Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

Research Article

Received: 23 February 2010 Accepted: 25 January 2011 Published online in Wiley Online Library: 1 April 2011

(wileyonlinelibrary.com) DOI 10.1002/jrs.2930

A novel combinative Raman and SEM mapping


method for the detection of exfoliation
of graphite in electrodes at very positive
potentials
Andreas Hintennach† and Petr Novák∗
The random exfoliation at very positive potentials (>5 V vs Li/Li+ ) of graphite (used as a conductive component in positive
electrodes of lithium-ion batteries) was investigated with in situ Raman microscopy and post mortem scanning electron
microscopy (SEM). A novel semiautomated computational method for the data analysis of both characterization methods was
developed to correlate Raman and SEM information with good lateral resolution, in order to locate exfoliated graphite particles.
Proof is given that the exfoliating particles detected via the semiautomatic in situ Raman microscopy mappings correctly
describes exfoliated areas, as confirmed via post mortem SEM pictures. Copyright  c 2011 John Wiley & Sons, Ltd.
Supporting information may be found in the online version of this article.

Keywords: SEM; Raman microscopy; lithium-ion battery; positive electrodes; exfoliation

Introduction mapping methods provide an overview of a selected area of the


graphite electrode (typically about 100 µm × 100 µm), providing
Electrochemical energy storage[1 – 4] is one of the key technologies
information, e.g. about the homogeneity of intercalated lithium
in a clean energy economy. The need for the reduction of carbon
into the graphene layers. There have been many reports on in situ
dioxide emission together with the increase in crude oil prices has
Raman spectroscopy with both time and space resolution focused
led to the exigency of augmentation of the energy efficiency of
on lithium-ion batteries and, frequently, on graphite electrodes
technical devices, e.g. cars. The current technology of choice for
and single graphite particles. However, in general the published
the near future is hybrid electric vehicles.[5,6] With the need for
results are focused on the negative graphite electrodes[15,29,38 – 43]
increase in the percentage of electric power (and a downsized
and/or on the use of graphite as conductive additive for standard
combustion engine) in mind, rechargeable lithium-ion batteries
‘4 V’ positive electrodes containing other active materials than
are considered to be the best energy storage devices. Compared
graphite.[44 – 49]
to nickel–metal hydride batteries, lithium-ion batteries can store
On the other hand, graphite is also used as an additive in the
about 2–3 times more energy per unit mass (energy density).[1 – 3]
positive electrodes of lithium-ion batteries, normally to increase
Besides the energy density, the long-term stability (high cycling
the electronic conductivity of the electrode mass. Of current
stability and long calendar life) of the batteries is a key factor for
interest is the behavior of graphite as the conductive additive in
their use in cars.
the ‘5 V’ future positive electrodes. Such electrodes offer a higher
Graphitic carbon is a key component of lithium-ion batteries. On
energy density because of the higher voltage. But they work in a
one hand, because of its high reversible specific charge of 372 Ah
potential region where (1) the intercalation of anions into graphite
kg−1 , high electronic conductivity, and good cycling stability,
is possible (as known from the so called ‘dual graphite’ cells – e.g.
graphitic materials are widely used in the negative electrodes of
the publication of Seel and Dahn[39] where the electrochemical
lithium-ion batteries.[7 – 10] For graphite serving as the negative
and structural aspects of anion intercalation into graphite are
working electrode, degradation effects due to exfoliation of the
discussed) and (2) as a possible deterioration mechanism the
graphene layers have been reported, with the consequence of
exfoliation of graphite can occur.[50] In this paper, we focus on
a deterioration of the cycling stability.[11 – 13] Several mechanisms
the detection of the exfoliation of graphite particles cycled under
have been proposed to explain the exfoliation as a consequence
of graphite surface characteristics,[14 – 17] the amount of defects at
the graphite surface,[18] and the surface chemistry.[15] To describe
∗ Correspondence to: Petr Novák, Paul Scherrer Institut, Electrochemistry
the amount of defects, earlier works have demonstrated that the
Laboratory, CH-5232 Villigen PSI, Switzerland. E-mail: petr.novak@psi.ch
active surface area of graphite[19,20] plays an important role in its
tendency for exfoliation. † Current address: Department of Mechanical Engineering, Electrochemical
Raman microscopy has proved to be a method of choice Energy Laboratory, Massachusetts Institute of Technology, 77 Massachusetts
for the in situ characterization of surfaces of working graphite Avenue, Building 31, Room 146, Cambridge, MA 02139, USA.
electrodes.[13,20 – 29] For this, in situ and post mortem surface
1754

Paul Scherrer Institut, Electrochemistry Laboratory, CH-5232 Villigen PSI,


mapping techniques have been developed.[22,28 – 37] These Raman Switzerland

J. Raman Spectrosc. 2011, 42, 1754–1760 Copyright 


c 2011 John Wiley & Sons, Ltd.
Raman and SEM mapping method

such rather abuse conditions at very positive potentials (>5 V using standard electrochemical equipment to study exfoliation
vs Li/Li+ ). The effect of electrochemical cycling is analyzed with because of repeated excursions to very positive potentials (>5 V vs
quantitative and qualitative methods of in situ and post mortem Li/Li+ ). The specific charge for the C-rate calculation was assumed
Raman mapping and is correlated with post mortem scanning to be 140 Ah kg−1 . The galvanostatic cycling was performed at
electron microscopy (SEM) results. The novel technique described C/10 for the first three cycles with subsequent 17 cycles at C/20.
below helps to detect and understand the degradation effects There were no idling periods at the potential reversal points.
in lithium-ion batteries when they are related to the exfoliation
of the graphitic conductive additive. Clearly, this deterioration
Electrode characterization
mechanism is of very high practical importance for electrode
developers. Raman mappings of the graphite electrode were recorded (both
Finally, we wish to stress that our scientific focus was on the in situ during the electrochemical cycling and post mortem) using a
development of the method as described below. To this point, 10 × 10 matrix with a lateral working distances of 20 µm. For every
the Raman spectral characteristics were used as a tool to identify point of a Raman map five independent spectra were recorded to
the spots where the exfoliation occurred. Fundamental questions check for possible changes in the local structure of the graphite
related to the evolution of Raman spectra were out of the focus of electrode because of the thermal impact of the focused laser spot
this work. (filtered to about 0.9 mW).
Figure S1 (Supporting Information) shows two typical SEM
pictures of the surface of an exfoliated graphite electrode. As
Experimental standard SEM pictures do not offer the possibility to draw
a correlation between them and in situ Raman microscopy
Electrode and in situ cell preparation
results – with respect to the localization of exfoliated areas on
A standard graphite (90 wt.%, TIMREX SFG44, TIMCAL SA, Bodio, the same surface (and the same spot) of the electrode – a unique
Switzerland) was used as a model material to investigate its marker had to be found on the surface of the electrode. Laser
possible deterioration when used as an electrically conductive marking turned out to be a convenient option. Thus the same
additive in the positive electrodes of lithium-ion batteries. The optics could be used for the laser marking process and the
SFG44 sample was selected among many available graphites as an Raman characterization. To create these markers at the surface of
optimal model material, based on both our own X-ray diffraction the electrode in a reproducible way, three points were focused
(XRD)[50] data and also the spectroscopic work investigating the outside of the area of the Raman mapping below line no. 10 (in the
influence of different graphite types on the exfoliation behavior negative y direction and at positions 1, 2, and 7, in the x direction;
of graphite at different potentials. No electroactive oxide was Fig. 1). Normally, for about 4 min a laser power of about 4.5 mW
used in such model electrodes. Polyvinylidene fluoride (PVDF, was applied. The thermal impact on the surface of the electrode
8 wt.%, SOLEF 1015, Solvay) dissolved in N-methylpyrrolidinone with the concentrated spot of laser radiation burned visible marker
(NMP, Fluka) was used as a binder, and the surfactant Triton X- spots (holes) on the surface with known coordinates. Figure S2
305 (2 wt.%, Sigma-Aldrich) was used to avoid agglomerations shows a representative SEM picture of such a burned hole. These
of graphite particles during the preparation of the electrodes. three thermally created holes could be addressed again in the post
The ratio of the materials in the model electrodes was 90 : 8 : 2 mortem recorded pictures from the SEM. Thus, the same area of the
(graphite/PVDF/surfactant, by mass after drying). To prepare a surface of the electrode was characterized by Raman microscopy
suspension, all used materials were mixed with an ultrasonic stirrer and SEM. Note that the coordinates of all surface features are
(Hielscher UP200H, Teltow, Germany) for 2 min at 100% amplitude known because of the fact that the same objective of the Raman
without pulsing. The suspension was sliced with a doctor-blade on microscope is used for characterization and marking and therefore
an aluminum foil and then dried under vacuum at 100 ◦ C overnight the in situ Raman measurements and the markings have the same
to achieve a final ‘dry’ height of 250 µm of the electrode mass. coordinate system.
Afterwards circular working electrodes with a diameter of 13 mm After the in situ Raman mapping using the 10 × 10 matrix was
were punched out and dried further under vacuum at 120 ◦ C completed, six additional points were chosen on two different
overnight. Then, in an Ar-filled glove box with less than 1 ppm graphite particles (distance approximately 300 µm, three points
of oxygen, nitrogen and water, the in situ Raman cell[17,51 – 53] was on each particle) to acquire independent information from two
assembled and hermetically sealed. A fiber glass separator (1 mm defined particles and to verify the results of the 10 × 10 in situ
thick) soaked with 300 µl of electrolyte [1 M LiClO4 in ethylene Raman mapping.
carbonate (EC)/dimethyl carbonate (DMC) 1 : 1] was used. LiClO4 After dismounting the electrodes of the in situ Raman cell,
was used because of its less distracting influence on the in situ the electrodes were washed with pure EC/DMC 1 : 1 and then
Raman signal.[13,51,54] Metallic lithium served as both the reference vacuum-dried. SEM images were then recorded using either a
and counter electrode. In situ Raman spectra were collected at Zeiss LEO440, SE1 detector, 3 kV, or Zeiss Ultra, SE2 or InLens
25 ◦ C through a 200-µm thick glass window with a confocal detector, 3 kV, to obtain a post mortem overview of the surface of
Raman microscope (LabRAM HR, HoribaJobinYvon SA, Bensheim, the working electrode. As the laser-burned markers on the surface
Germany), using a He–Ne laser as exciting source (632.8 nm, 18 of the electrode were visible, the local coordinates of every single
mW). The laser light was focused through a small hole (d ≈ 1 mm, SEM image could be attached to the data files describing the SEM
punched before coating) in the current collector onto the surface images. For larger areas, an SEM mapping was performed. With the
of the graphite electrode with an 80× objective (Olympus) and coordinate information and the computer software ImageJ JAVA,
filtered to 0.9 mW (measured with a meter: Coherent FieldMate, all surface images could be evaluated and correlated with the
Santa Clara, CA, USA) during measurement and 0.13 mW during Raman spectral information (using Cobol85 script, function-based
focusing (built-in filter dial), to avoid thermal degradation. The communication with Raman software). An additional software
1755

electrodes were galvanostatically cycled from 2.0 to 5.5 V vs Li/Li+ package was written (Cobol85 script, file-based communication

J. Raman Spectrosc. 2011, 42, 1754–1760 Copyright 


c 2011 John Wiley & Sons, Ltd. wileyonlinelibrary.com/journal/jrs
A. Hintennach and P. Novák

Figure 1. Schematic representation of the arrangement of the in situ


Raman mapping technique. The measurement points within the shown
grid (an exemplary width of 10 × 10 points was chosen) are regular Figure 2. Exemplary post mortem Raman spectra of two arbitrarily selected
measurement points, whereas the three additional points are laser-burned points on a SFG44 graphite electrode. The electrode was electrochemically
markers on the surface of the electrode. cycled up to very positive potentials (2.0–5.5 V vs Li/Li+ , 20 cycles, 1 M
LiClO4 in EC/DMC 1 : 1). (a) The E band (exfoliation band) detected at the
exfoliated parts of the graphite electrode and (b) the G band (with the
port) to allow the automated evaluation of the spectra data sets weak shoulder of the D band) seen at the intact parts of the electrode.
with respect to band shifts and the ratio of the integrals of
the D, G, and E bands (see Section on Results). A subsequent E2g2 (b) could be observed.[4,55 – 60] If the graphite were intact, the
correlation of the SEM pictures with this Raman information typical G and D bands would be seen in the same wavenumber
permitted automated colorization of the electrode SEM pictures region [spectrum (b) of Fig. 2],[13,61] and in an in situ experiment
with respect to, e.g. the exfoliated areas. Different γ values (relation the Raman spectrum development follows the pattern known for
of the proportion of black to white color) of the monochromic anion intercalation.
information set in the second layer of the picture were used. A The respective bands around 1600 cm−1 can be used as a
following analysis of the false color information of the image was spectral fingerprint for the semiautomatic search for exfoliated
performed with standard image evaluation routines (ImageJ JAVA) areas of the electrode. Figure 3 shows a typical post mortem SEM
to evaluate the ratio of exfoliated areas to non-exfoliated areas. image with two representative in situ Raman spectra from different
Afterwards, an automated statistical analysis of the surface of the locations. Using the above-described coordinate information,
electrode was executed (S+ , Tibco Software Inc.). To avoid double changes in the ratio of the intensities of D , G, and E bands
counting of exfoliated areas, the distance of every single exfoliated of the local Raman spectra of the graphite electrodes can be
area at the surface of the electrode to all surrounding exfoliated evaluated and correlated with the surface morphology: at the
areas of the same electrode is included in a file describing the exfoliated areas on the surface of the graphite electrode, the
surface of the electrode with unique data sets. The size of the Raman spectrum features E bands instead of the G band.
total area of exfoliated parts of the electrode is determined by The change with potential in the magnitudes of the integrals
the number of pixels of the detector of the SEM detecting an of both parts of this E band is significantly different from the
exfoliated area (cf Fig. S3). change in the magnitude of the integrals of the fitted D and G
bands (Fig. 4). Note that the above statement is valid only after the
graphite electrode was partially exfoliated by potential excursions
Results and Discussion up to 5.5 V.
The assignment of the overlapping D and E bands of the
Highly exfoliated particles of graphite show a typical accordion- Raman spectrum can be performed as follows: As the shoulder
like morphology (Fig. S1). The Raman spectrum of the exfoliated of the G band (the D band) decreases in intensity with a
graphite shows a typical band called the E band (‘exfoliation progressing intercalation of anions into the graphite, the E-band
band’, Fig. 2). The wavenumber of the spectral signature of the feature remains visible as a result of the irreversible structural
E band is different from that of the G band of intact graphite changes caused by the exfoliation. The irreversible exfoliation
with intercalated ions. This is due to the different environments of the graphite material can be indeed detected post mortem
of the respective graphene layers – the presence of solvated ions spectroscopically after complete electrochemical cycling. It was
adjacent to the surface of the exfoliated particles is the obvious shown in an earlier work[13] for very reductive potentials (i.e. the
reason. intercalation of lithium and the reductive exfoliation of graphite)
When the G band starts to broaden with increasing electrode that the D and E bands of the Raman spectra of graphite electrodes
potential and two side bands begin to grow and become distinct can be clearly distinguished by using the additional information
bands, these additional peaks can be assigned to the intensities of of the ratio of the magnitude of the D and D bands. As the
the interior E2g2 (i) (left part) and bounding E2g2 (b) (right part) same behavior was observed for the anion intercalation at very
bands of the exfoliated graphite.[13] Owing to changed C–C positive potentials, the assignment was performed here in an
1756

bond force constants because of the exfoliation, the shift of the analogous way.

wileyonlinelibrary.com/journal/jrs Copyright 
c 2011 John Wiley & Sons, Ltd. J. Raman Spectrosc. 2011, 42, 1754–1760
Raman and SEM mapping method

Figure 3. Exemplary representation of two selected measurement points


for the combination of the in situ Raman mapping technique with a
post mortem recorded SEM image (Zeiss LEO440, 2 kV, SE1 detector). The Figure 5. Shift with the potential of the D band of the in situ Raman signal
appropriate G- and E-band region is shown in the assigned Raman spectra of a graphite electrode (2.0–5.5 V vs Li/Li+ , 1 M LiClO4 in EC : DMC 1 : 1)
for the exfoliated and the non-exfoliated areas on the surface of the after ten cycles.
electrode. Electrochemical cycling was performed between 2.0 and 5.5 V
(vs Li/Li+ , 1 M LiClO4 in EC/DMC 1 : 1).

Figure 4. Typical changes with potential of the in situ Raman spectrum of


a graphite electrode during electrochemical cycling up to very positive Figure 6. Shift with the potential of the D band of the in situ Raman signal
potentials (2.0–5.5 V vs Li/Li+ , 1 M LiClO4 in EC/DMC 1 : 1) after ten cycles. of a graphite electrode after ten cycles. Electrolyte 1 M LiClO4 in EC/DMC
The relation of the integrals of the G and E bands is used to identify the 1 : 1. An exemplary standard deviation is shown for the potential of 4.8 V
exfoliation process. (vs Li/Li+ ).

Interestingly, in the in situ Raman spectrum of graphite surface disorder increases. Consequently, the D band is more
electrodes, a significant and reversible shift of the D band with intense. As the shift with potential of the D band is reversible
the potential could be observed in areas where the E band (Fig. 6), it can be used in addition to the ratio of the integrals
was detected, i.e. where exfoliation occurred (Figs 5 and 6). The of the E and G bands to define a new marker, which will
magnitude of the D-band shift in areas without exfoliation was help in localizing the exfoliation process of graphite. Therefore,
less than half that in areas with detected exfoliation. The D band the semiautomated localization of the areas of the electrode
shifts between 1322 and 1346 cm−1 for increasing potentials where exfoliation occurred is facilitated straightforwardly using
between 3.0 and 5.5 V, with the most significant shift around the described full set of spectral information (shift of the D
4.8 V (vs Li/Li+ ). Obviously, the intensity of the Raman D band band, ratio of the integrals of D and G bands, ratio of the
increases with an increasing degree of exfoliation. This is not integrals of G and E bands, and the post mortem detection of
surprising because the origin of the D band lies in defects the E band after electrochemical cycling). In short, intact areas
in the graphite structure and in prismatic surfaces. It is clear of the graphite electrode intercalate and de-intercalate anions
1757

that exfoliation increases the number of defects and, thus, the and therefore behave reversibly during cycling and consequently

J. Raman Spectrosc. 2011, 42, 1754–1760 Copyright 


c 2011 John Wiley & Sons, Ltd. wileyonlinelibrary.com/journal/jrs
A. Hintennach and P. Novák

Figure 7. (a) Two typical mappings from in situ Raman experiments (10 × 10 mapping points, 250 µm × 250 µm) of a SFG44 graphite electrode during
electrochemical cycling (2.0–5.5 V vs Li/Li+ , 1 M LiClO4 in EC/DMC 1 : 1). Light gray colors indicate non-exfoliated areas, whereas gray and dark gray colors
indicate exfoliated areas. (b) Distribution of the number of exfoliated areas and the total sum of these areas.

show low fluctuation of the spectral data, whereas the exfoliated small areas forming together a larger area, a minimum angle can
(deteriorated) areas show high fluctuation of the spectral data be defined. This is symbolized in Fig. S4 by α ij .
during cycling. Automated Raman mapping techniques created the maps
The post mortem SEM images were evaluated as follows: The shown in the top row of Fig. 7. Different colors indicate exfoliated
points on the electrode surface correspond to the pixels of the and non-exfoliated parts with a given lateral resolution. After a
digital image. The pixels were treated as a grid, as schematically statistical evaluation of the number and size of exfoliated and non-
shown in Fig. S3. Each square represents a pixel of the false color exfoliated areas on the surface of the electrode, a quantification
images to be generated after processing the SEM pictures and could be drawn as shown in the bottom row of Fig. 7. This
the in situ Raman data. However, as the ability to find exfoliated analysis revealed that exfoliation leading to the degradation of
areas on the surface of the electrodes is limited by the effective the electrode occurred irregularly on some spots of the electrode
range of the standard deviation calculated from all the involved when exposed to highly positive potentials (>5 V vs Li/Li+ ). Indeed,
the exfoliation was confirmed by the examination of the respective
parameters, the total resolution of the SEM images was reduced
areas of the SEM pictures. Moreover, the assignment of the E band
accordingly by interconnecting defined pixel areas (grid cell:
of the in situ Raman spectrum was confirmed by the quantitative
in the example of Fig. S3, 2 × 2 pixels were interconnected,
and qualitative analysis of the post mortem SEM images. Figure 8
represented in the scheme by the gray area in the middle of the
shows a typical result confirming the quality of the correlation of
total square). the Raman and SEM result.
Figure S4 shows schematically how the exfoliated particles were Figure S5 shows that the proper selection of the grid size is
classified. The electrode area is described by the respective Raman essential. The automated recognition of exfoliated particles by
spectra. An automated analysis was performed with respect to the described statistical analysis is influenced by the defined grid
the fluctuations in the set of spectral parameters for D-band cell size used for the data analysis. The reason is that the useful
positions and G/E band ratio of intensities and maxima positions. lateral resolution of the Raman spectral information is related
An increasing deviation from a calculated average was assumed to to the surface morphology, especially to the coarseness of the
indicate exfoliated parts on the surface of the electrode. To every surface of the electrode. The coarseness of the surface depends
found exfoliated area (‘considered particle’) in Fig. S4, the area and on the degree of the exfoliation of graphite, the type of the used
the angles of all surrounding other exfoliated areas were assigned, graphite material, and its particle size distribution. The angle of
1758

resulting thus in a unique address. To avoid multiple counting of incidence of the laser beam is also relevant. If only single pixels

wileyonlinelibrary.com/journal/jrs Copyright 
c 2011 John Wiley & Sons, Ltd. J. Raman Spectrosc. 2011, 42, 1754–1760
Raman and SEM mapping method

(PSI and ETH Zurich) are greatly appreciated. We wish to thank Dr


F. Simmen for her help with the SEM images.

Supporting information
Supporting information may be found in the online version of this
article.

References
[1] J. M. Tarascon, M. Armand, Nature 2001, 414, 359.
[2] M. Armand, J. M. Tarascon, Nature 2008, 451, 652.
[3] P. G. Bruce, B. Scrosati, J. M. Tarascon, Angew. Chem. Int. Ed. 2008,
47, 2930.
[4] M. Winter, J. O. Besenhard, M. E. Spahr, P. Novák, Adv. Mater. 1998,
10, 725.
[5] E. Karden, P. Shinn, P. Bostock, J. Cunningham, E. Schoultz, D. Kok,
J. Power Sources 2005, 144, 505.
Figure 8. Exemplary representation of a post mortem recorded SEM image [6] J. Milliken, F. Joseck, M. Wang, E. Yuzugullu, J. Power Sources 2007,
(Zeiss LEO440, 2 kV, SE1 detector) with attached false color information 172, 121.
highlighting the exfoliated areas. Electrochemical cycling was performed [7] P. Novák, F. Joho, M. Lanz, B. Rykart, J. C. Panitz, D. Alliata, R. Kötz,
between 2.0 and 5.5 V (vs Li/Li+ , 1 M LiClO4 in EC/DMC 1 : 1). O. Haas, J. Power Sources 2001, 97 –98, 39.
[8] J. Yang, Y. Takeda, N. Imanishi, O. Yamamoto, Recent Res. Develop.
Solid State Ionics 2003, 1, 1.
of the final false color images are evaluated to maintain high [9] M. J. Zhao, G. S. Cao, J. Xie, X. B. Zhao, Rare Metal Mater. Eng. 2005,
34, 1857.
resolution, weak peaks might be too weak with respect to the [10] W. Chen, G. Wu, C. Wang, H. Yang, P. He, X. Zhang, Z. Xu, W. Li,
noise level and therefore not detected. This is why a limit of the Y. Yang, Huaxue Wuli Xuebao 2001, 14, 88.
minimum of the grid cell size has to be defined to ensure that [11] D. Aurbach, M. Koltypin, H. Teller, Langmuir 2002, 18, 9000.
the fluctuation of the signal of the Raman spectra can be evaluated [12] T. Abe, N. Kawabata, Y. Mizutani, M. Inaba, Z. Ogumi, J. Electrochem.
Soc. 2003, 150, A257.
using statistical means (i.e. without a positive or negative tendency
[13] L. J. Hardwick, H. Buqa, M. Holzapfel, W. Scheifele, F. Krumeich,
of the fluctuation of the Raman signal because of the coarseness P. Novák, Electrochim. Acta 2007, 52, 4884.
of the electrode). [14] M. E. Spahr, H. Buqa, A. Würsig, D. Goers, L. J. Hardwick, P. Novák,
Finally, Fig. S6 shows the typical development of an in situ Raman F. Krumeich, J. Dentzer, C. Vix-Guterl, J. Power Sources 2006, 153,
spectrum with the cycle number. Typically, the E band (exfoliation 300.
[15] M. E. Spahr, H. Wilhelm, T. Palladino, N. Dupont-Pavlovsky, D. Goers,
band) evolves with the cycle number and then remains constant F. Joho, P. Novák, J. Power Sources 2003, 119–121, 543.
over additional cycles. This can be rationalized by the fact that [16] M. E. Spahr, H. Wilhelm, F. Joho, J. C. Panitz, J. Wambach, P. Novák,
every excursion into the very positive potential region increases N. Dupont-Pavlovsky, J. Electrochem. Soc. 2002, 149, A960.
the deterioration of the electrode because of the exfoliation until [17] P. Novák, F. Joho, R. Imhof, J. C. Panitz, O. Haas, J. Power Sources
1999, 81–82, 212.
a ‘final destructed steady state’ is reached. Figure S7 shows typical [18] H. Buqa, A. Würsig, J. Vetter, M. E. Spahr, F. Krumeich, P. Novák,
voltage profiles of graphite electrodes cycled up to such very J. Power Sources 2006, 153, 385.
positive potentials at constant charge and discharge currents, [19] P. Novák, J. Ufheil, H. Buqa, F. Krumeich, M. E. Spahr, D. Goers,
respectively. The evolution of the voltage profiles with time H. Wilhelm, J. Dentzer, R. Gadiou, C. Vix-Guterl, J. Power Sources
corresponds to what is expected and can be compared with 2007, 174, 1082.
[20] H. Buqa, A. Würsig, A. Goers, L. J. Hardwick, M. Holzapfel, P. Novák,
the data published by us and others.[39,50,62] The reader is referred F. Krumeich, M. E. Spahr, J. Power Sources 2005, 146, 134.
to our related work for a more detailed discussion.[50,62] [21] L. J. Hardwick, H. Buqa, P. Novák, Solid State Ionics 2006, 177, 2801.
[22] L. J. Hardwick, M. Hahn, P. Ruch, M. Holzapfel, W. Scheifele, H. Buqa,
F. Krumeich, P. Novák, R. Kötz, Electrochim. Acta 2006, 52, 675.
[23] C. Vix-Guterl, M. Couzi, J. Dentzer, M. Trinquecoste, P. Delhaes,
Conclusions J. Phys. Chem. B 2004, 108, 19361.
[24] P. Delhaes, M. Couzi, M. Trinquecoste, J. Dentzer, H. Hamidou,
In this work, a novel Raman microscopy-based semiautomatic C. Vix-Guterl, Carbon 2006, 44, 3005.
method for detection of exfoliated graphite particles in battery [25] L. J. Rendek, G. S. Chottiner, D. A. Scherson, Electrochem. Solid State
electrodes was developed. In an example, exfoliated areas of Lett. 2002, 5, A77.
[26] L. J. Rendek, G. S. Chottiner, D. A. Scherson, J. Electrochem. Soc. 2002,
model TIMREX SFG44 graphite electrodes were semiautomatically
149, E408.
detected after repeated excursions to very positive potentials [27] L. J. Rendek, G. S. Chottiner, D. A. Scherson, Langmuir 2002, 18,
(>5 V vs Li/Li+ ) with in situ Raman mapping and post mortem 6554.
SEM. The results of both techniques were combined using a new [28] Y. Luo, W. B. Cai, D. A. Scherson, J. Electrochem. Soc. 2002, 149,
correlating method for the data analysis. Finally, the Raman E A1100.
[29] Y. Luo, W. B. Cai, X. K. Xing, D. A. Scherson, Electrochem. Solid State
band describing the exfoliation process was detected at exfoliated Lett. 2004, 7, E1.
graphite particles. [30] L. J. Hardwick, P. Novák, GDCh-Monogr. 2005, 32, 220.
[31] L. J. Hardwick, M. Holzapfel, A. Wokaun, P. Novák, J. Raman
Spectrosc. 2007, 38, 110.
Acknowledgements [32] D. Goers, H. Buqa, L. J. Hardwick, A. Wuersig, P. Novák, Ionics 2003,
9, 258.
Technical assistance of Mr W. Scheifele (PSI) and Mr H. Kaiser (PSI) [33] R. Baddour-Hadjean, J. P. Pereira-Ramos, J. Power Sources 2007, 174,
1759

is highly appreciated. Scientific discussions with Prof A. Wokaun 1188.

J. Raman Spectrosc. 2011, 42, 1754–1760 Copyright 


c 2011 John Wiley & Sons, Ltd. wileyonlinelibrary.com/journal/jrs
A. Hintennach and P. Novák

[34] M. Kerlau, M. Marcinek, V. Srinivasan, R. M. Kostecki, Electrochim. [49] C. Buhrmester, L. M. Moshurchak, R. L. Wang, J. R. Dahn,
Acta 2007, 53, 1385. J. Electrochem. Soc. 2006, 153, A1800.
[35] R. Kostecki, J. L. Lei, F. McLarnon, J. Shim, K. Striebel, J. Electrochem. [50] W. Märkle, N. Tran, D. Goers, M. E. Spahr, P. Novák, Carbon 2009, 47,
Soc. 2006, 153, A669. 2727.
[36] R. Kostecki, F. McLarnon, Electrochem. Solid State Lett. 2004, 7, A380. [51] J. C. Panitz, F. Joho, P. Novák, Appl. Spectrosc. 1999, 53, 1188.
[37] J. L. Lei, F. McLarnon, R. Kostecki, J. Phys. Chem. B 2005, 109, 952. [52] P. Novák, J. C. Panitz, F. Joho, M. Lanz, R. Imhof, M. Coluccia, J. Power
[38] M. E. Spahr, T. Palladino, H. Wilhelm, A. Würsig, D. Goers, H. Buqa, Sources 2000, 90, 52.
M. Holzapfel, P. Novák, J. Electrochem. Soc. 2004, 151, A1383. [53] J. C. Panitz, P. Novák, O. Haas, Appl. Spectrosc. 2001, 55, 1131.
[39] J. A. Seel, J. R. Dahn, J. Electrochem. Soc. 2000, 147, 892. [54] A. Würsig, H. Buqa, M. Holzapfel, F. Krumeich, P. Novák, Electrochem.
[40] Y. Luo, W.-B. Cai, D. A. Scherson, J. Electrochem. Soc. 2002, 149, Solid State Lett. 2005, 8, A34.
A1100. [55] M. Inaba, H. Yoshida, Z. Ogumi, T. Abe, Y. Mizutani, M. Asano,
[41] M. M. Doeff, J. D. Wilcox, R. Yu, A. Aumentado, M. Marcinek, J. Electrochem. Soc. 1995, 142, 20.
R. Kostecki, J. Solid State Electrochem. 2008, 12, 995. [56] S. A. Solin, GraphiteIntercalationCompounds, Springer Verlag: Berlin,
[42] L. J. Hardwick, M. Marcinek, L. Beer, J. B. Kerr, R. Kostecki, 1990.
J. Electrochem. Soc. 2008, 155, A442. [57] M. Inaba, H. Yoshida, Z. Ogumi, J. Electrochem. Soc. 1996, 143, 2572.
[43] R. Kostecki, F. McLarnon, J. Power Sources 2003, 119–121, 550. [58] Z. Wang, X. Huang, R. Xue, L. Chen, Carbon 1999, 37, 685.
[44] K. A. Striebel, J. Shim, E. J. Cairns, R. Kostecki, Y. J. Lee, J. Reimer, [59] M. Endo, C. Kim, T. Karaki, T. Fujino, M. J. Matthews, S. D. M. Brown,
T. J. Richardson, P. N. Ross, X. Song, G. V. Zhuang, J. Electrochem. M. S. Dresslhaus, Synth. Met. 1998, 98, 17.
Soc. 2004, 151, A857. [60] T. Itoh, H. Sato, T. Nishina, T. Matue, I. Uchida, J. Power Sources 1997,
[45] K. A. Striebel, A. Sierra, J. Shim, C. W. Wang, A. M. Sastry, J. Power 68, 333.
Sources 2004, 134, 241. [61] M. A. Pimenta, G. Dresselhaus, M. S. Dresselhaus, L. G. Cancado,
[46] C. W. Wang, Y. B. Yi, A. M. Sastry, J. Shim, K. A. Striebel, J. Electrochem. A. Jorio, R. Saito, Phys. Chem. Chem. Phys. 2007, 9, 1276.
Soc. 2004, 151, A1489. [62] W. Märkle, J. F. Colin, D. Goers, M. E. Spahr, P. Novák, Electrochim.
[47] J. D. Wilcox, M. M. Doeff, M. Marcinek, R. Kostecki, J. Electrochem. Acta 2010, 55, 4964.
Soc. 2007, 154, A389.
[48] M. M. Doeff, Y. Hu, F. McLarnon, R. Kostecki, Electrochem. Solid-State
Lett. 2003, 6, A207.
1760

wileyonlinelibrary.com/journal/jrs Copyright 
c 2011 John Wiley & Sons, Ltd. J. Raman Spectrosc. 2011, 42, 1754–1760

You might also like