Subscriber access provided by Western Michigan University
Article
Polynanocrystalline Graphite: A New Carbon Anode with Superior Cycling Performance for K-ion Batteries Zhenyu Xing, Yitong Qi, Zelang Jian, and Xiulei Ji ACS Appl. Mater. Interfaces, Just Accepted Manuscript • DOI: 10.1021/acsami.6b06767 • Publication Date (Web): 03 Aug 2016 Downloaded from http://pubs.acs.org on August 7, 2016
Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.
ACS Applied Materials & Interfaces is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.
Page 1 of 25
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
ACS Applied Materials & Interfaces
Polynanocrystalline Graphite: A New Carbon Anode with Superior Cycling Performance for K-ion Batteries Zhenyu Xing, Yitong Qi, Zelang Jian and Xiulei Ji* Department of Chemistry, Oregon State University, Corvallis, OR, 97331, USA
Abstract We synthesized a new type of carbon—polynanocrystalline graphite—by chemical vapor deposition on a nanoporous graphenic carbon as an epitaxial template.
This
carbon is composed of nanodomains being highly graphitic along c-axis and very graphenic along ab plane directions, where the nanodomains are randomly packed to form micron-sized particles, thus forming a polynanocrystalline structure. polynanocrystalline
graphite
is
very
unique,
structurally
different
The from
low-dimensional nanocrystalline carbon materials, e.g. fullerenes, carbon nanotubes and graphene, nanoporous carbon, amorphous carbon and graphite, where it has a relatively low specific surface area of 91 m2/g as well as a low Archimedes density of 0.92 g/cc.
The structure is essentially hollow to a certain extent with randomly
arranged nanosized graphite building blocks.
This novel structure with disorder at
nanometric scales but strict order at atomic scales enables substantially superior long-term cycling life for K-ion storage as an anode, where it exhibits 50% capacity
1
ACS Paragon Plus Environment
ACS Applied Materials & Interfaces
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 2 of 25
retention over 240 cycles, whereas, for graphite, it is only 6% retention over 140 cycles.
Keywords Polynanocrystalline graphite; Disorder, K-ion batteries; Anode; Cycling life
Introduction Li-ion batteries (LIBs) have gained great successes as the power source in portable electronics and electric vehicles due to their high energy density.1-8
However, the
scarcity and uneven global distribution of lithium reserve raise the concern of long-term sustainability for LIBs.9,10
Furthermore, there exists tremendous demand
for inexpensive alternative energy storage technologies that can enable widespread deployment of intermittent renewable energy sources, such as solar and wind. Hence, it is critical that battery systems on the basis of Earth-abundant elements be explored and investigated.
To date, several battery technologies based on
Earth-abundant elements have been under intense investigation, such as Na-ion batteries (NIBs),11-35 Mg-ion batteries (MIBs),36-38 Al-ion batteries (AIBs),39,40 and K-ion batteries (KIBs).41-50
Potassium is the 7th most abundant element in Earth’s
crust, nearly 900 times richer than lithium.
The abundance potentially renders the
production cost of such batteries much lower compared to Li-batteries. Recently, our group and others reported that graphite can reversibly store K-ions electrochemically, which opens up the possibility of equipping KIBs with insertion-type carbon anodes.51-53 Additionally, our group further reported K-ion storage properties of 2
ACS Paragon Plus Environment
Page 3 of 25
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
ACS Applied Materials & Interfaces
non-graphitic carbons, including soft carbon and hard carbon.54
Among these bulk
carbon materials, at the same specific conditions, graphite suffers the fastest capacity fading, where it loses 50% of its capacity after 50 cycles at C/2 (1C-rate is defined as 279 mA/g), while the capacity retentions for soft carbon and hard carbon are 81% after 50 cycles at 2C and 83% after 100 cycles at C/10, respectively.
Here, we hypothesize that the superior cycling stability of hard carbon and soft carbon to graphite in KIBs is due to the disordered arrangement of turbostratic nanodomains in these carbons.
We postulate that it is the volume expansion (61%) of the compact
graphite structure with ABAB registry that is responsible for the fast capacity fading. To test such a hypothesis, herein, we design a new type of carbon material for K-ion storage, which resembles the non-graphitic carbon structures at long-range scales but maintains a highly ordered graphitic registry locally for a few nanometers along the c-axis.
Namely, ordered graphitic nanodomains instead of turbostratic nanodomains
are randomly oriented in bulk-sized carbon particles.
Such a carbon would be,
indeed, short-range ordered but long-range disordered polynanocrystalline graphite. The closest analogue to our proposed carbon structure that exists would be some special hard carbons if they are annealed up to a very high temperature, i.e., > 2000 ºC. However, such hard carbon, or often referred to as “glassy carbon”, is not locally graphitic as it still does not have the short-range order along the c-axis 55.
In this contribution, we report a facile synthesis of polynanocrystalline graphite by
3
ACS Paragon Plus Environment
ACS Applied Materials & Interfaces
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 4 of 25
chemical vapor deposition (CVD) using ethanol vapor as a precursor onto the surface of nanoporous graphenic carbon (NG) as an epitaxial template.
NG was synthesized
by our previously reported method, where a spontaneous metallothermic reduction of gaseous CO2 takes place on an Mg/Zn mixture as the reductant at high temperatures.56 We select NG as the template because CVD carbon deposition may occur toward different directions at nanometric scales on NG’s highly curved graphenic surface in the nanopores, thus generating a polynanocrystalline graphite.
The resulting
polynanocrystalline graphite we prepared, hereafter referred to as PG, shows very stable cycling life as an anode in KIBs with 50% of capacity retention over 240 cycles, in contrast to 6% over only 140 cycles by the commercial graphite.
The improved
cycling stability of PG is attributed to the random arrangements of graphitic nanodomains, which better accommodates the strain caused by K-ion insertion during cycling.
Experiment section Preparation of PG and NG NG is prepared following the method we reported before.1
To prepare PG by CVD
of carbon on NG, we place NG powder in an Al2O3 boat that is heated in a tube furnace at 1100 ºC for 20 h under ethanol vapor carried by Argon gas at a flow rate of 90 CCM. Characterization Methods: X-ray diffraction (XRD) patterns were recorded using a Rigaku Ultima IV 4
ACS Paragon Plus Environment
Page 5 of 25
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
ACS Applied Materials & Interfaces
Diffractometer with Cu Kα irradiation (λ= 1.5406 Å).
WITec confocal Raman
spectrometer with a 514 nm laser source was used to record the Raman spectra.
The
morphology was studied by field emission scanning electron microscopy (FESEM) using a Hitachi S5500 at 0.5 kV.
Transmission electron microscopy (TEM) images
were collected by FEI Titan 80-200.
Nitrogen sorption measurements were
performed on Micromeritics TriStar II 3020 analyzer.
Electrochemical Measurements: All the electrodes consist of 80 wt% carbon active mass, 10 wt% carbon black (C45) and 10 wt% sodium carboxymethyl cellulose (CMC) as the binder. The slurry was cast onto Cu foil and dried at 80 °C under vacuum for 12 h.
Coin cells were
assembled in an argon-filled glove box, with the as-prepared carbon electrode as the working electrode, potassium metal as the counter/reference electrode and glass fiber membrane as a separator.
The electrolyte was prepared by dissolving 0.8 M KPF6
into a mixture of ethylene carbonate (EC) and diethyl carbonate (DEC) (volume ratio: 1/1).
The active mass loading was around 2 mg/cm2.
The electrochemical
measurements were performed on an Arbin BT2000 system at room temperature, where the voltage range was from 0.01 to 2 V versus K+/K.
Results and discussion A polynanocrystalline material should exhibit a relatively low surface area and large particle sizes, which differentiates such a material from nanoporous materials or nanocrystalline materials.
Despite the low surface area and large particle sizes, its
5
ACS Paragon Plus Environment
ACS Applied Materials & Interfaces
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 6 of 25
density should be much lower than the bulk single crystalline or polycrystalline analogs.
After the CVD treatment of NG to form PG, the Brunauer-Emmett-Teller
(BET) surface area dramatically decreases from 1900 to 91 m2/g.
A comparison of
the isotherms of NG (Type II) and PG (Type III) in Figure S1a and c shows that micropores and mesopores of NG observed at low and intermediate relative pressures, respectively, are nearly filled up after CVD.
By considering the scales of absorbed
volumes (y-axis), after CVD PG only maintains a small portion of macropores observed at high relative pressures (Figure S1b,d).
SEM images illustrate the macroscopic morphology change from NG to PG, where NG exhibits an irregular hollow semi-spherical morphology with very smooth surface (Figure 1a,b), whereas PG particles are of very solid texture with much rougher surface after CVD (Figure 1c,d).
Furthermore, we measured the Archimedes
density of NG and the seemingly “non-porous” PG, where the density of PG is only 0.95 g/cm3, compared with 0.54 g/cm3 for NG.
Considering that graphite’s density
is above 2 g/cm3, one can gain from the above results two important insights: (1) the CVD carbon deposition seals the nanoporosity of NG effectively, resulting in a low surface area of PG; (2) PG contains a hollow structure, but the internal structure is clearly not accessible to N2 molecules during sorption measurements.
6
ACS Paragon Plus Environment
Page 7 of 25
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
ACS Applied Materials & Interfaces
Figure 1. SEM images of NG (a) and PG (c). (b) and (d) are the enlarged sections marked in the yellow boxes in (a) and (c).
To characterize the structures of NG and PG, we conducted x-ray diffraction (XRD), transmission electron microscopy (TEM) as well as Raman spectroscopy measurements.
The goal is to determine whether PG exhibits a high degree of local
order but much disorder in nanometric or longer scales.
As shown in the XRD
patterns (Figure 2a), PG exhibits a much sharper (002) peak with a smaller value of full width at half maximum (FWHM) and a more pronounced (100) peak, compared to NG. PG.
The broad bump to the left of (002) peak of NG is completely vanished in
The d(002) of PG is calculated to be 0.345 nm, which is fairly close to that of
7
ACS Paragon Plus Environment
ACS Applied Materials & Interfaces
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
graphite—0.335 nm.
Page 8 of 25
Based on the Scherrer equation, the coherence lengths: Lc
along c axis and La along ab planes are calculated to be 1.70 nm and 6.86 nm, respectively, for NG, and 3.71 nm and 8.42 nm, respectively, for PG.
The XRD
results showing a longer range of order along c-axis and ab planes indicate that PG comprises more graphitic and more graphenic domains than NG.
However, the
coherence range is still less than 10 nm, where the graphitic/graphenic domains in PG are, in fact, nanocrystalline; when considering the bulk particle size and low surface area of PG, PG is essentially polynanocrystalline.
Figure 2. (a,b) XRD patterns and Raman spectra of NG, PG and graphite (G).
Fitted
Raman data and their deconvolution into TPA (red), D (blue), A (green) and G (pink) bands for NG (c) and PG (d).
As shown in Figure 2b, graphite, PG and NG all give rise to both G band and D band, 8
ACS Paragon Plus Environment
Page 9 of 25
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
ACS Applied Materials & Interfaces
where the G-band corresponds to the sp2 carbon with E2g vibration, and the D-band is attributed to the A1g vibration of finite-sized graphenic domains.
To further analyze
the Raman results, we fit the experimental spectra of PG and NG by using TPA (transpolyacetylene) band, D band, A band, and G band, shown in Figure 2c and d.57 The TPA band can be observed as the shoulder to the left of the D band, and the A band contributes to the intensity between G band and D band.58-61
The peak
positions of G band for NG and PG are 1596 cm-1 and 1590 cm-1, respectively.
Here,
the redshift of G band is attributed to the increased graphenic degree, as revealed in the XRD results, which agrees with the trend reported before.62,63
The presence of A
band is commonly observed from disordered or non-graphitic carbon, where this band is attributed to point defects and/or graphene curvature.64,65
Interestingly, the fairly
pronounced A band in NG completely disappears in PG, clearly demonstrating the transition from a locally defective carbon to a much more graphenic carbon.
The
TPA band that is often attributed to transpolyacetylene-like structure also diminishes from NG to PG, also showing a better degree of order.66
The 2D band, the overtone
of D band, whose existence requires vibration from defect-free graphene domains, become much more pronounced from NG to PG, where the 2D band of PG shares the same wavenumber as graphite, further confirming the good extent of order along ab planes
67
.
The ratio between the integrated intensities of the G and D Raman bands
(IG/ID) along with the excitation laser wavelength are used to calculate the graphenic domain size, La, namely the coherence size along ab planes. The general equation correlating these parameters was introduced by Cancado et al.,68 as shown below:
9
ACS Paragon Plus Environment
ACS Applied Materials & Interfaces
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
2.4 ⋅ 10 ⋅ where the excitation laser wavelength (λnm) is 514 nm. The IG/ID value is 0.25 for NG and 0.81 for PG, corresponding to La values of 4.18 nm and 13.56 nm for NG and PG, respectively.
The calculated La values from Raman
spectra further confirm the enhanced local order of PG compared to NG.
Figure 3.
HRTEM images of NG (a) and PG (b).
Inset of (b) is the SAED pattern
of PG. 10
ACS Paragon Plus Environment
Page 10 of 25
Page 11 of 25
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
ACS Applied Materials & Interfaces
High-resolution
TEM
studies
provide
unequivocal
evidence
for
the
polynanocrystalline structure of PG, which is short-range ordered but long-range disordered.
As shown in Figure 3b, PG possesses vividly displayed lattice fringes
of graphitic nanodomains, where these domains are packed along random orientations, vastly different from the long-range ordered structure of graphite and completely disordered structure of amorphous carbon, i.e., hard carbon, indicating a unique polynanocrytalline structure.
Between these nanodomains, there are tiny voids,
which explain the low density of this material.
To provide a more complete
landscape for the unique structure of this polynanocrystalline graphite, additional TEM images are shown in Figure S2.
The formation of PG’s structure is attributed to the porosity and local atomic arrangement of NG as a CVD template, which exhibits a highly tortuous porous structure with graphenic walls (Figure 3a) and successfully serves as an epitaxial substrate for the growth of PG’s structure.
The polynanocrystalline structure of PG
is further confirmed by selected area electron diffraction (SAED), where the resolved diffraction rings indicate the isotropic nature in nanometric scales, very different from the typical SAED patterns of graphite with preferential crystallite orientation (Figure S3) and amorphous carbon that exhibits poorly resolved diffraction rings.
11
ACS Paragon Plus Environment
ACS Applied Materials & Interfaces
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 12 of 25
Figure 4. (a) The first galvanostatic potassiation/depotassiation potential profiles of PG and Graphite in the potential range of 0.1–2 V vs. K+/K at a current rate of 20 mA/g.
(b) Rate performance for PG and Graphite at current densities from 20, 50,
100, 200, 500 to 1000 mA/g.
CV curves of the first three cycles for PG (c) and
Graphite (d).
For the electrodes with a potentially large volume change during cycling, such as silicon anode in LIBs, suitable binders can provide a structural buffer or a high adhesion force to retain the electrode’s integrity during cycling.
Sodium
carboxymethyl cellulose (CMC), also known as CMC, is a linear cellulose derivative, which is widely used in LIBs for silicon anodes.
In this study, we use CMC as the
binder for all carbon anodes, where their K-ion storage performance is studied.
12
ACS Paragon Plus Environment
As
Page 13 of 25
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
ACS Applied Materials & Interfaces
shown in Figure 4a, PG exhibits more sloping potassiation/depotassiation potential profiles compared to those of graphite. clearly higher than that of graphite.
The average potassiation potential of PG is
The difference that PG presents a sloping profile
while graphite shows a plateau shape is more evident during depotassiation.
The
plateau profiles of graphite are attributed to the evolving stages of K-graphite intercalation compounds (K-GICs), whereas potassiation and depotassiation of PG occur in a manner more like a solid-solution process due to the polynanocrystalline structure.
If the stage-one K-graphite intercalation compound (GIC), KC8, is formed, the theoretical capacity of graphite and polynanocrystalline graphite is 279 mAh/g.
In
the first cycle, PG delivers a potassiation capacity of 414 mAh/g and a depotassiation capacity of 224 mAh/g with a coulombic efficiency of 54.1 % lower than 78.0 % of graphite, where PG exhibits a high level of reversibility in the following cycles (Figure S4a).
It is worth pointing out that the higher surface area of PG, 91 m2/g,
may play a role in the disparity of coulombic efficiency between PG and graphite (BET surface area: 1.2 m2/g).
We also compared the rate capability between PG and
graphite, as shown in Figure 4b, where PG delivers a reversible capacity of 221, 189, 142, 99, 43.2 and 13.6 mAh/g at current rates of 20, 50, 100, 200, 500 and 1000 mA/g, respectively, slightly lower than graphite at the same rates.
We hypothesize that the
tortuosity inside PG may slow down the ionic transport, which in turn lowers the rate performance.
13
ACS Paragon Plus Environment
ACS Applied Materials & Interfaces
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 14 of 25
It should be pointed out that the capacities of PG and graphite mentioned above have excluded the contribution from carbon black (C45). We tested C45/K half-cells separately by having C45 as the sole active mass with CMC as the binder (mass ratio: 90/10).
The first two galvanostatic potassiation/depotassiation potential profiles of
C45 in the potential range of 0.1–2 V vs. K+/K at a current rate of 20 mA/g are shown in Figure S5.
The capacity values of potassiation and depotassiation in the first
cycle are 1707 mAh/g and 276 mAh/g, respectively.
The coulombic efficiency of
the first cycle is only 16.2%, which may partially explain the low coulombic efficiency of the graphite (78.0%) and PG (54.1%) electrodes we tested because C45 was added.
C45’s more stabilized capacity values of potassiation and depotassiation
in the second cycle are 387 and 227 mAh/g, respectively.
Here, we have factored the
capacity value of C45 in the 2nd depotassiation sweep out of the observed capacity values of the mixtures of PG/C45 and Graphite/C45 to calculate our results.
The
calculated depotassiation capacity of PG is 224 mAh/g, which is lower than the observed 252 mAh/g when not considering that C45 contributes to the total capacity.
To further understand the electrochemical behavior of PG, we also performed cyclic voltammetry (CV) at a scan rate of 0.1 mV/s in the potential range of 0.01-2 V. Surprisingly, there is no cathodic peak until 0.1 V vs. K+/K (Figure 4c). During the anodic scan for depotassiation, a single broad current peak centers at 0.5 V, which is in contrast to the multiple anodic peaks for graphite accounting for sequential
14
ACS Paragon Plus Environment
Page 15 of 25
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
ACS Applied Materials & Interfaces
transitions between staged K-GIC phases (Figure 4d).
The lack of phase
transformation for PG electrode may lead to minimal volume change during charge/discharge.
We notice that there is a large irreversible capacity during the first galvanostatic cycle, whereas there is a lack of massive cathodic current corresponding to the SEI formation in the first CV curves, which seems contradictory. If SEI were formed, there should be an obvious slope/plateau at relatively high potentials, i.e., > 0.2 V vs K+/K in the potassiation/depotassiation profiles. However, such a behavior was not observed, which is further confirmed by the CV results. It is thus worth emphasizing that the lack of SEI formation on carbon anodes in KIBs could be a significant advantage for potential KIBs.
Then, the question is whether the irreversible capacity in
galvanostatic cycling is reflected by CV curves. If we integrate the areas for the cathodic and anodic current over time in the CV curves for the PG electrode and graphite electrode, we find the ratios are 42.8% and 54.0% for PG and graphite, respectively.
These values, representing the degrees of the irreversible cathodic
reactions, are very close to the coulombic efficiency values obtained from the 1st galvanostatic potassiation/depotassiation potential profiles.
That being said, the
exact cause that leads to the irreversibility is currently under investigation. It is most likely due to the trapping mechanism, where some trapped K-ions cannot be extracted during depotassiation.
15
ACS Paragon Plus Environment
ACS Applied Materials & Interfaces
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 16 of 25
It is also worth pointing out that high surface area of electrode materials typically induces more pronounced formation of solid electrolyte interphase (SEI).
NG with a
surface area of 1900 m2/g exhibits the first-cycle coulombic efficiency of only 10.9% (Figure S6a), which is confirmed by the massive irreversible cathodic current of the first CV curve (Figure S6c).
Also, its coulombic efficiency remains below 90% in
the first 20 cycles (Figure S6b, d).
Furthermore, NG presents linear galvanostatic
charge/discharge potential profiles and rectangular CV curves, which is typical capacitive charge storage behavior (Figure S6a, c).
Since the faster capacitive
mechanism plays a more dominant role than the intercalation mechanism, NG exhibits relatively good rate performance (Figure S6b), and fairly stable cycling performance especially after the SEI has been formed (Figure S6d).
The comparison of physical
and electrochemical properties of NG and PG is summarized in Table S1 in the supporting information.
16
ACS Paragon Plus Environment
Page 17 of 25
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
ACS Applied Materials & Interfaces
Figure 5. (a,b) Ex situ XRD patterns and ex situ Raman spectra of graphite and PG collected after 100 cycles. (c) Long-term cycling performance of graphite and PG at a current density of 100 mA/g.
PG exhibits much improved long-term cycling performance than graphite, where the cycling results from three PG/K cells, PG-1, PG-2, and PG-3 are shown in Figure 5c. Graphite retains ca. 6% of its original capacity after 140 cycles (from two cells, G-1 and G-2), whereas PG retained ca. 50% of its original capacity after 240 cycles.
The
capacity fading for PG is a linear curve, while graphite’s capacity remained stable for the first 50 cycles or so, and then went through an abrupt degradation afterwards. Figure
S7
shows
the
1st,
100th,
200th
and
17
ACS Paragon Plus Environment
300th
galvanostatic
ACS Applied Materials & Interfaces
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
potassiation/depotassiation potential profiles of PG.
Page 18 of 25
Besides the capacity fading, the
polarization increases upon cycling, which may be related to the varying surface properties of the carbon electrode and/or the possible degradation of the electrolyte, considering that potassium counter electrode can be very reactive toward organic electrolyte.
To understand the fading mechanism, we conducted ex situ XRD to investigate the structural change of PG and graphite after 100 cycles, as shown in Figure 5a.
The
sharp (002) peak for pristine graphite evolves into a compounded peak, which is composed of the pure graphite peak in the middle, a disordered carbon shoulder on the left and the other shoulder on the right that may be attributed to a high-stage GIC phase.69
In fact, the shoulder on the right shows up even after the first cycle.52
The
long-term capacity fading of graphite may be attributed to the newly generated amorphous substructures that shut down the access of K-ions into the internal galleries of the large graphite crystals with rigid structures.
In contrast, interestingly,
the originally ordered local structure of PG along c-axis becomes completely amorphous after 100 cycles based on the XRD results (Figure 5a).
The (002) peak
of PG shifted from 25.8º to 17.4º after 100 cycles, where the d-spacing increases from 0.345 to 0.510 nm, and the much broadened peak with a large value of FWHM suggests effective exfoliation of graphitic nanodomains possibly into single or double layers of graphenes with no coherence along c-axis.
18
ACS Paragon Plus Environment
Page 19 of 25
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
ACS Applied Materials & Interfaces
Considering the disparity of cycling performance of graphite and PG as well as the ex situ post-cycling XRD results, it appears that the cycled graphite containing amorphous and crystalline portions does not provide efficient accessibility for K-ions any longer; however, the exfoliation of graphitic nanodomains taking place inside PG, albeit turning PG completely non-graphitic, enables much better long-term cyclability. This confirms our hypothesis that the disordered arrangement of graphene layers inside non-graphitic carbons leads to superior cycling performance, which has been observed in hard carbon and soft carbon in our prior reports.
Interestingly, there is
barely any change of the Raman spectra for both G and PG before and after cycling in Figure 5b, indicating that the ordered atomic structure along the ab planes remain unaffected during the long cycling in KIBs.
Conclusion In summary, we synthesized polynanocrystalline graphite by chemical vapor deposition on a nanoporous graphenic carbon.
This new type of carbon is composed
of randomly oriented graphitic nanodomains, facilitating a unique structure with a low density of 0.92 g/cc and a relatively low surface area of 91 m2/g.
The
polynanocrystalline graphite is structurally unique, compared to low-dimensional nanocrystalline carbon materials, nanoporous carbon, amorphous carbon and graphite. This polynanocrystalline graphite shows much improved long-term cycling life in KIBs, where it exhibits 50% capacity retention over 240 cycles in contrast to 6% over 140 cycles for graphite.
The improved cycling performance is attributed to the
polynanocrystalline nature of PG with disorder at nanometric scales, which allows for 19
ACS Paragon Plus Environment
ACS Applied Materials & Interfaces
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 20 of 25
exfoliation of graphitic nanodomains inside the bulk particle of PG, thus still retaining the electrode structural integrity.
Our results provide new insights on carbon design
by differentiating structures at nanometric scales and atomic scales, while maintaining relatively low surface area.
This new type of carbon may well find applications
beyond KIBs.
ASSOCIATED CONTENT
Supporting Information.
BET results and supporting data are in Figure S1-S4.
This material is available free of charge via the Internet at http://pubs.acs.org.
AUTHOR INFORMATION Corresponding Author *E-Mail:
[email protected] Author Contributions The manuscript was written through contributions of all authors.
All authors have
given approval to the final version of the manuscript. Acknowledgments X. J. are thankful for the financial supports from National Science Foundation Award No.1551693.
We would like to thank Dr. Peter Eschbach and Ms. Teresa Sawyer for
the SEM measurements at the OSU Electron Microscopy Facility.
Additional
acknowledgements extend out to Mr. Joshua Razink for the TEM measurements at the Center for Advanced Materials Characterization at Oregon (CAMCOR) at the University of Oregon. We are thankful to Professor Chih-Hung Chang and Mr. 20
ACS Paragon Plus Environment
Page 21 of 25
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
ACS Applied Materials & Interfaces
Changqing Pan for Raman analysis.
References (1)
(2)
(3)
(4)
(5)
(6) (7) (8)
Yuan, L.-X.; Wang, Z.-H.; Zhang, W.-X.; Hu, X.-L.; Chen, J.-T.; Huang, Y.-H.; Goodenough, J. B. Development and Challenges of LiFePO4 Cathode Material for Lithium-ion Batteries. Energy Environ. Sci. 2011, 4 (2), 269-284. Liang, X.; Garsuch, A.; Nazar, L. F. Sulfur Cathodes Based on Conductive MXene Nanosheets for High-Performance Lithium-Sulfur Batteries. Angew. Chem., Int. Ed. 2015, 54 (13), 3907-3911. Weng, W.; Pol, V. G.; Amine, K. Ultrasound Assisted Design of Sulfur/Carbon Cathodes with Partially Fluorinated Ether Electrolytes for Highly Efficient Li/S Batteries. Adv. Mater. 2013, 25 (11), 1608-1615. Huang, H.; Yin, S. C.; Kerr, T.; Taylor, N.; Nazar, L. F. Nanostructured Composites: a High Capacity, Fast Rate Li3V2(PO4)3/Carbon Cathode for Rechargeable Lithium Batteries. Adv. Mater. 2002, 14 (21), 1525-1528. Huang, Y.-H.; Goodenough, J. B. High-rate LiFePO4 Lithium Rechargeable Battery Promoted by Electrochemically Active Polymers. Chem. Mater. 2008, 20 (23), 7237-7241. Tarascon, J.-M.; Armand, M. Issues and Challenges Facing Rechargeable Lithium Batteries. Nature 2001, 414 (6861), 359-367. Whittingham, M. S. Lithium Batteries and Cathode Materials. Chem. Rev. 2004, 104 (10), 4271-4302. Ma, H.; Jiang, H.; Jin, Y.; Dang, L.; Lu, Q.; Gao, F. Carbon Nanocages@ Ultrathin Carbon Nanosheets: One-Step Facile Synthesis and Application as Anode Material for Lithium-ion Batteries. Carbon 2016, 105, 586-592.
(9) (10) (11)
(12)
(13)
(14)
(15)
Ellis, B. L.; Nazar, L. F. Sodium and Sodium-ion Energy Storage Batteries. Curr. Opin. Solid State Mater. Sci. 2012, 16 (4), 168-177. Pan, H.; Hu, Y.-S.; Chen, L. Room-temperature Stationary Sodium-Ion Batteries for Large-Scale Electric Energy Storage. Energy Environ. Sci. 2013, 6 (8), 2338-2360. Yin, C.; He, L.; Wang, Y.; Liu, Z.; Zhang, G.; Zhao, K.; Tang, C.; Yan, M.; Han, Y.; Mai, L. Pyrolyzed Carbon with Embedded NiO/Ni Nanospheres for Applications in Microelectrodes. RSC Adv. 2016, 6 (49), 43436-43441. Jo, C.; Park, Y.; Jeong, J.; Lee, K. T.; Lee, J. Structural Effect on Electrochemical Performance of Ordered Porous Carbon Electrodes for Na-Ion Batteries. ACS Appl. Mater. Interfaces 2015, 7 (22), 11748-11754. Kim, Y.; Kim, Y.; Choi, A.; Woo, S.; Mok, D.; Choi, N. S.; Jung, Y. S.; Ryu, J. H.; Oh, S. M.; Lee, K. T. Tin Phosphide as a Promising Anode Material for Na-Ion Batteries. Adv. Mater. 2014, 26 (24), 4139-4144. Kim, Y.; Park, Y.; Choi, A.; Choi, N. S.; Kim, J.; Lee, J.; Ryu, J. H.; Oh, S. M.; Lee, K. T. An Amorphous Red Phosphorus/Carbon Composite as a Promising Anode Material for Sodium Ion Batteries. Adv. Mater. 2013, 25 (22), 3045-3049. Liu, Y.; Xu, Y.; Han, X.; Pellegrinelli, C.; Zhu, Y.; Zhu, H.; Wan, J.; Chung, A. C.;
21
ACS Paragon Plus Environment
ACS Applied Materials & Interfaces
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
(16)
(17)
(18)
(19) (20)
(21)
(22)
(23)
(24)
(25)
(26)
(27)
(28) (29)
Vaaland, O.; Wang, C. Porous Amorphous FePO4 Nanoparticles Connected by Single-Wall Carbon Nanotubes for Sodium Ion Battery Cathodes. Nano Lett. 2012, 12 (11), 5664-5668. Mu, L.; Xu, S.; Li, Y.; Hu, Y. S.; Li, H.; Chen, L.; Huang, X. Prototype Sodium-Ion Batteries Using an Air-Stable and Co/Ni-Free O3‐Layered Metal Oxide Cathode. Adv. Mater. 2015, 27 (43), 6928-6933. Wang, Y.; Xiao, R.; Hu, Y.-S.; Avdeev, M.; Chen, L. P2-Na0.6[Cr0.6Ti0.4]O2 Cation-Disordered Electrode for High-Rate Symmetric Rechargeable Sodium-Ion Batteries. Nat. Commun. 2015, 6. Wang, Y.; Yu, X.; Xu, S.; Bai, J.; Xiao, R.; Hu, Y.-S.; Li, H.; Yang, X.-Q.; Chen, L.; Huang, X. A Zero-Strain Layered Metal Oxide as the Negative Electrode for Long-Life Sodium-Ion Batteries. Nat. Commun. 2013, 4. Wen, Y.; He, K.; Zhu, Y.; Han, F.; Xu, Y.; Matsuda, I.; Ishii, Y.; Cumings, J.; Wang, C. Expanded Graphite as Superior Anode for Sodium-Ion Batteries. Nat. Commun. 2014, 5. Wu, L.; Hu, X.; Qian, J.; Pei, F.; Wu, F.; Mao, R.; Ai, X.; Yang, H.; Cao, Y. Sb-C Nanofibers with Long Cycle Life as an Anode Material for High-Performance Sodium-Ion Batteries. Energy Environ. Sci. 2014, 7 (1), 323-328. Fang, Y.; Xiao, L.; Qian, J.; Ai, X.; Yang, H.; Cao, Y. Mesoporous Amorphous FePO4 Nanospheres as High-Performance Cathode Material for Sodium-Ion Batteries. Nano Lett. 2014, 14 (6), 3539-3543. Yang, C.; Li, W.; Yang, Z.; Gu, L.; Yu, Y. Nanoconfined Antimony in Sulfur and Nitrogen Co-doped Three-Dimensionally (3D) Interconnected Macroporous Carbon for High-Performance Sodium-Ion Batteries. Nano Energy 2015, 18, 12-19. Xu, X.; Yan, M.; Tian, X.; Yang, C.; Shi, M.; Wei, Q.; Xu, L.; Mai, L. In Situ Investigation of Li and Na Ion Transport with Single Nanowire Electrochemical Devices. Nano Lett. 2015, 15 (6), 3879-3884. Dong, Y.; Li, S.; Zhao, K.; Han, C.; Chen, W.; Wang, B.; Wang, L.; Xu, B.; Wei, Q.; Zhang, L. Hierarchical Zigzag Na1.25V3O8 Nanowires with Topotactically Encoded Superior Performance for Sodium-Ion Battery Cathodes. Energy Environ. Sci. 2015, 8 (4), 1267-1275. Ding, J.; Wang, H.; Li, Z.; Cui, K.; Karpuzov, D.; Tan, X.; Kohandehghan, A.; Mitlin, D. Peanut Shell Hybrid Sodium Ion Capacitor with Extreme Energy–Power Rivals Lithium Ion Capacitors. Energy Environ. Sci. 2015, 8 (3), 941-955. Kohandehghan, A.; Cui, K.; Kupsta, M.; Ding, J.; Memarzadeh Lotfabad, E.; Kalisvaart, W. P.; Mitlin, D. Activation with Li Enables Facile Sodium Storage in Germanium. Nano Lett. 2014, 14 (10), 5873-5882. Li, S.; Dong, Y.; Xu, L.; Xu, X.; He, L.; Mai, L. Effect of Carbon Matrix Dimensions on the Electrochemical Properties of Na3V2(PO4)3 Nanograins for High-Performance Symmetric Sodium-Ion Batteries. Adv. Mater. 2014, 26 (21), 3545-3553. Zhu, C.; Kopold, P.; van Aken, P. A.; Maier, J.; Yu, Y. High Power–High Energy Sodium Battery Based on Threefold Interpenetrating Network. Adv. Mater. 2016, 28, 2409-2416. Liu, J.; Kopold, P.; Wu, C.; van Aken, P. A.; Maier, J.; Yu, Y. Uniform Yolk–Shell Sn4P3@C Nanospheres as High-Capacity and Cycle-Stable Anode Materials for Sodium-Ion Batteries. Energy Environ. Sci. 2015, 8 (12), 3531-3538.
22
ACS Paragon Plus Environment
Page 22 of 25
Page 23 of 25
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
ACS Applied Materials & Interfaces
(30)
(31)
(32)
(33)
(34)
(35)
(36)
(37)
(38)
(39) (40)
(41)
(42)
(43) (44)
Pol, V. G.; Lee, E.; Zhou, D.; Dogan, F.; Calderon-Moreno, J. M.; Johnson, C. S. Spherical Carbon as a New High-Rate Anode for Sodium-Ion Batteries. Electrochim. Acta 2014, 127, 61-67. Wang, J.; Luo, C.; Gao, T.; Langrock, A.; Mignerey, A. C.; Wang, C. An Advanced MoS2/Carbon Anode for High-Performance Sodium-Ion Batteries. Small 2015, 11 (4), 473-481. Komaba, S.; Murata, W.; Ishikawa, T.; Yabuuchi, N.; Ozeki, T.; Nakayama, T.; Ogata, A.; Gotoh, K.; Fujiwara, K. Electrochemical Na Insertion and Solid Electrolyte Interphase for Hard-Carbon Electrodes and Application to Na-Ion Batteries. Adv. Funct. Mater. 2011, 21 (20), 3859-3867. Wang, X.; Niu, C.; Meng, J.; Hu, P.; Xu, X.; Wei, X.; Zhou, L.; Zhao, K.; Luo, W.; Yan, M. Novel K3V2(PO4)3/C Bundled Nanowires as Superior Sodium‐Ion Battery Electrode with Ultrahigh Cycling Stability. Adv. Energy Mater. 2015, 5 (17), 1500716-1500723. Meng, J.; Liu, Z.; Niu, C.; Xu, X.; Liu, X.; Zhang, G.; Wang, X.; Huang, M.; Yu, Y.; Mai, L. A Synergistic Effect Between Layer Surface Configurations and K Ions of Potassium Vanadate Nanowires for Enhanced Energy Storage Performance. J. Mater. Chem. A 2016, 4 (13), 4893-4899. Qu, Y.; Zhang, Z.; Du, K.; Chen, W.; Lai, Y.; Liu, Y.; Li, J. Synthesis of Nitrogen-Containing Hollow Carbon Microspheres by a Modified Template Method as Anodes for Advanced Sodium-Ion Batteries. Carbon 2016, 105, 103-112. Shao, Y.; Rajput, N. N.; Hu, J.; Hu, M.; Liu, T.; Wei, Z.; Gu, M.; Deng, X.; Xu, S.; Han, K. S. Nanocomposite Polymer Electrolyte for Rechargeable Magnesium Batteries. Nano Energy 2015, 12, 750-759. Shao, Y.; Gu, M.; Li, X.; Nie, Z.; Zuo, P.; Li, G.; Liu, T.; Xiao, J.; Cheng, Y.; Wang, C. Highly Reversible Mg Insertion in Nanostructured Bi for Mg Ion Batteries. Nano Lett. 2013, 14 (1), 255-260. Shao, Y.; Liu, T.; Li, G.; Gu, M.; Nie, Z.; Engelhard, M.; Xiao, J.; Lv, D.; Wang, C.; Zhang, J.-G. Coordination Chemistry in Magnesium Battery Electrolytes: How Ligands Affect Their Performance. Sci. Rep. 2013, 3, 3130-3136. Zhang, X.; Tang, Y.; Zhang, F.; Lee, C. S. A Novel Aluminum-Graphite Dual‐Ion Battery. Adv. Energy Mater. 2016. Lin, M.-C.; Gong, M.; Lu, B.; Wu, Y.; Wang, D.-Y.; Guan, M.; Angell, M.; Chen, C.; Yang, J.; Hwang, B.-J. An Ultrafast Rechargeable Aluminium-Ion Battery. Nature 2015, 520 (7547), 324-328. Xing, Z.; Jian, Z.; Luo, W.; Qi, Y.; Bommier, C.; Chong, E. S.; Li, Z.; Hu, L.; Ji, X. A Perylene Anhydride Crystal as a Reversible Electrode for K-Ion Batteries. Energy Storage Materials 2016, 2, 63-68. Wessells, C. D.; Peddada, S. V.; Huggins, R. A.; Cui, Y. Nickel Hexacyanoferrate Nanoparticle Electrodes for Aqueous Sodium and Potassium Ion Batteries. Nano Lett. 2011, 11 (12), 5421-5425. Pasta, M.; Wessells, C. D.; Huggins, R. A.; Cui, Y. A High-Rate and Long Cycle Life Aqueous Electrolyte Battery for Grid-Scale Energy Storage. Nat. Commun. 2012, 3, 1149. Wessells, C. D.; Huggins, R. A.; Cui, Y. Copper Hexacyanoferrate Battery Electrodes with Long Cycle Life and High Power. Nat. Commun. 2011, 2, 550.
23
ACS Paragon Plus Environment
ACS Applied Materials & Interfaces
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
(45)
(46)
(47)
(48) (49) (50)
(51)
(52) (53)
(54) (55) (56)
(57)
(58)
(59)
(60)
Chen, Y.; Luo, W.; Carter, M.; Zhou, L.; Dai, J.; Fu, K.; Lacey, S.; Li, T.; Wan, J.; Han, X.; Bao, Y.; Hu, L., Organic Electrode for Non-Aqueous Potassium-Ion Batteries. Nano Energy 2015, 18, 205-211. McCulloch, W. D.; Ren, X.; Yu, M.; Huang, Z.; Wu, Y. Potassium-Ion Oxygen Battery Based on a High Capacity Antimony Anode. ACS Appl. Mater. Interfaces 2015, 7 (47), 26158-26166. Ren, X.; Lau, K. C.; Yu, M.; Bi, X.; Kreidler, E.; Curtiss, L. A.; Wu, Y. Understanding Side Reactions in K–O2 Batteries for Improved Cycle Life. ACS Appl. Mater. Interfaces 2014, 6 (21), 19299-19307. Ren, X.; Wu, Y. A Low-Overpotential Potassium-Oxygen Battery Based on Potassium Superoxide. J. Am. Chem. Soc. 2013, 135 (8), 2923-2926. Eftekhari, A. Potassium Secondary Cell Based on Prussian Blue Cathode. J. Power Sources 2004, 126 (1), 221-228. Liu, Y.; Fan, F.; Wang, J.; Liu, Y.; Chen, H.; Jungjohann, K. L.; Xu, Y.; Zhu, Y.; Bigio, D.; Zhu, T. In Situ Transmission Electron Microscopy Study of Electrochemical Sodiation and Potassiation of Carbon Nanofibers. Nano Lett. 2014, 14 (6), 3445-3452. Luo, W.; Wan, J. Y.; Ozdemir, B.; Bao, W. Z.; Chen, Y. N.; Dai, J. Q.; Lin, H.; Xu, Y.; Gu, F.; Barone, V.; Hu, L. B., Potassium Ion Batteries with Graphitic Materials. Nano Lett. 2015, 15 (11), 7671-7677. Jian, Z.; Luo, W.; Ji, X. Carbon Electrodes for K-Ion Batteries. J. Am. Chem. Soc. 2015, 137 (36), 11566-11569. Komaba, S.; Hasegawa, T.; Dahbi, M.; Kubota, K. Potassium Intercalation into Graphite to Realize High-Voltage/High-Power Potassium-Ion Batteries and Potassium-Ion Capacitors. Electrochem. Commun. 2015, 60, 172-175. Jian, Z.; Xing, Z.; Bommier, C.; Li, Z.; Ji, X. Hard Carbon Microspheres: Potassium-Ion Anode Versus Sodium-Ion Anode. Adv. Energy Mater. 2015, 6, 1501874-1501878. Harris, P. J. F. Fullerene-Like Models for Microporous Carbon. J. Mater. Sci. 2013, 48 (2), 565-577. Xing, Z.; Wang, B.; Gao, W.; Pan, C.; Halsted, J. K.; Chong, E. S.; Lu, J.; Wang, X.; Luo, W.; Chang, C.-H. Reducing CO2 to Dense Nanoporous Graphene by Mg/Zn for High Power Electrochemical Capacitors. Nano Energy 2015, 11, 600-610. Hu, C.; Sedghi, S.; Silvestre-Albero, A.; Andersson, G. G.; Sharma, A.; Pendleton, P.; Rodríguez-Reinoso, F.; Kaneko, K.; Biggs, M. J. Raman Spectroscopy Study of the Transformation of the Carbonaceous Skeleton of a Polymer-Based Nanoporous Carbon Along the Thermal Annealing Pathway. Carbon 2015, 85, 147-158. Bogdanov, K.; Fedorov, A.; Osipov, V.; Enoki, T.; Takai, K.; Hayashi, T.; Ermakov, V.; Moshkalev, S.; Baranov, A. Annealing-Induced Structural Changes of Carbon Onions: High-Resolution Transmission Electron Microscopy and Raman Studies. Carbon 2014, 73, 78-86. Osipov, V. Y.; Baranov, A.; Ermakov, V.; Makarova, T.; Chungong, L.; Shames, A.; Takai, K.; Enoki, T.; Kaburagi, Y.; Endo, M. Raman Characterization and UV Optical Absorption Studies of Surface Plasmon Resonance in Multishell Nanographite. Diamond Relat. Mater. 2011, 20 (2), 205-209. Cuesta, A.; Dhamelincourt, P.; Laureyns, J.; Martinez-Alonso, A.; Tascón, J. D. Raman
24
ACS Paragon Plus Environment
Page 24 of 25
Page 25 of 25
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
ACS Applied Materials & Interfaces
(61) (62)
(63)
(64) (65) (66) (67)
(68)
(69)
Microprobe Studies on Carbon Materials. Carbon 1994, 32 (8), 1523-1532. Yoshikawa, M.; Katagiri, G.; Ishida, H.; Ishitani, A.; Akamatsu, T. Raman Spectra of Diamondlike Amorphous Carbon Films. J. Appl. Phys. 1988, 64 (11), 6464-6468. Dresselhaus, M.; Jorio, A.; Souza Filho, A.; Saito, R. Defect Characterization in Graphene and Carbon Nanotubes Using Raman Spectroscopy. Philos. Trans. R. Soc., A 2010, 368 (1932), 5355-5377. Cançado, L. G.; Jorio, A.; Ferreira, E. M.; Stavale, F.; Achete, C.; Capaz, R.; Moutinho, M.; Lombardo, A.; Kulmala, T.; Ferrari, A. Quantifying Defects in Graphene via Raman Spectroscopy at Different Excitation Energies. Nano Lett. 2011, 11 (8), 3190-3196. Jawhari, T.; Roid, A.; Casado, J. Raman Spectroscopic Characterization of Some Commercially Available Carbon Black Materials. Carbon 1995, 33 (11), 1561-1565. Dallas, T. E. Structural Phases of Disordered Carbon Materials. Ph.D. Thesis, Texas Tech University, Lubbock, United States of America, 1996. Ferrari, A.; Robertson, J. Origin of the 1150 cm-1 Raman Mode in Nanocrystalline Diamond. Phys. Rev. B 2001, 63 (12), 121405. Calizo, I.; Balandin, A.; Bao, W.; Miao, F.; Lau, C. Temperature Dependence of the Raman Spectra of Graphene and Graphene Multilayers. Nano Lett. 2007, 7 (9), 2645-2649. Cancado, L.; Takai, K.; Enoki, T.; Endo, M.; Kim, Y.; Mizusaki, H.; Jorio, A.; Coelho, L.; Magalhaes-Paniago, R.; Pimenta, M. General Equation for the Determination of the Crystallite Size La of Nanographite by Raman Spectroscopy. Appl. Phys. Lett. 2006, 88 (16), 163106-163106. Dresselhaus, M. S.; Dresselhaus, G. Intercalation Compounds of Graphite. Adv. Phys. 2002, 51 (1), 1-186.
Table of Contents Graphic
25
ACS Paragon Plus Environment