Nano Research - Tailoring uniform γ-MnO2 nanosheets on ...Nano Res 1 Tailoring uniform γ-MnO 2...

4
Solero Professional 2013 GEZELLIG WAT DRINKEN OF UITGEBREID DINEREN? NEDERLAND EET STEEDS VAKER BUITEN DE DEUR. LETTERLIJK WEL TE VERSTAAN. BUITEN IS NAMELIJK VEEL MEER TE ZIEN EN TE BELEVEN. DE IDEALE PLEK OM TE ONTSPANNEN NA EEN DRUKKE DAG OF SAMEN MET VRIENDEN HET WEEKEND TE VIEREN. BIJ EEN PRIEMENDE ZON OF MAARTSE BUI HOEVEN UW GASTEN NIET MEER NAAR BINNEN. DE SOLERO PROFESSIONAL PARASOLS ZORGEN VOOR EEN AANGENAME BESCHUTTING. EN STAAN OOK HUN MANNETJE BIJ ZWAARDER WEER. EEN BLIKVANGER EN HERKENNINGSPUNT, WAARONDER UW GASTEN GRAAG ZULLEN VERTOEVEN.

Transcript of Nano Research - Tailoring uniform γ-MnO2 nanosheets on ...Nano Res 1 Tailoring uniform γ-MnO 2...

  • Nano Res

    1

    Tailoring uniform γ-MnO2 nanosheets on highly

    conductive three-dimensional current collectors for

    high-performance supercapacitor electrodes

    Sangbaek Park1†, Hyun-Woo Shim2†, Chan Woo Lee1, Hee Jo Song1, Ik Jae Park1, Jae-Chan Kim2, Kug

    Sun Hong1, and Dong-Wan Kim2()

    Nano Res., Just Accepted Manuscript • DOI: 10.1007/s12274-014-0581-1

    http://www.thenanoresearch.com on September 11, 2014

    © Tsinghua University Press 2014

    Just Accepted

    This is a “Just Accepted” manuscript, which has been examined by the peer-review process and has been

    accepted for publication. A “Just Accepted” manuscript is published online shortly after its acceptance,

    which is prior to technical editing and formatting and author proofing. Tsinghua University Press (TUP)

    provides “Just Accepted” as an optional and free service which allows authors to make their results available

    to the research community as soon as possible after acceptance. After a manuscript has been technically

    edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP

    article. Please note that technical editing may introduce minor changes to the manuscript text and/or

    graphics which may affect the content, and all legal disclaimers that apply to the journal pertain. In no event

    shall TUP be held responsible for errors or consequences arising from the use of any information contained

    in these “Just Accepted” manuscripts. To cite this manuscript please use its Digital Object Identifier (DOI®),

    which is identical for all formats of publication.

    Nano Research DOI 10.1007/s12274-014-0581-1

  • 1

    TABLE OF CONTENTS (TOC)

    Tailoring uniform γ-MnO2 nanosheets on highly

    conductive three-dimensional current collectors for

    high-performance supercapacitor electrodes

    Sangbaek Park1, Hyun-Woo Shim2, Chan Woo Lee1, Hee

    Jo Song1, Ik Jae Park1, Jae-Chan Kim2, Kug Sun Hong1,

    and Dong-Wan Kim2*

    1 Department of Materials Science and Engineering, Seoul

    National University, Seoul 151-744, Korea

    2 School of Civil, Environmental and Architectural

    Engineering, Korea University, Seoul 136-713, Korea

    Ultrathin 2-D MnO2 nanosheets with highly conductive 3-D

    current collectors was fabricated and assembled to obtain

    excellent electrochemical performance.

    Dong-Wan Kim, http://dwkim.ajou.ac.kr

  • 2

    Tailoring uniform γ-MnO2 nanosheets on highly conductive three-dimensional current collectors for high-performance supercapacitor electrodes

    Sangbaek Park1†, Hyun-Woo Shim2†, Chan Woo Lee1, Hee Jo Song1, Ik Jae Park1, Jae-Chan Kim2, Kug Sun Hong1, and Dong-Wan Kim2()

    1 Department of Materials Science and Engineering, Seoul National University, Seoul 151-744, Korea 2 School of Civil, Environmental and Architectural Engineering, Korea University, Seoul 136-713, Korea

    † These authors contributed equally

    Received: day month year / Revised: day month year / Accepted: day month year (automatically inserted by the publisher)

    © Tsinghua University Press and Springer-Verlag Berlin Heidelberg 2011

    ABSTRACT Recent efforts have focused on the fabrication and application of 3-D nanoarchitectured electrodes, which can

    exhibit excellent electrochemical performance. Herein, a novel strategy towards the design and synthesis of

    size- and thickness-tunable two-dimensional (2-D) MnO2 nanosheets on highly conductive 1-D backbone arrays

    were developed via a facile, one-step enhanced chemical bath deposition (ECBD) method at a low temperature

    (~ 50 °C). Inclusion of an oxidizing agent, BrO3-, in solution was crucial in controlling the heterogeneous

    nucleation and growth of the nanosheets, and in inducing the tailored and uniformly arranged nanosheet

    arrays. We fabricated supercapacitor devices based on 3-D MnO2 nanosheets with conductive Sb-doped SnO2

    nanobelts as the backbone. They achieved a specific capacitance of 162 F g-1 at an extremely high current

    density of 20 A g-1, and good cycling stability that shows the capacitance retention of ~92 % of its initial value,

    along with a coulombic efficiency of almost 100% after 5,000 cycles in an aqueous solution of 1 M Na2SO4. The

    results were attributed to the unique hierarchical structures, which provided a short diffusion path of

    electrolyte ions by 2-D sheets and direct electrical connections to the current collector by 1-D arrays as well as

    aggregation prevention by the well-aligned 3-D structure.

    KEYWORDS Nanosheets, Manganese oxide, Chemical bath deposition, SnO2 nanobelts, Supercapacitor

    1. Introduction

    Manganese dioxide (MnO2) is one of the most

    attractive inorganic materials and is a stable

    compound with excellent chemical and physical

    properties [1-4]. Due to its low cost, environmental

    friendliness, non-toxicity, as well as its structural

    flexibility and rich polymorphism (α-, β-, γ-, δ-, λ-,

    and ε-type), wide applications have been reported

    including as a catalyst [5], ion-sieve [6], ion exchange

    [7], biosensor [8], lithium ion battery [9], and

    Nano Res DOI (automatically inserted by the publisher) Review Article/Research Article Please choose one

    ————————————

    Address correspondence to D.-W. Kim, [email protected]

  • 3

    supercapacitor [10, 11]. In particular, MnO2 is a

    promising material as a replacement for RuO2 in

    pseudo-capacitors owing to its high theoretical

    capacitance of 1233 F g-1 for a one-electron transfer

    and complete reduction of MnIV to MnIII [12].

    Moreover, it can be utilized in mild neutral aqueous

    electrolytes, which ensures environmental

    compatibility, safety, non-flammability, and

    convenient assembly in air [13]. Among various

    phases, γ-MnO2 is most widely used in energy

    storage devices because its electrical activity

    decreases more slowly than other forms during the

    electrochemical process [14]. However, only 100 ~

    300 F g-1 have been reported for MnO2 powders [15],

    owing to their poor electrical conductivity (10-5 to 10-6

    S cm-1)[16] and low accessible surface areas [17].

    Thus, a MnO2 electrode that provides a short

    diffusion path for electrolyte ions and a high

    electrical conductivity of the active materials is

    needed.

    Several approaches have utilized in an attempt to

    meet the essential criteria in the design of

    high-performance MnO2-based electrodes for

    supercapacitors. Given that pseudo-capacitance is

    primarily generated by surface Faradaic redox

    reactions, it has been quite a challenge to develop

    MnO2 nanostructures with an ultrathin morphology

    for advanced supercapacitors [18-21]. Additionally,

    hybrid composite structures with highly conductive

    materials have been explored in order to improve the

    rate capabilities of MnO2 electrodes for high power

    performance [22-31]. Carbon-based materials such as

    carbon nanotubes, graphene sheets, and activated

    carbon have been investigated as hybrid composites

    with MnO2 [22-27]. Nevertheless, the aggregation of

    carbon caused by van der Waals interactions

    prevents its use in practical applications [24].

    Recently, three-dimensional (3-D) MnO2

    nanostructures with conductive backbones, e.g.,

    SnO2 or Zn2SnO4, showed improved electrochemical

    capacitance under carbon and binder-free conditions

    [28, 29]. Thus, the design and synthesis of 3-D MnO2

    nanostructures composed of 1-D metallic backbones

    and 2-D ultrathin MnO2 nanosheets are attractive for

    the development of superior supercapacitors.

    A key challenge in the fabrication of 3-D MnO2

    nanostructures is a facile and controllable synthetic

    route for well-defined MnO2 nanosheets on the 1-D

    backbone. Some efforts have been devoted to

    obtaining high-quality MnO2 nanosheets by using

    various methods, including delamination (exfoliation)

    [32, 33], reduction [34, 35], and templating [36, 37].

    For example, Liu et al. successfully deposited MnO2

    nanosheets on various metal oxide nanowires using

    sacrificial carbon templates coated on nanowires, in

    which the carbon layers functioned as reducing

    agents [38]. In spite of its wide applicability [35], this

    process is limited by the inevitable carbon layer

    deposition step. Although the structural design of 2D

    nanosheets on 1D conductive array with high

    performance in supercapacitor also has been

    extensively explored in other materials [38-40], most

    of them are synthesized in high temperature or

    pressure, which make it hard to modulate the

    morphology. A decade ago, Unuma et al. reported

    that MnO2 thin layers were heterogeneously grown

    on a substrate by chemical bath deposition (CBD)

    with NaBrO3 as the oxidizing agent [41]. Because

    CBD is a low-temperature and surfactant-free

    technique, it is very useful in tailoring 3-D

    nanoarchitectures [42]. Inspired by these pioneering

    studies, we designed a broad and general approach

    for forming MnO2 nanosheets on 1-D backbones by

    CBD.

    Herein, we present an alternative route by

    organizing a 3-D MnO2 nanoarchitectured electrode

    with 2-D nanosheets and 1-D conductive backbones

    for supercapacitor devices. The synthetic route

    consists of two stages: (i) construction of 1-D

    backbones by conventional methods such as

    vapor-liquid-solid, anodizing, or AAO templating,

    and (ii) decoration of 2-D MnO2 nanosheets on 1-D

    arrays by CBD (Scheme 1). Moreover, we adopted

    enhanced CBD (ECBD) composed of an oxidation

    reaction, which provided a facile, mass-producible,

    and controllable way for forming MnO2 nanosheets

    on arbitrary 1-D backbones. As a proof of concept,

    Sb-doped SnO2 (ATO) nanobelt arrays having a

    metallic conductivity [43] were selected as the

    conductive 1-D backbone. In neutral aqueous

    solutions, this 3-D ATO@MnO2 nanostructure

    showed superior capacitance and stability at high

    currents without any binders or conductors owing to

    the highly conductive pathway generated by the 1-D

    current collector, fast ion transport by the 2-D sheets,

  • 4

    Scheme 1 Illustration of 3-D hierarchical γ-MnO2 nanosheets.

    Fabrication of conductive backbone arrays directly grown on the

    current collector, and controllable γ-MnO2 nanosheets on each

    1-D backbone.

    and aggregation prevention by the well-defined 3-D

    structure.

    2. Results and discussion

    2.1. Morphology and structure

    The 3-D nanostructure decorated with 2-D MnO2

    nanosheets was verified by morphological analysis.

    As shown in Figure 1, ultrathin (< 10 nm) MnO2

    nanosheets with typical diameters of 500 nm were

    successfully formed on the ATO nanobelt surfaces by

    ECBD. Specifically, the reaction was carried out at 50 oC for 3.5 h in a solution containing 0.05 M Mn2+ and

    0.3 M BrO3-. It can be clearly seen from Figure 1b that

    all nanobelts were uniformly covered with MnO2

    nanosheets. Interestingly, aligned nanobelt arrays

    (Figure S1) were preserved (Figure 1a and b),

    suggesting that little damage was caused by the mild

    ECBD.

    Typical XRD patterns and Raman spectra of ATO

    and ATO@MnO2 are presented in Figure 2. With the

    exception of the SnO2- and Ti-related peaks, the most

    intensive peak of the MnO2 nanosheets at around 2θ

    = 36.9o corresponded to the (131) peak of the nsutite

    γ-MnO2 (JCPDS 14-0644, a = 6.36 Å, b = 10.15 Å, c =

    4.09 Å), consistent with previous studies for MnO2

    thin films fabricated in solution containing Mn2+ and

    Figure 1 Typical SEM images of (a) ATO nanobelts and (b)

    ATO@MnO2 3-D nanostructures grown on Ti substrates.

    BrO3- [41]. Relatively weak and broad MnO2 peaks

    represent the amorphous nature of nanosheets. To

    further identify the local Mn environment in the

    γ-MnO2 nanosheets, Raman analysis was carried out

    (Figure 2b). Three main bands were observed in the

    ATO nanobelts; the band near 627 cm-1 was the A1g

    mode of SnO2 and bands near 452 and 245 cm-1 were

    related to doping of Sb in SnO2 lattice [42]. The

    Raman spectra of ATO@MnO2 revealed two main

    vibration bands (ν1 = 644 and ν2 = 575 cm-1) in the

    wavenumber range of 500 – 700 cm-1, characteristic

    features of γ-MnO2 [14, 44]. The crystal structure of

    γ-MnO2 can be described as an arrangement of

    octahedral MnO6 units with a random intergrowth of

    pyrolusite (β-MnO2) layers within a ramsdellite

    (R-MnO2) matrix. Thus, the fraction of single chain

    slabs (pyrolusite) in the double chain (ramsdellite)

    framework affects vibrational interactions in the

    lattice of γ-MnO2, inducing the blue shift in the

    vibration peak (ν1) of γ-MnO2 from that of R-MnO2

    (ν1 = 630 cm-1) to that of β-MnO2 (ν1 = 665 cm-1) [44].

    Therefore, the amount of pyrolusite intergrowth,

    known as DeWolff defects, can be determined by

    linear interpolation; γ-MnO2 nanosheets have about

    32 % pyrolusite in the ramsdellite. All these features

    confirm that the MnO2 nanosheets were synthesized

    as nsutite γ-MnO2.

    The morphology and crystal structure of the

    ATO@MnO2 nanostructures were further

    investigated by TEM (Figure 3). ATO nanobelts were

    regularly enclosed by MnO2 nanosheets, which were

    perpendicular to the growth axis of the nanobelt and

    formed circular bands surrounding the nanobelt. To

    characterize the nanobelt and sheet separately,

    respective selected area electron diffraction (SAED)

    patterns were obtained. The nanobelt was assigned

    to the SnO2 rutile phase according to the measured

    SAED pattern with [102̅] zone axis (Figure 3b). From

  • 5

    Figure 2 (a) XRD graphs of ATO nanobelts and ATO@MnO2

    3-D nanostructures. The line pattern show reference JCPDS

    #41-1445 (SnO2) and #14-0644 (MnO2). (b) Raman spectra of

    ATO and ATO@MnO2.

    the SAED ring patterns (Figure 3c), it is clear that the

    nanosheet is γ-MnO2 with an orthorhombic lattice of

    d131 = 0.242 nm. Moreover, EDS mapping (Figure 3d)

    revealed that oxygen existed in the core and sheet

    components, whereas tin and manganese were only

    detected in the core and sheet, respectively.

    Therefore, 2-D γ-MnO2 nanosheet arrays could be

    formed normal to the 1-D backbone axis by one-step

    ECBD.

    2.2. The function of the oxidizing agent

    The key step in the presented synthesis was the use

    of an oxidizing agent (BrO3-), which played a

    significant role in the formation of the MnO2

    nanosheets on the 1-D backbone arrays. In the

    presence of BrO3-, the oxidation of Mn2+ to MnO2

    occurred via two reaction steps:

    6Mn2+ + BrO3– + 6H+ → 6Mn3+ + 3H2O + Br– (1)

    2Mn3+ +2H2O → Mn2+ + MnO2 + 4H+ (2)

    Thus, the overall reaction was as follows:

    6Mn2+ + BrO3– + 3H2O → 3MnO2 + Br– + 6H+ (3)

    From the Pourbaix diagram for manganese [41],

    reaction (3) has a negative Gibbs free energy (∆G) in

    all pH ranges, indicating that it is

    thermodynamically feasible. Moreover, given that

    there was no Mn3+ in the starting solution, the rates

    of reaction (1) & (2) were predominately determined

    by the concentration of Mn2+ and BrO3-. According to

    thermodynamics, the Gibbs free energy of

    heterogeneous nucleation is smaller than that of

    homogeneous nucleation, implying that

    Figure 3 (a) A TEM image of an ATO@MnO2 3-D

    nanostructure, (b) SAED pattern of ATO nanobelt, (c) SAED

    pattern of γ-MnO2 nanosheets, and (d) EDS mapping of a

    ATO@MnO2 3-D nanostructure.

    heterogeneous nucleation can be predominant when

    the rate of (1) is slow enough. Thus, heterogeneous

    nucleation and growth of MnO2 nanosheets can be

    controlled by the concentration of the oxidizing

    agent. Figure 4 shows the various ATO@MnO2

    nanostructures synthesized with different

    concentrations of BrO3- and Mn2+. Nanosheets were

    irregularly attached on the nanobelts at low

    concentrations of BrO3- (0.1 M) (Figure 4a, d, and g).

    As the concentration of BrO3- increased (0.2 M), the

    deposition amount and uniformity of the nanosheets

    increased (Figure 4b, e, and h). At high

    concentrations of BrO3- (0.3 M), nanosheets randomly

    wrapped the nanobelts (Figure 4c and f) and some

    precipitates were observed (Figure 4i). These results

    were also confirmed by XRD and EDS analysis

    (Figure S2&Table S1). The loading amount was

    obviously different with different concentration of

    BrO3-. Based on nucleation and growth theory, the

    dependency on the concentrations of the oxidizer can

    be interpreted as follows: First, with low

    concentrations of BrO3-, only a few monomers were

    formed in solution due to the very slow reaction rate,

    which induced the slow heterogeneous nucleation

    rate. Accordingly, the formed monomers were

    mostly consumed by the growth of already attached

    MnO2, rather than by the formation of other

    heterogeneous nucleation on the nanobelts, inducing

    abnormal growth and irregular deposition of MnO2

    nanosheets. Second, with moderate concentrations of

    BrO3-, the solution was filled with enough monomers

    for heterogeneous nucleation to dominate, inducing

    the thorough covering of nanosheets on the

    nanobelts. Moreover, the nucleated MnO2 on the

  • 6

    Figure 4 SEM images of γ-MnO2 nanosheets synthesized in the

    presence of MnCl2 and KBrO3 at different concentrations: (a-c)

    MnCl2 0.02 M, (d-f) MnCl2 0.05 M, (g-i) MnCl2 0.1 M. (a, d, g)

    KBrO3 0.1 M, (b, e, h) KBrO3 0.2 M, (c, f, i) KBrO3 0.3 M. The

    reaction temperature is fixed at 50 oC. Scale bar of images and

    insets is 5 μm and 1 μm, respectively. M and K refers to MnCl2

    and KBrO3, respectively.

    nanobelt was uniformly grown by sufficient

    monomers. Lastly, with high concentrations of BrO3-,

    fast reaction rates produced large amounts of

    monomers, higher than the supersaturation level,

    and homogeneous nucleation as well as

    heterogeneous nucleation become pronounced.

    Therefore, a large amount of MnO2 attached and

    some precipitates settled. These behaviors were also

    confirmed by the transparency of the solutions

    (Figure S3): with low concentrations of BrO3-, the

    solution was clear, meaning that only heterogeneous

    nucleation occurred, and with high concentrations of

    BrO3-, the solution became cloudy after the reaction,

    revealing that homogeneous nucleation also

    occurred. Therefore, fast heterogeneous nucleation

    and moderate growth rates by adjusting the oxidizer

    concentration are essential in growing uniform and

    regular nanosheet arrays on 1-D backbones.

    2.3. Controllable synthesis of γ-MnO2 nanosheets

    To control the size of the MnO2 nanosheets, the

    formation as a function of reaction time was

    observed by ex situ SEM (Figure 5). The

    concentrations of Mn2+ and BrO3- were fixed at 0.05

    M and 0.2 M, respectively. As the reaction time

    increased, the MnO2 nanosheets grew on the

    nanobelts with diameters ranging from 400 nm to 1.5

    Figure 5 SEM images of γ-MnO2 nanosheets prepared in the

    presence of 0.05 M MnCl2 and 0.2 M KBrO3 with different

    reaction times: (a) 2 h 15 m, (b) 2 h 30 m, (c) 2 h 45 m, (d) 3 h,

    (e) 3 h 30 m, and (f) 4 h. The reaction temperature is fixed at 50 oC. Scale bar of images and insets is 5 μm and 1 μm,

    respectively.

    μm. This was consistent with the XRD graphs, in

    which the (131) peak intensity of γ-MnO2 increased

    as the reaction time increased (Figure S4). This

    fine-tuning technique was attributed to the slow

    growth rate by the low temperature in ECBD, which

    is beneficial for the fabrication of 3-D hierarchical

    heterostructures. Interestingly, when the ECBD

    reaction was long enough (63 h), skein-like

    microspheres were obtained with an average

    diameter of 2.5 μm (Figure S5). Because irregular

    powders precipitated when the ATO nanobelt was

    not inserted in the solution, the formation of the

    mesoporous microspheres was due to the

    aggregation of the nanosheets on the nanobelt. This

    result also supports the fact that nanosheets were

    formed by heterogeneous nucleation and growth on

    the surface of the nanobelts. In addition, the

    thickness of the nanosheets could also be controlled

    by temperature variation in ECBD. Figure S6 shows

    the SEM images of the nanosheets formed at

    different reaction temperatures and concentrations of

    MnCl2. At 50 oC, all nanosheets synthesized with

    various concentrations of MnCl2 were thin layers

    with a thickness less than 10 nm. However, as the

    temperature increased, the thickness of the

    nanosheets increased up to 100 nm due to the fast

    growth rate. While only 2-D growth dominantly

    occurs at low temperatures, 3-D growth could occur

    at high temperatures, inducing an increase in

    thickness. These results demonstrated that size- and

    dimension-tunable 2-D MnO2 nanosheets can be

    formed on 1-D backbones by ECBD.

  • 7

    2.4. Supercapacitive performance of 3-D

    hierarchical ATO@MnO2 nanosheets

    To highlight the advantages of the 3-D ATO@MnO2

    hetero-nanostructures as electrode materials for

    supercapacitors, their electrochemical performance

    was investigated using cyclic voltammetry (CV) and

    galvanostatic charge-discharge (CD) test in a

    three-electrode system in 1 M Na2SO4. Figure 6a

    shows the typical CV curves of the ATO@MnO2

    electrode fabricated for 3.5 h (Figure 1) in the

    potential range of 0 to 1.0 V (vs. Ag/AgCl) at scan

    rates from 10 up to 500 mV s-1. This CV is an effective

    experimental technique to evaluate the characteristic

    capacitive behaviors of electrode materials in

    supercapacitors [45]. The CV responses presented

    quasi-rectangular shapes, indicating fast

    charging-discharging processes characteristic such as

    ideal electrical double-layer capacitance (EDLC)

    behavior. Besides, it should be also noted that a

    broad oxidation/reduction couple was observed at

    around 0.6 and 0.3 V. This means that the

    ATO@MnO2 electrode exhibited a pseudo-capacitive

    behavior and a reversible Faradaic transition of Mn

    oxide; that is, the ATO@MnO2 electrode undergoes

    an electrochemical charge transfer reaction in basic

    electrolytes, making it a potential candidate for

    pseudo-capacitor applications. According to the

    literature, such charge-discharge processes or the

    capacitive reaction of MnO2-based electrodes near

    the surface during CV or CD measurements in

    aqueous solutions can be expressed by the following

    faradaic reactions [46-48]:

    (Mn4+O2)surface + xC+ + xe- ↔ C+x(Mn1-x4+Mnx3+O2)surface (4)

    (Mn4+O2)bulk + xC+ + xe- ↔C+x(Mn1-x4+Mnx3+O2)bulk (5)

    (C+ = K+, Li+, Na+, or H+)

    This first mechanism (reaction (4)) illustrates the

    adsorption/desorption of electrolyte cations (C+) on

    the MnO2 surface and reaction (5) involves the

    insertion of cations dissolved in the electrolyte. The

    literature reveals that a pair of redox peaks,

    MnO2/MnOOC, are generally involved in this system

    and demonstrate that the overall redox reaction

    given by reactions (4) and (5) are involved in the

    charge storage mechanism. In the present study, the

    CV characteristics of the ATO@MnO2 electrode are

    mainly consistent with those previously reported for

    MnO2 in Na2SO4 electrolytes [47, 49-52], even though

    such a CV shape may vary from sample to sample

    and may strongly depend on the morphology and

    surface properties of the electrodes. Furthermore, the

    linear relationship of the plot of the anodic peak

    current density versus the scan rate up to 500 mV s-1

    (inset in Figure 6a and Figure S7) also indicated the

    occurrence of surface redox reactions. Therefore,

    these CV pattern characteristics substantiate

    pseudo-capacitive behaviors of the ATO@MnO2

    electrode; that is, the charge storage of the

    ATO@MnO2 electrode is characteristic of the

    pseudo-capacitive process originating from the

    reversible redox reaction.

    More importantly, the CV curves showed no

    obvious distortion from 10 to 200 mV s-1 and

    exhibited a nearly rectangular shape, indicating that

    the redox reaction is fast enough for the ATO@MnO2

    electrode at the measured scan rate. Even at the scan

    rate of 500 mV s-1, the CV retained a similar shape

    without the significant change, indicating not only

    excellent ion diffusion in the electrode but also a

    better high-rate response of the ATO@MnO2 material.

    In other words, this result suggests that the 3-D

    hierarchical formations as well as the presence of the

    self-supported ATO nanowires (NWs) on the current

    collector and thin 2-D MnO2 nanosheets combined

    loosely with ATO NWs can actually facilitate

    electronic and ion transport and thus improve the

    kinetics of the capacitive reaction. The mass-specific

    capacitance (SC, F g-1) of ATO@MnO2 electrode was

    estimated from the CV curves, and was denoted as

    SCCV (see Supporting Information; Calculation of

    specific capacitance). As a function of scan rate, the

    calculated SCCV values of the ATO@MnO2 electrode

    were not only the maximum specific capacitance of

    278 F g-1 at a scan rate of 10 mV s-1, but also 183 F g-1

    at a scan rate as high as 200 mV s-1. Furthermore,

    even at a high scan rate of 500 mV s-1, the

    ATO@MnO2 electrode retained a specific capacitance

    of 178 F g-1.

    Figure 6b shows the CD study of the ATO@MnO2

    electrode carried out at current densities of 0.5, 1, 2, 5,

    10, and 20 A g-1 within the potential range of 0 to 1.0

    V (vs. Ag/AgCl). The sloped variation with some

    broad and short plateaus in the discharge profiles,

    which illustrated the near linearity, demonstrates the

    representative pseudo-capacitive behavior of the

  • 8

    Figure 6 Electrochemical performance of ATO@MnO2 electrode. (a) Cyclic voltammetry (CV) behaviors measured at six different scan

    rates of 10, 20, 50, 100, 200, and 500 mV s-1. The inset shows the linearity of the anodic current density with scan rates, (b) Evolution of

    the galvanostatic discharge profiles obtained at current densities of 0.5, 1, 2, 5, 10, and 20 A g-1, (c) Corresponding mass-specific

    capacitance (SCCD) as a function of discharge current densities, (d) Long-term cycling performance (closed-circle) measured at a

    constant current density of 20 A g-1. The closed-squares indicate a superior Coulombic efficiency over average 99%, and each inset of (d)

    shows highly symmetric charge-discharge profiles and excellent SCCD retention until 5,000 consecutive cycling tests. All measurements

    were recorded on the potential window of 0 to 1.0 V (vs. Ag/AgCl).

    ATO@MnO2 electrode. Although the time

    dependence discharge curves in this work exhibited

    overall linear variation with broad and short

    plateaus, they clearly indicated suitable and fast

    pseudo-capacitive characteristics caused by the

    surface redox reaction of ATO@MnO2, which

    corresponded to the results of the broad redox pair

    and quasi-rectangular shapes observed in the CV

    profiles. We also calculated the mass-specific

    capacitance values of the ATO@MnO2 electrode from

    the CD curves, and were denoted as SCCD (see

    Supporting Information; Calculation of specific

    capacitance), and plotted as a function of discharge

    current densities as shown in Figure 6c. The SCCD

    values were 216, 203, 192, 180, 171, and 162 F g-1 at

    current densities of 0.5, 1, 2, 5, 10, and 20 A g-1,

    respectively. In addition, the calculated area-specific

    capacitances (SCCD,area) of ATO@MnO2 electrode were

    also shown in Figure S8. Importantly, the

    ATO@MnO2 electrode exhibited a high SCCD of up to

    216 F g-1, and 75% of the specific capacitance was

    retained even at a high current density of 20 A g-1. In

    particular, an increase in the discharge current

    generally leads to a large potential drop (i.e., iR drop)

    that results in a decrease in the capacitive

    performance, but the SCCD value of the ATO@MnO2

    electrode showed nevertheless over 90% retention

    even at 20 A g-1 as compared to the value at 5 A g-1. Among ATO@MnO2 heterostructures prepared with

    various synthetic condition, the sample prepared from

    0.05 M Mn2+ and 0.3 M BrO3- at 50 oC for 3.5 h

    showed the best capacitance performance (Figure S9),

    which might be originated from relatively higher

    surface area of 2D nanosheets (Figure 4), shorter

    diffusion path to 1D ATO (Figure 5), and larger loading

    amount of MnO2 active material (Table S1). Here, the

    highest specific capacitances of the ATO@MnO2

    electrode in terms of high-rate capability were 178 F

    g-1 at a scan rate of 500 mV s-1 (for CV) and 162 F g-1

    at a current density of 20 A g-1 (for CD); these values

    exceeded other MnO2-based electrodes in previous

    reports (Table S2) [46, 47, 49, 50, 52-72, 76]. In

    addition, we showed the Ragone plot (power density

  • 9

    vs. energy density) for the ATO@MnO2 electrode

    (Figure S10). It is also impressive that the

    ATO@MnO2 electrode delivered a specific power

    density from 250.6 to 10,625 W kg-1 and a specific

    energy density from 108 to 85 Wh kg-1 as the

    galvanostatic charge-discharge current density was

    increased from 0.5 to 20 A g-1.

    Long-term cycling performance of electrode

    materials at higher rate operation conditions is

    essential for real supercapacitor operations. Figure

    6d presents the cycle number dependence of the

    specific capacitance of the ATO@MnO2 electrode. As

    can be seen, the SC value of the ATO@MnO2

    electrode was over 165 F g-1 and maintained

    approximately 91.5% of its initial value after 5,000

    charge-discharge cycles (the inset on the far right in

    Figure 6d), demonstrating the excellent cycling

    stability of the ATO@MnO2 hetero-nanostructures as

    a supercapacitor electrode. Furthermore, the ~100%

    Coulombic efficiency (η) for the CD tests of 5,000

    cycles, a measure of the feasibility of the redox

    process calculated from the discharge and charging

    time (denoted as tD and tC, respectively) in CD

    measurements [73, 74]: 𝜂(%) =𝑡𝐷

    𝑡𝐶× 100 , clearly

    demonstrates the electrochemical suitability of the

    ATO@MnO2 hetero-nanostructures and the higher

    feasibility of the redox process. It can be also seen,

    the charge curves are symmetric to their

    corresponding discharge counterparts for 5,000

    cycles (the inset on the far left in Figure 6d), further

    indicating its excellent reversibility. The similar

    symmetric triangular charge-discharge curves (as

    compared to the initial profile) and high Coulombic

    efficiency demonstrate that no significant structural

    or phase change of the ATO@MnO2

    hetero-nanostructures were induced during the CD

    processes, contributing to its long-term

    electrochemical stability. The microstructural

    stability of the ATO@MnO2 hetero-nanostructures

    after the long-term cycling tests was further

    confirmed using TEM and EDS elemental mapping

    analyses (Figure S11) as well as SEM observation

    (Figure S12), and it illustrated that the 3-D

    hierarchical ATO@MnO2 hetero-nanostructures were

    retained rather well.

    In addition, we also investigated the

    electrochemical performance of bare ATO nanobelt

    electrode as a reference sample of comparison in

    order to emphasize the supercapacitive performance

    of ATO@MnO2 nanostructure electrode (Figure S13).

    Importantly, it is noted that the bare ATO nanobelt

    electrode did not showed the notable characteristics

    in the CV curves, and its specific capacitance values

    calculated from the CD measurement were

    approximately ~1 F g-1 and below, which is negligible.

    Therefore, it is believed that the ATO nanobelts

    (backbone) here can only offer the effective electron

    pathway, but their capacitive performance is not

    affected on the supercapacitive performance of

    ATO@MnO2 nanostructure electrode.

    It is obvious that the ATO@MnO2 electrode

    exhibits high-capacitive performance, which may be

    due to the unique and novel formation of the 3-D

    hierarchical hetero-nanostructures (Figure 7a). First,

    the loosely packed, thin 2-D layered architectures of

    MnO2 and macro-/mesoporous characteristic caused

    by the open space between each 2-D MnO2

    nanosheet as well as the neighboring individual

    ATO@MnO2 hetero-nanostructure can offer a large

    electroactive surface area to electrolyte ions for fast

    Figure 7 (a) Schematic illustration of facile accession of ions in

    electrolyte and electron “superhighways” in the 3-D hierarchical

    self-supported ATO@MnO2 nanosheet electrode, (b) The

    complex-plane impedance plots (Nyquist plots; imaginary part,

    Z" versus real part, Z') of the ATO@MnO2 electrode before and

    after 3,000 cycles. The inset shows the equivalent circuit for the

    EIS spectra, (c) Frequency-dependent specific capacitance values

    of the ATO@MnO2 electrode calculated from the impedance

    measurement. The EIS measurement was carried out by

    imposing a sinusoidal alternating voltage frequency of 100 kHz

    to 0.01 Hz at open circuit voltage (OCV), an alternating current

    (AC) amplitude of 5 mV, and a constant direct current (DC) bias

    potential of 1.0 V (vs. Ag/AgCl) under the same conditions: a

    three-electrode system with an aqueous electrolyte solution (1 M

    Na2SO4) at room temperature.

    surface sorption reactions and easy accessibility of

  • 10

    ions for highly feasible redox reactions, which may

    ensure sufficient electrochemical utilization of the

    MnO2 during the capacitive reaction. Furthermore,

    this morphology can also serve as an ‘ion-buffering

    reservoir’ of electrolyte ions [73, 75], resulting in

    better penetration and occurrence of efficient

    Faradaic reactions even at very high-rates, thereby

    leading to not only reduced internal and charge

    transfer resistance, but also enhanced power

    characteristic by shortening the diffusion length for

    both electrolyte ions and charges during the

    charge-discharge process. Second, the self-supported

    conductive ATO NBs array combined with the thin

    MnO2 nanosheets can provide excellent interfacial

    contacts and highly conductive paths throughout the

    MnO2 nanosheets for rapid electronic transport as a

    “superhighway”. In our previous study, the

    conductivity of an individual ATO NB was about

    1.03×104 S m-1, which corresponds to the metallic

    property [43]. Thus, ATO/Ti electrode can provide

    1-D conductive path, which is also confirmed by low

    sheet resistance of the electrode (6.74 mΩ/□). Even

    after MnO2 nanosheets coating, the electrode has

    quite low sheet resistance (42.7 mΩ/□). Third, the

    3-D hierarchical framework and voids give the

    microstructural stability and fast ion diffusion into

    the entire electrode matrix, thus enhancing the

    electrochemical kinetics, and also accommodating

    the strain arising due to high-rate insertion and

    extraction of electrolyte ions, which are

    synergistically beneficial for a strong cycling life

    along with high-rate capability.

    To gain further insight into the origin of superior

    electrochemical performance of the ATO@MnO2

    electrode, electrochemical impedance spectroscopy

    (EIS) measurements were carried out before

    (denoted as after 1st cycle in Figure 7b) and after the

    CD tests of 3,000 cycles in the frequency range of

    0.01–105 Hz at room temperature, and the obtained

    impedance spectra were analyzed by the CNLS

    fitting method based on a Randles equivalent circuit

    depicted in the inset of Figure 7b [51, 55].

    Importantly, the measured specific internal

    resistance (Rs) as well as the charge-transfer

    resistance (Rct) fitting values of the ATO@MnO2

    electrode were similar until 3,000 cycles (Rs: 13.19

    and 13.22 Ω mg-1 and Rct: 34.17 and 36.12 Ω mg-1,

    respectively), indicating not only a good electrolyte

    conductivity and electrochemical stability of the

    ATO@MnO2 electrode, but also the significant

    feasibility of the redox reactions even after long-term

    CD processes at high current densities, because of

    the enhancement of the diffusivity of the electrolyte

    ions in the electroactive sites and highly conductive

    paths for rapid electronic transport, which in turn,

    reduced the Rct without any appreciable structural

    change during high-rate cycling, thereby leading to

    an excellent pseudo-capacitive performance.

    Additionally, the profile exhibited higher angles than

    45 °, indicating the suitability as electrode materials

    for supercapacitors. The detailed EIS analyses are

    described in the Supporting Information:

    Electrochemical impedance spectroscopy (EIS)

    analysis. For a more informative representation, we

    converted the measured EIS data of the ATO@MnO2

    electrode before and after 3,000 cycles into specific

    capacitance (denoted as SCEIS, F g-1) using the

    following equation [68, 73]:

    𝑆𝐶𝐸𝐼𝑆 =1

    2𝜋𝑓𝑍" (6)

    and plotted the data as a function of frequency, “f”,

    as shown in Figure 7c. The calculated SCEIS values

    before and after 3,000 cycles were relatively similar

    at all frequency regions, even though the Rct is

    slightly increased after 3,000 cycles. Moreover, for

    the frequency range of 0.01–0.1 Hz, the ATO@MnO2

    electrode showed a maximum SCEIS value. This is

    because electrolyte ions could easily penetrate into

    the ATO@MnO2 hetero-nanostructures with a 3-D

    hierarchical form and are electrochemically more

    accessible over a reasonable frequency range

    measured for the electrode. However, a minimum of

    ~5% decrease in the SCEIS value after 3,000 cycles was

    also observed at the lower frequency region, which

    well corroborates with the result from the CD study.

    Interestingly, no decrease in the SCEIS value at higher

    operating frequency regions after long-term cycling

    tests at elevated current density condition is very

    crucial for the ATO@MnO2 as a possible electrode

    material for high-performance supercapacitor

    applications.

    3. Conclusion

    In conclusion, we successfully developed 3-D MnO2

  • 11

    nanostructures composed of 2-D ultrathin MnO2

    nanosheets and a 1-D highly conductive backbone by

    ECBD. Two-dimensional MnO2 nanosheets were

    uniformly grown perpendicular to the z-axis of the

    backbone and had a circular band shape

    surrounding the backbone. In this system, the

    addition of an oxidizer, particularly BrO3-, acted as a

    novel morphology controller and compensated for

    the demerit of the CBD method, namely the

    non-uniform formation and slow reaction time,

    enabling the one-step synthesis of the 2-D MnO2

    nanosheets on an arbitrary 1-D backbone. Moreover,

    the size and thickness of the nanosheets were easily

    tunable over a wide range (diameter: 0.4 – 1.5 μm

    and thickness: 10 – 100 nm) via the reaction duration

    and temperature. Furthermore, the ATO@MnO2

    electrode showed not only a high-rate capability of

    178 F g-1 at a scan rate of 500 mV s-1 (for CV) and 162

    F g-1 at a current density of 20 A g-1 (for CD), but also

    excellent long-term cycle stability as it retained

    approximately 91.5% of its initial value after 5,000

    charge-discharge cycles. This can be attributed to the

    unique 3-D hetero-nanostructure formation, namely

    loosely packed, thin MnO2 nanosheets combined

    with self-supported ATO NWs, and thereby, this

    structure leaded to enhanced electrochemical

    kinetics of the capacitive reaction. This work

    suggests not only the development of tailored

    nanosheets in 3-D hierarchical heterostructures, but

    also the possibility of oxidizers, which can act as

    controlling agents and as growth accelerators in

    low-temperature CBD. We envisage that the ECBD

    method can be expanded to achieve 2-D nanosheets

    engineering in 3-D nanostructures in other fields,

    such as photo-conversion devices, energy storage

    systems, etc.

    4. Methods

    4.1. Synthesis

    1-D Conductive ATO backbones were prepared by

    the vapor-liquid-solid (VLS) mechanism using a

    conventional procedure [42, 43]. In this study, 10 µm

    ATO nanobelt arrays were deposited on Ti foil

    (99.7%, thickness 0.127 mm, Sigma-Aldrich, USA) by

    thermal evaporation. For VLS growth, a 2 nm-thick

    Au catalyst was deposited on the Ti substrate by an

    evaporator. A mixed powder of Sn (99.5%, Samchun,

    Korea) and Sb (99.9%, high purity chemicals, Japan)

    in a 3:1 ratio was used as the source and was loaded

    on a quartz boat and placed into the center of a fused

    silica tube. The as-prepared Au/Ti substrates were

    inserted into the tube furnace 20 cm from the source.

    The source was heated to 1073 K under vacuum

    conditions (< 2×10-3 Torr) and the ATO nanobelts

    were grown on the Ti substrate at 873 K with an

    oxygen flow of 12 sccm.

    MnO2 were synthesized in a manner similar to that

    previously reported after a slight modification [41].

    2-D MnO2 nanosheets were formed on the surface of

    the ATO nanobelts via the oxidation reaction of Mn2+

    by one-step CBD with BrO3- as the oxidizer. In a

    typical synthesis, MnCl2·4H2O (99%, Sigma-Aldrich,

    USA) was dissolved in deionized water and KBrO3

    (99.8%, Sigma-Aldrich, USA) was added with

    various molar concentrations of Mn2+ (0.005 M to 0.1

    M) and BrO3- (0.1 to 0.3 M). The prepared suspension

    was stirred for 5 min and sonicated for 10 min at

    room temperature to dissolve all reagents. The ATO

    nanobelt electrodes at the bottom of the vial

    containing 10 mL of the solution were heated to

    different reaction temperatures (40 oC to 100 oC) for

    various amounts of time (2 h to 63 h). After the ECBD

    reaction, the samples were rinsed in deionized water

    and ethanol, and finally dried at room temperature.

    4.2. Characterization

    The crystallographic characteristics of the 3-D MnO2

    electrodes were investigated by XRD using a Bruker

    D8-Advance (Cu Kα1 radiation) instrument operated

    at 40 kV and 40 mA. The crystal structures were

    double checked by a Renishaw Raman spectrometer

    (inVia Raman microscope). The morphologies of the

    samples were investigated by a JEOL scanning

    electron microscope (JSM-6330F). TEM studies were

    performed using a JEM-2100F equipped with an

    energy dispersive X-ray analysis (EDS) system.

    4.3. Electrochemical characterization

    A home-made, three-electrode cell system was used

    to measure the response of the 3-D hierarchical,

    self-supported ATO@MnO2 nanosheets on the Ti foil,

    which were used directly as the working electrodes

    (~1 cm2) without cohesive binders and conductive

  • 12

    additives. An aqueous solution of 1 M Na2SO4 was

    utilized as the electrolyte with Pt gauze (2 × 3 cm2) as

    the counter electrode and an Ag/AgCl (saturated KCl)

    electrode as the reference electrode. Using this

    three-electrode configuration, the electrochemical

    characteristics were determined on an

    electrochemical workstation (model Ivium-n-Stat

    electrochemical analyzer, Ivium Technologies B. V.,

    Netherlands). Cyclic voltammetry (CV) and

    galvanostatic charge-discharge (CD) techniques were

    employed to evaluate the electrochemical

    performance of the supercapacitor electrodes. The

    CVs were measured in the voltage window between

    0 and 1.0 V (vs. Ag/AgCl) at scan rates from 10 to 500

    mV s-1, and the CDs were conducted in the same

    potential window range with current densities of 0.5,

    1, 2, 5, 10, and 20 A g-1. The cycle stability was also

    evaluated using CD measurement at a constant

    current density of 20 A g-1 for over 5,000 cycles. All

    operating current densities were calculated based on

    the mass of the active material (mass of MnO2

    nanosheets). Electrochemical impedance

    spectroscopy (EIS) was carried out in the frequency

    range of 0.01 Hz–100 kHz with a perturbation

    amplitude of 5 mV versus the open circuit potential

    (OCV), and the obtained impedance spectra were

    analyzed quantitatively by curve fitting with Z-view

    software (Version. 2.90, Scribner Associates Inc.). All

    electrochemical measurements were performed at

    room temperature.

    Acknowledgements

    This research was supported by Basic Science

    Research Program through the National Research

    Foundation of Korea(NRF) funded by the Ministry of

    Science, ICT and future Planning

    (2012R1A2A2A01045382 and 2010-0029027).

    Electronic Supplementary Material: Supplementary

    material (further TEM, SEM, XRD and EIS analysis)

    is available in the online version of this article at

    http://dx.doi.org/10.1007/s12274-***-****-*

    References [1] Zordan, T.; Hepler, L. G. Thermochemistry and oxidation

    potentials of manganese and its compounds. Chem. Rev.

    1968, 68, 737-745.

    [2] Cheng, F.; Zhao, J.; Song, W.; Li, C.; Ma, H.; Chen, J.;

    Shen, P. Facile controlled synthesis of MnO2

    nanostructures of novel shapes and their application in

    batteries. Inorg. Chem. 2006, 45, 2038-2044.

    [3] Subramanian, V.; Zhu, H.; Wei, B. Alcohol-assisted room

    temperature synthesis of different nanostructured

    manganese oxides and their pseudocapacitance properties

    in neutral electrolyte. Chem. Phys. Lett. 2008, 453,

    242-249.

    [4] Zhang, H.; Cao, G.; Wang, Z.; Yang, Y.; Shi, Z.; Gu, Z.

    Growth of manganese oxide nanoflowers on

    vertically-aligned carbon nanotube arrays for high-rate

    electrochemical capacitive energy storage. Nano Lett. 2008,

    8, 2664-2668.

    [5] Débart, A.; Paterson, A. J.; Bao, J.; Bruce, P. G. α‐MnO2 Nanowires: A Catalyst for the O2 Electrode in

    Rechargeable Lithium Batteries. Angew. Chem. Int. Edit.

    2008, 120, 4597-4600.

    [6] Li, W. N.; Yuan, J.; Shen, X. F.; Gomez‐Mower, S.; Xu, L. P.; Sithambaram, S.; Aindow, M.; Suib, S. L.

    Hydrothermal Synthesis of Structure‐and Shape‐Controlled Manganese Oxide Octahedral Molecular Sieve

    Nanomaterials. Adv. Func. Mater. 2006, 16, 1247-1253.

    [7] Long, J. W.; Rhodes, C. P.; Young, A. L.; Rolison, D. R.

    Ultrathin, protective coatings of poly (o-phenylenediamine)

    as electrochemical proton gates: making mesoporous MnO2

    nanoarchitectures stable in acid electrolytes. Nano Lett.

    2003, 3, 1155-1161.

    [8] Luo, X.; Morrin, A.; Killard, A. J.; Smyth, M. R.

    Application of nanoparticles in electrochemical sensors and

    biosensors. Electroanal 2006, 18, 319-326.

    [9] Sayle, T. X.; Maphanga, R. R.; Ngoepe, P. E.; Sayle, D. C.

    Predicting the electrochemical properties of MnO2

    nanomaterials used in rechargeable Li batteries: simulating

    nanostructure at the atomistic level. J. Am. Chem. Soc.

    2009, 131, 6161-6173.

    [10] Ghodbane, O.; Pascal, J. L.; Favier, F. Microstructural

    effects on charge-storage properties in MnO2-based

    electrochemical supercapacitors. ACS Appl. Mater.

    Interfaces 2009, 1, 1130-1139.

    [11] Fan, Z.; Yan, J.; Wei, T.; Zhi, L.; Ning, G.; Li, T.; Wei, F.

    Asymmetric Supercapacitors Based on Graphene/MnO2

    and Activated Carbon Nanofiber Electrodes with High

    Power and Energy Density. Adv. Func. Mater. 2011, 21,

    2366-2375.

    [12] Brousse, T.; Toupin, M.; Dugas, R.; Athouël, L.; Crosnier,

    O.; Bélanger, D. Crystalline MnO2 as possible alternatives

    to amorphous compounds in electrochemical

    supercapacitors. J. Electrochem. Soc. 2006, 153,

    A2171-A2180.

    [13] Xu, C.; Du, H.; Li, B.; Kang, F.; Zeng, Y. Asymmetric

    activated carbon-manganese dioxide capacitors in mild

    aqueous electrolytes containing alkaline-earth cations. J.

    Electrochem. Soc. 2009, 156, A435-A441.

    [14] Xiong, Y.; Xie, Y.; Li, Z.; Wu, C. Growth of Well‐Aligned γ‐MnO2 Monocrystalline Nanowires through a Coordination‐Polymer‐Precursor Route. Chem.-Eur. J. 2003, 9, 1645-1651.

    http://dx.doi.org/10.1007/s12274-***-****-*

  • 13

    [15] Jones, D. J.; Wortham, E.; Rozière, J.; Favier, F.; Pascal,

    J.-L.; Monconduit, L. Manganese oxide nanocomposites:

    preparation and some electrochemical properties. J. Phys.

    Chem. Solids 2004, 65, 235-239.

    [16] Lee, S. W.; Kim, J.; Chen, S.; Hammond, P. T.; Shao-Horn,

    Y. Carbon nanotube/manganese oxide ultrathin film

    electrodes for electrochemical capacitors. ACS Nano 2010,

    4, 3889-3896.

    [17] Dong, X.; Shen, W.; Gu, J.; Xiong, L.; Zhu, Y.; Li, H.; Shi,

    J. MnO2-embedded-in-mesoporous-carbon-wall structure

    for use as electrochemical capacitors. J. Phys. Chem. B

    2006, 110, 6015-6019.

    [18] He, Y.; Chen, W.; Li, X.; Zhang, Z.; Fu, J.; Zhao, C.; Xie,

    E. Freestanding three-dimensional graphene/MnO2

    composite networks as ultralight and flexible

    supercapacitor electrodes. ACS Nano 2012, 7, 174-182.

    [19] Jiang, H.; Li, C.; Sun, T.; Ma, J. A green and high energy

    density asymmetric supercapacitor based on ultrathin

    MnO2 nanostructures and functional mesoporous carbon

    nanotube electrodes. Nanoscale 2012, 4, 807-12.

    [20] Lei, Z.; Zhang, J.; Zhao, X. S. Ultrathin MnO2 nanofibers

    grown on graphitic carbon spheres as high-performance

    asymmetric supercapacitor electrodes. J. Mater. Chem.

    2012, 22, 153.

    [21] Zheng, H.; Wang, J.; Jia, Y.; Ma, C. a. In-situ synthetize

    multi-walled carbon nanotubes@MnO2 nanoflake core–

    shell structured materials for supercapacitors. J. Power

    Sources 2012, 216, 508-514.

    [22] Zhou, R.; Meng, C.; Zhu, F.; Li, Q.; Liu, C.; Fan, S.; Jiang,

    K. High-performance supercapacitors using a nanoporous

    current collector made from super-aligned carbon

    nanotubes. Nanotechnology 2010, 21, 345701.

    [23] Chen, S.; Zhu, J.; Wu, X.; Han, Q.; Wang, X. Graphene

    oxide−MnO2 nanocomposites for supercapacitors. ACS

    Nano 2010, 4, 2822-2830.

    [24] Wu, Z.-S.; Ren, W.; Wang, D.-W.; Li, F.; Liu, B.; Cheng,

    H.-M. High-energy MnO2 nanowire/graphene and

    graphene asymmetric electrochemical capacitors. ACS

    Nano 2010, 4, 5835-5842.

    [25] Yu, G.; Hu, L.; Liu, N.; Wang, H.; Vosgueritchian, M.;

    Yang, Y.; Cui, Y.; Bao, Z. Enhancing the supercapacitor

    performance of graphene/MnO2 nanostructured electrodes

    by conductive wrapping. Nano Lett.2011, 11, 4438-4442.

    [26] Gao, H.; Xiao, F.; Ching, C. B.; Duan, H.

    High-performance asymmetric supercapacitor based on

    graphene hydrogel and nanostructured MnO2. ACS Appl.

    Mater. Interfaces 2012, 4, 2801-2810.

    [27] Mao, L.; Zhang, K.; On Chan, H. S.; Wu, J.

    Nanostructured MnO2/graphene composites for

    supercapacitor electrodes: the effect of morphology,

    crystallinity and composition. J. Mater. Chem. 2012, 22,

    1845-1851.

    [28] Yan, J.; Khoo, E.; Sumboja, A.; Lee, P. S. Facile coating of

    manganese oxide on tin oxide nanowires with

    high-performance capacitive behavior. ACS Nano 2010, 4,

    4247-4255.

    [29] Bao, L.; Zang, J.; Li, X. Flexible Zn2SnO4/MnO2 core/shell

    nanocable-carbon microfiber hybrid composites for

    high-performance supercapacitor electrodes. Nano Lett.

    2011, 11, 1215-1220.

    [30] He, S.; Chen, W. High performance supercapacitors based

    on three-dimensional ultralight flexible manganese oxide

    nanosheets/carbon foam composites. J. Power Sources

    2014, 262, 391-400.

    [31] He, S.; Hu, C.; Hou, H.; Chen, W. Ultrathin MnO2

    nanosheets supported on cellulose based carbon papers for

    high-power supercapacitors. J. Power Sources 2014, 246,

    754-761.

    [32] Omomo, Y.; Sasaki, T.; Wang, L.; Watanabe, M.

    Redoxable nanosheet crystallites of MnO2 derived via

    delamination of a layered manganese oxide. J. Am. Chem.

    Soc. 2003, 125, 3568-3575.

    [33] Yang, X.; Makita, Y.; Liu, Z.-h.; Sakane, K.; Ooi, K.

    Structural characterization of self-assembled MnO2

    nanosheets from birnessite manganese oxide single crystals.

    Chem. Mater. 2004, 16, 5581-5588.

    [34] Deng, R.; Xie, X.; Vendrell, M.; Chang, Y.-T.; Liu, X.

    Intracellular glutathione detection using

    MnO2-nanosheet-modified upconversion nanoparticles. J.

    Am. Chem. Soc. 2011, 133, 20168-20171.

    [35] Sun, X.; Li, Q.; Lu, Y.; Mao, Y. Three-dimensional

    ZnO@MnO2 core@shell nanostructures for

    electrochemical energy storage. Chem. Comm. 2013, 49,

    4456-4458.

    [36] Zhao, G.; Li, J.; Jiang, L.; Dong, H.; Wang, X.; Hu, W.

    Synthesizing MnO2 nanosheets from graphene oxide

    templates for high performance pseudosupercapacitors.

    Chem. Sci. 2012, 3, 433-437.

    [37] Peng, L.; Peng, X.; Liu, B.; Wu, C.; Xie, Y.; Yu, G.

    Ultrathin two-dimensional MnO2/graphene hybrid

    nanostructures for high-performance, flexible planar

    supercapacitors. Nano Lett. 2013, 13, 2151-2157.

    [38] Liu, J.; Jiang, J.; Cheng, C.; Li, H.; Zhang, J.; Gong, H.;

    Fan, H. J. Co3O4 Nanowire@ MnO2 Ultrathin Nanosheet

    Core/Shell Arrays: A New Class of High‐Performance Pseudocapacitive Materials. Adv. Mater. 2011, 23,

    2076-2081.

    [39] Tian, W.; Wang, X.; Zhi, C.; Zhai, T.; Liu, D.; Zhang, C.;

    Golberg, D.; Bando, Y. Ni(OH)2 nanosheet@Fe2O3

    nanowire hybrid composite arrays for high-performance

    supercapacitor electrodes. Nano Energ. 2013, 2, 754-763.

    [40] Yang, Q.; Lu, Z.; Chang, Z.; Zhu, W.; Sun, J.; Liu, J.; Sun,

    X.; Duan, X. Hierarchical Co3O4 nanosheet@ nanowire

    arrays with enhanced pseudocapacitive performance. RSC

    Adv. 2012, 2, 1663-1668.

    [41] Unuma, H.; Kanehama, T.; Yamamoto, K.; Watanabe, K.;

    Ogata, T.; Sugawara, M. Preparation of thin films of MnO2

    and CeO2 by a modified chemical bath

    (oxidative-soak-coating) method. J. Mater. Sci. 2003, 38,

    255-259.

    [42] Park, S.; Lee, S.; Seo, S. W.; Seo, S.-D.; Lee, C. W.; Kim,

    D.; Kim, D.-W.; Hong, K. S. Tailoring nanobranches in

    three-dimensional hierarchical rutile heterostructures: a

    case study of TiO2–SnO2. CrystEngComm 2013, 15, 2939.

    [43] Park, S.; Seo, S.-D.; Lee, S.; Seo, S. W.; Park, K.-S.; Lee,

    C. W.; Kim, D.-W.; Hong, K. S. Sb:SnO2@TiO2

    Heteroepitaxial Branched Nanoarchitectures for Li Ion

  • 14

    Battery Electrodes. J. Phys. Chem. C 2012, 116,

    21717-21726.

    [44] Julien, C.; Massot, M.; Rangan, S.; Lemal, M.; Guyomard,

    D. Study of structural defects in γ-MnO2 by Raman

    spectroscopy. J. Raman Spectrosc. 2002, 33, 223-228.

    [45] Conway, B. E. Electrochemical Supercapacitors: Scientific

    Fundamentals and Technological Applications, Kluwer

    Academic/Plenum Press, 1999.

    [46] Devaraj, S.; Munichandraiah, N. Effect of crystallographic

    structure of MnO2 on its electrochemical capacitance

    properties. J. Phys. Chem. C 2008, 112, 4406-4417.

    [47] Toupin, M.; Brousse, T.; Bélanger, D. Charge storage

    mechanism of MnO2 electrode used in aqueous

    electrochemical capacitor. Chem. Mater. 2004, 16,

    3184-3190.

    [48] Pang, S. C.; Anderson, M. A.; Chapman, T. W. Novel

    electrode materials for thin‐film ultracapacitors: comparison of electrochemical properties of sol‐gel‐derived and electrodeposited manganese dioxide. J.

    Electrochem. Soc. 2000, 147, 444-450.

    [49] Qu, Q.; Zhang, P.; Wang, B.; Chen, Y.; Tian, S.; Wu, Y.;

    Holze, R. Electrochemical performance of MnO2 nanorods

    in neutral aqueous electrolytes as a cathode for asymmetric

    supercapacitors. J. Phys. Chem. C 2009, 113,

    14020-14027.

    [50] Li, S.; Qi, L.; Lu, L.; Wang, H. Facile preparation and

    performance of mesoporous manganese oxide for

    supercapacitors utilizing neutral aqueous electrolytes. RSC

    Adv.2012, 2, 3298-3308.

    [51] Kim, J. S.; Shin, S. S.; Han, H. S.; Oh, L. S.; Kim, D. H.;

    Kim, J. H.; Hong, K. S.; Kim, J. Y. 1-D structured flexible

    supercapacitor electrodes with prominent electronic/ionic

    transport capabilities. ACS Appl. Mater. Interfaces 2014, 6,

    268-274.

    [52] Xie, X.; Zhang, C.; Wu, M. B.; Tao, Y.; Lv, W.; Yang, Q.

    H. Porous MnO2 for use in a high performance

    supercapacitor: replication of a 3D graphene network as a

    reactive template. Chem. Commun. 2013, 49,

    11092-11094.

    [53] Yeager, M.; Du, W.; Si, R.; Su, D.; Marinković, N.; Teng,

    X. Highly Efficient K0.15MnO2 Birnessite Nanosheets for

    Stable Pseudocapacitive Cathodes. J. Phys. Chem. C 2012,

    116, 20173-20181.

    [54] Li, W.; Liu, Q.; Sun, Y.; Sun, J.; Zou, R.; Li, G.; Hu, X.;

    Song, G.; Ma, G.; Yang, J.; Chen, Z.; Hu, J. MnO2

    ultralong nanowires with better electrical conductivity and

    enhanced supercapacitor performances. J. Mater. Chem.

    2012, 22, 14864.

    [55] Huang, M.; Zhang, Y.; Li, F.; Zhang, L.; Ruoff, R. S.; Wen,

    Z.; Liu, Q. Self-assembly of mesoporous nanotubes

    assembled from interwoven ultrathin birnessite-type MnO2

    nanosheets for asymmetric supercapacitors. Sci. Rep. 2014,

    4, 3878.

    [56] Xu, C.; Zhao, Y.; Yang, G.; Li, F.; Li, H. Mesoporous

    nanowire array architecture of manganese dioxide for

    electrochemical capacitor applications. Chem.

    Commun.2009, 7575-7.

    [57] Feng, Z.-P.; Li, G.-R.; Zhong, J.-H.; Wang, Z.-L.; Ou,

    Y.-N.; Tong, Y.-X. MnO2 multilayer nanosheet clusters

    evolved from monolayer nanosheets and their predominant

    electrochemical properties. Electrochem. Commun. 2009,

    11, 706-710.

    [58] Subramanian, V.; Zhu, H.; Wei, B. Nanostructured MnO2:

    Hydrothermal synthesis and electrochemical properties as a

    supercapacitor electrode material. J. Power Sources 2006,

    159, 361-364.

    [59] Xiao, W.; Xia, H.; Fuh, J. Y. H.; Lu, L. Growth of

    single-crystal α-MnO2 nanotubes prepared by a

    hydrothermal route and their electrochemical properties. J.

    Power Sources 2009, 193, 935-938.

    [60] Xu, M.; Kong, L.; Zhou, W.; Li, H. Hydrothermal

    synthesis and pseudocapacitance properties of α-MnO2

    hollow spheres and hollow urchins. J. Phys. Chem. C 2007,

    111, 19141-19147.

    [61] Jiang, R.; Huang, T.; Liu, J.; Zhuang, J.; Yu, A. A novel

    method to prepare nanostructured manganese dioxide and

    its electrochemical properties as a supercapacitor electrode.

    Electrochim. Acta 2009, 54, 3047-3052.

    [62] Vargas, O. A.; Caballero, A.; Hernán, L.; Morales, J.

    Improved capacitive properties of layered manganese

    dioxide grown as nanowires. J. Power Sources 2011, 196,

    3350-3354.

    [63] Sung, D.-Y.; Kim, I. Y.; Kim, T. W.; Song, M.-S.; Hwang,

    S.-J. Room Temperature Synthesis Routes to the 2D

    Nanoplates and 1D Nanowires/Nanorods of Manganese

    Oxides with Highly Stable Pseudocapacitance Behaviors. J.

    Phys. Chem. C 2011, 115, 13171-13179.

    [64] Subramanian, V.; Zhu, H.; Vajtai, R.; Ajayan, P.; Wei, B.

    Hydrothermal synthesis and pseudocapacitance properties

    of MnO2 nanostructures. J. Phys. Chem. B 2005, 109,

    20207-20214.

    [65] Zhu, G.; Li, H.; Deng, L.; Liu, Z.-H. Low-temperature

    synthesis of δ-MnO2 with large surface area and its

    capacitance. Mater. Lett. 2010, 64, 1763-1765.

    [66] Ragupathy, P.; Park, D. H.; Campet, G.; Vasan, H.; Hwang,

    S.-J.; Choy, J.-H.; Munichandraiah, N. Remarkable

    capacity retention of nanostructured manganese oxide upon

    cycling as an electrode material for supercapacitor. J. Phys.

    Chem. C 2009, 113, 6303-6309.

    [67] Yuan, J.; Liu, Z.-H.; Qiao, S.; Ma, X.; Xu, N. Fabrication

    of MnO2-pillared layered manganese oxide through an

    exfoliation/reassembling and oxidation process. J. Power

    Sources 2009, 189, 1278-1283.

    [68] Yuan, C.; Gao, B.; Su, L.; Zhang, X. Interface synthesis of

    mesoporous MnO2 and its electrochemical capacitive

    behaviors. J. Colloid Interface Sci. 2008, 322, 545-550.

    [69] Xu, M.-W.; Jia, W.; Bao, S.-J.; Su, Z.; Dong, B. Novel

    mesoporous MnO2 for high-rate electrochemical capacitive

    energy storage. Electrochim. Acta 2010, 55, 5117-5122.

    [70] He, X.; Yang, M.; Ni, P.; Li, Y.; Liu, Z.-H. Rapid synthesis

    of hollow structured MnO2 microspheres and their

    capacitance. Colloid Surf. A-Physicochem. Eng. Asp. 2010,

    363, 64-70.

    [71] Yu, P.; Zhang, X.; Chen, Y.; Ma, Y. Self-template route to

    MnO2 hollow structures for supercapacitors. Mater. Lett.

    2010, 64, 1480-1482.

  • 15

    [72] Tang, N.; Tian, X.; Yang, C.; Pi, Z. Facile synthesis of

    α-MnO2 nanostructures for supercapacitors. Mater. Res.

    Bull. 2009, 44, 2062-2067.

    [73] Meher, S. K.; Rao, G. R. Ultralayered Co3O4 for

    High-Performance Supercapacitor Applications. J. Phys.

    Chem. C 2011, 115, 15646-15654.

    [74] Meher, S. K.; Justin, P.; Rao, G. R. Nanoscale morphology

    dependent pseudocapacitance of NiO: Influence of

    intercalating anions during synthesis. Nanoscale 2011, 3,

    683-692.

    [75] Shim, H. W.; Lim, A. H.; Kim, J. C.; Jang, E.; Seo, S. D.;

    Lee, G. H.; Kim, T. D.; Kim, D. W. Scalable one-pot

    bacteria-templating synthesis route toward hierarchical,

    porous-Co3O4 superstructures for supercapacitor electrodes.

    Sci. Rep. 2013, 3, 2325.

    [76] Peng, Y.; Chen, Z.; Wen, J.; Xiao, Q.; Weng, D.; He, S.;

    Geng, H.; Lu, Y. Hierarchical manganese oxide/carbon

    nanocomposites for supercapacitor electrodes. Nano Res.

    2011, 4, 216-225.

  • 16

    Electronic Supplementary Material

    Tailoring uniform γ-MnO2 nanosheets on highly conductive three-dimensional current collectors for high-performance supercapacitor electrodes

    Sangbaek Park1†, Hyun-Woo Shim2†, Chan Woo Lee1, Hee Jo Song1, Ik Jae Park1, Jae-Chan Kim2, Kug Sun Hong1, and Dong-Wan Kim2()

    1 Department of Materials Science and Engineering, Seoul National University, Seoul 151-744, Korea 2 School of Civil, Environmental and Architectural Engineering, Korea University, Seoul 136-713, Korea

    † These authors contributed equally

    Supporting information to DOI 10.1007/s12274-****-****-* (automatically inserted by the publisher)

    Calculation of specific capacitance

    The following equations were used to derive the mass-specific capacitance (SC, F g-1), measured in Faradays

    per gram, from the CV and CD profiles, respectively. Each SC value is denoted as SCCV and SCCD. First, the SCCV

    values were calculated by integrating the area under the I–V curves and then dividing by the sweep rate ν (mV

    s-1), the mass of the loaded active material (MnO2) in the electrode, and the working potential window (Va to Vc),

    using the following equation [48,51,53,73]:

    𝑆𝐶𝐶𝑉 =1

    𝜈𝑚(𝑉𝑐−𝑉𝑎)∫ 𝑖(𝑉)𝑑𝑉𝑉𝑐𝑉𝑎

    (1)

    The other mass-specific capacitance, SCCD values were also calculated from the galvanostatic discharge curves

    using the following equation [51,53,55]:

    𝑆𝐶𝐶𝐷 =𝑖

    𝑚(∆𝑉/∆𝑡) (2)

    where, i (mA) is the applied constant discharge current, ∆t (s) is the discharge time, ∆V (V) is the potential

    sweep window (1.0 V), and m (mg) is the mass of the loaded active material (MnO2) in the electrode.

    -----------------------------------------------------------------------------------------------------------------------------------------------------------------

    References

    48. S. -C. Pang, M. A. Anderson and T. W. Chapman, J. Electrochem. Soc., 2000, 147, 444.

    51. J. S. Kim, S. S. Shin, H. S. Han, L. S. Oh, D. H. Kim, J. -H. Kim, K. S. Hong and J. Y. Kim, ACS Appl. Mater. Interfaces, 2014,

    6, 268.

    53. M. Yeager, W. Du, R. Si, D. Su, N. Marinković and X. Teng, J. Phys. Chem. C, 2012, 116, 20173.

    55. M. Huang, Y. Zhang, F. Li, L. Zhang, R. S. Ruoff, Z. Wen and Q. Liu, Sci. Rep., 2014, 4, 3878.

    73. S. K. Meher and G. R. Rao, J. Phys. Chem. C, 2011, 115, 15646.

    ————————————

    Address correspondence to D.-W. Kim, [email protected]

  • 17

    Electrochemical impedance spectroscopy (EIS) analysis

    The EIS spectra are composed of two major characteristic features in the frequency regions: i) a depressed arc

    (partial semicircle) in the high- and middle-frequency regions, and ii) an inclined line (straight, sloped line

    along the imaginary axis (Z″)) in the low-frequency region with a transition between the two regions called

    the “knee frequency”, which illustrates typical capacitor behavior. These EIS characteristics are attributed to

    various resistance phenomena during different interfacial processes in Faradaic reactions. The obtained

    impedance spectra were analyzed by the CNLS fitting method based on a Randles equivalent circuit

    depicted in the inset of Figure 8b, which was the same as the circuit recently employed for the analysis of a

    MnO2-based electrode by Huang et al.[55] and Kim et al.,[51] where Rs and Rct are the internal resistance and

    Faradaic interfacial charge-transfer resistance, respectively, and constant phase element (CPEn) represents the

    capacitive performance. The interfacial diffusive resistance (Warburg impedance) has been denoted as “W”.

    In this equivalent series resistance (ESR), the intercept at the real impedance (Z) axis in the high-frequency

    region first corresponds to Rs, which represents the total resistance related to a combination of the ionic

    resistance of the electrolyte, intrinsic resistance of the active materials, and contact resistance at the active

    material-current collector interface. The higher Rs value indicates a lower electrical conductivity of the

    electrode materials and vice versa [S1]. Meanwhile, the semicircular arc in the high- and middle-frequency

    range, corresponding to the impedance behavior, also indicates contributing resistance from the Faradaic

    redox process, named the charge-transfer resistance (Rct) at the electrode-electrolyte interface. This results

    from the diffusion of electrons and can be calculated from the diameter of each semicircle. Such

    charge-transfer resistance is directly related to the surface phenomena of the electrode materials: i) the

    charge-transfer barriers, including the “electron transfer” between the current collector and the active

    materials, and ii) the “ion-transfer” between the active materials and the electrolyte.49 In addition, the linear

    parts of the Nyquist plot in the lower frequency range correspond to the Warburg impedance, W, which is

    described as the diffusive resistance of the electrolyte ions in the host materials (active sites). Each profile

    exhibited higher angles than 45 °, indicating the suitability as an electrode material for supercapacitors. In

    particular, the profile after 3,000 cycles almost tended to a vertical asymptote along the imaginary line axis,

    indicating the good electrochemical capacitance of the 3-D hierarchical ATO@MnO2 nanosheets in the

    Na2SO4 aqueous electrolyte. Therefore, the overall EIS characteristics suggest superior accessibility of the

    electrolyte ions and highly conductive paths for rapid electronic transport through the 3-D hierarchical

    self-supported hetero-nanostructures and explicit contribution of the pseudo-capacitance to the energy

    storage performance of the ATO@MnO2 material.

    -----------------------------------------------------------------------------------------------------------------------------------------------------------------

    References

    51. J. S. Kim, S. S. Shin, H. S. Han, L. S. Oh, D. H. Kim, J. -H. Kim, K. S. Hong and J. Y. Kim, ACS Appl. Mater. Interfaces, 2014,

    6, 268.

    55. M. Huang, Y. Zhang, F. Li, L. Zhang, R. S. Ruoff, Z. Wen and Q. Liu, Sci. Rep., 2014, 4, 3878.

    S1. G. A. Snook, P. Kao and A. S. Best, J. Power Sources, 2011, 196, 1.

  • 18

    Table S1. Quantitative elemental composition for ATO@MnO2 heterostructures synthesized in the presence of various

    MnCl2 and KBrO3 concentrations. Energy dispersive spectroscopy (EDS) analysis was conducted to evaluate the atomic

    concentration of Mn and Sn, and the ratio of Mn to Sn was calculated by obtained data.

    Synthetic condition Mn (at%) Sn (at%) Mn/Sn

    MnCl2 0.02 M and KBrO3 0.1 M 0.02 7.18 0.00279

    MnCl2 0.02 M and KBrO3 0.2 M 3.14 8.46 0.371

    MnCl2 0.02 M and KBrO3 0.3 M 10.9 5.86 1.87

    MnCl2 0.05 M and KBrO3 0.1 M 8.90 6.11 1.46

    MnCl2 0.05 M and KBrO3 0.2 M 29.5 1.05 28.1

    MnCl2 0.05 M and KBrO3 0.3 M 36.9 -0.09 ―

    MnCl2 0.1 M and KBrO3 0.1 M 22.9 2.94 7.79

    MnCl2 0.1 M and KBrO3 0.2 M 39.0 -0.30 ―

    MnCl2 0.1 M and KBrO3 0.3 M 40.0 -0.38 ―

  • 19

    Table S2. Comparison of the electrochemical performance with various reported MnO2-based electrode materials for

    supercapacitors utilizing Na2SO4 aqueous electrolytes in the three-electrode system.

    Electrode materials

    C

    (mol

    L-1

    )

    Operating

    voltage

    (∆V)

    High rate capability

    Long-term cycle stability

    Ref. Current

    density

    or

    Scan rate

    SC

    Cycles

    Current

    density

    or

    Scan rate

    SC

    Mesoporous MnO2 walnuts 1 1 50 mV s

    -1 ~25 F g

    -1 2,000 834 mA g

    -1 50

    Porous MnO2 1 1 15 A g

    -1

    200 mV s-1

    154 F g-1

    ~130 F g-1

    4,000 2 A g-1

    ~160 F g-1

    52

    K0.15

    MnO2 birnessite

    nanosheets/PEDOT

    0.1 0.9 10 A g-1

    200 mV s-1

    ~ 30 F g-1

    ~ 50 F g-1

    53

    MnO2 1 0.8 100 mV s

    -1 135 F g

    -1 1,000 2 mA cm

    -2 176 F g

    -1 61

    MnO2 powder 0.1 0.9 5 mV s

    -1 150 F g

    -1 47

    α-MnO2 ultralong nanowires 0.5 1 50 A g

    -1

    100 mV s-1

    ~60 F g-1

    137 F g-1

    2,000 50 mV s-1

    ~180 F g-1

    54

    α-MnO2 hollow spheres and

    urchins

    1 1 10 mA g-1

    124 F g-1

    350 5 mA g-1

    ~130 F g-1

    60

    MnO2 1 1 50 mV s

    -1 72 F g

    -1 100 200 mA g

    -1 ~135 F g

    -1 58

    MnO2 nanosheets 1 0.9 100 mV s

    -1 ~25 F g

    -1 1,000 10 mV s

    -1 ~395 F g

    -1 57

    Mesoporous MnO2 nanowire

    array

    0.5 1 12 A g-1

    84 F g-1

    800 4 A g-1

    ~475 F g-1

    56

    α-MnO2 nanotubes 1 0.9 200 mV s

    -1 136 F g

    -1 59

    Mesoporous MnO2 nanotubes 1 1 10 A g

    -1 ~203 F g

    -1 3,000 5 A g

    -1 ~202 F g

    -1 55

    Birnessite-type MnO2 nanowires 0.5 1 100 mV s

    -1 ~85 F g

    -1 62

    α-MnO2 1D nanostructures 1 1 20 mV s

    -1 127 F g

    -1 1,000 20 mV s

    -1

    0.5 mA cm-2

    63

    δ-MnO2 2D nanoplates 180 F g

    -1

    γ-MnO2-structured 3D urchins 100 F g

    -1

    MnO2 1 1 5 mV s

    -1

    200 mA g-1

    168 F g-1

    64

    MnO2 nanorods 0.5 1 100 mV s

    -1 ~60 F g

    -1 49

  • 20

    Mesoporous MnO2 1 1 100 mV s

    -1

    2.3 A g-1

    220 F g-1

    244 F g-1

    1,000 2 A g-1

    ~195 F g-1

    68

    MnO2 hollow spheres 1 0.9 2 A g

    -1 ~88 F g

    -1 71

    α, β, γ, and δ-MnO2 0.1 1 10 mA cm

    -2 ~260 F g

    -1 500 0.5 mA cm

    -2 ~210 F g

    -1 46

    Mesoporous MnO2 1 1 10 mA cm

    -2 202 F g

    -1 500 5 mA cm

    -2 ~205 F g

    -1 69

    MnO2 0.1 1 10 mV s

    -1 2,000 0.5 mA cm

    -2 ~230 F g

    -1 66

    α-MnO2 1 1 10 mA g

    -1 114 F g

    -1 100 500 mA g

    -1 ~131 F g

    -1 72

    Hollow α-MnO2 spheres 1 1 500 10 mV s

    -1 84 F g

    -1 70

    δ-MnO2 1 1 50 mV s

    -1 1,000 5 mV s

    -1 263 F g

    -1 65

    MnO2-pillared layered structures 1 1 50 mV s

    -1 600 10 mV s

    -1 ~215 F g

    -1 67

    MnO2/carbon nanocomposites 1 0.8 20 mV s-1

    2 A g-1

    ~152 F g-1

    100 1 A g-1 ~170 F g-1 76

    ATO@MnO2 hetero-structures 1 1 500 mV s

    -1

    20 A g-1

    178 F g-1

    162 F g-1

    5,000 20 A g-1

    ~152 F g-1

    This

    work

    -----------------------------------------------------------------------------------------------------------------------------------------------------------------

    References

    46. S. Devaraj and N. Munichandraiah, J. Phys. Chem. C, 2008, 112, 4406.

    47. M. Toupin, T. Brousse and D. Bélanger, Chem. Mater., 2004, 16, 3184.

    49. Q. Qu, P. Zhang, B. Wang, Y. Chen, S. Tian, Y. Wu and R. Holze, J. Phys. Chem. C, 2009, 113, 14020.

    50. S. Li, L. Qi, L. Lu and H. Wang, RSC Adv., 2012, 2, 3298.

    52. X. Xie, C. Zhang, M. -B. Wu, Y. Tao, W. Lv and Q. -H. Yang, Chem. Commun., 2013, 49, 11092.

    53. M. Yeager, W. Du, R. Si, D. Su, N. Marinković and X. Teng, J. Phys. Chem. C, 2012, 116, 20173.

    54. W. Li, Q. Liu, Y. Sun, J. Sun, R. Zou, G. Li, X. Hu, G. Song, G. Ma, J. Yang, Z. Chen and J. Hu, J. Mater. Chem., 2012, 22,

    14864.

    55. M. Huang, Y. Zhang, F. Li, L. Zhang, R. S. Ruoff, Z. Wen and Q. Liu, Sci. Rep., 2014, 4, 3878.

    56. C. Xu, Y. Zhao, G. Yang, F. Li and H. Li, Chem. Commun., 2009, 7575.

    57. Z. -P. Feng, G. -R. Li, J. -H. Zhong, Z. -L. Wang, Y. -N. Ou and Y. -X. Tong, Electrochem. Commun., 2009, 11, 706.

    58. V. Subramanian, H. Zhu and B. Wei, J. Power Sources, 2006, 159, 361.

    59. W. Xiao, H. Xia, J. Y. H. Fuh and L. Lu, J. Power Sources, 2009, 193, 935.

    60. M. Xu, L. Kong, W. Zhou and H. Li, J. Phys. Chem. C, 2007, 111, 19141.

    61. R. Jiang, T. Huang, J. Liu, J. Zhuang and A. Yu, Electrochim. Acta, 2009, 54, 3047.

    62. O. A. Vargas, A. Caballero, L. Hernán and J. Morales, J. Power Sources, 2011, 196, 3350.

    63. D. -Y. Sung, I. Y. Kim, T. W. Kim, M. -S. Song and S. -J. Hwang, J. Phys. Chem. C, 2011, 115, 13171.

    64. V. Subramanian, H. Zhu, R. Vajtai, P. M. Ajayan and B. Wei, J. Phys. Chem. B, 2005, 109, 20207.

    65. G. Zhu, H. Li, L. Deng and Z. -H. Liu, Mater. Lett., 2010, 64, 1763.

    66. P. Ragupathy, D. H. Park, G. Campet, H. N. Vasan, S. -J. Hwang, J. -H. Choy and N. Munichandraiah, J. Phys. Chem. C, 2009,

  • 21

    113, 6303.

    67. J. Yuan, Z. -H. Liu, S. Qiao, X. Ma and N. Xu, J. Power Sources, 2009, 189, 1278.

    68. C. Yuan, B. Gao, L. Su and X. Zhang, J. Colloid Interface Sci., 2008, 322, 545.

    69. M. -W. Xu, W. Jia, S. -J. Bao, Z. Su and B. Dong, Electrochim. Acta, 2010, 55, 5117.

    70. X. He, M. Yang, P. Ni, Y. Li and Z. -H. Liu, Colloid Surf. A-Physicochem. Eng. Asp., 2010, 363, 64.

    71. P. Yu, X. Zhang, Y. Chen and Y. Ma, Mater. Lett., 2010, 64, 1480.

    72. N. Tang, X. Tian, C. Yang and Z. Pi, Mater. Res. Bull., 2009, 44, 2062.

    76. Y. Peng, Z. Chen, J. Wen, Q. Xiao, D. Weng, S. He, H. Geng and Y. Lu, Nano Res., 2011, 4, 216.

  • 22

    Figure S1. Cross-sectional SEM images of the ATO nanobelt arrays.

    Figure S2. XRD graphs of γ-MnO2 nanosheets synthesized in the presence of MnCl2 and KBrO3 at different

    concentrations: (a-c) MnCl2 0.02 M, (d-f) MnCl2 0.05 M, (g-i) MnCl2 0.1 M. (a, d, f) KBrO3 0.1 M, (b, e, h) KBrO3 0.2

    M, (c, f, i) KBrO3 0.3 M.

  • 23

    Figure S3. Photographs of the MnCl2 solution and reacted Ti coin cell under various concentrations of MnCl2 and

    KBrO3. Reaction temperature and time was fixed to 50 oC and 6 h, respectively.

    Figure S4. XRD graphs of γ-MnO2 nanosheets prepared in the presence of 0.05 M MnCl2 and 0.2 M KBrO3 with

    different reaction times: (a) 2 h 15 m, (b) 2 h 30 m, (c) 2 h 45 m, (d) 3 h, (e) 3 h 30 m, and (f) 4 h.

  • 24

    Figure S5. (a) XRD graphs and (b) SEM images of γ-MnO2 particles precipitated in the solution of 0.05 M MnCl2 and

    0.2 M KBrO3 with and without insertion of ATO nanobelt/Ti substrate. Reaction temperature and time was fixed to 40

    oC and 63 h, respectively. Scale bar in the inset of b is 500 nm.

    Figure S6. SEM images of nanosheets synthesized in the presence of 0.2 M KBrO3 with different concentrations of

    MnCl2 and reaction temperatures. Reaction time was fixed to 6 h. Scale bar: 1 μm.

  • 25

    Figure S7. Enlarged electrochemical behavior graph of the inset of Figure 6a showing the linearity of the anodic current

    density with scan rates.

    Figure S8. The area-specific capacitance (SCCD,area) of ATO@MnO2 nanostructure electrode plotted as a function of

    discharge current densities.

  • 26

    Figure S9. Comparison of the mass-specific capacitance of ATO@MnO2 nanostructures prepared from the different

    synthetic condition.

  • 27

    Figure S10. Ragone plot derived from the galvanostatic discharge curves for the ATO@MnO2 electrode measured at six

    different current densities from 0.5 to 20 A g-1. Here, the energy density (E, Wh kg-1) and power density (P, kW kg-1) are

    calculated using the following equations [75]: E = (C × ∆V2)/2 and P = E/∆t, where C (F g-1) represents the specific

    capacitance of the active material, ∆V (V) represents the cutoff voltage across the electrode, E represents the energy

    density, and ∆t (s) represents the discharge time, respectively.

    -----------------------------------------------------------------------------------------------------------------------------------------------------------------

    References

    75. H. -W. Shim, A. -H. Lim, J. -C. Kim, E. Jang, S. -D. Seo, G. -H. Lee, T. D. Kim and D. -W. Kim, Sci. Rep., 2013, 3, 2325.

  • 28

    Figure S11. (a) TEM, (b) SAED, and (c) EDS mapping of an ATO@MnO2 nanostructure after 5,000 cycles. Scale bar of c

    is 100 nm.

    Figure S12. Enlarged SEM images of the ATO@MnO2 nanostructure before and after 5,000 cycles.

  • 29

    Figure S13. Electrochemical performance of bare ATO nanobelt electrode. (a) CV behaviors, (b) Corresponding

    mass-specific capacitance as a function of discharge current densities. Inset of (b) shows the galvanostatic discharge

    profile obtained at six different current densities. All measurements were carried out using the three-electrodes system, and

    were recorded on the potential window of 0 to 1.0 V (vs. Ag/AgCl).

    a0581Revised NR_MS & Supporting_2014 0816