Tải bản đầy đủ (.pdf) (16 trang)

Polysaccharides for sustainable energy storage – A review

Bạn đang xem bản rút gọn của tài liệu. Xem và tải ngay bản đầy đủ của tài liệu tại đây (8.99 MB, 16 trang )

Carbohydrate Polymers 265 (2021) 118063

Contents lists available at ScienceDirect

Carbohydrate Polymers
journal homepage: www.elsevier.com/locate/carbpol

Review

Polysaccharides for sustainable energy storage – A review
Werner Schlemmer a, Julian Selinger a, b, Mathias Andreas Hobisch a, Stefan Spirk a, *
a
b

Institute of Bioproducts and Paper Technology, Graz University of Technology, 8010, Graz, Austria
Department of Bioproducts and Biosystems, Aalto University, P. O. Box 16300, 00076, Aalto, Finland

A R T I C L E I N F O

A B S T R A C T

Keywords:
Battery
Polysaccharides
Cellulose
Alginate
Chitosan
Nanocellulose
Binder
Separator


The increasing amount of electric vehicles on our streets as well as the need to store surplus energy from
renewable sources such as wind, solar and tidal parks, has brought small and large scale batteries into the focus
of academic and industrial research. While there has been huge progress in performance and cost reduction in the
past years, batteries and their components still face several environmental issues including safety, toxicity,
recycling and sustainability. In this review, we address these challenges by showcasing the potential of
polysaccharide-based compounds and materials used in batteries. This particularly involves their use as electrode
binders, separators and gel/solid polymer electrolytes. The review contains a historical section on the different
battery technologies, considerations about safety on batteries and requirements of polysaccharide components to
be used in different types of battery technologies. The last sections cover opportunities for polysaccharides as
well as obstacles that prevent their wider use in battery industry.

1. Scope of the review
This review aims at summarizing the use of polysaccharides in en­
ergy storage systems. Central to this review is to focus on energy storage
elements, i.e., active material, separator, binders. The intention of the
review is not to list all types of materials but to focus on requirements of
the respective energy storage component and why polysaccharides can
be versatile candidates in the development of such components. We also
discuss limitations of polysaccharides in this area as well as obstacles
that prevent them from wider use in energy storage systems. In the end
of the review, the challenges and opportunities for polysaccharides in
battery systems will be highlighted and discussed.
2. Introduction
We are facing a global crisis as the use of fossil fuels has been
emitting huge quantities of greenhouse gases such as CO2 and methane

to the atmosphere. The increasing concentration of these compounds
into the atmosphere led to global warming at unprecedented rate in the
history of mankind. During the lifetime of Carbohydrate Polymers
(September 1981 until now), the CO2 content in the atmosphere

increased from 340 to 419 ppm (NOAA, 2021). As a consequence, policy
makers in Europe and in other parts of the world aim at limiting these
emissions to reduce global warming (European-Commission, 2020a,
2020b). A core element of this policy is to change the energy supply from
fossil based towards green sources such as solar, wind and tidal energy.
Some countries have made remarkable efforts in this respect, e.g., Ger­
many shut down a high number of coal fueled plants and massively
invested in wind and solar energy, which accounted for 42.1 % of overall
energy production in 2019 (Fraunhofer-ISE, 2021). Similarly, the Nordic
countries have been aiming at converting their energy supply to more
sustainable sources, thereby curbing CO2 emissions. A major challenge
of renewable energy, however, originates from fluctuations in energy
production, caused by weather conditions and seasonal changes. In

Abbreviations: AA, agar-agar; ANF, aramid nanofiber; BC, bacterial cellulose; CA, cellulose acetate; CaAlg, calcium alginate; CG, carrageenan; ChNF, chitosan
nanofibrils; CMC, carboxymethyl cellulose; CMCh, carboxymethyl chitosan; CN-CMCh, cyanoethyl carboxymethyl chitosan; CNF, cellulose nanofibrils; CNC, cellulose
nanocrystals; EC, ethyl cellulose; GA, gum arabic; GM, galactomannan; GG, guar gum; GPE, gel polymer electrolyte; HEC, hydroxyethyl cellulose; HNT, halloysite
nanotube; HPC, hydroxypropyl cellulose; LIB, lithium ion battery; LPC, lignosulfonate–polyamide-epichlorohydrin complex; LTO, lithium titanate, Li4Ti5O12; LNMO,
lithium nickel manganese oxide; m-CA, modified cellulose acetate; NaAlg, sodium alginate; PBI, poly(oxyphenylene benzimidazole); PE, polyethylene; PET, poly­
ethylene terephthalate; PMMA, poly(methyl methacrylate); PP, polypropylene; PPS, polypropylene sulfide; PPy, polypyrrol; PSA, polysulfonamide; PSS, polystyrene
sulfonate; PVDF, polyvinylidene fluoride; PVDF-HFP, polyvinylidene fluoride-co-hexafluoropropylene; SBR, styrene-butadiene rubber; SPE, solid polymer electro­
lytes; TG, tara gum; TEMPO, (2,2,6,6-tetramethylpiperidin-1-yl)oxyl; TOCN, TEMPO oxidized cellulose nanofibrils; XG, xanthan gum.
* Corresponding author.
E-mail address: (S. Spirk).
/>Received 11 March 2021; Received in revised form 7 April 2021; Accepted 8 April 2021
Available online 20 April 2021
0144-8617/© 2021 The Author(s). Published by Elsevier Ltd. This is an open access article under the CC BY license ( />

W. Schlemmer et al.


Carbohydrate Polymers 265 (2021) 118063

addition, the energy is often produced at locations which are not in close
spatial proximity to the end users. Moreover, energy is produced at times
when the supply exceeds the demand. These challenges lead to signifi­
cant instabilities in the grid, requiring back-up plants to ensure the
power supply. The most promising solution to this problem lies in the
installation of large-scale storage facilities, which can release the energy
when it is needed. Different approaches do exist that rely on storage in
the form of mechanical (e.g., flywheels), potential (hydropower, com­
pressed air storage) or chemical energy (e.g. batteries, hydrogen). All of
these technologies enable storage capacity ranging from two-digit MWh
(flywheels) to GWh (hydropower) (Mongird et al., 2019). Lithium ion
batteries (LIBs) have been proven to be an integral part of stationary
storage with the largest battery being the Hornsdale Facility in Australia
(100 MW h). However, there are several challenges with battery com­
ponents. They mostly rely on depletable sources, their recycling is
challenging and resources are not regionally available, causing a high
CO2 footprint. These issues have to be seen in the context of a steadily
increasing share of electric vehicles, boosting the demand particularly
for LIBs, accompanied with a rapid decline in costs which will probably
soon cross the 100 $/kWh threshold.
In this review, the emphasis is put on energy storage components
based on polysaccharides, comprising separators, electrolytes, and
binders. We highlight the specific advantages which polysaccharides can
offer for each application. Only batteries are covered, while super­
capacitors as well as electrode materials from biomass and lignin are not
investigated in this review; here we refer to other reviews for further
reading (Liedel, 2020; Nirmale, Kale, & Varma, 2017; Yu, Li, Chen, Wu,
& Peng, 2019). The principle of batteries in combination with a his­

torical evolution of technology will be explained in brief in the following
section.

dioxide (cathode) and uses sulfuric acid as electrolyte. The acid reacts
with both electrodes to form lead sulfate causing an electron flow from
one electrode to the other, which are both reversible processes, enabling
recharging of the battery. Interestingly, the design of the lead-acid
battery has not significantly changed since its invention. In some ap­
plications, such as in car batteries, they are still in use.
The dry cell technology was the next step towards new application
areas of batteries. The basic concept is a paste-like electrolyte, that
contains a small amount of moisture to allow for current flow. As the
electrolyte is a paste, these cells can be oriented in any direction as there
is nothing that could potentially spill off. Portable applications were
realized, such as the zinc carbon battery (1.5 V). It is made of a zinc
casing that serves at the same time as anode and a manganese(IV) oxide
cathode which is connected to a conductive carbon rod. Commonly used
gel electrolytes in today’s dry cells contain a moist paste of ammonium
or zinc chloride impregnated paper. The paper serves as separator be­
tween the zinc anode and the manganese(IV) oxide cathode. This type of
cell has a significant market share (roughly 20 %) for portable batteries.
An even larger share of portable batteries is occupied by dry alkaline
batteries. In such batteries, zinc and carbon are used with potassium
hydroxide as gel electrolyte and a separator, often made from cellulose.
The cell features a voltage of 1.4 V and is one of the most commonly used
primary batteries with current market shares of roughly 50 %.
In the past two decades, battery development was boosted by several
factors. Portable electronics require high energy density while being
rechargeable. In addition, battery technology was boosted by massive
investments in electric cars.

Lithium ion batteries are capable of delivering voltages of over 3 V
while having a high energy density. Developed by G.N. Lewis more than
100 years ago, it took several decades until LIBs became commercially
exploitable, with significant efforts of Besenhard and later Goodenough
(Besenhard & Eichinger, 1976; Eichinger & Besenhard, 1976; Mizush­
ima, Jones, Wiseman, & Goodenough, 1980). The first commercialized
LIB was brought to the market by Sony in 1991. The main advantage of
LIBs is that they operate at high voltages, requiring the use of electro­
lytes other than water, typically ethylene carbonate, propylene car­
bonate or diethyl carbonate. In such electrolytes, the lithium species are
dissolved (e.g. LiPF6, LiBF4). Commercial electrode materials consist of
layered oxides, spinels, polyanions (anode) and graphite (cathode).
During operation of the battery the lithium ions move to the other
electrode, where they intercalate into the electrode material (Fig. 1) (Lu,
Han, Li, Hua, & Ouyang, 2013). While LIBs have been used to power

3. Batteries-a short historical survey
The first battery was developed in the late 18th century when Luigi
Galvani observed a phenomenon he later termed ‘animal electricity’.
During the dissection of frog legs he realized that they twitched when
the iron scalpel touched them (Galvani & Volta, 1791). However, his
friend Alessandro Volta connected these observations to the metal sur­
faces of the scalpel rather than the frog legs and led the foundation for
the invention of the voltaic pile-the first battery. In such a battery, piled
copper and zinc disks were assembled. In order to avoid a short circuit,
the metals were separated by cloth or cardboard, which were presoaked
in an electrolyte solution to ensure conductivity (Dibner, 1964). The
battery was fully functioning (stable supply of electricity, current,
hardly any self-discharge) but had some disadvantages (electrolyte
leakages, short battery lifetime) caused by the weight of the piles and

parasitic reactions leading to hydrogen evolution during operation.
These issues were solved by the development of the Daniell cell, which
had an operating voltage of 1.1 V. This type of cell is based on a zinc
anode, and a copper vessel, which was filled with a solution of copper
sulfate. Into the vessel, a porous, ceramic tray filled with sulfuric acid
was placed which enabled an exchange of ions. Therefore hydrogen
evolution was suppressed and over time only conducting copper metal is
deposited on the anode (Spencer, Bodner, & Rickard, 2010). The Daniell
cell was a major accomplishment in battery technology and its impor­
tance can be seen in the unit of the electromotive force (Volt), since the
operating voltage of the Daniell cell was roughly 1 V (Hamer, 1965).
Later on, different improvements such as the Bird cell (Bird, 1838),
Gravity cell and the Poggendorf (Ayrton & Mather, 1911) cell were
introduced, which we will not cover in more detail here. It is noteworthy
that the Gravity cell was widely used in British and American telegraph
networks until the mid 1950s.
The next milestone in battery technology was the development of
rechargeable batteries. Until then, battery lifetime was limited by the
amount of redox active compounds in the battery. The lead-acid battery
was a game changer in this respect. It consists of lead (anode) and lead

Fig. 1. Schematics of a lithium ion battery using LiCoO2 and graphite as
electrode materials.
Reproduced from (Roy & Srivastava, 2015) with permission from The Royal
Society of Chemistry.
2


W. Schlemmer et al.


Carbohydrate Polymers 265 (2021) 118063

consumer electronics for the past 20 years, the increasing number of
electrically powered vehicles is driving innovation in this area. How­
ever, the past years showed that the current LIB technology features
environmental drawbacks ranging from non-sustainable raw material
supply (e.g., mining of lithium and cobalt) to challenging recycling.
Alternative, evolving battery technologies involve metals such as so­
dium and magnesium which are highly abundant and available in many
places of the world. Their operation is also safer than corresponding
lithium ion batteries.

2012). This practice would be acceptable if all experimental details are
made available in scientific papers, however, precise details about the
cell geometry are often not provided in publications. This problem has
been recognized for both supercapacitors and batteries (Freunberger,
2015; Gogotsi & Simon, 2011; Obrovac & Chevrier, 2014) but best
practice publishing standards as recently proposed for solar cells are yet
to be defined (Editorial, 2015). A material may show exceptional per­
formance when values are referenced to a single component, although
the values might be biased. In this respect, particularly the properties of
nanomaterials such as graphene can be misleading, having a very low
packing density, that allows for high electrolyte loading (which typically
does have a Faradaic contribution). For a better comparability of device
performance, for different fields of applications, energy and power
density can be compared in a Ragone plot (Fig. 2).

4. Requirements for battery components
Battery components must ensure a safe, economic and performance
driven operation of the device. Lately, the recycling, particularly of LIBs,

has also become an issue which again is connected to the steady increase
of electric vehicles on our streets. Safety aspects are manifold and cover
the cell chemistry, electrolytes, cell design, and separators. Good articles
covering these basic safety aspects are provided by Abada et al. (2016)
and Rezvanizaniani, Liu, Chen, and Lee (2014). Lately, car batteries that
caught fire either after accidents or during charging sensitized the main
public and policy makers in this respect as well.
The cell chemistry determines the energy density of the battery.
Mobile devices require a high energy density while stationary use works
as well with systems having a lower energy density. High energy density
is related to higher chemical reactivity which increases the risk of un­
desired reactions in a battery. Therefore, in commercial applications, the
cell chemistry is usually optimized to ensure a safe operation rather than
pushing the performance limits. A prominent example is (high energy
density) lithium cobalt oxide, which is subsequently replaced by other
less reactive lithium species such as lithium iron phosphates to over­
come safety issues. Electrolyte stability is another aspect that must be
considered, as its decomposition/degradation is more likely to take
place at high voltages and elevated temperatures. In addition, heat
transport must be included in the design of the cells to avoid local
hotspots in high energy density devices. Among other issues, these
hotspots can lead to separator failure as most of separators are made
from polymeric materials. When the separator melts, the cell undergoes
a thermal runaway associated with severe fires. This can be mitigated by
the use of thermally stable separators. A review on this aspect has been
recently published (Yuan & Liu, 2020).
Second only to safety, performance (e.g., energy density per volume/
mass, shelf-life, self-discharge, cycle life, device level) is of utmost
importance for batteries (Nitta, Wu, Lee, & Yushin, 2015). Here, the cell
chemistry determines the energy density and nominal voltage. Another

important aspect is shelf life, which however, again depends on battery
chemistry, ranging from 2 to 3 years for Leclanche cells to approx. 5
years for alkaline batteries and up to ten years for LIBs. Moreover,
self-discharge, i.e. the loss of battery capacity when not operated should
be minimized. The typical self-discharge rate for LIB is 2–3 % per month,
while lead acid batteries feature about 5 % loss within the same time. In
this context, temperature is an important factor, as elevated tempera­
tures typically accelerate undesired reactions, and promote
self-discharge. In addition, discharge rates, cycle life (number of char­
ge/discharge cycles until the capacity drops by 20 %) should be
optimized.
The surface area of the electrodes is a crucial parameter, as the
battery capacity, energy and power can be expressed as normalized
values by weight or volume. Often, such values are provided based on
the mass of the electron conductor. Although this is technically correct,
it can be easily misleading as the true performance may be very different
at the device level. The correlation of performance metrics of electro­
chemical energy storage devices to the mass or volume of a certain
“active” component has been become common for energy storage sys­
tems. Often, the reported electrochemical performance parameters may
represent just a part or even a negligible fraction of the total device mass
or volume (Bruce, Freunberger, Hardwick, & Tarascon, 2012; Choi et al.,

5. Polysaccharide based components in batteries
In the historical section, we already briefly described that poly­
saccharides have been crucial parts in batteries already from the very
beginning of battery technology development. However, research
extended from a mere, electrically insulating barrier (separator) in the
form of cloth or cardboard to different applications. In the following, we
will give a brief overview thereof.

5.1. Binder
The function of the binder is to prevent electrode swelling and me­
chanical damages as well as to protect the active material against the
electrolyte, while enabling ion transport throughout the binder.
Particularly for high energy density devices such as LIBs, this of crucial
importance so it is little surprising that significant efforts have been
made to improve binder performance. Here, mostly nanosized active
electrode materials have been in the focus as any degradation will lead
to drastic performance losses. The challenge is that nanoparticles (e.g.
Si) undergo volume changes upon lithiation/delithiation in the range of
300 % and more. As a consequence, the binder, which covers the
nanoparticles, is exposed to mechanical stress (Fig. 3). Cracking or
stripping of the binder leads to increasing contact area of silicon and the
electrolyte, causing increasing solid electrolyte interface (SEI) formation
and electrode gravelling (i.e. pulverization). Consequently, the nonconductive surface area increases which leads to a strong reduction in

Fig. 2. Ragone plot illustrating the performances of specific power vs specific
energy for different electrical energy-storage technologies. Times shown in the
plot are the discharge time, obtained by dividing the energy density by the
power density.
Reprinted with permission from (Shao et al., 2018). Copyright (2018) American
Chemical Society.
3


W. Schlemmer et al.

Carbohydrate Polymers 265 (2021) 118063

Fig. 3. Schematic of the volume expansion of graphite (MCMB, mesocarbon microbeads), silicon, and silicon–MCMB graphite composite electrodes.

Reproduced from (Yim, Courtel, & Abu-Lebdeh, 2013) with permission from The Royal Society of Chemistry.

performance, with battery failure being the final stage.
Polyvinylidene fluoride (PVDF) is often used in commercial battery
designs as it can be assembled in both anode and cathode materials.
Advantages of PVDF include high (electro-)chemical stability, and good
adhesion to electrode materials and current collectors. While these
properties are important, rather poor mechanical flexibility of PVDF is
unfavorable for a use as binder.
The advantage of many polysaccharides is that they readily form
homogeneous films and layers onto different materials. Moreover, many
polysaccharides can be processed from aqueous solutions. While envi­
ronmentally friendly, this also poses a risk during the drying as
shrinkage may induce cracks in the binder or reduction of the interfacial
contact areas by stripping. However, the use of organic solvents during
battery assembly can be avoided in case of aqueous polysaccharide
systems, which is a benefit for the production. A review on biopolymeric
binders gives more details on these points (Bresser, Buchholz, Moretti,
Varzi, & Passerini, 2018).
In the past decades, polysaccharides have been proposed in binder
formulation with carboxymethyl cellulose (CMC) as the most prominent
example, which is even in commercial use. Although one may think that
CMC may form a ‘soft’ elastic layer on electrode materials to account for
volume changes, this is actually not the case. As known from paper­
making, where CMC can be used to improve paper strength, it is a rigid
molecule. Typically, CMC films do not exceed 5–8 % elongation at break
(Lestriez, Bahri, Sandu, Rou´e, & Guyomard, 2007). Consequently, the
rigidity of the binder is only one important aspect in their performance.
This has been demonstrated by Li et al. who compared pure CMC and
styrene-butadiene-rubber (SBR-CMC) blends as binders. One would

argue that the more elastic SBR–CMC system should yield more stable
performance, which was however not the case. The stiffer CMC binder
showed superior characteristics regarding all relevant electrochemical
parameters – despite being more brittle (Li, Lewis, & Dahn, 2007). One
reason was that CMC improves the processability of the electrode

slurries as the encapsulation of silicon nanoparticles by CMC facilitates
their dispersibility. As many polysaccharides with carboxyl groups, the
CMC can act as crosslinker in slurries, creating three-dimensional net­
works between particles and polymer suspensions. Thereby, the inter­
action relies on the entanglements of the CMC macromolecules with the
particles, which is further governed by its conformation (coiled vs
extended), the degree of polymerization and the degree of substitution
with carboxymethyl groups (Fig. 4). It is known that high molar mass
and a high extent of coiling favors crosslinking, which is beneficial for
binders. CMC based binders are capable of compensating for up to 400 %
in volume change while maintaining stable performance (Lestriez et al.,
2007).
However, the mode of interaction of CMC with the electrode surface
is not merely physical. This has been extensively studied for silicon
nanoparticle-based electrodes. Also covalent bonds are formed between
the carboxyl groups of CMC and the nanoparticle surface, exhibiting
forces which are one order of magnitude larger than caused by pure
physisorption as shown by AFM (Maver, Znidarsic, & Gaberscek, 2011).
For silicon nanoparticles, the tendency of surface Si− OH groups to un­
dergo condensation reactions strongly depends on the pH value. ATR-IR
spectroscopy and 13C solid state NMR spectroscopy demonstrated that
the condensation reactions between the surface Si− OH groups could be
modulated; at pH 3 condensation was favored (Bridel, Azais, Morcrette,
Tarascon, & Larcher, 2011; Hochgatterer et al., 2008; Mazouzi, Lestriez,

Rou´e, & Guyomard, 2009). Several authors provide data on the influ­
ence of the pH value during processing, and in all cases slurries pro­
cessed at a pH value of 3 performed better than those processed at pH 7
(Tranchot, Idrissi, Thivel, & Roue, 2016). However, there are further
conclusions from these results, namely that a defined silicon oxide/­
hydroxide layer on the nanoparticles is crucial for the battery perfor­
mance (Delpuech et al., 2014). In the absence of such a defined layer,
preparation of the slurry at pH 3 did not improve performance. Addi­
tionally, CMC features self-healing. Upon intercalation, physical or
4


W. Schlemmer et al.

Carbohydrate Polymers 265 (2021) 118063

Fig. 4. Schematic illustration of the proposed interactions occurring between the carboxyl groups of CMC and Si particles. Both H-bonds and ester bonds are
established with the silanol (Si− OH) groups on the Si particle surface. The volume expansion occurring upon lithiation causes the rupture of the bonds. During the dealloying process, the Si particles shrink, but the ester bonds are not re-established. On the other hand, the self-healing property of the weaker H-bonds permits the
contact between the binder and the Si particles to be retained.
Reproduced from the Royal Society of Chemistry from Bresser, D., Buchholz, D., Moretti, A., Varzi, A., & Passerini, S., 2018, Energy & Environmental Science, 11,
3096− 3127. under a Creative Common License 3.0.

covalent bonds between the carboxyl groups and the electrode material
are cleaved. However, new bonds accomplishing for the volume changes
can be reformed commonly termed healing in this respect (Bridel, Azaïs,
Morcrette, Tarascon, & Larcher, 2010).
One may argue that the diffusion of Li+ ions is restricted by
complexation with the carboxyl groups of the CMC, as a combination of
coordination modes was observed. It was, however, shown by NMR
spectroscopy that the diffusion of Li+ was not compromised by CMC,

showing Li-exchange rates in the nanosecond range (Casalegno et al.,
2016).
CMC may contribute also in another way to maintain battery per­
formance. As the electrolyte degrades over time, a solid electrolyte
interface (SEI) layer is formed. As a consequence, pores are blocked,
preventing the intercalation of lithium ions into the anode material
(Mazouzi et al., 2012; Oumellal et al., 2011; Radvanyi et al., 2014).
There is some evidence that CMC forms an artificial SEI layer on com­
posite electrodes, which is in contrast to PVDF based binder systems
(Delpuech et al., 2014; Jeschull et al., 2016). This seems to be also
promising for batteries having very high cell potentials (4 V), where
CMC (and also alginate) exhibited superior binder performance
compared to PVDF (Liang et al., 2021).
CMC has been used in different setups and electrode configurations
and a detailed electrochemical discussion would go beyond the scope of
this review. Apart from CMC, there is a decent amount of data available
for other polysaccharides acting as binders. In the following, we elab­
orate a few examples which are noteworthy either in terms of material
or the binding mechanism to the electrode material.
An inexhaustive list of examples of electrode materials and poly­
saccharides as binders is depicted in Table 1.

For instance, alginates have been widely used as binders as they are
capable of forming 3D networks upon crosslinking with divalent ions
such as Ca2+. They have been used in conjunction with many electrode
materials such as Li–S or silicon as binders and there are many cases
known where they outperformed PVDF and even CMC-based binders
(Bao, Zhang, Gan, Wang, & Lia, 2013; Zhang, Zhang et al., 2014).
Particularly the crosslinking allows for an entrapment of the silicon
nanoparticle electrode materials in the alginate network, while simul­

taneously increasing its mechanical strength. Furthermore, the flexi­
bility of the alginate caused by its molecular structure may play an
important role as well. Alginate is a non-random copolymer consisting of
two monosaccharides (D-mannuronic acid, and its C5 epimer L-gulur­
onic acid) that show different connectivity. While the D-mannuronic
acids are connected in a rigid β (1→4)-fashion, the α (1→4)-linkage of
the L-guluronic acid moieties provides flexibility. Consequently, volume
changes upon lithium insertion can be easier compensated, while
maintaining electrical and structural integrity of the electrode, yielding
better battery performance. Like CMC, alginates feature carboxyl groups
and a similar self-healing mechanism was proposed (Fig. 5) (Liu et al.,
2014; Yoon, Oh, Jo, Lee, & Hwang, 2014).
Recently, the mechanism of alginate-silicon anode interactions was
further explored by in operando AFM measurements in the presence of
electrolyte. The authors showed that after immersion in the dimethyl
carbonate electrolyte and operation for several hours the Young’s
modulus was 56 times higher than those of PVDF (Lee et al., 2020).
Crosslinking also improved the binder performance of other poly­
saccharides and battery chemistries. For instance, chitosan was cross­
linked using glutaraldehyde, creating a 3D network, acting as binder for
antimony anode materials in sodium batteries (Gao, Zhou, Jang, &
5


W. Schlemmer et al.

Carbohydrate Polymers 265 (2021) 118063

Table 1
Polysaccharides used in binders in different battery technologies and electrode

materials.
Polysaccharide

Technology

Electrode

Reference

Agarose

LIB
Zn-air
LIB
LIB

Graphite
Zn
Si
Graphite

(Cuesta et al., 2015)
(Masri & Mohamad, 2009)
(Hwang et al., 2016)
(Cuesta et al., 2015; Soeda,
Matsui, Yamagata, &
Ishikawa, 2013)
(Feng, Xiong, Qian, & Yin,
2014)
(Mitra, Veluri,

Chakraborthy, & Petla,
2014)
(Kumar et al., 2014)
(Zhang, Ren et al., 2019)

Alginate

CdO
CoFe2O4
CoFe2O4/rGO
Dilithium
terephtalate
Fe2O3@C (hard
C)
Fe2O3 nanotubes
LiMn2O4

LiNi1/3Co1/3Mn1/
3O2
LTO

MnO2
Si/graphite
Si/pristine C
Si/rGO
Si-NP
LIB

Li-S


SIB

P
P2-Na2/3MnO2
SnS

Amylopectin

Zn
LIB

TiO2
MnO2
Si

Amylose

LIB

Si

SIB

Si
P
S
Anthraquinone
Graphite

λ-Carrageenan

CMC

LIB
LIB

LTO
MoO3
MoS2
nano-Sn/PPy
NiFe2O4/rGO
SnS nanorods
Si
SnS2
SIB

Hard carbon
CuO
MoS2/ N-doped
hard carbon
Na3V2(PO4)2F3

Table 1 (continued )
Polysaccharide

Technology

Electrode

Reference


SnS

(Dogrusoz & Demir-Cakan,
2020)
(Kumar, Krishnan, Samal,
Mohanty, & Nayak, 2018)
(Chai et al., 2013)
(Gao et al., 2016)
(Chen, Lee et al., 2016; Lee
et al., 2018)
(Cao et al., 2018; Zhao,
Yim, Du, & Abu-Lebdeh,
2018)
(Prasanna et al., 2019)
(Kuenzel et al., 2020)
(Chen et al., 2015; Feng
et al., 2021; Kim, Cho, &
Lee, 2020; Kim, Kim, Cho,
Lee, & Lee, 2020)
(Sun, Zhong, Jiao, Shao, &
Zhang, 2014)
(Zhong et al., 2017)
(Wu et al., 2019; Yue,
Zhang, & Zhong, 2014)
(Zhong et al., 2014)
(He, Wang, Zhong, Ding, &
Zhang, 2015)
(Jeong et al., 2014; Kwon
et al., 2015)
(Wang et al., 2013)

(Cuesta et al., 2015)
(Carvalho et al., 2018)
(Yin et al., 2020)
(Carvalho et al., 2016;
Cuesta et al., 2015; Lee,
Kim, & Oh, 2014)
(Dufficy, Khan, & Fedkiw,
2015; Ling et al., 2015;
Wang et al., 2021)
(Kuruba et al., 2015)
(Li et al., 2016; Liu et al.,
2017; Lu et al., 2016)
(Hu, Cai, Huang, Zhang, &
Yu, 2019)
(Cuesta et al., 2015)
(Ling et al., 2015)
(Li, Ling et al., 2015;
Shaibani et al., 2020)
(Xu et al., 2017)
(Qi et al., 2019)
(Bie, Yang, Nuli, & Wang,
2017)
(Carvalho et al., 2016)
(Wang, Wan, & Hong, 2020;
Yoon et al., 2016)
(Gendensuren, He, & Oh,
2020)
(Zhao et al., 2019)
(Bie, Yang, Nuli, & Wang,
2016; Hapuarachchi et al.,

2020; Jin et al., 2021;
Rohan et al., 2018; You
et al., 2019)
(Masri, Nazeri, Ng, &
Mohamad, 2015)
(Cuesta et al., 2015)
(He, Zhong, Wang, & Zhang,
2017)
(L´eonard & Job, 2019)
(Chen et al., 2014)
(Liu et al., 2017)

SnS2
Chitosan

LIB

Graphite
Sb
Si
Si/graphite

(Li et al., 2019)

CMCh

(Veluri & Mitra, 2013)
(Kim, De bruyn et al., 2019;
Kong et al., 2015; Ryou,
Hong, Winter, Lee, & Choi,

2013)
(Soeda et al., 2013; Zhang,
Deng et al., 2018)
(He, Gendensuren, Kim,
Lee, & Oh, 2020;
Phanikumar et al., 2019;
Toigo, Arbizzani, Pettinger,
& Biso, 2020)
(Li, Zhao, Wang, Ding, &
Guan, 2012)
(Gendensuren & Oh, 2018)
(Liu et al., 2014)
(Huang, Chen et al., 2019)
(Kovalenko et al., 2011; Wu
et al., 2017; Zhang, Zhang
et al., 2014)
(Bao et al., 2013; Luo et al.,
2020)
(Xiao et al., 2020)
(Xu et al., 2019)
(Dogrusoz & Demir-Cakan,
2020)
(Ling et al., 2018)
(Chang et al., 2020)
(Ling et al., 2020; Murase
et al., 2012)
(Murase et al., 2012; Wen &
Zhang, 2020)
(Yoon et al., 2016)
(Zhang, Lv et al., 2020)

(Ling et al., 2017)
(Xie et al., 2012)
(Chang et al., 2019; Cuesta
et al., 2015; Shin, Park, &
Paik, 2017)
(Carvalho et al., 2016; He
et al., 2020)
(Wang, Madhavi, & Lou,
2012)
(Sen & Mitra, 2013; Volkov,
Eliseeva, Tolstopjatova, &
Kondratiev, 2020)
(Chou et al., 2011)
(Li et al., 2017)
(Tripathi & Mitra, 2014)
(You et al., 2019)
(Zhong, Zhou, Yue, Tang, &
Zhang, 2014)
(Dahbi et al., 2014)
(Fan, Yu, & Chen, 2017)
(Pang et al., 2017)

Li/S

Si/hard carbon
LNMO
Li-S

LIB


LiFePO4
LNMO
Si-NP

CN-CMCh

LIB

SnS2
LiFePO4

β-Cyclodextrin

LIB

Si

Galactomannan

LIB
LIB

Li-S
Graphite
NMC
LNMO
LTO
Si

LIB


Si/graphite
Li-S

Gellan gum

LIB

Si

Gum arabic

LIB
LIB

Graphite
Si
Li-S

SIB
LiS
LIB

Fe2O3
S
Si

Karaya gum
Pectin


LTO
Si
Si/graphite

Starch

Xanthan gum

SIB
LIB

NVP
Si

Zn-air

Zn

LIB

Graphite
LiFePO4

LIB

LiFePO4/LTO
Si/graphene
Li-S

Goodenough, 2016). As for the alginates, the network was capable of

accommodating volume changes upon sodium insertion/removal. The
same crosslinking strategy can also be applied for LIB and Si anode
materials (Chen, Lee, Cho, Kim, & Lee, 2016).
An interesting paper that compared different binders in one study

(Zhao, Yang et al., 2018)

6


W. Schlemmer et al.

Carbohydrate Polymers 265 (2021) 118063

Fig. 5. (a) Molecular structure of alginate and Ca-mediated “egg-box’’ like cross-links in Ca-alginate. (b) Expansion and self-healing mechanism of Ca-alginatecontaining silicon anodes during charge–discharge cycles.
Reproduced from (Yoon et al., 2014) with permission of the Royal Society of Chemistry.

Fig. 6. Voltammograms of cycle 1, (a) and (c), and cycle 5, (b) and (d), of the binder electrodes and bare Cu. (c) and (d) show the SG/PVDF electrode for comparison
purposes.
Reproduced from (Cuesta et al., 2015) with permission of Elsevier.
7


W. Schlemmer et al.

Carbohydrate Polymers 265 (2021) 118063

´n, Antun
˜ a, &
was published by Cuesta et al. (Cuesta, Ramos, Camea

García, 2015). The direct comparison of the binders in the same testing
setup is crucial to assess the true performance of the investigated ma­
terials (Fig. 6). They used PVDF, CMC, alginate, gum arabic (GA), xan­
than gum (XG), guar gum (GG), agar-agar (AA) and carrageenan (CG)
for graphitic electrodes in LIBs (Cuesta et al., 2015). All chosen binder
systems were electrochemically and thermally stable under the
employed experimental conditions. Alginate, CMC, XG and GG binders
(5 wt.% content) exhibited good to excellent electrochemical perfor­
mance in galvanostatic cycling experiments, which were superior to
those of SG/PVDF electrodes with higher binder content (8 wt.%). The
other systems had issues with adhesion on the electrode leading to worse
performance.
Another interesting account deals with xanthan gum as binder for Si/
graphene electrodes in LIBs (Chen et al., 2014). They manufactured the
nanocomposite using high-energy ball milling combined with thermal
treatment. Anodes prepared with XG binders exhibited a better cycling
and rate performance compared to electrodes prepared using CMC as
binder. As mentioned above, the performance increase was caused by
larger binder stiffness and the strong adhesion of the binder to the
Si-based particles, accomplishing efficiently volume changes in the
anode material.
A somehow unusual binder material is polymerized β-cyclodextrin.
This material possesses an inherent 3D network, which offers manifold
coordination sites and hydrogen bonds to Si nanoparticles (Jeong et al.,
2014). Also this binder featured self-healing as binder-Si nanoparticles
interactions recovered during cycling, thereby maintaining good elec­
trode performance. The cyclodextrin could be readily oxidized using
H2O2 yielding a highly water soluble binder which showed promising
results for sulfur composite cathodes (Wang, Yao, Monroe, Yang, & Nuli,
2013).


were equivalent to conventional separators in their proof of concept in
LiCoO2 based batteries. Later on, a variety of nanopapers, i.e. paper
made from cellulose nanofibrils (either from lignocellulose or bacterial
cellulose), were described as separators (Chun et al., 2012; Kim et al.,
2013; Jiang, Yin, Yu, Zhong, & Zhang, 2015; Yu, Park, Jang, & Good­
enough, 2016). The nanofibrils in turn featured several intrinsic ad­
vantages as their high aspect ratio and high crystallinity contributed to
mechanical and thermal stability. As cellulose nanofibrils were amphi­
philic, they were able to accommodate a large variety of electrolytes. In
addition, the morphology and porosity of nanopapers could be
fine-tuned depending on the precise electrolyte composition. They can
be produced as rather thin sheets and their flexibility allows them to be
used as separators for bendable batteries. This makes them perfect
candidates for emerging technologies such as wearable electronics,
active radio-frequency identification tags and bendable reading devices
(Leijonmarck, Cornell, Lindbergh, & Wagberg, 2013). A common
approach to fibrillate cellulosic pulps into CNF is TEMPO oxidation,
which separates the fibrils by introduction of carboxyl groups. The
resulting materials are termed TEMPO-oxidized cellulose nanofibrils
(TOCN). Depending on the pH value TOCN feature either Na+ or H+ as a
counterion, which has an influence on its separator performance as
porosity is affected by the change of the cation (Fig. 7, Table 2).
While cellulose nanofibrils have been widely employed in separators,
the use of cellulose nanocrystals is more challenging. One of the few
examples involves cellulose nanocrystals (CNC) from Cladophora which
have been subjected to a paper making process to give sheets with a
thickness of 35 μm, an average pore size of about 20 nm, and a Young’s
modulus of ca 5.9 GPa (Pan et al., 2016). Their good availability, solu­
tion processability and mesoporous structure (~20 nm average pore

size) makes them promising as future separator materials (Zhou,
Nyholm, Strømme, & Wang, 2019), suppressing the formation of lithium
dendrites. The described electrolyte contained 1 M LiPF6 in ethylene
carbonate; the separator was thermally stable up to 150 ◦ C and elec­
trochemically inert in a potential range between 0 and 5 V vs. Li+/Li.
Similar results to those with Cladophora cellulose were described for
mesoporous CNC separators produced by coassembly with tetramethyl
orthosilicate from aqueous solutions with subsequent silica removal
(Gonỗalves, Lizundia, Silva, Costa, & Lanceros-M´endez, 2019). How­
ever, challenges of CNC in separator applications were their high degree
of brittleness when processed, causing the danger of pinholes. Those and
other problems in purely cellulosic separators might be overcome by
blending with other components.
Composite membranes aim at increasing performance by combining
different types of materials. CNC for instance are promising materials
but they have shortcomings when used without additives. Examples
include composites involving PVDF-co-hexafluoropropylene (PVDFHFP), a commonly used polymer in LIB separator technology (Kelley,
Simonsen, & Ding, 2013). PVDF-HFP/CNC nanocomposite films
exhibited improved Young’s modulus and tensile strength of the mem­
branes (Bolloli et al., 2016). Another approach to improve PVDF based
separators is compounding with CMC. One example includes the use of
an Al2O3/ PVDF-HFP/CMC slurry that was deposited on a polyethylene
(PE) substrate to form a separator which showed properties superior to a
separator containing only PE (Deng et al., 2016).
MFC/polysulfonamide composites were also suitable separator
membranes for LIBs (LiCoO2, LiFePO4), working at temperatures up to
120 ◦ C (Xu, Kong, Liu, Wang et al., 2014). Similarly, cellulose/poly­
dopamine (CPD) membranes with a compact porous structure were re­
ported that exhibited superior mechanical strength and excellent
thermal dimensional stability compared to commercial separators (Xu,

Kong, Liu, Zhang et al., 2014). Lithium cobalt oxide/graphite cells with
CPD separators showed a good cycling stability and rate capability
compared to commercially available polypropylene and neat cellulose
separators. Another approach is to deposit a polymer on a commercial
separator (e.g., polypropylene). This was accomplished with hydrox­
yethyl cellulose (HEC) aerogels deposited on commercial polypropylene

5.2. Separators
Although the separator is not an electrochemically active component
in a battery, its morphology as well as structural and physical properties
have a large impact on battery performance as it provides a physical
barrier between the anode and cathode materials (Lee, Yanilmaz, Top­
rakci, Fu, & Zhang, 2014). It must be permeable to ions during charge
and discharge and chemically inert towards the electrolyte and the
electrode materials even at extreme electrochemical conditions. Sepa­
rators must be able to withstand elevated temperatures and/or corrosive
environment while maintaining their mechanical properties. The ideal
separator features a high ionic conductivity and thus low internal
resistance, which can be achieved when the electrolyte uptake of the
separator is high. For a comprehensive, general overview on separators
the reader is referred to some review articles (Arora & Zhang, 2004; Lee,
Yanilmaz et al., 2014) involving also cellulose materials (Fu & Zhong,
2019; Sheng, Tong, He, & Yang, 2017; Zhang et al., 2021; Zhang, Tian,
Shen, Song, & Yao, 2019).
Classical separators can be classified into microporous membranes,
nonwovens, polymer electrolyte membranes and composite membranes.
Each of these systems has advantages and limitations and its usage de­
pends on the final application (Lee, Yanilmaz et al., 2014; Yang & Hou,
2012).
Microporous membranes comprise the classic category for cellulose

based separators and have been used already in the first Leclanche cells,
and later in lead acid batteries and alkaline batteries. They comprise
conventional papers, paper composites, as well as nanopapers and
membranes from nanocellulose and cellulose derivatives. In 1996, Asahi
Chemical Industries (Kuribayashi, 1996) described the usage of fines
(0.5–50 μm) which were deposited on a regenerated microporous cel­
lulose film (pore sizes: 10–200 nm, thickness 39–85 μm). The membrane
exhibited good mechanical strength when immersed in different elec­
trolytes (ethylene carbonate, propylene carbonate and γ-butyrolactone).
The electrochemical properties determined by impedance spectroscopy
8


W. Schlemmer et al.

Carbohydrate Polymers 265 (2021) 118063

Fig. 7. SEM images of TOCN membranes
before (TOCN-COO–Na+, a and c) and after
protonation (TOCN-COOH, b and d), and STEM
images of a TOCN single microfiber in trans­
mission mode at low (e) and high magnification
(f). Among SEM images a–d, first and second
rows correspond to the top and bottom sides of
the membrane, respectively. The black scale
bars in the bottom right-hand corners corre­
spond to 400 μm, whereas white ones to
500 nm.
Reproduced from the American Chemical Soci­
ety from Kim, H., Guccini, V., Lu, H., SalazarAlvarez, G., Lindbergh, G., & Cornell, A.,

2019, ACS Appl. Energy Mater., 2, 1241–1250
under a Creative Commons license 3.0.

based separators (Liao et al., 2016). The performance of the separator in
a Li/LiFePO4 cell in terms of thermal stability, electrolyte uptake, ionic
conductivity, cycling performance, was superior than the neat poly­
propylene. A polyvinyl alcohol-CNF-Li composite membrane was man­
ufactured by the NIPS techniques and yielded a highly porous material
(>60 %) with good ionic conductivity (ca. 1.1 mS cm− 1) and remarkable
electrolyte uptake (Liu, Shao, Wang, Lu, & Wang, 2016). Actually, the
idea was to use CNF‒Li from mechanically treated Lyocell fibers for
combining the properties of CMC-Li like separators with the advanta­
geous properties of nanofibers. The electrolyte uptake of the separator
was increased by CNF‒Li, promoting ion transport compared to com­
mercial membranes. This concept was further improved by incorpora­
tion of ceramic particles (size 0.1–3 μm). Upon deposition of the CNF‒Li
to a polyethylene terephthalate (PET) nonwoven, better results were
obtained (Long, Wang, Zhang, Hu, & Wang, 2016).
Electrospinning is a simple method to make non‒wovens with
defined porosity and fiber diameter. Consequently, a huge variety of
separator systems is known from electrospinning with limited practical
relevance as electrospinning still suffers from low throughput, making it
an expensive method on large scale. However, the described materials
showed potential to be used and given the progress of electrospinning in
the past years (e.g., parallelization, needle-less electrospinning, coaxial
electrospinning (Huang et al., 2015)), it still could offer many oppor­
tunities for new materials in the future. As for composite membranes,
also for the electrospun substrates often mixtures of polymers were used.

For instance, PVDF/ Poly (methyl methacrylate) (PMMA)/cellulose ac­

etate (CA) composite nanofibrous mats have been manufactured in
different ratios (100:0:0, 90:10:0, 90:5:5 and 90:0:10, all m:m:m)
(Yvonne, Zhang, Zhang, Omollo, & Ncube, 2014). Pulp fibers, sodium
alginate and silica have been reported to yield a heat-resistant and
flame-retardant separator with good mechanical strength (Zhang, Yue
et al., 2014).
Further details and various other examples are listed in Table 2.
Another method to obtain nanofibrous mats is force spinning, which
is based on centrifugal forces (Weng, Xu, Alcoutlabi, Mao, & Lozano,
2015). The advantages over electrospinning include higher throughput
and easier operation conditions (Sarkar et al., 2010). However, litera­
ture on force spinning of polysaccharides in general is still limited and
this is even more pronounced for separator applications (Alcoutlabi, Lee,
& Zhang, 2015).
Polymer electrolytes are very well-suited components in high energy
density battery applications, as well as in electrochromic devices, dye
sensitized solar cells and sensors. Here, we will focus exclusively on
battery applications although requirements for the other applications
are usually not so different. There are several types of polymer elec­
˜ ez,
trolytes as depicted in Fig. 8 (Boaretto, Meabe, Martinez-Iban
Armand, & Zhang, 2020).
Here, we will focus on gel and solid polymer electrolytes. Polymer
electrolytes enable ion conduction via local polymer chain relaxation.
Such relaxation, however, is more pronounced for amorphous polymers,
9


W. Schlemmer et al.


Carbohydrate Polymers 265 (2021) 118063

Table 2
Overview on non-woven and composite separators involving polysaccharides.
Note that the last three examples contain chitin and its derivatives in contrast to
the other, cellulose based systems. ES: electrospinning; FS: force spinning; PI:
phase inversion; PM: paper making; EISA: evaporation induced self assembly;
VF: vacuum filtration; VIPS: vapor induced phase inversion, NIPS: nonsolvent
induced phase inversion; CJS: centrifugal jet spinning, NIPS: nonsolvent induced
phase inversion; CJA: centrifugal jet spinning, n.r.; not reported.
Meth.

Separator
Material

d [μm]

P
[%]

EU
[%]

T
[◦ C]

Reference

ES


Cellulose

25

75

340

(Zhang et al., 2013)

CA/PVDF

30

88

177

Cell./PVDF/
HFP
CA/PVDF/
HFP
CA/PVDF/
HNT
PVDF/CA

27

65


280

n.r.

66

355

30

86

311

n.r.

99

323


200

150

200

200

150

n.r.

PVDF/
PMMA/CA
PVDF/TPP/
CA
Cell./PAN

n.r.

94

315

n.r.

58

90

301

21

n.r.

205

Cell./PANAl2O3
m-CA


50

n.r.

286

100

87

517

FS

Cellulose

50

76

370

PI

PVDF/HFP/
Cell.
PVDF/HFP/
Cell./C/TiO2
Al2O3/MFC


n.r.

51

170

50− 60

64

210

28

68

n.r.

Al2O3/MFC/
PVDF
BC/ANF

28

56

n.r.

~ 30


84

n.r.

CaAlg/Pulp

70

68

270

TOCN− Na+

29

59

TOCN− H+

34

62

LPC/CNF

100

72



200

200
161


170

250

300

220

180
<
160

160

180

180

160

180
<150


Pulp/PPS

n.r.

61

260

n.r.

64

315

40

70

270

EISA

Pulp/PSA/
CNF
Pulp/NaAlg/
SiO2
CNC

150


~75

280

VF

BC

30

74

284

BC/HNT

30

83

369

CNF

37

70

333


CNF/PPy

10

60

n.r.

TOBC

29

91

339

TOCN

33

58

VIPS

EC/PBI

n.r.

82



200
478

NIPS

CMC/HEC

~ 100

70

131

PM

<
150
n.r.

200

180

150
n.r.

200


180
>
150

200

200

150

350

180

Table 2 (continued )
Meth.

Separator
Material

d [μm]

P
[%]

EU
[%]

T
[◦ C]


Reference

PM

C- ChNF

12

5.4

439

EISA

ChNF

~ 25

60

252

CJS

ChF

67

n.r.


n.r.


170

160

200

(Zhang, Chen et al.,
2019)
(Zhang, Shen et al.,
2017)
(Kim et al., 2017)

possessing a low Tg, i.e. amorphous matrices are favored to accomplish
for high ion conduction. Conductivity in such systems is in the range of
10− 4S cm− 1 which is below the required upper thresholds of most bat­
tery applications (approx. 10− 3 S cm− 1).
Gel polymer electrolytes (GPE) are a subclass of polymer electrolytes.
In GPE, the presence of plasticizers restrict the crystallization of the
macromolecule chains, thereby increasing ion mobility in crosslinked
polymer matrices such as polyacrylonitrile, poly(vinylidene fluoride),
poly(methyl methacrylate), and poly(ethylene oxide) derivatives. In this
context, polysaccharides offer a huge range of opportunities as they
readily form gels that do not feature a high degree of crystallinity. There
are a variety of strategies to adjust for the polysaccharide gel properties
for GPE. This involves crosslinking to form elastomeric structures, the
introduction of flexible side chains (e.g. poly(oxyethylene)), formation

of composites as well as combinations thereof.
The crosslinking strategy was one of the first approaches published
using polysaccharides. The formation of elastomeric networks leads to
extremely flexible structures with a low inherent tendency to crystallize,
while providing sufficient ionic conductivity. Schoenenberger, Le Nest
and Gandini were the pioneers in this field as they published already in
the 1990s an account on grafting oligoether on HEC followed by cross­
linking using urethane chemistry (Nest, Gandini, & Cheradame, 1988;
Schoenenberger, Le Nest, & Gandini, 1995). The role of the HEC in that
paper was to enhance the film formation – an important feature for
processability – while the oligoethers side chains and crosslinked sites
provided flexibility and sufficient relaxation to accomplish for ion
conduction.
In the years to follow, several reports proposed cellulose esters and
ethers as polymer electrolytes. These polysaccharide materials involve
hydroxypropyl cellulose (HPC), HEC (Regiani, De Oliveira Machado,
LeNest, Gandini, & Pawlicka, 2001), CA (Abidin, Yahya, Hassan, & Ali,
2014; Ahmad et al., 2020; Chen, Shi et al., 2016; Kang et al., 2016; Lee
et al., 2010), cellulose acetate butyrate (Liu, Li, Zuo, Liu, & Li, 2013; Pan
et al., 2016), cyanoethylated cellulose/cyanoethylated HPC, methyl
cellulose (Chinnam, Zhang, & Wunder, 2015; Li, Wang et al., 2015;
Mantravadi, Chinnam, Dikin, & Wunder, 2016), ethyl cellulose (Para­
cha, Ray, & Easteal, 2012) and CMC (Regiani et al., 2001; Stojadinovic,
Dushina, Trocoli, & La Mantia, 2014; Zhu et al., 2015). While intro­
duction of flexible side chains and plasticizers increases the amorphicity
and the ionic conductivity, they can negatively impact the mechanical
properties of the polymer electrolyte. For instance, an HPC-PEO with
lithium triflate as electrolyte and organic carbonates (ethylene carbon­
ate, propylene carbonate) as plasticizers, exhibited sufficient conduc­
tivity (ca. 10− 3S cm− 1) at a plasticizer content around 50 % (Yue,

McEwen, & Cowie, 2003). While conductivity was sufficient for that
battery application, the mechanical properties were insufficient. This
was overcome by crosslinking using 1,6‒diisocyanatohexane which then
allowed also to incorporate plasticizer contents up to 70 %. In addition,
the morphology of the gels (porous vs non‒porous) could be tuned by
variation of the amount of the crosslinking agents. Also pure cellulosic
materials have been used for GPE, such as CNF (Willgert, Leijonmarck,
Lindbergh, Malmstroem, & Johansson, 2014), MFC (Chiappone, Nair,
Gerbaldi, Bongiovanni, & Zeno, 2013) and macroscale materials
(Chiappone et al., 2011; Chiappone, Nair, Gerbaldi, Zeno, & Bongio­
vanni, 2014; Chiappone, Nair, Gerbaldi, Bongiovanni, & Zeno, 2015;
Jafirin, Ahmad, & Ahmad, 2013; Liu, Liu et al., 2016; Nair et al., 2009;

(Wang, Zhang,
Shao, & Liu, 2019)
(Zhang et al., 2013)
(Huang et al.,
2015)
(Wang et al., 2019)
(Yvonne et al.,
2014)
(Yvonne et al.,
2014)
(Chen, Qiu et al.,
2020)
(Jo et al., 2020)
(Jo et al., 2020)
(Chen et al., 2018)
(Weng et al., 2015)
(Li et al., 2020)

(Li et al., 2020)
(Huang, 2014)
(Huang, 2014)
(Yang et al., 2020)
(Tan et al., 2020)
(Kim, Guccini et al.,
2019)
(Kim, Guccini et al.,
2019)
(Zhang, Lan, Peng,
Hu, & Zhao, 2020)
(Zhu et al., 2020)
(Zhang, Liu et al.,
2018)
(Zhang, Yue et al.,
2014)
(Gonỗalves et al.,
2019)
(Wang et al., 2019)
(Huang, Ji et al.,
2019)
(Yoo et al., 2020)
(Wang et al., 2018)
(Huang et al.,
2020)
(Kim, Mattinen
et al., 2020)
(Chen, Zuo et al.,
2020)
(Casas,

Niederberger, &
Lizundia, 2020)

10


W. Schlemmer et al.

Carbohydrate Polymers 265 (2021) 118063

Fig. 8. Comparison and classification of electrolytes for rechargeable batteries.
Reproduced from (Boaretto et al., 2020) using a creative common license 4.0. Copyright (2020) the Electrochemical Society. Reproduced by permission of IOP
Publishing. All rights reserved.

Navarra et al., 2015; Song et al., 2004), some exhibiting competitive
electrochemical performance.
Apart from cellulose, also other polysaccharides have been investi­
gated for GPE applications but to a much lesser extent. Here, mostly
chitosan and its derivatives such as chitosan acetate, plasticized chitosan
acetate, chitosan acetate containing electrolyte, and plasticized chitosan
acetate-electrolyte complexes were manufactured into films by casting
(Osman, Ibrahim, & Arof, 2000). Regardless of the procedure, the films
were well suited for GPE applications in batteries. For other poly­
saccharides such as pectin (Andrade, Raphael, & Pawlicka, 2009), acaia
gum (Arora, Sharma, Kumar, & Kumar, 2018), guar gum (Zhang, Sudre
et al., 2017), inulin (Gachuz et al., 2020), cationic starch (Lobregas &
Camacho, 2019) or xanthan (Migliardini, Di Palma, Gaele, & Corbo,
2018), limited information is available, despite their potential for GPE
applications.


performance related to the production costs is often not known to a
sufficient extent. As the research at such a TRL is often not anymore
covered by basic research funds, participation of companies/industries
in R&D&I activities is needed. These commercial partners, from both
raw material and battery side, however need to know about the expected
costs and the performance on device level before investing into new
technologies. As it is difficult to predict these parameters, technologies
are often not further pursued despite their technological potential – a
typical chicken–or-the–egg dilemma. However, it can be expected that
environmental and sustainability aspects of batteries and their compo­
nents will become more and more important in the upcoming years,
further driving research in this area. At the moment, societal discussion
is focused on the active electrode materials (e.g. Li, Co) but it can be
expected that other components such as fluorinated long-lived, envi­
ronmentally persistent separator and polymer electrolyte materials as
well as binders will also become an issue for battery (component) pro­
ducers. At the same time, this will be an excellent opportunity to replace
these materials (at least partially) by polysaccharides.
In general, polysaccharides could be integral parts of many battery
systems as they are abundant and offer huge structural diversity already
in their natural form. For some applications, e.g., polymer electrolytes,
chemical modifications are required to adjust for the needs of the battery
component. Although polysaccharide chemistry is challenging, signifi­
cant improvements in terms of available chemical functionality, sub­
stitution patterns, and purity at reasonable cost have been accomplished
in the past years. Particularly the use of tailormade polysaccharide
components for polymer electrolytes and binders offers a huge potential
which is waiting to be investigated. However, systematic approaches to
study the influence of functional groups and side chains on battery
component performance have not been reported so far.

Another opportunity for polysaccharides comes from the materials’
side. The increasing commercialization efforts for nanocellulose and
microfibrillated cellulose production may unravel their potential in
battery components once processing has been optimized. Here, concrete
application examples for commercialization include the use of cellulose

6. Challenges and opportunities for polysaccharides in batteries
The previous chapter showed that polysaccharides have the potential
to be used in basically all components of batteries such as separator,
binder, polymer electrolyte and – not discussed in this review – pre­
cursors for carbonaceous electrode materials. However, only a few
materials are used in commercial applications, such as cellulose based
materials in separators (paper, fines, cellulose acetate) and carbox­
ymethyl cellulose as component in SBR–CMC binder systems. There are
several obstacles that prevent the commercialization of other
polysaccharide-based components in battery systems, i.e. to translate
new technologies from technology readiness levels (TRL) 2–4 to TRL
7–9. For many materials, translation from single cells into functioning
devices is challenging, which is often neglected in scientific literature.
Also, the scale up of the material production often is an issue as it re­
quires pilot lines for battery production and – in some cases –from raw
material side. Such pilot line trials, however, are needed to estimate the
production costs as well as to identify problems of the materials in terms
of processability (e.g. drying, viscosity, conductivity). Consequently, the
11


W. Schlemmer et al.

Carbohydrate Polymers 265 (2021) 118063


nanofibrils in ultrathin separators, which could be commercialized
within the next few years.
While chemistry is one way to adjust material properties, we also
briefly discussed that the formation of composites with other materials
can be an asset to make tailor-made battery components. The possibil­
ities to combine inorganic and organic materials with polysaccharides
represents another huge toolbox that still needs to be further elaborated.
Formation of composites may be necessary to achieve electronic con­
ductivity (e.g. in binders) or to adjust for amorphicity as in polymer
electrolytes. Polymer electrolytes are probably the least explored area in
this context and it can be expected that new exciting materials will be
developed in the upcoming years.

Bao, W., Zhang, Z., Gan, Y., Wang, X., & Lia, J. (2013). Enhanced cyclability of sulfur
cathodes in lithium-sulfur batteries with Na-alginate as a binder. The Journal of
Energy Chemistry, 22, 790–794.
Besenhard, J. O., & Eichinger, G. (1976). High energy density lithium cells. Journal of
Electroanalytical Chemistry and Interfacial Electrochemistry, 68, 1–18.
Bie, Y., Yang, J., Nuli, Y., & Wang, J. (2016). Oxidized starch as a superior binder for
silicon anodes in lithium-ion batteries. RSC Advances, 6, 97084–97088.
Bie, Y., Yang, J., Nuli, Y., & Wang, J. (2017). Natural karaya gum as an excellent binder
for silicon-based anodes in high-performance lithium-ion batteries. Journal of
Materials Chemistry A, 5, 1919–1924.
Bird, G. (1838). Observation of teh cyrstallization of metal salts by voltaic action
independent on the proximity of electrodes. In Report of the seventh meeting of the
British society for the advancement of science (Vol. 8). London: J. Murray. retrieved at
/>Boaretto, N., Meabe, L., Martinez-Iba˜
nez, M., Armand, M., & Zhang, H. (2020).
Review—polymer electrolytes for rechargeable batteries: From nanocomposite to

nanohybrid. Journal of the Electrochemical Society, 167, Article 070524.
Bolloli, M., Antonelli, C., Molmeret, Y., Alloin, F., Iojoiu, C., & Sanchez, J.-Y. (2016).
Nanocomposite poly(vinylidene fluoride)/nanocrystalline cellulose porous
membranes as separators for lithium-ion batteries. Electrochimica Acta, 214, 38–48.
Bresser, D., Buchholz, D., Moretti, A., Varzi, A., & Passerini, S. (2018). Alternative
binders for sustainable electrochemical energy storage – The transition to aqueous
electrode processing and bio-derived polymers. Energy & Environmental Science, 11,
3096–3127.
Bridel, J. S., Azaïs, T., Morcrette, M., Tarascon, J. M., & Larcher, D. (2010). Key
parameters governing the reversibility of Si/Carbon/CMC electrodes for Li-Ion
batteries. Chemistry of Materials, 22, 1229–1241.
Bridel, J. S., Azais, T., Morcrette, M., Tarascon, J. M., & Larcher, D. (2011). In situ
observation and long-term reactivity of Si/C/CMC composites electrodes for Li-ion
batteries. Journal of the Electrochemical Society, 158, A750–A759.
Bruce, P. G., Freunberger, S. A., Hardwick, L. J., & Tarascon, J.-M. (2012). Li-O2 and Li-S
batteries with high energy storage. Nature Materials, 11, 19–29.
Cao, P.-F., Naguib, M., Du, Z., Stacy, E., Li, B., Hong, T., et al. (2018). Effect of binder
architecture on the performance of silicon/graphite composite anodes for lithium ion
batteries. ACS Applied Materials & Interfaces, 10, 3470–3478.
Carvalho, V. D., Loeffler, N., Kim, G.-T., Marinaro, M., Wohlfahrt-Mehrens, M., &
Passerini, S. (2016). Study of water-based lithium titanate electrode processing: The
role of pH and binder molecular structure. Polymers, 8, 276.
Carvalho, D. V., Loeffler, N., Hekmatfar, M., Moretti, A., Kim, G.-T., & Passerini, S.
(2018). Evaluation of guar gum-based biopolymers as binders for lithium-ion
batteries electrodes. Electrochimica Acta, 265, 89–97.
Casalegno, M., Castiglione, F., Passarello, M., Mele, A., Passerini, S., & Raos, G. (2016).
Association and diffusion of Li+ in carboxymethylcellulose solutions for
environmentally friendly Li-ion batteries. ChemSusChem, 9, 1804–1813.
Casas, X., Niederberger, M., & Lizundia, E. (2020). A sodium-ion battery separator with
reversible voltage response based on water-soluble cellulose derivatives. ACS Applied

Materials & Interfaces, 12, 29264–29274.
Chai, L., Qu, Q., Zhang, L., Shen, M., Zhang, L., & Zheng, H. (2013). Chitosan, a new and
environmental benign electrode binder for use with graphite anode in lithium-ion
batteries. Electrochimica Acta, 105, 378–383.
Chang, W. J., Lee, G. H., Cheon, Y. J., Kim, J. T., Lee, S. I., Kim, J., et al. (2019). Direct
observation of carboxymethyl cellulose and Styrene–Butadiene rubber binder
distribution in practical graphite anodes for Li-ion batteries. ACS Applied Materials &
Interfaces, 11, 41330–41337.
Chang, H. J., Rodríguez-P´erez, I. A., Fayette, M., Canfield, N. L., Pan, H., Choi, D., et al.
(2020). Effects of water-based binders on electrochemical performance of
manganese dioxide cathode in mild aqueous zinc batteries. Carbon Energy. https://
doi.org/10.1002/cey2.84
Chen, D., Yi, R., Chen, S., Xu, T., Gordin, M. L., & Wang, D. (2014). Facile synthesis of
graphene–silicon nanocomposites with an advanced binder for high-performance
lithium-ion battery anodes. Solid State Ionics, 254, 65–71.
Chen, Y., Liu, N., Shao, H., Wang, W., Gao, M., Li, C., et al. (2015). Chitosan as a
functional additive for high-performance lithium-sulfur batteries. Journal of
Materials Chemistry A, 3, 15235–15240.
Chen, W., Zhang, L., Feng, X., Zhang, J., Guan, L., Liu, C., et al. (2018). Electrospun
flexible cellulose acetate-based separators for sodium-ion batteries with ultralong
cycle stability and excellent wettability: The role of interface chemical groups. ACS
Applied Materials & Interfaces, 10, 23883–23890.
Chen, C., Lee, S. H., Cho, M., Kim, J., & Lee, Y. (2016). Cross-linked chitosan as an
efficient binder for Si anode of Li-ion batteries. ACS Applied Materials & Interfaces, 8,
2658–2665.
Chen, Y., Qiu, L., Ma, X., Dong, L., Jin, Z., Xia, G., et al. (2020). Electrospun cellulose
polymer nanofiber membrane with flame resistance properties for lithium-ion
batteries. Carbohydrate Polymers, 234, Article 115907.
Chen, W., Shi, L., Wang, Z., Zhu, J., Yang, H., Mao, X., et al. (2016). Porous cellulose
diacetate-SiO2 composite coating on polyethylene separator for high-performance

lithium-ion battery. Carbohydrate Polymers, 147, 517–524.
Chen, Q., Zuo, X., Liang, H., Zhu, T., Zhong, Y., Liu, J., et al. (2020). A heat-resistant poly
(oxyphenylene benzimidazole)/Ethyl cellulose blended polymer membrane for
highly safe lithium-ion batteries. ACS Applied Materials & Interfaces, 12, 637–645.
Chiappone, A., Nair, J. R., Gerbaldi, C., Jabbour, L., Bongiovanni, R., Zeno, E., et al.
(2011). Microfibrillated cellulose as reinforcement for Li-ion battery polymer
electrolytes with excellent mechanical stability. Journal of Power Sources, 196,
10280–10288.

7. Conclusion and outlook
Since the invention of batteries by Volta in the 18th century, poly­
saccharides have been integral elements of energy storage systems. Back
then, there were hardly any alternatives to biobased materials in bat­
teries as organic chemistry was still in its infancy and synthetic polymers
(as well as the concept’ polymer’) were not known at all. With the rise of
low-cost synthetic polymers, the use of polysaccharides in battery
components has been compromised and currently only a few poly­
saccharide (materials) are commercially used. However, there is a huge
potential as shown in this review to develop new polysaccharide-based
battery components due to their rich structural diversity and innu­
merous possibilities to alter the polysaccharide backbone. Particularly,
the field of polymer electrolytes is a still underexplored area, where
synthetic polysaccharide chemists, composite engineers and electro­
chemists can work together to design more sustainable battery
components.
CRediT authorship contribution statement
Werner Schlemmer: Data curation, Investigation, Writing - review
& editing. Julian Selinger: Investigation, Data curation, Writing - re­
view & editing. Mathias Andreas Hobisch: Writing - original draft,
Data curation. Stefan Spirk: Data curation, Conceptualization, Funding

acquisition, Supervision, Writing - original draft.
Declaration of Competing Interest
The authors report no declarations of interest.
Acknowledgments
This work was partially funded by the Academy of Finland’s Flagship
Programme under Projects No. 318890 and 318891 (Competence Center
for Materials Bioeconomy, FinnCERES).
References
Abada, S., Marlair, G., Lecocq, A., Petit, M., Sauvant-Moynot, V., & Huet, F. (2016).
Safety focused modeling of lithium-ion batteries: A review. Journal of Power Sources,
306, 178–192.
Abidin, S. Z. Z., Yahya, M. Z. A., Hassan, O. H., & Ali, A. M. M. (2014). Conduction
mechanism of lithium bis(oxalato)borate-cellulose acetate polymer gel electrolytes.
Ionics, 20, 1671–1680.
Ahmad, M. H., Selvanathan, V., Azzahari, A. D., Sonsudin, F., Shahabudin, N., &
Yahya, R. (2020). The impact of acetylation on physical and electrochemical
characteristics of cellulose-based quasi-solid polymer electrolytes. Journal of Polymer
Research, 27.
Alcoutlabi, M., Lee, H., & Zhang, X. (2015). Nanofiber-based membrane separators for
lithium-ion batteries. MRS Online Proceedings Library (OPL), 1718, 1–5.
Andrade, J. R., Raphael, E., & Pawlicka, A. (2009). Plasticized pectin-based gel
electrolytes. Electrochimica Acta, 54, 6479–6483.
Arora, N., Sharma, V., Kumar, R., & Kumar, R. (2018). Conductivity modification of gum
acacia-based gel electrolytes. Emerging Materials Research, 7, 89–94.
Arora, P., & Zhang, Z. (2004). Battery separators. Chemical Reviews, 104, 4419–4462.
Ayrton, W. E., & Mather, T. (1911). Practical electricity. London: Cassell and Company.

12



W. Schlemmer et al.

Carbohydrate Polymers 265 (2021) 118063

Chiappone, A., Nair, J. R., Gerbaldi, C., Bongiovanni, R., & Zeno, E. (2013). Nanoscale
microfibrillated cellulose reinforced truly-solid polymer electrolytes for flexible, safe
and sustainable lithium-based batteries. Cellulose, 20, 2439–2449.
Chiappone, A., Nair, J. R., Gerbaldi, C., Zeno, E., & Bongiovanni, R. (2014). Cellulose/
acrylate membranes for flexible lithium batteries electrolytes: Balancing improved
interfacial integrity and ionic conductivity. European Polymer Journal, 57, 22–29.
Chiappone, A., Nair, J. R., Gerbaldi, C., Bongiovanni, R., & Zeno, E. (2015). UV-cured
Al2O3-laden cellulose reinforced polymer electrolyte membranes for Li-based
batteries. Electrochimica Acta, 153, 97–105.
Chinnam, P. R., Zhang, H., & Wunder, S. L. (2015). Blends of pegylated
polyoctahedralsilsesquioxanes (POSS-PEG) and methyl cellulose as solid polymer
electrolytes for Lithium batteries. Electrochimica Acta, 170, 191–201.
Choi, N.-S., Chen, Z., Freunberger, S. A., Ji, X., Sun, Y.-K., Amine, K., et al. (2012).
Challenges facing lithium batteries and electrical double-layer capacitors.
Angewandte Chemie International Edition, 51, 9994–10024.
Chou, S.-L., Gao, X.-W., Wang, J.-Z., Wexler, D., Wang, Z.-X., Chen, L.-Q., et al. (2011).
Tin/polypyrrole composite anode using sodium carboxymethyl cellulose binder for
lithium-ion batteries. Dalton Transactions, 40, 12801–12807.
Chun, S.-J., Choi, E.-S., Lee, E.-H., Kim, J. H., Lee, S.-Y., & Lee, S.-Y. (2012). Eco-friendly
cellulose nanofiber paper-derived separator membranes featuring tunable
nanoporous network channels for lithium-ion batteries. Journal of Materials
Chemistry, 22, 16618–16626.
Cuesta, N., Ramos, A., Came´
an, I., Antu˜
na, C., & García, A. B. (2015). Hydrocolloids as
binders for graphite anodes of lithium-ion batteries. Electrochimica Acta, 155,

140–147.
Dahbi, M., Nakano, T., Yabuuchi, N., Ishikawa, T., Kubota, K., Fukunishi, M., et al.
(2014). Sodium carboxymethyl cellulose as a potential binder for hard-carbon
negative electrodes in sodium-ion batteries. Electrochemistry Communications, 44,
66–69.
Delpuech, N., Mazouzi, D., Dupr´
e, N., Moreau, P., Cerbelaud, M., Bridel, J. S., et al.
(2014). Critical role of silicon nanoparticles surface on lithium cell electrochemical
performance analyzed by FTIR, Raman, EELS, XPS, NMR, and BDS spectroscopies.
The Journal of Physical Chemistry C, 118, 17318–17331.
Deng, Y., Song, X., Ma, Z., Zhang, X., Shu, D., & Nan, J. (2016). Al2O3/PVdF-HFP-CMC/
PE separator prepared using aqueous slurry and post-hot-pressing method for
polymer lithium-ion batteries with enhanced safety. Electrochimica Acta, 212,
416–425.
Dibner, B. (1964). Alessandro Volta and the electric battery. Franklin Watts.
Dogrusoz, M., & Demir-Cakan, R. (2020). Mechanochemical synthesis of SnS anodes for
sodium ion batteries. The International Journal of Energy Research, 44, 10809–10820.
Dufficy, M. K., Khan, S. A., & Fedkiw, P. S. (2015). Galactomannan binding agents for
silicon anodes in Li-ion batteries. Journal of Materials Chemistry A, 3, 12023–12030.
Editorial. (2015). A checklist for photovoltaic research. Nature Materials, 14, 1073.
Eichinger, G., & Besenhard, J. O. (1976). High energy density lithium cells. Journal of
Electroanalytical Chemistry and Interfacial Electrochemistry, 72, 1–31.
European-Commission. (2020a). A Policy Framework for Climate and Energy in the Period
from 2020 to 2030.
European-Commission. (2020b). A roadmap for moving to a competitive low carbon
economy in 2050.
Fan, M., Yu, H., & Chen, Y. (2017). High-capacity sodium ion battery anodes based on
CuO nanosheets and carboxymethyl cellulose binder. Materials Technology, 32,
598–605.
Feng, J., Xiong, S., Qian, Y., & Yin, L. (2014). Synthesis of nanosized cadmium oxide

(CdO) as a novel high capacity anode material for Lithium-ion batteries: Influence of
carbon nanotubes decoration and binder choice. Electrochimica Acta, 129, 107–112.
Feng, J., Yi, H., Lei, Z., Wang, J., Zeng, H., Deng, Y., et al. (2021). A three-dimensional
crosslinked chitosan sulfate network binder for high-performance Li–S batteries. The
Journal of Energy Chemistry, 56, 171–178.
Fraunhofer-ISE. (2021). o/?l=de&c=DE [Accessed 5.3.2021].
Freunberger, S. A. (2015). The aprotic lithium-air Battery: New insights into materials
and reactions. In ECS Conference on Electrochemical Energy Conversion & Storage.
Fu, X., & Zhong, W.-H. (2019). Biomaterials for high-energy lithium-based batteries:
Strategies, challenges, and perspectives. Advanced Energy Materials, 9, Article
1901774.
Gachuz, E. J., Castillo-Santill´
an, M., Juarez-Moreno, K., Maya-Cornejo, J., MartinezRicha, A., Andrio, A., et al. (2020). Electrical conductivity of an all-natural and
biocompatible semi-interpenetrating polymer network containing a deep eutectic
solvent. Green Chemistry, 22, 5785–5797.
Galvani, L., & Volta, A. (1791). De viribus electricitatis in motu musculari commentarius.
retrieved from />Gao, H., Zhou, W., Jang, J.-H., & Goodenough, J. B. (2016). Cross-linked chitosan as a
polymer network binder for an antimony anode in Sodium-Ion Batteries. Advanced
Energy Materials, 6, Article 1502130.
Gendensuren, B., & Oh, E.-S. (2018). Dual-crosslinked network binder of alginate with
polyacrylamide for silicon/graphite anodes of lithium ion battery. Journal of Power
Sources, 384, 379–386.
Gendensuren, B., He, C., & Oh, E.-S. (2020). Preparation of pectin-based dual-crosslinked
network as a binder for high performance Si/C anode for LIBs. The Korean Journal of
Chemical Engineering, 37, 366–373.
Gogotsi, Y., & Simon, P. (2011). True performance metrics in electrochemical energy
storage. Science, 334, 917918.
Gonỗalves, R., Lizundia, E., Silva, M. M., Costa, C. M., & Lanceros-M´
endez, S. (2019).
Mesoporous cellulose nanocrystal membranes as battery separators for

environmentally safer lithium-ion batteries. ACS Applied Energy Materials, 2,
3749–3761.

Hamer, W. J. (1965). Standard cells: Their construction, maintenance, and characteristics.
National Bureau of Standards Monograph #84: US National Bureau of Standards.
Hapuarachchi, S. N. S., Wasalathilake, K. C., Nerkar, J. Y., Jaatinen, E., O’Mullane, A. P.,
& Yan, C. (2020). Mechanically robust tapioca starch composite binder with
improved ionic conductivity for sustainable lithium-ion batteries. ACS Sustainable
Chemistry & Engineering, 8, 9857–9865.
He, J., Wang, J., Zhong, H., Ding, J., & Zhang, L. (2015). Cyanoethylated carboxymethyl
chitosan as water soluble binder with enhanced adhesion capability and
electrochemical performances for LiFePO4 cathode. Electrochimica Acta, 182,
900–907.
He, J., Zhong, H., Wang, J., & Zhang, L. (2017). Investigation on xanthan gum as novel
water soluble binder for LiFePO4 cathode in lithium-ion batteries. The Journal of
Alloys and Compounds, 714, 409–418.
He, C., Gendensuren, B., Kim, H., Lee, H., & Oh, E. S. (2020). Electrochemical
performance of polysaccharides modified by the introduction of SO3H as binder for
high-powered Li4Ti5O12 anodes in lithium-ion batteries. Journal of Electroanalytical
Chemistry, 876, Article 114532.
Hochgatterer, N. S., Schweiger, M. R., Koller, S., Raimann, P. R., Wă
ohrle, T., Wurm, C.,
et al. (2008). Silicon/graphite composite electrodes for high-capacity anodes:
Influence of binder chemistry on cycling stability. Electrochem. Solid State Lett., 11,
A76–A80.
Hu, S., Cai, Z., Huang, T., Zhang, H., & Yu, A. (2019). A modified natural polysaccharide
as a high-performance binder for silicon anodes in lithium-ion batteries. ACS Applied
Materials & Interfaces, 11, 4311–4317.
Huang, X. (2014). Performance evaluation of a non-woven lithium ion battery separator
prepared through a paper-making process. Journal of Power Sources, 256, 96–101.

Huang, F., Xu, Y., Peng, B., Su, Y., Jiang, F., Hsieh, Y.-L., et al. (2015). Coaxial
electrospun cellulose-core fluoropolymer-shell fibrous membrane from recycled
cigarette filter as separator for high performance lithium-ion battery. ACS Sustainable
Chemistry & Engineering, 3, 932–940.
Huang, C., Ji, H., Yang, Y., Guo, B., Luo, L., Meng, Z., et al. (2020). TEMPO-oxidized
bacterial cellulose nanofiber membranes as high-performance separators for lithiumion batteries. Carbohydrate Polymers, 230, Article 115570.
Huang, H., Chen, R., Yang, S., Zhang, W., Fang, Y., Li, L., et al. (2019). High-performance
Si flexible anode with rGO substrate and Ca2+ crosslinked sodium alginate binder
for lithium ion battery. Synthetic Metals, 247, 212–218.
Huang, C., Ji, H., Guo, B., Luo, L., Xu, W., Li, J., et al. (2019). Composite nanofiber
membranes of bacterial cellulose/halloysite nanotubes as lithium ion battery
separators. Cellulose, 26, 6669–6681.
Hwang, G., Kim, J.-M., Hong, D., Kim, C.-K., Choi, N.-S., Lee, S.-Y., et al. (2016).
Multifunctional natural agarose as an alternative material for high-performance
rechargeable lithium-ion batteries. Green Chemistry, 18, 2710–2716.
Jafirin, S., Ahmad, I., & Ahmad, A. (2013). Potential use of cellulose from kenaf in
polymer electrolytes based on MG49 rubber composites. BioResources, 8, 5947–5964.
Jeong, Y. K., Kwon, T.-w., Lee, I., Kim, T.-S., Coskun, A., & Choi, J. W. (2014).
Hyperbranched β-Cyclodextrin polymer as an effective multidimensional binder for
silicon anodes in Lithium rechargeable batteries. Nano Letters, 14, 864870.
Jeschull, F., Lindgren, F., Lacey, M. J., Bjă
orefors, F., Edstră
om, K., & Brandell, D. (2016).
Influence of inactive electrode components on degradation phenomena in nano-Si
electrodes for Li-ion batteries. Journal of Power Sources, 325, 513–524.
Jiang, F., Yin, L., Yu, Q., Zhong, C., & Zhang, J. (2015). Bacterial cellulose nanofibrous
membrane as thermal stable separator for lithium-ion batteries. Journal of Power
Sources, 279, 21–27.
Jin, B., Wang, D., Song, L., Cai, Y., Ali, A., Hou, Y., et al. (2021). Biomass-derived
fluorinated corn starch emulsion as binder for silicon and silicon oxide based anodes

in lithium-ion batteries. Electrochimica Acta, 365, Article 137359.
Jo, J. H., Jo, C.-H., Qiu, Z., Yashiro, H., Shi, L., Wang, Z., et al. (2020). Nature-derived
cellulose-based composite separator for sodium-ion batteries. Frontiers in Chemistry,
8, 153.
Kang, W., Ma, X., Zhao, H., Ju, J., Zhao, Y., Yan, J., et al. (2016). Electrospun cellulose
acetate/poly(vinylidene fluoride) nanofibrous membrane for polymer lithium-ion
batteries. Journal of Solid State Electrochemistry, 20, 2791–2803.
Kelley, J., Simonsen, J., & Ding, J. (2013). Poly(vinylidene fluoride-cohexafluoropropylene) nanocomposites incorporating cellulose nanocrystals with
potential applications in lithium ion batteries. Journal of Applied Polymer Science,
127, 487–493.
Kim, J.-H., Kim, J.-H., Choi, E.-S., Yu, H. K., Kim, J. H., Wu, Q., et al. (2013). Colloidal
silica nanoparticle-assisted structural control of cellulose nanofiber paper separators
for lithium-ion batteries. Journal of Power Sources, 242, 533–540.
Kim, J.-K., Choi, B., Jin, J., Kim, D. H., Joo, S. H., Cha, A., et al. (2017). Hierarchical
chitin fibers with aligned nanofibrillar architectures: A nonwoven-mat separator for
lithium metal batteries. ACS Nano, 11, 6114–6121.
Kim, S., Cho, M., & Lee, Y. (2020). Multifunctional Chitosan-rGO network binder for
enhancing the cycle stability of Li-S batteries. Advanced Functional Materials, 30,
Article 1907680.
Kim, S., De bruyn, M., Louvain, N., Alauzun, J. G., Brun, N., Macquarrie, D. J., et al.
(2019). Green electrode processing using a seaweed-derived mesoporous carbon
additive and binder for LiMn2O4 and LiNi1/3Mn1/3Co1/3O2 lithium ion battery
electrodes. Sustainable Energy & Fuels, 3, 450–456.
Kim, H., Guccini, V., Lu, H., Salazar-Alvarez, G., Lindbergh, G., & Cornell, A. (2019).
Lithium ion battery separators based on carboxylated cellulose nanofibers from
wood. ACS Applied Energy Materials, 2, 1241–1250.
Kim, S., Kim, D. H., Cho, M., Lee, W. B., & Lee, Y. (2020). Fast-charging Lithium–Sulfur
batteries enabled via lean binder content. Small, 16, Article 2004372.

13



W. Schlemmer et al.

Carbohydrate Polymers 265 (2021) 118063

Kim, H., Mattinen, U., Guccini, V., Liu, H., Salazar-Alvarez, G., Lindstroem, R. W., et al.
(2020). Feasibility of chemically modified cellulose nanofiber membranes as
lithium-ion battery separators. ACS Applied Materials & Interfaces, 12, 41211–41222.
Kong, F., Longo, R. C., Yeon, D.-H., Yoon, J., Park, J.-H., Liang, C., et al. (2015).
Multivalent Li-Site doping of Mn oxides for Li-Ion batteries. The Journal of Physical
Chemistry C, 119, 21904–21912.
Kovalenko, I., Zdyrko, B., Magasinski, A., Hertzberg, B., Milicev, Z., Burtovyy, R., et al.
(2011). A major constituent of brown algae for use in high-capacity Li-Ion batteries.
Science, 334, 75.
Kuenzel, M., Porhiel, R., Bresser, D., Asenbauer, J., Axmann, P., Wohlfahrt-Mehrens, M.,
et al. (2020). Deriving structure-performance relations of chemically modified
chitosan binders for sustainable high-voltage LiNi0.5Mn1.5O4 cathodes. Batteries &
Supercaps, 3, 129.
Kumar, P. R., Kollu, P., Santhosh, C., Eswara Varaprasada Rao, K., Kim, D. K., &
Grace, A. N. (2014). Enhanced properties of porous CoFe2O4-reduced graphene
oxide composites with alginate binders for Li-ion battery applications. New Journal of
Chemistry, 38, 3654–3661.
Kumar, S., Krishnan, S., Samal, S. K., Mohanty, S., & Nayak, S. K. (2018). Toughening of
petroleum based (DGEBA) epoxy resins with various renewable resources based
flexible chains for high performance applications: A review. Industrial & Engineering
Chemistry Research, 57, 2711–2726.
Kuribayashi, I. (1996). Characterization of composite cellulosic separators for
rechargeable lithium-ion batteries. Journal of Power Sources, 63, 87–91.
Kuruba, R., Datta, M. K., Damodaran, K., Jampani, P. H., Gattu, B., Patel, P. P., et al.

(2015). Guar gum: Structural and electrochemical characterization of natural
polymer based binder for silicon–Carbon composite rechargeable Li-ion battery
anodes. Journal of Power Sources, 298, 331–340.
Kwon, T.-w., Jeong, Y. K., Deniz, E., AlQaradawi, S. Y., Choi, J. W., & Coskun, A. (2015).
Dynamic cross-linking of polymeric binders based on host–Guest interactions for
silicon anodes in Lithium ion batteries. ACS Nano, 9, 11317–11324.
Lee, J. M., Nguyen, D. Q., Lee, S. B., Kim, H., Ahn, B. S., Lee, H., et al. (2010). Cellulose
triacetate-based polymer gel electrolytes. Journal of Applied Polymer Science, 115,
32–36.
Lee, S. H., Lee, J. H., Nam, D. H., Cho, M., Kim, J., Chanthad, C., et al. (2018). Epoxidized
natural rubber/chitosan network binder for silicon anode in lithium-ion battery. ACS
Applied Materials & Interfaces, 10, 16449–16457.
Lee, M., Reddi, R. K. R., Choi, J., Liu, J., Huang, X., Cho, H., et al. (2020). In-operando
AFM characterization of mechanical property evolution of Si anode binders in liquid
electrolyte. ACS Applied Energy Materials, 3, 1899–1907.
Lee, B.-R., Kim, S.-j., & Oh, E.-S. (2014). Bio-derivative galactomannan gum binders for
Li4Ti5O12Negative electrodes in lithium-ion batteries. Journal of the Electrochemical
Society, 161, A2128–A2132.
Lee, H., Yanilmaz, M., Toprakci, O., Fu, K., & Zhang, X. (2014). A review of recent
developments in membrane separators for rechargeable lithium-ion batteries. Energy
& Environmental Science, 7, 3857–3886.
Leijonmarck, S., Cornell, A., Lindbergh, G., & Wagberg, L. (2013). Single-paper flexible
Li-ion battery cells through a paper-making process based on nano-fibrillated
cellulose. Journal of Materials Chemistry A, 1, 4671–4677.

eonard, A. F., & Job, N. (2019). Safe and green Li-ion batteries based on LiFePO4 and
Li4Ti5O12 sprayed as aqueous slurries with xanthan gum as common binder.
Materials Today Energy, 12, 168–178.
Lestriez, B., Bahri, S., Sandu, I., Rou´
e, L., & Guyomard, D. (2007). On the binding

mechanism of CMC in Si negative electrodes for Li-ion batteries. Electrochemistry
Communications, 9, 2801–2806.
Li, J., Lewis, R. B., & Dahn, J. R. (2007). Sodium carboxymethyl cellulose: A potential
binder for Si negative electrodes for Li-Ion batteries. Electrochemical and Solid-State
Letters, 10, A17–A20.
Li, J., Zhao, Y., Wang, N., Ding, Y., & Guan, L. (2012). Enhanced performance of a MnO2graphene sheet cathode for lithium ion batteries using sodium alginate as a binder.
Journal of Materials Chemistry, 22, 13002–13004.
Li, Q., Yang, H., Xie, L., Yang, J., Nuli, Y., & Wang, J. (2016). Guar gum as a novel binder
for sulfur composite cathodes in rechargeable lithium batteries. Chemical
Communications, 52, 13479–13482.
Li, C., Wang, X., Li, S., Li, Q., Xu, J., Liu, X., et al. (2017). Optimization of NiFe2O4/rGO
composite electrode for lithium-ion batteries. Applied Surface Science, 416, 308–317.
Li, C., Sarapulova, A., Zhao, Z., Fu, Q., Trouillet, V., Missiul, A., et al. (2019).
Understanding the lithium storage mechanism in core–shell Fe2O3@C hollow
nanospheres derived from metal–organic frameworks: An in operando synchrotron
radiation diffraction and in operando X-ray absorption spectroscopy study. Chemistry
of Materials, 31, 5633–5645.
Li, L., Li, H., Wang, Y., Zheng, S., Zou, Y., & Ma, Z. (2020). Poly(vinylidenefluoridehexafluoropropylene)/cellulose/carboxylic TiO2 composite separator with high
temperature resistance for lithium-ion batteries. Ionics, 26, 4489–4497.
Li, G., Ling, M., Ye, Y., Li, Z., Guo, J., Yao, Y., et al. (2015). Acacia Senegal–Inspired
bifunctional binder for longevity of lithium–sulfur batteries. Advanced Energy
Materials, 5, Article 1500878.
Li, M., Wang, X., Wang, Y., Chen, B., Wu, Y., & Holze, R. (2015). A gel polymer
electrolyte based on composite of nonwoven fabric and methyl cellulose with good
performance for lithium ion batteries. RSC Advances, 5, 52382–52387.
Liang, J., Chen, D., Adair, K., Sun, Q., Holmes, N. G., Zhao, Y., et al. (2021). Insight into
prolonged cycling life of 4 V all-solid-state polymer batteries by a high-voltage stable
binder. Advanced Energy Materials, 11, Article 2002455.
Liao, H., Zhang, H., Hong, H., Li, Z., Qin, G., Zhu, H., et al. (2016). Novel cellulose
aerogel coated on polypropylene separators as gel polymer electrolyte with high

ionic conductivity for lithium-ion batteries. The Journal of Membrane Science, 514,
332–339.

Liedel, C. (2020). Sustainable battery materials from biomass. ChemSusChem, 13,
2110–2141.
Ling, M., Xu, Y., Zhao, H., Gu, X., Qiu, J., Li, S., et al. (2015). Dual-functional gum arabic
binder for silicon anodes in lithium ion batteries. Nano Energy, 12, 178–185.
Ling, M., Zhang, L., Zheng, T., Feng, J., Guo, J., Mai, L., et al. (2017). Nucleophilic
substitution between polysulfides and binders unexpectedly stabilizing lithium
sulfur battery. Nano Energy, 38, 82–90.
Ling, L., Bai, Y., Wang, Z., Ni, Q., Chen, G., Zhou, Z., et al. (2018). Remarkable effect of
sodium alginate aqueous binder on anatase TiO2 as high-performance anode in
sodium ion batteries. ACS Applied Materials & Interfaces, 10, 5560–5568.
Ling, H. Y., Wang, C., Su, Z., Chen, S., Chen, H., Qian, S., et al. (2020). Amylopectin from
Glutinous Rice as a sustainable binder for high-performance silicon anodes. Energy &
Environmental Materials. />Liu, J., Li, W., Zuo, X., Liu, S., & Li, Z. (2013). Polyethylene-supported polyvinylidene
fluoride-cellulose acetate butyrate blended polymer electrolyte for lithium ion
battery. Journal of Power Sources, 226, 101–106.
Liu, J., Zhang, Q., Wu, Z.-Y., Wu, J.-H., Li, J.-T., Huang, L., et al. (2014). A highperformance alginate hydrogel binder for the Si/C anode of a Li-ion battery.
Chemical Communications, 50, 6386–6389.
Liu, J., Galpaya, D. G. D., Yan, L., Sun, M., Lin, Z., Yan, C., et al. (2017). Exploiting a
robust biopolymer network binder for an ultrahigh-areal-capacity Li–S battery.
Energy & Environmental Science, 10, 750–755.
Liu, K., Liu, M., Cheng, J., Dong, S., Wang, C., Wang, Q., et al. (2016). Novel cellulose/
polyurethane composite gel polymer electrolyte for high performance lithium
batteries. Electrochimica Acta, 215, 261–266.
Liu, C., Shao, Z., Wang, J., Lu, C., & Wang, Z. (2016). Eco-friendly polyvinyl alcohol/
cellulose nanofiber-Li+ composite separator for high-performance lithium-ion
batteries. RSC Advances, 6, 97912–97920.
Lobregas, M. O. S., & Camacho, D. H. (2019). Gel polymer electrolyte system based on

starch grafted with ionic liquid: Synthesis, characterization and its application in
dye-sensitized solar cell. Electrochimica Acta, 298, 219–228.
Long, J., Wang, X., Zhang, H., Hu, J., & Wang, Y. (2016). A nano-based multilayer
separator for lithium rechargeable battery. International Journal of Electrochemical
Science, 11, 6552–6563.
Lu, L., Han, X., Li, J., Hua, J., & Ouyang, M. (2013). A review on the key issues for
lithium-ion battery management in electric vehicles. Journal of Power Sources, 226,
272–288.
Lu, Y.-Q., Li, J.-T., Peng, X.-X., Zhang, T., Deng, Y.-P., Wu, Z.-Y., et al. (2016). Achieving
high capacity retention in lithium-sulfur batteries with an aqueous binder.
Electrochemistry Communications, 72, 79–82.
Luo, C., Hu, E., Gaskell, K. J., Fan, X., Gao, T., Cui, C., et al. (2020). A chemically
stabilized sulfur cathode for lean electrolyte lithium sulfur batteries. PNAS, 117,
14712–14720.
Mantravadi, R., Chinnam, P. R., Dikin, D. A., & Wunder, S. L. (2016). High conductivity,
high strength solid electrolytes formed by in situ encapsulation of ionic liquids in
nanofibrillar methyl cellulose networks. ACS Applied Materials & Interfaces, 8,
13426–13436.
Masri, M. N., & Mohamad, A. A. (2009). Effect of adding potassium hydroxide to an agar
binder for use as the anode in Zn–air batteries. Corrosion Science, 51, 3025–3029.
Masri, M. N., Nazeri, M. F. M., Ng, C. Y., & Mohamad, A. A. (2015). Tapioca binder for
porous zinc anodes electrode in zinc–air batteries. Journal of King Saud University Engineering Sciences, 27, 217–224.
Maver, U., Znidarsic, A., & Gaberscek, M. (2011). An attempt to use atomic force
microscopy for determination of bond type in lithium battery electrodes. Journal of
Materials Chemistry, 21, 4071–4075.
Mazouzi, D., Lestriez, B., Rou´e, L., & Guyomard, D. (2009). Silicon composite electrode
with high capacity and long cycle life. Electrochem. Solid State Lett., 12, A215–A218.
Mazouzi, D., Delpuech, N., Oumellal, Y., Gauthier, M., Cerbelaud, M., Gaubicher, J.,
et al. (2012). New insights into the silicon-based electrode’s irreversibility along
cycle life through simple gravimetric method. Journal of Power Sources, 220,

180–184.
Migliardini, F., Di Palma, T. M., Gaele, M. F., & Corbo, P. (2018). Solid and acid
electrolytes for Al-air batteries based on xanthan-HCl hydrogels. Journal of Solid State
Electrochemistry, 22, 2901–2916.
Mitra, S., Veluri, P. S., Chakraborthy, A., & Petla, R. K. (2014). Electrochemical
properties of spinel cobalt ferrite nanoparticles with sodium alginate as interactive
binder. ChemElectroChem, 1, 1068–1074.
Mizushima, K., Jones, P. C., Wiseman, P. J., & Goodenough, J. B. (1980). LixCoO2
(0Research Bulletin, 15, 783–789.
Mongird, K., Fotedar, V., Viswanathan, V., Koritarov, V., Balducci, P., Hadjerioua, B.,
et al. (2019). Energy storage technology and cost characterization report. Richland, WA,
USA: Pacific Northwest National Lab and HydroWIRES.
Murase, M., Yabuuchi, N., Han, Z.-J., Son, J.-Y., Cui, Y.-T., Oji, H., et al. (2012). Cropderived polysaccharides as binders for high-capacity Silicon/Graphite-Based
electrodes in lithium-ion batteries. ChemSusChem, 5, 2307–2311.
Nair, J. R., Gerbaldi, C., Chiappone, A., Zeno, E., Bongiovanni, R., Bodoardo, S., et al.
(2009). UV-cured polymer electrolyte membranes for Li-cells: Improved mechanical
properties by a novel cellulose reinforcement. Electrochemistry Communications, 11,
1796–1798.
Navarra, M. A., Dal Bosco, C., Moreno, J. S., Vitucci, F. M., Paolone, A., & Panero, S.
(2015). Synthesis and characterization of cellulose-based hydrogels to be used as gel
electrolytes. Membranes, 5, 810–823.
Nest, J.-F. L., Gandini, A., & Cheradame, H. (1988). Crosslinked polyethers as media for
ionic conduction. British Polymer Journal, 20, 253–268.

14


W. Schlemmer et al.


Carbohydrate Polymers 265 (2021) 118063

Nirmale, T. C., Kale, B. B., & Varma, A. J. (2017). A review on cellulose and lignin based
binders and electrodes: Small steps towards a sustainable lithium ion battery.
International Journal of Biological Macromolecules, 103, 1032–1043.
Nitta, N., Wu, F., Lee, J. T., & Yushin, G. (2015). Li-ion battery materials: Present and
future. Materials Today, 18, 252–264.
NOAA. (2021). />imate-change-atmospheric-carbon-dioxide [Accessed 5.3.2021].
Obrovac, M. N., & Chevrier, V. L. (2014). Alloy negative electrodes for Li-Ion batteries.
Chemical Reviews, 114, 11444–11502.
Osman, Z., Ibrahim, Z. A., & Arof, A. K. (2000). Conductivity enhancement due to ion
dissociation in plasticized chitosan based polymer electrolytes. Carbohydrate
Polymers, 44, 167–173.
Oumellal, Y., Delpuech, N., Mazouzi, D., Dupre, N., Gaubicher, J., Moreau, P., et al.
(2011). The failure mechanism of nano-sized Si-based negative electrodes for lithium
ion batteries. Journal of Materials Chemistry, 21, 6201–6208.
Pan, R., Cheung, O., Wang, Z., Tammela, P., Huo, J., Lindh, J., et al. (2016). Mesoporous
Cladophora cellulose separators for lithium-ion batteries. Journal of Power Sources,
321, 185–192.
Pang, Y., Zhang, S., Liu, L., Liang, J., Sun, Z., Wang, Y., et al. (2017). Few-layer MoS2
anchored at nitrogen-doped carbon ribbons for sodium-ion battery anodes with high
rate performance. Journal of Materials Chemistry A, 5, 17963–17972.
Paracha, R. N., Ray, S., & Easteal, A. J. (2012). Grafting of LiAMPS on ethyl cellulose: A
route to the fabrication of superior quality polyelectrolyte gels for rechargeable
lithium ion batteries. Journal of Materials Science, 47, 3698–3705.
Phanikumar, V. V. N., Rikka, V. R., Das, B., Gopalan, R., Appa Rao, B. V., & Prakash, R.
(2019). Investigation on polyvinyl alcohol and sodium alginate as aqueous binders
for lithium-titanium oxide anode in lithium-ion batteries. Ionics, 25, 2549–2561.
Prasanna, K., Subburaj, T., Jo, Y. N., Santhoshkumar, P., Karthikeyan, S. K. S. S.,
Vediappan, K., et al. (2019). Chitosan complements entrapment of silicon inside

nitrogen doped carbon to improve and stabilize the capacity of Li-ion batteries.
Scientific Reports, 9, 3318.
Qi, Q., Deng, Y., Gu, S., Gao, M., Hasegawa, J.-y., Zhou, G., et al. (2019). L-Cysteinemodified Acacia gum as a multifunctional binder for Lithium–Sulfur batteries. ACS
Applied Materials & Interfaces, 11, 47956–47962.
Radvanyi, E., Porcher, W., De Vito, E., Montani, A., Franger, S., & Jouanneau Si Larbi, S.
(2014). Failure mechanisms of nano-silicon anodes upon cycling: An electrode
porosity evolution model. Physical Chemistry Chemical Physics (PCCP), 16,
17142–17153.
Regiani, A. M., De Oliveira Machado, G., LeNest, J.-F., Gandini, A., & Pawlicka, A.
(2001). Cellulose derivatives as solid electrolyte matrixes. Macromolecular Symposia,
175, 45–53.
Rezvanizaniani, S. M., Liu, Z., Chen, Y., & Lee, J. (2014). Review and recent advances in
battery health monitoring and prognostics technologies for electric vehicle (EV)
safety and mobility. Journal of Power Sources, 256, 110–124.
Rohan, R., Kuo, T.-C., Chiou, C.-Y., Chang, Y.-L., Li, C.-C., & Lee, J.-T. (2018). Low-cost
and sustainable corn starch as a high-performance aqueous binder in silicon anodes
via in situ cross-linking. Journal of Power Sources, 396, 459–466.
Roy, P., & Srivastava, S. K. (2015). Nanostructured anode materials for lithium ion
batteries. Journal of Materials Chemistry A, 3, 2454–2484.
Ryou, M.-H., Hong, S., Winter, M., Lee, H., & Choi, J. W. (2013). Improved cycle lives of
LiMn2O4 cathodes in lithium ion batteries by an alginate biopolymer from seaweed.
Journal of Materials Chemistry A, 1, 15224–15229.
Sarkar, K., Gomez, C., Zambrano, S., Ramirez, M., de Hoyos, E., Vasquez, H., et al.
(2010). Electrospinning to forcespinning™. Materials Today, 13, 12–14.
Schoenenberger, C., Le Nest, J. F., & Gandini, A. (1995). Polymer electrolytes based on
modified polysaccharides. 2. Polyether-modified cellulosics. Electrochimica Acta, 40,
2281–2284.
Sen, U. K., & Mitra, S. (2013). High-rate and high-energy-density lithium-ion battery
anode containing 2D MoS2 nanowall and cellulose binder. ACS Applied Materials &
Interfaces, 5, 1240–1247.

Shaibani, M., Mirshekarloo, M. S., Singh, R., Easton, C. D., Cooray, M. C. D., Eshraghi, N.,
et al. (2020). Expansion-tolerant architectures for stable cycling of ultrahigh-loading
sulfur cathodes in lithium-sulfur batteries. Science Advances, 6, Article eaay2757.
Shao, Y., El-Kady, M. F., Sun, J., Li, Y., Zhang, Q., Zhu, M., et al. (2018). Design and
mechanisms of asymmetric supercapacitors. Chemical Reviews, 118, 9233–9280.
Sheng, J., Tong, S., He, Z., & Yang, R. (2017). Recent developments of cellulose materials
for lithium-ion battery separators. Cellulose, 24, 4103–4122.
Shin, D., Park, H., & Paik, U. (2017). Cross-linked poly(acrylic acid)-carboxymethyl
cellulose and styrene-butadiene rubber as an efficient binder system and its
physicochemical effects on a high energy density graphite anode for Li-ion batteries.
Electrochemistry Communications, 77, 103–106.
Soeda, K., Matsui, Y., Yamagata, M., & Ishikawa, M. (2013). Application of alginate
binders to graphite electrodes and characterization of their lithium-ion battery
performance. ECS Transactions, 53, 93–102.
Song, M.-K., Kim, Y.-T., Cho, J.-Y., Cho, B. W., Popov, B. N., & Rhee, H.-W. (2004).
Composite polymer electrolytes reinforced by non-woven fabrics. Journal of Power
Sources, 125, 10–16.
Spencer, J. N., Bodner, G. M., & Rickard, L. H. (2010). Chemistry: Structure and dynamics.
John Wiley & Sons.
Stojadinovic, J., Dushina, A., Trocoli, R., & La Mantia, F. (2014). Electrochemical
characterization of gel electrolytes for aqueous lithium-ion batteries. ChemPlusChem,
79, 1507–1511.
Sun, M., Zhong, H., Jiao, S., Shao, H., & Zhang, L. (2014). Investigation on
carboxymethyl chitosan as new water soluble binder for LiFePO4 cathode in Li-Ion
batteries. Electrochimica Acta, 127, 239–244.

Tan, L., Li, Z., Shi, R., Quan, F., Wang, B., Ma, X., et al. (2020). Preparation and
properties of an alginate-based fiber separator for lithium-ion batteries. ACS Applied
Materials & Interfaces, 12, 38175–38182.
Toigo, C., Arbizzani, C., Pettinger, K.-H., & Biso, M. (2020). Study on different waterbased binders for Li4Ti5O12 electrodes. Molecules, 25, 2443.

Tranchot, A., Idrissi, H., Thivel, P. X., & Roue, L. (2016). Impact of the slurry pH on the
expansion/contraction behavior of silicon/carbon/carboxymethylcellulose
electrodes for Li-ion batteries. Journal of the Electrochemical Society, 163,
A1020–A1026.
Tripathi, A. M., & Mitra, S. (2014). Tin sulfide (SnS) nanorods: Structural, optical and
lithium storage property study. RSC Advances, 4, 10358–10366.
Veluri, P. S., & Mitra, S. (2013). Enhanced high rate performance of [small alpha]-Fe2O3
nanotubes with alginate binder as a conversion anode. RSC Advances, 3,
15132–15138.
Volkov, A. I., Eliseeva, S. N., Tolstopjatova, E. G., & Kondratiev, V. V. (2020). Enhanced
electrochemical performance of MoS 2 anode material with novel composite binder.
Journal of Solid State Electrochemistry, 24, 1607–1614.
Wang, Z., Madhavi, S., & Lou, X. W. (2012). Ultralong α-MoO3 nanobelts: Synthesis and
effect of binder choice on their Lithium storage properties. The Journal of Physical
Chemistry C, 116, 12508–12513.
Wang, J., Yao, Z., Monroe, C. W., Yang, J., & Nuli, Y. (2013). Carbonyl-β-cyclodextrin as
a novel binder for sulfur composite cathodes in rechargeable lithium batteries.
Advanced Functional Materials, 23, 1194–1201.
Wang, Z., Pan, R., Ruan, C., Edstroem, K., Stromme, M., & Nyholm, L. (2018). Redoxactive separators for lithium-ion batteries. Advancement of Science, 5, Article
1700663.
Wang, S., Zhang, D., Shao, Z., & Liu, S. (2019). Cellulosic materials-enhanced sandwich
structure-like separator via electrospinning towards safer lithium-ion battery.
Carbohydrate Polymers, 214, 328–336.
Wang, J.-T., Wan, C.-C., & Hong, J.-L. (2020). Polymer blends of pectin/poly(acrylic
acid) as efficient binders for silicon anodes in lithium-ion batteries.
ChemElectroChem, 7, 3106–3115.
Wang, H., Wei, D., Wan, Z., Du, Q., Zhang, B., Ling, M., et al. (2021). Epoxy and amide
crosslinked polarity enhanced polysaccharides binder for silicon anode in lithiumion batteries. Electrochimica Acta, 368.
Wen, Y., & Zhang, H. (2020). Highly stretchable polymer binder engineered with
polysaccharides for silicon microparticles as high-performance anodes.

ChemSusChem, 13, 3887–3892.
Weng, B., Xu, F., Alcoutlabi, M., Mao, Y., & Lozano, K. (2015). Fibrous cellulose
membrane mass produced via forcespinning for lithium-ion battery separators.
Cellulose, 22, 1311–1320.
Willgert, M., Leijonmarck, S., Lindbergh, G., Malmstroem, E., & Johansson, M. (2014).
Cellulose nanofibril reinforced composite electrolytes for lithium ion battery
applications. Journal of Materials Chemistry A, 2, 13556–13564.
Wu, Z.-Y., Deng, L., Li, J.-T., Huang, Q.-S., Lu, Y.-Q., Liu, J., et al. (2017). Multiple
hydrogel alginate binders for Si anodes of lithium-ion battery. Electrochimica Acta,
245, 371–378.
Wu, Z.-H., Yang, J.-Y., Yu, B., Shi, B.-M., Zhao, C.-R., & Yu, Z.-L. (2019). Self-healing
alginate–carboxymethyl chitosan porous scaffold as an effective binder for silicon
anodes in lithium-ion batteries. Rare Metals, 38, 832–839.
Xiao, W., Sun, Q., Banis, M. N., Wang, B., Li, W., Li, M., et al. (2020). Understanding the
critical role of binders in phosphorus/carbon anode for sodium-ion batteries through
unexpected mechanism. Advanced Functional Materials, 30, Article 2000060.
Xie, L., Zhao, L., Wan, J.-l., Shao, Z.-q., Wang, F.-j., & Lv, S.-y. (2012). The
electrochemical performance of carboxymethyl cellulose Lithium as a binding
material for anthraquinone cathodes in Lithium batteries. Journal of the
Electrochemical Society, 159, A499–A505.
Xu, L., Sitinamaluwa, H., Li, H., Qiu, J., Wang, Y., Yan, C., et al. (2017). Low cost and
green preparation process for α-Fe2O3@gum arabic electrode for high performance
sodium ion batteries. Journal of Materials Chemistry A, 5, 2102–2109.
Xu, H., Jiang, K., Zhang, X., Zhang, X., Guo, S., & Zhou, H. (2019). Sodium alginate
enabled advanced layered manganese-based cathode for sodium-ion batteries. ACS
Applied Materials & Interfaces, 11, 26817–26823.
Xu, Q., Kong, Q., Liu, Z., Wang, X., Liu, R., Zhang, J., et al. (2014). Cellulose/
polysulfonamide composite membrane as a high performance lithium-ion battery
separator. ACS Sustainable Chemistry & Engineering, 2, 194–199.
Xu, Q., Kong, Q., Liu, Z., Zhang, J., Wang, X., Liu, R., et al. (2014). Polydopamine-coated

cellulose microfibrillated membrane as high performance lithium-ion battery
separator. RSC Advances, 4, 7845–7850.
Yang, M., & Hou, J. (2012). Membranes in Lithium ion batteries. Membranes, 2, 367–383.
Yang, Y., Huang, C., Gao, G., Hu, C., Luo, L., & Xu, J. (2020). Aramid nanofiber/bacterial
cellulose composite separators for lithium-ion batteries. Carbohydrate Polymers, 247,
Article 116702.
Yim, C.-H., Courtel, F. M., & Abu-Lebdeh, Y. (2013). A high capacity silicon–graphite
composite as anode for lithium-ion batteries using low content amorphous silicon
and compatible binders. Journal of Materials Chemistry A, 1, 8234–8243.
Yin, Z.-W., Zhang, T., Zhang, S.-J., Deng, Y.-P., Peng, X.-X., Wang, J.-Q., et al. (2020).
Understanding the role of water-soluble guar gum binder in reducing capacity fading
and voltage decay of Li-rich cathode for Li-ion batteries. Electrochimica Acta, 351,
Article 136401.
Yoo, G., Pyo, S., Gong, Y. J., Cho, J., Kim, H., Kim, Y. S., et al. (2020). Highly reliable
quinone-based cathodes and cellulose nanofiber separators: Toward eco-friendly
organic lithium batteries. Cellulose, 27, 6707–6717.
Yoon, J., Oh, D. X., Jo, C., Lee, J., & Hwang, D. S. (2014). Improvement of desolvation
and resilience of alginate binders for Si-based anodes in a lithium ion battery by

15


W. Schlemmer et al.

Carbohydrate Polymers 265 (2021) 118063
Zhang, S., Ren, S., Han, D., Xiao, M., Wang, S., & Meng, Y. (2019). Aqueous sodium
alginate as binder: Dramatically improving the performance of dilithium
terephthalate-based organic lithium ion batteries. Journal of Power Sources, 438,
Article 227007.
Zhang, T.-W., Shen, B., Yao, H.-B., Ma, T., Lu, L.-L., Zhou, F., et al. (2017). Prawn shell

derived chitin nanofiber membranes as advanced sustainable separators for Li/Naion batteries. Nano Letters, 17, 4894–4901.
Zhang, B., Sudre, G., Quintard, G., Serghei, A., David, L., Bernard, J., et al. (2017). Guar
gum as biosourced building block to generate highly conductive and elastic ionogels
with poly(ionic liquid) and ionic liquid. Carbohydrate Polymers, 157, 586–595.
Zhang, T.-W., Tian, T., Shen, B., Song, Y.-H., & Yao, H.-B. (2019). Recent advances on
biopolymer fiber based membranes for lithium-ion battery separators. Compos.
Comm., 14, 7–14.
Zhang, J., Yue, L., Kong, Q., Liu, Z., Zhou, X., Zhang, C., et al. (2014). Sustainable, heatresistant and flame-retardant cellulose-based composite separator for highperformance lithium ion battery. Scientific Reports, 4, 3935.
Zhang, L., Zhang, L., Chai, L., Xue, P., Hao, W., & Zheng, H. (2014). A coordinatively
cross-linked polymeric network as a functional binder for high-performance silicon
submicro-particle anodes in lithium-ion batteries. Journal of Materials Chemistry A, 2,
19036–19045.
Zhao, J., Yang, X., Zhang, Y., Loh, X. J., Hu, X., Chen, G., et al. (2019). Utilization of
biomass pectin polymer to build high efficiency electrode architectures with sturdy
construction and fast charge transfer structure to boost sodium storage performance
for NASICON-type cathode. Journal of Materials Chemistry A, 7, 1548–1555.
Zhao, J., Yang, X., Yao, Y., Gao, Y., Sui, Y., Zou, B., et al. (2018). Moving to aqueous
binder: A valid approach to achieving high-rate capability and long-term durability
for sodium-ion battery. Advancement of Science, 5, Article 1700768.
Zhao, X., Yim, C.-H., Du, N., & Abu-Lebdeh, Y. (2018). Crosslinked chitosan networks as
binders for silicon/graphite composite electrodes in Li-ion batteries. Journal of the
Electrochemical Society, 165, A1110–A1121.
Zhong, H., Zhou, P., Yue, L., Tang, D., & Zhang, L. (2014). Micro/nano-structured SnS2
negative electrodes using chitosan derivatives as water-soluble binders for Li-ion
batteries. Journal of Applied Electrochemistry, 44, 45–51.
Zhong, H., Lu, J., He, A., Sun, M., He, J., & Zhang, L. (2017). Carboxymethyl chitosan/
poly(ethylene oxide) water soluble binder: Challenging application for 5V
LiNi0.5Mn1.5O4 cathode. Journal of Material Science and Technology Research, 33,
763–767.
Zhou, S., Nyholm, L., Strømme, M., & Wang, Z. (2019). Cladophora cellulose: Unique

biopolymer nanofibrils for emerging energy, environmental, and life science
applications. Accounts of Chemical Research, 52, 2232–2243.
Zhu, Y. S., Xiao, S. Y., Li, M. X., Chang, Z., Wang, F. X., Gao, J., et al. (2015). Natural
macromolecule based carboxymethyl cellulose as a gel polymer electrolyte with
adjustable porosity for lithium ion batteries. Journal of Power Sources, 288, 368–375.
Zhu, C., Zhang, J., Xu, J., Yin, X., Wu, J., Chen, S., et al. (2020). Facile fabrication of
cellulose/polyphenylene sulfide composite separator for lithium-ion batteries.
Carbohydrate Polymers, 248, Article 116753.

calcium-mediated cross-linking. Physical Chemistry Chemical Physics (PCCP), 16,
25628–25635.
Yoon, D.-E., Hwang, C., Kang, N.-R., Lee, U., Ahn, D., Kim, J.-Y., et al. (2016).
Dependency of electrochemical performances of silicon lithium-ion batteries on
glycosidic linkages of polysaccharide binders. ACS Applied Materials & Interfaces, 8,
4042–4047.
You, R., Han, X., Zhang, Z., Li, L., Li, C., Huang, W., et al. (2019). An environmental
friendly cross-linked polysaccharide binder for silicon anode in lithium-ion batteries.
Ionics, 25, 4109–4118.
Yu, B.-C., Park, K., Jang, J.-H., & Goodenough, J. B. (2016). Cellulose-based porous
membrane for suppressing Li dendrite formation in lithium-sulfur battery. ACS
Energy Letters, 1, 633–637.
Yu, F., Li, S., Chen, W., Wu, T., & Peng, C. (2019). Biomass-derived materials for
electrochemical energy storage and conversion: Overview and perspectives. Energy &
Environmental Materials, 2, 55–67.
Yuan, M., & Liu, K. (2020). Rational design on separators and liquid electrolytes for safer
lithium-ion batteries. Journal of Energy Chemistry, 43, 58–70.
Yue, Z., McEwen, I. J., & Cowie, J. M. G. (2003). Novel gel polymer electrolytes based on
a cellulose ester with PEO side chains. Solid State Ionics, 156, 155–162.
Yue, L., Zhang, L., & Zhong, H. (2014). Carboxymethyl chitosan: A new water soluble
binder for Si anode of Li-ion batteries. Journal of Power Sources, 247, 327–331.

Yvonne, T., Zhang, C., Zhang, C., Omollo, E., & Ncube, S. (2014). Properties of
electrospun PVDF/PMMA/CA membrane as lithium based battery separator.
Cellulose, 21, 2811–2818.
Zhang, J., Liu, Z., Kong, Q., Zhang, C., Pang, S., Yue, L., et al. (2013). Renewable and
superior thermal-resistant cellulose-based composite nonwoven as lithium-ion
battery separator. ACS Applied Materials & Interfaces, 5, 128–134.
Zhang, Z., Fang, Z., Xiang, Y., Liu, D., Xie, Z., Qu, D., et al. (2021). Cellulose-based
material in lithium-sulfur batteries: A review. Carbohydrate Polymers, 255, Article
117469.
Zhang, T.-W., Chen, J.-L., Tian, T., Shen, B., Peng, Y.-D., Song, Y.-H., et al. (2019).
Sustainable separators for high-performance lithium ion batteries enabled by
chemical modifications. Advanced Functional Materials, 29, Article 1902023.
Zhang, S.-J., Deng, Y.-P., Wu, Q.-H., Zhou, Y., Li, J.-T., Wu, Z.-Y., et al. (2018). Sodiumalginate-Based binders for lithium-rich cathode materials in lithium-ion batteries to
suppress voltage and capacity fading. ChemElectroChem, 5, 1321–1329.
Zhang, F., Lan, X., Peng, H., Hu, X., & Zhao, Q. (2020). ’’Trojan horse’’ camouflage
strategy for high-performance cellulose paper and separators. Advanced Functional
Materials, 30, Article 2002169.
Zhang, H., Liu, J., Guan, M., Shang, Z., Sun, Y., Lu, Z., et al. (2018). Nanofibrillated
cellulose (NFC) as a pore size mediator in the preparation of thermally resistant
separators for lithium ion batteries. ACS Sustainable Chemistry & Engineering, 6,
4838–4844.
Zhang, H., Lv, Z., Liang, Q., Xia, H., Zhu, Z., Zhang, W., et al. (2020). Highly elastic
binders incorporated with helical molecules to improve the electrochemical stability
of black phosphorous anodes for sodium-ion batteries. Batteries & Supercaps, 3,
101–107.

16




×