6
A86 Journal of The Electrochemical Society, 162 (1) A86-A91 (2015) 0013-4651/2015/162(1)/A86/6/$31.00 © The Electrochemical Society High Rate Capacity through Redox Electrolytes Confined in Macroporous Electrodes R. Narayanan a and P. R. Bandaru a,b, z a Department of Nanoengineering, University of California, San Diego, La Jolla, California 92093, USA b Program in Materials Science, Department of Mechanical and Aerospace Engineering, University of California, San Diego, La Jolla, California 92093, USA The capacitive characteristics of a model macroporous electrode, comprised of vertically aligned carbon nanotubes, with a relatively large volume /unit mass (of 40 cm 3 /g) are discussed. The faradaic characteristics of the capacity were evidenced through a voltage plateau in the galvanostatic discharge experiments. It is shown that the confinement of electrolyte within the spacing between the nanostructures with respect to an appropriate choice of the voltage scan rate yields a regime where both high capacity and rate capability may be realized. Additionally, the absence of diffusional processes allows for large capacitance retention even at discharge current densities of the order of 200 mA/cm 2 . For 1M K 3 Fe(CN) 6 , the maximum observed capacity was found to be 26 mAh/cm 3 (normalized to the pore volume) or 0.28 mAh/cm 2 (normalized to the projected area of the electrode) and it was noted that the capacity could be cycled at very high rates for 5000 cycles. © 2014 The Electrochemical Society. [DOI: 10.1149/2.0461501jes] All rights reserved. Manuscript submitted October 1, 2014; revised manuscript received November 4, 2014. Published November 14, 2014. Electrochemical capacitors (ECs) typically store charge either (a) electrostatically (in the double layer at the electrode/electrolyte in- terface, as an electrochemical double-layer capacitor: EDLC) or (b) pseudocapacitively 1 (at the electrode surface through redox reaction mediated mechanisms). The latter mode of charge storage in ECs re- lies on faradaic charge transfer, may occur over a relatively smaller voltage range compared to EDLCs, and varies as a function of the sur- face coverage area. Such pseudocapacitive characteristics have been observed in transition metal oxides (such as RuO 2 , MnO 2 , etc.) as well as conductive polymers (such as polyaniline) 24 and seem to be related to specific functional groups. At the very outset, a major benefit of capacitive methods of stor- ing charge is the high power density that may be obtained due to the charge being predominantly on the surface. However, this very same attribute results in a lower energy density and consequently, there have been significant efforts to raise the energy density to the level of batteries through using interstitial spaces within the mate- rial. For instance, an intercalation pseudocapacitance 5,6 has been re- cently proposed as due to nanostructuring, in oxides such as TiO 2 and Nb 2 O 5 , and closely mimics battery-like behavior, while avoid- ing diffusional limitations. Such charge storage may also be realized through the use of high surface area nanostructured materials, such as activated carbon (AC) and carbon nanotubes (CNTs) 79 in con- junction with redox electrolytes. Optimized pore diameter or struc- ture spacing mitigates diffusional constraints and could yield high charge/discharge rates without a detrimental effect on the cycling sta- bility, which is an issue when using surface functional groups 10 for pseudocapacitance based storage. Concomitantly, a combination of re- dox electrolytes, 11,12 e.g., KI (for the positive electrode) and vanadyl sulfate (VOSO 4 , for the negative electrode), or through the use of an electrochemically active compound such as hydroquinone (HQ) added to a sulfuric acid electrolyte, 13,14 were investigated, with AC consti- tuting the electrodes. However, it was seen 12 that the charge storage capacity was diminished at high discharge rates. Computational mod- eling of faradaic process in porous electrodes has found that low pore volume may indicate that reduced capacity may be due to planar diffu- sion to the exterior surface 15 and not within the pores of the electrode itself. It was also intriguing to consider that faradaic processes in porous materials may apparently possess superior kinetics otherwise absent in bulk materials. 16 Such electrocatalytic effects have been proposed to be due to thin layer electrochemistry (TLE). 17,18 When the nanos- tructure spacing or pore diameter (d) is smaller than the equivalent diffusion layer thickness (δ π Dt π D(V /v)) with a rele- vant ion diffusion coefficient, D, over a given time (t), and a voltage z E-mail: [email protected] range V (with ν being the voltage scan rate) TLE conditions 19 could be induced. Then, the charge contained in the enclosed electrolyte could be harnessed through faradaic reactions inside the spaces due to homogenous depletion/accretion of redox species. In addition to the enhanced kinetics that could be obtained due to the absence of dif- fusion, the capacity would be directly proportional to the electrolyte volume confined in the pore (V pore ) and its bulk concentration (C o ). Consequently, planar diffusion would not be a kinetically limiting step. It has been derived that the accompanying TLE based peak cur- rents (I TLE ) = F 2 4RT νV pore C o , would vary directly with ν (where F is the Faraday constant 96,485 C/mol, R is the gas constant = 8.31 J/mole K, and T is the room temperature of 298 K). 20 At a low scan rate (say, ν 1 ): see Figures 1a and 1b, implying a large δ, the diffusion layers at the top of the macro-electrode overlap giving rise to a planar diffusion limited current (I diff ), proportional to C 0 , the projected area (A proj ) of the electrode and ν 1/2 . Additionally, there would be a TLE based current (I TLE ) due to the electrolyte trapped within the pores, which would be proportional to the V pore and ν. At moderate scan rates (say, ν 2 ) the individual I diff contributions at the top of the nanotube array begins to get resolved. The resulting concentration gradients internal to the pores yields additional I diff while the I TLE current con- tribution reduces. At very high scan rates (say, ν 3 ), there would be a dominant I diff proportional to the net area (A net ) of the electrode owing to a complete decoupling of diffusion layers and the CNTs behaving as individual nanoscale electrodes. The amount of charge (or capac- ity) contributed by the solvated redox species confined in the V pore , in the TLE regime is given by Q = nFV pore C o , where n is the num- ber of electrons per redox reaction. Consequently, high capacity and rate capability may be realized in the TLE regime through designing electrodes with large V pore . While activated carbon based materials may possess a large active area (>2000 m 2 /g), their V pore 0.5 – 1 cm 3 /g would result in a smaller magnitude of accessible capacitive charge. 21 In this paper, we aim to assess in greater detail the capacitive characteristics, in the TLE regime, of a model macroporous electrode comprised of a large V pore (of 40 cm 3 /g) comprised of vertically aligned CNTs: Figure 1c, the spacing of which may correspond to an equivalent pore size. Experimental Vertically Aligned (VA)-CNT growth:-— Substrate preparation.— Substrates onto which CNTs were grown were first prepared by depositing a 5 nm thick Iron (Fe) catalyst layer on a silicon wafer (n- type silicon, 3 diameter) by electron beam evaporation using a Temescal BJD 1800 electron beam evaporator. The catalyst deposited wafer was then diced into 5 mm by 5 mm squares (yielding an equivalent projected area, A proj = 0.25 cm 2 ). ) unless CC License in place (see abstract). ecsdl.org/site/terms_use address. Redistribution subject to ECS terms of use (see 137.110.36.10 Downloaded on 2015-01-25 to IP

162 The Electrochemical Society High Rate Capacity through ...nanomaterials-dev.ucsd.edu/wp-content/uploads/2015/...High Rate Capacity through Redox Electrolytes Confined in Macroporous

  • Upload
    others

  • View
    4

  • Download
    0

Embed Size (px)

Citation preview

Page 1: 162 The Electrochemical Society High Rate Capacity through ...nanomaterials-dev.ucsd.edu/wp-content/uploads/2015/...High Rate Capacity through Redox Electrolytes Confined in Macroporous

A86 Journal of The Electrochemical Society, 162 (1) A86-A91 (2015)0013-4651/2015/162(1)/A86/6/$31.00 © The Electrochemical Society

High Rate Capacity through Redox Electrolytes Confined inMacroporous ElectrodesR. Narayanana and P. R. Bandarua,b,z

aDepartment of Nanoengineering, University of California, San Diego, La Jolla, California 92093, USAbProgram in Materials Science, Department of Mechanical and Aerospace Engineering, University of California,San Diego, La Jolla, California 92093, USA

The capacitive characteristics of a model macroporous electrode, comprised of vertically aligned carbon nanotubes, with a relativelylarge volume /unit mass (of ∼40 cm3/g) are discussed. The faradaic characteristics of the capacity were evidenced through a voltageplateau in the galvanostatic discharge experiments. It is shown that the confinement of electrolyte within the spacing between thenanostructures with respect to an appropriate choice of the voltage scan rate yields a regime where both high capacity and ratecapability may be realized. Additionally, the absence of diffusional processes allows for large capacitance retention even at dischargecurrent densities of the order of 200 mA/cm2. For 1M K3Fe(CN)6, the maximum observed capacity was found to be ∼26 mAh/cm3

(normalized to the pore volume) or 0.28 mAh/cm2 (normalized to the projected area of the electrode) and it was noted that thecapacity could be cycled at very high rates for 5000 cycles.© 2014 The Electrochemical Society. [DOI: 10.1149/2.0461501jes] All rights reserved.

Manuscript submitted October 1, 2014; revised manuscript received November 4, 2014. Published November 14, 2014.

Electrochemical capacitors (ECs) typically store charge either (a)electrostatically (in the double layer at the electrode/electrolyte in-terface, as an electrochemical double-layer capacitor: EDLC) or (b)pseudocapacitively1 (at the electrode surface through redox reactionmediated mechanisms). The latter mode of charge storage in ECs re-lies on faradaic charge transfer, may occur over a relatively smallervoltage range compared to EDLCs, and varies as a function of the sur-face coverage area. Such pseudocapacitive characteristics have beenobserved in transition metal oxides (such as RuO2, MnO2, etc.) aswell as conductive polymers (such as polyaniline)2–4 and seem to berelated to specific functional groups.

At the very outset, a major benefit of capacitive methods of stor-ing charge is the high power density that may be obtained due tothe charge being predominantly on the surface. However, this verysame attribute results in a lower energy density and consequently,there have been significant efforts to raise the energy density to thelevel of batteries through using interstitial spaces within the mate-rial. For instance, an intercalation pseudocapacitance5,6 has been re-cently proposed as due to nanostructuring, in oxides such as TiO2

and Nb2O5, and closely mimics battery-like behavior, while avoid-ing diffusional limitations. Such charge storage may also be realizedthrough the use of high surface area nanostructured materials, suchas activated carbon (AC) and carbon nanotubes (CNTs)7–9 in con-junction with redox electrolytes. Optimized pore diameter or struc-ture spacing mitigates diffusional constraints and could yield highcharge/discharge rates without a detrimental effect on the cycling sta-bility, which is an issue when using surface functional groups10 forpseudocapacitance based storage. Concomitantly, a combination of re-dox electrolytes,11,12 e.g., KI (for the positive electrode) and vanadylsulfate (VOSO4, for the negative electrode), or through the use of anelectrochemically active compound such as hydroquinone (HQ) addedto a sulfuric acid electrolyte,13,14 were investigated, with AC consti-tuting the electrodes. However, it was seen12 that the charge storagecapacity was diminished at high discharge rates. Computational mod-eling of faradaic process in porous electrodes has found that low porevolume may indicate that reduced capacity may be due to planar diffu-sion to the exterior surface15 and not within the pores of the electrodeitself.

It was also intriguing to consider that faradaic processes in porousmaterials may apparently possess superior kinetics otherwise absentin bulk materials.16 Such electrocatalytic effects have been proposedto be due to thin layer electrochemistry (TLE).17,18 When the nanos-tructure spacing or pore diameter (d) is smaller than the equivalentdiffusion layer thickness (δ ∼ √

πDt ∼ √πD(�V/v)) with a rele-

vant ion diffusion coefficient, D, over a given time (t), and a voltage

zE-mail: [email protected]

range � V (with ν being the voltage scan rate) TLE conditions19 couldbe induced. Then, the charge contained in the enclosed electrolytecould be harnessed through faradaic reactions inside the spaces due tohomogenous depletion/accretion of redox species. In addition to theenhanced kinetics that could be obtained due to the absence of dif-fusion, the capacity would be directly proportional to the electrolytevolume confined in the pore (Vpore) and its bulk concentration (Co).Consequently, planar diffusion would not be a kinetically limitingstep. It has been derived that the accompanying TLE based peak cur-rents (ITLE) = F2

4RT νVporeCo, would vary directly with ν (where F isthe Faraday constant −96,485 C/mol, R is the gas constant = 8.31J/mole K, and T is the room temperature of ∼298 K).20 At a low scanrate (say, ν1): see Figures 1a and 1b, implying a large δ, the diffusionlayers at the top of the macro-electrode overlap giving rise to a planardiffusion limited current (Idiff), proportional to C0, the projected area(Aproj) of the electrode and ν1/2. Additionally, there would be a TLEbased current (ITLE) due to the electrolyte trapped within the pores,which would be proportional to the Vpore and ν. At moderate scan rates(say, ν2) the individual Idiff contributions at the top of the nanotubearray begins to get resolved. The resulting concentration gradientsinternal to the pores yields additional Idiff while the ITLE current con-tribution reduces. At very high scan rates (say, ν3), there would be adominant Idiff proportional to the net area (Anet) of the electrode owingto a complete decoupling of diffusion layers and the CNTs behavingas individual nanoscale electrodes. The amount of charge (or capac-ity) contributed by the solvated redox species confined in the Vpore,in the TLE regime is given by Q = nFVporeCo, where n is the num-ber of electrons per redox reaction. Consequently, high capacity andrate capability may be realized in the TLE regime through designingelectrodes with large Vpore.

While activated carbon based materials may possess a large activearea (>2000 m2/g), their Vpore ∼ 0.5 – 1 cm3/g would result in asmaller magnitude of accessible capacitive charge.21 In this paper, weaim to assess in greater detail the capacitive characteristics, in the TLEregime, of a model macroporous electrode comprised of a large Vpore

(of ∼40 cm3/g) comprised of vertically aligned CNTs: Figure 1c,the spacing of which may correspond to an equivalent poresize.

Experimental

Vertically Aligned (VA)-CNT growth:-— Substrate preparation.—Substrates onto which CNTs were grown were first prepared bydepositing a 5 nm thick Iron (Fe) catalyst layer on a silicon wafer(n- type silicon, 3′′ diameter) by electron beam evaporation using aTemescal BJD 1800 electron beam evaporator. The catalyst depositedwafer was then diced into 5 mm by 5 mm squares (yielding anequivalent projected area, Aproj = 0.25 cm2).

) unless CC License in place (see abstract).  ecsdl.org/site/terms_use address. Redistribution subject to ECS terms of use (see 137.110.36.10Downloaded on 2015-01-25 to IP

Page 2: 162 The Electrochemical Society High Rate Capacity through ...nanomaterials-dev.ucsd.edu/wp-content/uploads/2015/...High Rate Capacity through Redox Electrolytes Confined in Macroporous

Journal of The Electrochemical Society, 162 (1) A86-A91 (2015) A87

Figure 1. Electrochemical kinetics in macroporous electrodes constitutedfrom nanostructure arrays are manifested through: (a) overlap of the diffu-sion layer contours as the voltage scan rate (ν) decreases, i.e., ν3 > ν2 > ν1,resulting in planar diffusion to the top of the electrode, while (b) inside thepores, a decreased concentration gradient results in thin layer behavior at lowerscan rates, and more of the electrolyte volume can be accessed. (c) Verticallyaligned carbon nanotube (CNT) arrays were used as prototypes of macroporouselectrodes in this study.

Chemical vapor deposition (CVD).—CNTs were grown onto the sil-icon substrates in thermal CVD furnace with 1′′ diameter quartztube (33′′ in length). Prior to growth, the tube was pumped toless than 100 mTorr to evacuate oxygen. The system was thenbrought to atmospheric pressure (750 Torr) under an argon atmosphere

(350 SCCM flow rate). The furnace was heated at a rate 50◦C/minto 675◦C and then hydrogen and ethylene were introduced intothe quartz tube (100 and 200 SCCM respectively.) The tempera-ture and gas flow was held for a particular growth time, as it wasfound that that the CNT length was proportional to the time. Postgrowth, the furnace was cooled to room temperature under an argonatmosphere.Structural characterization.—The length (lCNT) and diameter (dCNT)of the CNTs were measured using a Phillips XL30 environmentalscanning electron microscope (ESEM) at multiple points on the sam-ple with average values determined to be 100 ± 10 μm and 40 ± 5nm, respectively: Figure 1c. The weight of the CNTs was determinedby weighing before and after growth with a Metler Toledo AG285balance and was determined to be in the range of 60 ± 15 μg. Thespacing between the CNTs was measured to be 350 ± 50 nm, and wastaken to yield an equivalent pore size (per convention, the electrodeswould then be classified as macroporous). The total volume of theelectrode, Vel was then estimated through the product of the projectedarea (=Aproj) of 0.25 cm2 and the CNT length of 100 μm to be 0.0025cm3 with a mass loading of 0.24 mg/cm2 (=60 μg/0.25 cm2).Electrochemical characterization.—All electrochemical characteriza-tion reported in this study was performed in a 3- electrode cell withVA-CNT arrays on Si substrate as the working electrode, SaturatedCalomel Electrode (SCE) as the reference electrode, and platinumwire as the counter electrode using a Gamry PCI4 potentiostat. Inthe experiments, the front side of the silicon wafer constituted elec-trode (with the CNT arrays) was immersed into the electrolyte, whilea copper wire was soldered (using indium) to the back of the sub-strate. The wire and the solder were kept outside the electrolyte. Itwas ensured through cyclic voltammetry (CV) conducted under sim-ilar conditions that the silicon substrate has a negligible contributionto the electrochemical performance. For electrolytes, we used aque-ous redox systems consisting of both ferricyanide (potassium hexa-cyanoferrate (III) (K3Fe(CN)6), with a diffusion coefficient,22 D = 6.8× 10−6 cm2/s, and ruthenium hexaamine trichloride (Ru(NH3)6Cl3)- with a D = 7.9 × 10−6 cm2/s, in 1M KCl supporting electrolyte.23

The former and the latter provide prototypes of an inner and outersphere systems, respectively for monitoring the sensitivity to sur-face conditions and electron transfer kinetics.24,25 All chemicals werepurchased from Sigma-Aldrich Inc. and were >97% pure. Prior toelectrochemical characterization, to ensure complete wetting by theelectrolyte, the CNT electrodes were rinsed in acetone, isopropyl al-cohol and water (in that order for 30 s in each solvent). CV wasemployed to record the current-voltage curves, the area of whichdivided by the scan rate (ν) would be directly proportional to thetransferred charge. The electrolytes used in every experiment were<2 hours old and in the case of CV, the potential was scanned be-tween +0.6 and −0.6 V vs. SCE at various scan rates. PresentedCVs at any particular ν consist of the average of at least 5 cycles.Chronoamperometry (CA) considering the redox reactions under avoltage step was used to probe the time scales in the variation ofthe electrical current. Galvanostatic discharge experiments were con-ducted at varying current densities and redox electrolyte concentra-tions to evaluate the rate capability of the investigated capacitivesystems.

Results and Discussion

The electrochemical performance of the CNT array electrodes ina given electrolyte, i.e., K3Fe(CN)6 or Ru(NH3)6Cl3, was first charac-terized using CV through the variation of the cathodic (or anodic) peakcurrents: Ip, and the ν: Figure 2a. From the average CNT spacing of∼350 nm and with 0.04 V/s < ν < 4 V/s, and � V ∼26 mV (=RT/F)the δ was estimated to be ∼3.7 μm even at the largest scan rate. Sucha value was significantly larger than the spacing between the tubesand hence the investigated macroporous system could exhibit TLEbehavior at all scan rates.

) unless CC License in place (see abstract).  ecsdl.org/site/terms_use address. Redistribution subject to ECS terms of use (see 137.110.36.10Downloaded on 2015-01-25 to IP

Page 3: 162 The Electrochemical Society High Rate Capacity through ...nanomaterials-dev.ucsd.edu/wp-content/uploads/2015/...High Rate Capacity through Redox Electrolytes Confined in Macroporous

A88 Journal of The Electrochemical Society, 162 (1) A86-A91 (2015)

Figure 2. (a) Cyclic Voltammograms (CV) depicting increasing peak currents with increasing ν (0.04 V/s, 0.1 V/s, 0.4 V/s and 1 V/s), in 10 mM K3Fe(CN)6/ 1 M KCl (aq.) electrolyte. (b) Calculated peak currents due to the combined contributions from the thin layer current and the planar diffusion current(in red), (c) effect of Ohmic distortion (in red) due to the solution resistance on the estimated thin layer peak currents observed in the CV. Parameters used:Vpore = 0.0024 cm3, Aproj = 0.25 cm2, Co = 10 mM, Ru = 5 �.

In the introductory section, it was stated that that the current ob-served in CV could have contributions from both (a) planar diffusionlimited current at the top of the electrode (Idiff) and (b) a TLE current(ITLE) in between the CNTs. For a reversible, one-electron redox re-action, over a time, t, the total peak current (after baseline correction)is20 the sum:

I = Idi f f + IT L E = F Aproj Co

(πDF

RT

)1/2

χ

(Fνt

RT

)· ν1/2

+(

n2 F2

RT

(1 + ξ)2VporeCo · ν [1]

In the above expression, χ(

FνtRT

)denotes the normalized current for

a reversible reaction under planar diffusion while ξ = e(E−Eo )F

RT whereE is the applied voltage and E0 is the formal potential of the redoxcouple. The deconvolution of a particular I-V curve, obtained fromCV, into the thin layer (black curve) and the diffusion constituents (redcurve) is depicted in Figure 2b. While it may be possible to estimateAproj and Vpore for a given electrode using [1], the voltammetric cur-rents at any given potential may not necessarily follow due to uncom-pensated resistance induced ohmic distortion.26 It was noted that thepositive feedback compensation used in commercial instrumentationyields a residual, small resistance of ∼0.5 �, which at larger currents(>10 mA) and at high ν , could still distort the I-V curves, through acorrespondingly large voltage drop and preclude clear electrochemi-cal interpretation. To separate ohmic and mass transport effects, thevoltammograms were numerically computed from an expression pre-viously derived by de Tacconi.27 Figure 2c compares the thin layercurrents (ITLE) with (red curve) and without (black curve) ohmic dis-tortion, assuming an Ohmic resistance, Ru of 2.5 �. It was seen thatthe distorted peak was lower amplitude but smaller but wider com-

pared to the undistorted peak. However, if the faradaic reaction isallowed to completion, i.e., if the voltammetric peak is well withinthe scanned potential window (Vi = −0.6 V to Vf = +0.6 V, in ourcase), the voltammetric charge Qtot will be unaffected by ohmic ef-fects. Consequently, analysis of Qtot could provide greater insightsinto the kinetics.28

Qtot = Qdi f f + QT L E =

V f∫Vi

(Idi f f + IT L E )dV

ν

= F Aproj Co

(πDF

RT

)1/2V f∫

Vi

χ

(Fνt

RT

)· ν−1/2dV + FVporeCo

[2]

Such a variation was clearly seen experimentally from a plot ofQtot with ν−1/2: Figure 3a. The intercept (=FVporeCo at a ν−1/2 = 0) was∼2.3 mC, using 10 mM K3Fe(CN)6. The volume of the electrolytetrapped within the pores (=Vpore) was estimated to be 0.0024 cm3,which is 96% of the total volume of the electrode of 0.0025 cm3 (=Vel)as previously estimated (see Experimental section) and indicates theextent of CNT constituted electrode porosity.

Further chronoamperometric experiments were used to note thatdiffusion to the electrodes dominates above ∼0.3 s, as evidenced bythe I ∼ t−1/2 variation – corresponding to a Cottrell mechanism,implying purely planar diffusion limited response through:29

ICot = F Aproj Co

(D

πt

)1/2

[3]

) unless CC License in place (see abstract).  ecsdl.org/site/terms_use address. Redistribution subject to ECS terms of use (see 137.110.36.10Downloaded on 2015-01-25 to IP

Page 4: 162 The Electrochemical Society High Rate Capacity through ...nanomaterials-dev.ucsd.edu/wp-content/uploads/2015/...High Rate Capacity through Redox Electrolytes Confined in Macroporous

Journal of The Electrochemical Society, 162 (1) A86-A91 (2015) A89

Figure 3. (a) Voltammetric charge (Qtot) as a function of ν−1/2 (ν: voltage scan rate) for the macroporous electrodes .The intercept at ν−1/2 = 0, constitutes thethin layer charge (QTLE), (b) A current-time plot from chronoamperometry conducted on the macroporous electrodes. It was noted that at short times, that therewas rapid depletion of the redox species in the confined thin layer followed by a purely diffusion limited Cottrell response at longer times. 10 mM K3Fe(CN)6) in1 M KCl (aq.) was used as the redox electrolyte.

The electrode area Aproj was estimated from the slope to be∼0.28 cm2, close to the estimated Aproj of 0.25 cm2. From Figure 3b,it was also observed that the slope of the current at very short timescales is clearly not diffusion limited owing to the rapid depletion ofthe electrolyte in the pores, i.e., corresponding to the thin layer regime.At large times, planar diffusion to the top of the clusters would be ob-tained reflecting the Cottrell ideal. The noted supra-linear dependency,i.e., I ∼ t−γ, with 0.5 < |γ | ≤ 1, at smaller time scales point to elec-trolyte depletion and at even shorter times Cottrell behavior shouldagain occur, now related to the regions between the nanotubes.30,15

A key aspect of the TLE regime is that the electrochemical responseseems to be identical to that of species adsorption on the electrode31

from the linearity of the observed I with the ν. Such species adsorptionmay be undesirable and form a rate-limiting step if the correspondingkinetics are slow. The charge stored in this case is capacitive due tothe absence of mass transport and the confinement of electroactivespecies on or close to the surface.32,33 This is particularly relevant forK3Fe(CN)6 which has been shown to adsorb onto carbon electrodeswith surface chemistry sensitive transfer kinetics.24 Consequently, for

further verification of the TLE regime, we conducted CV experimentsusing (Ru(NH3)6Cl3), an outer-sphere redox couple that is insensitiveto surface chemistry. From Figure 4a, it was inferred that the inter-cept charge was of similar magnitude (∼2.2 mC) to that observedfor K3Fe(CN)6. As a additional control experiment, to rule out ad-sorption, we transferred samples exposed to the K3Fe(CN)6 redoxspecies to the bare electrolyte (1M KCl) and compared the voltammo-grams in both the cases. The observed voltammetric response (inset toFigure 4b) was that of a double layer capacitance without any redoxpeaks that could be associated with adsorbed K3Fe(CN)6. The absenceof such memory effects effectively precludes adsorption.34

To evaluate the capacity of the examined electrode/electrolyte sys-tem, galvanostatic discharge experiments were carried out with differ-ent concentrations of the redox additives at a constant current density.A large plateau was observed in the voltage at close to the formalpotential (∼0.23 V vs. SCE) of the redox couple indicative of faradaiccharacteristics: Figure 5a. The recorded charge capacity, in mAh/cm3

or mAh/cm2, was obtained through a chronopotentiogram, by dividingthe charge by either Vel or Aproj, respectively. The discharge current

Figure 4. (a) Voltammetric charge (Qtot) as a function of ν−1/2, for the macroporous electrodes, with 10 mM Ru(NH3)6Cl3) in 1 M KCl (aq.) as the redoxelectrolyte, (b) The lack of a current peak when the macroporous electrode is transferred from a 10 mM constituted redox electrolyte (K3Fe(CN)6 in 1 M KCl(aq.) to the supporting electrolyte (1 M KCl, devoid of K3Fe(CN)6, i.e., 0 M), indicates the lack of adsorption/desorption (at 0.4 V/s). The inset indicates thebackground double-layer current in the absence of K3Fe(CN)6.

) unless CC License in place (see abstract).  ecsdl.org/site/terms_use address. Redistribution subject to ECS terms of use (see 137.110.36.10Downloaded on 2015-01-25 to IP

Page 5: 162 The Electrochemical Society High Rate Capacity through ...nanomaterials-dev.ucsd.edu/wp-content/uploads/2015/...High Rate Capacity through Redox Electrolytes Confined in Macroporous

A90 Journal of The Electrochemical Society, 162 (1) A86-A91 (2015)

Figure 5. (a) Galvanostatic discharge curves (at 200 mA/cm2) for various K3Fe(CN)6 concentrations (in 1 M KCl aq. electrolyte), inset: shows that the faradaiccapacity varies linearly with the concentration, (b) The capacity, normalized to the nominal thin layer capacity, as a function of discharge current density. It wasseen that the capacity contribution from diffusion processes is negligible beyond 50 mA/cm2 and mostly is due to thin layer attributes at 200 mA/cm2.

density was ∼200 mA/cm2 (corresponding to the ratio of the appliedcurrent of 50 mA and Aproj). For 1M K3Fe(CN)6, the maximum ob-served capacity, from Figure 5a, was ∼26 mAh/cm3 or 0.28 mAh/cm2. Such values were close to or relatively better (considering thedischarge current densities) than those previously observed and esti-mated for macroporous electrodes constituted from, e.g., (1) Nickelfoam based electrodes with CoO:35 0.8 mAh/cm2 (at 40 mA/cm2),(2) hybrid Polypyrrole/CoO nanowire array36 of 0.15 mAh/cm2 (at50 mA/cm2), (3) conductive hybrid metal oxides (Ni(OH)2/NiCo2O4)on carbon fiber paper: 0.25 mAh/cm2 (at 150 mA/cm2),37 (4)Co3O4/NiO core/shell nanowire arrays:38 ∼0.18 mAh/cm2 (at120 mA/cm2), (5) (Co, Ni) based compounds on hollow core nanorodarray:39 ∼0.14 mAh/cm2 (at 10 mA/cm2), etc. It was also observedfrom the inset of Figure 5a, that the observed capacity was linearlyproportional to the concentration of the redox species, with the impli-cation that a further improvement in capacity would be possible usingredox additives with a greater solubility, e.g., KI, with a solubilitylimit of ∼8 M compared to K3Fe(CN)6 with an aqueous solubility of∼1M.40

Moreover, in our experiments, the capacity normalized to the thinlayer capacity (through the calculated Vpore) seems to be constantbeyond ∼50 mA/cm2: Figure 5b, implying that much of the observedcapacity (at high current densities) is within the electrolyte confinedin Vpore. At lower current density, diffusional contributions would alsoadd to the capacity. It was noted that the TLE based capacity of thesingle macroporous electrode constituted device could be cycled atvery high rates for up to 5000 cycles (at 200 mA/cm2): Figure 6.The capacity dropped by 10% within the first ten cycles after whichit was constant while exhibiting excellent Coulombic efficiency, i.e.,Qdischarge/Qcharge of ∼97%.

Conclusions

We have shown that under conditions where the nanostructurespacing/pores in a macroporous electrode is smaller than the equiva-lent diffusion layer thickness that capacity enhancement correspond-ing to thin layer electrochemistry could be induced. Consequently,faradaic reactions from the charge in the enclosed electrolyte couldcontribute. In addition to the enhanced kinetics that could be obtaineddue to the relative absence of diffusion, the measured capacity wouldbe directly proportional to the electrolyte volume confined in the poresand its bulk concentration. Galvanostatic discharge experiments, car-ried out with different concentrations of the redox additives, yielded avoltage plateau further indicative of the faradaic characteristics of thecapacity. For 1M K3Fe(CN)6, the maximum observed capacity was

Figure 6. Capacity cycling studies (at 200 mA/cm2 discharge current den-sity) indicates relative stability (after an initial drop of ∼10% over the first10 cycles) of the capacity for over 5000 cycles. The Coulombic efficiency(=Qdischarge/Qcharge) over all the cycles was ∼97%. The inset shows therepresentative charge/discharge cycling behavior over three cycles. 100 mMK3Fe(CN)6) in 1 M KCl (aq.) was used for the redox electrolyte.

found to be ∼26 mAh/cm3 (normalized to the pore volume) or 0.28mAh/cm2 (normalized to the projected area of the electrode). For anaqueous system, with a density of ∼1 g/cm3 the capacity could beof the order of 26 mAh/g. It was also noted that the capacity couldbe cycled at very high rates for up to 5000 cycles (at 200 mA/cm2).Further work in increasing the capacity could focus on electrolyteswith a greater concentration of the redox species as well as elec-trodes with an increased pore volume. Additionally, self-dischargephenomena inherent to the considered faradaic processes41 must beconsidered. Recent experiments on the use of electrode specific elec-trolytes confined through the use of ion-selective membranes,12 orthrough enhanced polarization42 indicate efforts in this direction.

Acknowledgments

We acknowledge the financial support (through CMMI 1246800)of the National Science Foundation (NSF). Discussions with H. Ya-mada are deeply appreciated.

) unless CC License in place (see abstract).  ecsdl.org/site/terms_use address. Redistribution subject to ECS terms of use (see 137.110.36.10Downloaded on 2015-01-25 to IP

Page 6: 162 The Electrochemical Society High Rate Capacity through ...nanomaterials-dev.ucsd.edu/wp-content/uploads/2015/...High Rate Capacity through Redox Electrolytes Confined in Macroporous

Journal of The Electrochemical Society, 162 (1) A86-A91 (2015) A91

List of Symbols

Symbol Unit Meaning

Anet cm2 total area of the nanostructuredelectrode

Aproj cm2 projected area of the electrodesCo mol/cm3 bulk concentration of redox speciesD cm2/s diffusion coefficientd nm distance between nanotubes/Pore

length scale/δ cm diffusion layer thicknessE V applied potential vs. SCE (standard

calomel electrode, reference)E0 V formal potential vs. SCEF 96500 C/mol Faraday constant

Idiff A electrical current due to diffusionalprocesses

ITLE A current due to thin layerelectrochemistry (TLE)

n number of electrons in redox reactionξ normalized thin layer voltammetric

currentQ C voltammetric chargeR 8.31 J/mol.K Gas constantRu � uncompensated resistanceT K temperaturet s experimental time

Vel cm3 total volume of the electrolyteVpore cm3 total pore volume

ν V/s voltage scan rateχ normalized diffusion limited

voltammetric current

References

1. B. E. Conway, in Electrochemical Supercapacitors- Scientific Fundamentals andTechnological Applications, Kluwer- Academic (1999).

2. J. P. Zheng, P. J. Cygan, and T. R. Jow, J. Electrochem. Soc., 142, 2699 (1995).3. E. Raymundo-Pinnero, V. Khomenko, E. Frackowiak, and F. BeFguin, J. Electrochem.

Soc., 152, A229 (2005).4. J. Yan, et al., J. Power Sources, 195, 3041 (2010).5. J. Wang, J. Polleux, J. Lim, and B. Dunn, J. Phys. Chem. C, 2, 14925 (2007).

6. V. Augustyn, et al., Nat. Mater., 12, 518 (2013).7. S. T. Senthilkumar, R. K. Selvan, N. Ponpandian, J. S. Melo, and Y. S. Lee, J. Mater.

Chem. A, 1, 7913 (2013).8. G. Lota, K. Fic, and E. Frackowiak, Electrochem. commun., 13, 38 (2011).9. S. Roldan, et al., Electrochim. Acta, 56, 3401 (2011).

10. V. Augustyn, P. Simon, and B. Dunn, Energy Environ. Sci., 7, 1597 (2014).11. M. Meller, J. Menzel, K. Fic, D. Gastol, and E. Frackowiak, Solid State Ionics, 265,

61 (2014).12. E. Frackowiak, K. Fic, M. Meller, and G. Lota, ChemSusChem, 5, 1181 (2012).13. S. Roldan, C. Blanco, M. Granda, R. Menendez, and R. Santamarıa, Angew. Chem.

Int. Ed. Engl., 50, 1699 (2011).14. S. Roldan, et al., J. Phys. Chem. C, 115, 17606 (2011).15. E. O. Barnes, X. Chen, P. Li, and R. G. Compton, J. Electroanal. Chem., 720–721,

92 (2014).16. G. Keeley and M. Lyons, Int J Electrochem Sci, 4, 794 (2009).17. M. C. Henstridge, E. J. F. Dickinson, M. Aslanoglu, C. Batchelor-McAuley, and

R. G. Compton, Sensors Actuators B Chem., 145, 417 (2010).18. K. R. Ward, L. Xiong, N. S. Lawrence, R. S. Hartshorne, and R. G. Compton, J.

Electroanal. Chem., 702, 15 (2013).19. A. T. Hubbard and F. C. Anson, in Electroanalaytical Chemistry: A Series of Ad-

vances, p. 129 (1970).20. A. J. Bard and L. R. Faulkner, Electrochemical Methods: fundamentals and Applica-

tions, 2nd ed., John Wiley, New York, (2001).21. J. Gamby, P. Taberna, P. Simon, J. Fauvarque, and M. Chesneau, J. Power Sources,

101, 109 (2001).22. S. J. Konopka and B. McDuffie, Anal. Chem., 42, 1741 (1970).23. C. Beriet and D. Pletcher, J. Electroanal. Chem., 375, 213 (1994).24. P. Chen, M. A. Fryling, and R. L. McCreery, Anal. Chem., 67, 3115 (1995).25. K. K. Cline, M. T. McDermott, and R. L. McCreery, J. Phys. Chem., 98, 5314 (1994).26. J. Weidner and P. Fedkiw, J. Electrochem. Soc., 138, 2514 (1991).27. N. R. De Tacconi, A. J. Calandra, and A. J. Arvıa, Electrochim. Acta, 18, 571

(1973).28. S.-B. Yoon, H.-K. Song, K. C. Roh, and K.-B. Kim, J. Electrochem. Soc., 161, A137

(2013).29. R. G. Compton and C. E. Banks, Understanding Voltammetry, Imperial College Press,

London, UK (2011).30. D. Sanchez-Molas, J. P. Esquivel, N. Sabate, F. X. Munoz, and F. J. del Campo, J.

Phys. Chem. C, 116, 18831 (2012).31. J. Eskusson, A. Janes, A. Kikas, L. Matisen, and E. Lust, J. Power Sources, 196, 4109

(2011).32. E. Laviron, J. Electroanal. Chem, 52, 355 (1974).33. S. Srinivasan and E. Gileadi, Electrochim. Acta, 11, 321 (1966).34. R. T. Kachoosangi, G. G. Wildgoose, and R. G. Compton, Analyst, 133, 888

(2008).35. C. Guan, et al., Adv. Mater., 24, 4186 (2012).36. C. Zhou, Y. Zhang, Y. Li, and J. Liu, Nano Lett., 13, 2078 (2013).37. L. Huang, et al., Nano Lett., 13, 3135 (2013).38. X. Xia, et al., ACS Nano, 6, 5531 (2012).39. L. Wan, J. Xiao, F. Xiao, and S. Wang, ACS Appl. Mater. Interfaces, 6, 7735 (2014).40. G. Lota and E. Frackowiak, Electrochem. commun., 11, 87 (2009).41. B. E. Conway, W. G. Pell, and T.-C. Liu, J. Power Sources, 65, 53 (1997).42. B. Wang, et al., J. Electrochem. Soc., 161, A1090 (2014).

) unless CC License in place (see abstract).  ecsdl.org/site/terms_use address. Redistribution subject to ECS terms of use (see 137.110.36.10Downloaded on 2015-01-25 to IP