You are on page 1of 39

Journal of Power Sources

Modulating the Electrochemical Performance of MnO2 Nanowires by Na+ ions Pre-


intercalation For Next Generation Energy Storage Devices
--Manuscript Draft--

Manuscript Number: POWER-D-23-01390

Article Type: Research Paper

Section/Category: Supercapacitors: SCITECH

Keywords: MnO2; Ultrathin nanowires; Na; pre-pre-intercalation; Electrochemical performance.

Corresponding Author: Muhammad Farooq Warsi, Ph.D


THE ISLAMIA UNIVERSITY BAHAWALPUR
Bahawalpur, PAKISTAN

First Author: Muhammad Usman Khalid

Order of Authors: Muhammad Usman Khalid

Muhammad Farooq Warsi, Ph.D

Muhammad Shahid

Sonia Zulfiqar

Manuscript Region of Origin: PAKISTAN

Abstract: Supercapacitors are state of the art technology to overcome the energy crisis globally,
owing to their fast charging/discharging rates and higher power density. One
dimensional morphology (nanorods, nanowires etc.) boosts the inherent low
conductivity of transition metal oxides including MnO2 by confining charge transport
only in one direction. Here we have pre-intercalated Na+ ions into MnO2 nanowires as
conductivity booster in conjunction with α-MnO2 tunnel stabilizing agent. Morphological
analysis reveals sub 50 nm nanowires whose surface gets cracked with Na+ pre-
intercalation, offering less dead area. LSV results reveal an increase in oxygen
evolution overpotential by Na pre-intercalation which can enable supercapacitor to
operate at an extended potential window. Na pre-intercalation and control on
morphology not only increase the conductivity but also shield the electrode
pulverization against tedious charging/discharging cycles and reduce the electrolyte
diffusion pathway. These features enable Na0.10MnO2 NWs to exhibit a specific
capacitance of 1046 Fg-1 @ 1Ag-1, excellent rate capability of 86% at 9Ag-1 along
with 96.9% capacitance retention after 6000 charging-discharging cycles at 12 Ag-1.
This study shows that Na pre-intercalation in MnO2 can improve electrochemical
performance and opens up new horizon to manufacture high-performance next-
generation supercapacitors.

Suggested Reviewers: Aamir Rasheed, PhD


aamir786@skku.edu

Imran Shakir, PhD


Professor
mshakir@ucla.edu

Powered by Editorial Manager® and ProduXion Manager® from Aries Systems Corporation
Manuscript

The Islamia University of Bahawalpur


INSTITUTE OF CHEMISTRY

April 22, 2023


The Editor

Subject: Submission of Full Length Article

Dear Editor, we would like to submit a full-length article entitled “Modulating the electrochemical

performance of MnO2 Nanowires by Na+ ions pre-intercalation for next generation energy storage

devices” in your journal. We herein confirm that the present manuscript and its contents have not been

published previously in some other forms or by any other authors and are not under consideration for

publication in another journal at the time of submission. The article is original and has been written by the

stated authors, who are all aware of its contents and approve its submission. If accepted, the article will not

be published elsewhere in any form.

Corresponding author

Muhammad Farooq Warsi, PhD / Associate Professor

==========================================================
Responses to Technical Check Results

Response to Technical Results


Journal of Power Sources
Manuscript No: POWER-D-23-01390

Comment 1: The figures are not built properly in a manuscript file. Kindly revise your
submission
Response to Comment 1: We have resolved the mentioned issue regarding Figures insertion in
main manuscript body.
Graphical Abstract
Manuscript Click here to view linked References

1
2
3
4 Modulating the Electrochemical Performance of MnO2 Nanowires by Na+ ions
5
6 Pre-intercalation For Next Generation Energy Storage Devices
7
8
9 Muhammad Usman Khalida, Muhammad Farooq Warsi*a, Muhammad Shahidb,
10
11 Sonia Zulfiqarc,d*
12
a
13 Institute of Chemistry, Baghdad-ul-Jadeed Campus, The Islamia University of Bahawalpur,
14
15 63100, Pakistan
16
17 b
18
Department of Chemistry, College of Science, University of Hafr Al Batin, P.O. Box 1803, Hafr
19 Al Batin, Saudi Arabia
20
21
c
22 Department of Chemistry, Faculty of Science, University of Ostrava, 30. Dubna 22, Ostrava,
23 701 03, Czech Republic
24
25 d
26
Department of Chemical and Biological Engineering, Iowa State University, Sweeney Hall, 618
27 Bissell Road, Ames, Iowa 50011, United States
28
29 Corresponding author: farooq.warsi@iub.edu.pk, zulfiqar@istate.edu
30
31
32
33
34
Abstract
35
36
37 Supercapacitors are state of the art technology to overcome the energy crisis globally, owing to
38
39 their fast charging/discharging rates and higher power density. One dimensional morphology
40
41
42
(nanorods, nanowires etc.) boosts the inherent low conductivity of transition metal oxides
43
44 including MnO2 by confining charge transport only in one direction. Here we have pre-intercalated
45
46 Na+ ions into MnO2 nanowires as conductivity booster in conjunction with α-MnO2 tunnel
47
48
49 stabilizing agent. Morphological analysis reveals sub 50 nm nanowires whose surface gets cracked
50
51 with Na+ pre-intercalation, offering less dead area. LSV results reveal an increase in oxygen
52
53
54 evolution overpotential by Na pre-intercalation which can enable supercapacitor to operate at an
55
56 extended potential window. Na pre-intercalation and control on morphology not only increase the
57
58
59 conductivity but also shield the electrode pulverization against tedious charging/discharging cycles
60
61
62 1
63
64
65
1
2
3
4 and reduce the electrolyte diffusion pathway. These features enable Na0.10MnO2 NWs to exhibit a
5
6
7 specific capacitance of 1046 Fg-1 @ 1Ag-1, excellent rate capability of 86% at 9Ag-1 along with
8
9 96.9% capacitance retention after 6000 charging-discharging cycles at 12 Ag-1. This study shows
10
11
12 that Na pre-intercalation in MnO2 can improve electrochemical performance and opens up new
13
14 horizon to manufacture high-performance next-generation supercapacitors.
15
16
17 Keywords: MnO2; Ultrathin nanowires; Na; pre-pre-intercalation; Electrochemical Performance.
18
19
20 1. Introduction
21
22
23
24 Supercapacitors (SCs) are state of the art technology to overcome the energy crisis around the
25
26 globe, as a trade-off of their fast charging/discharging rates and longer cyclic life span [1, 2].
27
28 Specifically, their high power density enables them to run hybrid electric vehicles and power up
29
30
31 the off grid homes. Yet, their low energy density is a key impediment to their widespread practical
32
33 deployment. To take the advantage of SCs' high power density for continuous power supply,
34
35
36 enhancing their energy density is critical to the advancement of high-performance supercapacitors.
37
38 Since, the energy density is directly related to the capacitance and square of potential window by
39
40 1
41 the equation E = 2CV2. Thus, a supercapacitor with high energy density can be established either
42
43
44 by improving the capacitance of the electrode material [3, 4] or by extending potential window [5,
45
46 6]. Capacitance can be manipulated by the faradaic reaction at surface or near the surface. Better
47
48
49
control over morphology of the electrode material can provide basis for efficient faradaic redox
50
51 reaction at surface as well as near the surface. One dimensional (1D) nanostructures, particularly
52
53 nanowires of sub (50 nm) diameter offer the advantages that can substantially meet expectations.
54
55
56 Nanowires under 50 nm diameter offer; (1) rapid ambling of electrolyte ions (2) more exposed
57
58 sites due to deep percolation of electrolyte ions [7, 8], resulting in an increased contact area
59
60
61
62 2
63
64
65
1
2
3
4 between electrode and electrolyte, that decreases charging/discharging time [9, 10]. Additionally,
5
6
7 nanowires can manage the volume expansion, impeding degradation, and therefore prolong cyclic
8
9 life span [11]. Nanowires also feature high aspect ratio and as well as reduced dimensions,
10
11
12 permitting charge carrier transport only in one dimension [12]. Thus, nanowires have been
13
14 envisaged as efficient electrode material thanks to their reduced ion diffusion length leading
15
16
17 towards increased charging/discharging rates. Mehmood et al. prepared ultrathin MnO2 nanowires
18
19 coupled with MXene that showed a specific capacitance of 611.5 Fg-1 in conjunction with 96%
20
21 capacitance retention after 1000 cycles [13]. Mane et al. investigated the supercapacitor
22
23
24 performance of MnO2 nanowires grown on NF and observed a specific capacitance of 641 Fg-1 @
25
26 2 Ag-1 current density with 97% capacitance retention [14]. Wang et al. hydrothermally
27
28
29 synthesized ultralong α-MnO2 nanowires having 118 Fg-1 specific capacitance @ 2 mVs-1 scan
30
31 rate and showed 95.3% capacitance retention after 1000 cycles [15].
32
33
34 In the recent times, manganese dioxide (MnO2) has shown great promise towards electrochemical
35
36
37 properties owing to its nontoxicity, ease in morphology control, cost effectiveness, and higher
38
39 theoretical capacitance (1370 Fg-1) [16-18]. MnO2 exists in a number of crystal phases including
40
41
42 α-MnO2, β-MnO2, γ-MnO2, δ-MnO2, and λ-MnO2, differ in the arrangement of MnO6 octahedra
43
44 creating interstitial sites with different tunnel size. α-MnO2 because of its larger tunnel structure
45
46
47
(4.62Å) is believed to be the best due to the ease of insertion/disinsertion of electrolyte ions [19,
48
49 20]. The α-MnO2 crystal phase stability is very important as it is transformed to other phases. Water
50
51 as well as ions with larger ionic radii such as (K+, Ag+, Na+ etc.) are finest choices as stabilizing
52
53
54 species [21, 22]. The pre-intercalation into MnO2 framework also offers (1) superior cyclic
55
56 stability thanks to the stabilization of polymorph structure, (2) a contributor towards the
57
58
59 capacitance improvement (3) Such a pre-intercalation can increase the oxygen evolution
60
61
62 3
63
64
65
1
2
3
4 overpotential because of energy needed for cation pre-intercalation, which is a competitive process
5
6
7 against oxygen evolution potential [23, 24]. Furthermore, cations pre-intercalation into MnO2
8
9 framework has been demonstrated as beneficial to expedite the ion diffusion process. Na+ pre-
10
11
12 intercalation into MnO2 framework has been reported with superior electrochemical performance.
13
14 Zong et al. successfully synthesized Na pre-intercalated MnO2 nanosheets hydrothermally. They
15
16
17 observed a specific capacitance of 265.4 mF cm-2 along with excellent energy density 178 μWh
18
19 cm-2 in 3 M Na2SO4 [25]. Jabeen et al. investigated the supercapacitor performance of Na pre-
20
21 intercalated MnO2 nanowalls. They concluded that Na0.5MnO2 electrode showed 366 Fg-1 specific
22
23
24 capacitance with 96% capacitance retention after 10000 cycles @ 4Ag-1 current density [21]. Ting
25
26 et al. reported an increase in oxygen evolution overpotential of MnO2 by doping Na+ ions into
27
28
29 MnO2 tunnels. The as synthesized supercapacitor demonstrated an energy density of 61 WhKg-1
30
31 at 982 WKg-1 power density in addition with 94% capacitance retention.
32
33
34 The current work aims to synthesize ultrathin MnO2 nanowires (NWs) with pre-intercalation of
35
36
37 alkali ion (Na+) into the (2×2) tunnel framework. This pre-intercalation process is expected to
38
39 increase the overpotential for oxygen evolution, resulting in an increase in the operational voltage
40
41
42 of supercapacitors and hence the energy density. The pre-intercalation of Na+ ions into the tunnels
43
44 can also tune the structural, electrical, and electrochemical properties, as confirmed by XRD,
45
46
47
FTIR, and electrochemical studies. Additionally, ultrathin morphology and cracks in MnO2 NWs
48
49 by Na+ doping was confirmed by FESEM analysis, which resulted in an increase of surface area
50
51 shown by BET analysis. Alkali ions pre-intercalated MnO2 NWs were supported on 3-D nickel
52
53
54 foam (NF) in order to get more exposed surface area. This configuration along with ultrathin and
55
56 cracked morphology facilitate the electrolyte ions deep percolation into bulk giving rise to an
57
58
59 increase in capacitance. Overall, this work provides a strategy for improving the energy density of
60
61
62 4
63
64
65
1
2
3
4 supercapacitors by tuning the structural, electrical, and electrochemical properties of MnO2 NWs
5
6
7 through alkali ion pre-intercalation.
8
9
10 2. Experimental
11
12
13 2.1. Materials
14
15
16 All the chemicals used were of analytical grade and employed as received. Potassium
17
18
19 permanganate (KMnO4, Merck 99%), ammonium persulfate [(NH4)2S2O8, Sigma Aldrich, 98%],
20
21 nitric acid (HNO3, Daejung, 70%), sodium sulfate (Na2SO4, Sigma Aldrich, 99%), absolute
22
23
24 ethanol (C2H5OH, Daejung, 99%) and deionized (DI) water.
25
26
27 2.2. Synthesis of MnO2 and NaxMnO2
28
29
30 Bare and Na pre-intercalated MnO2 nanowires were synthesized by a simple hydrothermal method.
31
32 Typically, homogenous aqueous solution of KMnO4 (4g), in DI water was obtained by stirring.
33
34
35 After that, ammonium persulfate (2g) was added to the above solution by the subsequent addition
36
37 of 1 mL nitric acid. The mixture was stirred for half an hour for complete mixing. The
38
39
40 homogenized mixture was moved to 100 mL Teflon lined autoclave and preheated to obtain the
41
42 required temperature (180 °C). Subsequently, autoclave was heated for 16 hours at 180 °C.
43
44
Afterwards, autoclave was cooled to room temperature and precipitates were collected for washing
45
46
47 prior to drying at 100 °C for 15 hours. The dried brown color precipitates were grinded into fine
48
49 powder using pestle and mortar and stored in glass viols for further characterization. A schematic
50
51
52 illustration of bare and Na pre-intercalated MnO2 NWs is shown in Figure 1.
53
54
55
56
57
58
59
60
61
62 5
63
64
65
1
2
3
4
5
6
7
8
9
10
11 Stirring Autoclave Washing
12
13
14
15
16
17
18
19
20
21
22
23
24 Nanowires Drying
25
26
27
28
29
30
Figure 1: schematic illustration of MnO2 NWs synthesis by hydrothermal method.
31
32
33
34
35 2.3. Electrode preparation
36
37
38 Binder free electrodes of MnO2 NWs and Na-MnO2 NWs on nickel foam was prepared by
39
40 sonication. In a typical synthesis, MnO2 NWs slurry in DI water was prepared by sonication for
41
42
43 an hour. The homogenized slurry was then loaded on washed nickel foam using micropipette.
44
45 Subsequently, electro-active material loaded nickel foam was dried in an oven at 50 °C. The as-
46
47
48 obtained electrode was then used for electrochemical supercapacitor studies.
49
50
51 2.4. Electrochemical studies
52
53
54 Electrochemical measurements of MnO2 NWs and Na-MnO2 NWs were carried out using 3
55
56 electrode system in 1M Na2SO4. Cyclic voltametric (CV) measurements were performed in a
57
58
59 potential window 0-0.6 V. GCD measurements were taken at 1 A/g current density for the sake of
60
61
62 6
63
64
65
1
2
3
4 comparison among the electrodes. Moreover, impact of different current densities on Na0.1MnO2
5
6
7 was also observed. Electrochemical impedance spectroscopic (EIS) measurements were taken in
8
9 the frequency range 0.1 Hz-1MHz.
10
11
12 2.5. Characterization
13
14
15
16
Structural analysis of MnO2 NWs and Na-MnO2 NWs was performed on Lab-X XRD 6100 X-ray
17
18 diffractometer. Functional group investigations were carried out on IR Affnity-1S
19
20 spectrophotometer. Morphological studies were performed on ZEISS LEO SUPRA 55
21
22
23 field emission scanning electron microscope. JEOL JCM-6000 Plus SEM was used to assess the
24
25 elemental presence. N2 adsorption/desorption investigations were obtained using Micromeritics
26
27
28 ASAP 2020 Physisorption analyzer to calculate the surface area. Current-voltage (I-V) response
29
30 was checked using Keithly 6400. Electrochemical measurements were performed on GAMRY
31
32
33 interface 5000E.
34
35
36 3. Results and Discussion
37
38
39 3.1. X-ray diffraction (XRD) analysis
40
41
42 Structural investigations of the as-synthesized MnO2 NWs and Na-MnO2 NWs are carried out by
43
44
XRD analysis. XRD patterns of MNO2 NWs and NaxMnO2 NWs are recorded in the 2θ range 20°-
45
46
47 70° and are shown in Figure 2.
48
49
50 All XRD diffraction peaks of MnO2 NWs and Na-MnO2 NWs are well corresponded to the
51
52
53
tetragonal crystal symmetry of α-MnO2 with I4/m space group. In a typical XRD profile of MnO2
54
55 NWs, the diffraction peaks observed at 2θ = 25.71°, 28.83°, 36.68°, 37.53°, 38.95°, 41.20°, 41.97°,
56
57 47.33°, 49.84°, 56.33°, 60.26°, 65.13°and 69.69° are indexed to the (220), (310), (211), (221),
58
59
60 (420), (330), (301), (510), (411), (600),(521), (002) and (541) hkl diffraction planes respectively
61
62 7
63
64
65
1
2
3
4 having the most intense diffraction peak at 2θ = 37.53° that well match with previous literature
5
6
7 and ICDD card No. 00-044-0141 [26]. It appears that increasing the concentration of Na in MnO2
8
9 leads to a reduction in the intensity of XRD diffraction peaks. This reduction is attributed to a
10
11
12 decrease in the crystallinity of the material due to lattice distortion caused by the presence of Na
13
14 [27, 28]. However, it is also mentioned that no impurity peak related to any Na phase is observed,
15
16
17 indicating that Na is well dispersed throughout the MnO2 lattice. It suggests that Na is not present
18
19 as a separate phase within the material, but is instead incorporated into the MnO2 lattice structure.
20
21 Enlarged view of the diffraction peak (221) shows XRD peak shifting towards smaller diffraction
22
23
24 angle in Figure 2b. This peak shifting is ascribed to the Na+ ions pre-intercalation into (2×2)
25
26 tunnels of MnO2 resulting in an expansion of lattice planes [29-31]. For Na0.15MnO2 NWs the peak
27
28
29 shifting towards lower angle is more pronounced. It may be due to the replacement of hydrated K+
30
31 ions with hydrated Na+ ions as the dopant concentration exceed 0.1 molar. Since the ionic radius
32
33
34 of hydrated Na+ ions is bigger than hydrated K+ ions, thus causes an expansion of crystal lattice.
35
36 Overall, the information suggests that the addition of Na to MnO2 results in lattice distortion,
37
38
39
reduced crystallinity and XRD peak shifting towards smaller diffraction angle, but without the
40
41 formation of separate Na phase.
42
43
44 The crystallographic parameters including crystallize size and dislocation density are computed
45
46
47
using according to the following equation and the results are given in Table 1.
48
49 𝐾𝜆
50 D= (1)
𝛽𝑐𝑜𝑠𝜃
51
52
53 1
54 𝛿 = D2 (2)
55
56
57
58
59
60
61
62 8
63
64
65
1
2
3
4 Crystallite size is decreased for Na pre-intercalated MnO2 NWs as compared to the bare MnO2
5
6
7 NWs. This decrease in crystallite size is ascribed to the expansion in crystallite lattice resulted
8
9 from larger ionic radii of pre-intercalated Na+ ions [32, 33].
10
11
12 Table 1: Crystallite size and dislocation density of bare and Na pre-intercalated MnO2 NWs.
13
14
15
16 Sr. No. Sample Name Crystallite size Dislocation density
17
18 (nm) (nm-2)
19
20
21
1 MnO2 21 0.0022
22
23 2 Na0.05MnO2 18 0.0030
24
25
3 Na0.10MnO2 16 0.0039
26
27
28 4 Na0.15MnO2 12 0.0069
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
61
62 9
63
64
65
1
2

(a (b
3
4
5
6
7
8
) )
9
10
11
12
13
14
15
16
17
18
19
20

(c
21
22
23

)
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42 Figure 2: (a) XRD profile of bare and Na pre-intercalated MnO2 NWs, (b) Enlarged view of
43
44
45 XRD profile, and (c) FTIR spectra of bare and Na pre-intercalated MnO2 NWs.
46
47
48 3.2. Fourier transform infrared spectroscopy (FTIR)
49
50
51 Presence of functional groups as well as metal-oxygen bond in MnO2 NWs and NaxMnO2 NWs
52
53
54 are explored by FTIR spectra. FTIR spectra for bare MnO2 NWs and NaxMnO2 NWs are recorded
55
56 in the wavenumber ranges between 4000-400 cm-1 and are shown in Figure 2c. There exist five
57
58
59
characteristics absorption bands in bare MnO2 NWs and well corresponded with the previous
60
61
62 10
63
64
65
1
2
3
4 literature. The absorption bands positioned at 461 cm-1, 523 cm-1, and 706 cm-1 are ascribed to the
5
6
7 metal-oxygen bond (Mn-O) of MnO2 NWs, confirming the successful synthesis of MnO6
8
9 octahedron [34]. An absorption band with moderate intensity is observed at 1530 cm-1. It is
10
11
12 credited to the existence of OH functional group of water either adsorbed on the MnO2 surface or
13
14 connected to Mn atoms or reside into the MnO2 tunnels to stabilize α-MnO2 [35, 36]. A broad
15
16
17 absorption band positioned around 3500 cm-1 is ascribed to the O-H stretching vibration.
18
19
20 For all Na pre-intercalated MnO2 NWs, the absorption bands similar to bare MnO2 NWs are
21
22 observed. However, the position of metal-oxygen bands and O-H functionality intensity are
23
24
25 changed. The absorption band shift towards lower wavenumber is ascribed to the weakening of
26
27 metal-oxygen bond due to the crystal lattice distortion with Na pre-intercalation [32, 37], on the
28
29
other hand O-H functional group absorption band gets more intense. Since, the charge to size ratio
30
31
32 and hydrated ionic radii of Na+ is greater than that of hydrated K+, the ion dipole interaction is
33
34 greater, thus requiring more solvent molecules to be attracted by single Na+ ion, thereby increasing
35
36
37 the water content in Na pre-intercalated MnO2 NWs [24].
38
39
40 3.3. Scanning electron microscopy (SEM)
41
42
43 Morphological investigations of MnO2 NWs and all NaxMnO2 NWs are evaluated by FESEM.
44
45
46
Figure 3 shows the FESEM micrographs of all the samples. Figure 3a illustrates ultrathin one-
47
48 dimensional morphology (sub 50 nm) with aspect ratio more than 50. Hence, these are considered
49
50 to be the NWs. Along with being ultrathin, MnO2 NWs are homogeneously distributed and
51
52
53 ultralong (micrometer range). Ultrathin MnO2 NWs are entangled with each other giving the
54
55 appearance of straw’s nest. These long range interconnected NWs facilitate the charge transfer by
56
57
58 offering uninterrupted charge carrier’s transport. It can be realized from Figure 3(b-d) that all Na
59
60 pre-intercalated MnO2 NWs show the same one-dimensional morphology with cracked surface.
61
62 11
63
64
65
1
2
3
4 The cracked surface is expected to be beneficial because it offers more active sites and less dead
5
6
7 area for electrolyte ions adsorption. However, diameter of all Na pre-intercalated MnO2 NWs
8
9 becomes marginally thicker than bare MnO2 NWs but still below 50 nm. This increase in NWs
10
11
12 diameter is ascribed to the stronger diffusion of Na+ ions in MnO2 [38]. In Na0.15MnO2, a few
13
14 NWs are observed with shorter lengths.
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
Figure 3: (a) SEM micrograph of MnO2 NWs, (b) SEM micrograph of Na0.05MnO2 NWs,(c) SEM
47
48 micrograph of Na0.10MnO2 NWs, (d) SEM micrograph of Na0.15MnO2 NWs. (inset shows
49
50 diameter distribution and average diameter).
51
52
53
54
55
56
57
58
59
60
61
62 12
63
64
65
1
2
3
4 3.4. Elemental analysis
5
6
7
8
Elemental composition, purity, and chemical homogeneity of MnO2 NWs and NaxMnO2 NWs are
9
10 evaluated by EDX analysis. Figure 4a illustrates the MnO2 NWs EDX profile indicating only the
11
12 presence of Mn and O atoms. An additional peak between 3-4 KeV is also observed. This peak is
13
14
15 associated with the K atoms. The K atoms from KMnO4 reside into the α-MnO2 tunnels in hydrated
16
17 form. When an electron from outermost orbit of “Mn” falls to the ‘K’ or ‘L’ orbit the energy loss
18
19
20 is measured in KeV. The greater energy loss is noted when an electron falls to K orbit as compared
21
22 to L orbit from higher energy orbit. In EDX profile of NaxMnO2 NWs an additional spike along
23
24
25 with Mn, K, and O spikes is also observed at 1 KeV. This peak is ascribed to the de-excitation of
26
27 Na atoms, thus confirming the successful pre-intercalation of Na in MnO2. It is clear from Figure
28
29
4 that when the dopant concentration is up to 0.1 molar, no significant decrease in spike related to
30
31
32 K atom is observed. A significant decrease in peak related to ‘K’ atom is observed when the dopant
33
34 concentration exceed 0.1 molar. It suggest that as the dopant concentration increases beyond 0.1
35
36
37 molar, hydrated Na+ ions start replacing hydrated K+ due to higher water content of hydrated Na+
38
39 ions for greater stabilization of (2×2) tunnels [39-41]. It also supports the pronounced XRD peak
40
41
42 shifting towards lower diffraction angle for Na0.15MnO2. The decrease in ionic conductivity for
43
44 Na0.15MnO2 as compared to Na0.10MnO2 exhibited by EIS measurements is also due to bandgap
45
46
47
broadening and lesser mobility of hydrated Na+ ions because of greater hydrated ionic radius and
48
49 well consistent with XRD and EDX results.
50
51
52
53
54
55
56
57
58
59
60
61
62 13
63
64
65
1
2
3
4
5
6
7
8
( (
9
10
11
12
13
14
15
16

(c (
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33 Figure 3: (a) EDX profile of MnO2 NWs, (b) EDX profile of Na0.05MnO2 NWs,(c) EDX profile
34
35 of Na0.10MnO2 NWs, (d) EDX profile of Na0.15MnO2 NWs.
36
37
38 3.5. N2 adsorption-desorption Isotherm analysis
39
40
41
42 Braunner Emmet Teller (BET) analysis is conducted to evaluate the consequences that Na pre-
43
44 intercalation has on surface area of MnO2 NWs. The N2 adsorption desorption isotherms measured
45
46 at 77K are illustrated in Figure 5. Figure 5a shows that N2 adsorption desorption isotherms
47
48
49 resemble a type III isotherm with an H3 hysteresis loop, which is a common type of hysteresis
50
51 loop observed in mesoporous materials with slit-shaped pores. This suggests that the MnO2 NWs
52
53
54 with alkali ions doping have a mesoporous structure with slit-shaped pores as shown by FESEM
55
56 micrographs. The BET analysis is a common method used to determine the specific surface area
57
58
59 of a material. It involves measuring the amount of gas adsorbed onto the surface of a solid at
60
61
62 14
63
64
65
1
2
3
4 different pressures and then analyzing the data to determine the surface area. The BET equation is
5
6
7 commonly used to calculate the surface area, and it assumes that the gas adsorbs onto a
8
9 homogeneous surface. Overall, the N2 adsorption desorption isotherms analysis provides important
10
11
12 information about the surface area and pore structure of the MnO2 NWs with alkali ion pre-
13
14 intercalation. Figure 5 shows that although bare and Na pre-intercalated MnO2 NWs have the same
15
16
17 adsorption pattern however the amount of gas adsorbed by Na pre-intercalated MnO2 NWs is much
18
19 greater than bare MnO2 NWs. It suggests that alkali ions pre-intercalation have been beneficial to
20
21 improve the BET surface area. BET surface area of bare and Na pre-intercalated MnO2 NWs is
22
23
24 given in Table 2.
25
26
27 3.6. Thermogravimetric analysis (TGA)
28
29
30 The thermal stability of MnO2 NWs and impact of Na+ pre-intercalation on thermal stability of
31
32
33 MnO2 NWs are assessed by TGA analysis. Moreover phase evolution of MnO2 NWs is also studied
34
35 by TGA. The as-synthesized bare and pre-intercalated MnO2 NWs are heated in the range 35-950
36
37
°C and their spectra are illustrated in Figure 5c. Weight loss up to 230 °C is associated with the
38
39
40 removal of water that is adsorbed on the surface or present into the crystal lattice [42]. Water
41
42 removal for Na0.10MnO2 NWs (4.7%) is higher than MnO2 NWs (3.6%). It shows the presence of
43
44
45 more water content in pre-intercalated samples and well consistent with the FTIR results. This
46
47 increased water content is ascribed to the hydrated Na+ ions, which accommodate more water in
48
49
50 the tunnels for stabilization. The weight loss between 230-705 °C is associated with the MnO2
51
52 phase transformation to Mn2O3 with the evolution of oxygen. The Na0.10MnO2 phase is
53
54
55
transformed to the Mn2O3 at higher temperature as compared to the MnO2 NWs. It is due to the
56
57 stabilization of tunnel structure by Na+ ions [43]. The weight loss at relatively higher temperature
58
59 is associated with Mn2O3 phase transformation to Mn3O4 phase, as a result of more oxygen
60
61
62 15
63
64
65
1
2
3
4 evolution [44]. A summary of MnO2 phase transformation as a function of temperature is given in
5
6
7 Table 2.
8
9
10 Table 2: BET surface area and phase transformation of bare Na pre-intercalated MnO2 NWs.
11
12
13
Sr. No. Sample Name BET surface area Phase Phase Electrical
14
15 (m2/g) transformation transformation conductivity
16
17 temperature temperature (Scm-1)
18
19 (°C) (°C)
20
21
MnO2 Mn2O3
22
23 Mn2O3 Mn3O4
24
25 1 MnO2 55 589 886 2.93×10-5
26
27
28 2 Na0.05MnO2 - - - 1.91×10-4
29
30 3 Na0.10MnO2 81 705 950 3.5×10-3
31
32
33 4 Na0.15MnO2 - - - 9.49×10-4
34
35
36
37
38 3.7. Current voltage (I-V) measurements
39
40
41 Current-voltage measurements are performed to calculate the electrical conductivity of MnO2
42
43 NWs and Na pre-intercalated MnO2 NWs by two probe method in the voltage range between -5-5
44
45
46 V. I-V measurements are carried out by pressing fine powdered samples into circular pellets. I-V
47
48 spectra of bare and Na pre-intercalated MnO2 NWs are illustrated in Figure 5d.
49
50
51 It can be inferred from Figure 5d that Na pre-intercalated MnO2 NWs have better I-V response in
52
53
54 comparison with bare MnO2 NWs. The conductivity values for all the samples are calculated using
55
56 the following equation and are given in Table 2.
57
58
59 𝐿
60 𝜎 = 𝑅𝐴 (3)
61
62 16
63
64
65
1
2
3
4 The I-V measurements reveales that conductivity of MnO2 NWs is increased by alkali ion pre-
5
6
7 intercalation from 2.93×10-5 to 3.5×10-3. It is ascribed to the hybridization among Mn d, O p and
8
9 Na s orbitals resulting in an increase in carrier density near the fermi level giving rise to narrowing
10
11
12 of bandgap which eases the charge transference. An interesting change in the trend of increased
13
14 conductivity by increasing Na content is observed as the Na concentration exceeds 0.1 molar.
15
16
17 Conductivity for Na0.10MnO2 NWs is more than Na0.15MnO2 NWs. It is well known that α-MnO2
18
19 has 12.5 wt% tunnels. As the dopant concentration exceeds 0.1 molar, the Na+ ions start replacing
20
21 K+ ions from tunnels [41]. K+ ions have greater ionic radius as compared to the Na+ ions. Due to
22
23
24 greater ionic radius, K+ ions are capable to lower the bandgap more than relatively smaller ionic
25
26 radius Na+ ions. Consequently, electron transfer from valence band to conduction band becomes
27
28
29 more difficult [45, 46]. Thus, electrical conductivity is decreased. Overall, the results suggest that
30
31 the pre-intercalation of alkali ions can enhance the conductivity of MnO2, but the effect depends
32
33
34 on the concentration of the pre-intercalated ions.
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
61
62 17
63
64
65
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
(a) (b)
21
22
23
24
25
26
27
28
29
30
31
32
33
34

(c) (d)
35
36
37
38
39
40
41
42 Figure 4: (a) N2 adsorption desorption curve of MnO2 NWs, (b) N2 adsorption desorption curve
43
44 of Na0.1MnO2 NWs,(c) TGA analysis of bare and Na intercalated MnO2 NWs, (d) I-V plot of bare
45
46 and Na intercalated MnO2 NWs,.
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62 18
63
64
65
1
2
3
4
5
6
7
8
9
10
11 4. Electrochemical measurements
12
13
14 To investigate the electrochemical behavior of bare and pre-intercalated MnO2 NWs, the electro-
15
16 active materials are decorated on HCl treated 3D nickel foam. CV, GCCD and electrochemical
17
18
19 impedance spectroscopy are executed in a neutral electrolyte (1M Na2SO4). To probe suitable
20
21 potential window, rough estimation of specific capacitance, and faradaic redox reaction of all the
22
23
24 MnO2 samples, CV measurements are performed in the potential window 0-0.6V. CV profile of
25
26 bare and pre-intercalated MnO2 NWs on NF recorded at 50 mVs-1 in the positive potential window
27
28 are depicted in Figure 6. HCl treated NF CV profile is also given in Figure 6a and shows a minimal
29
30
31 enclosed area. It shows insignificant contribution of NF towards the specific capacitance as
32
33 compared to the bare and pre-intercalated MnO2 NWs on NF.
34
35
36
37 CV profile of bare and Na pre-pre-intercalated MnO2 show the oxidation (charging) and reduction
38
39 (discharging) peak, suggesting the faradaic redox reaction contribution towards specific
40
41
42 capacitance. An increased anodic (charging) peak current and cathodic (discharging) peak current
43
44 associated with Na pre-pre-intercalated MnO2 NWs suggest enhanced specific capacitance up to
45
46
47
Na content 0.1 M. A morphological change associated with Na+ pre-intercalation is observed.
48
49 Consequently, a large number of active sites on MnO2 NWs surface might be generated by Na+
50
51 doping. This change is supposed to provide ion rich environment, which is expected to facilitate
52
53
54 the cations diffusion. Na+ ions from the electrolyte penetrate and instantly incorporate inside the
55
56 surface of NaxMnO2 NWs during charging, suggesting that electrolyte ions (Na+) do not require
57
58
59 long distance diffusion to contribute in the faradaic redox process [41]. Additionally, As a result
60
61
62 19
63
64
65
1
2
3
4 of enhanced mobility of cations, pre-intercalation activate more sites contributing towards greater
5
6
7 specific capacitance. Additionally, due to ultrathin morphology of nanowires, there is an increase
8
9 in electrochemically active sites, which has reduced the energy transfer channel and the electrolyte
10
11
12 ions' diffusion pathway, thus optimizing the electrochemical performance. Furthermore, as
13
14 observed in the SEM micrographs, the pre-intercalation of Na+ ions breaks the lattice alignment,
15
16
17 leading to a wide turnout of grain boundaries. The large number of grain boundaries provide more
18
19 active sites for redox reaction as well as a convenient diffusion channel for ions and electrons,
20
21 permitting for more competent usage of the electrode materials. Additionally, larger number of
22
23
24 grain boundaries promote the creation of lattice defects, which increase the amount of oxygen
25
26 vacancies and thus improve NaxMnO2's electrical conductivity . However, as the dopant
27
28
29 concentration exceeds 0.1 molar, a decrease in specific capacitance is observed. This decrease
30
31 might be associated with the decrease in conductivity with an increase in dopant content. When
32
33
34 Na concentration increase beyond 0.1 M, hydrated Na+ ions start replacing hydrated K+ ions from
35
36 MnO2 lattice. Due to lower mobility of hydrated Na+ ions as compared to the hydrated K+ ions as
37
38
39
a consequent of higher hydrated radius of Na+ ions is responsible for decrease in conductivity. CV
40
41 profile of Na0.10MnO2 NWs on NF at different scan rates are illustrated in Figure 6b. It is observed
42
43 that as the scan rate increases the anodic and cathodic peaks current increases. However, the shape
44
45
46 of respective redox peak is not changed suggesting the good stability of the electrode material. The
47
48 overall redox reaction involved during charging and discharging can be summarized as follows.
49
50
51
52 NaxMnO2 + Na+ + e- NaxMnOONa
53
54
55 As charge storage in a supercapacitor involves two storage mechanisms, i.e, electric double layer
56
57
58 charge storage (EDLC) and pseudocapacitive charge storage (faradaic redox reaction). To
59
60
61
62 20
63
64
65
1
2
3
4 determine the leading charge storage process in this work a statistical analysis (power law) was
5
6
7 used.
8
9
10 i = 𝑎𝑣 𝑏 (4)
11
12
13
14 OR
15
16
17 log(i) = blog(𝑣) + log⁡(𝑎) (5)
18
19
20
21 A graph is plotted by taking log of current on y-axis and log of scan rate on x-axis and is shown in
22
23 Figure 6c. The linear fitting of the data gives the value of the slope (b) and intercept (a). Value of
24
25
26 slope determines either the dominant charge storage process is EDLC or pseudocapacitor [47]. If
27
28 the slope value is 0.5, the dominant charge storage process is psudocapacitor, and if the slope value
29
30
31 is unity the major contribution towards charge storage is capacitive controlled. The slope value for
32
33 Na0.10MnO2 comes out to be 0.6. It suggests that the dominant charge storage mechanism is
34
35
36
diffusion controlled (Pseudocapacitor).
37
38
39 Further, evaluation of the dominant charge storage mechanism and its dependence on scan rate is
40
41
42
studied by Wang’s method according to the following relation.
43
44
45 𝑖 = 𝐾1 𝑣 + ⁡ 𝐾2 𝑣 0.5 (6)
46
47
48
49
OR
50
51
52 𝑖𝑣 −0.5 = 𝐾1 𝑣 0.5 + ⁡ 𝐾2 (7)
53
54
55
56 The profile of dominant charge storage at various scan rate is illustrated in Figure 6d. At lower
57
58 scan rate (5 mVs-1) the major charge storage contribution is diffusion controlled. At the highest
59
60
61
62 21
63
64
65
1
2
3
4 scan rate (50 mVs-1) the diffusion control charge storage contribution is decreased. These results
5
6
7 are well consistent with the results obtained from Figure 7b. In fact, at lower scan rate the
8
9 electrolyte ions have adequate time to percolate into the structure of the electroactive material
10
11
12 resulting an increase in faradaic redox reaction.
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27

(b
28
29
30 (a
31
32

(d
33
34
35
36
(c
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53 Figure 5: (a) CV profile of bare and Na intercalated Mno2 NWs, (b) CV profile of Na0.10MnO2
54
55 NWs @ various scan rate, (c) power law plot, (d) Impact of scan rate on capacitive and diffusion
56
57 control capacitance.
58
59
60
61
62 22
63
64
65
1
2
3
4 To calculate the exact value of specific capacitance for bare and Na pre-intercalated MnO2 NWs,
5
6
7 galvanostatic cyclic charge discharge (GCCD) measurements are carried out. Comparative GCCD
8
9 profile of bare and Na pre-intercalated MnO2 NWs recorded at 1 Ag-1 are given in Figure 7. The
10
11
12 discharging curve for Na0.1MnO2 shows the longest discharging time (628 s) among all the electro-
13
14 active materials. The discharging time showed by MnO2 NWs, Na0.05MnO2 NWs, and Na0.15MnO2
15
16
17 NWs are 319 s, 456 s, and 530 s respectively. It concludes that Na0.10MnO2 on NF is the best
18
19 electro-active material among all the electrodes. Moreover, the precise and exact specific
20
21 capacitance for all the samples is calculated by using the following equation that is based on current
22
23
24 density and discharging time.
25
26
27 𝐼×△𝑡
𝐶𝑠𝑝 = 𝑚×△𝑉⁡ (8)
28
29
30
31 The specific capacitance of all the samples calculated using eq. 8 are shown in Figure 7d. The
32
33 increase in specific capacitance by increasing Na concentration is due to the ultrathin morphology
34
35
36 and Na pre-intercalation. Ultrathin morphology and Na pre-intercalation reduces the electrolyte
37
38 ions diffusion pathway. Thus, resulting in an increase in specific capacitance by increasing Na
39
40
41
concentration. The impact of varying current density on charging/discharging of Na0.10MnO2 is
42
43 shown in Figure 7b. As the current density increases, the charging/discharging curve shrinks,
44
45 however the shape of the charging/discharging curves remains intact. It suggests the excellent
46
47
48 stability of the electro-active material. It is evident from Table 3 that as the current density
49
50 amplifies, the specific capacitance decreases. At higher current density electrolyte ions are not
51
52
53 capable to assess the bulk of the electro-active material rather they adsorb on electrode surface
54
55 only. Consequently, redox active sites in the bulk are not fully accessed resulting in a decrease in
56
57
58 specific capacitance. On the other hand at lower current density the increased specific capacitance
59
60 thanks to the adequate time that electrolyte ions have for deep percolation into bulk, enabling full
61
62 23
63
64
65
1
2
3
4 access to the redox active sites in conjunction with adsorption at electrode surface. Capacitance
5
6
7 retention after 6000 cycles and specific capacitance of all the samples are shown in Figure 7c and
8
9 7d respectively.
10
11
12
13
14
15
16
17
18
19
20
21
22 Table 3: Impact of current density on specific capacitance of Na0.1MnO2 NWs.
23
24
25 Sr. No. Current density Specific capacitance Capacitance retention
26
27
(Ag-1) (Fg-1) (%)
28
29
30 1 1 1046 100
31
32 2 3 1005 96
33
34
35 3 5 966 92
36
37 4 7 933 89
38
39
40 5 9 900 85
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62 24
63
64
65
1
2
3
4
5
6
7
(a (b
8
9
10
) )
11
12
13
14
15
16
17
18
19
20
21
22
23
24
(c (d
25
26 ) )
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42 Figure 6: (a) GCD curves of MnO2 and Na pre-intercalated MnO2 @ 1 Ag-1, (b) GCD curves of
43
44 Na0.10MnO2 @ various current densities, (c) Cyclic stability and capacitance retention after
45
46 6000 cycles, (d) capacitance of bare and Na intercalated MnO2 NWs.
47
48
EIS measurements are conducted to evaluate the resistance offered by the electro-active material
49
50
51 and electrolyte to the flow of alternating current. The Nyquist plot is drawn by taking the real and
52
53 imaginary resistances on x-axis and y-axis respectively and is shown in Figure 8. Nyquist plot
54
55
56 usually consists of three portions, (i) x-intercept in the high frequency region (ii) semi-circle in the
57
58 high frequency region (iii) sloping line in the low frequency region. Bare MnO2/NF electrode has
59
60
61
62 25
63
64
65
1
2
3
4 the largest semicircle suggesting the highest charge transfer resistance (Rct) and x-intercept far
5
6
7 away from origin among all the sample indicating the highest solution resistance (RES).
8
9 Na0.1MnO2/NF electrode has the smallest semicircle arc and x-intercept closest to the origin
10
11
12 indicating lowest charge transfer resistance as well as smallest solution resistance respectively.
13
14 Values of Rct and RS and Rw for all the samples are given in Table 4. An interesting change in
15
16
17 the trend of increased conductivity by increasing doping content was observed as the doping
18
19 concentration exceed 0.1 molar. Conductivity for Na0.10MnO2 NWs is more than Na0.15MnO2
20
21 NWs. It is well known that α-MnO2 has 12.5 wt% tunnels. As the dopant concentration exceed
22
23
24 from 0.1 molar, the Na+ ions start replacing K+ ions from tunnels. Hydrated radius of Na+ ions
25
26 (3.58 Å) is greater that hydrated K+ ions radius (3.31 Å). Due to greater hydrated ionic radius, Na+
27
28
29 ions have lesser mobility as compared to the hydrated K+ ions due to stronger interaction between
30
31 Na+ ions and water originating from higher charge to size ratio, thus leading to a decrease in
32
33
34 conductivity by increasing dopant concentration beyond 0.1 molar [24].
35
36
37 Randle circuit diagram for Na 0.10MnO2/NF electrode is shown in inset of Figure 8a. In the fitted
38
39 circuit Rct represents the charge transfer resistance, RS is the solution resistance, Rw symbolizes
40
41
42 the Warburg resistance, and Cp represents capacitance by faradaic redox reaction.
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62 26
63
64
65
1
2
3
4 Table 4: Specific capacitance, and various resistances of bare and Na pre-intercalated MnO2 NWs.
5
6
7
8 Sr. No. Sample Name Rct Rs Carrier charge density
9
10 Ω (Ω) (cm-3)
11
12
13
1 MnO2 57 4.73 1.52 × 1026
14
15 2 Na0.05MnO2 40 4.17 -
16
17 3 Na0.10MnO2 15 2.32 1.29 × 1027
18
19
20 4 Na0.15MnO2 23 3.16 -
21
22
23
24
25
26 5. Oxygen evolution overpotential analysis
27
28
29 Impact of Na pre-intercalation onto oxygen evolution overpotential of MnO2 NWs is studied by
30
31 linear scanning voltammetry (LSV). LSV profile of bare MnO2 NWs and Na pre-intercalated
32
33
34 MnO2 NWs are shown in Figure 8b. The oxygen evolution overpotential for bare MnO2 NWs and
35
36 Na pre-intercalated MnO2 NWs calculated at current density 20 mA cm-2 is shown in Figure 8b.
37
38 By Na pre-intercalation into MnO2 NWs the oxygen evolution overpotential is increased since the
39
40
41 pre-intercalation is competitive process against the oxygen evolution overpotential. These results
42
43 are also consistent with the CV profile and previous results which show an increase in potential of
44
45
46 the redox peaks by Pre-intercalation [21, 22]. This increase in oxygen evolution overpotential is
47
48 beneficial since the Na+ pre-intercalated MnO2 supercapacitor electrode can operate at higher
49
50
51 operating voltage [24]. The LSV results suggest that Na pre-intercalated MnO2 electrode coupled
52
53 with appropriate counter electrode can be useful to assemble a device that can operate at wider
54
55
56
potential window.
57
58
59
60
61
62 27
63
64
65
1
2
3
4 6. Mott-Schottky studies
5
6
7
8
Impact of Na+ ions pre-intercalation on the production of oxygen vacancies and carrier charge
9
10 density in MnO2 lattice are evaluated by Mott-Schottky (M-S) analysis [48]. The M-S profiles of
11
12 MnO2 NWs and Na0.1MnO2 NWs are shown in Figure 8 (c, d). Flat band potential for bare and Na
13
14
15 pre-intercalated MnO2 NWs are 0.022 and -0.048 respectively. The M-S plot of MnO2 NWs shows
16
17 negative slope which confirms that it is a p-type semiconductor material. The slope calculated
18
19
20 from M-S plot of Na pre-intercalated MnO2 is positive. The slope value change from negative to
21
22 positive confirm the transformation of MnO2 NWs from p-type to n-type materials. This
23
24
25 transformation is ascribed to the production of oxygen vacancies. The oxygen vacancies are
26
27 produced to balance the charge. These results are well consistent with the previous literature [49-
28
29
51]. The carrier charge density is also calculated and given in Table 4. The increases in carrier
30
31
32 charge density by Na pre-intercalation into MnO2 lattice also suggests the formation of oxygen
33
34 vacancies [52]. The oxygen vacancies have been reported to be advantageous for electrochemical
35
36
37 studies by offering more active sites [53].
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62 28
63
64
65
1
2
3
4
5
6
7
8
9
10 C
11
12 R
13 R R
14
15
16
17
(a (b
18
19
20
) )
21
22
23
24
25

(c (d
26
27
28
29
30
31
) )
32
33
34
35
36
37
38
39
40
41
42
43
44
45 Figure 7: (a) EIS plot of bare and Na intercalated MnO2 NWs, (b) LSV profile of bare and Na
46
47 intercalated MnO2 NWs, (c) M-S plot of bare MnO2 NWs, (d) M-S plot of Na0.10MnO2 NWs.
48
49 Conclusion
50
51
52
53
In summary, bare and Na pre-intercalated MnO2 NWs with special control over morphology were
54
55 synthesized by a facile hydrothermal method. Successful synthesis and Na pre-intercalation were
56
57 validated by physical-chemical methods. Moreover, the impact of Na pre-intercalation on
58
59
60 structural, morphological, electrical, thermal, and electrochemical properties were also
61
62 29
63
64
65
1
2
3
4 investigated. SEM micrographs confirms the successful synthesis of nanowires with diameter (sub
5
6
7 50 nm). I-V and BET studies reveal that Na pre-intercalation improves the electrical conductivity
8
9 and surface area. Thermal stability test expresses the enhanced thermal stability as a function of
10
11
12 alkali ion pre-intercalation. Mott-Schottky analysis shows the electro-active material
13
14 transformation from p-type to n-type with increased carrier charge density. Electrochemical
15
16
17 studies were performed by decorating electrode material on porous 3-D nickel foam. CV
18
19 investigations show the respective redox peak in positive and negative current response with
20
21 pseudocapacitor as preferred charge storage mechanism. Na0.10MnO2 NWs exhibit superior
22
23
24 specific capacitance of 1046 Fg-1 @ 1Ag-1 than bare MnO2 (531 Fg-1). Na0.10MnO2 also exhibit the
25
26 lower resistances among all the samples and excellent capacitance retention (96.9% @ 12 Ag-1
27
28
29 after 6000 cycles). It is concluded from the above discussion that the improved electrochemical
30
31 performance of MnO2 is ascribed to the ultrathin morphology of nanowires, improved
32
33
34 conductivity, Na+ ions interaction into tunnels, and reduced diffusion path length of electrolyte
35
36 ions.
37
38
39 Acknowledgements: Authors are thankful to their respective institutes / organisations.
40
41
42 References
43
44
45
46 [1] C. Liu, Z. Yu, D. Neff, A. Zhamu, B.Z. Jang, Graphene-based supercapacitor with an ultrahigh
47 energy density, Nano letters, 10 (2010) 4863-4868.
48
49 [2] J. Yan, Q. Wang, T. Wei, Z. Fan, Recent advances in design and fabrication of electrochemical
50
51
supercapacitors with high energy densities, Advanced Energy Materials, 4 (2014) 1300816.
52
53 [3] T. Zhai, L. Wan, S. Sun, Q. Chen, J. Sun, Q. Xia, H. Xia, Phosphate ion functionalized Co3O4
54 ultrathin nanosheets with greatly improved surface reactivity for high performance
55 pseudocapacitors, Advanced Materials, 29 (2017) 1604167.
56
57
58
[4] P. Shi, L. Li, L. Hua, Q. Qian, P. Wang, J. Zhou, G. Sun, W. Huang, Design of amorphous
59 manganese oxide@ multiwalled carbon nanotube fiber for robust solid-state supercapacitor, ACS
60 nano, 11 (2017) 444-452.
61
62 30
63
64
65
1
2
3
4 [5] K. Fic, G. Lota, M. Meller, E. Frackowiak, Novel insight into neutral medium as electrolyte
5
6 for high-voltage supercapacitors, Energy & Environmental Science, 5 (2012) 5842-5850.
7
8 [6] G. Hasegawa, K. Kanamori, T. Kiyomura, H. Kurata, T. Abe, K. Nakanishi, Hierarchically
9 porous carbon monoliths comprising ordered mesoporous nanorod assemblies for high-voltage
10 aqueous supercapacitors, Chemistry of Materials, 28 (2016) 3944-3950.
11
12
13 [7] C.K. Chan, H. Peng, G. Liu, K. McIlwrath, X.F. Zhang, R.A. Huggins, Y. Cui, High-
14 performance lithium battery anodes using silicon nanowires, Nature nanotechnology, 3 (2008) 31-
15 35.
16
17 [8] S.H. Ko, D. Lee, H.W. Kang, K.H. Nam, J.Y. Yeo, S.J. Hong, C.P. Grigoropoulos, H.J. Sung,
18
19
Nanoforest of hydrothermally grown hierarchical ZnO nanowires for a high efficiency dye-
20 sensitized solar cell, Nano letters, 11 (2011) 666-671.
21
22 [9] Y. Xia, P. Yang, Y. Sun, Y. Wu, B. Mayers, B. Gates, Y. Yin, F. Kim, H. Yan, One‐
23 dimensional nanostructures: synthesis, characterization, and applications, Advanced materials, 15
24
(2003) 353-389.
25
26
27 [10] H.M. Chen, C.K. Chen, R.-S. Liu, L. Zhang, J. Zhang, D.P. Wilkinson, Nano-architecture and
28 material designs for water splitting photoelectrodes, Chemical Society Reviews, 41 (2012) 5654-
29 5671.
30
31
[11] P. Meduri, C. Pendyala, V. Kumar, G.U. Sumanasekera, M.K. Sunkara, Hybrid tin oxide
32
33 nanowires as stable and high capacity anodes for Li-ion batteries, Nano letters, 9 (2009) 612-616.
34
35 [12] X. Lu, W. Zhang, C. Wang, T.-C. Wen, Y. Wei, One-dimensional conducting polymer
36 nanocomposites: Synthesis, properties and applications, Progress in Polymer Science, 36 (2011)
37 671-712.
38
39
40 [13] M. Mahmood, A. Rasheed, I. Ayman, T. Rasheed, S. Munir, S. Ajmal, P.O. Agboola, M.F.
41 Warsi, M. Shahid, Synthesis of Ultrathin MnO2 Nanowire-Intercalated 2D-MXenes for High-
42 Performance Hybrid Supercapacitors, Energy & Fuels, 35 (2021) 3469-3478.
43
44 [14] S.D. Raut, H.R. Mane, N.M. Shinde, D. Lee, S.F. Shaikh, K.H. Kim, H.-J. Kim, A.M. Al-
45
46 Enizi, R.S. Mane, Electrochemically grown MnO2 nanowires for supercapacitor and
47 electrocatalysis applications, New Journal of Chemistry, 44 (2020) 17864-17870.
48
49 [15] W. Yao, J. Wang, H. Li, Y. Lu, Flexible α-MnO2 paper formed by millimeter-long nanowires
50 for supercapacitor electrodes, Journal of Power Sources, 247 (2014) 824-830.
51
52
53 [16] M.U. Khalid, K.M. Katubi, S. Zulfiqar, Z.A. Alrowaili, M. Aadil, M.S. Al-Buriahi, M.
54 Shahid, M.F. Warsi, Boosting the electrochemical activities of MnO2 for next-generation
55 supercapacitor application: Adaptation of multiple approaches, Fuel, 343 (2023) 127946.
56
57 [17] Y. Xie, C. Yang, P. Chen, D. Yuan, K. Guo, MnO2-decorated hierarchical porous carbon
58
59
composites for high-performance asymmetric supercapacitors, Journal of Power Sources, 425
60 (2019) 1-9.
61
62 31
63
64
65
1
2
3
4 [18] S. He, C. Hu, H. Hou, W. Chen, Ultrathin MnO2 nanosheets supported on cellulose based
5
6 carbon papers for high-power supercapacitors, Journal of Power Sources, 246 (2014) 754-761.
7
8 [19] Y. De Luna, A. Alsulaiti, M.I. Ahmad, H. Nimir, N. Bensalah, Electrochemically stable
9 tunnel-type α-MnO2-based cathode materials for rechargeable aqueous zinc-ion batteries,
10 Frontiers in Chemistry, 11 (2023).
11
12
13 [20] M.U. Khalid, M.F. Warsi, I. Shakir, M.F. Aly Aboud, M. Shahid, S.S. Shar, S. Zulfiqar,
14 Al3+/Ag1+ induced phase transformation of MnO2 nanoparticles from α to β and their enhanced
15 electrical and photocatalytic properties, Ceramics International, 46 (2020) 9913-9923.
16
17 [21] N. Jabeen, A. Hussain, Q. Xia, S. Sun, J. Zhu, H. Xia, High‐performance 2.6 V aqueous
18
19
asymmetric supercapacitors based on in situ formed Na0. 5MnO2 nanosheet assembled nanowall
20 arrays, Advanced materials, 29 (2017) 1700804.
21
22 [22] T. Xiong, T.L. Tan, L. Lu, W.S.V. Lee, J. Xue, Harmonizing energy and power density toward
23 2.7 V asymmetric aqueous supercapacitor, Advanced Energy Materials, 8 (2018) 1702630.
24
25
26
[23] A.C. Thenuwara, S.L. Shumlas, N.H. Attanayake, Y.V. Aulin, I.G. McKendry, Q. Qiao, Y.
27 Zhu, E. Borguet, M.J. Zdilla, D.R. Strongin, Intercalation of Cobalt into the Interlayer of Birnessite
28 Improves Oxygen Evolution Catalysis, ACS Catalysis, 6 (2016) 7739-7743.
29
30 [24] T. Xiong, W.S.V. Lee, J. Xue, K+-Intercalated MnO2 Electrode for High Performance
31
Aqueous Supercapacitor, ACS Applied Energy Materials, 1 (2018) 5619-5626.
32
33
34 [25] Q. Zong, Q. Zhang, X. Mei, Q. Li, Z. Zhou, D. Li, M. Chen, F. Shi, J. Sun, Y. Yao, Z. Zhang,
35 Facile Synthesis of Na-Doped MnO2 Nanosheets on Carbon Nanotube Fibers for Ultrahigh-
36 Energy-Density All-Solid-State Wearable Asymmetric Supercapacitors, ACS Applied Materials
37 & Interfaces, 10 (2018) 37233-37241.
38
39
40 [26] M. Shi, P. Xiao, C. Yang, Y. Sheng, B. Wang, J. Jiang, L. Zhao, C. Yan, Scalable gas-phase
41 synthesis of 3D microflowers confining MnO2 nanowires for highly-durable aqueous zinc-ion
42 batteries, Journal of Power Sources, 463 (2020) 228209.
43
44 [27] J. Wang, R. Li, Y. Xu, S. Lou, S. Zhou, Enhanced thermoelectric properties of Ag-doped
45
46 MnO2 single crystal nanowires for room-temperature application, Materials Research Express, 6
47 (2019).
48
49 [28] B.R. Kumar, B. Hymavathi, X-ray peak profile analysis of Sb2O3-doped ZnO nanocomposite
50 semiconductor, Advances in Natural Sciences: Nanoscience and Nanotechnology, 9 (2018)
51
52
035018.
53
54 [29] H. Usui, S. Suzuki, Y. Domi, H. Sakaguchi, Impacts of MnO2 Crystal Structures and Fe
55 Doping in Those on Photoelectrochemical Charge–Discharge Properties of TiO2/MnO2
56 Composite Electrodes, ACS Sustainable Chemistry & Engineering, 8 (2020) 9165-9173.
57
58
59
60
61
62 32
63
64
65
1
2
3
4 [30] I. Hussain, T. Hussain, S. Yang, Y. Chen, J. Zhou, X. Ma, N. Abbas, C. Lamiel, K. Zhang,
5
6 Integration of CuO nanosheets to Zn-Ni-Co oxide nanowire arrays for energy storage applications,
7 Chemical Engineering Journal, 413 (2021) 127570.
8
9 [31] I. Hussain, S.G. Mohamed, A. Ali, N. Abbas, S.M. Ammar, W. Al Zoubi, Uniform growth of
10 Zn-Mn-Co ternary oxide nanoneedles for high-performance energy-storage applications, Journal
11
12
of Electroanalytical Chemistry, 837 (2019) 39-47.
13
14 [32] M. Aadil, S. Zulfiqar, M. Shahid, S. Haider, I. Shakir, M.F. Warsi, Binder free mesoporous
15 Ag-doped Co3O4 nanosheets with outstanding cyclic stability and rate capability for advanced
16 supercapacitor applications, Journal of Alloys and Compounds, 844 (2020) 156062.
17
18
19
[33] M.F. Warsi, B. Bashir, S. Zulfiqar, M. Aadil, M.U. Khalid, P.O. Agboola, I. Shakir, M.A.
20 Yousuf, M. Shahid, Mn1-xCuxO2/ reduced graphene oxide nanocomposites: Synthesis,
21 characterization, and evaluation of visible light mediated catalytic studies, Ceramics International,
22 47 (2021) 5044-5053.
23
24
[34] E. Hastuti, A. Subhan, P. Amonpattaratkit, M. Zainuri, S. Suasmoro, The effects of Fe-doping
25
26 on MnO2: phase transitions, defect structures and its influence on electrical properties, RSC
27 Advances, 11 (2021) 7808-7823.
28
29 [35] A. Khan, A.M. Toufiq, F. Tariq, Y. Khan, R. Hussain, N. Akhtar, S.u. Rahman, Influence of
30 Fe doping on the structural, optical and thermal properties of α-MnO2 nanowires, Materials
31
32 Research Express, 6 (2019) 065043.
33
34 [36] D.P. Dubal, W.B. Kim, C.D. Lokhande, Galvanostatically deposited Fe: MnO2 electrodes for
35 supercapacitor application, Journal of Physics and Chemistry of Solids, 73 (2012) 18-24.
36
37 [37] M. Aadil, W. Hassan, H.H. Somaily, S.R. Ejaz, R.R. Abass, H. Jasem, S.K. Hachim, A.H.
38
39 Adhab, E.S. Abood, I.A. Alsafari, Synergistic effect of doping and nanotechnology to fabricate
40 highly efficient photocatalyst for environmental remediation, Journal of Alloys and Compounds,
41 920 (2022) 165876.
42
43 [38] S.M. Jadhav, R.S. Kalubarme, N. Suzuki, C. Terashima, J. Mun, B.B. Kale, S.W. Gosavi, A.
44
45
Fujishima, Cobalt-Doped Manganese Dioxide Hierarchical Nanostructures for Enhancing
46 Pseudocapacitive Properties, ACS Omega, 6 (2021) 5717-5729.
47
48 [39] C. Wei, C. Xu, B. Li, H. Du, D. Nan, F. Kang, Anomalous effect of K ions on electrochemical
49 capacitance of amorphous MnO2, Journal of Power Sources, 234 (2013) 1–7.
50
51
52
[40] N. Jabeen, Q. Xia, S.V. Savilov, S.M. Aldoshin, Y. Yu, H. Xia, Enhanced Pseudocapacitive
53 Performance of α-MnO2 by Cation Preinsertion, ACS Applied Materials & Interfaces, 8 (2016)
54 33732-33740.
55
56 [41] M. Feng, Q. Du, L. Su, G. Zhang, G. Wang, Z. Ma, W. Gao, X. Qin, G. Shao, Manganese
57
oxide electrode with excellent electrochemical performance for sodium ion batteries by pre-
58
59 intercalation of K and Na ions, Scientific reports, 7 (2017) 2219.
60
61
62 33
63
64
65
1
2
3
4 [42] A.K. Worku, D.W. Ayele, N.G. Habtu, Influence of nickel doping on MnO 2 nanoflowers as
5
6 electrocatalyst for oxygen reduction reaction, SN Applied Sciences, 3 (2021) 1-16.
7
8 [43] C.M. Julien, A. Mauger, Nanostructured MnO2 as electrode materials for energy storage,
9 Nanomaterials, 7 (2017) 396.
10
11 [44] S. Asiri, E. Çevik, H. Sabit, A. Bozkurt, Alginate-guided size and morphology-controlled
12
13 synthesis of MnO 2 nanoflakes, Soft Materials, 18 (2019) 1-9.
14
15 [45] X. Liu, B. Sui, P.H. Camargo, J. Wang, J. Sun, Tuning band gap of MnO2 nanoflowers by
16 Alkali metal doping for enhanced Ferroptosis/phototherapy synergism in Cancer, Applied
17 Materials Today, 23 (2021) 101027.
18
19
20 [46] S. Hu, X. Liu, C. Wang, P.H. Camargo, J. Wang, Tuning Thermal Catalytic Enhancement in
21 Doped MnO2–Au Nano-Heterojunctions, ACS applied materials & interfaces, 11 (2019) 17444-
22 17451.
23
24 [47] S.J. Patil, N.R. Chodankar, Y.-K. Han, D.W. Lee, Carbon alternative pseudocapacitive V2O5
25
26
nanobricks and δ-MnO2 nanoflakes@ α‐MnO2 nanowires hetero-phase for high-energy
27 pseudocapacitor, Journal of Power Sources, 453 (2020) 227766.
28
29 [48] Y. Yang, H. Wang, L. Wang, Y. Ge, K. Kan, K. Shi, J. Chen, A novel gas sensor based on
30 porous α-Ni (OH) 2 ultrathin nanosheet/reduced graphene oxide composites for room temperature
31
detection of NO x, New Journal of Chemistry, 40 (2016) 4678-4686.
32
33
34 [49] Y. Liu, P. Zhang, Catalytic decomposition of gaseous ozone over todorokite-type manganese
35 dioxides at room temperature: Effects of cerium modification, Applied Catalysis A: General, 530
36 (2017) 102-110.
37
38 [50] J. Ma, C. Wang, H. He, Transition metal doped cryptomelane-type manganese oxide catalysts
39
40 for ozone decomposition, Applied Catalysis B: Environmental, 201 (2017) 503-510.
41
42 [51] J. Jia, W. Yang, P. Zhang, J. Zhang, Facile synthesis of Fe-modified manganese oxide with
43 high content of oxygen vacancies for efficient airborne ozone destruction, Applied Catalysis A:
44 General, 546 (2017) 79-86.
45
46
47 [52] S. Yousaf, S. Zulfiqar, H.H. Somaily, M.F. Warsi, A. Rasheed, M. Shahid, I. Ahmad, An
48 efficient and stable iodine-doped nickel hydroxide electrocatalyst for water oxidation: synthesis,
49 electrochemical performance, and stability, RSC Advances, 12 (2022) 23454-23465.
50
51 [53] D. Gong, H. Tong, J. Xiao, Y. Wu, X. Chen, Y. Zhou, X. Zhang, Enhanced Reaction Kinetics
52
53 of N–MnO2 Nanosheets with Oxygen Vacancies via Mild NH3·H2O Bath Treatment for
54 Advanced Aqueous Supercapacitors, ACS Applied Energy Materials, 5 (2022) 7490-7502.
55
56
57
58
59
60
61
62 34
63
64
65
Manuscript

Declaration of interests

☑ The authors declare that they have no known competing financial interests or personal
relationships that could have appeared to influence the work reported in this paper.

☐ The authors declare the following financial interests/personal relationships which may be considered
as potential competing interests:

Corresponding authors

You might also like