• No results found

On the Capacities of Freestanding Vanadium Pentoxide-Carbon Nanotube-Nanocellulose Paper Electrodes for Charge Storage Applications

N/A
N/A
Protected

Academic year: 2021

Share "On the Capacities of Freestanding Vanadium Pentoxide-Carbon Nanotube-Nanocellulose Paper Electrodes for Charge Storage Applications"

Copied!
11
0
0

Loading.... (view fulltext now)

Full text

(1)

On the Capacities of Freestanding Vanadium Pentoxide

Carbon Nanotube

–Nanocellulose Paper Electrodes for

Charge Storage Applications

Ahmed S. Etman,* Zhaohui Wang, Youyou Yuan, Leif Nyholm,* and Johanna Rosen*

1. Introduction

The demand on electrochemical energy storage devices, nowa-days, is not only for operating our portable electronic devices and automobiles, but also for establishing stationary clean and sustainable energy resources.[1–4]For portable and wearable devi-ces, the commercial interest is on the development offlexible and lightweight electrode materials with good rate performances, so that the electrodes can be charged rapidly.[5,6]Rechargeable

bat-teries and supercapacitors are two categories of energy storage

devices with different charge storage mech-anisms. The charge storage mechanism in supercapacitors mainly originates from the charging of electric double layers (EDL), which yields high power densities but rela-tively low energy densities. The charge stored using the EDL mechanism is com-monly reported in terms of specific capaci-tance with the unit F g1. Batteries, on the other hand, feature high energy densities but low power densities, as the full volume of the electrode material (and not only the electrode surface as in a supercapacitor) is assumed to be involved in the electrochem-ical reaction. The charge stored in battery-like materials is usually expressed in terms of specific capacity with the unit C g1or mAh g1. Many studies have consequently been conducted to increase the energy den-sities of supercapacitors, including the use of nonaqueous electrolytes to expand the operating potential window,[4] and hybrid electrode materials consisting of both carbonaceous (double-layer charging) materials and redox active materials.[7]

The 2D transition metal oxides,[8–13]chalcogenides,[14,15]and carbides (MXenes)[16–19]are materials with promising properties compared with their bulk counterparts, including high aspect ratios, high ion diffusion rates between the nanosheets, low elec-tronic resistances, as well as good solution-based processabilities. Hydrated vanadium pentoxide (V2O5⋅ nH2O) nanosheets are

Dr. A. S. Etman, Prof. J. Rosen

Department of Physics, Chemistry and Biology (IFM) Linköping University

SE-58183 Linköping, Sweden

E-mail: ahmed.etman@liu.se; johanna.rosen@liu.se Dr. A. S. Etman

Department of Chemistry Faculty of Science Alexandria University

Ibrahimia, Alexandria 21321, Egypt

The ORCID identification number(s) for the author(s) of this article can be found under https://doi.org/10.1002/ente.202000731. © 2020 The Authors. Energy Technology published by Wiley-VCH GmbH. This is an open access article under the terms of the Creative Commons Attribution License, which permits use, distribution and reproduction in any medium, provided the original work is properly cited.

DOI: 10.1002/ente.202000731

Dr. Z. Wang, Prof. L. Nyholm

Department of Chemistry-Ångström Laboratory Uppsala University

SE-75121 Uppsala, Sweden E-mail: leif.nyholm@kemi.uu.se Dr. Z. Wang

College of Materials Science and Engineering Hunan University

Changsha 410082, China Dr. Y. Yuan

College of Chemistry and Molecular Engineering Peking University

Yiheyuan Road 5, Beijing 100871, China

Herein, a one-step protocol for synthesizing freestanding 20μm thick cellulose paper electrodes composed of V2O5⋅ H2O nanosheets (VOx), carbon nanotubes

(CNTs), andCladophora cellulose (CC) is reported. In 1.0MNa2SO4, the VOx–

CNT–CC electrodes deliver capacities of about 200 and 50 C g1at scan rates of 20 and 500 mV s1, respectively. The obtained capacities are compared with the theoretical capacities and are discussed based on the electrochemical reactions and the mass loadings of the electrodes. It is shown that the capacities are diffusion rate limited and, consequently, depend on the distribution and thick-ness of the V2O5⋅ H2O nanosheets, whereas the long-term cycling stabilities

depend on vanadium species dissolving in the electrolyte. The electrodes feature high mass loadings (2 mg cm2), good rate performances (25% capacity retention at 500 mV s1), and capacity retentions of 85% after 8000 cycles. A symmetric VOx–CNT–CC energy storage device with a potential window of

about 1 V exhibits a capacity of 40 C g1at a scan rate of 2 mV s1.

(2)

often assumed to be promising 2D materials based on the theo-retical capacities of 965 or 1930 C g1calculated, assuming two or four electrons per V2O5unit (i.e., a complete reduction of Vþ5

to Vþ4and Vþ3, respectively) and one water molecule per V2O5

unit. The V2O5⋅ nH2O nanosheet material is commonly

composed of only a few layers, which has been reported to give a lithium-ion battery electrode capacity of 1728 C g1 for 4μm thick electrode at a current density of 10 mA g1.[20] The V2O5⋅ nH2O interlayer distance (d-spacing) can vary

between 6.34 and 13.80 Å, depending on the number of water molecules (n) present between the layers,[21]which is why the

double layer of V2O5⋅ nH2O can host a variety of ions.[11,20,21]

It has been shown that V2O5⋅ nH2O can be used as electrode

materials for supercapacitors[22] and rechargeable metal-ion batteries, e.g., lithium-ion batteries,[11,12,20] sodium-ion

batteries,[23,24]magnesium-ion batteries, and zinc batteries.[25,26] However, V2O5⋅ nH2O electrodes generally exhibit low electronic

conductivities and rapid capacity fading.[27,28]The latter capacity fading commonly stems from a combination of structural changes[29]and dissolution of vanadium in the electrolyte, espe-cially, in aqueous electrolytes.[30]Many authors have, therefore,

proposed possible solutions to the conductivity problem based on composite electrodes containing conducting polymers[31]or

car-bon additives, such as graphene,[32]carbon nanotubes (CNTs),[33] or carbon dots.[34] Likewise, some other nanostructured

vana-dium oxide and hydroxide materials were also encapsulated in carbonaceous network and showed good electrochemical perfor-mance.[35–39] A solid electrolyte has also been proposed to address the vanadium dissolution problem.[32,40]Conventional

cast electrodes, which contain organic binders and carbon black, on the other hand, typically suffer from problems such as decreased fractions of the electroactive compounds in the electro-des and slow diffusion of ions in the electrode. One promising approach is to use freestanding electrodes with relatively thin electroactive layers immobilized onto a porousfiber matrix with a large surface areas as such electrodes should be less affected by the abovementioned problems.[41]

The electrode dimensions and mass loadings are also impor-tant factors, which can influence the electrochemical behavior of the electrodes under investigation. In general, the use of electro-des with very small areas and, hence, very low mass loadings can lead to misleading conclusions about the electrochemical behav-ior (e.g., rate capability and long-term cycling) of a supercapacitor or battery material. The iR drop is, for example, directly propor-tional to the absolute applied current used, which depends on the amount of active materials (mass loading), whereas the electro-active material may not be fully accessible when using thick electrodes. When quite small electrodes are used, the iR drop will, hence, remain rather small even at high current densities. Significantly higher specific capacities are also expected for elec-trodes containing monolayers or only a few layers of the electro-active material in comparison with electrodes containing μm thick layers of the same material. This is a result of the much longer ion-diffusion lengths needed for the thicker electrodes, which imply that much longer times are needed to fully oxidize or reduce the thick electrodes. When comparing the performan-ces of electrodes with different mass loadings, the capacity of the electrode (in, e.g., C or mAh) should, hence, also be considered. While monolayer-based electrodes often exhibit high specific

capacities (e.g., in C g1of mAh g1), their capacities are gener-ally too low to be of practical importance unless their mass load-ings can be increased sufficiently without causing a significant decrease in their specific capacity. The mass loading of the elec-trode is likewise important when the active material undergoes dissolution in the electrolyte, because the capacity retention during long-term cycling then can depend on the dimensions and porosities of the electrodes. There is, therefore, a general need to study the influence of the electrode dimensions and mass loadings on the electrochemical behaviors of potential electrode materials.

Cellulose is a highly abundant sustainable material, which can be found in, e.g., trees, algae, and bacteria.[42–44]As cellulose has many interesting properties, such as biodegradability, solution-based processability, high flexibility, and high mechanical strength,[5,6,45–50]cellulose has been frequently used to fabricate electrode materials for wearable energy storage devices. This development has benefitted significantly from access to straight-forward manufacturing approaches. Wang et al.[51,52]have, for

example, described a straightforward vacuumfiltration approach for the manufacturing of freestanding cellulose paper electrodes using a mixture of a redox active material, CNTs, and Cladophora cellulose (CC). As the latter electrodes are porous, high specific capacities as well as fast charge and discharge rates can be obtained when the thickness of the electroactive coating on the cellulose fibers and CNT fibers is thin enough (e.g., <50 nm).[5,53–55] Recently, this approach was also used in the

manufacturing of nanostructured reduced molybdenum trioxide MoO3–x–CNT–nanocellulose paper electrodes, which were found to show fast charge and discharge rates in supercapacitor applications (e.g., exhibiting the capacities of 30 C g1 at a current density of 78 A g1).[56]As it is reasonable to assume that this synthesis approach can also be used together with other solution-processable nanostructured materials, this possibility should, consequently, be further investigated.

Hydrated vanadium pentoxide (V2O5⋅ nH2O, with n 1)

nanosheets,[20]composed of three to six layers, were recently

pre-pared using water-based techniques and commercial vanadium oxide precursors.[11,12,57]The synthesis involved a reflux of V

2O5

in hot water for 24 h in the presence of VO2or oxalic acid, after

which the formed nanostructured V2O5⋅ H2O was collected

using centrifugation and, subsequently, dried in air for a few hours. The process, which is efficient, fast, and provides a high yield, can be scaled-up to produce about 50 g.[57] Drop-casting of an aqueous suspension of the nanosheet onto hydrophobic substrates, such as CNT paper, has been found to produce a surface layer of V2O5⋅ nH2O on a CNT paper.[20] Although

V2O5⋅ nH2O nanosheets have previously been used as electrode

materials for lithium and sodium-ion batteries,[11,20,23]the

acces-sible capacity was found to depend on the electrode manufactur-ing method. There is, therefore, a need to explore the use of other hydrophilic substrates facilitating the attainment of a conformal distribution of the nanosheets. Furthermore, the electrochemical activity of the V2O5⋅ nH2O nanosheets synthesized by aqueous

exfoliation approach has not yet been explored in an aqueous electrolyte.

Herein, we describe a straightforward approach for the synthe-sis of freestanding V2O5⋅ H2O–CNT–CC paper electrodes for

(3)

The capacities of the electrodes are discussed based on the nature of the electrochemical reactions, the electrode mass loadings, and the distribution and thickness of the V2O5⋅ H2O nanosheets

in the electrode. The electrodes show promising rate capabilities and long-term cycling stabilities. Furthermore, a symmetric VOx–CNT–CC energy storage device is tested and shown to

exhibit a good capacity and rate capability within a potential window of about 1 V.

2. Results and Discussion

2.1. Synthesis and Morphology of the V2O5⋅ H2O–CNT–CC

Paper Electrodes

The V2O5⋅ H2O nanosheets were prepared using water-based

exfoliation of V2O5in the presence of oxalic acid, and the

chemi-cal, thermal, and structural analyses of the obtained V2O5⋅ H2O

nanosheets have been described elsewhere.[11,57]The as-prepared V2O5⋅ H2O is composed of three to six layers and contains about

20% V4þas indicated by previous high resolution transmission electron microscopy, X-ray photoelectron spectroscopy (XPS), Nuclear magnetic resonance (NMR), and X-ray absorption near-edge structure (XANES) data.[12,13,20] Furthermore, previous findings indicate that there is approximately one water molecule per V2O5unit at room temperature.[20]The V2O5⋅ n H2O–CNT–CC,

hereafter denoted VOx–CNT–CC, paper electrodes were

prepared using a single step vacuumfiltration method, as sche-matically shown in Figure 1. In a typical experiment, given amounts of V2O5⋅ nH2O, CNTs, and CC in the weight ratios

of 7:2:1 were dispersed by sonication in a 3:1 water–ethanol mix-ture. The formed suspension was then vacuumfiltrated to form flexible and freestanding VOx–CNT–CC paper. Figure 1d shows

the as-prepared paper electrode in different bending states, reflecting the high flexibility of the VOx–CNT–CC paper. The flexibility of the paper electrodes commonly increases the life-span of the electrodes during long-term electrochemical cycling via decreasing the internal strain inside the electrodes.[50,58] This goodflexibility of the paper electrode can be attributed to the good mechanical properties of the cellulose nanofibers. The X-ray diffraction (XRD) pattern of the VOx–CNT–CC paper

(see Figure S1a, Supporting Information) shows that the struc-ture of the V2O5⋅ H2O within the cellulose paper electrodes is

analogous to that of pristine V2O5⋅ H2O.

Scanning electron microscopy (SEM) images of the VOx–

CNT–CC freestanding paper from side (cross section) and top views are shown in Figure 1e,f, respectively. The side view dis-plays that the thickness of the cellulose paper electrode is about 20μm, whereas the top view indicates that the V2O5⋅ H2O is

dis-tributed within the randomly oriented CNT–CC matrix. Here, it should be noted that it is hard to distinguish between the CC fibers and the CNTs, as they have very similar morphologies.[52]

The energy-dispersive X-ray (EDX) mapping likewise indicates an even distribution of V, O, and C (see Figure S1b–e,

Figure 1. Synthesis and morphology of the VOx–CNT–CC paper: a–c) schematic illustration of the synthesis method used to make the VOx–CNT–CC

freestanding paper. d) Photographs of the freestanding VOx–CNT–CC paper showing its flexibility. e) Cross-sectional SEM image of the VOx–CNT–CC

(4)

Supporting Information), further supporting a uniform distribu-tion of V2O5⋅ H2O within the CNT–CC matrix. This finding is,

however, challenged by the results of the electrochemical characterization, which indicates inhomogeneous depth distri-butions of V2O5⋅ H2O in the electrodes (see below). The latter

indicates that some of the V2O5⋅ H2O was less electrochemically

accessible even though a homogeneous distribution was seen in the EDX mapping results.

2.2. Electrochemical Behavior of the VOx–CNT–CC Paper

Electrodes

The electrochemical behavior of the cellulose paper electrodes was investigated using cyclic voltammetry and constant current measurements. Figure 2a–c shows cyclic voltammograms (CVs) for the VOx–CNT–CC paper electrodes recorded at different scan

rates. As indicated by the arrows in Figure 2a, the potential was initially scanned from the open circuit potential (OCP) of about 0.32 V down to0.45 V and then back to 0.55 V. In some experi-ments, an extended potential window between0.8 and 0.8 V was, however, also used. The OCP value suggests (see below) that the pristine VOx–CNT–CC electrodes contained mainly Vþ5in agreement with previous XPS and NMR results for nanostruc-tured V2O5⋅ nH2O.[11,13,20] The pristine electrodes showed an

initial increase in the capacity during thefirst few cycles at a scan rate of 10 mV s1, indicating that the cycling opened up the struc-ture and increased the interfacial surface area accessed by the electrolyte (see Figure 2a). The latter assumption was confirmed by electrochemical impedance spectroscopy (EIS) measurements on the cellulose paper electrodes prior (blue line) to and after 10 (red line) and 75 (black) electrochemical cycles, as a smaller charge-transfer resistance (Rct) was obtained after the cycling

(see the Nyquist plots in Figure 2g,h). Furthermore, the slope of the straight line in the low-frequency region of Nyquist plots increases by cycling, reflecting the increase in the ion transport rate during the electrochemical cycling.

The electrochemistry of vanadium-based electrodes in aqueous electrolytes is quite complicated due to the presence of several vanadium oxidation states. At pH 7, thermodynamic data[59,60] indicate that V2O5 (i.e., Vþ5), VO2 (i.e., Vþ4), and

V2O3(i.e., Vþ3) should be the dominating solid phases in the

potential region of main interest here, i.e.,0.8 to 0.8 V versus Ag/AgCl (1M KCl). The reduction reactions of interest here

should, therefore, be the reduction of Vþ5(i.e., V2O5) to Vþ4

(i.e., VO2) and the reduction of Vþ4(i.e., VO2) to Vþ3(i.e., V2O3).

These can be described using the reactions and standard reduc-tion potentials (calculated usingΔG0¼ nFE0) presented in the following, which indicate that V2O5should be stable at potentials

above about 0.5 V versus Ag/AgCl (1MKCl), whereas VO2should

be found above 0.1 V versus Ag/AgCl (1MKCl), assuming

stan-dard conditions (e.g., pH¼ 0). V2O5þ 2 Hþþ 2 e⇌ 2 VO2þ H2O

¼ 0:54 V versus Ag=AgCl ð1 м KClÞ (1)

2VO2þ 2 Hþþ 2 e⇌ V2O3þ H2O

¼ 0:14 V versus Ag=AgCl ð1 м KClÞ (2)

Note also that both Reactions (1) and (2) should exhibit a pH dependence of59 mV pH1. The latter is important as meas-urements with pH indicator paper (see Figure S4e, Supporting Information) indicated that the pH in the electrolyte was roughly 4. This acidic pH can be explained by the presence of dissolved CO2and the low acid/base buffer capacity of the 1MNa2SO4

elec-trolyte. At room temperature and normal pressure, the pH should be about 4 in a solution saturated with CO2.[61]At pH

4, the equilibrium potentials should, consequently, be about 0.2 V more negative than the standard potentials (see Table S1, Supporting Information). This means that V2O5should

exist above about 0.3 V versus Ag/AgCl (1MKCl), whereas VO2

should be found above about0.1 V versus Ag/AgCl (1MKCl). The Pourbaix diagram,[60]however, also suggests that the

pres-ence of soluble Vþ2species may complicate the electrochemical behavior of the electrodes at potentials below about0.6 V versus Ag/AgCl (1MKCl) and pH< 4. Such a generation of soluble Vþ2

species should, however, only affect the electrochemical behavior of Vþ3in the extended potential region between0.8 and 0.8 V versus Ag/AgCl (1MKCl). As shown in Figure 2a, the couple

redox peaks were clearly seen in the CVs obtained with the VOx–CNT–CC paper electrodes recorded using a scan rate of

10 mV s1. It is reasonable to assume that the redox pairs seen at about 0 and 0.17 V versus Ag/AgCl (1MKCl), i.e., peaks 2 and

20, were due to the reduction of V2O5(i.e., Vþ5) to VO2(i.e., Vþ4)

and the corresponding oxidation, respectively. The sharp redox peaks observed at about 0.246 V (peak 1, reduction) and 0.07 V (peak 10, oxidation) versus Ag/AgCl (1MKCl) can

anal-ogously be assigned to the VO2/V2O3(i.e., Vþ4/Vþ3) redox couple.

To explore the possibility of widening the potential window during electrochemical cycling of the cellulose paper electrodes, the performances of the electrodes were studied using different upper and lower cutoff potentials (see Figure 2b). When the potential window was expanded to cover 0.8 to 0.8 V versus Ag/AgCl (1M KCl), the accessible capacity was increased (see Figure 2b, black curve). The reversibility of the system, however, decreased, as is shown by the increased coulombic efficiency values, especially at low scan rates (i.e., about 2 and 5 mV s1; see Figure S2a–c, Supporting Information). The reduction observed below 0.45 V versus Ag/AgCl (1M KCl) was most

likely due to hydrogen evolution, because the equilibrium poten-tial for this reaction should be about 0.4 V versus Ag/AgCl (1M KCl) at pH 4. It should be noted that the conversion of

V2O5to metallic vanadium (see Reaction (3)) is unlikely to in

flu-ence the voltammogram as the equilibrium potential for this reaction should be about0.64 V versus Ag/AgCl (1MKCl) at

pH 4. Moreover, V2O5 should not exist at 0.45 V versus

Ag/AgCl (1MKCl) due to its stepwise reduction to VO2 and

V2O3during the cathodic scan. The reduction of V2O3to VO[62]

and metallic vanadium, respectively (see Reactions (4) and (5)) are also improbable based on their standard potentials of 0.65 and 0.90 V versus Ag/AgCl (1M KCl) and their pH

dependence of 59 mV pH1 (see Table S1, Supporting Information), indicating that these reactions should take place outside the used potential window at pH 4.

V2O5þ 10 Hþþ 10 e⇌ 2 V þ 5 H2O

(5)

Figure 2. Electrochemical performance of the VOx–CNT–CC paper electrodes: a) CVs showing the first (blue curve) and tenth (black curve) cycles at a

scan rate of 10 mV s1. b) Third scan of CVs obtained at a scan rate of 2 mV s1using different upper and lower cutoff potentials. c) Third scan of CVs recorded using the scan rates of 2, 5, 10, and 20 mV s1. The arrows in (a–c) indicate the scan direction. d) The discharge (i.e., reduction) and charge (i.e., oxidation) capacities and the coulombic efficiency as a function of the logarithm of the scan rate. e) Potential versus capacity plots at the current densities of 1, 3, 5, 10, and 40 A g1. f ) Potential versus capacity plots for the VOx–CNT–CC paper and CNT–CC paper at a current density of 1 A g1.

The inset in (f ) shows a schematic description of the cellulose paper electrodes. g,h) Nyquist plots for the pristine electrode (blue) and after 10 (red line) and 75 (black line) electrochemical cycles.

(6)

V2O3þ 2 Hþþ 2 e⇌ 2 VO þ H2O E°¼ 0.65 V versus Ag=AgClð1 м KClÞ (4) V2O3þ 6 Hþþ 6 e⇌ 2 V þ 3 H2O E°¼ 0.90 V versus Ag=AgClð1 м KClÞ (5) V2O5þ H2⇌ 2 VO2þ H2O (6) 2 VO2þ H2⇌ V2O3þ H2O (7)

On the subsequent anodic scan, two main oxidation peaks are seen at about 0.15 and 0.65 V versus Ag/AgCl (1M KCl) in Figure 2b. These oxidation peaks most likely correspond to the oxidation of Vþ3 (i.e., V2O3) to Vþ4 (i.e., VO2) and Vþ4

(i.e., VO2) to Vþ5(i.e., V2O5), respectively. The increased

oxida-tion charge compared with for cycling between0.45 and 0.55 V versus Ag/AgCl (1M KCl) (see the red voltammogram in

Figure 2b) indicates that the generated hydrogen assisted in the reduction of the electrode, as indicated in Reactions (6) and (7). The more positive oxidation potentials seen for the black voltammogram compared with the red voltammogram in Figure 2b further indicate that the local pH at the electrode sur-face was increased due to the preceding hydrogen evolution.

Based on the electrochemical performance shown in Figure 2b, a potential window of0.45 to 0.55 V versus Ag/AgCl (1MKCl) was used in the cycling experiments discussed in the following. The results obtained with the VOx–CNT–CC paper electrodes (see Figure 2d) indicated that the accessible capacity was almost approaching 200 C g1at the scan rates of 2, 5, 10, and 20 mV s1. As the theoretical capacity of V2O5⋅ H2O,

assum-ing a complete reduction of V2O5to V2O3and one water

mole-cule, is about 1930 C g1, it is immediately clear that the measured capacity of the VOx–CNT–CC paper electrodes corre-sponded to about 10% of the theoretical capacity. This is, how-ever, not unexpected as it is very unlikely that the theoretical capacity is reached experimentally, especially for this type of a two-step reduction process. As has been demonstrated for SnO2 and Cu2O electrodes in Li-ion batteries,[63–65] both the

oxide reduction and reformation steps are rarely complete unless the oxide layer is very thin (on the order of nanometers). The reason for this is that a reduction of V2O5to VO2 should give

rise to a passivating (i.e., reduced) VO2layer on top of the

remain-ing V2O5while a passivating oxide layer of V2O5would be formed

on top of the VO2during the subsequent oxidation step. In the

V2O5case, this effect (which incidentally should result in broader

reduction and oxidation peaks) should appear both during the V2O5 and VO2 reduction and oxidation steps. After the first

reduction scan, there could, hence, be a layer of V2O3 on top

of a layer of VO2on top of a layer of the pristine V2O5if the

V2O5layer was too thick to be completely reduced during the

scan prior to the onset of the reduction of VO2 to V2O3.

Analogously, a layer of V2O5on top of a layer of VO2 on top

of any remaining V2O3(on top of any VO2and V2O5still present

after the reduction step) would be expected to be found after the first oxidation cycle. As is indicated in Reaction (1) and (2), Hþ

would need to diffuse from the solution and into the material during the reduction steps and out toward the solution during the oxidation steps to maintain charge neutrality in the material. As a result of the diffusion-controlled reactions, the shapes of the

voltammograms (particularly at higher scan rates) become anal-ogous to those of voltammograms obtained with supercapacitor electrodes (i.e., EDL charge storage mechanism). This fact inci-dentally explains why capacitances (expressed in F g1) rather than capacities (expressed in C g1 or mAh g1) are often reported for vanadium oxide-based electrodes in conjunction with energy storage devices. In the present case, the currents are, however, clearly due to redox reactions with inherent specific redox potentials, although the peaks become more and more drawn out as a result of the diffusion-controlled processes discussed earlier. This, consequently, means that the capacity (or capacitance) will be strongly dependent on the thickness of the V2O5⋅ H2O nanosheets and the time scale of the

experi-ments. A restacking of the V2O5⋅ H2O nanosheets to form

thicker layers, hence creating long diffusion paths in the oxide layers, cannot be excluded. The present results, consequently, indicate that the relatively low capacity was due to the fact that mainly the surface of the VOx–CNT–CC material was

electro-chemically active (see inset in Figure 2f ), and that the specific capacity should decrease when the thickness of the V2O5⋅ H2O nanosheets is increased. Results obtained with an

electrode coated with a thin nanosheet layer should, hence, only be used to predict the capacity of an electrode with a mass loading of mg cm2if a restacking of the layers can be prevented during the upscaling of the electrode mass loading.

The mass loading of the electrode should also affect the iR (i.e., the ohmic) drop in the electrodes as the currents obtained in the voltammograms, or used in the constant current experi-ments, should increase when increasing the mass loading. If the iR drop is increased, the capacity should decrease as a result of a decrease in the width of the potential window as the potential or cutoff limits would be reached prematurely. Interestingly, the peak-to-peak separations did not increase significantly when increasing the scan rate from 2 to 20 mV s1(see Figure 2c). However, when higher scan rates were used (i.e., 50–500 mV s1), the capacity decreased with increasing rate (see Figure 2d), and the voltammograms also became significantly distorted (see Figure S2d, Supporting Information) due to the increased iR drop. Voltammograms with analogous shapes were also obtained when the potential window was expanded from0.8 to 0.8 V ver-sus Ag/AgCl (1MKCl) (see Figure S2b, Supporting Information).

The coulombic efficiency was about 95% at the low rates (due to the increased chance of the slow parasitic reactions such as H2

evolution) and 100% at the high rates, indicating a high reversibil-ity of the redox reactions. This is in a good agreement with previ-ousfindings for the abovementioned SnO2 electrodes[63–65] for

which a steady-state situation was seen to develop as the initially larger reduction charge gradually decreased tofinally match the oxidation charge even though the obtained capacity was signi fi-cantly lower than the theoretical capacity.

Galvanostatic charge–discharge experiments were likewise conducted in the potential window between0.45 and 0.55 V versus Ag/AgCl (1MKCl). Overall, the galvanostatic results were

analogous to the cyclic voltammetric results. The sloping reduc-tion (i.e., discharge) curves shown in Figure 2e, as well as the broad voltammetric peaks in Figure 2c, can be explained by the poor crystallinity of the V2O5⋅ H2O nanosheets, reflected

in the XRD pattern (see Figure S1a, Supporting Information) fea-turing only broad 00l peaks.[11,20,23] The reduction (discharge)

(7)

capacities were about 250, 200, 170, 136, and 44 C g1with the current densities of about 1, 3, 5, 10, and 40 A g1. These values are comparable to those obtained in the voltammetric experiments even though a strict comparison is difficult to carry out, because the effective reduction time in the voltammetric case depended on the reduction onset potential. The capacity retention observed at high rates (i.e., 5–40 A g1), which ranged between 70% and 20%, was comparable to that obtained in the voltammetric experiment (see Figure 2d). It was also found that the contribution of the CNTs to the measured capacity was about 20 C g1compared with the 250 C g1obtained with the VOx–CNT–CC electrodes at a

cur-rent density of 1 A g1(see Figure 2f ). This demonstrates that the major part of the capacity stemmed from the V2O5⋅ H2O

nano-sheets. This is in a good agreement with the corresponding com-parison of the voltammetric behaviors of the VOx–CNT–CC and CNT–CC electrodes shown in Figure S2e,f, Supporting Information. The observed redox peaks at about 0.36 and 0.45 V versus Ag/AgCl (1MKCl) can, most likely, be explained

by surface confined reduction and oxidation of the CNT surfaces, respectively. These CNT redox peaks cannot be observed in the VOx–CNT–CC voltammograms due to the larger the current orig-inating from the V2O5⋅ H2O. A comparison of the results

obtained with the present VOx–CNT–CC electrodes and other

V2O5-based electrodes is presented in Table S2, Supporting

Information. The latter illustrates the good rate performance and long-term cycling stability of the present cellulose paper elec-trodes as well as the high specific capacity that could be obtained despite the use of a mass loading of about 2 mg cm2.

2.3. Insights Regarding the Long-Term Cycling of the VOx–CNT–CC Paper Electrodes

To explore the factors affecting the electrochemical behavior of the VOx–CNT–CC electrode during long-term cycling as well as

the reproducibility of these results, experiments were conducted using three electrodes with the same size and mass loading (i.e., 0.32 cm2, 650μg), and the detailed experiments are summa-rized and discussed in Note S1, Supporting Information, and Figure S3 and S4. The results showed that the capacity retention of VOx–CNT–CC electrodes varied between 85% and 46% after

8000 cycles. This wide range may be attributed to a difference in the concentration of V2O5⋅ H2O close to the surface of the

different electrodes (see CVs in Figure S3a, Supporting Information). Another possibility is that the thicknesses of the V2O5⋅ H2O layers were not the same everywhere in the three

electrodes, e.g., due to different degrees of restacking, even though their mass loadings were very similar. As the V2O5⋅ H2O nanosheets most likely were thinner and/or located

closer to the surface of the electrodes with fast fading, therefore, the difference in the capacity retention is unlikely to have been due to the electrochemical cycling effects involving the formation of passivating oxide layers[63–65]discussed earlier. This suggests that the capacity losses were more likely to have been caused by vanadium oxide dissolution in the electrolyte, further motivated in Note S1, Supporting Information.

One way to reduce a fading capacity could be to use a surface coating, which is able to slow down the vanadium dissolution rate while not obstructing the diffusion of the charge

compensating protons. MXenes are promising 2D materials, which can exhibit high conductivities and capacitances, which, in turn, make them good candidates for use as surface coatings on the VOx–CNT–CC paper electrodes.[17,66]Therefore, Mo

1.33C

MXene surface coatings (containing 80μg cm2, implying a thickness of about 120 nm based on a density of Mo1.33C[17]of

3.4 g cm3) were here deposited on VOx–CNT–CC electrodes by drop-casting an MXene aqueous suspension (0.7– 1.0 mg mL1) on both sides of the electrodes; see SEM and EDX mapping in Figure S5, Supporting Information. The elec-trodes werefirst dried in air for 30–60 min, at room temperature, after which they were stored in an inert argon atmosphere. This MXene-coating approach did, however, not result in a sig-nificantly improved long-term cycling performance as analogous capacity losses were also seen for the Mo1.33C-coated electrodes

(see Figure S6, Supporting Information).

2.4. Influence of the Mass Loading on the Rate Capability and Long-Term Cycling Stability

The performance of electrodes with different dimensions and mass loadings was also compared using three electrodes referred to as large (0.50 cm2, 950μg), moderate (0.32 cm2, 650μg), and small (0.08 cm2, 150–240 μg), and schematically described in

Figure 3a. As shown in Figure 3b, the CVs for the large and small electrodes had similar overall shapes even though the sharper peaks seen for the small electrode indicated a higher surface con-centration of the V2O5⋅ H2O nanosheets. The shifts in the peak

potentials as well as the distortions of the voltammograms shown in Figure 3b,c with increasing the scan rate can be explained by the increased iR drop.

As expected, the capacity depended on the mass loading of the electrode (see Figure 3d), indicating that the electroactive fraction of the V2O5⋅ H2O nanosheets was about the same in the large

and small electrodes. The capacity retentions at 500 mV s1were also analogous, i.e., 25% and 28% of the capacity at 2 mV s1for the large and small electrode, respectively. The latter indicates that the capacity losses seen at high scan rates mainly were a result of the diffusion limited redox reactions, although the iR drop also gave rise to smaller active potential windows and, hence, smaller capacities. The larger specific capacity seen for the small electrode compared with that of the large one at the lowest scan rates can most likely be explained by the abovemen-tioned differences between the surface concentrations of the V2O5⋅ H2O nanosheets. The latter hypothesis is supported by

the fact that the rate dependences of the moderate and large electrodes (see Figure 3e) did not differ significantly.

While the large electrode showed a capacity retention of about 80% (see Figure S7a, Supporting Information) during 5000 cycles at a current density of about 10 A g1, a capacity retention of about 25% was obtained for the small electrode after 1000 cycles at a current density of 1 or 6.25 A g1 (see Figure S7b,c, Supporting Information). This difference was most likely a result of the previously mentioned formation of dissolv-able vanadium species and the higher surface concentration of the V2O5⋅ H2O nanosheets for the small electrode. The influence

of the latter effect was also enhanced by the lower amount of active material in the small electrode (i.e., 150–240 μg) as this

(8)

should decrease the likelihood of obtaining a saturated solution in the vicinity of the electrode. The initial increase in the capacity seen for the small electrode in Figure S7c, Supporting Information, was probably caused by a gradual wetting of the electrode yielding an increasing electrode area. All in all, the opti-mum loadings and dimensions of the electrodes will depend on thefinal use of the charge storage device; however, they should be considered carefully during the interpretation of the electro-chemical behavior.

2.5. Electrochemical Performance of VOx–CNT–CC Paper

Electrodes Symmetric Device

To investigate the performance of the V2O5⋅ H2O–CNT–CC

paper electrodes in an energy storage device, a symmetric

two-electrode device was assembled and cycled over a cell voltage window of 1 V. As seen from the obtained CVs recorded at dif-ferent scan rates and displayed in Figure 4a,b, the behavior of the device was analogous to that seen for a supercapacitor, which is why both the specific electrode capacities and capacitances have been reported in Table S2, Supporting Information. Here, it should be pointed out that the mass loadings used in the experi-ment should be considered when comparing different values as very high specific capacities or capacitances can be obtained using electrodes with very low mass loadings even though the capacities or capacitances of such electrodes would be too small to be of any practical interest. The device capacities were about 50 mC (40 C g1) and 12 mC (10 C g1) at the scan rates of about 2 and 1000 mV s1, respectively (see Figure 4e) after normaliza-tion using the masses of the active materials in both two electro-des. In agreement with the voltammetric results, the constant

Figure 3. Effect of the electrode mass loading on the rate capability: a) schematic illustration of the large, moderate, and small electrodes. b,c) CVs for the large (solid lines) and small (dotted lines) electrodes at a scan rate of 10 mV s1(blue), 20 mV s1(grey), 100 mV s1(green), and 200 mV s1(purple), respectively. The d) capacity and e) specific capacity, as a function of the logarithm of the scan rate for the large (black squares), moderate (red diamonds), and small (blue triangles) electrodes.

(9)

current measurements gave rise to sloping charge–discharge profiles, as shown in Figure 4c. The device capacity evaluated from the latter curves was about 40 mC (34 C g1) and 13 mC (11 C g1) using the current densities of about 0.5 and 10 A g1, respectively (see Figure 4f ). The low normalized device capacities compared with the theoretical capacity of about 480 C g1can be attributed to the diffusion limitations associated with the redox reactions discussed earlier. The long-term cycling of the symmetric device showed (see Figure 4d) that the capacity retentions after 1000 and 2000 cycles were of about 86% and 60%, respectively. As discussed earlier, this decrease in the capacity retention can be explained by vanadium species under-going dissolution in the electrolyte. Despite this, the coulombic

efficiency remained close to 100% during 2000 cycles, reflecting the high reversibility of the electroactive parts of the electrodes. It can, consequently, be concluded that the coulombic efficiency constitutes a rather blunt instrument when evaluating the perfor-mance of this type of energy storage devices.

3. Conclusion

A straightforward and novel approach for fabricating vanadium oxide-based electrodes (VOx–CNT–CC), using a porous CNT–CC matrix, was described for use in aqueous energy storage devices. The VOx–CNT–CC electrodes with the mass loadings of about

Figure 4. Electrochemical performance of the symmetric device: a,b) CVs recorded at different scan rates. c) Potential versus capacity curves for different current densities. d) The capacity (red diamonds), capacity retention (black circles), and coulombic efficiency (blue circles) as a function of the cycle number. e,f ) The normalized capacity (red diamonds) and absolute capacity (blue squares) as a function of the logarithm of the scan rate and current density, respectively.

(10)

2 mg cm2exhibited the capacities of about 250 and 44 C g1in 1M Na2SO4 at the current densities of about 1 and 40 A g1,

respectively. The long-term cycling of the electrodes was in flu-enced by a dissolution of vanadium species in the electrolyte; consequently, the capacity retention varied between 46% and 85% after 8000 cycles, depending on the thickness of the V2O5⋅ H2O layer and/or the V2O5⋅ H2O concentration at the

sur-face of the VOx–CNT–CC electrodes. The use of thin coating of Mo1.33C MXene applied on the surface of the VOx–CNT–CC

elec-trodes was also investigated; however, it did not efficiently mini-mize the dissolution. The vanadium oxide was well distributed within the matrix as a whole; however, variations in the depth distribution and the thickness of the V2O5⋅ H2O nanosheets

influenced the electrochemical results significantly due to the accessible capacity being determined by diffusion controlled redox reactions involving the reduction of V2O5 to VO2 and

VO2 to V2O3. A symmetric device containing two VOx–CNT–

CC paper electrodes was found to exhibit a capacity of 40 C g1at a scan rate of 2 mV s1, which can be attributed to the surface of the electrode being electrochemically active as a result of diffusion limitations. To realize the full potential of these materials, further studies are needed to develop approaches for immobilizing nanometer thin layers of vanadium oxide (e.g., vanadium oxide nanosheets) on porous,flexible, and conducting substrates, and to prevent possible restacking of the vanadium oxide nanosheets used herein during the electrode fabrication.

4. Experimental Section

Synthesis of Nanostructured V2O5⋅ nH2O and VOx–CNT–CC Paper

Electrodes: V2O5(1 mmol, i.e., 0.182 g) (Beijing chemical reagent industry)

was dispersed in deionized water by sonication for 10 min, after which 0.5 mmol (i.e., 0.630 g) of oxalic acid, H2C2O4⋅ 2H2O (Xilong chemical

Ltd.), was added. The suspension was purged with argon for 5 min and stirred under reflux at 80C for 24 h. The resulting black-greenish

suspension of V2O5⋅ nH2O was dried at 60C for 10 h to obtain the

nano-structured V2O5⋅ nH2O (withn1 at 25C). The synthesis process was

robust and could be scaled-up to yield 50 g.[57]

The paper electrodes were fabricated similar to the previously reported protocol.[51,52,56]In particular, 70 mg of V

2O5⋅ nH2O nanosheet powder,

20 mg of multi-walled CNTs, and 10 mg of CC were dispersed in a water– ethanol mixture (60 mL of water and 20 mL of ethanol) and sonicated for 20 min. The formed mixture wasfiltrated under vacuum on a polypropyl-enefilter to produce a paper sheet. The paper sheet was allowed to dry at room temperature and then peeled off the polypropylenefilter yielding a freestanding paper electrode. The overall loading including the nanocellu-lose was about 2 mg cm2, and active mass loading (mass of V2O5⋅ nH2O

and CNTs) was about 1.8 mg cm2. Similarly, the CNT–CC paper electro-des were synthesized using a mixture of CNT and CC of 1:1 weight ratio. Material Characterization and Electrochemical Measurements: An SEM (LEO 1550 Gemini) equipped with an EDX detector was used to study the morphology and thickness of the VOx–CNT–CC paper. The structure

of VOx–CNT–CC paper was also examined by the XRD analysis using a

PANalytical diffractometer equipped with copper Kα radiation source

(λ ¼ 1.54 Å).

All electrochemical experiments were performed in a stainless-steel Swagelok cell. Gold foil was used as a current collector, whereas a piece ofCelgard 3501 soaked in 1.0MNa2SO4(500μL) was used as the

separa-tor. Cyclic voltammetry and galvanostatic charge–discharge experiments were used to examine the electrochemical behavior of the VOx–CNT–CC

paper electrodes. In the three-electrode measurements, the VOx–CNT–CC

paper electrodes were used as the working electrodes, whereas a circular piece of activated carbon (8 mm diameter, with a mass loading of 5–8 mg,

YP-50, Kuraray, Japan) was used as the counter electrode, and a Ag/AgCl (1MKCl) electrode was used as the reference electrode. The potential was generally scanned between0.45 and 0.55 V versus Ag/AgCl (1MKCl), but some experiments were also conducted in the potential window 0.8 to 0.8 V versus Ag/AgCl (1M KCl). The overall working electrode mass was varied between 150 and 950μg, and the capacity was normal-ized with respect to the VOx–CNT mass (i.e., 90% of the total electrode

mass). The working electrodes had the diameters of about 8.0 mm corre-sponding to an area of about 0.5 cm2(large electrode) and 6.4 mm cor-responding to an area of about 0.32 cm2 (moderate electrode). Some

experiments were also done using rectangular electrodes with the dimen-sions of 2–3 mm corresponding to an area of about 0.08 cm2

(small electrode).

The two-electrode symmetric device experiments were done using a couple of symmetric circular (6.4 mm diameter) VOx–CNT–CC paper

elec-trodes. The total dry mass of each electrode (prior to the assembly of the device) was about 650μg, and the mass normalization was done with respect to total the active mass of the two electrodes (i.e., 2 650  0.9 μg). Cyclic voltammetry and constant current experiments were used to explore the electrochemical performances of the symmetric devices for cell voltages between 0.0 and 1.0 V.

The standard potentials indicted in association with the reactions given as follows were calculated based on the change in ΔG0, using ΔG0¼ nFE0

and publishedΔG0values.[67]

Supporting Information

Supporting Information is available from the Wiley Online Library or from the author.

Acknowledgements

J.R. acknowledges support from the Swedish Foundation for Strategic Research (SSF) for Project Funding (EM16-0004) and the Knut and Alice Wallenberg (KAW) Foundation for a Fellowship grant. Z.W. acknowl-edges support from the Swedish Research Council (grant No. VR-2019-03492).

Conflict of Interest

The authors declare no conflict of interest.

Keywords

carbon nanotubes, cellulose nanofibers, energy storage, freestanding paper electrodes, vanadium oxide nanosheets

Received: August 13, 2020 Revised: September 18, 2020 Published online: October 26, 2020

[1] M. Winter, R. J. Brodd,Chem. Rev. 2004, 104, 4245. [2] M. S. Whittingham,Chem. Rev. 2004, 104, 4271.

[3] N. Yabuuchi, K. Kubota, M. Dahbi, S. Komaba,Chem Rev. 2014, 114, 11636.

[4] C. Choi, D. S. Ashby, D. M. Butts, R. H. DeBlock, Q. Wei, J. Lau, B. Dunn,Nat. Rev. Mater. 2020, 5, 5.

[5] L. Nyholm, G. Nyström, A. Mihranyan, M. Strømme, Adv. Mater. 2011,23, 3751.

[6] L. Yuan, X. Xiao, T. Ding, J. Zhong, X. Zhang, Y. Shen, B. Hu, Y. Huang, J. Zhou, Z. L. Wang,Angew. Chem., Int. Ed. 2012, 51, 4934.

(11)

[7] F. Wang, X. Wu, X. Yuan, Z. Liu, Y. Zhang, L. Fu, Y. Zhu, Q. Zhou, Y. Wu, W. Huang,Chem. Soc. Rev. 2017, 46, 6816.

[8] C. Tan, X. Cao, X.-J. Wu, Q. He, J. Yang, X. Zhang, J. Chen, W. Zhao, S. Han, G.-H. Nam, M. Sindoro, H. Zhang,Chem. Rev. 2017, 117, 6225. [9] H. Zhao, Y. Zhu, F. Li, R. Hao, S. Wang, L. Guo,Angew. Chem., Int. Ed.

2017,56, 8766.

[10] A. S. Etman, L. Wang, K. Edström, L. Nyholm, J. Sun,Adv. Funct. Mater. 2019, 29, 1806699.

[11] A. S. Etman, A. K. Inge, X. Jiaru, R. Younesi, K. Edström, J. Sun, Electrochim. Acta 2017, 252, 254.

[12] A. S. Etman,Aqueous Exfoliation of Transition Metal Oxides for Energy Storage and Photocatalysis Applications: Vanadium Oxide and Molybdenum Oxide Nanosheets, Stockholm University, Stockholm 2019.

[13] A. S. Etman, A. J. Pell, P. Svedlindh, N. Hedin, X. Zou, J. Sun, D. Bernin,ACS Omega 2019, 4, 10899.

[14] W. Choi, N. Choudhary, G. H. Han, J. Park, D. Akinwande, Y. H. Lee, Mater. Today 2017, 20, 116.

[15] M. Chhowalla, H. S. Shin, G. Eda, L.-J. Li, K. P. Loh, H. Zhang, Nat. Chem. 2013, 5, 263.

[16] B. Anasori, M. R. Lukatskaya, Y. Gogotsi,Nat. Rev. Mater. 2017, 2, 16098.

[17] Q. Tao, M. Dahlqvist, J. Lu, S. Kota, R. Meshkian, J. Halim, J. Palisaitis, L. Hultman, M. W. W. Barsoum, P. O. Å. Persson, J. Rosen,Nat. Commun. 2017, 8, 14949.

[18] Z. Fu, N. Wang, D. Legut, C. Si, Q. Zhang, S. Du, T. C. Germann, J. S. Francisco, R. Zhang,Chem. Rev. 2019, 119, 11980.

[19] M. Ghidiu, M. R. Lukatskaya, M.-Q. Zhao, Y. Gogotsi, M. W. Barsoum,Nature 2014, 516, 78.

[20] A. S. Etman, H. D. Asfaw, N. Yuan, J. Li, Z. Zhou, F. Peng, I. Persson, X. Zou, T. Gustafsson, K. Edström, J. Sun,J. Mater. Chem. A 2016, 4, 17988.

[21] A. Moretti, S. Passerini,Adv. Energy Mater. 2016, 6, 1600868. [22] J. Zhu, L. Cao, Y. Wu, Y. Gong, Z. Liu, H. E. Hoster, Y. Zhang,

S. Zhang, S. Yang, Q. Yan, P. M. Ajayan, R. Vajtai, Nano Lett. 2013,13, 5408.

[23] A. S. Etman, J. Sun, R. Younesi,J. Energy Chem. 2019, 30, 145. [24] A. Moretti, M. Secchiaroli, D. Buchholz, G. Giuli, R. Marassi,

S. Passerini,J. Electrochem. Soc. 2015, 162, A2723.

[25] M. Yan, P. He, Y. Chen, S. Wang, Q. Wei, K. Zhao, X. Xu, Q. An, Y. Shuang, Y. Shao, K. T. Mueller, L. Mai, J. Liu, J. Yang, Adv. Mater. 2018, 30, 1703725.

[26] P. Senguttuvan, S.-D. Han, S. Kim, A. L. Lipson, S. Tepavcevic, T. T. Fister, I. D. Bloom, A. K. Burrell, C. S. Johnson,Adv. Energy Mater. 2016, 6, 1600826.

[27] V. Augustyn, P. Simon, B. Dunn,Energy Environ. Sci. 2014, 7, 1597. [28] D. Majumdar, M. Mandal, S. K. Bhattacharya, ChemElectroChem

2019,6, 1623.

[29] D. McNulty, D. N. Buckley, C. O’Dwyer, J. Power Sources 2014, 267, 831.

[30] Y. Zhang, J. Zheng, Q. Wang, S. Zhang, T. Hu, C. Meng,Appl. Surf. Sci. 2017, 423, 728.

[31] W. Bi, E. Jahrman, G. Seidler, J. Wang, G. Gao, G. Wu, M. Atif, M. Alsalhi, G. Cao,ACS Appl. Mater. Interfaces 2019, 11, 16647. [32] J. Bao, X. Zhang, L. Bai, W. Bai, M. Zhou, J. Xie, M. Guan, J. Zhou,

Y. Xie,J. Mater. Chem. A 2014, 2, 10876.

[33] B. Pandit, D. P. Dubal, P. Gómez-Romero, B. B. Kale, B. R. Sankapal, Sci. Rep. 2017, 7, 43430.

[34] R. Narayanan,J. Solid State Chem. 2017, 253, 103.

[35] J. Zheng, Y. Zhang, Q. Wang, H. Jiang, Y. Liu, T. Lv, C. Meng,Dalt. Trans. 2018, 47, 452.

[36] T. Hu, Y. Liu, Y. Zhang, M. Chen, J. Zheng, J. Tang, C. Meng, J. Colloid Interface Sci. 2018, 531, 382.

[37] M. Chen, Y. Zhang, Y. Liu, Q. Wang, J. Zheng, C. Meng,ACS Appl. Energy Mater. 2018, 1, 5527.

[38] Y. Zhang, M. Chen, T. Hu, C. Meng,ACS Appl. Nano Mater. 2019, 2, 2934.

[39] J. Zheng, Y. Zhang, C. Meng, X. Wang, C. Liu, M. Bo, X. Pei, Y. Wei, T. Lv, G. Cao,Electrochim. Acta 2019, 318, 635.

[40] M. J. Deng, L. H. Yeh, Y. H. Lin, J. M. Chen, T. H. Chou,ACS Appl. Mater. Interfaces 2019, 11, 29838.

[41] X. Rui, J. Zhu, W. Liu, H. Tan, D. Sim, C. Xu, H. Zhang, J. Ma, H. H. Hng, T. M. Lim, Q. Yan,RSC Adv. 2011, 1, 117.

[42] J. Shojaeiarani, D. Bajwa, A. Shirzadifar, Carbohydr. Polym. 2019, 216, 247.

[43] J. Wang, J. Tavakoli, Y. Tang,Carbohydr. Polym. 2019, 219, 63. [44] O. Nechyporchuk, M. N. Belgacem, J. Bras,Ind. Crops Prod. 2016,

93, 2.

[45] D. Tobjörk, R. Österbacka,Adv. Mater. 2011, 23, 1935.

[46] A. Razaq, L. Nyholm, M. Sjödin, M. Strømme, A. Mihranyan, Adv. Energy Mater. 2012, 2, 445.

[47] G. Nyström, A. Razaq, M. Strømme, L. Nyholm, A. Mihranyan,Nano Lett. 2009, 9, 3635.

[48] H. Olsson, G. Nyström, M. Strømme, M. Sjödin, L. Nyholm, Electrochem. commun. 2011, 13, 869.

[49] B. Yao, J. Zhang, T. Kou, Y. Song, T. Liu, Y. Li,Adv. Sci. 2017, 4, 1700107.

[50] Z. Wang, P. Tammela, M. Strømme, L. Nyholm,Adv. Energy Mater. 2017,7, 1700130.

[51] Z. Wang, C. Xu, P. Tammela, P. Zhang, K. Edström, T. Gustafsson, M. Strømme, L. Nyholm,Energy Technol. 2015, 3, 563.

[52] Z. Wang, C. Xu, P. Tammela, J. Huo, M. Strømme, K. Edström, T. Gustafsson, L. Nyholm,J. Mater. Chem. A 2015, 3, 14109. [53] Z. Wang, P. Tammela, M. Strømme, L. Nyholm,Nanoscale 2015,

7, 3418.

[54] Z. Wang, P. Tammela, P. Zhang, J. Huo, F. Ericson, M. Strømme, L. Nyholm,Nanoscale 2014, 6, 13068.

[55] S. Zhou, L. Nyholm, M. Strømme, Z. Wang,Acc. Chem. Res. 2019, 52, 2232.

[56] A. S. Etman, Z. Wang, A. El Ghazaly, J. Sun, L. Nyholm, J. Rosen, ChemSusChem 2019, 12, 5157.

[57] A. S. Etman, J. Sun, Y. Yuan, Synthesis of Vanadium Pentoxide Nanosheets, 2018, WIPO WO2018013043 A8.

[58] Z. Wang, Y. Lee, S. Kim, J. Seo, S. Lee, L. Nyholm,Adv. Mater. 2020, 10, 2000892.

[59] K. Kanamori, K. Tsuge,Inorganic Chemistry of Vanadium, Springer Netherlands, Dordrecht 2012.

[60] M. Pourbaix,Atlas of Electrochemical Equilibria in Aqueous Solutions, Pergamon Press, Oxford, New York, 1966.

[61] R. K. Haghi, A. Chapoy, L. M. C. Peirera, J. Yang, B. Tohidi,Int. J. Greenh. Gas Control 2017, 66, 190.

[62] Z.-D. Geng, Y. Wang,J. Solid State Electrochem. 2015, 19, 3131. [63] S. Böhme, K. Edström, L. Nyholm,J. Electroanal. Chem. 2017, 797, 47. [64] S. Böhme, B. Philippe, K. Edström, L. Nyholm,J. Phys. Chem. C 2017,

121, 4924.

[65] D. Rehnlund, M. Valvo, C.-W. Tai, J. Ångström, M. Sahlberg, K. Edström, L. Nyholm,Nanoscale 2015, 7, 13591.

[66] M. Naguib, M. Kurtoglu, V. Presser, J. Lu, J. Niu, M. Heon, L. Hultman, Y. Gogotsi, M. W. Barsoum,Adv. Mater. 2011, 23, 4248. [67] G. H. Aylward, T. J. V. Findlay,SI Chemical Data, Wiley, Milton, QLD

References

Related documents

Industrial Emissions Directive, supplemented by horizontal legislation (e.g., Framework Directives on Waste and Water, Emissions Trading System, etc) and guidance on operating

46 Konkreta exempel skulle kunna vara främjandeinsatser för affärsänglar/affärsängelnätverk, skapa arenor där aktörer från utbuds- och efterfrågesidan kan mötas eller

Both Brazil and Sweden have made bilateral cooperation in areas of technology and innovation a top priority. It has been formalized in a series of agreements and made explicit

För att uppskatta den totala effekten av reformerna måste dock hänsyn tas till såväl samt- liga priseffekter som sammansättningseffekter, till följd av ökad försäljningsandel

The increasing availability of data and attention to services has increased the understanding of the contribution of services to innovation and productivity in

Generella styrmedel kan ha varit mindre verksamma än man har trott De generella styrmedlen, till skillnad från de specifika styrmedlen, har kommit att användas i större

Parallellmarknader innebär dock inte en drivkraft för en grön omställning Ökad andel direktförsäljning räddar många lokala producenter och kan tyckas utgöra en drivkraft

Närmare 90 procent av de statliga medlen (intäkter och utgifter) för näringslivets klimatomställning går till generella styrmedel, det vill säga styrmedel som påverkar