Nanostructured Negatrodes for High-Energy - ACS Publications

Feb 27, 2018 - active sites in the previous electrode than in the later electrodes due to its interconnected-type architecture and a high surface area...
2 downloads 7 Views 2MB Size
Subscriber access provided by MT ROYAL COLLEGE

Article

Polycrystalline and Mesoporous 3-D Bi2O3 Nanostructured Negatrodes for High Energy and Power Asymmetric Supercapacitors: Superfast Room-Temperature Direct Wet Chemical Growth Nanasaheb Shinde, Qi Xun Xia, Je Moon Yun, Rajaram Mane, and Kwang Ho Kim ACS Appl. Mater. Interfaces, Just Accepted Manuscript • DOI: 10.1021/acsami.8b00260 • Publication Date (Web): 27 Feb 2018 Downloaded from http://pubs.acs.org on February 27, 2018

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

ACS Applied Materials & Interfaces is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

Polycrystalline and Mesoporous 3-D Bi2O3 Nanostructured Negatrodes for High Energy and Power Asymmetric Supercapacitors: Superfast RoomTemperature Direct Wet Chemical Growth

Nanasaheb M. Shindeab, Qi Xun Xiaab, Je Moon Yunb*, Rajaram S. Maneb,c*, Kwang Ho Kimab*

a

Department of Materials Science and Engineering, Pusan National University, San 30

Jangjeon-dong, Geumjeong-gu, Busan 609-735, Republic of Korea. b

Global Frontier R&D Center for Hybrid Interface Materials, Pusan National University, 30,

Jangjeon-dong, Geumjung-gu, Busan 609-735, Republic of Korea. c

School of Physical Sciences, SRTM, University, Nanded, India.

1 ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 32

ABSTRACT

Superfast (10 min) room-temperature (300 K) chemical synthesis of 3-D polycrystalline and mesoporous bismuth(III) oxide (Bi2O3) nanostructured negatrode (as an abbroviation of negative electrode) materials, viz. coconut-shell, marigold, rose, and honey-nest cross-sections with different surface areas, charge transfer resistances, and electrochemical performances essential for energy storage, harvesting, and even catalysis devices, are directly grown onto Ni-foam without and with poly(ethylene glycol), ethylene glycol, and ammonium fluoride surfactants, respectively. Smaller diffusion lengths, caused by the involvement of irregular crevices, allow electrolyte ions to infiltrate deeply, increasing the utility of inner active sites for the following electrochemical performance. A marigold 3-D Bi2O3 electrode of 58 m2.g–1 surface area demonstrated 447 F.g–1 specific capacitance at 2 A.g–1 and 85% chemical stability even after 5000 redox cycles at 10 A.g–1 in 6 M KOH electrolyte solution, which was higher than that of other morphology negatrode materials. An asymmetric supercapacitor (AS) device assembled with marigold Bi2O3 negatorde and manganese(II) carbonate quantum dots/nickel

hydrogencarbonate–manganese(II)–carbonate

(MnCO3QDs/NiH–Mn–CO3)

positrode corroborates as high as 51 Wh.kg–1 energy at 1500 W.kg–1 power and nearly 81% cycling stability even after 5000 cycles. The obtained results were comparable or superior to the values reported previously for other Bi2O3 morphologies. This AS assembly has glowed as a red light-emitting diode for 20 min, demonstrating the scientific and industrial credentials of developed superfast Bi2O3 nanostructured negatrodes in assembling various energy storage devices.

2 ACS Paragon Plus Environment

Page 3 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

KEYWORDS: Polycrystalline and mesoporous 3D Bi2O3 nanostructures; Growth Mechanism; Negatrodes; Quasifaradaic redox reaction; Superfast room-temperature chemical synthesis; Asymmetric Supercapacitor; Mass production and commercial viability 1. INTRODUCTION Supercapacitors with 3-D nanostructures have attracted considerable attention in developing miniaturized energy storage devices due to their high power and energy, fast recharge capability, and long-cycle life.1,2 In addition to redox states, the performance of supercapacitor electrode materials is morphology, surface area, and crystallinity dependent.3 Nanostructures of transition metal oxides (TMOs)/chalcogenides with different crystallinities and phases have developed a niche place in designing energy storage/harvesting devices. For example, in dye-sensitized solar cells (DSSCs), anatase, but not rutile or brookite, titanium (IV) oxide with high crystallinity and porosity is essential.4-5 Thereby, different research groups are actively engaged in developing different nanostructures, including nanorods,6 nanoflakes,7 needles,8 nanosheets,9 nanowalls,10 nanospheres,11 flowers, etc.,12 which not only increase the electrochemical performance but also helps in understanding the charge transport kinetics involved. As an example, because of the involvement of hopping-free charge transportation, Law et al.13 reported faster charge transportation in nanorods than in nanospheres of ZnO-based DSSCs. The reported synthesis time and sintering temperature were 2.5 h and 92 °C, respectively. Available chemical/physical methods for obtaining TMO nanostructures are either expensive, time-consuming, air-annealing-mediated, or templateassisted.14,

15

Binder-inspired (if the product is a powder) nanostructures require various

annealing steps to remove the binders, otherwise, their presence blocks or covers the pores and surfaces by suppressing the dye-loading capacity, gas molecule adsorption, and electrolyte in electrochemical devices with diminished performance. Chemical methods produce relatively amorphous nanostructures where several annealing steps are required to 3 ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 32

achieve high crystallinity. Second, at any given time, either physical or chemical methods produce only one type of nanostructure under the given experimental conditions whose modification requires additional steps. Third, there is no guarantee that all these nanostructures shall be 3-D and mesoporous because planar nanostructures demonstrate limited surface area and high sheet resistance.16, 17 Moreover, the metal oxide nanostructures, synthesized using chemical methods envisaged in electrochemical supercapacitor (ES) applications

are

time-consuming,

and

complex/binder

and

template-mediated.18-21

Polycrystalline and 3-D hierarchical architectures of TMOS have provided high surface area, strong electrolyte/electrode interfacial contact, and smaller diffusion length for electrolyte ions for obtaining enhanced energy and power storage capacity.22 Other than carbon-based materials, a very few polycrystalline and mesoporous TMOs as negatrode materials, with a simple and fast synthesis methods and remrkable electrochemical storage performance, are reported.23 Because of the low-cost, lack of toxicity, and existence of various phases, bismuth(III) oxide (Bi2O3) is primarily used in dye-sensitized solar cell,24, 25 catalyst,26, 27 gas sensor,28, 29 fuel cell,30 and ES31 applications. As far as ES applications are concerned, Sun et al.32 showed the use of solvothermally prepared 3-D hierarchical Bi2O3 microspheres with 832 F.g–1 specific capacitance (SC) at 1 A.g–1. and still maintained 90% of the initial level at 20 A.g-1. Liu et al.33 successfully proved the deposition of Bi2O3 flowers on worm-like mesoporous carbon with 386 F.g–1 at 250 mA.g–1 SC using a microwave-assisted chemical method in addition to a superior rate capability and 68% cycle stability at 500 cycles . Su et al.34 demonstrated a systematic study of rod-like Bi2O3 with 1350 F.g–1 SC at 0.1 mA.g–1 (only 2.4 % capacitance decline after 1,000 cycles). A composite structure fabricated by electrodeposition of Bi2O3 onto titania nanotube arrays exhibited an aerial SC of 430 mF.cm–2 at 5 mA.cm–2 and 75% cycle stability at 500 cycles.35 Rod-like Bi2O3, obtained by a sol–gel method, showed 528 F.g–1 SC at 5 mV.s–1 and 62.33% rate capability at 1000 cycles.36 Electrochemically deposited Bi2O3 thin films demonstrated electrochemical reversibility with 4 ACS Paragon Plus Environment

Page 5 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

98 F.g−1 SC at 20 mV.s–1.37 Tong’s group prepared hierarchical ripple-type Bi2O3 nanobelts with 250 F.g−1 SC at 100 mV.s–1 and a good electrochemical stability (without providing numerical data).38 Sankpal et al. used a successive ionic layer adsorption and reaction method to obtain needle-like Bi2O3 on a stainless-steel substrate, which produced 329 F.g–1 SC at 100 mV.s–1, 39 when used in an SC application. Recently, we reported a wet chemical synthesis on Ni foam and the SC application of flower-type Bi2O3 with 557 F.g–1 (at 1 mA.cm–2) SC and 85% retention over 2000 cycles,

40

where the synthesis time of several hours was a serious

drawback. In fabricating supercapacitor negatrode (negative)

and positrode (positive) electrode

materials41 for SC application, chemical/physical methods need to be simple and direct, i.e., binder and template-free, annealing-free, not time-consuming, and capable of producing large-scale crystalline and mesoporous nanostructures with good energy storage ability.42 In the present work, Ni foam-supported direct synthesis of polycrystalline and mesoporous 3-D Bi2O3 nanostructured negatrodes (without and with poly(ethylene glycol, PEG), ethylene glycol (EG), and ammonium fluoride (AF)) surfactants using a soft chemical method at room temperature (300 K) in a deposition time of less than 10 min was revealed. These 3-D nanostructured negatrodes are characterized by their crystal structures, morphologies, surface areas, pore-size distributions, and further, are envisaged as electrode materials (half- and fullcell electrode configurations) in ES application studies. 2. EXPERIMENTAL SECTION 2.1 Synthesis 3-D Bi2O3 Nanostructured Negatrodes. 3-D nanostructures, viz. coconutshell, marigold, honey-nest cross-section, and rose of Bi2O3 were synthesized with no surfactant and with poly(ethylene glycol) (PEG), ammonium fluoride (AF), and ethylene glycol (EG) surfactants, at room temperature respectively, (see experimental conditions, reaction mechanisms, and demonstrating images in supporting information S1) on Ni foam, 5 ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

named as A, B, C, and D, respectively. In brief, 0.1 M Bi(NO3)3 (bismuth(III) nitrate) was dissolved in 50 ml double-distilled water and an appropriate amount of 1 M HNO3 (nitric acid) was added to obtain a clear and transparent solution. Aqueous ammonia solution was added to each beaker to maintain the pH  8. Four beakers of the same kind were arranged. The PEG, AF, and EG surfactants were added (1 wt. %) to three beakers, and nothing to the fourth, while stirring the solutions. Ni foam was inserted in each solution using a long crocodile pin that was fixed to the substrate holder. Finally, 500 l hydrochloric acid (HCl) was added to initiate oxidation reaction in each beaker. All reactions were completed in less than 10 min deposition time, which eventually protected oxidation of Ni foam to form NiO@Ni (discussed later). Notably, all samples used for physical as well as electrochemical measurement characterizations were free from the air/inert gas-annealing step. 2.2 Characterizations The structural elucidation and morphological evolution

studies of nanoscale modulations

Bi2O3 were carried out by X-ray diffraction (XRD, D8-Discovery Bruker, 40 kV, 40 mA, Cu Kα, λ = 1.5406 Å) pattern and field emission scanning electron microscopy (FE-SEM, Hitachi, S-4800, 15 kV) equipped with energy dispersive X-ray spectroscopy (EDX) and highresolution transmission electron microscopy (HRTEM, JEOL 2100F) techniques, respectively. The X-ray photoelectron spectroscopy (XPS, VG Scientifics ESCALAB250) measurement was performed to analyze the chemical bonding status of Bi2O3-Ni-F. Micromeritics ASAP 2010 analyzer was used to measure the pore-size using nitrogen absorption at 77 K. The poresize distribution was confirmed from the Barrett–Joyner–Halenda (BJH) plot. Electrochemical measurements Cyclic voltammetry (CV) curves, galvanostatic charge/discharge tests and electrochemical impedance spectroscopy (EIS) were carried out using IVIUM electrochemical workstation system (Ivium, State). All electrochemical measurements were carried out in a

6 ACS Paragon Plus Environment

Page 6 of 32

Page 7 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

conventional three–electrode system with a 6 M KOH aqueous electrolyte at room temperature.

3. RESULT AND DISCUSSION 3.1 Preparation and Reaction Mechanism.

Figure 1. Schematic illustration presenting formation of Bi2O3 nanostructures with their featured morphlogies (digital photographs are taken from google search engine).

The following plausible chemical reactions could be responsible for the formation of Bi2O3 with various morphologies (Figure 1). The source of Bi+3 such as Bi(NO3)3.5H2O was dissolve in HNO3 and H2O with adding NH3 to increase the pH to ~8 of solution (equ. 1). Four experiments were pereferred out of which one was without surfactant and three were with ethylene glycol, poly ethylene glycol and NH4F surfactants. In view of the fact that {O-[CH2CH2-]-OH}-, {O-[-CH2CH2-]n-OH}- and F- could be good promoters to different morphologies which might attached to the Bi3+ for [Bi{O-[-CH2CH2-]-OH}x](x-3)-, [Bi{O-[CH2CH2-]n-OH}x](x-3)- and [BiFx] (x-3)- (equ. 2-4) unstable intermidiate complexes formation, respectivily. At the same time, reactions were processed in same container by attaching OHto Bi3+ and all intermidiate complexes could lost their stability by forming [Bi(OH4)]-, resulting with the formation of Bi2O3 in different morphologies due to change of chair density

7 ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 32

(equ. 5-7). Finally, the H+, present in all precursors, attached to [Bi(OH4)]- by forming Bi2O3, in accordance with equation. Bi(NO3)3.5H2O → Bi+3 + NO3- + H+ + OH-

..............

(1)

Bi+3 + x{HO-[-CH2CH2-]-OH}→ [Bi{O-[-CH2CH2-]-OH}x](x-3)- + H+

..............

(2)

Bi+3 + x{HO-[-CH2CH2-]n-OH}→ [Bi{O-[-CH2CH2-]n-OH}x](x-3)- + H+

..............

(3)

Bi+3 + xNH4F →[BiFx] (x-3)- + N H4+

..............

(4)

[Bi{O-[-CH2CH2-]-OH}x](x-3)- + OH- → [Bi(OH4)]-

..............

(5)

..............

(6)

[BiFx] (x-3)- + OH- → [Bi(OH4)]-

..............

(7)

2[Bi(OH4)]- + 2 H+→ Bi2O3 + 5H2O

..............

(8)

[Bi{O-[-CH2CH2-]n-OH}x](x-3)-

+ OH- → [Bi(OH4)]-

3.2 3-D Bi2O3 Nanostructured Negatrodes 3-D Bi2O3 nanostructured negatrodes obtained through this simple and superfast chemical method were tested with field emission scanning electron microscope (FE–SEM) images (Figure 1A–D) where the morphologies such as coconut-shell, marigold, honey-nest crosssection (from cross-sectional point of view and not from the top-view, which is generally hexagonal), and rose for surfactant-free, PEG, AF, and EG surfactants, respectively, were corroborated, suggesting the importance of surfactants in developing 3-D Bi2O3 morphologies (see supporting information for more details). From Figure 2A, where a1 and a2 are highmagnification FE–SEM images, it was inferred that the obtained morphology was the microsphere-type with diameters ranging from 5 to 8 (2) m. However, under close inspection, the several upright-standing thin nanoplate arrays of various diameters (and an average width of 250 (50) nm) were observed to be stacked to form microspheres of nearly the same size. The diameter of each circular nanoplate, a part of the microsphere, increased from one side to the centre and then decreased from the center to the other side (Figure 2A, 8 ACS Paragon Plus Environment

Page 9 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

a1). The average width of the nanoplatelets in the microsphere was 150 (50) nm and the distance between two nearest nanoplates was 80 (20) nm (Figure 2A, a2). The surfactantmediated nanostructures (Figure 2B, C, and D) of Bi2O3 confirm a complete change in the surface morphology.

Figure 2. (A–D) 50 µm, 10 µm (a1-d1) and 1 µm (a2-d2) bar-scale magnifications FE– SEM images of the Bi2O3 nanostructured negatrodes obtained without and with PEG, AF, and EG surfactants (fall images are converted from original FE-SEM images for better physical comfortability). 9 ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

In the presence of PEG, marigold-like globes composed of interconnected upright-standing nanoplatelets were observed (Figure 2B). These globes are 100 (50) nm in diameter and nanoplatelets are 100 (50) nm in width. A hollow air cavity of 800 (200) nm was identified (Figure 2B, b2). After incorporation of AF in the precursor solution, the FE–SEM image confirms the honey-nest cross-sectional view (Figure 2C, c1, c2). Several rectangular disks of 18 (2)  18 (2)  1.5 (0.2) m3 dimensions reside on the Ni foam. Each disk is composed of three morphologies, viz. upright and downright standing and parallel to the edges of the respective disk. Upright and downright nanoplates show the same heights, i.e., 0.40 (0.05) m. However, their lengths are unclear as they seem to be well connected to one another. Parallel plates, the third morphology, are incredible. They demonstrate parallel as well as inter-branched signatures. Because of the deep insertion, their dimensions, except 40 (10) nm thickness, are unprecedented (Figure 2C, c2). Furthermore, rose-type Bi2O3 of 80 (20) m in size was composed of loosely packed curly nanoplatelet arrays (Figure 2D). These nanoplatelets were merged into one another to form a rose-like architecture. An enlarged image shows 600 (100) nm air spacing separated by 150 (5) nm plate width (Figure 2D, d2). As stated above, FE–SEM results support the controlled evolution of Bi2O3 morphologies by simply adjusting the surfactant additives, where the role of different surfactants in obtaining the Bi2O3 3-D nanostructures has been proposed in SI S2. The chemical compositions of these nanostructures were obtained from the SEM–EDX and corresponding data are given in Figure S2. The XPS survey and de-colvonated spectrums of A, B, C, and D samples qualitatively have supported the formation of Bi2O3 (Figure S3).

10 ACS Paragon Plus Environment

Page 10 of 32

Page 11 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

3.3 Polycrystallinity and Mesoporosity The phase of the Bi2O3 samples was obtained from the X-ray diffraction (XRD) patterns (Figure 3a). On adding different surfactants, the (110) plane gradually weakens and finally, disappears, supporting -Bi2O3.32, 39 The change in crystallinity of Bi2O3 is responsible for cupping behaviour when applied in electrochemical measurements (discussed later). All obtained diffraction peaks were indexed to (110), (210), (201), (220), (212), (222), (203), (421) and (402) diffraction planes, confirming the existence of -Bi2O3 (JCPDS, no. 270052).28, 36, 40 The peaks marked with “* [(111), (200) and (220)]” and ‘♣ (113)’ belonged to metallic Ni and Bi-O-Cl (JCPDS, no. 06-0249), respectively, indicating presence of a trace amount of. Bi-O-Cl in Bi2O3 crystal structure matrix.

Figure 3. (a) XRD patterns, (b) nitrogen adsorption–desorption isotherms, and (c) poresize distribution plots of A–D Bi2O3 negatrodes. Specific surface area and pore-size distribution studies of Bi2O3 nanostructures obtained from the sediment powders were performed by N2-adsorption/desorption isotherms (Figure 3b, A– D), where H3-type hysteresis in the range of ca. 0.6–1.0 P/P0 confirms the involvement of the mesoporous-type character in all nanostructures.32, 40 The BET specific surface areas of A, B, C, and D Bi2O3 3-D nanostructures were found to be 31, 58, 44, and 37 m2.g–1, respectively. The highest surface area was assigned to the B-type nanostructure, i.e., marigold, which is densely populated and free from wide voids. However, average pore-size distributions of A, C, 11 ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

and D nanostructures have supported a mesoporous signature with 2–12 nm pore sizes. The reported BET surface areas of as-obtained polycrystalline and mesoporous Bi2O3 negatrodes are comparable/superior to those reported to similar (67.89,32 29.9,36 2438 m2.g–1) and other (V2O5 (133),43 MoO3 (4.78),44 Fe2O3 (24 m2.g–1)45, ) negatrodes materials. Notably, carbonousbased negatrode materials demonstrated higher surface area i.e. 1000-2000 m2.g–1 2,46 signifies an opening of new era where synthesis of hybrid negatrodes (with either carbon based materials, conducting polymers or TMOs of same signature) of high surface and fast charge transportation followed electrochemical performance than pristine Bi2O3 can be targeted. The average pore-size distribution maxima of A, B, C, and D nanostructures and average pore volumes were 2.53 and 3.89 nm (0.11 cm3/g), 3.20 nm (0.25 cm3/g), 2.8 nm (0.21 cm3/g), and 2.30, 3.48, and 5.16 nm (0.14 cm3/g), respectively. The BET results demonstrate different pore diameters for each nanostructure, which was due to the presence of internal separation between nanoplatelets and/or porosity of agglomerated nanoplatelets. The B electrode of Bi2O3 reveals higher specific surface area and pore volume/diameter (mesoporous), which can find more suitability as an electrode material when used in energy storage applications over A, C, and D Bi2O3 nanostructure electrodes. TEM scans were applied to understand the obtained Bi2O3 nanostructures better (Figure 4(A– D)) where each nanostructure is composed of several polished nanoplatelets with average thickness 15 (±5) nm, which is in good agreement with the results obtained from the FE-SEM images (Figure 2 (A-D)). All samples demonstrate clear lattice fringes of 0.305 nm interplanar spacing, which is in accordance with (210) plane of Bi2O3. The HR-TEM images (Figure 4(a1-d1)) of nanostructures are in accordance with their respective morphologies revealed in FE-SEM surface analysis. It is worth noting that despite the low-intensity XRD peaks, all nanostructures demonstrate considerable crystallinity because they show sharp spots instead of foggy background and circular lines in selected area diffraction patterns (Figure 3(a212 ACS Paragon Plus Environment

Page 12 of 32

Page 13 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

d2)).38, 40 Notably, XRD results showed polycrystalline and SAED images confirmed single crystalline nature of all Bi2O3 nanostructures. Presence of stacking faults, trace amount of BiO-Cl as an impurity, grain boundaries etc., in few millimetre XRD scale dominated previous signature than in few nanometer SAED scale later one, which is one of the common practices. The chemical bonding over surfactant-inspired 3-D Bi2O3 nanostructures was confirmed with the help of A, B, C, and D FT-IR measurements (Figure S4), confirming the critical role of surfactants in obtaining different Bi2O3 nanostructures.

Figure 4. (A-D) TEM, and (a1-d1) HR-TEM images with corrosponding (a2-d2) SEAD pattern of the Bi2O3 samples.

13 ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 32

3.4 Supercapacitive Performance Figure 5 shows the cyclic voltammograms (CVs) of A–D nanostructured electrodes from −1.0 to 0.0 V and at 5–100 mV.s–1 scan rates in an aqueous 6 M KOH electrolyte solution. All CVs exhibited a pseudocapacitance signature with redox peaks. As seen in Figure 5A–D, anodic peaks and cathodic peaks were found in all nanostructures at expected potential locations in the CV curves.32–37 However, because of different redox reaction rates of participation, the electrochemical redox peak positions for each nanostructure are different. The possible mechanism of the oxidation and reduction peaks in CV can be understood from the following equation. 40, 47, 48 Bi2O3  2 H 2O  2e   Bi2O2  2OH 

Bi2 O2  2 H 2O  4e   2 Bi  4OH 

--------- (9)

--------- (10)

A probable reversible energy storage process for the formation of pseudocapacitance Bi2O3 was proposed by considering the following reaction: Bi2O3  3H 2O  6e   2 Bi  6OH 

-------- (11)

It is clear from Figure 5a that the current density response of Bi2O3 3-D mesoporous nanostructured electrodes follows the B >C >D >A trend. Area under the CV curve of Nifoam is negligibly small, indicating an infinitesimal contribution in electrochemical energy storage, when Bi2O3 is loaded on it whose performance, if there is any, was same in all electrodes. Because of the higher surface area, in contrast to other nanostructure electrodes, the CV curves of electrode B (Figure 5B) shows maximum current density and area, whereas A, C, and D demonstrate relatively lower surface areas as well as porosities. With increasing scan rate, the oxidation peak potentials in each nanostructured electrode were shifted toward a positive direction and the reduction peak potentials were to a negative direction, which was 14 ACS Paragon Plus Environment

Page 15 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

mainly attributed to a change in a crystallite IR-drop component that dominates at higher current values.40 Notably, at high scan rate, the cupping-type behavior of the reduction peak dominates, suggesting phase instability of Bi2O3 at higher scan rates.40 The kinetics of the interfacial quasi-Faradaic redox reaction may be rapid enough, as an increase in current response at higher scan rates has already been observed and reported elsewhere.48

Figure 5. (A–D) CVs of Bi2O3 nanostructured negatrodes at different scan rates. (a) comparative CVs of Bi2O3 electrodes scanned at 100 mV.s−1 fixed scan rate.

15 ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 6. (A–D) Charge–discharge curves (CDC) of A–D electrodes at different current densities (2–10 A.g–1). (a) CDC of A–D electrodes at 2 A.g–1. (b) SC values of each electrode (A–D) at different current densities.

Figure 7. (a) Schematic presenting the hierarchical marigold structure (with inset as a photograph of a marigold and extnesion as possible channels enrolled therein). (b) Nyquist plots of A–D electrodes in the 100 kHz ―100 MHz frequency range. (c) Typical CDC profile of D-electrode presenting the kinetics of interfacial quasi quasi-faradaic redox reaction responsible for IR drop.

16 ACS Paragon Plus Environment

Page 16 of 32

Page 17 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

The charge–discharge curves (CDC) of the A, B, C, and D electrodes in 6 M KOH aqueous solution at different current densities and in the –1.0 to 0.0 V voltage window are shown in Figure 6A–D. The CDC are nonsymmetrical with IR drop, mainly due to a pseudocapacitive signature of Bi2O3. Each discharge voltage curve is divided into two regions: a) internal resistance known as steep voltage drop and b) a prolonged voltage plateau, which is due to the involvement of the Faradaic process in Bi2O3.40,

49

With increasing current density, the

prolonged voltage plateau is shifted the same as that of the CV curves, which again is a symptom of fast kinetics for the quasi-Faradaic redox reaction of Bi2O3.34–36 The SC, energy density (ED), and power density (PD) equations given on page S9 of SI were used for calculations. The masses of deposited A, B, C, and D electrodes were 1.24 (0.01), 1.58 (0.01), 1.03 (0.01), and 1.91 (0.01) mg, respectively. From Figure 6a, it was seen that; a) the sleep voltage drop is considerably high and efforts need to be pluged in to resolve this issue for better performance and b) the “B” electrode possesses a significantly longer charging/discharging time than that of the A, C, and D electrodes, illustrating larger capacitance. The SC of 186 F.g−1 (A/coconut-shell), 447 F.g−1 (B/marigold), 281 F.g−1 (C/honey-nest cross-section), and 263 F.g−1 (D/rose) at a current density of 2 A.g−1 are obtained. The SC value of individual nanostructure electrodes decreases with increase in the current density (Figure 6b), which is attributed to the intercalation/deintercalation of electrolyte ions at low current density taking a longer time and can transfer more charge compared with the higher current density, leading to maximum utilization of nanostructureactive electrode material at low current density and smaller SC at higher current density. The higher SC value of the B electrode is due to the interactions between the highly uniformly distributed marigold network (as shown in the FE–SEM image in Figure 2B, b1, b2, the inset includes the photograph and schematics in Figure 7a). The higher SC performance in B electrode was mainly ascribed to the presence of; (a) large number of interwoven nanoplatelet-arrays, providing considerable wider spaces which we called meso/macro17 ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

chanels act as ion-buffering reservoirs, could open doors for easy percorcolation to electrolyte ions, and (b) micro/meso-pore distributions, confirmed from BET analysis, of Bi2O3 nanoplates (micro/meso-channels) could allow electrolyte ions to pass interior for excessive quasi-faradaic redox reactions (Figure 7a), ensuring smaller diffusion lengths for enormous redox reactions at high current densities.49, 50 Figure 7b shows the Nyquist plots of the A, B, C, and D electrodes in the 100 kHz – 100 MHz frequency range, which is one of the popular ways to evaluate the fundamental behavior of electrode material in electrochemical measurements. All nanostructured electrodes confirm a typical semicircle in the highfrequency region and a straight line in the low-frequency region, suggesting capacitive behavior.6-11, 50 The high-frequency semicircle diameter in the Nyquist plots is related to the charge-transfer resistance (Rct). The measured values of Rct for the A, B, C, and D nanostructures are 5.40 (0.20), 0.48 (0.20), 1.31 (0.20), and 3.32 (0.20) , respectively. The value of Rct for the B Bi2O3 electrode is smaller than those for the A, C, and D electrodes, which suggests faster charge transportation and utilization of more active sites in the previous electrode than in the later electrodes due to its interconnected-type architecture and high surface area, respectively. However, CDC data-supoorted Rs values of 400 (± 100) Ω calculated using Vdrop/2I relation were differet to those measured from the EIS spectra (Figure 7c). The plausible reasions for this descripency are as follows; 1) due to a quasireversible redox reaction of Bi2O3, the CDC curve of B (taken as reference) negatrode (soild line) is different than that of triguler symmeteric (dased line) curve i.e. conventional electrochemical supercapacitor, implying supercapacitor mechanism of either under potential deposition, redox pseudocapacitance, or intercalation pseudocapacitance type which is assigned to its irreversible process with sluggish or slow electrolyte ions exchange over electrode surface. Hence, charging and discharge curves became non-symmetric, and 2) due to various issues for example; kinetic barriers for electron transfer, ohmic resistance and ion transport difficulty which would exhibit an oxidation current peak at a more positive potential 18 ACS Paragon Plus Environment

Page 18 of 32

Page 19 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

than that of the reduction current peak, the rate of redox reactions over electrode surface is much less reversible.40, 41, 51, 52 To make the B Bi2O3 electrode available in energy storage device applications, it is essential to check its chemical stability. Hence, the electrochemical stability of the B Bi2O3 electrode evaluated by CDC measurement at a current density of 10 A.g–1 has been provided in SI S5, where 85% capacitance, even after 5000 cycles, was retained. Finally, in addition to this, a chart of the results obtained in the present work compared with previously reported Bi2O3 nanostructures synthesized using time-consuming chemical/physical methods is listed in Table S2. The 3-D polycrystalline Bi2O3 nanostructures obtained in the present work have demonstrated excellent performance compared with previously reported nanostructures showing the importance of the provided method and the potential of developed nanostructures in electrochemical energy storage devices. 3.5 Asymmetric Supercapacitor A complete asymmetric supercapacitor (AS) energy storage device, using Bi2O3 (B) as a negatrode and MnCO3QDs/NiH–Mn–CO3 as cathode, was assembled and investigated. Electrochemical properties of the MnCO3QDs/NiH–Mn–CO3 electrode were reported previously,8 where as high as 2641.3 F.g–1 SC at 3 A.g–1 with 91.3% stability after 10,000 cycles was obtained. Figure 8a shows the comparative CV curves of the Bi2O3 and MnCO3QDs/NiH–Mn–CO3 at a scan rate of 100 mV.s–1. The operating cell voltage greatly increases when both electrodes are assembled into the AS device. The CV curves of the Bi2O3//MnCO3QDs/NiH–Mn–CO3 AS device in the 0.8―1.5 V range at 100 mV.s–1 scan rate are shown in Figure 8b. The voltage window of the AS device was dramatically extended to 1.5 V, which proves the high ED of the assembled AS. Figure 8c presents the CV curves of the Bi2O3//MnCO3QDs/NiH–Mn–CO3 AS at scan rates from 5 to 100 mV.s–1 between 0 and 1.5 V, where the shapes of the CV curves are not close to the ideal shape due to the 19 ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

pseudocapacitance-type of assembled AS.10 In addition, the CDC scanned at different current densities are shown in Figure 8d. The SCs calculated from the CDC are shown in Figure 7d, which decreases from 164 to 73 F.g–1 as the current density increases from 2 to 10 A.g–1. Figure 8e is the Ragone plot, evidenced by various ED and PD values, for the Bi2O3//MnCO3QDs/NiH–Mn–CO3 AS device, confirming as high as 51 Wh.kg–1 ED at 1500 W.kg–1 PD.

Figure 8. (a) Comparison CV curves of Bi2O3 (electrode B) and Bi2O3//MnCO3QDs/NiH– Mn–CO3 electrodes at a scan rate of 100 mV.s–1. (b) CV curves for different voltages (0.8–1.5 V) at a fixed scan rate of 100 mV.s–1. (c) CV curves at different scan rates (5– 100 mV.s–1) at 1.5 V fixed voltage. (d) CDC curves at different current densities in the 0– 1.5 V voltage range. (e) SC at different current densities, and (f) Ragone plots of the Bi2O3//MnCO3QDs/NiH–Mn–CO3 AS presenting performance better than other literature reports. Compared with similar systems (and also with another negative quality of AS devices reported previously, whose details can be obtained from SI Table S3),53–59 the present work displays significantly higher ED and PD values. The cycle stability of the B20 ACS Paragon Plus Environment

Page 20 of 32

Page 21 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

Bi2O3//MnCO3QDs/NiH–Mn–CO3 AS device assembly was tested by performing continuous charge–discharge cycles at a constant discharge current density of 10 mA.cm–2 for 5000 cycles (Figure 9). The SC of the AS was maintained at the 81% level of its original value even after 5000 cycling tests (inset presents the plots for the 1st and 5000th cycles). The decrease of the capacitance with cycle number is attributed to structural alternation/disintegration around the active sites, by virtue of repeated ingression and depletion of ions in the electrode materials, during the Faradaic reactions.24 The inset right image provides evidence for the practical potential of the B-Bi2O3//MnCO3QDs/NiH–Mn–CO3 AS device where a red lightemitting diode (LED) was ON for 20 minutes with an intensity that was sufficiently high to be observed in a dark room.

Figure 9. The cycle stability of the B-Bi2O3//MnCO3QDs/NiH–Mn–CO3 AS device; Inset (left) shows the curves at the 1st and 5000th cycles and (right) provides a photograph of the lit red LED from Bi2O3//MnCO3QDs/NiH–Mn–CO3.

21 ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

4. CONCLUSION In conclusion, incredible polycrystalline and mesoporous 3-D Bi2O3 nanostructured negatrodes viz. coconut-shell, marigold, honey-nest cross-section and rose are directly grown onto a Ni-foam in presence of various surfactants using room-temperature wet chemical method in less than ten minute time duration.

Role of surfactants in evolving various

morphologies of Bi2O3 electrodes is clearly elaborated and can be applied for other metal oxide/chalcogenide nanostructures too. Nevertheless, all nanostructures confirm air-crevices with different dimensions. Marigold Bi2O3 electrode i.e. B demonstrates higher surface area and interconnected architecture due to which excess faradaic reactions with fast charge transportation rate can be obtained. This leads to produce 447 F.g-1 specific capacitance at a 2 A.g-1 current density and 85% of the capacitive stability after 5000 cycles, which is better than other morphology electrodes. Asymmetric supercapacitor of B-Bi2O3//MnCO3QDs/NiH–Mn– CO3 configuration operated at 1.5 V delivers as high as 164 F. g-1 SC and 51 Wh.kg-1 ED at 1500 W.kg-1 PD, which is better/comparable to results obtained in other studies reported previously. Biogenic, photocatalysis and glucose sensing60 applications of these nanostructure will be fascinating, and are undre progress. Efforts are also in progress to prepare hybrid negatrodes by combining these polycrystalline and mesoporous 3-D Bi2O3 nanostructures with; a) high surface area sulfide-carbon-based materials for sodium storage devices,42, 61 and b) several redox active TMOs with electronically conducting polymers for enhnacing reversible electron transfer in a wide potential range as high energy and power density electrochemical supercapacitors.

22 ACS Paragon Plus Environment

Page 22 of 32

Page 23 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

ASSOCIATED CONTENT Supporting Information Experimenta, elemental mapping (EDX and XPS) and FTIR details, fomulae used (half and full supercapacitor cells), chemical stability, performance comparative charts (half and full cells) etc., of polycrsyatlline and mesoporous 3-D Bi2O3 negatrodes. This material is available free of charge via the Internet at http://pubs.acs.org. AUTHOR INFORMATION CORRESPONDING AUTHOR: [email protected] (Je Moon Yun, Dr.), [email protected] (R.S. Mane, Prof.) and [email protected] (Kwang Ho Kim, Prof.) ACKNOWLEDGMENTS This study was supported by; a) The Global Frontier Program through the Global Frontier Hybrid Interface Materials (GFHIM) of the National Research Foundation of Korea (NRF) funded by the Ministry of Science, ICT & Future Planning (2013 M3A6B1078874),

and

b)

National

Core

Research

Centre

(NCRC)

grant

2015M3A6B1065262. Authors indebted to Ms. Seonghee Jeong, Sr. FE-SEM operator Technician, for skilful FESEM operations, and Dr. Pritam Shinde, SRTM University, Nanded, India, for scientific discussions followed reaction mechanism proposal.

23 ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

REFERENCES AND NOTES 1. Conway, B., Electrochemical Supercapacitors: Scientific Fundamentals and Technological Applications, Kluwer Academic/Plenum Publishers, New York, 1999. 2. Liu. M.; Kong, L.; Kang, L.; Li, X.; Walsh, F.; Xing, M.; Lu, C.; Ma, X.; Luo, Y. Synthesis and Characterization of M3V2O8 (M = Ni or Co) Based Nanostructures: A New Family of High Performance Pseudocapacitive Materials, J. Mater. Chem. A, 2014, 2, 4919–4926. 3. Wang, G.; Zhang, L.; Zhang, J. A Review of Electrode Materials for Electrochemical Supercapacitors, Chem. Soc. Rev., 2012, 41, 797–828. 4. Shaikh, S.; Mane, R.; Min, B.; Hwang, Y.; Joo, O. D-Sorbitol-Induced Phase Control of TiO2 Nanoparticles and Its Application for Dye-Sensitized Solar cells, Sci. Rep., 2016, 6, 1–6. 5. Kusumawati, Y.; Hosni, M.; Martoprawiro, M.; Cassaignon, S.; Pauporte, Th. Charge Transport and Recombination in TiO2 Brookite-Based Photoelectrodes, J. Phys. Chem. C, 2014, 118, 23459−23467. 6. Xu, W.; Jianga, Z.; Yanga, Q.; Huoa, W.; Javeda, Muhammad S.; Lia, Y.; Li, H.; Xia, G.; Hua, C. Approaching the Lithium-Manganese Oxides Energy Storage Limit with Li2Mn- O3 Nanorods for High-performance Supercapacitor, Nano Energy, 2018, 43, 168–176 7. Li, F.; Chen, H.; Liu, X.; Zhu, S.; Jia, J.; Xu, C.; Xu, H.; Dong, F. Low-Cost HighPerformance Asymmetric Supercapacitors Based on Co2AlO4@MnO2 Nanosheets and Fe3O4 Nanoflakes, J. Mater. Chem. A, 2016, 4, 2096–2104. 24 ACS Paragon Plus Environment

Page 24 of 32

Page 25 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

8. Xia, Q.; Hui, K.; Hui, K.; Kim, S.; Lim, J.; Choi, S.; Zhang, L.; Mane, R.; Yun, J.; Kim, K. Facile Synthesis of Manganese Carbonate Quantumdots/Ni(HCO3)2–MnCO3 Composites as Advanced Cathode Materials for High Energy Density Asymmetric Supercapacitors, J. Mater. Chem. A, 2015, 3, 22102–22117. 9. Xia, Q.; Hui, K.; Hui, K.; Hwang, D.; Lee, S.; Zhou, W.; Cho,Y.; Kwon, S.; Wang, Q.; Son, Y. A Facile Synthesis Method of Hierarchically Porous NiO Nanosheets, Materials Lett., 2012, 69, 69–71. 10. Guan, C.; Li, X.; Wang, Z.; Cao, X.; Soci, C.; Zhang, H.; Fan, H. Nanoporous Walls On Macroporous Foam: Rational Design of Electrodes to Push Areal Pseudocapacitance, Adv. Mater., 2012, 24, 4186–4190. 11. Wang, Q.; Jiao, L.; Du, H.; Si.; Wang, Y.; Yuan, H. Co3S4 Hollow Nanospheres Grown On Graphene as Advanced Electrode Materials for Supercapacitors, J. Mater. Chem., 2012, 22, 21387–21391. 12. Prevot, V.; N, Caperaa.; Taviot-Guého, C.; Forano, C. Glycine-Assisted Hydrothermal Synthesis of NiAl-Layered Double Hydroxide Nanostructures, Cryst. Growth Des., 2009, 9, 3646–3654. 13. Law, M.; Greenel, L.; Johnson, J.; Saykally, R.; Yang, P. Nanowire Dye-Sensitized solar cells, Nat. Mater., 2005, 4, 455–459. 14. Bang, J.; Suslick, K. Applications of Ultrasound to The Synthesis of Nanostructured Materials, Adv. Mater., 2010, 22, 1039–1059. 15. Xu, S.; Wang, Z. One-dimensional ZnO Nanostructures: Solution Growth and Functional Properties, Nano Res., 2011, 4, 1013–1098.

25 ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

16. Mane, R.; Lokhande, C. Chemical Deposition Method for Metal Chalcogenide Thin Films, Mater. Chem. Phys., 2000, 65, 1–31. 17. Xu, W.; Jiang, Z.; Yang, Q.; Huo, W.; Javed, M.; Li, Y.; Huang, L.; Gu, X.; Hu, C. Approaching The Lithium-Manganese Oxides' Energy Storage Limit With Li2MnO3 Nanorods for High-Performance Supercapacitor, Nano Energy, 2018, 43, 168–176.

18. He, Y..; Li. G.; Wang, Z.; Su, C.; Tong, Y. Single-crystal ZnO Nanorod/Amorphous and Nanoporous Metal Oxide Shell Composites: Controllable Electrochemical Synthesis and Enhanced Supercapacitor Performances, Energy Environ. Sci., 2011, 4, 1288–1292. 19. Su, Z.; Yang, Y.; Xie, B.; Lin, Z.; Zhang, Z.; Liu, J.; Li, B.; Kang, F.; Wong, C. Scalable Fabrication of MnO2 Nanostructure Deposited on Free-Standing Ni Nanocone Arrays for Ultrathin, Flexible, High-performance Micro-supercapacitor, Energy Environ. Sci., 2014, 7, 2652–2659. 20. Ji, J.; Zhang, L.; Ji, H.; Li, Y.; Zhao, X.; Bai, X.; Fan, X.; Zhang, F.; Ruoff, R. Nanoporous Ni(OH)2 Thin Film on 3D Ultrathin-Graphite Foam for Asymmetric Supercapacitor, ACS Nano, 2013, 7, 6237–6243. 21. Li, J.; Zhao, W.; Huang, F.; Manivannanc, A.; Wu, N. Single-Crystalline Ni(OH)2 and NiO Nanoplatelet Arrays as Supercapacitor Electrodes, Nanoscale, 2011, 3, 5103–5109. 22. Mu, J.; Chen, B.; Guo, Z.; Zhang, M.; Zhang, Z.; Zhang, P.; Shao, C.; Liu, Y. Highly Dispersed Fe3O4 Nanosheets On One-Dimensional Carbon Nanofibers: Synthesis, Formation Mechanism, and Electrochemical Performance as Supercapacitor Electrode Materials, Nanoscale, 2011, 3, 5034–5040.

26 ACS Paragon Plus Environment

Page 26 of 32

Page 27 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

23. Xia, X.; Zhang, Y.; Chao, D.; Guan, C.; Zhang, Y.; Li, L.; Ge, X.; Bacho, I.; Tu, J.; Fan, H. Solution Synthesis Of Metal Oxides for Electrochemical Energy Storage Applications, Nanoscale, 2014, 6, 5008–5048. 24. Shaikh, S.; Rahman, G.; Mane, R.; Joo, O. Bismuth Oxide Nanoplates-Based Efficient DSSCs: Influence of ZnO Surface Passivation layer, Electrochim. Acta, 2013, 111, 593–600. 25. Qin, F.; Zhao, H.; Li, G.; Yang, H.; Li, J.; Wang, R.; Liu, Y.; Hu, J.; H, Sund.; Chen, R. Size-Tunable Fabrication of Multifunctional Bi2O3 Porous Nanospheres for Photocatalysis, Bacteria Inactivation and Template-synthesis, Nanoscale, 2014, 6, 5402–5409. 26. Zhou, L.; Wang, W.; Xu, H.; Sun, S.; Shang, M. Bi2O3 Hierarchical Nanostructures: Controllable Synthesis, Growth Mechanism, and Their Application in Photocatalysis. Chem.– Eur. J., 2009, 15, 1776–1782. 27. Zhu, S.; Lu, Lilin.; Zhao, Z.; Wang, T.; Liu, X.; Zhang, H.; Don, F.; Zhan, Y. Mesoporous Ni-Doped δ-Bi2O3 Microspheres for Enhanced Solar-Driven Photocatalysis: A Combined Experimental and Theoretical Investigation, J. Phys. Chem. C 2017, 121, 9394−9401. 28. Bhande, S.; Mane, R.; Ghule, A; Han, S. A Bismuth Oxide Nanoplate-Based Carbon Dioxide Gas Sensor, Scripta Materialia, 2011, 65, 1081–1084. 29. Cabot, A.; Marsal, A.; Arbiol, J.; Morante, J. Bi2O3 as a Selective Sensing Material for NO Detection, sensors and Actuators B, 2004, 99, 74–89. 30. Lee, K.; Lidie, A.; Jeon, S.; Hitz, G.; Song, S.; Wachsman, E. Highly Functional NanoScale Stabilized Bismuth Oxides Via Reverse Strike Co-Precipitation for Solid Oxide Fuel cells, J. Mater. Chem. A, 2013, 1, 6199–6207.

27 ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

31. Ganesh, T.; Ham, D.; Chang, J.; Cai, G.; Kil, B.; Min, S.; Mane, R.; Han, S. pH Dependent Morphological Evolution of Bi2O3/PANI Composite for Supercapacitor Applications, J. Nanosci. Nanotechnol., 2011, 11, 589–592. 32. Sun, J.; Wang, J.; Li, Z.; Yang, Z.; Yang, S. Controllable Synthesis of 3D Hierarchical Bismuth Compounds With Good Electrochemical Performance for Advanced Energy Storage Devices, RSC Adv., 2015, 5, 51773–51778. 33. Xia, N.; Yuan, D.; Zhou, T.; Chen, J.; Mo, S.; Liu, Y. Microwave Synthesis and Electrochemical Characterization of Mesoporous Carbon@Bi2O3 Composites, Mater. Res. Bull., 2011, 46, 687–691. 34. Su, H.; Cao, S.; Xia, N.; Huang, X.; Yan, J.; Liang, Q.; Yuan, D. Controllable growth of Bi2O3 With Rod-Like Structures Via The Surfactants and Its Electrochemical Properties. J. Appl. Electrochem., 2014, 44, 735–740. 35. Sarma, B.; Jurovitzki, A.; Smith, Y.; Mohanty, S.; Misra, M. Redox-Induced Enhancement in Interfacial Capacitance of the Titania Nanotube/bismuth Oxide Composite Electrode, ACS Appl. Mater. Interfaces, 2013, 5,1688–1697. 36. Huang, X.; Zhang, W.; Tan, Y.; Wu, J.; Gao, Y.; Tang, B. Facile Synthes is of Rod-like Bi2O3 Nanoparticles as an Electrodematerial for Pseudocapacitors, Ceram. Int., 2016, 42, 2099–2105. 37. Gujar, T.; Shinde, V.; Lokhande, C.; Han, S. Electrosynthesis of Bi2O3 Thin Films and Their Use in Electrochemical Supercapacitors, J. Power Sources, 2006, 161, 1479–1485. 38. Zheng, F.; Li, G.; Ou, Y.; Wang, Z.; Su, C.; Tong, Y. Synthesis of Hierarchical Rippled Bi2O3 Nanobelts for Supercapacitor Applications, Chem. Commun., 2010, 46, 5021–5023.

28 ACS Paragon Plus Environment

Page 28 of 32

Page 29 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

39. Raut, S.; Bisen, O.; Sankapal, B. Synthesis of Interconnected Needle-Like Bi2O3 Using Successive Ionic Layer Adsorption and Reaction Towards Supercapacitor Application, Ionics, 2017, 23, 1831–1837. 40. Shinde, N.; Xia, Q.; Yun, J.; Singh, S.; Mane, R.; Kim, K. A Binder-Free Wet Chemical Synthesis Approach to Decorate Nanoflowers of Bismuth Oxide on Ni-foam for Fabricating Laboratory Scale Potential Pencil-type Asymmetric Supercapacitor Device, Dalton Trans., 2017, 46, 6601–6611. 41. Chen, G. Supercapacitor and Supercapattery as Emerging Electrochemical Energy Stores, Int. Mater. Rev., 2017, 4, 173-202. 42. Peng, G.; Cao, X.; Hou, H.; Li, S.; Ji, X. Rodlike Sb2Se3 Wrapped with Carbon: The Exploring of Electrochemical Properties in Sodium-Ion Batteries, ACS Appl. Mater. Interfaces, 2017, 9, 34979−34989. 43. J, Zhu.; Cao, L.; Wu,Y.; Gong, Y.; Liu, Z.; Hoster, H.; Zhang, Y.; Zhang, S.; Yang, S.; Yan, Q.; Ajayan, P.; Vajtai, R. Building 3D Structures of Vanadium Pentoxide Nanosheets and Application as lectrodes in supercapacitors, Nano Lett., 2013, 13, 5408−5413. 44. Maheswari, N.; Muralidharan, G. Controlled Synthesis of Nanostructured Molybdenum Oxide Electrodes for High Performance Supercapacitor Device, Appl. Surf. Sci., 2017, 416, 461-469. 45. Wanga, H.; Xua, Z.; Yia, H.; Weib, H.; Guob, Z.; Wang, X. One-step Preparation of Single-crystalline Fe2O3 Particles/graphene Composite Hydrogels as High Performance Anode Materials for Supercapacitors, Nano Energy, 2014, 7, 86–96. 46. Zhang, H.; Gao, G.; Yang, Y. Carbon Nanotube Arrays and Their Composites for Electrochemical Capacitors and Lithium-ion Batteries, Energy Environ. Sci., 2009, 2, 932-943. 29 ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

47. Chitrada, K.; Raja, K. Nanoporous Anodic Bismuth Oxide for Electrochemical Energy Storage, ECS Trans., 2014, 61, 55–67. 48. Zuo, W.; Zhu, W.; Zhao, D.; Sun, Y.; Li, Y.; Liu, J.; Lou, X. Bismuth Oxide: a Versatile High-Capacity Electrode Material for Rechargeable Aqueous Metal-ion Batteries, Energy Environ. Sci., 2016, 9, 2881–2891. 49. Khan, Z.; Bhattu, S.; Haram, S.; Khushalani, D. SWCNT/BiVO4 Composites As Anode Materials for Supercapacitor Application, RSC Adv., 2014, 4, 17378–17381. 50. Lu, Y.; Yan, H.; Qiu, K.; Cheng, J.; Wang, W.; Liu, X.; Tang, C.; Kim, J.; Luo, Y. Hierarchical Porous CuO Nanostructures with Tunable Properties For High Performance Supercapacitors, RSC Adv., 2015, 5, 10773–10781. 51. George, C. Understanding Supercapacitors Based on Nano-hybrid Materials with Interfacial Conjugation, Prog. Nat. Sci., 2013, 23, 245–255. 52. Li, G.; Yu, L.; George, C. Capacitive and Non-Capacitive Faradaic Charge Storage, Electrochim. Acta, 2016, 206, 464–478G, 53. Senthilkumar, S.; Kalai Selvan, R.; Ulaganathan, M.; Melo, J. Fabrication of Bi 2O3//AC Asymmetric Supercapacitor with Redox Additive Aqueous Electrolyte and Its Improved Electrochemical Performances, Electrochimica Acta, 2014, 115, 518–524. 54. Xu, H.; Hu, X.; Yang, H.; Sun, Y.; Hu, C.; Huang, Y. Flexible Asymmetric MicroSupercapacitors Based on Bi2O3 and MnO2 Nanoflowers: Larger Areal Mass Promises Higher Energy Density, Adv. Energy Mater., 2015, 5, 1–7. 55. Li, L.; Zhang, X.; Zhang, Z.; Zhang, M.; Cong, L.; Pan, Y.; Lin, S. A Bismuth Oxide Nanosheet-Coated Electrospun Carbon Nanofiber Film: A Free-Standing Negative Electrode for Flexible Asymmetric Supercapacitors, J. Mater. Chem. A, 2016, 4, 16635–16644. 30 ACS Paragon Plus Environment

Page 30 of 32

Page 31 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

56. Wu, Z.; Ren, W.; Wang. D.; Li. F.; Liu. B.; Cheng. H. High-Energy MnO2 Nanowire/Graphene and Graphene Asymmetric Electrochemical Capacitors, ACS Nano, 2010, 4, 5835–5842. 57. Jin, W.; Cao, G.; Sun, J. Hybrid Supercapacitor Based on MnO2 and Columned FeOOH Using Li2SO4 Electrolyte Solution, J. Power Sources, 2008, 175, 686–691. 58. Barzegar, F.; Bello, A.; Momodu, D.; Dangbegnon, J.; Taghizadeh, F.; Madito, M.; Masikhwa, T.; Manyala, N. Asymmetric Supercapacitor Based on an α-MoO3 Cathode and Porous Activated Carbon Anode Materials, RSC Adv., 2015, 5, 37462–37468. 59. Qua, Q.; Shi, Y.; Li, L.; Guo, W.; Wu, Y.; Zhang, H.; Guan, S.; Holze, R. V 2O5·0.6 H2O Nanoribbons as Cathode Material for Asymmetric Supercapacitor in K2SO4 Solution, Electrochem. Commun., 2009,11, 1325–1328.

60. Weina, X.;

Liu,

G.;

Wang,

C.;

Chenguo,

Hu.;

Xue,

W. Novel

β -MnO2

Micro/Nanorod Arrays Directly Grown on Flexible, Carbon Fiber Fabric For HighPerformance Enzymeless Glucose Sensing, Electrochim. Acta., 2017, 225, 121–128. 61. Ge, P.; Hou, H.; Ji, X.; Huang, Z.; Li, S.; Huang, L. Enhanced Stability of Sodium Storage Exhibited by Carbon Coated Sb2S3 Hollow Spheres, Mater. Chem. Phys., 2018, 203, 185– 192.

31 ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 32 of 32

Table of Contents Graphic    

 

ACS Paragon Plus Environment