You are on page 1of 5

Letter

pubs.acs.org/NanoLett

New Mechanistic Insights on Na-Ion Storage in Nongraphitizable


Carbon
Clement Bommier, Todd Wesley Surta, Michelle Dolgos, and Xiulei Ji*
Department of Chemistry, Oregon State University, Corvallis, Oregon 97331-4003, United States
*
S Supporting Information

ABSTRACT: Nongraphitizable carbon, also known as hard carbon, is considered one


of the most promising anodes for the emerging Na-ion batteries. The current
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

mechanistic understanding of Na-ion storage in hard carbon is based on the “card-


house” model first raised in the early 2000s. This model describes that Na-ion insertion
occurs first through intercalation between graphene sheets in turbostratic nanodomains,
followed by Na filling of the pores in the carbon structure. We tried to test this model
Downloaded via ESS INFLIBNET PCA 1 on July 25, 2022 at 13:23:44 (UTC).

by tuning the sizes of turbostratic nanodomains but revealed a correlation between the
structural defects and Na-ion storage. Based on our experimental data, we propose an
alternative perspective for sodiation of hard carbon that consists of Na-ion storage at
defect sites, by intercalation and last via pore-filling.

KEYWORDS: Na-ion batteries, intercalation, hard carbon, reaction mechanism, defect sites

I n today’s world, there is an ever-increasing need to move


beyond Li-ion batteries (LIBs) for more sustainable and
affordable electrochemical energy storage (EES) solutions.1
electrochemical performances in accordance with the card-
house model reported in the early 2000s.19−22 Under this
model, the Na-ion storage is attributed to two distinct
While LIBs represent the state-of-the-art technology, the mechanisms: (1) storage of ions between the graphene sheets
relative scarcity and uneven global distribution of lithium inside the TNs and (2) Na-ion/atom filling of the “pores”
reserves greatly hinders their use in large-scale applications such between the domains. This model further assigns the Na-ion
as grid-level EES.2 storage between the graphene sheets to the sloping voltage
To date, Na-ion batteries (NIBs) represent one of the most region of a potentiogram while ascribing the storage in the
promising alternatives due to the similar chemical and pores to the plateau region, as illustrated in Figure 1.
electrochemical properties between Na and Li and the vast In this study, we tuned the atomic structure of a model hard
abundance of Na resources.3−6 However, some differences carbon in order to confirm the electrochemical structure−
between the two types of atoms have presented considerable
property relationships given in the card-house model. We
challenges in the development of NIBs, especially in the case of
anode materials.7,8 LIBs utilize a graphite anode, as Li-ions can
reversibly form the LiC6 stage-one graphite intercalation
compound (GIC). However, Na-ion’s larger radii of 102 pm
vs 76 pm of Li-ion, along with its relatively high ionization
potential only allows for the reversible formation of the NaC64
GIC.9−12
In pursuing functional NIB anodes, much attention has been
devoted to nongraphitizable carbon, often referred to as hard
carbon. Unlike highly ordered graphite, which forms ABAB
stacked graphene sheets with an interlayer spacing of 0.335 nm,
hard carbon is composed of disordered turbostratic nano-
domains (TNs) and empty space (referred to as pores
hereafter) between these domains. This leads to three distinct
chemical environments for the storage of Na-ions: edge/defect
sites on pore surfaces, e.g., carbenes, vacancies, and dangling Figure 1. Visual representation of the card-house model on Na-ion
bonds on the edges of TNs, the interlayer space enclosed inside storage in hard carbon. The two distinct phases: intercalation inside
TNs and pore filling are seen.
the TNs, and last, the empty pores.13 This specific structure
allows for stoichiometry of NaC7.4 or 300 mAh g−1, which is
nearly 10-fold that of graphite. Received: May 19, 2015
In the literature, most studies on hard carbon and Revised: July 3, 2015
nongraphitic carbons14−18 as NIB anodes rationalize their Published: August 4, 2015

© 2015 American Chemical Society 5888 DOI: 10.1021/acs.nanolett.5b01969


Nano Lett. 2015, 15, 5888−5892
Nano Letters Letter

hypothesized that annealing sucrose-derived hard carbon at also observed through Raman spectroscopy by using the
progressively increasing temperatures would yield the desired equation below (Figure 2b).25
“tunable” hard carbon structures by controlling the domain
dimensions of the all-important TNs along the axial axis, Lc, ⎛I ⎞
and along the ab planes, La. Sucrose-derived carbons were La(nm) = (2.4 × 10−10)λnm 4⎜ G ⎟
⎝ ID ⎠
compared to a commercial glassy carbon (Aldrich) that is
essentially the “hardest” available hard carbon, as it is typically
where λnm4 is from the wavelength of the laser in nanometers,
obtained through pyrolysis at temperatures of over 2000 °C.23
ID is the integrated D band at 1350 cm−1 attributed to sp2
However, as our results will later show, we found some
carbon atoms with defects, and IG is the integrated G band at
experimental discrepancies that could not be reconciled with
the card-house model. This leads us to propose an alternative 1580 cm−1, which arises from planar sp2 configured carbon
perspective on the storage of Na-ions in hard carbon materials. atoms and is typical in pristine graphene (Figure S6). Results of
In order to examine the Lc and La parameters of the TNs in the Raman spectra show a similar trend as the XRD results,
different samples, we conducted X-ray diffraction (XRD), with La values of ∼9.2, 10.5, 13.3, and 19.4 nm for S-1100, S-
Raman spectroscopy, and total scattering via neutron diffraction 1400, S-1600, and glassy carbon, respectively (Table S2). The
measurements. Surface area and porosity measurements, scale discrepancy between the XRD and Raman La values has
alongside TEM images, are also included in the Supporting been noted before, as Raman measurements tend to over-
Information (Table S1 and Figures S1−S5). In the XRD estimate the La values.26 However, when performing a linear
patterns, we observed the (002) peaks, indicative of the d- regression of the Raman La values vs the XRD La values, we find
spacing between the graphene sheets in the TNs, to be an R2 of 0.95, thus confirming the increasing trend in La in the
contracting from ∼0.375 nm for pyrolyzed sucrose annealed at carbon materials (Figure S7). Additional information from the
1100 °C (S-1100), to ∼0.371 nm for S-1400, ∼0.370 nm for S- Raman data includes an ID/IG ratio, which can be used to
1600, and ∼0.352 nm for glassy carbon (Figure 2a). Average Lc quantify the concentration of defects along the graphene
sheets.27,28 These ratios indicate that the S-1100 material has
the greatest concentration of defects, while the glassy carbon
contains the smallest concentration.
While these traditional measurements are insightful, they fall
short of revealing local atomic structures. Thus, to probe the
short and midrange order of the carbon materials, we
performed total neutron scattering measurements and the
associated pair distribution function (PDF) analysis. The peaks
up to 5 Å represent the short-range correlations of the C−C
bonds of one hexagon unit within a sheet of graphene (Figures
2c and S8). Regardless of the annealing temperatures, all the
peaks are well aligned and are consistent with the graphene
lattice.29 There is an absence of a C−C correlation at 3.35−3.45
Å, which is found in highly ordered, graphitic carbon. This lack
of ordering between the graphene sheets supports the existence
of turbostratic disorder. The size of the graphene nanodomains
can be estimated in the mid/long-range where well-resolved
peaks vanish at certain r values. The S-1100 has notable peak
features until ∼15−17 Å, the S-1400 exhibits features
diminishing around ∼18−21 Å, while the glassy carbon has
features that persist past 30 Å (Figures 2c and S9−S11). These
data are consistent with XRD and Raman results, as the glassy
carbon exhibits the largest domains, while the S-1100 has the
smallest.
Figure 2. (a,b) XRD patterns and Raman spectra of the different In addition to the size of the graphene sheets in TNs,
carbon materials. (c) PDF results for total neutron scattering, with the neutron PDF results also provide information on the sp2 defect
inset showing the short-range order. concentration present in the TNs. Importantly, we note that
the intensity of the PDF peaks differs from sample to sample. In
PDF patterns, the area underneath a peak is proportional to the
and La values were estimated through the PDXL software by coordination number of that pair correlation.30,31 At 1.42 Å the
use of the Scherrer equation on (002) and (100) peaks, intensity of S-1400 and glassy carbon is virtually identical, and
respectively.24 The results reveal that the S-1100 is likely to only S-1100 has a slightly lower intensity. However, as r
have the smallest TN dimensions with Lc and La values of 1.15 increases, the difference in peak intensity becomes readily
and 2.54 nm, respectively, while these values increase to 1.17 apparent (Figure 2c). In particular, the first three peaks in the
and 3.18 nm for S-1400, 1.19 and 3.50 nm for S-1600, and 1.25 S-1100 pattern have lower intensities, meaning that it possesses
and 6.13 nm for the glassy carbon (Table S2). a significant number of nonhexagonal carbon rings. These
While the XRD trend is difficult to unambiguously defects likely result in a bending of the graphene sheets, which
distinguish between the different sucrose carbons, Raman diminishes the long-range ordering and therefore the size of the
spectroscopy was used as a local probe for more reliable domains.32 The PDF data show S-1100 has the greatest defect
measurements of amorphous materials. The increase in La is concentration, the glassy carbon has the least, while S-1400 is
5889 DOI: 10.1021/acs.nanolett.5b01969
Nano Lett. 2015, 15, 5888−5892
Nano Letters Letter

somewhere in the middle. The trend regarding the defect which corresponds to the high sodiation potentials in the
concentration corroborates the Raman spectrum results. sloping region.33−36 Furthermore, the “sloping” nature of this
Using several techniques, we have demonstrated that TN region can be attributed to the fact that there is no clear-cut
sizes increase upon annealing and are the largest for glassy definition of defect sites in amorphous carbon. They range from
carbon. According to the card-house model, capacity obtained dangling bonds at the edge of TNs, to monovacancies,
in the sloping region, defined as the region above 0.115 V vs divacancies, stone-wales defects, and extreme curvature in
Na+/Na of the potentiogram curve, is due to intercalation of graphene sheets; along with the presence of the sp3 “linking”
sodium-ions in the TNs. With larger TN sizes, the card-house carbons that connect the neighboring TNs. The plethora of
model predicts increasing sloping capacity from S-1100 to S- different morphological defects indicates that each type may
1600. However, the sloping capacity decreases from 120 mAh have its own respective sodiation voltage, thereby giving the
g−1 to 102 mAh g−1 and to 92 mAh g−1 for S-1100, S-1400, and sodiation of defects a slope-like shape on the potentiogram
S-1600, respectively (Figure 3a). One may be concerned about curve. Furthermore, considering the delocalized nature of
electrons of graphene,37 binding of a sodium-ion at defect sites
is more energetically favorable, as the defect sites have low
energy unfilled molecular orbitals that can effectively store extra
electrons. This increases the binding energy with sodium and
thus allows the sodiation to happen at higher voltages vs Na+/
Na.
We further investigated the sodiation mechanism through the
evaluation of kinetic properties using galvanostatic intermittent
titration (GITT) measurements (Figure S13).38,39 The Na-ion
diffusivity in hard carbon calculated as a function of potential
(Figure 3c) shows that diffusion associated with the sloping
potentials is much faster than that with the plateau. This
suggests that initial sodiation happens on easily accessible sites
in the carbon structure.
It is reasonable to say that the surface sites of TNs are more
accessible than the interlayer space in the TNs. As these sites
are progressively sodiated, Na-ions should then diffuse inside
the TNs. However, in order to do so, the Na-ions have to
overcome a repulsive charge gradient from the previously
bound Na-ions on defect sites in order to diffuse inside the
Figure 3. (a) Sodiation potentiograms for different carbons. Complete TNs. This explains the steep drop in diffusivity in the plateau
electrochemical data in Figures S14−S20 and Table S3−S5. (b) Plot of region of the potentiogram.
the sloping capacity vs the ID/IG ratio from Raman spectra. (c) GITT These kinetic results not only support our hypothesis that
profile and diffusivity as a function of states of charge (inset). The the sloping capacity is due to defect sites on the TN surfaces
diffusivity values are clearly overestimated as the geometric area of
but also give further insights into the sodiation mechanism of
electrodes is used for calculation. Our primary purpose is to
demonstrate the trends of evolving diffusivity at different state of the plateau region of the potentiogram curve. This has been the
charge. (d) dQ/dV plot from 0.12 to 0.01 V with corresponding subject of ambiguity, as recently published experimental results
diffusivity values. have suggested that the plateau region is due to Na insertion
into TNs as opposed to the pore filling mechanism suggested
whether it is the decreased d-spacing from S-1100 to S-1600 by the card-house model.40−42 Similarly, we also observed
that causes the lower capacity. As estimated from the XRD reversible expansion and contraction of d-spacing due to
patterns, we notice that the d-spacings are only different by sodiation and desodiation at the plateau region, i.e., from 0.2 to
0.005 nm between S-1100 and S-1600. We are aware that this 0.01 V, while conducting ex situ XRD. This suggests that
average d-spacing difference is beyond relevant meaning in intercalation occurs at lower voltages (Figure 4a,b).
physics. Despite our hypothesis that the plateau region could be
Instead of being dependent on the size of TNs, the capacity attributed to intercalation of sodium-ions in TNs, the newly
in the sloping region seems to be directly correlated to the obtained diffusion results, combined with the rest of our
defect concentration present. When plotting the slope capacity experimental data suggest that the plateau region is in fact a
versus defect concentration in the carbon samples, as expressed function of both intercalation into TNs as well as pore-filling
by the integrated ID/IG ratio, we observe a linear relationship suggested by the card-house model. When observing the values
with an R2 value of 0.90 (Figures 3b and S12). This leads to our at low voltages, we can observe that diffusion values reach a
hypothesis that the defected carbon sites in hard carbon rather minimum at 0.05 V; the same voltage where dQ/dV values
than the TNs are responsible for the capacity in the sloping reach a maximum. From 0.05 V to the cutoff voltage, the
region. The obtained PDF data also supports this hypothesis as diffusion values gradually increase, while dQ/dV values decrease
it shows that the S-1100 material has the greatest defect (Figure 3d). This observation begs the question: if intercalation
concentration and the largest sloping capacity, while the glassy is the one and only mechanism in the plateau region, why do
carbon has the lowest defect concentration and the smallest diffusivity values begin to rise at the very end, displaying an
sloping capacity. Prior computational studies also support this almost U-turn like reversal? If intercalation were the only
hypothesis as they show that binding energy between Na-ion mechanism, diffusivity values should become progressively
and carbon is highest at graphene defect sites and vacancy sites, lower and lower, as the diffusion length to storage sites keeps
5890 DOI: 10.1021/acs.nanolett.5b01969
Nano Lett. 2015, 15, 5888−5892
Nano Letters Letter

bilayer.34 This is a logical claim, as a sodium-ion between two


graphene sheets is more coordinated than one ion interacting
with only a single graphene sheet.
To summarize, we have revealed some discrepancies with the
card-house model on the Na-ion storage in hard carbon. We
have demonstrated that the storage mechanism in the sloping
part of the potentiogram curve can be better explained through
storage at defect sites, as opposed to intercalation between
graphene sheets. Furthermore, our additional diffusion, ex situ
XRD, and Na-plating experiments lead us to hypothesize that
the storage mechanism in the low voltage plateau region is due
to the intercalation between graphene sheets and minor
phenomenon of Na-ion adsorption on pore surfaces as shown
in the graphical representation in Figure 4d. Lastly, we realize
that this Letter might be regarded as controversial in the field of
sodium-ion storage in amorphous carbon, some of its
indications are at odds with long-standing principles. Our
Figure 4. (a) Ex situ XRD profiles at various stage of sodiation (S) and sincere hope is to present another angle of perspective to this
issue.


desodiation (D) with the (002) peak indicated. (b) Corresponding d-
spacing plots. (c) Potentiogram of sodiation until Na-metal plating is
induced. (d) Potentiogram and schematic of proposed Na-ion three- ASSOCIATED CONTENT
part storage mechanism. Please note that the figure is meant as a
schematic representing the three different types of binding sites as
*
S Supporting Information

opposed to an accurate representation of the hard carbon structure. The Supporting Information is available free of charge on the
ACS Publications website at DOI: 10.1021/acs.nano-
lett.5b01969.
increasing. However, the continuous drop in diffusivity values is
Surface area and porosity, HRTEM, additional PDF data,
not observed.
This suggests that the sodiation mechanism is changing in complete electrochemical characterization, and exper-
the voltage range right before reaching the cutoff potential. At imental methods (PDF)
such low voltages, storage sites can be characterized by a weak
binding energy. Furthermore, the increasing diffusivity values
suggest a more facile diffusion than intercalation. Considering
■ AUTHOR INFORMATION
Corresponding Author
the weak binding energy and facile diffusion, we postulate that *E-mail: david.ji@oregonstate.edu.
storage at low voltages is due to the Na-atom adsorption on the
sp2 configured pore surfaces. First off, a sp2 graphene surface Notes
The authors declare no competing financial interest.


has been shown in computational studies to be the least
energetically favorable binding site for sodium-ions.33−36 Thus,
it is logical that storage of sodium-ions on pore surfaces should ACKNOWLEDGMENTS
only occur at low voltages close to that of sodium plating. X.J. acknowledges the financial supports from Advanced
Additionally, unlike intercalation, storage of on pore surface Research Projects Agency-Energy (ARPA-E), DOE of the
does not require the expansion of two adjacent graphene layers United States, Award number: DE-AR0000297TDD. M.D.
to make insertion possible, which may account for its faster gratefully acknowledges the financial support from Oregon
diffusivity. State University. Research at ORNL’s Spallation Neutron
To test the above hypothesis, it is critical to determine Source was sponsored by the Scientific User Facilities Division,
whether such adsorption is of the same potential as the Na Office of Basic Energy Sciences of the U.S. DOE. X.J. thanks
plating. When running a half-cell to voltages below 0.0 V vs Dr. Yuliang Cao from Wuhan University for discussion.


Na+/Na, we found that the onset of Na-metal plating actually
occurs at −0.02 V before reaching a steady value of −0.015 V REFERENCES
(Figure 4c). When looking at the potentiogram near 0.0 V, we (1) Choi, N. S.; Chen, Z. H.; Freunberger, S. A.; Ji, X. L.; Sun, Y. K.;
notice that the slope becomes much steeper starting at 0.04 ± Amine, K.; Yushin, G.; Nazar, L. F.; Cho, J.; Bruce, P. G. Angew. Chem.,
0.01 V and stays linear until plating begins. We hypothesize that Int. Ed. 2012, 51, 9994−10024.
this is a result of continuous deposition of sodium atoms on the (2) Thackeray, M. M.; Wolverton, C.; Isaacs, E. D. Energy Environ.
pore surfaces until a critical point is reached, and the sodium Sci. 2012, 5, 7854−7863.
atoms begin to agglomerate into metallic clusters. Therefore, (3) Pan, H. L.; Hu, Y. S.; Chen, L. Q. Energy Environ. Sci. 2013, 6,
considering the increase in diffusivity values at the very end of 2338−2360.
the sodiation process, we find it reasonable to propose that the (4) Kim, S. W.; Seo, D. H.; Ma, X. H.; Ceder, G.; Kang, K. Adv.
last storage mechanism that occurs during the sodiation of hard Energy Mater. 2012, 2, 710−721.
(5) Yabuuchi, N.; Kubota, K.; Dahbi, M.; Komaba, S. Chem. Rev.
carbon is the deposition of sodium atoms on pore surfaces. 2014, 114, 11636−11682.
This hypothesis of storage on pore surfaces is in good (6) Slater, M. D.; Kim, D.; Lee, E.; Johnson, C. S. Adv. Funct. Mater.
agreement with the original one made by the card-house model 2013, 23, 947−958.
and is also supported by the computational results; which show (7) Bommier, C.; Ji, X. Isr. J. Chem. 2015, 55, 486−507.
that the binding energy between planar graphene sheets and Na (8) Chevrier, V. L.; Ceder, G. J. Electrochem. Soc. 2011, 158, A1011−
is much smaller than binding on defect sites or in a graphene A1014.

5891 DOI: 10.1021/acs.nanolett.5b01969


Nano Lett. 2015, 15, 5888−5892
Nano Letters Letter

(9) Doeff, M. M.; Ma, Y. P.; Visco, S. J.; Dejonghe, L. C. J.


Electrochem. Soc. 1993, 140, L169−L170.
(10) Enoki, T.; Endo, M.; Suzuki, M. Graphite Intercalation
Compounds and Applications; Oxford University Press: Oxford, U.K.,
2003.
(11) Okamoto, Y. J. Phys. Chem. C 2014, 118, 16−19.
(12) Wang, Z.; Selbach, S. M.; Grande, T. RSC Adv. 2014, 4, 3973−
3983.
(13) Franklin, R. E. Proc. R. Soc. London, Ser. A 1951, 209, 196−218.
(14) Li, Y.; Xu, S.; Wu, X.; Yu, J.; Wang, Y.; Hu, Y.-S.; Li, H.; Chen,
L.; Huang, X. J. Mater. Chem. A 2015, 3, 71−77.
(15) Alcantara, R.; Lavela, P.; Ortiz, G. F.; Tirado, J. L. Electrochem.
Solid-State Lett. 2005, 8, A222−A225.
(16) Wenzel, S.; Hara, T.; Janek, J.; Adelhelm, P. Energy Environ. Sci.
2011, 4, 3342−3345.
(17) Ponrouch, A.; Goni, A. R.; Palacin, M. R. Electrochem. Commun.
2013, 27, 85−88.
(18) Wen, Y.; He, K.; Zhu, Y.; Han, F.; Xu, Y.; Matsuda, I.; Ishii, Y.;
Cumings, J.; Wang, C. Nat. Commun. 2014, 5, 4033−4042.
(19) Thomas, P.; Billaud, D. Electrochim. Acta 2002, 47, 3303−3307.
(20) Stevens, D. A.; Dahn, J. R. J. Electrochem. Soc. 2001, 148, A803−
A811.
(21) Stevens, D. A.; Dahn, J. R. J. Electrochem. Soc. 2000, 147, 1271−
1273.
(22) Stevens, D. A.; Dahn, J. R. J. Electrochem. Soc. 2000, 147, 4428−
4431.
(23) Hishiyama, Y.; Inagaki, M.; Kimura, S.; Yamada, S. Carbon 1974,
12, 249−258.
(24) Warren, B. Phys. Rev. 1941, 59, 693−698.
(25) Cançado, L. G.; Takai, K.; Enoki, T.; Endo, M.; Kim, Y. A.;
Mizusaki, H.; Jorio, A.; Coelho, L. N.; Magalhães-Paniago, R.; Pimenta,
M. A. Appl. Phys. Lett. 2006, 88, 163106.
(26) Vázquez-Santos, M. B.; Geissler, E.; László, K.; Rouzaud, J.-N.;
Martínez-Alonso, A.; Tascón, J. M. D. J. Phys. Chem. C 2012, 116,
257−268.
(27) Ferrari, A. C.; Robertson, J. Philos. Trans. R. Soc., A 2004, 362,
2477−2512.
(28) Ferrari, A. C.; Robertson, J. Phys. Rev. B: Condens. Matter Mater.
Phys. 2000, 61, 14095−14107.
(29) Zhou, P.; Lee, R.; Claye, A.; Fischer, J. E. Carbon 1998, 36,
1777−1781.
(30) Egami, T.; Billinge, S. J. Underneath the Bragg Peaks: Structural
Analysis of Complex Materials; Elsevier: New York, 2003.
(31) Keen, D. A. J. Appl. Crystallogr. 2001, 34, 172−177.
(32) Petkov, V.; Difrancesco, R. G.; Billinge, S. J. L.; Acharya, M.;
Foley, H. C. Philos. Mag. B 1999, 79, 1519−1530.
(33) Datta, D.; Li, J. W.; Shenoy, V. B. ACS Appl. Mater. Interfaces
2014, 6, 1788−1795.
(34) Tsai, P.-c.; Chung, S.-C.; Lin, S.-k.; Yamada, A. J. Mater. Chem. A
2015, 3, 9763−9768.
(35) Malyi, O. I.; Sopiha, K.; Kulish, V. V.; Tan, T. L.; Manzhos, S.;
Persson, C. Appl. Surf. Sci. 2015, 333, 235−243.
(36) Shen, H.; Rao, D.; Xi, X.; Liu, Y.; Shen, X. RSC Adv. 2015, 5,
17042−17048.
(37) Popov, I. A.; Bozhenko, K. V.; Boldyrev, A. I. Nano Res. 2012, 5,
117−123.
(38) Weppner, W.; Huggins, R. A. J. Solid State Chem. 1977, 22,
297−308.
(39) Prosini, P. P.; Lisi, M.; Zane, D.; Pasquali, M. Solid State Ionics
2002, 148, 45−51.
(40) Ding, J.; Wang, H.; Li, Z.; Cui, K.; Karpuzov, D.; Tan, X.;
Kohandehghan, A.; Mitlin, D. Energy Environ. Sci. 2015, 8, 941−955.
(41) Lotfabad, E. M.; Ding, J.; Cui, K.; Kohandehghan, A.; Kalisvaart,
W. P.; Hazelton, M.; Mitlin, D. ACS Nano 2014, 8, 7115−7129.
(42) Ding, J.; Wang, H. L.; Li, Z.; Kohandehghan, A.; Cui, K.; Xu, Z.
W.; Zahiri, B.; Tan, X. H.; Lotfabad, E. M.; Olsen, B. C.; Mitlin, D.
ACS Nano 2013, 7, 11004−11015.

5892 DOI: 10.1021/acs.nanolett.5b01969


Nano Lett. 2015, 15, 5888−5892

You might also like