polypyrrole-LiFePO4 composite cathode
materials for lithium-ion batteries. Electrochim Acta 2005;50:4649.
[46] Maranchi JP, Hepp AF, Evans AG, Nuhfer NT, Kumta PN. Interfacial properties of the a-Si/Cu: active–inactive thin-film
anode system for lithium-ion batteries. J Electrochem Soc 2006;153:A1246.
[47] Fan J, Wang T, Yu C, Tu B, Jiang Z, Zhao D. Ordered, nanostructured tin-based oxides/carbon composites as the negative
electrode material for lithium-ion batteries. Adv Mater 2004;16:1432.
[48] Kumar TP, Ramesh R, Lin YY, Fey GT-K. Tin-filled carbon nanotubes as insertion anode materials for lithium-ion batteries.
Electrochem Commun 2004;6:520.
[49] Bazin L, Mitra S, Taberna PL, Poizot P, Gressier M, Menu MJ, et al. High rate capability pure Sn-based nano-architectured
electrode assembly for rechargeable lithium batteries. J Power Sources 2009;188:578.
[50] Li H, Shi L, Wang Q, Chen L, Huang X. Nano-alloy anode for lithium ion batteries. Solid State Ionics 2002;148:247.
[51] Paek S-M, Yoo EJ, Honma I. Enhanced cyclic performance and lithium storage capacity of SnO2/graphene nanoporous
electrodes with three-dimensionally delaminated flexible structure. Nano Lett 2009;9:72.
112 A. Mukhopadhyay, B.W. Sheldon / Progress in Materials Science 63 (2014) 58–116

[52] Xiao X, Liu P, Verbrugge MW, Haftbaradaran H, Gao H. Improved cycling stability of silicon thin film electrodes through
patterning for high energy density lithium batteries. J Power Sources 2011;196:1409.
[53] Haftbaradaran H, Soni SK, Sheldon BW, Xiao X, Gao H. Modified stoney equation for patterned thin film electrodes on
substrates in the presence of interfacial sliding. J Appl Mech 2012;79:031018.
[54] Mukhopadhyay A, Guo F, Tokranov A, Xiao X, Hurt RH, Sheldon BW. Engineering of graphene layer orientation to attain
high rate capability and anisotropic properties in Li-ion battery electrodes. Adv Funct Mater 2013;23:2397.
[55] Guo F, Mukhopadhyay A, Sheldon BW, Hurt RH. Vertically aligned graphene layer arrays from chromonic liquid crystal
precursors. Adv Mater 2011;23:508.
[56] Kang B, Ceder G. Battery materials for ultrafast charging and discharging. Nature 2009;458:190.
[57] Kang K, Meng YS, Breger J, Grey CP, Ceder G. Electrodes with high power and high capacity for rechargeable lithium
batteries. Science 2006;311:977.
[58] Huang H, Yin S-C, Nazar LF. Approaching theoretical capacity of LiFePO4 at room temperature at high rates. Electrochem
Solid-State Lett 2001;4:A170.
[59] Buqa H, Goers D, Holzapfel M, Spahr ME, Novak KP. High rate capability of graphite negative electrodes for lithium-ion
batteries. J Electrochem Soc 2005;152:A474.
[60] Christensen J, Newman J. Stress generation and fracture in lithium insertion materials. J Solid State Electrochem
2006;10:293.
[61] Wang D, Wu X, Wang Z, Chen L. Cracking causing cyclic instability of LiFePO4 cathode material. J. Power Sources
2005;140:125.
[62] Gabrisch H, Wilcox J, Doeff MM. TEM study of fracturing in spherical and plate-like LiFePO4 particles. Electrochem Solid-
State Lett 2008;11:A25.
[63] Wang Y, Huang H-YS. Proceedings of the ASME 2011 international mechanical engineering congress & exposition
IMECE2011, November 11–17 2011, Denver, Colorado, USA.
[64] Liu Y, Zheng H, Liu XH, Huang S, Zhu T, Wang J, et al. Lithiation-induced embrittlement of multiwalled carbon nanotubes.
ACS Nano 2011;5:7245.
[65] Li J, Dozier AK, Li Y, Yang F, Cheng Y-T. Crack pattern formation in thin film lithium-ion battery electrodes. J Electrochem
Soc 2011;158:A689.
[66] Qi Y, Harris SJ. In situ observation of strains during lithiation of a graphite electrode. J Electrochem Soc 2010;157:A741.
[67] Zhu Y, Wang C. Strain accommodation and potential hysteresis of LiFePO4 cathodes during lithium ion insertion/
extraction. J Power Sources 2011;196:1442.
[68] Hu Y, Zhao X, Suo Z. Averting cracks caused by insertion reaction in lithium-ion batteries. J Mater Res 2010;25:1007.
[69] Huggins RA, Nix WD. Decrepitation model for capacity loss during cycling of alloys in rechargeable electrochemical
systems. Ionics 2000;6:57.
[70] Aifantis KE, Dempsey JP. Stable crack growth in nanostructured Li-batteries. J Power Sources 2005;143:203.
[71] Golmon S, Maute K, Dunn ML. Numerical modeling of electrochemical–mechanical interactions in lithium polymer
batteries. Comput Struct 2009;87:1567.
[72] Thomas JP, Qidwai MA. The design and application of multifunctional structure-battery materials systems. JOM
2005;57:18.
[73] Huang H-YS, Wang Y-X. Dislocation based stress developments in lithium-ion batteries. J Electrochem Soc
2012;159:A815.
[74] Cheng Y-T, Verbrugge MW. Evolution of stress within a spherical insertion electrode particle under potentiostatic and
galvanostatic operation. J Power Sources 2009;190:453.
[75] Harris SJ, Deshpande RD, Qi Y, Dutta I, Cheng Y-T. Mesopores inside electrode particles can change the Li-ion transport
mechanism and diffusion-induced stress. J Mater Res 2010;25:1433.
[76] Woodford WH, Chiang Y-M, Carter WC. ‘‘Electrochemical shock’’ of intercalation electrodes: a fracture mechanics
analysis. J Electrochem Soc 2010;157:A1052.
[77] Soni SK, Sheldon BW, Xiao X, Bower AF, Verbrugge MW. Diffusion mediated lithiation stresses in Si thin film electrodes. J
Electrochem Soc 2012;159:A1520.
[78] Aifantis KE, Hackney SA, Dempsey JP. Design criteria for nanostructured Li-ion batteries. J Power Sources 2007;165:874.
[79] Aifantis KE, Dempsey JP, Hackney SA. Cracking in Si-based anodes for Li-ion batteries. Rev Adv Mater Sci 2005;10:403.
[80] Zhang X, Shyy W, Sastry AM. Numerical simulation of intercalation-induced stress in Li-ion battery electrode particles. J
Electrochem Soc 2007;154:A910.
[81] Zhang X, Sastry AM, Shyy W. Intercalation-induced stress and heat generation within single lithium-ion battery cathode
particles. J Electrochem Soc 2008;155:A542.
[82] Zhang J, Lu B, Song Y, Ji X. Diffusion induced stress in layered Li-ion battery electrode plates. J Power Sources
2012;209:220.
[83] Bhandakkar TK, Gao H. Cohesive modeling of crack nucleation under diffusion induced stresses in a thin strip:
implications on the critical size for flaw tolerant battery electrodes. Int J Solids Struct 2010;47:1424.
[84] Purkayastha RT, McMeeking RM. An integrated 2-D model of a lithium ion battery: the effect of material parameters and
morphology on storage particle stress. Comput Mech 2012;50:209.
[85] Christensen J. Modeling diffusion-induced stress in Li-ion cells with porous electrodes. J Electrochem Soc 2010;157:A366.
[86] Cheng Y-T, Verbrugge MW. Diffusion-induced stress, interfacial charge transfer, and criteria for avoiding crack initiation
of electrode particles. J Electrochem Soc 2010;157:A508.
[87] Garcia RE, Chiang Y-M, Carter WC, Limthongkul P, Bishop CM. Microstructural modeling and design of rechargeable
lithium-ion batteries. J Electrochem Soc 2005;152:A255.
[88] Cheng Y-T, Verbrugge MW. Application of Hasselman’s crack propagation model to insertion electrodes. Electrochem
Solid State Lett 2010;13:A128.
[89] Deshpande R, Cheng Y-T, Verbrugge MW, Timmons A. Diffusion induced stresses and strain energy in a phase-
transforming spherical electrode particle. J Electrochem Soc 2011;158:A718.
A. Mukhopadhyay, B.W. Sheldon / Progress in Materials Science 63 (2014) 58–116 113

[90] Mao O, Dunlap RA, Dahn JR. Mechanically alloyed Sn–Fe(–C) powders as anode materials for Li-ion batteries; I. The Sn2Fe–
C system. J Electrochem Soc 1999;146:405.
[91] Mukaibo H, Momma T, Shacham-Diamand Y, Osaka T, Kodaira M. In situ stress transition observations of electrodeposited
Sn-based anode materials for lithium-ion secondary batteries. Electrochem Solid-State Lett 2007;10:A70.
[92] Huang JY, Zhong L, Wang CM, Sullivan JP, Xu W, Zhang LQ, et al. In situ observation of the electrochemical lithiation of a
single SnO2 nanowire electrode. Science 2010;330:1515.
[93] Wang CM, Xu W, Liu J, Zhang J-G, Saraf LV, Arey BW, et al. In situ transmission electron microscopy observation of
microstructure and phase evolution in a SnO2 nanowire during lithium intercalation. Nano Lett 2011;11:1874.
[94] Liu XH, Huang JY. In situ TEM electrochemistry of anode materials in lithium ion batteries. Energy Environ Sci
2011;4:3844.
[95] Chao S-C, Yen Y-C, Song Y-F, Chen Y-M, Wuc H-C, Wu N-L. A study on the interior microstructures of working Sn particle
electrode of Li-ion batteries by in situ X-ray transmission microscopy. Electrochem Commun 2010;12:234.
[96] Key B, Bhattacharyya R, Morcrette M, Seznéc V, Tarascon J-M, Grey CP. Real-time NMR investigations of structural
changes in silicon electrodes for lithium-ion batteries. J Am Chem Soc 2009;131:9239.
[97] Tian Y, Timmons A, Dahn JR. In situ AFM measurements of the expansion of nanostructured Sn–Co–C films reacting with
lithium. J Electrochem Soc 2009;156:A187.
[98] Wang C-M, Li X, Wang Z, Xu W, Liu J, Gao F, et al. In situ TEM investigation of congruent phase transition and structural
evolution of nanostructured silicon/carbon anode for lithium ion batteries. Nano Lett 2012;12:1624.
[99] Wang JW, Liu XH, Zhao K, Palmer A, Patten E, Burton D, et al. Sandwich-lithiation and longitudinal crack in amorphous
silicon coated on carbon nanofibers. ACS Nano 2012;6:9158.
[100] Chen WX, Lee JY, Liu Z. The nanocomposites of carbon nanotube with Sb and SnSb0.5 as Li-ion battery anodes. Carbon
2003;41:959.
[101] Li X, Meng X, Liu J, Geng D, Zhang Y, Banis MN, et al. Tin oxide with controlled morphology and crystallinity by atomic
layer deposition onto graphene nanosheets for enhanced lithium storage. Adv Funct Mater 2012;22:1647.
[102] Zhang L-S, Jiang L-Y, Yan H-J, Wang WD, Wang W, Song WG, et al. Mono dispersed SnO2 nanoparticles on both sides of
single layer graphene sheets as anode materials in Li-ion batteries. J Mater Chem 2010;20:5462.
[103] Saint J, Morcrette M, Larcher D, Laffont L, Beattie S, Pérès J-P, et al. Towards a fundamental understanding of the improved
electrochemical performance of silicon–carbon composites. Adv Funct Mater 2007;17:1765.
[104] Zhang Y, Zhang XG, Zhang HL, Zhao ZG, Li F, Liu C, et al. Composite anode material of silicon/graphite/carbon nanotubes
for Li-ion batteries. Electrochim Acta 2006;51:4994.
[105] Wang W, Kumta PN. Nanostructured hybrid silicon/carbon nanotube heterostructures: reversible high-capacity lithium-
ion anodes. ACS Nano 2010;4:2233.
[106] Cui L-F, Hu L, Choi JW, Cui Y. Light-weight free-standing carbon nanotube-silicon films for anodes of lithium ion batteries.
ACS Nano 2010;7:3671.
[107] Krishnan R, Lu T-M, Koratkar M. Functionally strain-graded nanoscoops for high power Li-ion battery anodes. Nano Lett
2011;11:377.
[108] Guoping W, Qingtang Z, Zuolong Y, MeiZheng Q. The effect of different kinds of nano-carbon conductive additives in
lithium ion batteries on the resistance and electrochemical behavior of the LiCoO2 composite cathodes. Solid State Ionics
2008;179:263.
[109] Cho J, Kim YJ, Kim T-J, Park B. Zero-strain intercalation cathode for rechargeable Li-ion cell. Angew Chem Int Ed
2001;40:3367.
[110] Chen Z, Dahn JR. Studies of LiCoO2 coated with metal oxides. Electrochem Solid-State Lett 2003;6:A221.
[111] Bouwman PJ. Lithium Intercalation in preferentially oriented submicron LiCoO2 films. PhD Thesis. University of Twente,
Enschede, Sweden; 2002.
[112] Chung KY, Ryu C-W, Kim K-B. Onset mechanism of Jahn–Teller distortion in 4 V LiMn2O4 and its suppression by
LiM0.05Mn1.95O4 (M = Co, Ni) coating. J Electrochem Soc 2005;152:A791.
[113] Pereira N, Klein LC, Amatucci GG. Particle size and multiphase effects on cycling stability using tin-based materials. Solid
State Ionics 2004;167:29.
[114] Wu H, Zheng G, Liu N, Carney TJ, Yang Y, Yi Cui. Engineering empty space between Si nanoparticles for lithium-ion battery
anodes. Nano Lett. 2012;12:904.
[115] Choi SH, Kim J, Yoon YS. A TEM study of cycled nano-crystalline HT-LiCoO2 cathodes for rechargeable lithium batteries. J
Power Sources 2004;135:286.
[116] Yu C, Li X, Ma T, Rong J, Zhang R, Shaffer J, et al. Silicon thin films as anodes for high-performance lithium-ion batteries
with effective stress relaxation. Adv Energy Mater 2012;2:68.
[117] Soni SK, Sheldon BW, Xiao X, Verbrugge MW, Ahn D, Haftbaradaran H, et al. Stress mitigation during the lithiation of
patterned amorphous Si islands. J Electrochem Soc 2012;159:A38.
[118] Xiao X, Wang L, Wang D, He X, Peng Q, Li Y. Hydrothermal synthesis of orthorhombic LiMnO2 nano-particles and LiMnO2
nanorods and comparison of their electrochemical performances. Nano Res 2009;2:923.
[119] Misra S, Liu N, Nelson J, Hong SS, Cui Y, Toney MF. In situ X-ray diffraction studies of (de)lithiation mechanism in silicon
nanowire anodes. ACS Nano 2012;6:5465.
[120] Beaulieu LY, Hatchard TD, Bonakdarpour A, Fleischauer MD, Dahn JR. Reaction of Li with alloy thin films studied by In Situ
AFM. J Electrochem Soc 2003;150:A1457.
[121] Sethuraman VA, Chon MJ, Shimshak M, Srinivasan V, Guduru PR. In situ measurements of stress evolution in silicon thin
films during electrochemical lithiation and delithiation. J Power Sources 2010;195:5062.
[122] Sethuraman VA, Srinivasan V, Bower AF, Guduru PR. In situ measurements of stress-potential coupling in lithiated silicon.
J Electrochem Soc 2010;157:A1253.
[123] Soni SK, Sheldon BW, Xiao X, Tokranov A. Thickness effects on the lithiation of amorphous silicon thin films. Scr Mater
2011;64:307.
[124] Sheldon BW, Soni SK, Xiao X, Que Y. Stress contributions to solution thermodynamics in Li–Si alloys. Electrochem Solid
State Lett 2012;15:A9.
114 A. Mukhopadhyay, B.W. Sheldon / Progress in Materials Science 63 (2014) 58–116

[125] Mukhopadhyay A, Tokranov A, Xiao X, Sheldon BW. Stress development due to surface processes in graphite electrodes
for Li-ion batteries: a first report. Electrochim Acta 2012;66:28.
[126] Sethuraman VA, Chon MJ, Shimshak M, Winkle NV, Guduru PR. In situ measurement of biaxial modulus of Si anode for Li-
ion batteries. Electrochem Commun 2010;12:1614.
[127] Chason E, Sheldon BW. Monitoring stress in thin films during processing. Surf Eng 2003;19:387.
[128] Chason E, Sheldon BW, Freund LB, Floro JA, Hearne SJ. Origin of compressive residual stress in polycrystalline thin films.
Phys Rev lett 2002;88. 156103-1.
[129] Arora P, Zhang Z. Battery separators. Chem Rev 2004;104:4419.
[130] Aurbach D, Talyosef Y, Markovsky B, Markevich E, Zinigrad E, Asraf L, et al. Design of electrolyte solutions for Li and Li-ion
batteries: a review. Electrochim Acta 2004;50:247.
[131] Zhang Y, Feng H, Wu X, Wang L, Zhang A, Xia T, et al. Progress of electrochemical capacitors electrode materials. Int J
Hydrogen Storage 2009;34:4889.
[132] Zheng T, Reimers JN, Dahn JR. Effect of turbostratic disorder in graphitic carbon hosts on the intercalation of lithium. Phys
Rev B 1995;51:734.
[133] Qi Y, Guo H, Hector LG, Timmons A. Threefold increase in the Young’s modulus of graphite negative electrode during
lithium intercalation. J Electrochem Soc 2010;157:A558.
[134] Sethuraman VA, Van Winkle N, Abraham DP, Bower AF, Guduru PR. Real-time stress measurements in lithium-ion battery
negative-electrodes. J Power Sources 2012;206:334.
[135] Kanamura K, Naito H, Yao T, Takehara Z-I. Structural change of the LiMn2O4 spinel structure induced by extraction of
lithium. J Mater Chem 1996;6:33.
[136] He Y, Yu X, Li G, Wang R, Li H, Wang Y, et al. Shape evolution of patterned amorphous and polycrystalline silicon
microarray thin film electrodes caused by lithium insertion and extraction. J Power Sources 2012;216:131.
[137] Inaba M, Iriyama Y, Ogumi Z, Todzuka Y, Tasaka A. Raman study of layered rock-salt and LiCoO2 and its electrochemical
lithium deintercalation. J Raman Spectrosc 1997;28:613.
[138] Amatucci GG, Tarascon JM, Klein LC. CoO2, the end member of the LiXCoO2 solid solution. J Electrochem Soc
1996;143:114.
[139] Boukamp BA, Lesh GC, Huggins RA. All-solid lithium electrodes with mixed-conductor matrix. J Electrochem Soc
1981;128:725.
[140] VanDerMarel C, Vinke GJB, VanDerLugt W. The phase diagram of the system lithium–silicon. Solid State Commnun
1985;54:917.
[141] Shenoy VB, Johari P, Qi Y. Elastic softening of amorphous and crystalline Li–Si phases with increasing Li concentration: a
first-principles study. J Power Sources 2010;195:6825.
[142] McAlister AJ. The Al–Li (aluminum–lithium) system. Bull Alloy Phase Diagram 1982;3:177.
[143] Chen G, Song X, Richardson TJ. Electron microscopy study of the LiFePO4 to FePO4 phase transition. Electrochem Solid-
State Lett 2006;9:A295.
[144] Woo KC, Kamitakahara WA, DiVincenzo DP, Robinson DS, Mertwoy H, Milliken JW, et al. Effect of in-plane density on the
structural and elastic properties of graphite intercalation compounds. Phys Rev Lett 1983;50:182.
[145] Hamon Y, Brousse T, Jousse F, Topart P, Buvat P, Schleich DM. Aluminum negative electrode in lithium ion batteries. J
Power Sources 2001;97–98:185.
[146] Reimers JN, Dahn JR. Electrochemical and in situ X-ray diffraction studies of lithium intercalation in LixCoO2. J
Electrochem Soc 1992;139:2091.
[147] Gabrisch H, Yazami R, Fultz B. The character of dislocations in LiCoO2. Electrochem Solid-State Lett 2002;5:A111.
[148] Besenhard JO, Winter M, Yang J, Biberacher W. Filming mechanism of lithium-carbon anodes in organic and inorganic
electrolytes. J Power Sources 1995;54:228.
[149] Aurbach D, Koltypin M, Teller H. In situ AFM imaging of surface phenomena on composite graphite electrodes during
lithium insertion. Langmuir 2002;18:9000.
[150] Jeong S-K, Inaba M, Abe T, Ogumi Z. Surface film formation on graphite negative electrode in lithium-ion batteries AFM
study in an ethylene carbonate-based solution. J Electrochem Soc 2001;148:A989.
[151] Rosolen JM, Decker F. Stress in carbon film electrodes during Li+ electrochemical intercalation. J Electrochem Soc
1996;143:2417.
[152] Buqa H, Wursig AW, Vetter J, Spahr ME, Krumeich F, Novak P. SEI film formation on highly crystalline graphitic materials
in lithium-ion batteries. J Power Sources 2006;153:385.
[153] Campana FP, Kotz R, Vetter J, Novak P, Siegenthaler H. In situ atomic force microscopy study of dimensional changes
during Li+ ion intercalation/de-intercalation in highly oriented pyrolytic graphite. Electrochem Commun 2005;7:107.
[154] Tokranov A, Sheldon BW, Lu P, Xiao X, Mukhopadhyay A. The origin of stress in the solid electrolyte interphase on carbon
electrodes for Li-ion batteries. J Electrochem Soc 2014;161:A58.
[155] Chason E, Sheldon BW, Freund LB, Floro JA, Hearne SJ. Origin of compressive residual stress in polycrystalline thin films.
Phy Rev Lett 2002;8. 156103-1.
[156] Qi Y, Sheldon BW, Guo H, Xiao X, Kothari AK. Impact of surface chemistry on grain boundary induced intrinsic stress
evolution during polycrystalline thin film growth. Phy Rev Lett 2009;102. 056101-1.
[157] Zane D, Antonini A, Pasquali M. A morphological study of SEI film on graphite electrodes. J Power Sources 2001;97–
98:146.
[158] Liu W-R, Yang M-H, Wu H-C, Chiao SM, Wu N-L. Enhanced cycle life of si anode for Li-ion batteries by using modified
elastomeric binder. Electrochem Solid-State Lett 2005;8:A100.
[159] Shiva K, Asokan S, Bhattacharyya AJ. Improved lithium cyclability and storage in a multi-sized pore (‘‘differential
spacers’’) mesoporous SnO2. Nanoscale 2011;3:1501.
[160] Beaulieu LY, Eberman KW, Turner RL, Krause LJ, Dahn JR. Colossal reversible volume changes in lithium alloys.
Electrochem Solid-State Lett 2001;4:A137.
[161] Graetz J, Ahn CC, Yazami R, Fultz B. Highly reversible lithium storage in nanostructured silicon. Electrochem Solid-State
Lett 2003;6:A194.
A. Mukhopadhyay, B.W. Sheldon / Progress in Materials Science 63 (2014) 58–116 115

[162] Jung H, Park M, Yoon Y-G, Kim G-B, Joo S-K. Amorphous silicon anode for lithium-ion rechargeable batteries. J Power
Sources 2003;115:346.
[163] Guo ZP, Zhao ZW, Liu HK, Dou SX. Electrochemical lithiation and de-lithiation of MWNT–Sn/SnNi nanocomposites.
Carbon 2005;43:1392.
[164] Zhao K, Tritsaris GA, Pharr M, Wang WL, Okeke O, Suo Z, et al. Reactive flow in silicon electrodes assisted by the insertion
of lithium. Nano Lett 2012;12:4397.
[165] Verbrugge M, Cheng Y-T. Automotive traction battery needs and the influence of mechanical degradation of insertion
electrode particles. ECS Trans 2009;19:1.
[166] Wachtler M, Besenhard JO, Winter M. Tin and tin-based intermetallics as new anode materials for lithium-ion cells. J
Power Sources 2001;94:189.
[167] Inaba M, Uno T, Tasaka A. Irreversible capacity of electrodeposited Sn thin film anode. J Power Sources 2005;146:473.
[168] Larche FC, Cahn JW. Overview No. 41. The interactions of composition and stress in crystalline solids. Acta Metall
1985;33:331.
[169] Chevrier VL, Dahn JR. First principles model of amorphous silicon lithiation batteries and energy storage. J Electrochem
Soc 2009;156:A454.
[170] Liu D, Wang Y, Xie Y, He L, Chen J, Wu K, et al. On the stress characteristics of graphite anode in commercial pouch
lithium-ion battery. J Power Sources 2013;232:29.
[171] Liu XH, Fan F, Yang H, Zhang S, Huang JY, Zhu T. Self-limiting lithiation in silicon nanowires. ACS Nano 2013;7:1495.
[172] Lee SW, McDowell MT, Berla LA, Nix WD, Cui Y. Fracture of crystalline silicon nanopillars during electrochemical lithium
insertion. PNAS 2012;109:4080.
[173] Pol VG, Thackeray MM. Spherical carbon particles and carbon nanotubes prepared by autogenic reactions: evaluation as
anodes in lithium electrochemical cells. Energy Environ Sci 2011;4:1904.
[174] DeLasCasas C, Li W. A review of application of carbon nanotubes for lithium ion battery anode material. J Power Sources
2012;208:74.
[175] Wen CY. Non-catalytic heterogeneous solid-fluid reaction models. Ind Eng Chem 1968;60:34.
[176] Stoney GG. The tension of metallic films deposited by electrolysis. Proc R Soc (Lond) A 1909;82:172.
[177] Pyun Su-I, Go J-Y, Shin H-C. Electrogravimetric study of the stress developed during lithium transport through the RF
sputtered Li1dCoO2 film electrode. J New Mater Electrochem Syst 2002;5:143.
[178] Chung KY, Kim K-B. Investigation of structural fatigue in spinel electrodes using in situ laser probe deflection technique. J
Electrochem Soc 2002;149:A79.
[179] Kepler KD, Vaughney JT, Thackeray MM. Copper–tin anodes for rechargeable lithium batteries: an example of the matrix
effect in an intermetallic system. J Power Sources 1999;81–82:383.
[180] Reddy MA, Varadaraju UV. NbSb2 as an anode material for Li-ion batteries. J Power Sources 2006;159:336.
[181] Hassoun J, Panero S, Mulas G, Scrosati B. An electrochemical investigation of a Sn–Co–C ternary alloy as a negative
electrode in Li-ion batteries. J Power Sources 2007;171:928.
[182] Li T, Cao YL, Ai XP, Yang HX. Cycleable graphite/FeSi6 alloy composite as a high capacity anode material for Li-ion
batteries. J Power Sources 2008;184:473.
[183] Mao O, Turner RL, Courtney IA, Fredericksen BD, Buckett MI, Krause LJ, et al. Active/inactive nanocomposites as anodes for
Li-ion batteries. Electrochem Solid-State Lett 1999;2:3.
[184] Fu ZW, Wang Y, Yue X-L, Zhao S-L, Qin Q-Z. Electrochemical reactions of lithium with transition metal nitride electrodes. J
Phys Chem B 2004;108:2236.
[185] Benedek R, Thackeray MM. Lithium reactions with intermetallic-compound electrodes. J Power Sources 2002;110:406.
[186] Kepler KD, Vaughey JT, Thackeray MM. LixCu6Sn5 (0 < x < 13): an intermetallic insertion electrode for rechargeable
lithium batteries. Electrochem Solid-State Lett 1999;2:307.
[187] Alcántara R, Fernández-Madrigal FJ, Lavela P, Tirado JL, Jumas JC, Olivier-Fourcade J. Electrochemical reaction of lithium
with the CoSb3 skutterudite. J Mater Chem 1999;9:2517.
[188] Mao O, Dahn J. Mechanically alloyed Sn–Fe(–C) powders as anode materials for Li-ion batteries II. The Sn–Fe system. J
Electrochem Soc 1999;146:414.
[189] Yang J, Takeda Y, Imanishi N, Yamamoto O. Ultrafine Sn and SnSb0.14 powders for lithium storage matrices in lithium-ion
batteries. J Electrochem Soc 1999;146:4009.
[190] Stjerndahl M, Bryngelsson H, Gustafsson T, Vaughey JT, Thackeray MM, Edstrom K. Surface chemistry of intermetallic
AlSb-anodes for Li-ion batteries. Electrochim Acta 2007;52:4947.
[191] Wang X, Wen Z, Lin B, Lin J, Wu X, Xu X. Preparation and electrochemical characterization of tin/graphite/silver composite
as anode materials for lithium-ion batteries. J Power Sources 2008;184:508.
[192] Shiva K, Rajendra HB, Subrahmanyam KS, Bhattacharyya AJ, Rao CNR. Improved lithium cyclability and storage in
mesoporous SnO2 electronically wired with very low concentrations (61%) of reduced graphene oxide. Chem Eur J
2012;18:4489.
[193] Idota Y, Kubota T, Matsufuji A, Maekawa Y, Miyasaka T. Tin-based amorphous oxide: a high-capacity lithium-ion–storage
material. Science 1997;276:1395.
[194] Idota Y, Nishima M, Miyaki Y, Kubota T, Miyasaki T. Can Pat Appl 1994;2(1134):053.
[195] Courtney TA, Dahn JR. Electrochemical and in situ X-ray diffraction studies of the reaction of lithium with tin oxide
composites. J Electrochem Soc 1997;144:2045.
[196] Magasinski A, Dixon P, Hertzberg B, Kvit A, Ayala J, Yushin G. High-performance lithium-ion anodes using a hierarchical
bottom-up approach. Nature Mater 2010;9:353.
[197] Khomenko VG, Barsukov VZ. Characterization of silicon- and carbon-based composite anodes for lithium-ion batteries.
Electrochim Acta 2007;52:2829.
[198] Yoshio M, Wang HY, Fukuda K, Umeno T, Dimov N, Ogumi ZE. Carbon-coated Si as a lithium-ion battery anode material. J
Electrochem Soc 2002;149:A1598.
[199] Shu J, Li H, Yang R, Shi Y, Huang X. Cage-like carbon nanotubes/Si composite as anode material for lithium ion batteries.
Electrochem Commun 2006;8:51.
116 A. Mukhopadhyay, B.W. Sheldon / Progress in Materials Science 63 (2014) 58–116

[200] Wang W, Epur R, Kumta PN. Vertically aligned silicon/carbon nanotube (VASCNT) arrays: hierarchical anodes for lithium-
ion battery. Electrochem Commun 2011;13:429.
[201] Hu L, Wu H, Gao Y, Cao A, Li H, McDough J, et al. Silicon–carbon nanotube coaxial sponge as Li-ion anodes with high areal
capacity. Adv Energy Mater 2011;1:523.
[202] Chen Z, Christensen L, Dahn JR. Comparison of PVDF and PVDF-TFE-P as binders for electrode materials showing large
volume changes in lithium-ion batteries. J Electrochem Soc 2003;150:A1073.
[203] Hariharan S, Saravanan K, Ramar V, Balaya P. A rationally designed dual role anode material for lithium-ion and sodium-
ion batteries: case study of eco-friendly Fe3O4. Roy Soc Chem 2013;15:2954.
[204] Chen Z, Dahn JR. Methods to obtain excellent capacity retention in LiCoO2 cycled to 4.5 V. Electrochim Acta
2004;49:1079.
[205] Balaya P, Saravan K, Hariharan S, Ramar V, Lee HS, Kuezma M, et al. Nanostructured mesoporous materials for lithium-ion
battery applications. In: Energy harvesting and storage: materials, devices and applications II, book series: proceedings of
SPIE, vol. 8035, article no.: 803503; 2011.
[206] Balaya P, Maier J. Nanostructured materials for reversible lithium batteries. Innovation 2007;7:32.
[207] Teki R, Datta MK, Krishnan R, Parker TC, Lu T-M, Kumta PN, et al. Nanostructured silicon anodes for lithium ion
rechargeable batteries. Small 2009;5:2236.
[208] Gleiter H. Nanostructured materials: basic concepts and microstructure. Acta Metal 2000;48:1.
[209] Mukhopadhyay A, Basu B. Consolidation-microstructure-properties of bulk nanoceramics and ceramic nanocomposites: a
review. Intl Mater Rev 2007;52:257.
[210] Liu XH, Zhong L, Huang S, Mao SX, Zhu T, Huang JY. Size-dependent fracture of silicon nanoparticles during lithiation. ACS
Nano 2012;6:1522.
[211] Okubo M, Hosono E, Kim J, Enomoto M, Kojima N, Kudo T, et al. Nanosize effect on high-rate Li-ion intercalation in LiCoO2
electrode. J Am Chem Soc 2007;129:7444.
[212] Hsu K, Tsay S, Hwang B. Synthesis and characterization of nano-sized LiFePO4 cathode materials prepared by a citric acid
based sol-gel route. J Mater Chem 2004;14:2690.
[213] Chen LB, Xie JY, Yu HC, Wang TH. An amorphous Si thin film anode with high capacity and long cycling life for lithium ion
batteries. J Appl Electrochem 2009;39:1157.
[214] Chan CK, Patel RN, O’Connell MJ, Korgel BA, Cui Y. Solution-grown silicon nanowires for lithium-ion battery anodes. ACS
Nano 2010;4:1443.
[215] Jung H, Park M, Han SH, Lim H, Joo S-K. Amorphous silicon thin-film negative electrode prepared by low pressure
chemical vapor deposition for lithium-ion batteries. Solid State Commun 2003;125:387.
[216] Armstrong AR, Armstrong G, Canales J, Garcia R, Bruce PG. Lithium-ion intercalation into TiO2-B nanowires. Adv Mater
2005;17:862.
[217] Li N, Martin CR, Scrosati B. Nanomaterial-based Li-ion battery electrodes. J Power Sources 2001;97–98:240.
[218] Kim DK, Muralidharan P, Lee H-W, Ruffo R, Yang Y, Chan CK, et al. Spinel LiMn2O4 nanorods as lithium ion battery
cathodes. Nano Lett 2008;8:3948.
[219] Saji VS, Kim Y-S, Kim T-H, Cho J, Song H-K. One-dimensional (1D) nanostructured and nanocomposited LiFePO4: its
perspective advantages for cathode materials of lithium ion batteries. Phys Chem Chem Phys 2011;13:19226.
[220] Shiva K, Asokan S, Bhattacharyya AJ. Improved lithium cyclability and storage in a multi-sized pore (‘‘differential
spacers’’) mesoporous SnO2. Nanoscale 2011;3:1501–3.
[221] Shiva K, Rajendra HB, Subrahmanyam KS, Bhattacharyya AJ, Rao CNR. Improved lithium cyclability and storage in
mesoporous SnO2 electronically wired with very low concentrations (61%) of reduced graphene oxide. Chem Eur J
2012;18:4489.
[222] Iwamura S, Nishihara H, Kyotani T. Fast and reversible lithium storage in a wrinkled structure formed from Si
nanoparticles during lithiation/delithiation cycling. J Power Sources 2013;222:400.
[223] Aravindan V, Jinesh KB, Prabhakar RR, Kale VS, Madhav S. Atomic layer deposited (ALD) SnO2 anodes with exceptional
cycleability for Li-ion batteries. Nano Energy 2013;2:720.

You might also like