You are on page 1of 61

______________________________

Structural Battery Electrolytes


Bachelor Thesis
KA103X

______________________________
Authors:
Max Bjerkensjö (maxbje@kth.se)
Amanda Halvarsson (amahal@kth.se)
Julia Rune (juliarun@kth.se)
Pernilla Öberg (perobe@kth.se)

Supervisors:
Samuel Emilsson (samemi@kth.se)
Lynn Schneider (lmsc@kth.se)

Examiner:
Mats Johansson (matskg@kth.se)

KTH Royal Institute of Technology


Department of Fibre and Polymer Technology

Stockholm, Sweden 2021


Abstract

Structural batteries are multifunctional; providing electrochemical energy storage synergistically with a
load-bearing function that enables their integration into the body panels of electric devices and vehicles.
Thus, massless energy can be achieved. As a composite material, it is composed of reinforcing carbon
fibre electrodes embedded in an electrolyte matrix. To realize this concept, the electrolyte must
simultaneously transfer mechanical load and transport ions between electrodes. The following study
builds on a phase-separated polymer electrolyte, created using polymerization-induced phase separation
via thermal curing, formulated by Schneider et al. and Ihrner et al.. The impact of the incorporation of
thiols for copolymerization and as cross-linking agents for the polymer network was researched along
with use of an EC:PC-based solvent. The three thiols studied were: trimethylolpropane
tris(3-mercaptopropionate) (3TMP), pentaerythritol tetrakis(3-mercaptopropionate) (4PER), and
dipentaerythritol hexakis-(3-mercaptopropionate) (6DPER). These differed in regard to the amount of
thiol functional groups present. Ionic conductivity, thermo-mechanical performance and
structure-property relationships were studied across 3 laboratory phases. The first phase concerned the
effect of thiol-functionality, the thiol functional group ratio relative to the allyl group present in the
primary monomer, and the solvent interaction. 6DPER was concluded to be the most promising
cross-linking agent. During the second phase, the effect of electrolyte content was evaluated with an
optimum of 45 weight% determined. The third phase concluded the study, wherein a half-cell was
assembled with the optimized electrolyte formulation showing improved capacity relative to previous
studies. The results developed here contribute to the understanding of structural battery electrolyte
systems and their continued research to meet application demands.

Key Words: structural battery electrolytes, polymerization-induced phase separation, bicontinuous


morphology, lithium-ion conductivity, thiol-ene chemistry

1
Sammanfattning

Strukturella batterier är multifunktionella; de tillhandahåller lagring av elektrokemisk energi samtidigt


som de bidrar med en lastbärande funktion. Tillsammans möjliggör detta att batteriet kan integreras i
karossen hos ett elektriskt fordon eller apparat. Denna multifunktionalitet möjliggör således en avsevärd
reducering i fordonets vikt. Kompositmaterialet är förstärkt av kolfiberelektroder, innesluten i en
elektrolytstruktur. För att förverkliga detta koncept måste batteriets elektrolyt kunna motstå mekanisk
belastning, samtidigt som den transporterar joner mellan batteriets elektroder. Denna studie syftar till att
bygga vidare på konceptet av fas-separerade polymerelektrolyter, skapade från polymerisationsinducerad
fasseparation via termisk härdning, vilket är en teknik utvecklad av Schneider et al. och Ihrner et al.
Vidare undersöks effekten av att dels använda en elektrolytlösning baserad på EC:PC, men även att
inkorporera tioler till polymernätverket. Tvärbindningsmolekylerna som användes i denna studie
inkluderade trimetylolpropan tris(3-merkaptopropionat) (3TMP), pentaerythritol
tetrakis(3-merkaptopropionat) (4PER), och dipentaerythritol hexakis-(3-merkaptopropionat) (6DPER).
Dessa skiljer sig i antal funktionella tiolgrupper. Konduktivitet, termo-mekanisk prestanda och
strukturberoende egenskaper undersöktes genom tre laborativa faser. Den första fasen behandlade
inverkan på elektrolytsystemet av ändrat lösningsmedel, tiol-funktionalitet samt tiolgruppförhållandet
gentemot allyl gruppen på den primära monomeren. Sampolymeren innehållandes 6DPER uppvisade bäst
multifunktionalitet, varpå denna utvecklades vidare i fas två där en optimal sammansättning fastställdes
som bestod utav 45 viktprocent jonlösning. I den slutliga fasen konstruerades en halv-cell baserat på den
tidigare optimerade elektrolytkompositionen; den uppmätta kapaciteten visar tydlig förbättring jämfört
med tidigare forskning. Resultatet som erhölls i denna studie bidrar till förståendet av strukturella
batteri-elektrolyter samt den forskning som en dag kan komma att förverkliga strukturella batterier och
dess tillämpningskrav.

Nyckelord: Strukturella batteri-elektrolyter, polymerisations-inducerad fasseparation, bikontinuerlig


morfologi, litium-jon konduktivitet, tiol-en kemi

2
Abbreviations

SBE Structural battery electrolyte


CFC Carbon fibre composite
Tg Glass transition temperature [℃]
SPE Solid polymer electrolyte
GPE Gel polymer electrolyte
PIPS Polymerization induced phase separation
LiTFS Lithium trifluoromethanesulfonate
Monomer A Bisphenol A dimethacrylate
Monomer B Bisphenol A ethoxylate dimethacrylate
DMMP Dimethyl methylphosphonate
EC Ethylene carbonate
PC Propylene carbonate
AIBN 2,2′-azobis(2-methylpropionitrile)
3TMP Trimethylolpropane tris(3-mercaptopropionate)
4PER Pentaerythritol tetrakis(3-mercaptopropionate)
6DPER Dipentaerythritol hexakis-(3-mercaptopropionate)
FG ratio Functional group ratio
EIS Electrochemical impedance spectroscopy
SEM Scanning electron microscopy
FTIR Fourier-Transform Infrared Spectroscopy
DMA Dynamic Mechanical Analysis

3
Table of Contents
1. Introduction 1
2.1 Structural Batteries: Components & Function 3
2.2 Electrochemistry of Lithium-Ion Batteries 4
2.3 Structural Battery Electrolytes: Parameters 5
2.3.1 Structural Battery Electrolyte Parameters 5
2.3.2 Thermal & Mechanical Properties 6
2.4 Structural Battery Electrolytes: Alternatives 7
2.4.1 Homogenous Electrolytes 7
2.2.3 Heterogeneous Phase-Separated Electrolytes 9
2.5 Polymerization-Induced Phase Separation 9
2.4 Thiol-ene Chemistry 13

3. Aim 15

4. Methodology 16
4.1 Materials Used and Their Origin 16
4.2 Techniques and Procedures 16
4.2.1 Structural Battery Electrolyte Preparation 17
4.2.2 Electrochemical Impedance Spectroscopy 17
4.2.3 Scanning Electron Microscopy & Gravimetric Analysis 18
4.2.4 Fourier-Transform Infrared Spectroscopy 18
4.2.5 Dynamic Mechanical Analysis 19
4.2.6 Half-Cell Preparation 19

5. Results & Discussion 20


5.1 Phase 1: The effect of thiol functionality on electrolyte properties 20
5.1.1 Thermal & Mechanical Data 20
5.1.2 Conductivity 23
5.1.3 Multifunctional Analysis 25
5.1.4 Curing Performance 26
5.1.5 Morphology 28
5.2 Phase 2: Effect of Liquid Electrolyte Concentration 33
5.2.1 Thermal & Mechanical Data 33
5.2.2 Conductivity 35
5.2.3 Multifunctional Analysis 36
5.2.4 Curing Performance 37
5.2.5 Morphology 38
5.3 Phase 3: Half-cell Performance 41

4
6. Conclusion 43

7. Potential errors 45

8. References 47

Appendix A 51

Appendix B 53

Appendix C 55

5
1. Introduction
In the past decade, the extensive use of larger lithium-ion batteries have become the norm for electric
automobile manufacturers. As the demand for non-fossil fuel engines increases in conjunction with
anthropogenic global warming, lithium-ion batteries are considered a viable alternative for a greener
vehicular industry. Tesla is a leading car producer actively striving towards this goal. However, despite
their persistence in making their products as efficient as possible, they still fall short regarding the
battery's mass. Currently, their Model S battery makes up 25% of the vehicle’s total weight, making it a
parasitic with regard to mass [1]. Reducing the overall weight of vehicles can lead to significant
decreases in energy usage. This would not only increase overall efficiency and reduce indirect carbon
emissions, but also lead to advancement for other mechanical transport modes. For example: (1) aircrafts,
for which no electrical commercial equivalent exists due to lack of energy density in relation to weight;
and (2) aerospace-crafts, such as satellites and space-probes, that require a significant amount of energy
when in orbit and are currently costly to launch [1] [2].

Structural batteries are a solution to these problems. By integrating a battery function into the body of a
device, the energy storage material will also fulfill the function of a structural load carrier. The result:
large mass savings without loss of structural integrity [1]. There are two approaches to the concept of
structural batteries: embedded vs laminated. In embedded structural batteries, the load-bearing qualities
are carried out by outer face-sheets, for example aluminum plates, that embed the battery in a sandwiched
composite structure. An industrial example is the variant that Tesla suggested in 2020, where an
embedded structural battery is incorporated as the vehicle’s floor [3]. Laminated structural battery, or
laminated structural electrodes, are where the electrode materials instead possess both load-bearing
properties in conjunction with a battery function [2]. The main difference between these two types is that
the embedded battery does not itself possess an inherent structural integrity like the laminated structural
battery, as this is provided by the stacking of face-sheets instead. Thus, the embedded structural battery
does not exhibit multifunctionality, which is the defining feature of laminated structural batteries. This is
achieved through the combination of carbon fibre electrodes and a separator embedded in an adhesive
electrolyte matrix to form a carbon fibre composite (CFC) [2].

1
These CFCs can be incorporated into the load-bearing body panels of an electric vehicle or device. This
multifunctionality enables the significant reduction in system level mass, improving electrical efficiency.
For example: a structural battery roof may achieve mass-savings of 62% in comparison to an original
steel-based design [4]. The following illustration shown in Figure 1 shows how a structural battery can be
integrated into the body of a car, and depicts the electrochemistry utilized.

Figure 1. Illustration of the laminated structural battery concept as applied to an electric vehicle [5].

The primary obstacle in the realization of laminated structural batteries is the antagonistic relationship
between mechanical properties and ionic conductivity present in the structural battery electrolyte (SBE)
[1]. For the ionic conductivity it is preferred to have a minimum conductivity of at least 10−3 S cm−1 in
order to obtain a functioning transport of ions from between electrodes [6]. As for mechanical properties,
a storage modulus at a minimum at 500 MPa is preferred [1]. Further research is required to meet these
application demands, which is the focal point of this study.

2
2. Theoretical Background
The following section is concerned with the theoretical underpinnings of this study. First, the components
of a laminated structural battery are presented with regard to their function; this introduces the conceptual
framework within which a multifunctional electrolyte must operate. A brief explanation of the
electrochemistry of lithium-ion batteries is then provided. Subsequently, the SBE concept is reviewed.
Included herein is an overview of relevant performance parameters and the composition of various
electrolyte materials in addition to their potential use in a structural battery application. This is followed
by a description of the previous research conducted on the formulation of SBEs using reaction-induced
phase separation (PIPS). Finally, thiol-ene chemistry is examined as pertinent to this study.

2.1 Structural Batteries: Components & Function


The realization of a structural battery requires that its components perform multifunctionally, providing
structural reinforcement synergistically with the electrochemical storage of energy. For this purpose,
CFCs provide the combined mechanical and electrical performance through which a battery function,
comparable to lithium-ion chemistry, may be introduced [7]. These primarily consist of reinforcing carbon
fibres impregnated with a thermoset polymer matrix. The combination of the macroscopically identifiable
materials therein contributes to a superior structural result, whereby the polymer matrix binds the carbon
fibres together [8] [9].

Components of a laminated CFC structural battery with relative placement is depicted in Figure 2,
showcasing a 3-layer composite. In this technology, commercially available polyacrylonitrile-based
carbon fibres constitute the electrodes [10]. The high strength-to-weight ratio and stiffness of the material
makes it favourable for the intended application, providing the skeletal backbone to the battery composite.
Moreover, its carbonaceous microstructure readily facilitates reversible Li-intercalations at low electrode
potentials. It is also conductive; this is beneficial for electrode performance, as it enables direct contact of
electrons with the external circuit connecting anode to cathode [7]. Electron flow is thus facilitated, and
the size of the adjoining current collectors can be minimized, allowing further mass-savings. The positive
electrode also requires surface coating with an electrochemically active material to create the necessary
potential difference; LiFePO4 can be used for this purpose, with polyvinylidene fluoride as binder and

3
carbon black to promote electric contact [11]. To prevent short-circuits, a permeable and electrically
insulating separator, composed of a thin glass-microfiber filter, divides the two electrodes. All
constituents are immersed in the SBE thermoset, which must exhibit Li-ion conduction in conjunction
with mechanical load transfer between the reinforcing fibres [1] [10]. The latter is dependent on the
interfacial adhesion between polymer and carbon fibre, which thus is an important factor for the
mechanical performance of the composite.

Figure 2. A schematic illustration of a laminated structural battery [12].

2.2 Electrochemistry of Lithium-Ion Batteries


The proposed CFC has a layered architecture with a carbon-based negative electrode much like that of the
state-of-the-art lithium-ion battery, that are currently favored by the relevant applications for their high
specific energy, slow self-discharge and lack of memory effect [12]. This similarity enables the
integration of Li-ion functionality into the structural battery, described as follows: Upon discharge,
lithium intercalated into the anode is oxidized and migrates to the cathode through the electrolyte
medium. The released electrons are extracted by the current collectors, and concurrently move through
the external circuit connecting the two electrodes. Thus, an electric current is provided by the battery that
can power e.g. an electric motor. Subsequently, recombination occurs at the cathode in a half-reduction
reaction. The reverse mechanism takes place during charge [13].

4
The controlled full charge and discharge of a battery unit in this manner constitutes a charge cycle, which
can be used to measure capacity at different current densities. This is indicative of the amount of
electrochemical energy available during discharge and is measured in Ampere-hours per gram (Ah g-1)
[14].

2.3 Structural Battery Electrolytes: Parameters


SBEs are multifunctional electrolytes that combine mechanical and conductive properties to achieve
mass-less energy storage, in contrast to monofunctional materials. Research has shown that the most
promising approach is to create a bicontinuous structure - providing mechanical stability from one phase
and ionic conductivity from the other [15] [16]. This section addresses relevant parameters necessary for
assessing the performance of an SBE; included are ionic conductivity as well as thermal and mechanical
properties.

2.3.1 Structural Battery Electrolyte Parameters


Upon applying an external electric potential, the motion of lithium-ions induces ion conduction ( currents
of electricity). The rate at which these lithium-ions migrate between electrodes is dependent on the ionic
conductivity (σ, Siemens per meter (S m-1)) of the electrolyte [17]. It is determined as stated in equation 1:

σ = ∑ 𝑞𝑛𝑢 (1)

Equation 1 displays how conductivity is dependent on n, the concentration of charge carrier species; q,
the carrier's charge (-1 for lithium); and u, the mobility of the carrier. Thus, ionic conductivity is
proportional to the charge carrier concentration and its ionic mobility [17]. This indicates that an
enhanced rate performance is achieved through an electrolyte environment that promotes these factors.

The transport mechanisms by which charge-carriers travel in a liquid are: 1) diffusion, dependent on
concentration polarizations in the electrolyte medium; and 2) the migration that occurs due to an applied
electric field. The resulting ionic mobility is estimated by the following correlation in equation 2:

𝑠 𝑞 𝑞𝐷𝑁𝐴
𝑢 = 𝐸
= 6πη𝑎
= 𝑅𝑇
(2)

5
Here, s is the velocity of the ion; E is the electric field present between two planar electrodes; ηis the
viscosity of the medium; D is the diffusion coefficient; NA is the molar concentration; T is the
temperature; and a is the radius of the spherically-estimated cation [18] [19]. The aforementioned
variables are all contributing factors to the degree of mobility attained by a Li-ion in the electrolyte
system.

The effective concentration of Li+ cations is related to the dissociation degree of lithium salts in the
electrolyte medium, which in turn is governed by the dielectric constant of the solvent; defined as the
ratio of the material’s electric permeability in relation to a vacuum [19]. A large constant is indicative of
strong solvation, and it is necessary to minimize drag from anionic interference that may disturb mobility.
Ion-solvent interactions are therefore a key factor to consider.

2.3.2 Thermal & Mechanical Properties


Storage modulus, E’, measures the capacity of a material to store mechanical energy imposed upon it as
an elastic deformation. Thus, stiffness can be quantified as a response to force exerted on the material
[20]. This is determined using Dynamic Mechanical Analysis (DMA) and expressed in the unit Pascal
(Pa). It is highly temperature dependent, as visualized in Figure 3, which allows mechanical properties to
be further investigated as a function of temperature.

Figure 3. Storage modulus (E´) as a function of temperature [21].

6
When increasing the temperature of a polymeric material, the structure undergoes several transitions due
to internal expansions stemming from increased molecular chain mobility. The storage modulus will be
affected accordingly - the curve depicts a decrease in stiffness upon a rise in temperature. When the glass
transition temperature (Tg) is reached, the storage modulus drops significantly (as evident from region
(3)); the polymer loses its original rigid structure and achieves a rubbery state. By further increasing the
temperature, the material melts (as represented by region (1)). As for thermosets - investigated in this
study- the material will not reach a melting point due to the high degree of cross-linking whereby covalent
bonds are formed between chains in the network [21]. Tg can either be determined from the storage
modulus, at the temperature at which the DMA curve drastically declines, or from a so called tan𝛿-curve .
Tan𝛿, the loss angle, as a function of temperature will provide a curve with the peak corresponding to the
Tg transition [22]. The characteristics of modulus and Tg are considerably indicative of the polymer
structure; they entail substantial information regarding rigidity, elasticity, and overall mechanical
resistance in terms of factors such as cross-linking density, chain mobility and degree of polymerization
[21].

2.4 Structural Battery Electrolytes: Alternatives


The following section contains a review of homogenous electrolyte systems currently established within
the scientific community, based on their advantages and deficiencies regarding practical application
within a structural battery. Expounding upon this, the necessity of a heterogenous electrolyte is motivated.

2.4.1 Homogenous Electrolytes


Traditionally, liquid electrolytes have been the preferred medium for use in lithium-ion batteries. The
electrochemical performance is excellent; at ambient temperatures, ionic conductivities of ­~10-2 S cm-1 are
attained, making it the industry benchmark for comparison [6]. This is due to the enhanced mobility of
charge-carrier species, as migration is facilitated by a low-viscosity system. The electrolyte-electrode
interface contact is also promoted, with the liquid easily extending into the porous electrode and thus
improving battery performance through enhanced intercalation [19]. However, as a liquid it provides no
mechanical integrity and requires heavy casing to prevent leakage and ignition of volatile organic

7
electrolytes [6]. This poses considerable security and functionality issues for the purposes of a structural
battery.

Previous work has investigated solid polymer electrolytes (SPE) and gel polymer electrolytes (GPEs) as
alternatives to liquid systems to mitigate the previously mentioned instability issues. In an SPE, soft
thermoset polymers commonly contain polar polyethylene glycol (PEG) segments to dissolve and
transport Li-ions through coordination with the electronegative oxygen atoms [6]. This occurs in the
amorphous regions, facilitated by the segmental motion of chains. By suppressing crystallization,
conductivity can be thus improved by promoting ion-hopping through increased chain mobility [19].
However, this simultaneously decreases the stiffness of the polymer network that provides the electrolyte
system with mechanical strength [23] [6]. Moreover, the Tg is lowered; this is contradictory to the
requirements of a structural battery, where a high Tg above operating temperature is necessary for
sufficient mechanical performance. Typically, state-of-the-art SPEs exhibit an E-modulus of only 100
MPa and ion conductivities in the range 10-5 – 10-8 S cm-1 at ambient temperatures [1] [6] [17]. The
practical use of SPEs in a structural battery application is therefore limited.

GPEs are similarly problematic. By adding a plasticising organic solvent to a solid polymer matrix,
commonly composed of PEG, peroxyacyl nitrates (PAN), polyvinylidene fluoride (PVDF) or poly(methyl
methacrylate) (PMMA), conductivity can be promoted to upwards of 10-4-10-2 S cm-1 through gelation
[17] [24]. This is due to the ion transport primarily occurring in the liquid plasticiser, that becomes
intermixed with the soft polymer in an integrated membrane. However, this occurs at the expense of the
mechanical properties otherwise provided by the polymer network. These are reduced as chain mobility
increases and the electrolyte becomes a soft amorphous solid which lacks load-bearing properties [6].
Hence, GPEs are also insufficient for use in a structural battery .

Notable, multiple studies report an improvement upon mechanical performance following


copolymerization and the incorporation of cross-linking agents for SPEs and GPEs [25] [26] [27]. This
creates a macroscopically more elastic and mechanically strengthened system, which results in an overall
improved dimensional and thermal stability of the electrolyte. Despite this, the multifunctional
performance remains too poor for the requirements of the structural battery application demands.

8
2.2.3 Heterogeneous Phase-Separated Electrolytes
Heterogeneous electrolytes are a promising alternative whereby the benefits of homogenous electrolytes
are combined in a two-phase liquid-solid system. Herein, percolating structures are formed on a
sub-micron scale. The solid phase is composed of a porous polymer that provides necessary mechanical
load transfer, and is impregnated by an ion-conductive liquid electrolyte [10] [11] [23]. This concept is
visualized in Figure 4.

Figure 4. A schematic of the bicontinuous SBE concept [28].

2.5 Polymerization-Induced Phase Separation


Polymerization-induced phase separation (PIPS) is used when developing heterogenous electrolytes for
structural batteries. In this method, an initially homogenous multicomponent mixture is separated into
different phases following the polymerization reaction of one or several components [29]. Solubility
parameters vary as monomers transform into polymers, which is the key concept utilized in PIPS; initially
miscible polar monomers can transform into a fully immiscible polymer network. The monomer has polar
functional groups and thus dissolves in a polar solvent, whilst the formed polymer becomes non-polar and
reaches a gel point where a high degree of crosslinking makes it non-dissolvable in the solvent. This
aspect of PIPS enables phase separation through curing of thermosetting polymers. One of PIPS’ many
advantages is the possibility of forming in-situ filled porous polymer systems. Thus, a
microphase-separated electrolyte with a high modulus phase and ion-conducting phase can be produced,
as seen in Figure 5.

9
Figure 5. Image depicting how a phase separation of a homogenous mixture will proceed to produce a
microphase-separated electrolyte, with the high modulus phase and ion-conducting phase defined [10].

There are different ways of approaching a phase-separated electrolyte system, using different monomer
chemistries (including vinyl ester or epoxy chemicals for chain or step-wise polymerization, respectively)
along with various combinations of lithium salts and either conventional electrolyte solvents or ionic
liquids [16] [23].

Shirshova et al. first investigated a series of bicontinuous ionic liquid–epoxy resin systems, with the aim
of optimizing the multifunctional performance of structural supercapacitor electrolytes based on
phase-composition. Thermal curing was used in their experimental approach. The study yielded an
optimized electrolyte composed of 70 wt% ionic liquid, with an ionic conductivity of 0.8∙10-4 S cm−1 at
standard ambient temperature and a modulus of 200 MPa. Moreover, they found that morphology and
multifunctional performance was influenced by the epoxy resin and the volume/weight fraction of the
resin used [15].

A further study was conducted by Shirshova et al. to determine how these properties of epoxy-based
dual-phase structural electrolytes can be regulated by composition. The results showed that morphology
was impacted to a greater extent by content rather than the curing temperature; this was attributed to the
miscibility of components prior to and during the PIPS process. Salt concentration was determined to be a
deciding factor: Upon increasing lithium salt concentration in the ionic liquid, morphology became
macroscopically homogeneous. Moreover, modulus increased whilst ionic conductivity decreased,
showcasing a typically inverse relationship. This was attributed to a change in microstructure and
effective cross-linking, and an increased viscosity, respectively. [16]

10
In another study, Ihrner et al. researched the formulation of a novel SBE using reaction induced phase
separation . A radical chain-wise reaction mechanism and UV-curing was used in their experimental
approach to create a vinyl ester-based thermoset. The monomer found to yield the best results was
bisphenol A ethoxylate dimethacrylate (Monomer B) compared to bisphenol A dimethacrylate (Monomer
A) [23]. Monomer B is shown in Figure 6.

Figure 6. Chemical structure of monomer B, bisphenol A ethoxylate dimethacrylate.

Monomer B displayed a higher conversion (95%) than Monomer A (60%), due to the detrimental
vitrification effects that follows when polymerization occurs at ambient temperatures for a high Tg system
such as Monomer A. Moreover, monomer B displayed the smaller morphological domain size; this
difference can be attributed to the extra ethylene glycol units present that delay the phase separation,
creating network constraints. Combining Monomer B with a solvent mixture of EC:Dimethyl
methylphosphonate (DMMP) containing lithium tri-fluorouoromethanesulfonate (LiTFS) formed an SBE
with with an ionic conductivity of 2∙10-4 S cm-1, a modulus of 360 MPa at ambient temperature, and a Tg
of 72 °C. Contrary to Shirshova et al., this study concluded that ionic conductivity is not correlated with
phase separation size – however, the studies differ in system composition [23].

Schneider et al. expounded upon the work of Ihrner et al. by investigating the optimization of thermally
initiated PIPS to prepare a bicontinuous SBE. UV-curing is a limiting factor for the fabrication of
structural batteries, as the light cannot penetrate through a non-transparent multiple layered composites.
Instead, Schneider et al. found that thermally initiated PIPS presents a “robust and scalable process route
to fabricate structural batteries” [10]. The study looked at the impact of curing temperature with regards to
multifunctional performance of a Monomer B–based electrolyte system. The study yielded an SBE with

11
ionic conductivity of 1.96∙10-4 S cm-1 and a modulus between 530-540 MPa at ambient temperature when
polymerization occurred at an optimized temperature of 90 ℃, where the highest conversion (96 %) and
fastest curing time was achieved. Morphology was found to increase slightly compared to UV-curing,
such that the phase separation occurred of an increased domain size. The difference is speculated to be
connected to the polymerization or phase separation rate. Furthermore, the thermally initiated samples
exhibited improved mechanical properties at higher temperature. Upon testing the performance of a UD
lamina half-cell produced through thermal curing at 80 ℃ for 75 minutes, a capacity of 105-110 mA g-1 at
a current density of 18.6 mA g -1 was recorded [10].

Parameters influencing the phase separation are: relative solubility, temperature, kinetics, and reaction
mechanism [10]. As pertinent to this study, a chain-wise polymerization diverges from a step-wise. For
chain-wise polymerization, large chains form almost immediately while for step-wise, it forms towards
the end of the reaction [30]. This is visualized in Figure 7 and is explained by the differing propagation
step: For chain-growth, addition of monomers only occurs on the active site present on a growing polymer
molecule; whereas for step-growth, reaction occurs throughout the matrix such that high molecular weight
is achieved when oligomers react to form polymers [31]. This is expected to affect the phase separation
process, as the length of the chain influences solubility.

Figure 7. Chain-wise polymerisation vs. step-wise growth [31].

12
2.4 Thiol-ene Chemistry
From studies related to the topic, variables influencing both ionic conductivity and mechanical properties
have been investigated. The addition of copolymerization and cross-linking agents to the SBE polymer
matrix is a promising alternative. Notably, a study by Willgert et al. demonstrated that through addition of
small amounts of thiol to a poly(ethylene-oxide)-methacrylate SPE, the multifunctional performance of
the electrolyte was maintained - the ionic conductivity increased whilst the modulus was not detrimentally
affected [7] [32].

The utilization of the free radical thiol-ene reaction for thermoset polymers is of special interest, with
advantages including a step-growth nature, high yield, lack of side products, and a low responsiveness to
oxygen inhibition [33]. Moreover, it simultaneously aids in surface adherence of thiol-ene polymers [34].
It is speculated that such an effect would be beneficial for the interphase between electrode and electrolyte
in a structural battery.

The thiol-ene system presents a type of polymerization technique where the thiol compounds functional
groups can be used in different stoichiometric ratios to the alkene groups of the monomer to impact the
cross-linking density of the copolymerized system [33]. A polymer network consisting of thiol-ether
bonds is thus formed. Ribca et al. conducted a study on this topic, investigating how the properties of
lignin thermosets change as a function of crosslinker functionalities. Cross-linkers are a necessary
proponent of lignin thermosets, by combining monomer units into a polymer network; thus, a 1:1 ratio is
necessary for the thiol and allylic groups on a lignin molecule. The thiol cross-linkers differed only in the
degree of functionality, giving thermosets of comparable morphology, but contrasting storage modulus
and Tg. The chosen thiols used for this research report are the following, presented in Figure 8:
trimethylolpropane tris(3-mercaptopropionate) (3TMP) (Figure 8a), pentaerythritol
tetrakis(3-mercaptopropionate) (4PER) (Figure 8b), and dipentaerythritol hexakis-(3-mercaptopropionate)
(6DPER) (Figure 8c). [33]

13
Figure 8. Chemical structures of thiols 3TMP (A), 4PER (B), and 6DPER (C).

The reaction mechanism of a free-radical thiol-ene reaction is propagated through a step-growth reaction
as is seen in Figure 9 . During the initiation step, the thiol group is deprotonated to create a free radical.
The double bond of the allyl group will then react with the free radical of the thiol group in the
propagation step. The reaction repeats through chain transfer, and is terminated through the combination
of radicals present on oligomers [35]. Observe: If a different monomer is present in reaction mixture,
homopolymerization will still occur for the competing monomer to an extent dependent on the
concentration of thiol groups present.

Figure 9. Schematic illustration of a general thiol-ene reaction mechanism [35].

14
3. Aim
To ensure that results are comparable to industry standards and thus reflect the applicability of the SBE, a
solvent consisting of PC and EC was used for this study. These are widely utilized within battery
manufacturing due to the carbonate group (-O-(C=O)-O-) possessing a high dielectric constant, which
promotes a high salt dissociation such that charge carrier concentration and ion mobility is increased [19].
Moreover, DMMP is highly flammable and a health hazard; it is thus also beneficial from a green
chemistry perspective to replace it with PC, which is merely an irritant [36] [37].

The current research project is concerned with the optimization of thermomechanical properties and ionic
conductivity; it expounds upon the phase-separated SBE-system developed by Schneider et al. and Ihrner
et al. to improve battery performance and meet the structural battery application demands. The aim of this
study is threefold. During the first phase, the functionality and effect of differing ratios of thiol
functionalities in the cross-linking agent relative allyl groups present was evaluated. Homogeneity was
investigated upon the addition of a step-growth mechanism provided by the thiol-ene reaction. In
addition, the impact of an EC:PC solvent on the electrolyte system developed by Ihrnere et al. and
Schneider et al. was examined. The second phase investigated the optimum weight percentage of the
ionically conductive liquid electrolyte for the developed SBE from phase one. Analysis was performed
through assessment of conductivity, thermomechanical stability, and the relationship therein. In addition,
structure-property relationships were established by investigation of morphology The third phase
concluded the study, wherein a CFC half-cell was assembled with the optimized SBE formulation
incorporated. Testing was subsequently performed to determine the practical electrochemical performance
of the SBE developed throughout this study.

15
4. Methodology

4.1 Materials Used and Their Origin


Monomer Bisphenol A ethoxylate dimethacrylate (Monomer B, Mn; 540 g mol-1) was donated by
Sartomer. The thermal initiator 2,2′-azobis(2-methylpropionitrile) (AIBN) was purchased from Sigma
Aldrich and used as received. Three cross-linking thiols with different amounts of functional groups (3, 4,
and 6) was used in this research. 3TMP and 4PER were, like the monomers, donated by Bruno Bock
Chemische Fabrik GmbH. 6DPER was purchased from Fujifilm. PC, EC and the electrolyte salt lithium
trifluoromethanesulfonate (LiTFS) were purchased from Sigma Aldrich.

4.2 Techniques and Procedures


During the first phase, three ratios for each thiol were prepared based on the different molar ratios
between the thiol and alkene functional groups present in the monomer; giving a total of nine sample
groups. The molar ratios for thiol:ene were fixed at: 0.1:1, 0.05:1, and 0.2:1. Molar mass and number of
functional groups are specified in Table A1 in Appendix A. The amount of liquid electrolyte was held
constant, corresponding to 39 wt% of the ingoing compounds and composed of a EC:PC solvent
containing LiTFS. The corresponding weight for the thiols and monomer are specified in Table A2 along
with the exact experimental measurements in Table A3 in Appendix A. Initiator and liquid electrolyte was
kept constant at 0.03 g and 1.95 g, or 1 wt% and 39 wt%, respectively for each reaction mixture
throughout the experimental phase. A reference sample without any cross-linking agents was prepared,
containing: Monomer B + Initiator + Liquid Electrolyte.

In phase two, one of the thiols was included for further experimentation following analysis of the physical
and electrochemical properties belonging to the different sample groups. 6DPER was chosen due to
showing more desirable results (see section 4.1.1). In this phase, the electrolyte concentration was
increased for two chosen thiol ratios (0.05 and 0.1). The electrolyte concentrations were increased to 45
and 50 wt% for both thiol functional group ratios (FG ratio), giving a total of four additional sample
groups. See Table A4 in the Appendix A for exact experimental measurements. The same analysis
procedures were performed for phase two as for phase one to allow comparison.

16
During phase three, a structural battery half-cell prototype was manufactured based on what was deemed
as the best performing electrolyte concentration from phase 2. The SBE consisted of Monomer B and
6DPER at an FG ratio of 0.05, with the administered liquid electrolyte at a 45 wt%. The electrochemical
performance of the half-cells were analysed via galvanostatic cycling.

4.2.1 Structural Battery Electrolyte Preparation


A stock solution of electrolyte liquid was prepared and used for all samples containing 1.8 M LiTFS in
1:1 wt ratio of the EC:PC solvent. The thermal initiator AIBN and the electrolyte solution were mixed until
the initiator was dissolved. The thiolic SBE solutions were obtained by mixing thiol cross-linkers,
monomer B, and the liquid electrolyte-AIBN mixture. Reference samples were prepared likewise
containing only monomer B. A magnet was added to the SBE solution and stirred using a magnetic stirrer
until a homogenous mixture was achieved. Remaining air bubbles were then removed. The solution was
transferred to a teflon-sprayed aluminum mold using a syringe (32 × 6 × 0.5 mm3). Simultaneously, a
glass disc was slid onto the mold. The aluminum mold with a glass disc was clamped on both sides to
prevent bubble formation. The samples were transferred to a preheated oven and thermally cured at 90°C
for 45 minutes. Observe that this procedure was performed in a fume hood; typically, it is done in an
argon-filled environment, as found in a glovebox, to hinder oxygen and moisture interference that may
otherwise negatively impact electrochemical performance. This introduces a source of error to our
subsequent measurements, discussed in section 7.

4.2.2 Electrochemical Impedance Spectroscopy


Electrochemical impedance spectroscopy (EIS) was used to determine the ionic conductivity of SBE films
at ambient temperature. The samples were measured directly after curing using Gamry Series G 750
potentiostat with gold wire as a four-point electrode. The impedance was measured within a frequency
interval 120 kHz to 1 Hz and an amplitude of 10mV. The conductivity was calculated using equation 3.

𝑙
σ= 𝑅𝑏·𝐴
(3)

Here, l (m) is the length between reference electrodes (55 mm), Rb (Ω) is the bulk resistance, and A ([2])
is the area of the sample, cross-sectioned. [7] σ is the conductivity, and A the cross-section area of the film

17
(the thickness of each sample was measured using a digital slide caliper). Rb is the bulk resistance, derived
from the Nyquist plot generated during EIS by extracting the low-frequency intercept with the real axis
[6].

4.2.3 Scanning Electron Microscopy & Gravimetric Analysis


The microstructure of the cured SBE films were investigated using scanning electron microscopy (SEM),
Hitachi S-4800 equipped with a cold field-emission electron source. The cross-sections of the films were
scanned for analysis. Beforehand, the samples were submerged in a water bath for 48h to extract the
liquid electrolyte. The baths were replaced with distilled water 3 times during extraction to aid the
concentration gradient. Then, samples were dried in a vacuum oven for 24h at a temperature of 50°C. To
perform gravimetric analysis, samples were weighed before putting them in the water bath and once more
after drying them. The samples were later freeze-fractured into smaller fragments, after submersion in
liquid nitrogen so as to minimize artifacts. The fragments were then mounted upon a conductive carbon
tape. The samples were sputtered with a Pt/Pd coating, using a Cressington sputter coater 208 HR to
ensure conductivity, before analysis with SEM commenced.

4.2.4 Fourier-Transform Infrared Spectroscopy


To study the final double-bond conversion, Fourier-Transform Infrared Spectroscopy (FTIR) was used on
the samples, before and after curing. The instrument used for both methods was a Perkin Elmer Spectrum
100, equipped with a triglycine sulfate (TGS) detector with a single reflection attenuated total reflection
(ATR) accessory unit. Contained was a diamond ATR crystal (Golden gate) with temperature control from
Graseby Specac Ltd. The software used during measurement was Spectrum software v. 10.5.1, also by
Perkin Elmer.

Spectra for both uncured and cured samples were subsequently produced in MATLAB. The conversion of
the acrylate groups was determined by qualitatively comparing the area under the vinyl peak at 1637 cm-1.
To determine the amount of electrolyte present after the drying process, spectra for undried and dried
samples were compared. The peaks that are uniquely inherent to the organic electrolyte components were
identified by studying their individual spectra [38] [39] [40] [41].

18
4.2.5 Dynamic Mechanical Analysis
Dynamic Mechanical Analysis (DMA) was used to investigate the mechanical properties of the SBE
samples. For the DMA analysis, a DMA Q800 V20.24 Build 43 was used. This was equipped with an air
cooling system and tensile film clamps. The thermally cured SBE films were clamped into the sample
holder. In order to allow the samples to equilibrate, the samples were kept in an ambient temperature
between 20° C to 200° C for 1 hour and were heated at a rate of 3° C per minute. The amplitude applied
was set at 10-15 µm. The T g was determined from maximum tan𝛿 signal.

4.2.6 Half-Cell Preparation


The half-cell for the third phase was prepared using a similar procedure as Schneider et al [6]. Carbon
fibre was used as the cathode material. The SBE resin was prepared in the same manner as section 3.2.1,
with the key difference that the electrolyte and initiator was prepared separately inside of a glove box in
an argon-filled environment. The monomer thiol mixture was later transferred to the glove box and the
two mixtures combined, and then stirred to homogeneity using a vortex mixer. The new SBE resin was
sealed with a septum and brought outside for assembly into the CFC, using vacuum infusing. The
composite was subsequently placed in a 90° C oven for 45 minutes to thermally cure. The CFC with the
thiol cross-linked matrix thermoset was then again transferred to the glove box and removed from the
vacuum package and cut into two separate electrodes. The composites were placed into their individual
pouch cells. The pouch cell assembly used a lithium foil as the counter electrode, copper and nickel foil as
the current collector and a glass microfiber placed between the two electrodes as the separator. The
structural electrode was also wetted with additional droplets of liquid electrolyte to ensure ionic
conductivity. The two pouch cells were later tested for their open circuit voltage potential and
electrochemically tested in the same manner as Schneider et al. The current densities were the following:
18.6 mA g-1 over cycles 0-10, 37.2 mA g-1 over cycles 10-15, 74.4 mA g-1 over cycles 15-35, and 18.6 mA
g-1 over cycles 35-40.

19
5. Results & Discussion

5.1 Phase 1: The effect of thiol functionality on electrolyte properties

Following are the results from investigating electrolyte performance with regards to the different thiol
additives, corresponding to phase 1one of the study. For each analysis, 3TMP, 4PER and 6DPER were
examined and compared in relation to the functional groups (FG) ratios of 1:0.05, 1:0.1 and 1:0.2. The
electrolyte concentration was fixed to 39 wt%.

5.1.1 Thermal & Mechanical Data

DMA curves, comparing the reference and different thiols for each FG ratio, are presented in Figure
10-11. They depict storage modulus and Tg as functions of temperature. Storage modulus and Tg were
subsequently extracted from the DMA data, compiled in Table B1 in Appendix B.

a) b) c)

Figure 10. DMA data depicting storage modulus (E’) [MPa] and tan𝛿 vs. temperature [℃] for thiol samples
containing different FG ratios (0.05, 0.1, 0.2), in consideration of thiol (a) 3TMP, (b) 4PER and (c) 6DPER.
Electrolyte concentration of 39 wt%.

20
a) b) c)

Figure 11. DMA data depicting storage modulus (E’) [MPa] and tan𝛿 vs. temperature [℃] for thiol samples, in
consideration of ratios (a) 0.05, (b) 0.1 and (c) 0.2. Electrolyte concentration of 39 wt%.

When comparing storage modulus between different thiols in Figure 10-11, no distinct difference is
noticeable; modulus between different thiols is almost equivalent at the same FG ratio - despite the
substantial change in functionality.

For 3TMP and 4PER the overall trend observed is that the modulus decreases with increasing FG ratio,
from values around 800 MPa at 0.05 to around 600 MPa at the 0.2 ratio. A possible explanation for the
decreasing modulus upon increasing the FG ratio is the fact that the stiff monomer is being replaced by
the more flexible thiols, thus giving rise to a less stiff polymer. For 6DPER, however, the modulus
decreases at 0.1 ratio and then increases to max at 0.2, presenting a clear deviation from remaining
samples.

When comparing results to the reference sample, it becomes evident that the storage modulus for both
0.05 and 0.1 FG ratios within all thiols exceed the reference modulus, whereas the 0.2 ratio generates
lower modulus. In reference to the same monomer system performed by Schneider et al. where the
modulus reached a value of 530-540 MPa at ambient temperature [10], all thiol-systems- as well as the
reference sample- have managed to improve upon modulus capacity. The larger storage modulus achieved
by all copolymers may partly be due to the achievement of a more homogeneous network, as gradients in
the network would detriment the mechanical performance rather than improve it. It can thus be

21
ascertained that the homogeneity of the network is not being compromised when integrating thiols into
the system, but rather improved. When analyzing the sample morphology (see SEM Figure 16-17), all
samples indeed appear to have become homogeneous. The degree of phase separation will further
corroborate this observation. Moreover, the degree of effective cross-linking may be correlated with the
number of functionalities per molecule in the thiol, explaining why 6DPER has the highest modulus;
6DPER possesses the most number of functional groups, implicating more cross-linkage per molecule.
The difference in the storage modulus that occurs in the higher temperature range between thiol additives
supports this theory, as 6DPER demonstrates the highest modulus across the entire temperature interval.
An important aspect to keep in mind when analyzing and observing trends with modulus however, is the
fact that both the polymer network and the phase separation is affecting it which makes it challenging to
draw conclusions solely from modulus data.

Regarding Tg analysis, there are no apparent trends between nor within thiols. Between 3TMP and 4PER
Tg increases with increasing functionality, where 3TMP exhibits a Tg of 159 ℃ compared to 4PER of 164
℃ at FG ratio 0.05, and 80 ℃ and 109 ℃ respectively at a 0.2 ratio. For 6DPER however, the results of
Tg do not follow the same trend (Figure 10-11). Why Tg increases slightly with increasing functionality
across the different thiol’s FG ratio can be related to the number of thiol molecules in each network; due
to the different number of FG within each thiol molecule, there are less molecules needed in, for example,
the 4PER network than in the 3TMP to reach the same FG ratio. As a result, the 4PER network with its
fewer number of thiol molecules is expected to (provided that the conversion is completed) have a tighter
network than 3TMP, thus possessing a higher Tg.

Furthermore, Tg for the reference sample is higher than all thiol samples, at 182 ℃. Generally, Tg
decreases with increasing FG ratio within the same thiol samples (Figure 10), which is expected from
what is indicated in the theory; the Tg reduction arising from adding thiols to the network is related to the
reduced relative amount of monomer B. The original monomer creates a stiff network with its bulky and
immobile benzene groups (see Figure 5), thus giving rise to a relatively high Tg. When replacing these
original monomers with thiols that provide flexibility to the polymer chains, the network is softened and
Tg decreases. Another clear difference between the networks with and without the added copolymer is the
fact that the reference system has a wider Tg transition (tan𝛿 peak) whereas the peaks in the thiol-systems
are more well defined (Figure 10-11). Given that the cross-linked thermoset is vastly heterogeneous and

22
bicontinuous - composed of both a liquid phase and a stiff polymer phase - the system will not behave as
a pure polymer, explaining the wide Tg transition of the reference. The thiol additive on the other hand
contributes to a more homogeneous polymer network relative to the reference, giving rise to more distinct
peaks. The homogeneity could be explained by the phase separation, where the polymerization
mechanism for the reaction is key. As opposed to the chain growth of Monomer B, the step-wise growth
that the thiols contribute appears to lead to a more homogeneous network and phase separation. This
would be due to the phase separation occurring at a later stage, restricting molecular mobility due to the
network formation imposing constraints. This may further indicate that the increase in thiol ratio would
increase homogeneity of the polymer network. However, the corresponding modulus appears to decrease.
This may be due to the effect of adding more thiol-bonds, which may overcome the benefits imposed by
the more homogenous system.

Based on solely the mechanical and thermal properties of these thiol-networks, all are considered
promising crosslinkers for increasing the mechanical performance of the electrolyte - outperforming the
reference sample in regard to E’ modulus, though not Tg. As Tg occurs well above the ambient
temperature relevant to this study, this is not a point of contention.

5.1.2 Conductivity

Results from EIS analysis are introduced in Table 1. Included are the measured ionic conductivities and
standard deviations. Apparent is that all samples, including the reference, exhibit lower ionic
conductivities than those presented in previous studies; the conductivity of 2∙10-4 S cm-1 demonstrated by
Ihrner et al. is over one order of magnitude greater than the samples in phase 1 [23]. It can be argued that
the deviations partly derive from not having performed the sample preparation under sterile conditions (i.e
outside the glove box), see section 7.

The main reason, however, may be due to the use of a different electrolyte mixture in this study; instead
of PC solvent, Ihrner et al. and Schneider et al. utilized DMMP. This may have a significant impact on the
phase separation as well as how the percolated system is formed. As DMMP is more polar than PC, the
phase separation is expected to occur at an earlier stage, reaching a percolation threshold with high
conductivity as opposed to PC where this balance is altered and a lower conductivity is obtained.

23
Table 1. Ionic conductivities [S cm-1] at ambient temperature for 3TMP, 4PER and 6DPER additives varying the FG
ratios. Electrolyte concentration at 39 wt%.

Thiol FG Ratio Ionic Conductivity at 25 ℃ [S cm-1]

0.05 9.44∙10-6 ± 7.8%

0.1 1.57∙10-5 ± 3.1%


3TMP

0.2 6.52∙10-6 ± 9.3%

0.05 1.19∙10-5 ± 10.6%

0.1 1.71∙10-5 ± 1.4%


4PER

0.2 9.58∙10-6 ± 1.2%

0.05 1.72∙10-5 ± 15.6%

0.1 2.41∙10-5 ± 4.6%


6DPER

0.2 2.86∙10-6 ± 0.0

Ref. - 1.66∙10-5 ±1.4%

Table 1 indicates that conductivity increases from FG ratio 0.05 to 0.1, and then decreases at 0.2 ratio. For
FG ratio 0.05 and 0.1 conductivity seem to increase with the number of functional groups, whereas for 0.2
there are no clear trends discernible. Compared with the reference conductivity of 1.66∙10-5 S cm-1, 4PER
at 0.1 ratio and 6DPER at 0.05 and 0.1 ratio all exhibit slightly higher conductivities, suggesting potential
promise for further studies. However, 3TMP did not show conductivities favorable for its intended
purpose as they were all lower than the reference.

A considerable change in conductivity is not present following addition of thiols, which can be related
back to the phase separation as it indicates that the extent of phase separation is comparable between the

24
systems. 6DPER, however, appears to outperform the other thiol groups in regard to ionic conductivity. A
possible explanation is due to the phase separation occurring earlier; 6DPER may be expected to
cross-link faster than remaining thiols due to it having more functionalities per molecule. Additionally,
the homogeneity of the system - discussed in the thermal and mechanical section 5.1.1 - is speculated to
have a considerable impact on conductivity; as previously observed, the phase separation would
theoretically become more homogeneous with increasing FG ratio, thus potentially explaining the
different conductivities. This theory is further discussed upon analyzing morphology.

5.1.3 Multifunctional Analysis

In Figure 12 the relationship between storage modulus and conductivity for each thiol is illustrated. As
clearly demonstrated by both Figure 12.a and 12.b, the storage modulus is considerably increased by the
addition of thiol cross-linkers whilst the ionic conductivity is kept in the same range as the reference - if
not lower. Striking is the result for 6DPER with FG ratio 0.1, with both considerably high conductivity
and storage modulus relative to the reference (Figure 12.b). Based on conductivity and
thermo-mechanical analysis, 6DPER at 0.1 and 0.05 ratio is the thiol cross-linker which displays the most
promising properties for use in a structural battery application.

a) b)
Figure 12. Ionic conductivity [S cm-1] vs. storage modulus [MPa] for 3TMP, 4PER and 6DPER additives with error
bars included. FG ratio is (a) 0.05 and (b) 0.1 with 39 wt% electrolyte.

25
As discussed in previous sections, a comparison between the reference sample to both Ihrner and
Schneider et al., modulus is improved by 200-300 MPa. The increased modulus can be related back to the
difference in phase separation between the systems; the PC solvent expectedly causes the phase
separation to occur later than for DMMP, giving rise to a more homogeneous system. The mechanical
analysis further demonstrated that a later phase separation appears to benefit homogeneity of the polymer
network. Concurrently, it is speculated that an earlier phase separation benefits conductivity; explaining
the decreased conductivity relative to Schneider et al. and Irhner et al.. This demonstrates a possible
antagonistic relationship between the two properties.

There are no clear trends between increasing functionality and SBE performance, as clearly illustrated in
Figure 12a. Worth bearing in mind is the fact that these percolated systems are complex; several
parameters, such as the electrolyte liquid and compounds, the monomer, and the thiol structure, are all
contributing to the network functionality in different ways - sometimes conflicting each other in ways that
prevent solid conclusions from being drawn.

5.1.4 Curing Performance


FTIR analysis was performed on both the reaction mixtures prior to curing and on the thermoset films.
The differences in double bond conversions between 3TMP, 4PER and 6DPER were plotted and
compared in MATLAB. The disappearance of the vinyl peak at 1637 cm-1 is visualized in Figure 13,
signifying a high conversion rate for all samples. In terms of curing behaviour, the conversion appears
qualitatively similar to the results of Schneider et al. where a 96 % conversion is achieved. This indicates
that the addition of thiol cross-linkers does not negatively impact curing, and that the curing performance
remains ideal for a structural battery application.

26
a) b) c)
-1
Figure 13. FTIR spectra depicting the vinyl peak at 1637 cm for uncured reaction mixtures relative to the cured
samples of different FG ratios, for a) 3TMP; b) 4PER; c) 6DPER.

The high degree of conversion of all samples lends credence to the theory regarding the effect of
increasing functionality on network compactness and Tg value, as mentioned in section 5.1.1.

Further analysis was performed to determine the peaks which are unique to the organic liquid electrolytes;
these are marked in Figure 14, which present the FTIR-spectra of a sample post-leaching compared to one
undried of the same concentration. As is illustrated, these peaks have been significantly reduced following
the drying procedure, indicating that liquid electrolyte has been effectively removed from the thermoset
within the sample area investigated. This signifies that the network is mostly percolated and the liquid
electrolyte can move freely through the matrix in this region.

27
Figure 14. FTIR spectra of a dried and undried sample 6DPER with a thiol ratio of 0.05, with peaks corresponding
to organic liquid electrolyte components marked.

5.1.5 Morphology
The results from the gravimetric analysis are available in Table 2 and illustrate a sample weight loss that
varies between 34 - 38 % among the different thiols and FG ratios.

As there are no apparent trends evident between thiols or FG ratios in regard to weight loss, no specific
conclusions can be drawn regarding implications of thiol additives to the network. Nonetheless, the
results indicate that a phase-separated, percolated, structure has been obtained in all samples to the same
extent- even though the gravimetric analysis indicates that isolated electrolyte compounds and solvent are
still present in the structure. This is especially notable in the sample with 6DPER at a ratio of 0.2, with
c.a. 38 % deviation from the weight of liquid electrolyte in the polymer thermoset. The minimal weight
loss can signify an incomplete percolating structure, which may also explain the low Tg value as
plasticing liquid electrolyte remains enclosed in the system. The reduction in porosity might inhibit ionic
mobility in the liquid phase, furthermore explaining the low ion-conductivity of this sample. Compared
with Schneider et al. the results are noticeably smaller, which display a weight loss close to 39 wt%
indicating full removal of the liquid electrolyte a more open morphology.

28
Table 2. Percentual weight loss from gravimetric analysis measured by FG ratio, with percental deviation from wt%
liquid electrolyte originally present.

Thiol Ratio of FG Sample weight loss [%] %-Deviation

0.05 35.31 9.47

3TMP 0.1 38.77 0.60

0.2 35.11 9.98

0.05 37.37 4.18

4PER 0.1 34.19 12.3

0.2 36.07 7.50

0.05 35.57 8.80

6DPER 0.1 35.24 9.63

0.2 24.24 37.9

Notably, this is contradictory to the FTIR analysis which indicates that the system is percolated and all
liquid has been removed. As FTIR only analyzes a small part of the sample, the gravimetric analysis does
indeed show that the system might still have closed pores that hinder leaching of all the liquid electrolyte.

SEM images comparing the electrolyte surfaces for the reference sample as well as thiols with different
FG ratios are presented in Figure 15-17.

Figure 15. SEM image of the reference sample in 5k and 35k resolution respectively.

29
Figure 16. SEM images in 5k resolution of electrolyte surface with (from left to right) 3TMP, 4PER and 6DPER
additives. FG ratio 1:0.05 and 39 wt% electrolyte.

Figure 17. SEM images in 35k resolution of electrolyte surface with (from left to right) 3TMP, 4PER and 6DPER
additives. FG ratio 1:0.05 and 39 wt% electrolyte.

The morphology of the different thiol additives at a ratio of 0.05 appear to be comparable, and no
considerable difference can be discerned from the SEM images displayed in Figure 16-17. Neither can a
change in homogeneity be determined. However, the divergence to the reference can be seen: a porous
structure is evident whereas the thiol copolymerized samples appear more dense and homogenous, with
smaller pores. This implies that a thiol cross-linked polymer structure creates a matrix with a more closed
morphology, which may be connected to the phase separation occurring at a later stage during the curing
process due to the incorporation of the slow step-wise growth. The conjoint increase in molecular weight
for all chains present in the network may imply that the phase separation is instantaneous, occurring
simultaneously for the entire network. This differs from the reference sample, which propagates only
through a chain-wise polymerisation; such that parts of the polymer undergo phase separation at different
points in time. When comparing the reference to the SEM images obtained by Ihrner et al., the
morphology is more closed and appears less porous; this indicates that a later phase separation , occurring

30
due to the changed polarity in the system stemming from the DMMP, generates a less percolated matrix.
This lends support to the above mentioned theory. Moreover, the lowered conductivity can thus be
explained by changed morphology. The increased tortuosity stemming from a decrease in porosity would
indeed be expected to limit the mobility of the charge-carriers navigating through the structure. However,
a more homogenous phase separation may concurrently result in a more uniform ion transport through the
network; this would benefit conductivity, and may explain why the thiol-copolymers at ratio 0.05 and 0.1
exhibit a conductivity that is equal or larger than the reference, despite demonstrating smaller pores and a
more closed morphology. This suggests that homogeneity both benefits and hinders conductivity.

Another contributing factor could be the polarity present in the polymer chain. Irhner et al. observed that
the monomer with less polarity (Monomer A) showed a significantly more open morphology in the SEM
analysis compared with the slightly more polar Monomer B [23]. Similarly, the more non-polar reference
sample is less closed off than the thiol copolymerized systems, where polar sulphur-atoms have been
added. It should be noted that analysis by SEM is hindered by the manipulation of the samples as required
by the method, which is detailed in section 7.

Pictures of the electrolyte films for all samples in phase 1 are presented in Figure 18-20.

Figure 18. Images of electrolyte films with increasing thiol FG ratio (3TMP): 0.05 (a), 0.1 (b), 0.2 (c).

31
Figure 19. Images of electrolyte films with increasing thiol FG ratio (4PER): 0.05 (a), 0.1 (b), 0.2 (c).

Figure 20. Images of electrolyte films with reference and increasing thiol FG ratio of (6DPER): Reference (a), 0.05
(b), 0.1 (c), 0.2 (d).

When examining the sample films shown in Figure 20, the reference displays a more opaque appearance
compared to the samples with added thiols, which became more transparent the higher the thiol ratio. This
trend is observed for all thiol functionalities in Figures 18-20. This difference can be linked to the smaller
domain size in the samples with higher thiol ratio that results in greater transparency. This is explained by
Ihrner et al., who conclude that a larger domain size leads to more opacity in the samples due to light’s
increased ability to scatter incoherently. [23] Thus, a transparent sample would suggest a more closed
morphology and smaller pores. This may further prove that slower phase separation gives way to a more
homogenous system that allows for a uniform distribution of light. Taking account of the pictures from
the SEM analysis, this seems to be the case as the percolated structure is more visible in the reference
sample than it is for the samples containing thiols. Furthermore, 6DPER appears more transparent at all

32
ratios in comparison with 3TMP and 4PER, indicating greater homogeneity - this may explain why it
outperforms in regard to conductivity at ratios 0.05 and 0.1, lending credence to the theory of
homogeneity impacting uniform ion mobility as theorized above. That the greatest modulus is achieved
by the 6DPER at the 0.2 ratio due to greater homogeneity is also substantiated, as it appears significantly
more see-through compared to 3TMP and 4PER. Here, however, the increased homogeneity has perhaps
also constricted ion mobility too much as demonstrated by the gravimetric analysis; suggesting again that
conductivity is connected to morphology.

5.2 Phase 2: Effect of Liquid Electrolyte Concentration


In the second phase of the present study, the influence of liquid electrolyte concentration (39, 45, and 50
wt%) on mechanical and electrochemical properties of the 6DPER copolymerized SBE was investigated.
A FG ratio of 0.05 and 0.1 is considered, based on the results of the previous investigative phase.

5.2.1 Thermal & Mechanical Data


Following are the results derived from DMA. Storage modulus and tan𝛿 as a function of temperature is
visualized in Figure 21, allowing juxtaposition of the different samples. Tg and storage modulus at
ambient temperature, as is pertinent to this study, has been extracted from the data and is available in
Table B2 in Appendix B. In comparison to the reference sample with a modulus of 701 MPa and Tg at 182
℃, the samples demonstrate values both above and below the reference when increasing the wt%; with
FG ratio 0.05 the modulus decreases from 823 MPa at 39 wt% to 462 MPa at 50 wt%, whereas for FG
ratio 0.1 the modulus decreases from 774 MPa at 39 wt% to 440 MPa at 50 wt%. Tg increases from 175
℃ to over 191 ℃ with increasing wt% for 0.05, and from 152 to 187 ℃ for FG ratio 0.1.

33
a) b)

Figure 21. DMA data depicting storage modulus (E’) and tan𝛿 vs. temperature for samples containing different wt%
liquid electrolyte (39, 45, 50), in consideration of FG ratios at a) 0.05 and b) 0.1.

The following trends are observed: storage modulus decreases with increasing wt% liquid electrolyte,
whilst Tg region appears to rise and widen. A plasticizing impact of the liquid electrolyte on the polymer
matrix is evident, as the mechanical performance of the SBE suffers upon increasing wt% liquid
electrolyte. Upon increasing the wt% liquid electrolyte, the ratio between the volume of solid and liquid is
amended too, however the change in modulus is notably larger than this amendment which indicates other
factors are at play. This may be attributed to the phase separation occurring at an earlier stage; by
increasing the electrolyte content relative to monomer and thiol, the polarity in the solution may
concurrently increase due to the greater presence of polar solvent molecules. Homogeneity may also thus
decrease, as previously discussed. This is partly visible in the broader Tg curves, which exhibit greater
similarity to the reference.

Morphology results will further confirm these trends; the more open morphology, the more likely is the
liquid electrolyte to percolate in larger pores - leading to the decrease in mechanical properties (see Figure
15-17). However, the increase in Tg with electrolyte concentration for the 0.05 and 0.1 FG ratio samples is

34
unexpected; as the flexibility of the polymer matrix increases due to the plasticisation, segmental chain
movement should begin at lower temperatures.

5.2.2 Conductivity
The measured ionic conductivities of relevant samples are displayed in Table 3, alongside their standard
deviations. As evident, increasing wt% liquid electrolyte to 45% results in a concurrent increase of ionic
conductivity of one order of magnitude in reference to samples in phase 1. An additional increment-
however not as substantial- is observed when increasing further to 50 wt%. Consequently, upon
increasing wt% liquid electrolyte the ionic conductivity has managed to improve to 2.58∙10-4 and 2.95∙10-4
S cm-1 for FG ratio 0.05 and 0.1 respectively. These are promising results as they are now comparable to
what both Schneider and Ihrner et al. obtained (1.96 and 2.0∙10-4 S cm-1 , respectively) [10][23].

Table 3: Ion conductivity measured by FG ratio and wt% liquid electrolyte for 6DPER.

wt% Ratio of FG Ionic Conductivity at 25 ℃ [S cm-1]

0.05 1.72∙10-5 ± 15.6%


39
0.1 2.41∙10-5 ± 4.6%

0.05 1.35∙10-4 ± 5.0%


45
0.1 1.22∙10-4 ± 7.1%

0.05 2.58∙10-4 ± 1.2%


50
0.1 2.95∙10-4 ± 4.4%

Conductivity increases concurrently with the amount of liquid electrolyte incorporated into the network,
due to the greater presence of charge carrier species. The morphology of the pores further promotes the
mobility of these carriers- increasing charge flow- which is further supported from the gravimetric
analysis (morphology section 5.2.5) implying the achievement of a more percolated network and an
earlier phase separation . As evident from Table 3, the improvement in ionic conductivity from 45 to 50
wt% is not as substantial as the one from 39 to 45 wt%. This could indicate that the sample composition is
approaching the percolation threshold of the system as discussed by Ihrner et al [23].

35
5.2.3 Multifunctional Analysis
Electrochemical and mechanical performance of the SBE’s is juxtaposed in Figure 22. A near inverse
relationship is displayed; modulus decreases with increased liquid electrolyte incorporated into the
polymer matrix, whilst conductivity is augmented. Relative to the reference samples, increasing wt%
electrolyte liquid detrimentally impacts mechanical performance and improves the electrochemical
properties: the multifunctional performance achieved supports that a completed phase separation is
reached. Notably, the electrochemical performance is greatly enhanced yet simultaneously maintains a
storage modulus above 500 MPa as desired for a structural battery [1]. This makes it a viable alternative
for further research, as the SBE formulation is more applicable from an industrial perspective.

a) b)
Figure 22. Ionic conductivity (logarithmic scale) vs. Storage modulus at different wt%, including reference, for a)
0.05 and b) 0.1 FG ratios.

Notably, first when increasing the wt% of electrolyte in phase 2, the conductivity is approaching values
closer to Ihrner et al. For Ihrner et al. the conductivity is more or less stable at 1.1-2.1∙10-4 S cm-1 when
altering the choice of monomer structures (Monomer B vs Monomer A), whereas the modulus varies
broadly between 360 and 750 MPa [23]. In the case of this study however, both the mechanical properties
and ionic conductivity are affected. This is further indicative of a morphology dependence, where the
mobility of ions appears to be affected by the morphology and phase separation.

36
5.2.4 Curing Performance
FTIR analysis was performed on both the reaction mixtures prior to curing and on the thermoset films.
The disappearance of the vinyl peak is visualized in Figure 23; it is reduced significantly, implying that a
high double-bond conversion has been achieved, similar to the results of phase 1. This further indicates
that increasing the electrolyte wt% content of the un-cured resin samples within the given bounds does
not negatively impact the curing efficiency.

Figure 23. FTIR spectra of reaction mixture containing a ratio of 0.1 6DPER and 39 wt% electrolyte liquid, in
comparison to cured films containing 45 and 50 wt% at the same thiol FG ratio.

In addition, the leaching of the organic electrolytes appears to be successful in the region studied; the
corresponding peaks are no longer visible, as shown in Figure 24 where the spectra of dried and undried
samples are compared. The limitations of the FTIR are however still to be considered, as discussed in
phase 1.

37
Figure 24. FTIR spectra of a dried sample with a thiol ratio of 0.05 and previously composed of 50 wt%; and
samples containing 45 and 50 wt% liquid electrolyte, at a respective thiol ratio of 0.1 and 0.05.

5.2.5 Morphology

The average weight loss following a gravimetric analysis is presented in Table 4; included is the percent
deviation from the wt% electrolyte initially introduced to the samples. This data indicates that whilst a
phase-separated, percolating structure has been achieved, the electrolyte analyte was not entirely removed
during the leaching process. However, compared to the gravimetric results from phase 1 it seems like
there are less closed pores in the samples with higher electrolyte content, which has allowed more
electrolyte to be removed than during phase 1. This can be seen as the percentile deviation is not as high
with higher electrolyte content than they are with different thiol ratio.

38
Table 4: Percentual weight loss from gravimetric analysis measured by FG ratio, with percentile deviation from
original wt% liquid electrolyte present for 6DPER.

wt% Ratio of FG Weight loss (%) %-Deviation

0.05 35.57 8.80


39
0.1 35.24 9.63

0.05 42.79 4.91


45
0.1 43.07 4.29

0.05 46.59 6.82


50
0.1 47.46 5.07

SEM images were procured for samples with a functional group ratio of 0.05 and 0.1; at 40 and 50 wt%
for the latter, and at 45 wt% for the former. These are displayed below in Figure 25. For additional SEM
images, see Appendix C.

Figure 25. SEM images of 6DPER thiol 0.05 ratio in 35k resolution of electrolyte surface. From left to right: 39
wt% electrolyte, 45 wt%, 50 wt%.

As evidenced by the images above, a porous percolating network is inherent to the morphology of all
samples. Furthermore, a relatively homogenous structure is visible. Upon comparison with SEM images
of samples previously containing electrolyte liquid at 39 wt%, available in Figure 16-17, it appears that
phase separation has increased in domain size and the morphology has become more open. This may be a

39
consequence of the increased electrolyte content that has had an impact on the phase separation rate.
Comparing SEM images between different electrolyte contents in Figure 25, 50 wt% indeed seem slightly
more open than remaining samples. This leads to more room for the liquid electrolyte to exist in the SBE,
which correlates with the increasing ionic conductivity that was presented during the EIS analysis.
However, as indicated by the similarity of the images in Figure 25, domain and pore size do not appear to
be significantly impacted by addition of liquid electrolyte beyond 45 wt%.

Photographs of the samples are pictured in Figure 26.

Figure 26. Film of electrolyte material with thiol 6DPER at 0.05 FG ratio with increasing electrolyte concentration
45 wt% (a) and 50 wt% (b), and at 0.1 FG ratio with increasing electrolyte concentration 45 wt% (c) and 50 wt%
(d).

When additional electrolyte was incorporated into the SBE the samples showed the materials’ opacity
increased along with the electrolyte wt% as shown in Figure 26. As these samples also showed a slightly
higher conductivity than the previous samples in phase one it further strengthens the idea that a more
opaque material suggests a more open morphology where the lithium ions can freely move through.

40
5.3 Phase 3: Half-cell Performance
Results from phase 3 are presented in Figure 27, displaying the capacity per cycle of lithiation and
delithiation. Each point represents the charging and discharging that constitute a cycle. Each series of
cycles utilized different current density (see section 4.2.6) leading to the variation in capacity seen in the
graph. The results from the two manufactured half-cells were very similar, signifying reproducibility of
the results.

The large drop in capacity observed in the first charge cycle is likely due to the formation of a solid
electrolyte interphase layer on the CF, as well as lithium ions being trapped in the CFs. [10] The same
trend in the first cycle is presented in the work done by Schneider and is common in lithium-ion batteries.

Figure 27. Capacity vs. cycle number.

From what can be seen in Figure 27, a maximum capacity of around 145 mAh g-1 was reached during the
final 5 cycles upon stabilization at a current density of 18.6 mA g-1 . Comparing this to the UD lamina
half-cell from Schneider et al. that reached a maximum capacity of 110 mAh g-1, it becomes evident that
the electrolyte in this study has managed to increase the half-cell capacity significantly. Stable cycling
behaviour and a satisfactory capacity retention were also achieved.

41
The successful performance of the electrochemical cycling further indicates that the SBE consists of a
bicontinuous, percolating structure; were this not achieved, capacity would be limited or nonexistent.
Moreover, it confirms the formation of a strong interphase which facilitates the transference of lithium
ions between electrode and electrolyte; transport through the electrolyte network is also substantiated.

In summary, the results illustrate an improvement in capacity compared to the results obtained by
Schneider et al.. Due to time limitations, there was no room for further analysis on the CF composite with
regards to the interfacial characteristics using, for example, SEM. Thus, for future studies it would be of
interest to perform a more extensive analysis of the half-cells electrochemical and mechanical
performance. What the thiols and solvents are contributing to here isn’t fully clear at this point; it may be
speculated that the high dielectric constant of the PC solvent facilitates ion mobility, whilst the interfacial
adhesion is benefited by the addition of thiol groups. However, this is outside the scope of this study, but
it can be concluded that a higher capacity in the half-cell is achieved.

42
6. Conclusion
The utilization of an EC:PC-based liquid electrolyte solvent in an SBE appeared to result in the phase
separation occurring at a later stage; this would contribute to a more homogenous network, which
promoted storage modulus but worsened conductivity relative to DMMP. Thus, this factor must be
considered upon applying a heterogenous structural battery on an industrial scale.

The addition of a step-growth mechanism to an electrolyte system that primarily propagates through a
chain-wise reaction feasibly caused the phase separation to occur later during the polymerization
reaction. This in turn is speculated to have contributed to a more homogenous structure with improved
mechanical properties for the thiol copolymers which exceeded that of previous studies on SBEs. Upon
increasing the ratio of thiol relative to the allyl groups, homogeneity appeared to increase further but at
the expense of storage modulus. The greater presence of flexible thiol-ether bonds is believed to have
surmounted the positive effect of increased homogeneity. Moreover, the results suggest that homogeneity
can promote ionic conductivity by creating a more uniform ion transport through the matrix.

As opposed to what Ihrner et al. established, not only mechanical properties are affected upon altering the
monomer structure of an SBE; the mobility of ions appears to be significantly dependent on the
morphology and phase separation. As these percolated systems are complex, multiple parameters have
different effects on the network and its functionality; certain factors will conflict with each other,
aggravating the settlement of definite conclusions.

6DPER is the most promising cross-linker for the application purposes of a structural battery; it exhibited
the highest elastic modulus and conductivity after incorporation into the SBE. Upon increasing wt%
liquid electrolyte in the 6DPER-based SBE, conductivity was improved by one order of magnitude whilst
storage modulus was reduced; the electrolyte is determined to have a plasticising effect on the network. It
also resulted in a more open morphology.

The half-cell created using the optimized formulation showed significant improvement in the energy
capacity as compared to the one by Schneider et al., reaching a maximum capacity of around 145 mAh

43
g-1. An improvement to the electrolyte system can thus be concluded. Stable cycling behaviour and a
satisfactory capacity retention were also observed.

6DPER with 45 wt% electrolyte and 0.05 FG ratio, with a solvent consisting of the greener and more
applicable EC:PC, appears promising with a modulus of 578 MPa for use in a structural battery
application. The conductivity still remains short of the target at a magnitude of 10-4 S cm-1 but the
understanding of SBE systems developed in this study may contribute to further research that can go on to
meet application demands.

44
7. Potential errors
During the curing stage of the experiment, the oven’s set temperature did not correspond accurately with
its actual inner temperature according to a measurement taken by a separate thermometer. No adjustment
was made as it could affect the curing process and the validity of the results.

The experiment was not conducted in a cleanroom environment, so contaminants were prevalent in the
laboratory. Contaminants could have entered the sample precursor or electrolyte solution before curing
despite precautions taken, as the presence of unknown particles were observed in the reaction mixtures.
This can cause an unexpected deviation in the results. Overall the presence of oxygen and moisture can
greatly affect the results as the electrolyte is sensitive to oxidation. Therefore, replicating the experiment
in a glove box is expected to yield more promising results.

There is also the potential error with faulty equipment - the EIS instrument needed to be assembled by
manually placing the samples onto the instrument, and then screwing it shut. It sometimes occurred that a
sample broke into splinters after detaching it from the instrument. This could lead to the EIS’ gold wires
being in too close contact with the samples to give accurate results. Additionally, the scale used for the
gravimetric analysis oscillated too frequently between measurements. This could signify a high margin of
error in the gravimetric analysis and should be repeated with a more accurate scale in the future.

For both SEM analysis of phase 1 and phase 2, there was an initial step where the samples had to be
submerged in liquid nitrogen in order to more easily break apart the samples without damaging the
microstructure. The whole process of extracting the liquid electrolyte, drying and later fracturing the
samples pose a risk of introducing foreign artifacts that vitiate the microstructure. Some of the samples
experienced burns on the surface during the time spent in the electron microscope. This is due to the
material being sensitive to the electron beam and a picture cannot be held for too long until the material
gets burnt. This makes reshoots of the same material difficult.

Then there is also the fact that SEM doesn’t cover the entirety of the sample and can leave critical parts of
the morphology out. Sample preparation can also have affected the mostly similar SEM pictures where it

45
could be argued that additional electrolyte beyond 45 wt% actually did result in larger and more open
pores from the first viewing. It would be best to test this in later experiments.

46
8. References

[1]L. E. Asp et al., “A Structural Battery and its Multifunctional Performance,” Advanced Energy and
Sustainability Research, vol. 2, no. 3, p. 2000093, Jan. 2021, doi: 10.1002/aesr.202000093.

[2]“Structural battery composites for mass-less energy storage | Chalmers,” Chalmers.se, 2021.
https://www.chalmers.se/en/projects/Pages/Structural-battery-composites-for-mass-less-energy-storage.as
px# (accessed Jun. 08, 2021).

[3]Chanan Bos, “Tesla’s New Structural Battery Pack — It’s Not Cell-to-Pack, It’s Cell-to-Body,”
CleanTechnica, Oct. 10, 2020.
https://cleantechnica.com/2020/10/10/teslas-new-structural-battery-pack-its-not-cell-to-pack-its-cell-to-bo
dy/ (accessed Jun. 08, 2021).

[4]W. Johannisson, D. Zenkert, and G. Lindbergh, “Model of a structural battery and its potential for
system level mass savings,” Multifunctional Materials, vol. 2, no. 3, p. 035002, Sep. 2019, doi:
10.1088/2399-7532/ab3bdd.

[5]“Structural battery electrolytes,” KA103X, vol. 2021.

[6]D. Zhou, D. Shanmukaraj, A. Tkacheva, M. Armand, and G. Wang, “Polymer Electrolytes for
Lithium-Based Batteries: Advances and Prospects,” Chem, vol. 5, no. 9, pp. 2326–2352, Sep. 2019, doi:
10.1016/j.chempr.2019.05.009.

[7]L. E. Asp and E. S. Greenhalgh, “Structural power composites,” Composites Science and Technology,
vol. 101, no. , pp. 41–61, Sep. 2014, doi: 10.1016/j.compscitech.2014.06.020.

[8]M. Puttegowda, S. M. Rangappa, M. Jawaid, P. Shivanna, Y. Basavegowda, and N. Saba, Sustainable


Composites for Aerospace Applications. Woodhead Publishing, 2018, pp. 315–351.

[9]A. K. Sharma, R. Bhandari, A. Aherwar, and R. Rimašauskienė, “Matrix materials used in composites:
A comprehensive study,” Materials Today: Proceedings, vol. 21, no. 3, pp. 1559–1562, 2020, doi:
10.1016/j.matpr.2019.11.086.

[10]L. M. Schneider, N. Ihrner, D. Zenkert, and M. Johansson, “Bicontinuous Electrolytes via Thermally
Initiated Polymerization for Structural Lithium Ion Batteries,” ACS Applied Energy Materials, vol. 2, no.
6, pp. 4362–4369, May 2019, doi: 10.1021/acsaem.9b00563.

47
[11]W. Johannisson et al., “Multifunctional performance of a carbon fiber UD lamina electrode for
structural batteries,” Composites Science and Technology, vol. 168, pp. 81–87, Nov. 2018, doi:
10.1016/j.compscitech.2018.08.044.

[12]L. E. Asp, M. Johansson, G. Lindbergh, J. Xu, and D. Zenkert, “Structural battery composites: a
review,” Functional Composites and Structures, vol. 1, no. 4, p. 042001, Nov. 2019, doi:
10.1088/2631-6331/ab5571.

[13]“How Does a Lithium-ion Battery Work?,” Energy.gov, 2017.


https://www.energy.gov/eere/articles/how-does-lithium-ion-battery-work (accessed Jun. 08, 2021).

[14]Jürgen Garche, Encyclopedia of electrochemical power sources. 3. Amsterdam: Elsevier, 2009.

[15]N. Shirshova et al., “Structural supercapacitor electrolytes based on bicontinuous ionic liquid–epoxy
resin systems,” Journal of Materials Chemistry A, vol. 1, no. 48, p. 15300, 2013, doi:
10.1039/c3ta13163g.

[16]N. Shirshova et al., “Composition as a Means To Control Morphology and Properties of Epoxy Based
Dual-Phase Structural Electrolytes,” The Journal of Physical Chemistry C, vol. 118, no. 49, pp.
28377–28387, Dec. 2014, doi: 10.1021/jp507952b.

[17]F. Baskoro, H. Q. Wong, and H.-J. Yen, “Strategic Structural Design of a Gel Polymer Electrolyte
toward a High Efficiency Lithium-Ion Battery,” ACS Applied Energy Materials, vol. 2, no. 6, pp.
3937–3971, May 2019, doi: 10.1021/acsaem.9b00295.

[18]P. Stilbs, Exempelsamling i Kemisk Dynamik KTH, 2021. 2021.

[19]M. Park, X. Zhang, M. Chung, G. B. Less, and A. M. Sastry, “A review of conduction phenomena in
Li-ion batteries,” Journal of Power Sources, vol. 195, no. 24, pp. 7904–7929, Dec. 2010, doi:
10.1016/j.jpowsour.2010.06.060.

[20]PerkinElmer and Inc., “Dynamic Mechanical Analysis (DMA) A Beginner’s Guide,” 2013. [Online].
Available:
https://www.perkinelmer.com/CMSResources/Images/44-74546GDE_IntroductionToDMA.pdf.

[21]P. M. V. Raja and A. R. Barron, “2.10: Dynamic Mechanical Analysis,” Chemistry LibreTexts, Jul. 13,
2016.
https://chem.libretexts.org/Bookshelves/Analytical_Chemistry/Book%3A_Physical_Methods_in_Chemist
ry_and_Nano_Science_(Barron)/02%3A_Physical_and_Thermal_Analysis/2.10%3A_Dynamic_Mechani
cal_Analysis (accessed Jun. 08, 2021).

48
[22]J. C. Hernandez-Mejia, “Dissipation Factor (Tan δ),” in CDFI Handbook, Georgia Institute of
Technology, 2016.

[23]N. Ihrner, W. Johannisson, F. Sieland, D. Zenkert, and M. Johansson, “Structural lithium ion battery
electrolytes via reaction induced phase separation ,” Journal of Materials Chemistry A, vol. 5, no. 48, pp.
25652–25659, 2017, doi: 10.1039/c7ta04684g.

[24]W. Li, Y. Pang, J. Liu, G. Liu, Y. Wang, and Y. Xia, “A PEO-based gel polymer electrolyte for lithium
ion batteries,” RSC Advances, vol. 7, no. 38, pp. 23494–23501, 2017, doi: 10.1039/c7ra02603j.

[25]H. Ben youcef, O. Garcia-Calvo, N. Lago, S. Devaraj, and M. Armand, “Cross-Linked Solid Polymer
Electrolyte for All-Solid-State Rechargeable Lithium Batteries,” Electrochimica Acta, vol. 220, pp.
587–594, Dec. 2016, doi: 10.1016/j.electacta.2016.10.122.

[26]W. Ren, C. Ding, X. Fu, and Y. Huang, “Advanced gel polymer electrolytes for safe and durable
lithium metal batteries: Challenges, strategies, and perspectives,” Energy Storage Materials, vol. 34, pp.
515–535, Jan. 2021, doi: 10.1016/j.ensm.2020.10.018.

[27]Q. Xiao, C. Deng, Q. Wang, Q. Zhang, Y. Yue, and S. Ren, “In Situ Cross-Linked Gel Polymer
Electrolyte Membranes with Excellent Thermal Stability for Lithium Ion Batteries,” ACS Omega, vol. 4,
no. 1, pp. 95–103, Jan. 2019, doi: 10.1021/acsomega.8b02255.

[28]W. Johannisson, “Exploring structural carbon fiber composites for mass-less energy and actuation,”
PhD dissertation, KTH Royal Institute of Technology, 2020.

[29]V. Tu, L. E. Asp, N. Shirshova, F. Larsson, K. Runesson, and R. Jänicke, “Performance of


bicontinuous structural electrolytes,” Multifunctional Materials, vol. 3, no. 2, p. 025001, Jun. 2020, doi:
10.1088/2399-7532/ab8d9b.

[30]“2.3: Step Growth and Chain Growth,” Chemistry LibreTexts, Jul. 24, 2020.
https://chem.libretexts.org/Bookshelves/Organic_Chemistry/Book%3A_Polymer_Chemistry_(Schaller)/0
2%3A_Synthetic_Methods_in_Polymer_Chemistry/2.03%3A_Step_Growth_and_Chain_Growth
(accessed Jun. 08, 2021).

[31]“Chain-Growth versus Step-Growth,” Polymerdatabase.com, 2021.


https://polymerdatabase.com/polymer%20chemistry/Chain%20versus%20Step%20Growth.html
(accessed Jun. 08, 2021).

[32]M. Willgert, M. H. Kjell, G. Lindbergh, and M. Johansson, “New structural lithium battery
electrolytes using thiol–ene chemistry,” Solid State Ionics, vol. 236, pp. 22–29, Apr. 2013, doi:
10.1016/j.ssi.2013.01.019.

49
[33]I. Ribca et al., “Exploring the Effects of Different Cross-Linkers on Lignin-Based Thermoset
Properties and Morphologies,” ACS Sustainable Chemistry & Engineering, vol. 9, no. 4, pp. 1692–1702,
Jan. 2021, doi: 10.1021/acssuschemeng.0c07580.

[34]C. E. Hoyle, T. Y. Lee, and T. Roper, “Thiol-enes: Chemistry of the past with promise for the future,”
Journal of Polymer Science Part A: Polymer Chemistry, vol. 42, no. 21, pp. 5301–5338, 2004, doi:
10.1002/pola.20366.

[35]H. Liu and H. Chung, “Visible-Light Induced Thiol–Ene Reaction on Natural Lignin,” ACS
Sustainable Chemistry & Engineering, vol. 5, no. 10, pp. 9160–9168, Aug. 2017, doi:
10.1021/acssuschemeng.7b02065.

[36]PubChem, “Dimethyl methylphosphonate,” pubchem.ncbi.nlm.nih.gov.


http://pubchem.ncbi.nlm.nih.gov/compound/Dimethyl-methylphosphonate#datasheet=LCSS%C2%A7ion
=GHS-Classification. (accessed Jun. 08, 2021).

[37]PubChem, “Propylene carbonate,” pubchem.ncbi.nlm.nih.gov.


https://pubchem.ncbi.nlm.nih.gov/compound/7924.

[38]“1,3-Dioxolan-2-one,” webbook.nist.gov.
http://webbook.nist.gov/cgi/cbook.cgi?ID=C96491&Type=IR-SPEC&Index=0 (accessed Jun. 08, 2021).

[39]“Propylene Carbonate,” webbook.nist.gov.


http://webbook.nist.gov/cgi/cbook.cgi?ID=C108327&Units=SI&Type=IR-SPEC&Index=1#IR-SPEC.

[40]“Lithium triflate(33454-82-9) IR2,” www.chemicalbook.com.


https://www.chemicalbook.com/SpectrumEN_33454-82-9_ir2.htm.

[41]“BISPHENOL A DIMETHACRYLATE(3253-39-2) IR1,” www.chemicalbook.com.


https://www.chemicalbook.com/SpectrumEN_3253-39-2_IR1.htm.

50
Appendix A

Table A1: Molecular weight of crosslinkers and distribution of functional groups.

Thiols Mw [g mol-1] No. of SH groups per molecule

3TMP 398,6 3

4PER 489,1 4

6DPER 783,1 6

Table A2. FG ratio between thiol and ene converted into mass [g]. 39wt% electrolyte

Ratio, thiol : ene 3TMP : ene [g] 4PER : ene [g] 6DPER : ene [g]

1 : 0.05 0.0720 : 2.9280 0.0664 : 2.9336 0.0708 : 2.9292

1 : 0.1 0.1407 : 2.8593 0.1300 : 2.8700 0.1383 : 2.8617

1 : 0.2 0.2688 : 2.7312 0.2491 : 2.7509 0.2645 : 2.7355

Table A3. Raw data from phase 1 of exact mass measurements of thiols, monomer B, initiator and liquid electrolyte
mixture.
FG ratio 0.1 m_Thiol m_monom.B m_initiator m_electrolyte
3TMP 0,1430 2,8592 0,031 1,951
4PER 0,1370 2,873 0,0303 1,951
6DPER 0,1437 2,8616 0,0303 1,9511
FG ratio 0,05 m_Thiol m_monom.B m_initiator m_electrolyte
3TMP 0,0710 2,929 0,03 1,945
4PER 0,0657 2,932 0,03 1,947
6DPER 0,0709 2,929 0,03 1,946
FG ratio 0,2 m_Thiol m_monom.B m_initiator m_electrolyte
3TMP 0,2690 2,7312 0,3 1,949
4PER 0,2500 2,7513 0,3 1,946
6DPER 0,2654 2,7356 0,3 1,95

51
Table A4. Raw data from phase 2 of exact mass measurements of thiols, monomer B, initiator and liquid electrolyte
mixture.
FG ratio 0,05 m_Thiol m_monom.B m_initiator m_electrolyte
45 wt% 0,071 2,932 0,03 2,483
50 wt% 0,071 2,93 0,03 3,032
FG ratio 0,1 m_Thiol m_monom.B m_initiator m_electrolyte
45 wt% 0,138 2,862 0,03 2,481
50 wt% 0,138 2,862 0,03 3,031

52
Appendix B

Table B1. Storage modulus [MPa] and Tg [℃] for 3TMP, 4PER and 6DPER additives varying the FG ratios.
Reference sample specified for comparison. Electrolyte concentration of 39wt%.

Thiol Ratio of FG Storage Modulus [MPa at 30 ℃] Tg [℃]

0.05 838.24 158.62

3TMP 0.1 757.10 127.62

0.2 626.03 79.93

0.05 853.54 163.93

4PER 0.1 757.56 179.18

0.2 669.11 108.62

0.05 823.38 174.64

6DPE 0.1 773.56 151.5


R
0.2 876.14 84.65

Ref - 701.60 182.42

53
Table B2: Results from DMA analysis, containing T g and storage modulus [MPa] at ambient temperature
and Tg [℃] for corresponding FG ratio and wt% liquid electrolyte for 6DPER.
wt% Ratio of FG Storage Modulus [MPa at 30 ℃] Tg [℃]

0.05 823.38 174.64


39
0.1 773.56 151.5

0.05 578.18 191.66


45
0.1 553.97 184.36

0.05 462.06 > 195.94


50
0.1 439.75 186.62

54
Appendix C

Figure C1: SEM images for sample 0.05 at 45 wt% at a magnification of: a) 5K, b) 35K, C) 70K

Figure C2: SEM images for sample 0.05 at 50 wt% at a magnification of: a) 5K, b & c) 35K

Figure C3: SEM images for sample 0.1 at 45 wt% at a magnification of: a) 5K, b & c) 35K

55

You might also like