Sei sulla pagina 1di 46

Journal Pre-proof

Measurement of Mechanical and Fracture Properties of Solid Electrolyte Interaphase


on Lithium Metal Anodes in Lithium Ion Batteries

Insun Yoon, Sunhyung Jurng, Daniel P. Abraham, Brett L. Lucht, Pradeep R. Guduru

PII: S2405-8297(19)31005-0
DOI: https://doi.org/10.1016/j.ensm.2019.10.009
Reference: ENSM 952

To appear in: Energy Storage Materials

Received Date: 2 September 2019


Revised Date: 9 October 2019
Accepted Date: 10 October 2019

Please cite this article as: I. Yoon, S. Jurng, D.P. Abraham, B.L. Lucht, P.R. Guduru, Measurement of
Mechanical and Fracture Properties of Solid Electrolyte Interaphase on Lithium Metal Anodes in Lithium
Ion Batteries, Energy Storage Materials, https://doi.org/10.1016/j.ensm.2019.10.009.

This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition
of a cover page and metadata, and formatting for readability, but it is not yet the definitive version of
record. This version will undergo additional copyediting, typesetting and review before it is published
in its final form, but we are providing this version to give early visibility of the article. Please note that,
during the production process, errors may be discovered which could affect the content, and all legal
disclaimers that apply to the journal pertain.

© 2019 Elsevier B.V. All rights reserved.


Measurement of Mechanical and Fracture Properties of Solid Electrolyte Interaphase on Lithium

Metal Anodes in Lithium Ion Batteries

Insun Yoon1, Sunhyung Jurng2, Daniel P. Abraham3, Brett L. Lucht2, Pradeep R. Guduru1*

1
School of Engineering, Brown University, 184 Hope st. Providence RI 02912

2
Department of Chemistry, University of Rhode Island,140 Flagg Rd. Kingston RI 02881

3
Chemical Sciences and Engineering Division, Argonne National Laboratory, 9700 South Cass. Ave.

Argonne IL 60439

*E-mail: pradeep_guduru@brown.edu
1 Measurement of Mechanical and Fracture Properties of Solid Electrolyte Interaphase on Lithium

2 Metal Anodes in Lithium Ion Batteries

3 Abstract

4 Mechanical integrity of the solid electrolyte interphase (SEI) plays an essential role in determining

5 the life and performance of lithium-ion batteries. Fracture and continued formation of the SEI contribute

6 to consumption of lithium, drying of electrolyte, increase in impedance, and growth of dendrites resulting

7 in capacity fade and premature failure. Electrolyte additives such as fluoroethylene carbonate (FEC) have

8 been known to improve performance, but the underlying reasons have been elusive. Despite its

9 importance, reliable methods for mechanical characterization of SEI have been lacking. Here, we present

10 a new experimental technique that combines atomic force microscopy and membrane-bulge configuration

11 to accurately measure the stress-strain behavior of SEI, including the onset of inelastic response and

12 evolution of fracture. We characterize the SEI formed with two ethylene carbonate-based electrolytes,

13 without and with fluoroethylene carbonate (FEC) additive. The measurements show a striking contrast;

14 SEI with FEC additive has 80% higher elastic modulus and a vastly higher resistance to fracture. These

15 findings offer a mechanical-behavior based rationale to understand how SEI controls battery performance.

16 Moreover, the experimental technique offers a robust diagnostic tool to design electrolytes that can form

17 SEI with the desired mechanical properties for optimal battery performance.

18

19 Keywords: elastic modulus, yield strength, fracture resistance, electrolyte additives, fluoroethylene

20 carbonate, atomic force microscopy

21

22 1. Introduction

1
23 The increasing use of LIB for transportation demands further advances in their energy density,1 which

24 has attracted interest in Li-alloying materials (e.g. Si, Sn, Ge) and renewed interest in Li metal as

25 anodes.2-5 However, insufficient life and reliability have limited their use in practice.6-9 Recent

26 investigations reveal the central role of mechanical deformation and fracture of SEI in the failure

27 mechanisms of these anodes.6,10-12 The SEI is a thin layer, consisting primarily of electrolyte reduction

28 products, that forms on anode surfaces.13 As the lithium alloying anode particles undergo large volume

29 changes (up to ~300%) during electrochemical cycling, the SEI on their surface is subjected to

30 correspondingly large cyclic tensile/compressive strains.12,14,15 Fracture of SEI induced by the large strains

31 causes continued formation of SEI, which consumes the available lithium and the electrolyte, increases

32 the internal cell impedance, and eventually degrades cycle-life.10,16,17 SEI formed on lithium metal anodes

33 is locally strained due to non-uniform lithium plating and the consequent surface roughness as a result of

34 its inhomogeneous ionic-conductivity.10 Fracture of SEI intensifies non-uniform plating, leading to the

35 formation of dendrites, “dead” lithium and a “mossy” electrode.2,10,18 Consequently, the mechanical

36 properties and fracture resistance of SEI has been widely acknowledged to be an important determining

37 factor for LIB performance and reliability; its significance has been emphasized in several recent

38 articles.8,9,12,19

39 However, despite its importance, there have not been reliable experimental methods for accurate

40 mechanical characterization of SEI. Prior experimental approaches aimed at measurement of elastic

41 modulus of SEI are mostly limited to indentation using atomic force microscopy (AFM)20-22 or

42 measurement of acoustic wave velocity.23 These approaches are prone to uncertainties such as substrate

43 influence, uncertain indenter-sample contact area, or assumptions on the gravimetric density of SEI.

44 Consequently, the reported SEI elastic modulus ranges from tens of MPa to tens of GPa, a spread of three

45 orders of magnitude. Moreover, these approaches measure only the elastic modulus of SEI, while other

46 relevant properties for SEI failure such as the yield strength and inelastic response have not been reported.

2
47 Here, (i) we describe a new experimental technique for accurate mechanical characterization of SEI

48 and (ii) provide a basis to understand the role of electrolyte composition on cell performance in terms of

49 the difference in mechanical properties of the corresponding SEI. The experimental technique employs a

50 micron-scale membrane bulge configuration integrated with an atomic force microscope (AFM). The

51 membrane bulge configuration24-28 consists of a free-standing film subjected to an increasing lateral

52 pressure while the resulting deflection is measured. An analysis of the pressure-deflection relation yields

53 the residual stress, the plane strain elastic modulus, the yield strength, and the inelastic response.24,26-30

54 We present the results for SEI formed with two model electrolytes: (i) 1.2 M LiPF6 in ethylene carbonate

55 (EC) and (ii) 1.2 M LiPF6 in EC+ FEC (8:2 by weight). The main difference between the two electrolytes

56 is the FEC additive, which has been broadly observed to improve the cycling performance of LIB

57 anodes.31-38 The results reveal that the presence of FEC increases the elastic modulus by about 80% and

58 vastly improves the resistance to fracture during inelastic deformation. Although the molecular and meso-

59 scale origins of the observations are unknown at this time, the findings suggest that enhancing the

60 mechanical properties of SEI through electrolyte design could be a way forward in improving the life and

61 reliability of emerging LIB technologies.

62

63 2. Experimental

64 Membrane bulge test

65 The deflection of a free-standing narrow rectangular membrane with constrained edges, subjected to

66 lateral pressure (see Fig. 1) is determined by the geometry of the membrane (thickness, t, width, 2a, and

67 length), the residual stress in the membrane σ0, and the plane strain modulus of the membrane material .

68 Here, = /(1 − ), where E is the Young’s modulus and ν is the Poisson’s ratio. For a sufficiently

69 large aspect ratio of the rectangular membrane (length-to-width ratio of greater than about 6), the

70 influence of the membrane length on the pressure-deflection relation can be neglected for regions

3
71 sufficiently away from the ends (Fig. 1a).24 When the membrane remains elastic, the pressure q and the

72 maximum deflection (shown in Fig. 1b) are related through the following equation.

73 = +
Equation (1)

74 The pressure-deflection relation deviates from the above equation when the stress/strain level is large

75 enough to induce inelastic deformation of the membrane. Hence, the onset of inelastic deformation can be

76 determined from the stress at which the pressure-deflection relation deviates from the cubic relation

77 expressed by Eq. 1 (see Fig. 1c for a schematic illustration). In the inelastic regime, the strain can be

78 evaluated by approximating the shape of the deformed membrane with a circular arc of radius =

79 as shown in Fig. 1b. The membrane can then be approximated to be a section of a thin-walled

80 cylinder under uniform internal pressure; the corresponding hoop stress and hoop strain can be calculated

81 through,

!" "
82 = , # = #$ + arcsin − 1 Equation (2)
"

83 where εo is the residual strain, equal to $/


+ . The accuracy of the above approximation, even in the

84 elastic range, has been verified through experiments and computational modelling,29 and is widely used to

85 characterize the mechanical properties and inelastic response of thin-film materials.29,30,39-42

4
(a) (c)

(b) (d)

Figure 1. (a) Schematic illustration of a rectangular free-standing membrane with constrained edges
and (b) the deflected membrane under a uniform lateral pressure q. The pressure-deflection ( – $ )
relation is governed by the width 2a, the thickness t, the residual stress σ0, and the plane strain
modulus , of the membrane material. (c) Schematic of the pressure-deflection curve. When the
membrane is elastic, it is a cubic relation (solid curve, Eq. 1). Deviation from the cubic relation
indicates onset of inelastic response (dashed line). (d) Schematic of the hoop stress-strain ( -# )
relation, which is obtained by approximating the deflected shape of the membrane as a circular arc and
using Eq. 2.
86

87 Specimen fabrication and experimental setup

88 Fig. 2 illustrates the specimen fabrication steps and a photograph of the experimental setup. First, a

89 free-standing silicon nitride (SiN) membrane is fabricated on a silicon wafer (double side polished, <100>

90 orientation, ~10 mm × 10 mm, thickness of 500 µm) through standard micro-nano fabrication steps (Fig.

91 2a). The free-standing SiN membrane has approximate dimensions of 20 – 40 µm in width, 400 µm in

92 length and 20 nm in thickness. See Section S1 of the supporting information for a detailed description of

93 the fabrication process. Next, an ultra-thin polydimethylsiloxane (PDMS) film is deposited on the SiN

94 film through the following steps. PDMS base elastomer and the curing agent are mixed at a ratio of 10:1

95 and the mixture is diluted by hexane at a ratio of 1:15. The solution is spin-coated on the substrate at 6000

5
96 rpm for 2 minutes and cured at room temperature for two hours, followed by curing at 90 °C for twelve

97 hours. The SiN layer in the free-standing area is etched by CF4 reactive ion etching, which creates an

98 ultrathin (~240 nm) free-standing PDMS (Fig. 2b) membrane. Such free-standing PDMS membranes

99 have been shown to be highly compliant and resilient under large deformation.43,44 The specimen is

100 attached to an annular ceramic (Macor®) disc for subsequent assembly.

101 Next, a lithium thin film is deposited on the PDMS layer in a custom-made physical vapor

102 deposition (PVD) chamber located in an argon-filled glovebox (Fig. 2c). Prior to loading into the

(a) (b) (e)

(c) (d)
(f)

Figure 2. (a) A schematic and an SEM image of a free-standing SiN membrane. A long rectangular
silicon nitride membrane (20 - 40 µm × 400 µm) is fabricated on a Si substrate. (b) Schematic steps of
fabricating a freestanding ultra-thin PDMS membrane and an optical microscope image of it. The
PDMS layer (~240 nm) is created by spin-coating PDMS:hexane (1:15) solution on the SiN layer. The
SiN layer in the free-standing area is removed by CF4 reactive ion etching. (c) A schematic and a
photograph of the specimen after lithium deposition. Prior to lithium deposition, the specimen is
epoxy bonded to an annular disc for cell assembly (shown in the photo) and a small PDMS mask is
placed near the edge to create a substrate-lithium step (highlighted in the dotted square). (d) A
schematic and a photograph of the specimen after lithium-to-SEI conversion. The thin film lithium is
completely converted to SEI by letting it react with the electrolyte, and the substrate-lithium step
becomes a substrate-SEI step. (e) A schematic of the final specimen assembled in a custom pressure
cell. The pressure cell design enables controlled lateral pressure to be applied on the specimen. Note
that the film thicknesses shown are not to scale. (f) A photograph of the micro-bulge test setup
integrated with an AFM located in an argon-filled glovebox. Accurately regulated compressed argon
is supplied to the pressure cell. The AFM measures the SEI thickness and the resulting SEI/PDMS
bulge topography.

6
103 deposition chamber, a small PDMS mask (2 mm × 2 mm × 500 µm) is placed near the edge of the

104 specimen as illustrated in Fig. 2c (highlighted by a dashed square). Following Li deposition, the PDMS

105 mask is removed to create a PDMS-Li step on the specimen surface, which facilitates the measurement of

106 the SEI thickness as described below. Li deposition is carried out at a nominal source temperature at 830

107 K and a chamber pressure of lower than 10-4 Pa.45 Further, prior investigations have established that

108 PDMS is inert to lithium.45-47

109 Next, the entire lithium film is converted into SEI by letting it react with the electrolyte (see

110 Section 2 of the supporting information). As discussed in the introduction, two electrolytes are prepared –

111 1.2 M LiPF6 in EC (EC electrolyte) and 1.2M LiPF6 in EC + 20 wt% FEC (EC+FEC electrolyte). In each

112 case, ~0.2 mL of the electrolyte is dropped on the specimen surface immediately after taking it out of the

113 deposition chamber. The specimens are placed in an argon-tight container (to prevent evaporation of the

114 electrolyte) for approximately 48 hours to completely convert the lithium thin-film to SEI. The sample is

115 then rinsed with anhydrous dimethyl carbonate (DMC) to remove any residual lithium salt (Fig. 2d). It is

116 worth noting the salient features of the final specimens (Fig. 2e); (i) the PDMS-Li step on the specimen is

117 converted to a PDMS-SEI step, which is measured by AFM to determine the SEI thickness accurately

118 (top left of Fig. 2e), and (ii) creation of a narrow rectangular free-standing SEI-PDMS membrane, which

119 would be subjected to a bulge test.

120 In order to determine the mechanical behavior of the SEI from the bulge test on the SEI-PDMS

121 double-layer membrane, it is necessary to subtract the response of the PDMS layer. First, the membrane

122 bulge test was carried out on bare PDMS films (no SEI on them) and their plane strain modulus and the

123 residual stress are determined to be 6.6 MPa and 0.15 MPa respectively (see Section S3 of the supporting

124 information for details). These values are similar to those reported by Thangawng et al.44 With a thickness

125 of 240 nm, the stretch stiffness of the PDMS film as measured by the product of its modulus and

126 thickness (referred from here onwards as the modulus-thickness) is 1.65 N/m; it would be subtracted from

127 combined stiffness of the SEI-PDMS membrane to extract the properties of the SEI. However, it will be

7
128 seen in the next section that it is only a small fraction, about 5%, of the combined stretch stiffness. Thus,

129 the pressure-deflection relation for the SEI/PDMS double-layer membrane is determined almost entirely

130 by the mechanical behavior of the SEI layer. Similarly, the residual stress – thickness product of the

131 PDMS layer is 0.038 N/m, which is subtracted to determine the residual stress in the SEI. It will be seen

132 in the next section that, similar to the stretch stiffness, the residual stress-thickness of the PDMS layer is

133 small compared to that of the SEI layer.

134 The specimens are assembled in a custom-designed pressure cell as illustrated in Fig. 2e. The

135 bottom chamber of the pressure cell can be filled with argon of controlled pressure at a resolution of 70

136 Pa (Fig. 2e), subjecting the SEI-PDMS membrane to lateral deflection. The pressure cell is integrated

137 with an AFM (Fig. 2f), which measures the SEI thickness, the resulting membrane bulge deflection and

138 the SEI surface topography evolution. The lithium deposition chamber and the AFM are located in a

139 single argon-filled glovebox (O2, H2O < 0.1 ppm); the samples remain in the same inert atmosphere

140 during fabrication, electrochemical reaction to form SEI and the pressure-deflection measurements in the

141 AFM.

142

143 Results and discussion

144 When the Li film is exposed to the electrolyte, the ensuing reactions convert it to a translucent SEI layer.

145 The composition of the resulting SEI may not be identical to that of the SEI formed under electrochemical

146 cycling, hence the SEI investigated here should be viewed as model SEI. Moreover, past investigations

147 have used the method of direct reaction of Li metal with the electrolyte in order to form SEI and

148 investigate its composition48,49. The method employed here is consistent with this prior body of work.

149 For the EC-based electrolytes used in the current investigation, the freestanding membrane remains

150 taut during SEI formation, which implies that the residual stress in the SEI, if any, is tensile. This is in

151 contrast to the case of the ionic liquid (IL) electrolyte reported previously45 where the residual stress in

8
152 the SEI is compressive, resulting in buckling and wrinkling of the underlying PDMS surface. The

153 membranes are subjected to lateral pressure in increments of about 700 Pa in the elastic range while the

154 AFM measures the resulting bulge topography at each pressure. Beyond the elastic limit, the pressure

155 increment is increased to 3500 Pa. In the inelastic regime, in addition to the bulge topography, a high-

156 resolution surface image (20 µm × 20 µm) is also measured at the center of the membrane at each

157 increment.

158 Fig. 3 presents a summary of the results for the two electrolytes considered; Figs. 3a-c (designated as

159 SEIEC) are for the EC electrolyte and Figs. 3d-f (designated as SEIEC+FEC) are for the EC+FEC electrolyte.

160 Figs. 3a and 3d show the deflected shapes of the membrane at a few representative values of the applied

161 pressure (for the sake of clarity, the ordinate scale is magnified by an order of magnitude). The SEI

162 thickness is determined by measuring the PDMS-SEI step height illustrated in Fig. 2e (see Section S4 of

163 the supporting information for details) and the width of the membrane is measured from the bulge

164 topography. The membrane width (2a) remains constant throughout the experiment for all cases

165 considered here, which implies that the interface between PDMS and the Si substrate does not delaminate

166 for the range of pressures applied in the experiments. The SEIEC sample in Figs. 3a-c has an SEI thickness

167 of 118 nm and a membrane width of 31.7 µm. The corresponding values for the SEIEC+FEC sample in Figs.

168 3d-f are 116 nm and 22.3 µm respectively. Figs. 3b and 3e show the pressure vs. deflection curves for the

169 SEI+PDMS. In order to analyze the pressure vs. deflection curves to extract the mechanical response of

170 the SEI, Eqs. 1 and 2 need to be modified to account for the double layer structure of the SEI+PDMS

171 membrane. The modified equations, details of the analysis and how the contribution of the PDMS layer is

172 subtracted to obtain the SEI response are shown in Section S4 of the supporting information. The results

173 of the analysis are shown in Figs. 3c and 3f. For the sake of clarity, in Fig. 3, the elastic range is shown in

174 blue and the inelastic range is shown in orange. It is worth noting that, as observed in Section 2, the

175 contribution of the PDMS layer to the response of the SEI+PDMS double layer is only a small fraction of

176 that of the SEI layer. The combined elastic modulus-thickness, + ∗ -, and the residual stress-thickness of

9
177 the SEIEC+PDMS membrane, σ ∗ -, are ~30.7 N/m and ~0.52 N/m respectively; these values for the

178 SEIEC+FEC/PDMS membrane are ~52.7 N/m and ~0.52 N/m. The corresponding values for the PDMS

179 layer alone are 0.038 N/m and 1.65 N/m. These are approximately 5% and 7% respectively of the

180 SEIEC/PDMS membrane, and 3% and 7% respectively of the SEIEC+FEC/PDMS membrane. Hence, it is

181 clear that the measured pressure-deflection behavior of the specimens is dominated by the response of the

182 SEI.

183 In the double layer membrane, the responses of the SEI and the PDMS layers are elastic initially. A

184 fit of the pressure vs. deflection data by assuming elastic response of the constituent layers is shown in

185 Figs. 3b and 3e (blue dashed curves). The fit agrees with the data initially, but it begins to deviate at

186 higher deflections (i.e. higher strains). It is well known that PDMS remains linear elastic up to a strain of

187 50% 50, i.e., the PDMS layer remains elastic in the strain range applied in this investigation. Hence, the

188 deviation between the elastic fit and the experimental data indicates the onset of inelastic deformation in

189 SEI (as noted schematically in Fig. 1b). From the elastic fits and following the procedure described in

190 Section S5 of the supporting information for the subtracting the contributions of the PDMS layer, the

191 residual stress, plane strain modulus, and the yield hoop stress of the SEI in each case are obtained. From

192 multiple repeated experiments, the plane strain modulus, elastic limit and residual stress of SEIEC are

193 measured to be approximately 240 MPa, 9 MPa and 4.5 MPa respectively; the corresponding values for

194 the SEIEC+FEC are 430 MPa, 9 MPa, and 3.9 MPa. These values are summarized in Table 1 along with the

195 estimated uncertainty/scatter.

196 Based on the modified form of Eq. 2 in Section S6 of the supporting information, the (hoop) stress vs.

197 strain relations for SEIEC and SEIEC+FEC are constructed and presented in Figs. 3c and 3f. Recall that Eq. 2

198 is based on approximating the bulged membrane to be a segment of a thin-walled cylinder. The accuracy

199 of this approximation in the elastic range is demonstrated in Section S6 of the supporting information.

200 Recently, Tanaka et al.51 suggested inelastic deformation behavior of SEI is essential to understand and

201 investigate the failure mechanisms52. The results presented in Figs. 3c, f show that the SEI deformation

10
202 response can be reasonably approximated as elastic – perfectly plastic. However, it is important to

203 recognize that the measured inelastic response is likely to have contributions from plasticity as well as

204 cracking, so the elastic-perfectly plastic description is strictly valid beyond the elastic limit until cracks

205 begin to appear. The deformation behavior investigated here can provide realistic property parameters for

206 models aimed at understanding Li dendrite-SEI interactions or the mesoscale structure of SEI53,54.

(a) (b) (c)

(d) (e) (f)

Figure 3. (a-c) Evolution of the bugle topography, pressure-deflection response, and hoop stress vs.
strain curve for SEIEC and (d-f) those for SEIEC+FEC. (a, d) Initially flat membranes deflect with
progressively increasing applied pressure. (b, e) The pressure-deflection relation initially follows the
elastic relation (blue dashed line represents the elastic fit) before deviating from it. The deviation
implies onset of inelastic deformation of the membrane. (c, f) Hoop stress vs. hoop strain of the SEI in
each case, which gives the residual stress, elastic limit, plane strain modulus and inelastic response.
The dashed lines are linear fits in the elastic range.
207

208 Table 1 Measured residual stress, plane strain modulus and yield hoop stress of SEIEC and SEIEC+FEC.

Residual stress Plane strain modulus Yield hoop stress

SEIEC 4.5 ± 0.8 MPa 238 ± 7.8 MPa 9 ± 1.0 MPa

11
SEIEC+FEC 3.9 ± 1.3 MPa 429 ± 10.0 MPa 9 ± 1.7 MPa

209

210 A striking observation is that the addition of FEC increases the plane strain elastic modulus by

211 approximately 80% (~240 MPa to ~430 MPa). It appears not to substantially alter the residual stress and

212 elastic limit. The increase in the elastic modulus may arise from difference in the composition55,56 and the

213 nano-structure34,57 of the SEI.

214 There have been several recent investigations aimed at measuring the elastic modulus of SEI

215 using AFM indentation approach: (i) Young’s modulus of SEI formed on highly oriented pyrolytic

216 graphite (HOPG) anode using 1 M LiPF6 in EC/DMC (1:1) was reported to be 3.8 ± 5.6 GPa.21 (ii) The

217 measurements on SEI formed on silicon thin film anodes using the same electrolyte composition showed

218 Young’s modulus varying between tens of MPa to several GPa.22 (iii) Mean values of Young’s modulus

219 of SEI formed on MnO electrode using 1 M LiPF6 in EC/propylene carbonate (PC) electrolyte were ~16

220 MPa and ~540 MPa for the inner layer and the outer layer respectively.20 Besides the AFM indentation

221 approach, Zhang et al.23 used laser acoustic wave technique to measure the Young’s modulus of SEI

222 formed on silicon electrode using 1M LiPF6 in EC/DMC (1:1). They assumed the gravimetric density of

223 SEI to be that of fully dense Li2CO3 to yield Young’s modulus of ~50 GPa and ~70 GPa at formation

224 potentials of 0.4 V and 0.2 V respectively. Although a direct comparison between the prior reports and the

225 present investigation cannot be made due to substantial uncertainties or assumptions in the former, the

226 plane strain moduli measured here (240 MPa for SEIEC and 430 MPa for SEIEC+FEC) are within the broad

227 range measured by the AFM indentation approaches. However, it should be emphasized that the approach

228 used in this study eliminates several uncertainties involved in the indentation technique such as the

229 influence of substrate and inaccurate contact area. Thus, the elastic moduli measured here are expected to

230 be precise and capture the influence of electrolyte composition. At the same time, the method presented

231 here yields thickness-averaged properties/parameters; in its current form, it cannot capture the variations

232 along the thickness. Our previous report using elastic buckling-based metrology yielded a value of

12
233 1.6GPa for the plane strain modulus of SEI formed with lithium and bis(trifluoromethylsulfonyl)imide

234 (TFSI) anion based ionic liquid (IL) electrolyte reactions.45 This value is approximately 3-5 times higher

235 than that of the SEI formed with carbonate-based electrolytes measured here. The difference may be

236 attributed to the compositional difference: IL electrolytes are suggested to form mostly inorganic SEI

237 components,58,59 while those of carbonate-base electrolytes are both organic and inorganic materials.9,19

238 Fig. 4 presents detailed surface topographies of SEIEC near the elastic limit and at several strain

239 values in the inelastic regime. Fig. 4a shows the points on the stress-strain curve that correspond to the

240 AFM images in Figs. 4b-f. The AFM images are obtained near the center of the membrane, in both the

241 length and width directions. Fig. 4b shows the AFM image of the SEIEC at a strain of ~3.6% (point b in

242 Fig. 4a), which shows the grainy structure of the SEI, with “grain” sizes of 70 - 120 nm and a roughness

243 Ra of ~6.2 nm (See Fig. S7 in the supporting information for high resolution local surface images).

244 Similar grainy structures were observed in previous reports using AFM60 and TEM.34 Fig. 4c shows the

245 surface at an inelastic strain of ~4.0% (point c in Fig. 4a), which shows the appearance of isolated cracks

246 in SEIEC. The incipient cracks are not obvious at the scale of Fig. 4c (a larger image is provided in the

247 supporting information, Fig. S8); Fig. 4d shows a magnified view of a few representative cracks in Fig. 4c

248 (the magnified regions are marked by dashed rectangles in Fig. 4c). Recall that the membrane has a

249 narrow rectangular geometry with an aspect ratio of greater than ten. In this geometry, the membrane is

250 subjected to hoop straining in the width direction (indicated by red arrows in Fig. 4d), whereas the strain

251 in the longitudinal direction (indicated by black arrows in Fig. 4d), is zero (for regions sufficiently far

252 away from the ends). As one would expect, the initial cracks in Fig. 4d are generally aligned with the

253 longitudinal direction.

13
(a)
(b) (c)

(d)

(e) (f)

Figure 4. Fracture/failure evolution in SEIEC. (a) SEIEC stress vs. strain curve showing the points that
correspond to the AFM topography images shown. (b, c) Detailed surface topographies at strains of
(b) ~3.6% and (c) ~4.0%. (d) Magnified local images revealing a crack broadly aligned with the length
direction (black arrow direction). The red arrows indicate the hoop (width) direction. (e, f) Detailed
surface topography at strains of ~4.6% and ~5.6%. Blue arrows and dashed lines are drawn to
highlight the failure morphology.
254

14
(a) (b) (c)

(f)
(d) (e)

Figure 5. Crack evolution in SEIEC+FEC. (a) Stress vs. strain curve showing the points where presented
membrane surface topography is measured. (b-e) Detailed surface topography at strains of ~2.3%,
~3.8%, ~4.8% and ~6.2%. (f) A magnified local surface topography of (d) indicated as dashed
rectangle which contains a possible crack.
255

256 Figs. 4e and 4f correspond to strains of 4.6%, 5.6% respectively; the corresponding

257 failure/fracture patterns evolve into complex 2D patterns. It appears that SEIEC starts to crack at an early

258 stage of inelastic deformation. At a strain of ~4.6% (Fig. 4e), initiation of lateral failure (highlighted in

259 blue dashed line) is observed. The failure morphology shown in Fig. 4f at a strain of ~5.6% corresponds

260 to a further evolved fragmentation pattern. The faint lateral failure in Fig. 4e becomes obvious in Fig. 4f

261 as highlighted in the blue dashed lines at the same location. Notice that the failure patterns appear as

262 bright lines indicating local elevation of topography, which is attributed to local bulging of the PDMS in

263 the cracked regions due to the applied pressure. The AFM topography images without the highlighting

264 lines are shown in the supporting information, Fig. S9. Fig. 5 presents the surface topography evolution of

15
265 SEIEC+FEC with increasing strain. Fig. 5a shows the points on the stress-strain curve that correspond to the

266 AFM images in the figure, which correspond to strains of ~2.3%, ~3.8%, ~4.8% and ~6.2% respectively.

267 SEIEC+FEC also shows grainy surface structure with grain size varying from 50 nm to 90 nm and a

268 roughness Ra of 6.2 nm (Fig. S6 in supporting information). The average grain size of SEIEC+FEC is smaller

269 than that of SEIEC, which implies an FEC-induced difference in the substructure. A striking feature of the

270 SEIEC+FEC samples is that cracking does not appear until a strain of ~4.8% (compared to 3.6% for SEIEC),

271 and the crack density is dramatically smaller compared to that of SEIEC at higher strains (a larger image is

272 provided in the supporting information, Fig. S10). The cracks are almost invisible at a strain as high as of

273 6.2% (Fig. 5e), whereas the SEI is completely fragmented at a strain of 5.6% in Fig. 4e. The absence of

274 cracking in SEIEC+FEC suggests that the inelastic response in Fig. 5a is predominantly due to plasticity.

275 Contrasting Figs. 4 and 5 reveals that in the presence of FEC, the SEI cracks very little and retains its

276 mechanical integrity even at high inelastic strains. The contrast in the crack evolution between SEIEC and

277 SEIEC+FEC is direct evidence of improved mechanical properties and integrity of the of SEI formed in

278 presence of FEC in the electrolytes.

279 To gain further insights, the membrane topographies are measured after fully releasing the

280 applied pressure, which are presented in Fig. 6 As expected from the crack evolution, the surface

281 topography of SEIEC has web-like crack pattern over the entire area of the membrane (Fig. 6a); in contrast,

282 the surface of SEIEC+FEC appears smooth and continuous due to the absence of cracking (Fig. 6b). The

283 shapes of the unloaded membranes for the two cases in Figs. 6c and 6d show that the membranes do not

284 recover the initial flat topography due to inelastic deformation. Whereas the inelastic deformation of

285 SEIEC has significant contribution from crack evolution, the case of SEIEC+FEC is almost entirely due to

286 plastic deformation.

16
Figure 6. AFM topography of the membranes after completely releasing the applied pressure. (a)
Web-like failure morphology can be observed in SEIEC. (b) The surface of SEIEC+FEC is smooth due to
the absence of cracks, indicating improved mechanical stability. (c, d) Cross-section of AFM
topographies presented in (a, b) respectively. Also plotted is the cross-section of as-prepared flat
membranes. The membranes do not recover the initial flat topography due to inelastic deformation of
SEI.
287

288 Clearly, the difference in the mechanical properties of the SEI in the two cases is due to the difference in

289 the composition and the micro/nano-structure of the SEI as a result of FEC. Determining the precise

290 micro/nano-structure of the SEI is a formidable problem, which has been an active topic of research. FEC

291 as an electrolyte additive has been studied for several years and its influence on the composition of SEI

292 has been reported. Numerous reports suggest that the FEC additive increases the fraction of polymeric

293 species.32,33,36,56,61 Additionally, the reduction products of FEC are thought to improve the cross-linking of

17
294 polymeric species in SEI.62-64 FEC is also reported to promote formation of LiF in SEI.55,65,66 To

295 complement the mechanical measurements, we have characterized SEIEC and SEIEC+FEC using X-ray

296 photoelectron spectroscopy (XPS), and the details are shown in Section S11 of the supporting information.

297 The key differences in the spectra due to the addition of FEC in the electrolyte are as follows: (i)

298 relatively larger amount C-C, C-H containing (polymeric) species, (ii) increased amount of LiF, and (iii)

299 less amount of carbonate (Li2CO3) species. These findings, as well as the spectra profiles, are in

300 agreement with those reported in the literature.32,37,56,61 Qualitatively, the high elastic modulus of LiF (~65

301 GPa)67 among the SEI components could account for the measured increase in the modulus reported here,

302 however it is not conclusive since the measured values are about two orders of magnitude smaller than

303 that of LiF. Differences in the nanostructure of the SEI may have as much influence on the mechanical

304 properties as the molecular species present, as suggested by Brown et al.34 As noted in the introduction

305 section, the mechanical properties of SEI are expected to play a significant role in determining the

306 electrochemical performance. SEI fracture or SEI delamination expose the electrode surface to the

307 electrolyte, resulting in more SEI formation and loss of acive lithium. FEC is known to result in better

308 electrochemical performance and our investigation reveals that FEC increases the fracture strain

309 substantially. Hence, one can reasonably conclude that FEC enhances the performance by endowing the

310 SEI with superior mechanical properties. It is possible to extend such reasoning by suggesting that further

311 enhancements in electrochemical performance can be achieved by choosing electrolyte and additive

312 combinations that would incrase the fracture strain of SEI and adhesion strength of SEI to the underlying

313 electrode surface. A fundamental understanding of these issues requires a systematic investigation of the

314 meso-scale structure of the SEI and how it is influenced by electrolyte composition, which remains open

315 for future research.

316

317 Conclusions

18
318 We have developed a new experimental technique for accurate mechanical property and fracture

319 characterization of SEI in which freestanding SEI+PDMS membranes are subjected to lateral pressure

320 while the corresponding deflection is measured with atomic force microscopy. The technique is

321 distinguished by its unprecedented accuracy in determining the SEI properties. The technique is used to

322 characterize SEI formed with two electrolytes, 1.2 M LiPF6 in EC and 1.2 M LiPF6 in EC/FEC (8:2) to

323 reveal the influence of the FEC additive. The key conclusions from this investigation are as follows. (i)

324 The plane strain elastic modulus of SEIEC is measured to be ~240 MPa and that of SEIEC+FEC is measured

325 to be ~430 MPa. (ii) The residual stress and the elastic limit of SEIEC are approximately 4.5 MPa and 9

326 MPa respectively; the corresponding values for SEIEC+FEC are approximately 3.9 MPa and 9 MPa. (iii)

327 Presence of FEC in the electrolyte appears to increase the elastic modulus by approximately 80% while

328 no substantial influence on the residual stress and the yield stress is observed. (iv) The strain at the elastic

329 limit of SEIEC is ~3.8%; cracking begins soon after entering the inelastic regime and the SEI severely

330 degrades by a strain of ~5.6%. In stark contrast, only a few cracks are present in SEIEC+FEC at a strain as

331 high as of 6.2%, which demonstrates dramatically enhanced resistance to fracture and mechanical

332 integrity. (v) The stress-strain response of SEIEC and SEIEC+FEC can be characterized as approximately

333 elastic – perfectly plastic. In the case of SEIEC, cracking begins shortly after entering the inelastic regime;

334 hence the inelastic regime seen in the stress-strain curve is due to a combination of plastic deformation

335 and crack evolution. However, it is predominantly due to plastic deformation in the case of SEIEC+FEC.

336 The findings suggest that the superior mechanical properties observed in the presence of FEC can

337 partly explain the widely observed performance and cycle life improvement of Li ion battery anodes with

338 FEC containing electrolytes. An important and unresolved question is – how does FEC (and other

339 additives) change the mechanical properties so substantially? At this stage, one can only point to the

340 increased fraction of LiF in the SEI as a contributing factor for the higher elastic modulus. A satisfactory

341 explanation requires determination of the precise nanostructure of the SEI, which is beyond the scope of

342 this investigation.

19
343 The approach and the experimental technique presented here can be extended to study deformation

344 and failure of SEI formed with other relevant electrolyte compositions and artificial SEI; it can be a useful

345 tool in developing engineered SEI with the desired mechanical properties to enhance stability and life of

346 high energy-density lithium batteries.

347

348 Acknowledgments:

349 The authors gratefully acknowledge financial support from the United States Department of Energy

350 EPSCoR Implementation award (grant # DE-SC0007074). The micro and nanofabrication work in this

351 investigation was carried out at the Center for Nanoscale Systems (CNS), which is part of the National

352 Nanotechnology Coordinated Infrastructure (NNCI).

353 Data availability

354 The raw/processed data required to reproduce these findings cannot be shared at this time due to technical

355 or time limitations.

356 References

357 1 Consortium, U. S. A. B. Electric vehicle battery test procedures manual. USABC, Jan (1996).
358 2 Wu Xu, Jiulin Wang, Fei Ding,c Xilin Chen, Eduard Nasybulin, Yaohui Zhangad, Ji-Guang Zhang.
359 Lithium metal anodes for rechargeable batteries. Energy & Environmental Science 7, 513-537
360 (2014).
361 3 McDowell, M. T., Lee, S. W., Nix, W. D. & Cui, Y. 25th anniversary article: understanding the
362 lithiation of silicon and other alloying anodes for lithium‐ion batteries. Adv. Mater. 25, 4966-
363 4985 (2013).
364 4 Beaulieu, L., Beattie, S., Hatchard, T. & Dahn, J. The electrochemical reaction of lithium with tin
365 studied by in situ AFM. J. Electrochem. Soc. 150, A419-A424 (2003).
366 5 Beaulieu, L., Eberman, K., Turner, R., Krause, L. & Dahn, J. Colossal reversible volume changes in
367 lithium alloys. Electrochemical and Solid-State Letters 4, A137-A140 (2001).
368 6 Aurbach, D., Zinigrad, E., Cohen, Y. & Teller, H. A short review of failure mechanisms of lithium
369 metal and lithiated graphite anodes in liquid electrolyte solutions. Solid state ionics 148, 405-
370 416 (2002).
371 7 Marom, R., Amalraj, S. F., Leifer, N., Jacob, D. & Aurbach, D. A review of advanced and practical
372 lithium battery materials. Journal of Materials Chemistry 21, 9938-9954 (2011).

20
373 8 Peled, E. & Menkin, S. SEI: past, present and future. Journal of The Electrochemical Society 164,
374 A1703-A1719 (2017).
375 9 Xu, K. Electrolytes and interphases in Li-ion batteries and beyond. Chem. Rev. 114, 11503-11618
376 (2014).
377 10 Lin, D., Liu, Y. & Cui, Y. Reviving the lithium metal anode for high-energy batteries. Nature
378 nanotechnology 12, 194 (2017).
379 11 Chan, C. K. et al. High-performance lithium battery anodes using silicon nanowires. Nature
380 nanotechnology 3, 31 (2008).
381 12 Zhang, W.-J. A review of the electrochemical performance of alloy anodes for lithium-ion
382 batteries. Journal of Power Sources 196, 13-24 (2011).
383 13 Peled, E. The electrochemical behavior of alkali and alkaline earth metals in nonaqueous battery
384 systems—the solid electrolyte interphase model. Journal of The Electrochemical Society 126,
385 2047-2051 (1979).
386 14 Winter, M. & Besenhard, J. O. Electrochemical lithiation of tin and tin-based intermetallics and
387 composites. Electrochimica Acta 45, 31-50 (1999).
388 15 Rezazadeh-Kalehbasti, S., Liu, L., Maris, H. & Guduru, P. In Situ Measurement of Phase Boundary
389 Kinetics during Initial Lithiation of Crystalline Silicon through Picosecond Ultrasonics.
390 Experimental Mechanics, 1-9 (2019).
391 16 Wu, H. et al. Stable cycling of double-walled silicon nanotube battery anodes through solid–
392 electrolyte interphase control. Nature nanotechnology 7, 310 (2012).
393 17 Oumellal, Y. et al. The failure mechanism of nano-sized Si-based negative electrodes for lithium
394 ion batteries. Journal of Materials Chemistry 21, 6201-6208 (2011).
395 18 Zhamu, A. et al. Reviving rechargeable lithium metal batteries: enabling next-generation high-
396 energy and high-power cells. Energy & Environmental Science 5, 5701-5707 (2012).
397 19 Xu, K. Nonaqueous liquid electrolytes for lithium-based rechargeable batteries. Chem. Rev. 104,
398 4303-4418 (2004).
399 20 Zhang, J. et al. Direct observation of inhomogeneous solid electrolyte interphase on MnO anode
400 with atomic force microscopy and spectroscopy. Nano Lett. 12, 2153-2157 (2012).
401 21 Shin, H., Park, J., Han, S., Sastry, A. M. & Lu, W. Component-/structure-dependent elasticity of
402 solid electrolyte interphase layer in Li-ion batteries: Experimental and computational studies. J.
403 Power Sources 277, 169-179 (2015).
404 22 Zheng, J. et al. 3D visualization of inhomogeneous multi-layered structure and Young's modulus
405 of the solid electrolyte interphase (SEI) on silicon anodes for lithium ion batteries. PCCP 16,
406 13229-13238 (2014).
407 23 Zhang, Q., Xiao, X., Zhou, W., Cheng, Y. T. & Verbrugge, M. W. Toward High Cycle Efficiency of
408 Silicon‐Based Negative Electrodes by Designing the Solid Electrolyte Interphase. Advanced
409 Energy Materials 5, 1401398 (2015).
410 24 Vlassak, J. & Nix, W. A new bulge test technique for the determination of Young's modulus and
411 Poisson's ratio of thin films. J. Mater. Res. 7, 3242-3249 (1992).
412 25 Wang, L., Williams, C. M., Boutilier, M. S., Kidambi, P. R. & Karnik, R. Single-layer graphene
413 membranes withstand ultrahigh applied pressure. Nano Lett. 17, 3081-3088 (2017).
414 26 Xin, H., Borduin, R., Jiang, W., Liechti, K. M. & Li, W. Adhesion energy of as-grown graphene on
415 copper foil with a blister test. Carbon 123, 243-249 (2017).
416 27 Xu, D., Liechti, K. M. & de Lumley-Woodyear, T. H. Closed form nonlinear analysis of the
417 peninsula blister test. The Journal of Adhesion 82, 831-866 (2006).
418 28 Liechti, K. & Shirani, A. Large scale yielding in blister specimens. International Journal of Fracture
419 67, 21-36 (1994).

21
420 29 Nicola, L., Xiang, Y., Vlassak, J., Van der Giessen, E. & Needleman, A. Plastic deformation of
421 freestanding thin films: experiments and modeling. Journal of the Mechanics and Physics of
422 Solids 54, 2089-2110 (2006).
423 30 Xiang, Y., Chen, X. & Vlassak, J. J. Plane-strain bulge test for thin films. J. Mater. Res. 20, 2360-
424 2370 (2005).
425 31 Lin, Y.-M. et al. High performance silicon nanoparticle anode in fluoroethylene carbonate-based
426 electrolyte for Li-ion batteries. Chem. Commun. 48, 7268-7270 (2012).
427 32 Nguyen, C. C. & Lucht, B. L. Comparative study of fluoroethylene carbonate and vinylene
428 carbonate for silicon anodes in lithium ion batteries. J. Electrochem. Soc. 161, A1933-A1938
429 (2014).
430 33 Etacheri, V. et al. Effect of fluoroethylene carbonate (FEC) on the performance and surface
431 chemistry of Si-nanowire Li-ion battery anodes. Langmuir 28, 965-976 (2011).
432 34 Brown, Z. L., Jurng, S., Nguyen, C. C. & Lucht, B. L. Effect of Fluoroethylene Carbonate
433 Electrolytes on the Nanostructure of the Solid Electrolyte Interphase and Performance of
434 Lithium Metal Anodes. ACS Applied Energy Materials 1, 3057-3062 (2018).
435 35 Kim, H. M., Hwang, J.-Y., Aurbach, D. & Sun, Y.-K. Electrochemical properties of sulfurized-
436 polyacrylonitrile cathode for lithium–sulfur batteries: effect of polyacrylic acid binder and
437 fluoroethylene carbonate additive. The Journal of Physical Chemistry Letters 8, 5331-5337 (2017).
438 36 Markevich, E., Salitra, G., Chesneau, F., Schmidt, M. & Aurbach, D. Very stable lithium metal
439 stripping–plating at a high rate and high areal capacity in fluoroethylene carbonate-based
440 organic electrolyte solution. ACS Energy Letters 2, 1321-1326 (2017).
441 37 Markevich, E. et al. High Performance LiNiO2 Cathodes with Practical Loading Cycled with Li
442 metal anodes in Fluoroethylene Carbonate Based Electrolyte Solution. ACS Applied Energy
443 Materials (2018).
444 38 Salitra, G. et al. High Performance Cells Containing Lithium Metal Anodes, LiNi0. 6Co0. 2Mn0.
445 2O2 (NCM 622) Cathodes and Fluoroethylene Carbonate Based Electrolyte Solution with
446 Practical Loading. ACS applied materials & interfaces (2018).
447 39 Xiang, Y. & Vlassak, J. Bauschinger and size effects in thin-film plasticity. Acta Mater. 54, 5449-
448 5460 (2006).
449 40 Xiang, Y., Tsui, T. & Vlassak, J. J. The mechanical properties of freestanding electroplated Cu thin
450 films. J. Mater. Res. 21, 1607-1618 (2006).
451 41 Wei, X., Lee, D., Shim, S., Chen, X. & Kysar, J. W. Plane-strain bulge test for nanocrystalline
452 copper thin films. Scripta Mater. 57, 541-544 (2007).
453 42 Preiß, E. I., Merle, B. & Göken, M. Understanding the extremely low fracture toughness of
454 freestanding gold thin films by in-situ bulge testing in an AFM. Materials Science and
455 Engineering: A 691, 218-225 (2017).
456 43 Ryoo, J., Jeong, G., Kang, E. & Lee, S. in 15th International Conference on Miniaturized Systems
457 for Chemistry and Life Sciences 2011, MicroTAS 2011.
458 44 Thangawng, A. L., Ruoff, R. S., Swartz, M. A. & Glucksberg, M. R. An ultra-thin PDMS membrane
459 as a bio/micro–nano interface: fabrication and characterization. Biomed. Microdevices 9, 587-
460 595 (2007).
461 45 Yoon, I., Jurng, S., Abraham, D. P., Lucht, B. L. & Guduru, P. R. In Situ Measurement of the Plane-
462 Strain Modulus of the Solid Electrolyte Interphase on Lithium-Metal Anodes in Ionic Liquid
463 Electrolytes. Nano letters 18, 5752-5759 (2018).
464 46 Yang, Y. et al. Transparent lithium-ion batteries. Proceedings of the National Academy of
465 Sciences 108, 13013-13018 (2011).
466 47 Zhu, B. et al. Poly (dimethylsiloxane) Thin Film as a Stable Interfacial Layer for High‐
467 Performance Lithium‐Metal Battery Anodes. Adv. Mater. 29 (2017).

22
468 48 Kanamura, K., Tamura, H., Shiraishi, S. & Takehara, Z. i. XPS analysis of lithium surfaces following
469 immersion in various solvents containing LiBF4. Journal of the Electrochemical Society 142, 340-
470 347 (1995).
471 49 Kanamura, K., Tamura, H. & Takehara, Z.-i. XPS analysis of a lithium surface immersed in
472 propylene carbonate solution containing various salts. Journal of Electroanalytical Chemistry 333,
473 127-142 (1992).
474 50 Khanafer, K., Duprey, A., Schlicht, M. & Berguer, R. Effects of strain rate, mixing ratio, and
475 stress–strain definition on the mechanical behavior of the polydimethylsiloxane (PDMS) material
476 as related to its biological applications. Biomedical microdevices 11, 503 (2009).
477 51 Tanaka, M., Hooper, J. B. & Bedrov, D. Role of Plasticity in Mechanical Failure of Solid Electrolyte
478 Interphases on Nanostructured Silicon Electrode: Insight from Continuum Level Modeling. ACS
479 Applied Energy Materials (2018).
480 52 Kumar, R. et al. In situ and operando investigations of failure mechanisms of the solid electrolyte
481 interphase on silicon electrodes. ACS Energy Letters 1, 689-697 (2016).
482 53 Hao, F., Liu, Z., Balbuena, P. B. & Mukherjee, P. P. Mesoscale elucidation of solid electrolyte
483 interphase layer formation in Li-ion battery anode. The Journal of Physical Chemistry C 121,
484 26233-26240 (2017).
485 54 Hao, F., Verma, A. & Mukherjee, P. P. Mechanistic insight into dendrite–SEI interactions for
486 lithium metal electrodes. Journal of Materials Chemistry A 6, 19664-19671 (2018).
487 55 Nie, M., Abraham, D. P., Chen, Y., Bose, A. & Lucht, B. L. Silicon solid electrolyte interphase (SEI)
488 of lithium ion battery characterized by microscopy and spectroscopy. The Journal of Physical
489 Chemistry C 117, 13403-13412 (2013).
490 56 Veith, G. M. et al. Determination of the Solid Electrolyte Interphase Structure Grown on a Silicon
491 Electrode Using a Fluoroethylene Carbonate Additive. Scientific reports 7, 6326 (2017).
492 57 Jurng, S., Brown, Z. L., Kim, J. & Lucht, B. L. Effect of electrolyte on the nanostructure of the solid
493 electrolyte interphase (SEI) and performance of lithium metal anodes. Energy & Environmental
494 Science 11, 2600-2608 (2018).
495 58 Budi, A. et al. Study of the initial stage of solid electrolyte interphase formation upon chemical
496 reaction of lithium metal and N-methyl-N-propyl-pyrrolidinium-Bis (Fluorosulfonyl) imide. The
497 Journal of Physical Chemistry C 116, 19789-19797 (2012).
498 59 Shkrob, I. A., Marin, T. W., Zhu, Y. & Abraham, D. P. Why bis (fluorosulfonyl) imide is a “magic
499 anion” for electrochemistry. The Journal of Physical Chemistry C 118, 19661-19671 (2014).
500 60 Cohen, Y. S., Cohen, Y. & Aurbach, D. Micromorphological studies of lithium electrodes in alkyl
501 carbonate solutions using in situ atomic force microscopy. The Journal of Physical Chemistry B
502 104, 12282-12291 (2000).
503 61 Markevich, E. et al. Fluoroethylene carbonate as an important component in electrolyte
504 solutions for high-voltage lithium batteries: Role of surface chemistry on the cathode. Langmuir
505 30, 7414-7424 (2014).
506 62 Shkrob, I. A., Wishart, J. F. & Abraham, D. P. What makes fluoroethylene carbonate different?
507 The Journal of Physical Chemistry C 119, 14954-14964 (2015).
508 63 Michan, A. L. et al. Fluoroethylene carbonate and vinylene carbonate reduction: Understanding
509 lithium-ion battery electrolyte additives and solid electrolyte interphase formation. Chem.
510 Mater. 28, 8149-8159 (2016).
511 64 Jin, Y. et al. Understanding Fluoroethylene Carbonate and Vinylene Carbonate Based
512 Electrolytes for Si Anodes in Lithium Ion Batteries with NMR Spectroscopy. J. Am. Chem. Soc.
513 140, 9854-9867 (2018).

23
514 65 Xu, C. et al. Improved performance of the silicon anode for Li-ion batteries: understanding the
515 surface modification mechanism of fluoroethylene carbonate as an effective electrolyte additive.
516 Chem. Mater. 27, 2591-2599 (2015).
517 66 Okuno, Y., Ushirogata, K., Sodeyama, K. & Tateyama, Y. Decomposition of the fluoroethylene
518 carbonate additive and the glue effect of lithium fluoride products for the solid electrolyte
519 interphase: an ab initio study. PCCP 18, 8643-8653 (2016).
520 67 Combes, L. S., Ballard, S. S. & McCarthy, K. A. Mechanical and thermal properties of certain
521 optical crystalline materials. JOSA 41, 215-222 (1951).
522

523

524

525

526

527

528

529

24
530 Graphical abstract

531

532 A new experimental method provides a route for systematic investigations on the influence of electrolyte
533 compositions on the mechanical stability of solid electrolyte interphase.
534

25
Supporting information
Measurement of Mechanical and Fracture Properties of Solid Electrolyte Interaphase on Lithium

Metal Anodes in Lithium Ion Batteries

S1. Detailed steps in the fabrication of free-standing silicon nitride membranes

(a) (b) (c) (d)

Figure S1. Schematic illustration of fabrication of free-standing silicon nitride membranes. The
schematic is not to scale.

(a) A non-stoichiometric silicon nitride (SiN) film (~20 nm) is deposited by low pressure chemical

vapor deposition (LPCVD) technique on a Si wafer (double side polished, <100> orientation, ~10

mm x 10 mm, thickness of 500 µm).

(b) A Photoresist layer is patterned to expose a rectangular area of the substrate with nominal

dimensions of 1200 µm × 720 µm.

(c) The SiN layer in the exposed region is removed by CF4 reactive ion etching (RIE) at chamber

pressure of ~ 100 mTorr and the radio frequency (RF) power of ~ 100 W. The SiN layer in the

photoresist masked area remains intact. The photoresist layer is cleaned by a series of acetone,

methanol, isopropanol and distilled water baths.

(d) The substrate is wet-etched in 30 wt% potassium hydroxide in water solution heated at ~90 °C.

The SiN layer is not etched by KOH solution1 and plays a role of a stable mask. Thus, the

anisotropic etching only progresses in the exposed Si substrate area. The resulting sample has a

free-standing SiN rectangular membrane with dimensions of 20-40 µm × 400 µm.


S2. Evidence for the complete conversion from Li film to SEI.

(a) (b)

Figure S2. Photographs of a lithium thin film on a bulk PDMS substrate before and after conversion
to SEI. (a) Electrolyte drops are placed on the lithium film immediately after taking the sample out of
the deposition chamber. (b) Lithium film in contact with the electrolyte drops converts to transparent
SEI layer.

S3. Measurement of residual stress and plane strain modulus of ultrathin PDMS membrane

The plane strain modulus and the residual stress of a PDMS layer are measured to analyze their

contribution to the pressure-deflection behavior of the SEI+PDMS double layer membranes.

(i) The PDMS membranes used in the experiments have a thickness of ~240 nm (Fig. S3a).

However, these have proven to be too compliant for our instrumentation (the extremely

high compliance results in AFM scan instability under pressurization). Instead, we have

characterized thicker (~790 nm) PDMS free-standing membranes (Fig. S3b) and assumed

that the properties are independent of thickness in this range.

(ii) The PDMS membrane is subjected to lateral pressure of up to ~1100 Pa in increments of

approximately 100 Pa. The corresponding pressure vs. deflection curve is shown in Fig.

S3c.

(iii) The pressure-deflection curve is analyzed to obtain the plane strain modulus and

the residual stress of PDMS, which are measured to be ~6.6 MPa and ~0.15 MPa
respectively (Fig. S4d). The corresponding modulus-thickness ∗ and

residual stress-thickness ∗ values for the thinner PDMS layer (~240 nm) in

the SEI/PDMS membrane samples are ~1.58 N/m and ~0.036 N/m respectively.

(iv) These values suggest that PDMS layer contributes ~5% and ~3% of modulus-thickness of

SEIEC/PDMS and SEIEC+FEC/PDMS membranes; the contribution of PDMS on the

residual stress-thickness is approximately 7% for both SEIEC and SEIEC+FEC.


(a) (b)

(c) (d)

Figure S3. (a, b) AFM scans to measure the thickness of PDMS layers. (a) ~240 nm thick PDMS
membranes to support SEI layers. (b) ~790 nm thick PDMS membranes for measurement of elastic
modulus and residual stress. (c) Pressure deflection curve and (d) the hoop stress-strain curve of the
PDMS membrane. As discussed in the manuscript, the relative stiffness of the PDMS support layer is
very small compared to that of SEI in the SEI+PDMS double layer membranes.

S4. Determination of the thickness of the SEI layer


Figure S4. AFM scan of the SEI-substrate step. The step height represents the SEI thickness as shown
in the schematic.

S5. Analysis of the pressure-deflection relation of double-layer SEI+PDMS membrane and

subtraction of the PDMS layer contribution

S5.1 Pressure-deflection relation of a double-layer membrane

Eq.1 of the manuscript describes the pressure-deflection relation for a single layer membrane. It

can be extended for a double-layer membrane in a straightforward way as shown below. For the

specimens used in this investigation, the pressure-deflection curve results predominantly from the

stretching of the membrane; the bending contribution can be neglected.2, 3. Following the work of Vlassak

and Nix4, the equilibrium equation and boundary/symmetry conditions of the membrane can be written as

=− + Equation (S1)

± =0 Equation (S2)

0 =0 Equation (S3)

! ± =0 Equation (S4)

Here, u(x) and w(x) are displacements in x and y directions respectively (coordinates are shown in Fig.

S5a). Superscripts ‘SEI’ and ‘PDMS’ indicate SEI layer and PDMS layer respectively.

The hoop strain in each layer is


# $ % ()*+,
" =" + + =" +" &' = - *+,
Equation (S5)
%

# $ % ()./0*
" =" + +% =" +" &' = - ./0*
Equation (S6)

where

# $ %
" &' = +% Equation (S7)

(1*+, (1./0*
Here, " = - *+,
and " = - ./0*
are the residual strains in the SEI layer and the PDMS layer

respectively; " &' is the strain induced by the deflection, which is the same in each layer. Integration of

Eq. S1 and applying the boundary condition (Eq. S4) yields,

2
!= %
− 5%
% ()*+, 3 *+, 4()./0* 3 ./0*
Equation (S8)

62
= 5
()*+, 3 *+, 4()./0* 3 ./0*
Equation (S9)

The maximum deflection w0 is at x=0, from which

2
w = %
% ()*+, 3 *+, 4()./0* 3 ./0*
Equation (S10)

Substituting in Eq. S7 and using Eq. S9,

# 2
=" &' − 5% Equation (S11)
% ()*+, 3 *+, 4()./0* 3 ./0*

Integrating Eq. S11 and applying boundary/symmetry conditions (Eqs. S2, S3),

2
" &' − %
=0 Equation (S12)
8 ()*+, 3 *+, 4()./0* 3 ./0*

Rewriting the second term on the left-hand side of Eq. S12 in terms of w0 and a using Eq. S10,

% 1
" &' = Equation (S13)
9:
From Eqs. S5 and S6,

= - " &' + Equation (S14)

= - " &' + Equation (S15)

Substituting Eqs. S14 and S15 in Eq. S10,

2
w = - *+, ;<=> 4(1*+, 3 *+, 4 - ./0* ;<=> 4(1./0* 3 ./0*
%
Equation (S16)
%

Combining Eqs. S13 and S16 and rearranging the terms results in the following expression.

*+, *+, ./0* ./0*


% (1*+, 3 *+, 4(1./0* 3 ./0* ? 3 4 3
= ! + !9

Equation (S17)
: 9: @

Thus, the effective residual stress-thickness and modulus-thickness of the double-layered membrane are

simply the sum of those values of each layer.

S5.2 Measurement of the elastic limit (i.e., the yield stress)

Dividing each side of Eq. S17 by w0 and re-definition of variables yield following linear equation.

*+, *+, ./0* ./0*


% (1*+, 3 *+, 4(1./0* 3 ./0* ? 3 4 3 2
A= + B, ℎEFE, A = GH B = ! % Equation (S18)

: 9: @ 1

A plot of the pressure-deflection data in this form is presented in Figs. S5 b, c. The onset of inelastic

deformation is determined to the point where the data deviates from linearity.

S5.3 Subtraction of PDMS layer contribution

The pressure-deflection data in elastic range are fit to the following cubic equation.

= I ∙ ! + K ∙ !9 Equation (S19)

where

*+, *+, ./0* ./0*


% (1*+, 3 *+, 4(1./0* 3 ./0* ? 3 4 3
I= and K =

: 9: @
The resulting elastic fits are shown as blue dashed curves in Figs. S5 d, e, yielding the coefficients A and

B. Using the membrane width 2a, the coefficients A and B are rearranged as follows.

:
∙I= + Equation (S20)
%

9
?
?
∙K = + Equation (S21)

The stress-thickness ∗ and modulus-thickness ∗ values of the PDMS layer

are measured to be 0.036 N/m and 1.58 N/m respectively, as described in Section S3 above. The

corresponding PDMS contribution is shown as black dashed curves in Fig. S5 d, e, which are subtracted

from Eqs. S20 and S21 to calculate the residual stress and the plane strain modulus of SEI. It

can be seen from Figs. S5 d, e that the pressure-deflection behavior is dominated by the response of the

SEI layer.
(a)

(b) (c)

(d) (e)

Figure S5. (a) Schematic of the specimen showing x, y coordinates. (b, c) A = !20 vs. β=
2
) plots for SEIEC and SEIEC+FEC to assess the elastic range and the yield point. (d) Pressure
1
q vs. deflection w0 plot of SEIEC/PDMS membrane and (e) that of SEIEC+FEC/PDMS
membrane. Also presented as black dashed curves are the contribution of PDMS layer on the
pressure-deflection response. Detailed procedures to account the PDMS influence as well as
to obtain residual stress and plane strain modulus of SEI are described in the supporting
information text.

S6. Accuracy of the geometric approximation, i.e., approximating the bulged membrane to a section

of a thin-walled cylinder

Note that the analysis described in the preceding section is valid in the elastic range only. Beyond

the elastic range, the bulged membrane is approximated to be a section of a thin-walled cylinder and the
corresponding stress and strain are described by Eq. 2. Here, we refer to this approximation as the

geometric approximation. The geometric approximation is useful in constructing the stress-strain curve in

the inelastic range where an analytical solution is unavailable. Here, we demonstrate that the geometric

approximation is quite accurate in describing the experimental data over the entire measurement range,

including the elastic range.

First, we construct the hoop stress and strain relation in the SEI layer based on the double layered

membrane analysis in Section 5. Rearranging Eqs. S6 and S10,

= +- "MNO Equation (S22)

2:
+ = Equation (S23)
% 1

Substituting PDMS stress and "MNO terms in Eq. S23 from Eqs. S14 and S22 and rearranging the terms

results in an equation for the hoop stress in the SEI layer. Note that the PDMS layer contribution is being

subtracted.

+-
2: % 1
=P −P Q Q/ Equation (S23)
% 1 :

From Eq. S6, the hoop strain in the SEI layer is given as

(1*+, % 1
" = - *+,
+ Equation (S24)
9:

The above equations are valid in the elastic range. Now, we present equations for the hoop stress and

strain in the SEI layer by approximating the deflected membrane to be a segment of a thin walled cylinder

(i.e., the geometric approximation). From equilibrium,

+ = S Equation (S25)

1 4:
Rearranging Eq. S25 and Eq. S22 and recalling that S = ,
% 1
+-
2 1 4: % 1
=P −P Q Q/ Equation (S26)
% 1 :

Under the geometric approximation, the hoop strain in the SEI layer is given by (recall Eq. 2 in the main

text),

(1*+, T :
" = - *+,
+ sin6$ −1 Equation (S27)
: T

Notice that the only difference between the expressions for the hoop stress in the SEI layer from the

geometric approximation (Eq. S26) and the elastic analysis in Section S5 (Eq. S23) is the presence of ! %

in the first term of the right-hand side. Therefore, the maximum difference in the hoop stress between the

two cases is only ~2%. Figs. S6a and S6b show the hoop strain as a function of deflection from Eqs. S24

(red) and S27 (blue) for the experimental data presented in Fig. 3 in the main text. The two curves in each

graph almost overlap on each other, indicating that the calculated strain calculated from the elastic

analysis and the geometric approximation are nearly the same.

Figs. S6c and S6d compare the (hoop) stress vs. strain plots constructed from the analysis in

Section 5 and the geometric approximation for SEIEC and SEIEC+FEC respectively. The proximity of the

two curves to each other in each case demonstrates the accuracy of the geometric approximation in the

elastic range. Further, the geometric approximation continues to be valid beyond the elastic range since its

derivation does not rely on the assumption that the material response be elastic.
(a) (b)

Hoop strain εθ
Hoop strain εθ

Deflection w0 (nm) Deflection w0 (nm)

(c) (d)
Hoop stress σθ (Pa)

Hoop stress σθ (Pa)

Hoop strain εθ Hoop strain εθ


Figure S6. (a, b) Hoop strain vs. deflection constructed based on the analysis in Section S5 (red)
and the geometric approximation (blue) for (a) SEIEC and (b) SEIEC+FEC. (c, d) Hoop stress vs.
strain constructed based on Section S5 analysis (red) and the geometric (thin-walled cylinder)
approximation (blue) for (c) SEIEC and (d) SEIEC+FEC.
S7. Surface structure and roughness of SEI

(a) (b)

Figure S7. Detailed surface topography (2 µm × 2 µm) of (a) SEIEC and (b) SEIEC+FEC. Each
topography shows a granular surface structure with grain size of 70 - 120 nm and 50 – 90 nm for
SEIEC and SEIEC+FEC respectively. The surface roughness (Ra) is nominally similar in the two
cases, with a value of 6.2 nm. These quantities are measured using the built-in AFM software
(Bruker, Nanoscope Analysis ver. 1.9).
S8. Magnified SEI surface image containing cracks

Figure S8. Magnified image of Fig. 4c (at a strain of 4.0%) for enhanced visualization of the
cracks in SEI. Some of the representative cracks are highlighted by arrows.
S9. SEI surface evolution without annotations

Figure S9. Fig.4 without annotations or overlaid lines.


S10. Magnified SEIEC+FEC surface image

Figure S10. Magnified image of Fig. 5d (at a strain of 4.8%.) for enhanced visualization of
a possible crack (highlighted by an arrow)
S11. XPS measurements on SEIEC and SEIEC+FEC

An identical batch of specimens to those used for mechanical characterization is prepared to

investigate the chemical compositions of SEI by X-ray photoelectron spectroscopy (XPS). The XPS

samples are transferred from the glovebox to the XPS chamber using a vacuum transfer module designed

for environment-sensitive samples (Thermo Scientific K-Alpha XPS system). XPS measurements are

conducted using Al Kα radiation source (kν = 1486.68 eV) at a chamber pressure of less than 1×10-7 Pa.

Sputter etching is performed at an ion beam energy of 200 eV. The binding energies are calibrated by

referencing the lithium fluoride (LiF) position to be at 685 eV.

Fig. S11 summarizes the detailed XPS spectra in C 1s O 1s and F 1s regions as well as the atomic

percent as a function of etch time for SEIEC and SEIEC+FEC. The species at each binding energy is

identified primarily referencing a review paper written by Verma et al.5 along with several supplementing

references6-8. The following key observations can be made from each spectrum and the depth profile.

Fig. S11 (a, d) C 1s: Note that the C 1s spectra of SEIEC are scaled to 300% to improve the clarity.

Carbonate species (290 eV), C-O (286.5 eV) and C-C / C-H (285 eV) containing species are identified in

this region. Absence of FEC increase the fraction of inorganic carbonate species, while the presence of

FEC increase the fraction of organic species. The observation is in good agreement with the report by

Veith et al.9. They observed that 5% of FEC additive in 1.2 M the fraction of polymeric species in SEI.

The observations reported here also agree with those made by Aurbach’s group10-14.

Fig. S11 (b, e) O 1s: Note that the O 1s spectra of SEIEC+FEC are scaled by 300%. The two spectra

commonly consist of a broad peak at 532 eV spreading to 533 eV. These are the characteristics of

carbonate species and C-O containing species respectively. Stronger peak associated to carbonate species

in the O 1s spectra of SEIEC is in good agreement with the observation made from C 1s spectra.
Fig. S11 (c, f) F 1s: Both spectra have main peak at 685 eV with a small bump at 687 eV. These are

the characteristics of LiF and LixPFyOz or LiPF6. These spectra suggest that the majority of fluorine in the

SEI is in the form of LiF.

Fig S11 (g) Atomic percent of SEIEC and SEIEC+FEC: The atomic percent of carbon is close between

SEIEC and SEIEC+FEC. This suggests that the absolute amount of carbonate species is larger in SEIEC, while

that of polymeric species are larger in SEIEC+FEC. The observation further supported by higher percent of

oxygen in SEIEC. The amount of fluorine is smaller in SEIEC compared to SEIEC+FEC. This suggests

increased formation of LiF in SEI by addition of FEC. This observation is consistent with the report by

Nie et al.7 which characterized chemical compounds SEI on Si formed with either EC electrolyte or FEC

electrolyte using TEM, XPS and NMR.

(b) (c) (g)


(a)

(e) (f)
(d)

Figure S11. (a)-(f) Detailed XPS spectra in C 1s, O 1s, F 1s regions for SEIEC and SEIEC+FEC and (g)
atomic percent profile of each element. SEIEC appears to contain more carbonate species, while
SEIEC+FEC appears to contain more polymeric species and LiF.

Supporting Information References

1. Williams, K. R.; Gupta, K.; Wasilik, M., Etch rates for micromachining processing-Part II. Journal
of microelectromechanical systems 2003, 12 (6), 761-778.
2. Timoshenko, S. P.; Woinowsky-Krieger, S., Theory of plates and shells. McGraw-hill: 1959.
3. Freund, L. B.; Suresh, S., Thin film materials: stress, defect formation and surface evolution.
Cambridge University Press: 2004.
4. Vlassak, J.; Nix, W., A new bulge test technique for the determination of Young's modulus and
Poisson's ratio of thin films. J. Mater. Res. 1992, 7 (12), 3242-3249.
5. Verma, P.; Maire, P.; Novák, P., A review of the features and analyses of the solid electrolyte
interphase in Li-ion batteries. Electrochimica Acta 2010, 55 (22), 6332-6341.
6. Zhang, Q.; Xiao, X.; Zhou, W.; Cheng, Y. T.; Verbrugge, M. W., Toward High Cycle Efficiency of
Silicon‐Based Negative Electrodes by Designing the Solid Electrolyte Interphase. Advanced Energy
Materials 2015, 5 (5), 1401398.
7. Nie, M.; Abraham, D. P.; Chen, Y.; Bose, A.; Lucht, B. L., Silicon solid electrolyte interphase (SEI)
of lithium ion battery characterized by microscopy and spectroscopy. The Journal of Physical Chemistry C
2013, 117 (26), 13403-13412.
8. Nie, M.; Chalasani, D.; Abraham, D. P.; Chen, Y.; Bose, A.; Lucht, B. L., Lithium ion battery
graphite solid electrolyte interphase revealed by microscopy and spectroscopy. The Journal of Physical
Chemistry C 2013, 117 (3), 1257-1267.
9. Veith, G. M.; Doucet, M.; Sacci, R. L.; Vacaliuc, B.; Baldwin, J. K.; Browning, J. F., Determination
of the Solid Electrolyte Interphase Structure Grown on a Silicon Electrode Using a Fluoroethylene
Carbonate Additive. Scientific reports 2017, 7 (1), 6326.
10. Markevich, E.; Salitra, G.; Chesneau, F.; Schmidt, M.; Aurbach, D., Very stable lithium metal
stripping–plating at a high rate and high areal capacity in fluoroethylene carbonate-based organic
electrolyte solution. ACS Energy Letters 2017, 2 (6), 1321-1326.
11. Markevich, E.; Salitra, G.; Fridman, K.; Sharabi, R.; Gershinsky, G.; Garsuch, A.; Semrau, G.;
Schmidt, M. A.; Aurbach, D., Fluoroethylene carbonate as an important component in electrolyte
solutions for high-voltage lithium batteries: Role of surface chemistry on the cathode. Langmuir 2014,
30 (25), 7414-7424.
12. Etacheri, V.; Haik, O.; Goffer, Y.; Roberts, G. A.; Stefan, I. C.; Fasching, R.; Aurbach, D., Effect of
fluoroethylene carbonate (FEC) on the performance and surface chemistry of Si-nanowire Li-ion battery
anodes. Langmuir 2011, 28 (1), 965-976.
13. Markevich, E.; Salitra, G.; Talyosef, Y.; Kim, U.-H.; Ryu, H.-H.; Sun, Y.-K.; Aurbach, D., High
Performance LiNiO2 Cathodes with Practical Loading Cycled with Li metal anodes in Fluoroethylene
Carbonate Based Electrolyte Solution. ACS Applied Energy Materials 2018.
14. Salitra, G.; Markevich, E.; Afri, M.; Talyosef, Y.; Hartmann, P.; Kulisch, J.; Sun, Y.-K.; Aurbach, D.,
High Performance Cells Containing Lithium Metal Anodes, LiNi0. 6Co0. 2Mn0. 2O2 (NCM 622) Cathodes
and Fluoroethylene Carbonate Based Electrolyte Solution with Practical Loading. ACS applied materials
& interfaces 2018.

Potrebbero piacerti anche