Download as pdf or txt
Download as pdf or txt
You are on page 1of 81

Journal of CO2 Utilization

Recent Advances and Perspectives in Carbon Nanotube Production from the


Electrochemical Conversion of Carbon Dioxide
--Manuscript Draft--

Manuscript Number:

Article Type: Review Article

Keywords: Molten salt-based CO2 electrolysis; CO2 reduction; Carbon nanotubes; Chemical
vapor deposition; Single-walled carbon nanotubes

Corresponding Author: Chan Woo Lee


Kookmin University
KOREA, REPUBLIC OF

First Author: Herayati Herayati

Order of Authors: Herayati Herayati

Chan Woo Lee

Abstract: Molten salt-based electrolysis of CO2 to a valuable carbon product is a promising


method to overcome the excessive concentration of CO2 in the atmosphere. Carbon
nanotubes (CNTs) are one of the most favorable carbon nanomaterials. Recently,
scientists have developed CNT production method using molten salt-based CO2
electrolysis and optimized reaction conditions to obtain desirable CNT products.
However, to date, the production of single-walled CNTs (SWCNTs), as a type of CNTs
with excellent chemical and physical properties, has not been realized. Here, we
review the fundamental principles and process mechanisms of molten salt-based CO2
electrolysis. In addition, we cover the key components for improving CNT production
performance, such as cathode, anode, and electrolyte. Furthermore, we investigate the
key strategies applied in chemical vapor deposition to produce commercial SWCNTs
and suggest a perspective for SWCNT production from CO2 based on material
strategies. This paper provides deep insights into the CNT synthesis mechanism as
well as a value-addition strategy through cost-efficient and environment-friendly CO2
conversion processes.

Suggested Reviewers: Yun Jeong Hwang


Seoul National University
yjhwang1@snu.ac.kr

Jin Young Kim


Ulsan National Institute of Science and Technology
jykim@unist.ac.kr

Dong Wan Kim


Korea University
dwkim1@korea.ac.kr

Kyoung Suk Jin


Korea University
kysjin@korea.ac.kr

Opposed Reviewers:

Powered by Editorial Manager® and ProduXion Manager® from Aries Systems Corporation
Cover Letter

December 02, 2023

Dear Editor:

We are pleased to submit our manuscript entitled “Recent Advances and Perspectives in
Carbon Nanotube Production from The Electrochemical Conversion of Carbon Dioxide”
for publication in Journal of CO2 Utilization. Formatted: Subscript

The increase in the atmospheric level of CO2 contributes to global warming, which has several
adverse effects on the environment. Carbon capture and utilization (CCU) technology is the
technology to overcome this issue. This technology applies electrochemical conversion of CO2
to produce more valuable products. In some recent studies, this method has been successfully
employed to produce carbon nanotubes (CNTs). However, to date, the production of single-
walled CNTs (SWCNTs), as a type of CNTs with excellent chemical and physical properties,
using CCU technology has not been achieved yet. By summarizing recent advancements that
have been reported, we attempt to describe the fundamental principles and process mechanisms
of molten salt-based CO2 electrolysis. In addition, we cover the key components, such as
cathode, anode, and electrolyte, necessary to improve the CNT production performance.
Furthermore, we investigate the important strategies applied in chemical vapour deposition
(CVD) for the production of commercial SWCNTs.

To produce SWCNTs, some main components in electrochemical conversion process should


be controlled. In this review, we highlight the utilization of transition metal-based electrodes
in molten salt-based CO2 electrolysis. The general nature of transition metals with d-orbitals is
possibly to interact with 𝜋 electron clouds from CNTs. Therefore, transition metals can
strongly bind and nucleate the growth of CNTs. In addition, we investigate some computational
simulations by using density functional theory to predict the orbital interaction occurring
between transition metals and CNTs.

Another essential component is the electrolyte. The molten salt electrolyte composition
determines the effectiveness of CO2 capture and selectivity of electrochemical products. In the
most recent findings, borate ions have been found to assist the carbon dioxide reduction
reaction to selectively produce a high yield of carbon rather than other products. In this review
paper, we also explain in detail how the molten salt component regulates the CO 2 capture and
electrochemical conversion process to achieve cost and energy efficiency.
Fig. 1 An overview of the main subjects discussed in this review.

Elaborating on the key strategies applied in CVD, we suggest a perspective for SWCNT
production from CO2 based on material strategies. We believe this review provides deep
insights into the CNT synthesis mechanism and also value-addition strategies through cost-
efficient and environmentally friendly CO2 conversion processes. Therefore, we anticipate that
this manuscript will pique the interest of a broad readership in Journal of CO2 Utilization, Formatted: Subscript
including those from academia and industry.

Thank you for your consideration. We look forward to hearing from you.

Your sincerely,
Chan Woo Lee

Chan Woo Lee, Ph.D.


Assistant Professor
Department of Applied Chemistry
Kookmin University
Seoul 02707, South Korea
Tel : +82-2-910-4733, E-mail: cwlee1@kookmin.ac.kr (C.W. Lee)
Website: https://sites.google.com/view/cwlee1/
Google Scholar : https://scholar.google.com/citations?user=My89yV4AAAAJ&hl=ko
Highlights

Highlights:
 Recent progress of molten salt-based electrolysis of CO2 to Carbon Nanotubes has
been reviewed
 Key components, such as cathode, anode, and electrolyte, have been discussed.
 Key strategies applied in chemical vapor deposition to produce commercial SWCNTs
were summarized.
 Prospectives and future research of molten salt-based electrolysis of CO2 to Carbon
Nanotubes were proposed.
Graphical Abstract

Graphical Abstract:
Manuscript File Click here to view linked References

1 Recent Advances and Perspectives in Carbon Nanotube Production from


1
2
3 2 the Electrochemical Conversion of Carbon Dioxide
4
5
6 3
7
8
9
10
4 Authors: Herayati,a Chan Woo Leea,*
11
12 a
13 5 Department of Chemistry, Kookmin University, Seoul 02707, Korea.
14
15
16 6 *Corresponding author: Chan Woo Lee; Email: cwlee1@kookmin.ac.kr
17
18
19 7
20
21
22 8 Abstract:
23
24
25 9 Molten salt-based electrolysis of CO2 to a valuable carbon product is a promising
26
27
28 10 method to overcome the excessive concentration of CO2 in the atmosphere. Carbon nanotubes
29
30 11 (CNTs) are one of the most favorable carbon nanomaterials. Recently, scientists have
31
32
33
12 developed CNT production method using molten salt-based CO2 electrolysis and optimized
34
35 13 reaction conditions to obtain desirable CNT products. However, to date, the production of
36
37 14 single-walled CNTs (SWCNTs), as a type of CNTs with excellent chemical and physical
38
39
40 15 properties, has not been realized. Here, we review the fundamental principles and process
41
42 16 mechanisms of molten salt-based CO2 electrolysis. In addition, we cover the key components
43
44
45 17 for improving CNT production performance, such as cathode, anode, and electrolyte.
46
47 18 Furthermore, we investigate the key strategies applied in chemical vapor deposition to produce
48
49
50 19 commercial SWCNTs and suggest a perspective for SWCNT production from CO2 based on
51
52 20 material strategies. This paper provides deep insights into the CNT synthesis mechanism as
53
54
21 well as a value-addition strategy through cost-efficient and environment-friendly CO2
55
56
57 22 conversion processes.
58
59
60 23
61
62
63
1
64
65
24 Keywords: Molten salt-based CO2 electrolysis; CO2 reduction; Carbon nanotubes; Chemical
1
2 25 vapor deposition; Single-walled carbon nanotubes
3
4
5
6 26
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
61
62
63
2
64
65
27 1. Introduction
1
2 28 Every year, human activities release more carbon dioxide than natural processes can
3
4
5 29 absorb, leading to an increase in the amount of carbon dioxide in the atmosphere. In May 2023,
6
7 30 the global average of carbon dioxide reached 424 ppm [1], a new record in history, and it
8
9
10 31 continues to increase owing to the combustion of fossil fuels for energy. The increase in the
11
12 32 atmospheric level of CO2 can contribute to global warming, which may (i) accelerate sea-level
13
14
15 33 rise through the melting of polar ice caps and expansion of the space oceans, (ii) change rainfall
16
17 34 and salinity patterns, and (iii) change the intensity and frequency of tropical storms and
18
19
35 hurricanes. Incidentally, to prevent these climate change issues, global regulations on
20
21
22 36 greenhouse gas emissions have been progressively adopted since the Paris Agreement in 2015.
23
24 37 In this respect, many scientists have devoted considerable efforts toward developing some
25
26
27 38 technologies to overcome this problem. One way to reduce atmospheric CO2 is the use of
28
29 39 carbon capture and utilization (CCU) technologies. In the CCU process, CO2 is captured from
30
31
32 40 the atmosphere or high-concentration CO2 sources and subsequently converted into various
33
34 41 valuable products. Recently, advanced CCU methods have been reported for converting CO2
35
36
37
42 into valuable carbon compounds such as diamond [2], graphene [3], fullerene [4], carbon
38
39 43 nanotubes (CNTs) [5], and carbon nanofibers (CNFs) [6]. Among them, CNTs are one of the
40
41 44 most favorable products owing to their exceptional properties.
42
43
44 45 CNTs have excellent electrical, chemical, thermal, and mechanical properties. They
45
46 46 exhibit high thermal conductivity, large surface area, and high current density, which make
47
48
49 47 them suitable for a wide range of applications such as photovoltaic devices, sensors, transparent
50
51 48 electrodes, supercapacitors, and conducting composites [7]. Presently, CNTs are manufactured
52
53
54 49 using the chemical vapor deposition (CVD) technique using hydrocarbon precursors by a gas-
55
56 50 phase reaction. Several chemical and material principles have been applied in CVD to control
57
58
51 the nucleation and growth of CNTs and successfully produce single-walled CNTs (SWCNTs)
59
60
61
62
63
3
64
65
52 with superior chemical and physical properties [8]. Another affordable and promising method
1
2 53 for producing CNTs is the electrochemical conversion of CO2 by molten salt-based electrolysis.
3
4
5 54 In this method, CO2 is captured by a molten salt in carbonate form and then reduced to carbon
6
7 55 at the cathode [9]. Some scientists have developed this method and accomplished monumental
8
9
10 56 achievements to optimize shape selectivity, reaction rate, and energy efficiency in the
11
12 57 production of multiwalled carbon nanotubes (MWCNTs) [10–12]. However, to date, SWCNTs
13
14
15 58 have not yet been synthesized using CCU technology, although it represents a high-value
16
17 59 carbon product with much higher commercial value compared to MWCNTs [13].
18
19
60 Herein, we review the fundamental principles and process mechanisms of molten salt-based
20
21
22 61 CO2 electrolysis. Furthermore, we discuss the essential components such as cathode, anode,
23
24 62 and electrolyte, required to improve CNT production performance. In addition, we investigate
25
26
27 63 the key strategies applied in CVD for the commercial production of SWCNTs. One of the main
28
29 64 parameters is the use of supported transition metals (TM) catalysts. Several TMs have been
30
31
32 65 explored regarding their function as CNT growth catalysts. Moreover, we aim to elaborate these
33
34 66 principles with the CO2 electrochemical reaction for gaining new perspectives to develop a
35
36
37
67 synthesis method for CNTs using molten-salt CO2 electrolysis, which is believed to be a
38
39 68 potential solution to overcome the excessive CO2 concentration and realize a carbon-neutral
40
41 69 energy cycle.
42
43
44 70
45
46 71 2. Basic Principle
47
48
49 72 In the electrochemical reaction of gaseous CO2 to produce carbon, the electrodes are
50
51 73 immersed in molten salt. CO2 is then absorbed by the oxide ion contained in the molten salt
52
53
54 74 and converted into CO32− ions (equation (1))[14]. This process is called CO2 capture. The CO2
55
56 75 capture is followed by the next reaction step, namely, the reduction of carbonate ion[14]. It can
57
58
59
76 be transformed into C or CO at the cathode (equations (2a) and (2b)), releasing oxide ions.
60
61
62
63
4
64
65
77 The oxide ions are oxidized at the anode to evolve oxygen gas (equation (3)). The generated
1
2
78 oxide ions also react with CO2 in the melt to regenerate the carbonate ions, leading to a cycle
3
4
5 79 in which CO2 can be continuously absorbed and then split into carbon and oxygen.
6
7
8
80 CO2 + O2− → CO3 2− 1
9
10 81 CO3 2− + 4e− → C + 3O2− 2a
11
CO3 2− + 2e− → CO + 2O2−
12
13
82 2b
14
15 83 2O2− → O2 + 4e− 3
16
17
18
84
19
20 85 According to the classical acid–base theory, O2− can be defined as a base and CO2 is
21
22 86 known as a Lewis acid. According to equations (2a) and (2b), the reduction potential of CO32−
23
24
25 87 is associated with the activity of O2− or alkalinity [15,16]. Fig. 1a–c show that the reduction
26
27 88 potentials of CO32− to form either C or CO both shift positively at a given temperature as the
28
29
30 89 alkalinity is reduced. Specifically, the carbon product derived from CO32− is predominant in
31
32 90 the relatively high alkalinity region, while the CO product derived from CO32− is predominant
33
34
35 91 in the relatively low alkalinity region, suggesting that product selectivity would change
36
37 92 significantly on reducing the activity of oxide ions. These observations further indicate that CO
38
39
40
93 can be preferentially obtained even at a relatively low temperature (e.g., 550 °C) if the
41
42 94 electrolyte alkalinity is low enough. Furthermore, the generation of CO is favored by an
43
44 95 increase in temperature because a larger CO-dominant region can be observed at higher
45
46
47 96 temperatures (Fig. 1d), which is consistent with the findings of previous works [17,18].
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
5
64
65
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 97
28
29 98 Fig. 1 Standard relationships (E–pO) between reduction potential (E) and activity of oxygen
30
31
32 99 anion (pO) under different conditions. Different CO pressures at (a) 550 °C, (b) 600 °C, (c)
33
34 100 and 650 °C and (d) under a CO pressure of 10−1 bar at different temperatures. Reproduced
35
36
37 101 with permission from ref. [19]; copyright 2020, Elsevier.
38
39
40 102
41
42 103 Multicompound eutectic salts, which are characterized by a high latent heat of fusion,
43
44 104 high heat capacity, wide working temperature range, and no chemical reaction between the
45
46
47 105 components, are utilized in this electrolysis system [20]. These mixtures possess a specific
48
49 106 eutectic temperature, the minimum operation temperature, which is influenced by the type of
50
51
52 107 molten salt. The melting point of molten salt mixtures depends on the composition of the
53
54 108 molten salts. Many works have discussed the thermodynamic model to optimize the eutectic
55
56
57
109 mixture. Calculation of Phase Diagram (CALPHAD) is used to predict some eutectic mixtures.
58
59 110 This method is based on the thermodynamic evaluation and optimization of mixture
60
61
62
63
6
64
65
111 subsystems [21,22]. For example, the calculated eutectic points of the LiF–LiCl, LiF–Li2CO3,
1
2 112 and LiCl–Li2CO3 binary systems (
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
35 113
36
37 114 Fig. 2) are located at 772, 886, and 789 K, respectively. The phase transition temperature is
38
39
40 115 influenced by electrostatic and van der Waals interactions [23,24]. Changing the anion will
41
42 116 change the ionic interactions. Therefore, the eutectic temperature increases with an increase in
43
44
45
117 the ionic interaction strength [25]. Meanwhile, the predicted eutectic point of the LiF–LiCl–
46
47 118 Li2CO3 ternary system is 735 K with XLiF = 28.41 mol%, XLiCl = 55.09 mol%, and XLi2CO3 =
48
49 119 16.50 mol%. The melting point can be shifted by changing the concentration ratio.
50
51
52
53
54
55
56
57
58
59
60
61
62
63
7
64
65
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
30 120
31
32
121 Fig. 2 Calculated phase diagram of binary (a) Li2CO3–LiCl, (b) Li2CO3–LiF, (c) LiCl–LiF and
33
34
35 122 (d) ternary LiCl–LiF–Li2CO3. Reproduced with permission from ref. [20]; copyright 2022,
36
37 123 Elsevier.
38
39
40 124
41
42 125
43
44
45 126 Table 1 Recent progress of molten salt-based CO2 electrolysis. Reproduced with permission
46
47 127 from ref. [26]; copyright 2019, Elsevier.
48
49
50 Electrolyte T (°C) Cathode Anode Curre Current Products Referen
51 nt efficien ce
52 densit cy
53 y
54 (A/cm2
55
56 )
57 LiCl–KCl– 450 Al Glass Carbon [27]
58 K2CO3 carbon film
59
60
61
62
63
8
64
65
Li2CO3– 450 Ni, glass Au or C Carbon [28]
1 Na2CO3– carbon Nano
2
K2CO3 Powders
3
4 (CNPs),
5 CNFs
6 NaCl–KCl, 550– Pt Glassy 0.005– CNTs [29]
7 NaCl– 750 carbon 0.2
8
9 KCl–CsCl
10 Li2CO3 750– Pt Ni 0.1– Carbon [30]
11 950 0.5 powder
12 Li2CO3 900 Ti Graphite 0.1 Close to CO [17]
13
14
100%
15 LiCl–Li2O, 650/90 SS C 0.075– 80%– CNTs [31]
16 CaCl2– 0 0.5 90%
17 CaO
18
19
Li2CO3– 450– Ni SnO2 0.1– 70%– Carbon [32]
20 Na2CO3– 650 2.1 90% powder
21 K2CO3
22 LiCl–KCl– 700 304 C 0.03 Cr–O–C [33]
23 K2CO3 stainless layer
24
25 steel (SS)
26 LiF–NaF– 750 Ni Graphite 0.16– Carbon [34]
27 Li2CO3 0.215 powder
28 CaCl2– 500– Ni, Pt; SnO2, 1.0–4 95% Carbon [9]
29
30 CaCO3– 800 mild steel graphite
31 LiCl–KCl,
32 Li2CO3–
33 K2CO3
34
35
Li2CO3– 540– Mild SS 1–6 65%– CNPs [35]
36 K2CO3 700 steel 100%
37 LiCl– 700 W Pt 0.3– Carbon [36]
38 Li2CO3 0.4 powder
39
40
CaCl2– 850 W TiO2.Ru 1.1 Graphite [37]
41 NaCl–CaO O2 sheet,
42 CNTs
43 CaCl2– 900 Graphite Graphite 1.25 CNPs, [38]
44 CaO CNTs
45
46 CaCl2– 900 SS C 0.1– 37% CO, C [39]
47 CaO 0.2 (for CO)
48 Li2CO3 700– Galvaniz Ir, Ni 1 80%– CNFs, [12,40]
49 800 ed steel 100% CNTs
50
51 Li2CO3– 577– Fe Ni–Cr 0.2– 70%– Honeycom [41]
52 Na2CO3– 677 0.4 80% b-like
53 K2CO3 carbon,
54 CNTs
55
56
LiCl–Li2O 630 Ti Pt 0.05– Carbon [42]
57 0.4 film
58
59
60
61
62
63
9
64
65
LiCl–KCl– 550 Ni Graphite Carbon [43]
1 CaCl2– film
2
CaC2
3
4 CaCl2– 600 Mo (1- 0.36– Carbon [44]
5 LiCl–CaO x)CaTiO 0.5 powder
6 3–xNi
7 LiCl– 650– W Pt 0.1– 50%– Carbon [45]
8
9 NaCl– 800 1.5 90% powder,
10 Na2CO3 CNTs, CO
11 CaCl2– 750 SS, Cu, TiO2.Ru 1.5 - Graphene [46]
12 NaCl–CaO Ni O2
13
14 Li2CO3– 450 Ni Graphite 0.025– 65%– CNPs, [47]
15 K2CO3– 0.1 90% carbon
16 Na2CO3, nanosheet
17 LiCl–KCl–
18
19
CaCO3,LiC
20 l–KCl-
21 Li2CO3
22 Binary or 600– Fe Ni 0.05– 60%– Carbon [48]
23 ternary 750 0.2 95% flakes
24
25 carbonates
26 Li2CO3 750 SS Ni 0.05– 90% Carbon [11]
27 (Al2O3 0.1 nanotubes
28 coating)
29
30 Binary or 500– Graphite, Graphite 0.09– >65% Carbon [49]
31 ternary 700 Au, Cu, 1.2 sheet,
32 carbonates Ti, SS carbon
33 nanoparticl
34 es
35
36 Li2CO3 770 Galvaniz Ni–Cr 1 CNTs [10]
37 ed Steel
38 LiCl– 550 Ni Graphite 0.035 93.8% CNFs [50]
39 Li2CO3–
40
41 LiBO2
42 NaCl– 750 C, Cu, Graphite 2 51%– Ge@CNT [51]
43 CaCl2– Ni, Fe 78%
44 CaO–GeO2
45
46 CaCl2– 800 NA Graphite NA 85.1% Sn@CNT [52]
47 NaCl–
48 CaO–SnO2
49 Li2CO3– 770 Inconel Muntz 0.08 98% MWCNT [53]
50
51
Li2O + 625 (58% Brass (41 layers)
52 0.1% Ni, 5% (60%
53 Fe2O3 Fe, 20% Cu, 40%
54 Cr, 8% Zn)
55 Mo,
56
57 4.15%
58 Tb)
59 Li2CO3– 650 Pt Ni 0.2 C or CO [54]
60 Na2CO3–
61
62
63
10
64
65
K2CO3–
1 BO33−
2
3
Na2CO3– 550 Fe–Ni Ni alloy 0.1 C or CO [55]
4 K2CO3 GDE
5 NaCl–KCl
6 128
7
8
9 129
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
61
62
63
11
64
65
130 3. Cathodes
1
2 131 Various transition metal (TMs) and alloys have been explored as cathode materials
3
4
5 132 (Table 1). TM catalysts have a different binding affinity for carbon, which determines the type
6
7 133 of formed carbon nanostructure [56]. In an investigation by Dey et al., density functional theory
8
9
10 134 (DFT) calculation showed that the π–d bonding between the nanotube and metal resulted in the
11
12 135 formation of a strong TM–C bond at the interface of the nanotube growth [56]. For example,
13
14
15 136 the ab initio molecular dynamics (MD) of SWCNTs adsorbed onto 13-atom Ni and Zn clusters
16
17 137 (
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
45 138
46
47 139 Fig. 3a-b) showed that the nanotubes remained strongly adsorbed onto the Ni cluster at the end
48
49
50 140 of 0.30 ps, while in the same duration they dissociated into individual atoms over a Zn cluster.
51
52 141 This is further evidence that the Ni cluster exhibits a stronger adsorption than the Zn cluster.
53
54
55 142 The energy of the Ni atom/cluster adsorption to the nanotube is stronger than that of Zn
56
57 143 atom/cluster owing to the overlap bonding between the vacant d orbital in the Ni (d7) and π
58
59
60 144 electron cloud in the SWCNT forming a dative covalent bond between them. This dative bond
61
62
63
12
64
65
145 is absent in the case of Zn, as it has fully occupied d10 orbitals. Hence, the adhesion energy is
1
2 146 weak. Meanwhile, in an alloying metal comprising Ni–Zn atoms, as the number of Ni atoms is
3
4
5 147 increased, the adhesion energy also increases, modifying the CNT production ability.
6
7 148
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
61
62
63
13
64
65
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
26 149
27
28 150 Fig. 3 (a) 0.30-ps MD simulation of (5,0) CNT adsorbed onto a 13-atom Ni cluster and (b)
29
30 151 0.25-ps MD simulation of (5,0) CNT adsorbed onto a 13-atom Zn cluster (reproduced with
31
32
33 152 permission from ref. [56]; copyright 2011, RSC). (c) Calculated adhesion energies per bond
34
35 153 for the first three TM rows. The shaded region marks the “Goldilocks zone” (reproduced with
36
37
38 154 permission from ref. [57]; copyright 2015, RSC). (d) Plots of highest occupied molecular
39
40 155 orbital (HOMO) and lowest unoccupied molecular orbital (LUMO) of the binding complexes
41
42
43
156 of various TM-decorated SWCNTs: (1) Fe-, (2) Ru-, (3) Os-, (4) Ni-, (5) Pd- and Pt-SWCNT
44
45 157 (reproduced with permission from ref. [58]; copyright 2013, Elsevier).
46
47 158
48
49
50 159 Silvearv et al. studied a wide range of TMs using DFT calculation with respect to the
51
52 160 carbon–metal adhesion energy per bond [57]. TMs for successful CNT production were
53
54
55 161 proposed according to the calculation result. For a metal to be catalytically active, it must form
56
57 162 particles that are able to fulfill three key parameters: (i) decompose the carbon feedstock gas,
58
59
60 163 (ii) form graphitic caps at their surface, and (iii) maintain the CNT hollow structure by
61
62
63
14
64
65
164 stabilizing the growing end [59,60]. To satisfy criterion (iii), the metal–carbon bonds must be
1
2 165 strong enough to allow the dissociation of the catalytic metal particle and CNT [59]. However,
3
4
5 166 too strong a metal–carbon bonding leads to the formation of metal carbides (which occurs for
6
7 167 Mo and W) [59]. Moreover, too weak a metal–carbon bond cannot stabilize the hollow structure
8
9
10 168 (which is the case for Cu, Au, and Pd) [59]. The medium adhesion energy is indicated by the
11
12 169 shaded region in
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
40 170
41
42
43
171 Fig. 3c, which is called the “Goldilocks zone.” In the first row of TMs, there are well-known
44
45 172 catalyst metals, namely, Fe, Co, and Ni, with desirable adhesion energies per M–C bond of
46
47 173 (−2.56/−2.78 eV), (−2.63/−2.93 eV), and (−2.65/−2.79 eV), respectively. Candidates for other
48
49
50 174 catalyst metals should be in the same energy range or at least close to it. Thus, other elements
51
52 175 such as Y, Zr, Rh, Pd, La, Ce, and Pt have promising adhesion energies for CNT growth.
53
54
55 176
56
57
58
59
60
61
62
63
15
64
65
177 Table 2 Binding energy and energy gaps of various TM-CNTs. Reproduced with permission
1
2 178 from ref. [58], copyright 2013, Elsevier Limited.
3
4
5 SWCNT + TM  TM-SWCNT ΔEZPE (kcal mol−1) Energy gaps (eV)
6 Fe-SWCNT −187.46 1.606
7 Rh-SWCNT −162.23 1.769
8
9 Pt-SWCNT −145.72 1.660
10 Ni-SWCNT −135.76 1.551
11 Co-SWCNT −125.37 1.905
12
Pd-SWCNT −86.05 1.606
13
14 179
15
16 180 The energetics and thermodynamic properties of the binding reaction of TM atoms
17
18
19 181 deposited on SWCNTs were reported by Tabtimsai et al [58]. The binding abilities of TMs onto
20
21 182 SWCNTs are in the order Fe > Rh > Pt > Ni > Co > Pd; their binding energies (ΔEZPE) are
22
23
24
183 given in Table 2. The overlapping orbital between the TM and SWCNT is shown in
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
52 184
53
54 185 Fig. 3d. The highest occupied molecular orbitals (HOMOs) and lowest unoccupied molecular
55
56 186 orbitals (LUMOs) of the Ru-, Os-, Ni-, Pd-, and Pt-decorated SWCNTs are delocalized over
57
58
59 187 metal atoms. For the Fe-decorated SWCNT, its HOMO is delocalized over C atoms, which are
60
61
62
63
16
64
65
188 placed at the opposite Fe-binding site. Thus, the Fe-decorated SWCNT is the most active
1
2 189 compound. Fe-SWCNTs has lower energy gaps compared to other TM-SWCNTs (Table 2).
3
4
5 190 The smaller the energy gap, the more likely it is for Fe to participate in electron transfer to
6
7 191 CNTs, which can affect its binding stability with CNTs [61].
8
9
10 192 Novoselova et al. initiated CNT synthesis using an electrolytic method at TM cathode
11 193 in the halide (Na–K/Na–K–Cs) molten salts at a temperature of 750 °C under an excess
12 194 pressure up to 15 bar. Pt and glassy-carbon electrodes were used as the cathode and anode,
13
14 195 respectively [29]. The use of Pt as an electrode resulted in the successful production of CNTs.
15 196 Typical morphologies of the nanotubes and nanoparticles were observed by transmission
16 197 electron microscopy (TEM). The patterns showed that 40% of the products produced in the
17
18 198 NaCl–KCl melt were CNTs, the majority of which were multiwalled and had a curved form
19 199 with structural defects (
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42 200
43 201 Fig. 4a). The CNTs frequently agglomerated into bundles, and more rarely they were
44
45
46 202 arranged as individual tubes. The outer diameter of CNTs varied from 5 to 250 nm, while the
47
48 203 internal diameter varied from 2 to 140 nm. However, owing to cost issues, the use of Pt as a
49
50
51 204 cathode is limited.
52
53 205
54
55
56
57
58
59
60
61
62
63
17
64
65
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21 206
22
23
24
207 Fig. 4 (a) TEM image of the carbon product obtained at the Pt cathode in molten NaCl–KCl.
25
26 208 Reproduced with permission from ref. [29], copyright 2008, Elsevier. (b) Scanning electron
27
28 209 microscopy (SEM) image of carbon products obtained and (c) product distributions under
29
30
31 210 various conditions at the Ni cathode in molten LiCl–Li2CO3–LiBO2. Reproduced with
32
33 211 permission from ref. [50], copyright 2020, Elsevier. (d) SEM image of carbon products
34
35
36 212 obtained at the Cu cathode, (e) Monel cathode, and (f) Ni–Cr cathode in molten Li2CO3.
37
38 213 Reproduced with permission from ref. [62], copyright 2017, Elsevier.
39
40
41 214
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
18
64
65
215 After the abovementioned development, several studies to optimize the properties of
1 216 CNTs have been reported. Hu et al. reported that carbon nanofibers (CNFs) (
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24 217
25
26 218 Fig. 4b) with a diameter of ~80 nm were successfully obtained at the Ni cathode from
27 219 CO2 electrolysis in LiCl–Li2CO3–LiBO2 molten salts at 550 °C [50]. This method showed a
28 220 selectivity of more than 90% (
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51 221
52
53 222 Fig. 4c) in producing highly crystalline CNF. Borate-assisted cathodic reactions, which
54
55 223 will be discussed in detail in Section 5, can act as a controller of electrolyte alkalinity to buffer
56
57
58 224 the concentration of O2− generated during the CO2 reduction reaction, positively shifting the
59
60 225 reduction potential of the captured CO2 and concurrently extending the product species.
61
62
63
19
64
65
226 However, at slightly higher temperatures of 650 °C, the formation of CO is more favorable. It
1
2 227 is consistent with the standard relationship between CO32− reduction potential and oxygen ion
3
4
5 228 activity as explained in the E-pO diagram in Section 2.
6
7 229 Johnson et al. reported that a high yield of CNTs could also be produced under
8
9 230 electrolysis at 0.2 A cm−2 for 1.5 h in Li2CO3 at 770 °C when a TM alloy such as Monel (Ni–
10 231 Cu alloy), Cu, or Ni–Cr alloy was used as the cathode [62]. However, different cathodes
11 232 showed different carbon morphologies. When a Cu cathode was used, thin tangled CNTs (
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34 233
35
36 234 Fig. 4d) were formed, while the use of Monel or Ni–Cr cathodes resulted in uniform,
37 235 thicker, and straight CNTs (
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 236
61
62
63
20
64
65
237 Fig. 4e-f) [63]. Initially, a thin carbon coating formed on the Monel cathode substrate
1
2 238 surface. As delineated, (i) certain combinations of TM (contained in Monel cathodes) catalysts
3
4
5 239 lead to CNT growth and repair that are 10–100-fold better compared to single metals alone, (ii)
6
7 240 Cu-containing cathodes can improve the uniformity of the base graphene layer to initiate CNT
8
9
10 241 growth, and (iii) the CNT growth may be dominated by tip growth from a catalyst that
11
12 242 maximizes exposure to the bulk electrolyte. This study suggests that TM alloys are better as
13
14
15 243 CNT catalysts than pure TMs.
16
17 244 Another TM alloy, namely, Zn-coated steel, was also reported to successfully produce
18
19
245 CNFs through a different mechanism [40]. The high-crystallinity CNFs (
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
50 246
51
52 247 Fig. 5) were prepared by electrolysis using a galvanized-steel-wire cathode and an oxygen-
53
54
248 generating nickel anode in molten Li2CO3 at 730 °C, initiated by a low current of 0.05 A,
55
56
57 249 followed by constant-current electrolysis at 1 A for 4 h. The faradaic efficiency was over 80%.
58
59
60
61
62
63
21
64
65
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
28 250
29
30 251 Fig. 5 (a) Calculated electrolysis potentials relevant to the high-yield electrolytic formation of
31
32
33
252 CNTs from molten carbonates. (b) Electrolytic CNT growth mechanism; i: high-voltage
34
35 253 electrolysis of alkali metals followed by exfoliation and alkali intercalation into a solid carbon
36
37 254 electrode, ii: spontaneous absorption of CO2 into molten carbonate at a low voltage, iii: lower-
38
39
40 255 voltage reduction of lithium carbonate without a catalyst, iv: metal catalyst nucleation of an
41
42 256 assortment of carbons (including CNTs) at the cathode, v: Zn-metal-assisted deposition of
43
44
45 257 catalyst and initial carbon to generate a high yield of electrolytic CNTs. (c) SEM image of
46
47 258 CNFs. The red arrows in the SEM image indicate typical Ni nucleation sites. The blue arrow
48
49
50 259 originates at one Ni site and moves along the CNF path. (d) Energy-dispersive X-ray
51
52 260 spectroscopy (EDS) composition mapping along the 6-μm blue arrow path shown in SEM.
53
54
261 Reproduced with permission from ref. [40]; copyright 2022, American Chemical Society.
55
56
57 262
58
59
60
61
62
63
22
64
65
263 Zn is beneficial for the formation of high-yield CNFs. As revealed by the calculated
1
2 264 electrolysis potentials in
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
33 265
34
35 266 Fig. 5a, Zn has the highest electrolysis energy compared to Ni, Cu, Co, or Fe in terms of the
36
37 267 energy required to form Zn metal from Zn ion. This energy is also greater than the energy
38
39
40 268 required to reduce CO32− to solid carbon. The existence of Zn metal is advantageous because
41
42 269 it is energetically sufficient to activate both (i) the spontaneous formation of solid carbon from
43
44
45 270 carbonate and (ii) the spontaneous formation of metal catalyst nuclei that assist the initiation
46
47 271 of the controlled growth of nanofibers at the nucleation site (
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
23
64
65
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
28 272
29
30 273 Fig. 5b – part v). Ren et al. reported that the oxygen evolution at the anode leads to Ni erosion.
31
32
33
274 Ni anode was dissolved in the molten salt because galvanic exchange reaction occurred
34
35 275 between the oxide ion and Ni atom. When the Ni anode is kept in contact with oxide ions in
36
37 276 molten salt, the Ni atoms from the anode surface can be quickly oxidized and dissolved in the
38
39
40 277 molten salt [64]. The Ni metal is deposited again on the cathode and acts as a nucleation site
41
42 278 for CNTs. The Ni metal deposition is preferable because it has lower electrolysis energy than
43
44
45 279 Zn. This is also evidenced by the negative electrolysis potential of Zn-assisted deposition of Ni
46
47 280 (
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
24
64
65
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
28 281
29
30 282 Fig. 5b – part v). Thus, Zn metal is known to be a critical activator. When Zn is absent from
31
32
33
283 the cathode, carbon formation initiates only at much higher potentials (
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
63
25
64
65
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
28 284
29
30 285 Fig. 5b – part iv), leading to the formation of nucleation sites at higher current densities, which
31
32
33
286 are not conducive to confined CNF growth, resulting in the observed profusion of mixed,
34
35 287 amorphous graphite, graphene, and carbon nanostructures of various shapes.
36
37 288 The electrolysis product shown in the SEM
38
39
40 289 of
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
26
64
65
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
28 290
29
30 291 Fig. 5c comprises controlled carbon fibers with metal nucleation points (bright spots). Most of
31
32
33
292 the Ni nanoparticles in the SEM are located at nanofiber tips. The characteristic CNF structure
34
35 293 was observed when the electrolysis was initiated at a gradually increasing current density or
36
37 294 with an initial low current (5 mA cm−2 for 1 h at the cathode) followed by an extended high-
38
39
40 295 current electrolysis such as at 100 mA cm−2 (for several hours). The linear EDS map in
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
27
64
65
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
28 296
29
30 297 Fig. 5d shows elemental variation along the 6-μm path of the EDS scan from pure Ni at the
31
32
33
298 beginning of the fiber to pure carbon along the remainder of the fiber. However, when the
34
35 299 electrolysis starts directly at only a high current density (100 mA cm−2), the cathode product is
36
37 300 principally amorphous (and only ~25% CNF). Ren et al. interpreted this mechanistically as
38
39
40 301 follows: owing to its low solubility and lower reduction potential, Ni (in this case, originating
41
42 302 from the anode) is preferentially deposited at low applied electrolysis currents (5 or 10 mA
43
44
45 303 cm−2). The high concentration of electrolytic [CO32−] ≫ [Ni2+] and mass diffusion dictates that
46
47 304 higher currents will be dominated by carbonate reduction, leading to CNF deposition [30].
48
49
50 305
51
52
53
54
55
56
57
58
59
60
61
62
63
28
64
65
1
2
3
4
5
6
7
8
9
10
11 306
12 307 Fig. 6 SEM elemental maps showing evolution of catalysts (a,c) as-deposited and (b,d) after
13
14 308 15 min of heating in the molten electrolyte; (a,b) show 0.5-nm Fe catalysts and (c,d) show 5-
15
16
17 309 nm Fe catalysts. (e) Scanning TEM (STEM) elemental map of 0.5-nm Fe-catalyst-grown CNT
18
19 310 showing the Fe catalyst, (f) galvanostatic electrolysis plot of potential vs. time with a schematic
20
21
22 311 illustration inlay of the proposed mechanism of catalyst formation and reduction. Reproduced
23
24 312 with permission from ref. [11]; copyright 2018, American Chemical Society.
25
26
313
27
28
29 314 Douglas et al. conducted a study to elucidate mechanisms that control electrochemical
30
31 315 CNT growth toward small-diameter CNT. In their study, the Fe catalyst layers were deposited
32
33
34 316 at different thicknesses (0.5–5 nm) on a stainless-steel (SS) cathode, which was paired with an
35
36 317 Al2O3-coated Ni anode [11]. The electrochemical reaction was conducted in lithium carbonate
37
38
39 318 molten salt, which was held at 750 °C. A CNT-coated cathode was formed after 60 min of
40
41 319 electrochemical growth. This duration of growth yielded ~50 mg of carbon deposited across an
42
43
44 320 electrode with a total area of 2.5 cm2, achieved under a constant current of 100 mA/cm2 (~90%
45
46 321 faradaic efficiency) [11]. This approach allowed the adjustment of the size of the catalyst
47
48 322 according to the thickness of the deposited Fe. As shown in Fig. 6a–e, SEM elemental analysis
49
50
51 323 was employed to study the evolution of the cathode surfaces. The as-deposited Fe thin films on
52
53 324 SS surfaces before and after prenucleation heat treatment to 750 °C in Li2CO3 for 15 min were
54
55
56 325 compared. As shown in Fig. 6a, the 5-nm Fe film appears uniformly distributed on the SS
57
58 326 surface until heated in the electrolyte. After heating, the film forms larger particles (Fig. 6b).
59
60
61
62
63
29
64
65
327 However, the 0.5-nm Fe film shows the formation of smaller well-distributed particles (Fig.
1
2 328 6c-d).
3
4
5 329 Scanning TEM energy-dispersive X-ray spectroscopy (STEM-EDS) mapping was
6
7 330 performed to confirm the catalyst formation prior to CNT nucleation and growth. The STEM-
8
9
10 331 EDS mapping analysis (Fig. 6e) showed that Fe remained embedded inside a CNT tip grown
11
12 332 from the catalyst layer. It confirmed that the growth of CNTs was catalyzed by Fe contained in
13
14
15 333 bulk substrates without any contribution of predeposited metal from the electrolyte or anode
16
17 334 owing to the inert nature of the passivated Ni anode [11]. The reaction mechanism was analyzed
18
19
335 by galvanostatic measurement. The galvanostatic electrolysis data presented in Fig. 6f reveal
20
21
22 336 two distinct plateaus, including the first at ~1.2 V for 6 s and then a higher potential plateau at
23
24 337 ~2.0 V, which extended for the duration of the hour-long electrolysis. The low-voltage initial
25
26
27 338 plateau was attributed to the reduction of the catalyst. This catalyst reduction was then followed
28
29 339 by CNT deposition at a voltage of 2.0 V [60].
30
31
32 340
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
63
30
64
65
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
39 341
40
41 342 Fig. 7 (a) Size distribution of CNTs grown from 0.5-nm Fe under varying growth times, (b–d)
42
43
44 343 representative SEM images of 0.5-nm Fe-grown CNTs at each growth time, (e) size
45
46 344 distributions of CNTs at the same growth time from a 5-nm Fe film, (f–h) representative SEM
47
48
49
345 images for 5-nm Fe-grown CNTs at each growth time, (i) representative Raman spectra with
50
51 346 fits shown by solid lines and raw data shown by open circles, and (j) ID/IG ratios as a function
52
53 347 of Fe catalyst thickness. Reproduced with permission from ref. [11]; copyright 2018, American
54
55
56 348 Chemical Society.
57
58 349
59
60
61
62
63
31
64
65
350 To understand the effect of the dynamic processes between the catalyst reduction (at
1 351 the first few seconds) and the total synthesis time of 60 min, time-stop experiments for
2
3 352 growth times of 3, 10, and 30 min were performed. The CNTs produced under these
4 353 conditions were compared with CNTs grown for 1 h. The CNT diameter distributions and
5 354 representative SEM images for each growth time and thickness are shown in
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
46 355
47 356 Fig. 7a–h. The smaller diameter distribution of CNTs was found to be correlated with
48
49
50 357 shorter growth times. These diameter distributions became wider under longer growth times.
51
52 358 When the growth time is extended, the particles have more time to interact, collide, and adhere
53
54
55 359 to each other. However, the rate of diffusion, which is the process of particles moving from
56
57 360 regions of higher to lower concentration, may not be sufficient to counteract the increased rate
58
59
60 361 of aggregation over extended growth time. Therefore, the key feature for achieving controlled
61
62
63
32
64
65
362 CNT growth has been linked to the fast reduction of the catalyst layer to produce the smallest
1
2 363 possible catalyst particles. It can be achieved only by adopting a thinner catalyst layer.
3
4
5 364 Raman spectroscopy (
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
45 365
46
47 366 Fig. 7i) was used to characterize the MWCNT products, which exhibited the characteristic
48 367 graphitized carbon G-peak at ~1580 cm−1, indicative of in-plane sp2-hybridized carbons, and
49
50 368 a D mode at ~1350 cm−1, which corresponded to out-of-plane defective sp3-hybridized
51 369 carbons. A higher D/G intensity ratio was observed for CNTs grown from thicker Fe catalyst
52 370 layers. This observation indicated a greater concentration of sp3 carbon materials in the
53
54 371 samples produced from the thicker Fe catalyst layer. The trend of the D/G intensity ratio
55
56
57
58
59
60
61
62
63
33
64
65
372 (ID/IG) as a function of the Fe thickness (
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
40 373
41
42 374 Fig. 7j) indicated that thicker Fe layers yielded larger-diameter CNTs with a larger defect
43
44
45
375 concentration [11]. Larger catalyst particles may lead to slower diffusion rates of reactants to
46
47 376 the active sites on the catalyst surface [66]. This reduced rate of diffusion can result in less-
48
49 377 ordered or less-crystalline CNTs because the carbon may not have sufficient exposure to the
50
51
52 378 active sites [67,68].
53
54
55
56
57
58
59
60
61
62
63
34
64
65
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
39 379
40
41
42
380 Fig. 8 Schematic illustration of the Fe–Ni gas diffusion CO2 electrode: (a) Digital photos of the
43
44 381 Fe–Ni gas diffusion electrode, (b) schematic diagram of CO2 direct reduction with the Fe–Ni
45
46 382 gas diffusion electrode in molten salt, (c) SEM images with EDS mappings of the carbon
47
48
49 383 products in molten Na2CO3–K2CO3, (d) SEM image with EDS mapping of the carbon products
50
51 384 in molten NaCl–KCl. Reproduced with permission from ref. [55]; copyright 2023, Elsevier.
52
53
54 385
55
56 386 In some recent reports, the use of gas diffusion electrodes has become an established
57
58
387 method for mitigating problems caused by the limited solubility of CO2 in electrolytes
59 388 [69,70]. Shi et al. reported the use of a porous Fe–Ni gas diffusion electrode to realize direct
60
61
62
63
35
64
65
389 CO2 electrochemical reduction in molten Na2CO3–K2CO3 (
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
41 390
42
391 Fig. 8a), whereby the enhanced contact between CO2 and the porous electrode
43
44
45 392 overcame the low CO2 solubility and facilitated the thermodynamically favorable reduction
46
47 393 rather than the reduction of CO32− and metal cations [55]. In the narrow pores of the Fe–Ni gas
48
49
50 394 diffusion electrode, CO2 migrated to the cathode quickly for direct reduction through the thin
51
52 395 electrolyte. By using this electrode, CO2 mass transfer distances to nearby electrode surfaces
53
54
55 396 were greatly reduced.
56
57
58
59
60
61
62
63
36
64
65
397 During a 10-min electrolysis (
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
41 398
42
399 Fig. 8b), the gas diffusion electrode with CO2 bubbling had a potential ~0.13 V higher
43
44 400 than that with Ar bubbling at an apparent current density of 100 mA cm−2. The potential
45 401 difference between the thermodynamic analysis of the direct reduction of CO2 and the
46 402 reduction of Na+ was extremely close to the theoretical value of 0.16 V. Thus, under CO2
47
48 403 bubbling, CO2 reduction and carbon deposition are thermodynamically more favorable,
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
37
64
65
404 which is confirmed by the SEM images shown in
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
41 405
42
406 Fig. 8c. A dense layer of carbon was obtained at the Fe–Ni gas diffusion cathode. The
43
44 407 morphology of products was dominated by flaky and amorphous carbon with a current
45 408 efficiency of 54.55%. The direct CO2 reduction to produce carbon was also verified in molten
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
38
64
65
409 NaCl–KCl using the same porous electrode with the same morphology of carbon products (
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
41 410
42
411 Fig. 8d). In combination with a low-cost Ni10Cu11Fe oxygen-evolution inert anode, the
43
44
45 412 molten-salt electrolyzer converted CO2 into C, CO, and O2. In addition, the use of molten
46
47 413 carbonate free of lithium and alkaline earth metal cations reduces the cost of electrolytes and
48
49
50 414 simplifies product separation. Thus, the reduction path and electrolyte selection range can be
51
52 415 regulated by mediating the electrode architecture, which is meaningful for further engineering
53
54
55 416 of optimal CO2 electrolyzers for achieving a negative-carbon technology with lower energy
56
57 417 consumption. However, further research is needed, particularly to increase selectivity in CNT
58
59
60
418 production.
61
62
63
39
64
65
419
1
2 420 4. Anode
3
4
5 421 Another crucial component in the electrochemical reaction of CO2 in molten salt is
6
7 422 anode. The anode promotes oxygen generation from accumulated oxide ions. Carbon anodes
8
9
10 423 have been used in electrochemical experiments owing to their good thermostability and
11
12 424 electrical conductivity [71]. However, in molten salt-based CO2 electrolysis, the released
13
14
15 425 oxygen ions react with the graphite anode to produce CO and CO2, resulting in a decrease in
16
17 426 the electrode size, drop in current efficiency, and an increase in energy consumption [72]. Thus,
18
19
427 novel non-carbon inert anodes have been explored. Physical stability at working temperature;
20
21
22 428 electrical conductivity; resistance to attack by molten salts, chlorine and oxygen gases, and
23
24 429 chloride and oxide ions; and resistance to thermal shock are all desirable characteristics of such
25
26
27 430 inert anodes [73]. Three general types of materials have been investigated as anodic materials:
28
29 431 pure metals, alloys, and oxides.
30
31
32 432 Among metals, Pt is known to be one of the most stable anodes (Fig. 9). In LiCl–Li2O
33
34 433 molten salts, on the surface of the Pt anode, a Li2PtO3 film is observed, indicating the
35
36
37
434 anticorrosion ability of the Pt electrode [74]. However, the high cost of Pt anodes restricts their
38
39 435 use. Apparently, covering a noble metal with a good corrosion-resistant film on the substrate
40
41 436 anode is an excellent technique to cut costs. Stable anode metals or other materials are often
42
43
44 437 coated on the conductive substrate to form a coated inert anode. For example, the Pt–Ti
45
46 438 composite inert anode which demonstrated good catalytic performance and high endurance in
47
48
49 439 Li2CO3–Na2CO3–K2CO3 molten salts at temperatures less than 500 °C [75]. However, long-
50
51 440 term electrolysis causes oxygen and molten salts to permeate the interface between the coating
52
53
54 441 and substrate, resulting in the corrosion and degradation of the coating [75]. Hence, further
55
56 442 fundamental studies are required to improve surface film–substrate adhesion.
57
58
443
59
60
61
62
63
40
64
65
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
44 444
45 445
46 446 Fig. 9 Polarization curves of metals in molten (a) pure CaCl2 and (b) CaCl2 – 2.5 mol% CaO
47
48
49
447 at 850 °C (potential scan rate of 2 mV s−1); (c) polarization curves of Ni, Pt, and pre-oxidized
50
51 448 Ni in molten CaCl2 at 850 °C (potential scan rate, 2 mV s−1). Reproduced with permission from
52
53 449 ref. [32]; copyright 2011, RSC. (d) Polarization curves of Pt, Ir, and SnO2 electrodes in molten
54
55
56 450 Li2CO3–Na2CO3–K2CO3 at 500 °C in a CO2 atmosphere; scan rate: 5 mV s−1. Reproduced with
57
58 451 permission from ref. [76]; copyright 2013, RSC. (e) Current efficiency and (f) energy
59
60
61
62
63
41
64
65
452 consumption calculated from experimental data under different electrolysis conditions (PC-
1
2 453 160s, PC-290s, PC-3120s, and DC-114400s used a RuO2–TiO2 anode and varied the electrolysis time;
3
4
5 454 PC-48:2 and PC-57:3 used a Ni-TiO2 anode and varied the composition). Reproduced with
6
7 455 permission from ref. [77]; copyright 2019, RSC.
8
9
10 456
11
12 457 The stability of other pure metals in molten salt systems is assessed by using the anodic
13
14 458 polarization potential. As reported by Yin et al. [32], in pure CaCl2, Mo was the most stable
15
16
17
459 among the tested metals in addition to Pt, followed by Ni, Co, Ag, Fe, and Cu (Fig. 9a).
18
19 460 However, in the presence of oxide ions, in the CaCl2–CaO molten salt system, Ni was more
20
21 461 stable than Mo (Fig. 9b) because of the NiO coating formed on the surface of the Ni metal,
22
23
24 462 effectively preventing corrosion by oxygen ions[32].
25
26 463 On the basis of the anodic polarization potential (Fig. 9c), nickel oxide has resistivity
27
28
29 464 comparable with that of Pt. However, Ge et al. reported that at temperatures lower than 575
30
31 465 °C, LixNi1−xO is the main oxidization product, which is soluble in carbonate electrolytes [78].
32
33
34 466 Accordingly, other metal oxides have been explored. The SnO2 anode shows polarization
35
36 467 curves similar to those of Pt and Ir in Li2CO3–Na2CO3–K2CO3 molten salt at 500 °C (Fig. 9d)
37
38 468 [76]. However, in CaCl2-based molten salts, the CaSnO3 film is formed on the surface, which
39
40
41 469 leads to a reduction in conductivity [79]. In addition, CaSnO3 could be reduced by alkali or
42
43 470 alkaline earth metals, leading to the mass loss of the SnO2 anode. Moreover, SnO2 has poor
44
45
46 471 conductivity, thermal cracking property, and is brittle at high temperatures [80].
47
48 472 Hu et al. reported that CaRuO3-based anodes showed low corrosion rates even after
49
50
51 473 long-term electrolysis in CaCl2–CaO molten salt [81]. Accordingly, a promising composite
52
53 474 oxide anode, i.e., RuO2·TiO2, was developed. The electrochemical behavior of the
54
55
56
475 RuO2·TiO2 anode in CaCl2–CaO (1.0 wt%) was examined. The initial polarization potential of
57
58 476 the RuO2·TiO2 anode was more positive than those of Pt and graphite [81]. In another study,
59
60 477 Hu et al. also evaluated the current efficiency and energy consumption under electrolytic
61
62
63
42
64
65
478 conditions using RuO2–TiO2 and Ni–TiO2 anodes [77] as shown in Fig. 9e. The Ru-based
1
2 479 anode showed high current efficiency. By contrast, the Ni–TiO2 anode showed a significant
3
4
5 480 drop in current efficiency because of the insulating oxide layer generated during electrolysis,
6
7 481 which considerably increased the energy consumption (Fig. 9f). Overall, RuO2·TiO2 inert
8
9
10 482 anode showed good performance in CaCl2-based molten salt. However, owing to the limited
11
12 483 Ru reserves on earth, the practicability of RuO2·TiO2 may be restricted.
13
14
15 484 Recently, the utilization of transition metal alloy-based anodes has been intensively
16
17 485 reported. Ni-based alloys are widely studied as anode materials because of the stability of Ni
18
19
486 in some molten salt systems [32]. Cheng et al. reported that after continuous electrolysis for
20
21
22 487 100 h in carbonate molten salt, NiFe2O4 was found on the surface of Ni–Fe–Cu alloy [82].
23
24 488 Thus, Ni–Fe-based alloys have good long-term stability in carbonate. However, in halide
25
26
27 489 molten salts, the stability of the alloy deteriorates. This issue can be solved by controlling the
28
29 490 composition of alloys. For the Ni–Fe–Cu alloy, the NiFe2O4 phase can be strengthened by Cu.
30
31
32 491 Accordingly, in a study, the penetration of the molten salt into oxide films was found to be
33
34 492 reduced and anode dissolution was effectively inhibited [83]. In addition, when other elements
35
36
37
493 such as Al were added to Ni–Fe alloy, the stability in molten salt was improved. Because Al is
38
39 494 more reactive than Fe–Ni, it is preferentially oxidized, which enables good adhesion between
40
41 495 the metal and oxide film [84]. Thus, the stability of the Ni–Fe alloy in molten salt can be
42
43
44 496 improved by doping elements such as in Inconel 718 (Ni–Fe–Cr), Inconel 625 (Ni–Fe–Cr–
45
46 497 Mo), nichrome C (Ni–Fe–Cr), Muntz brass, and galvanized steel.
47
48
49 498 In some electrolysis processes, O2 generation can promote anode metal erosion. The
50
51 499 erosion of metal anodes depends on molten salt electrolyte. In a study, Ni metal underwent
52
53
54 500 continuous erosion in a sodium and potassium carbonate electrolyte, and no erosion was
55
56 501 evident in barium/lithium carbonate electrolytes [85]. Meanwhile, in lithium carbonate, the rate
57
58
502 of Ni erosion at the anode is a function of anode current density, electrolysis time, temperature,
59
60
61
62
63
43
64
65
503 and lithium oxide concentration [86,87]. It is also worth mentioning that metal anode erosion
1
2 504 could assist the CNT growth process. During anode erosion, small anode fragments are slowly
3
4
5 505 dissolved in the melt and are transported to the cathode, nucleating the CNT growth. When a
6
7 506 galvanized steel anode is used as mentioned in the report of Arcaro et al. [10] and Ren et al.
8
9
10 507 [40], the anode could possibly erode, which facilitates the formation of an Fe–Ni oxide that is
11
12 508 also efficient in catalyzing the growth of CNTs with smaller walls and high crystallinity [88].
13
14
15 509 Thus, compared to pure metals, Ni–Fe-based alloys are better alternatives as the anode owing
16
17 510 to their stability and capability as a potential catalyst for CNT growth.
18
19
511
20
21
22 512 5. Molten-Salt Electrolyte
23
24 513 Electrolytes play crucial roles in the electrochemical reaction of CO2. Electrolytes are
25
26
27 514 essential for collecting CO2 from the incoming gas stream and transferring ions. In addition to
28
29 515 molten salts, electrolytes can be replaced by aqueous solutions, ionic liquids, and solid
30
31
32 516 electrolytes. Different electrolytes exhibit various selectivities, products, and efficiencies. As
33
34 517 summarized in Fig. 10, aqueous solutions are impractical as electrolytes owing to several
35
36
37
518 significant disadvantages, including poor electrochemical windows, water splitting, limited
38
39 519 solubility, and slow reaction kinetics [89,90]. Moreover, ionic liquids show excellent chemical
40
41 520 stability, high CO2 solubility, and high conversion yield; however, their high cost and low
42
43
44 521 reaction kinetics hinder their commercial applications [89,91]. Solid electrolytes have an
45
46 522 overall higher current efficiency, but their specific conductivity is less than that of molten salt
47
48
49 523 electrolytes [26]. Molten salt electrolyte refers to the ionic liquid formed after the salt is melted.
50
51 524 Ionic liquids do not require water as a supporting electrolyte and have no ion solvation process;
52
53
54 525 they can conduct electricity through the migration of their own ions. Molten salt electrolytes
55
56 526 are chosen owing to their satisfying properties. High-temperature molten salts exhibit low
57
58
59
60
61
62
63
44
64
65
527 toxicity, high heat capacity, wide electrochemical operating window, high ionic conductivity,
1
2 528 and low cost due to abundant availability [92].
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
38 529
39
40 530 Fig. 10 Comparison of the potential of the electrochemical process to reduce the atmospheric
41
42
43
531 CO2 concentration using different electrolytes. Reproduced with permission from ref. [26];
44
45 532 copyright 2019, Elsevier.
46
47 533
48
49
50 534 Selecting a molten salt electrolyte is considered more crucial than choosing an electrode
51
52 535 because electrolytes have a greater impact on the CO2 solubility and electrode reactions leading
53
54
55 536 to final carbon product properties. The molten salt suitable for the electrochemical reaction of
56
57 537 CO2 should be able to dissolve the O2− ion, which helps capture CO2. Carbonates, chlorides,
58
59
60
538 and oxides containing alkali and alkaline earth metals are the most widely researched
61
62
63
45
64
65
539 electrolytes owing to their wide electrochemical windows and good CO2 solubility [26]. They
1
2 540 are frequently mixed together to achieve optimal performance, e.g., lower melting points.
3
4
5 541 Chlorides show lower melting points than carbonates. However metal chlorides have
6
7 542 poor CO2 absorption ability [93]. Accordingly, chlorides need to be mixed with oxides or
8
9
10 543 carbonates to improve CO2 capture and conversion. Compared to carbonate electrolytes with
11
12 544 oxide additives, chlorides with oxide additives absorb CO2 faster. Deng et al. reported that the
13
14
15 545 Li2O-to-Li2CO3 conversion efficiency almost doubled in molten chloride electrolyte (~94%)
16
17 546 compared to that in molten carbonate electrolyte (~45%) [47]. The low activity of Li2O in
18
19
547 carbonates is due to the strong interaction between CO2 and carbonates, which leads to slower
20
21
22 548 CO2 diffusion.
23
24 549 In the molten carbonate system, metal cations dominate the properties of the system
25
26
27 550 because they can be reduced to elemental metal during electrolysis and affect the yield of
28
29 551 produced carbon. Therefore, the choice of metal cations has considerable impact on the
30
31
32 552 reaction. According to Table 3, in molten alkali carbonates, carbon deposition is
33
34 553 thermodynamically preferred to Li deposition; however, it is more difficult than Na or K
35
36
37
554 deposition [9]. Thermodynamic calculations also show that it is possible to use molten alkaline
38
39 555 earth metals, particularly in CaCO3. However, CaCO3 decomposes at temperatures only
40
41 556 slightly above its melting point (825 °C). To overcome this issue, additives are added to the
42
43
44 557 molten salt system. As reported by Cao et al., SnO2 and NaCl were added to the CaCO3–CaO
45
46 558 molten salt system. This mixture successfully produced Sn encapsulated in MWCNT at 800 °C
47
48
49 559 [52].
50
51 560
52
53 561 Table 3 Deposition potentials of metal cations and carbon at 600 °C. Reproduced with
54
55
56 562 permission from ref. [9]; copyright 2014, RSC.
57
58 Molten salt Deposition Deposition
59 potential of metal potential of carbon
60
61
62
63
46
64
65
Li2CO3 −2.964 V −1.719 V
1 Na2CO3 −2.546 V −2.551 V
2
3
K2CO3 −2.612 V −3.083 V
4 CaCO3 −3.033 V −1.349 V
5 BaCO3 −3.069 V −1.992 V
6 563
7
8
9 564 Additional metal oxides, including SnO, ZnO, Fe2O3, FeO, and Li2O, have been
10
11 565 explored to improve the nanostructure of the carbon produced. Metal cations of the additives
12
13
14 566 can be deposited at the cathode and serve as nucleation sites for CNT growth. Zn or ZnO
15
16 567 additives were reported to be favorable for CNF formation [40]. Another report by Weng et al.
17
18
19
568 mentioned that Ge was encapsulated in MWCNT as a result of the electrochemical reaction of
20
21 569 CO2 in the NaCl–CaCl2–CaO–GeO molten salt system [51].
22
23 570 Borate ion has excellent performance in molten salt-based CO2 electrolysis. Because of
24
25
26 571 their versatile oxo-structures and broad oxo-acidity, borates are promising candidates serving
27
28 572 as conjugate acid/base pairs (toward O2−) that can accommodate O2− and concurrently absorb
29
30
31 573 CO2 according to the Lux–Flood acid–base theory [94], where acidity is dependent on the
32
33 574 function and relative content of oxygen atoms in the molecular structures. For example, the
34
35
36 575 oxo-acidity sequence of borates follows the order B2O3 > B4O72− > BO2− > BO33− [95], among
37
38 576 which more oxo-acidic borate species (e.g., B2O3, B4O72−, and BO2−) acting as O2− acceptors
39
40
577 can spontaneously react with O2− to form more oxo-basic borate species (e.g., BO33−), while
41
42
43 578 BO33− exhibits strong CO2 absorption capability due to its suitable oxo-basicity [96].
44
45 579 Yang et al. reported the development of a Li-free Na–K carbonate by introducing earth-
46
47
48 580 abundant borax (Na2B4O7) [97]. The reactions can be described as follows [98–100]:
49
50 581 Initial ripening process of B4O72− :
51
52
53 582 B4 O7 2− + CO3 2− → BO3 3− + BO2 − + CO2 4
54
55 583 Transformation of O2-:
56
57
58 584 BO2 − + O2− → BO3 3− 5
59
60 585 CO2 capture by oxo-basic borate species:
61
62
63
47
64
65
586 BO3 3− + CO2 → CO3 2− + BO2 − 6
1
2
3
587 By combining equations (2) and (5), the borate-assisted cathodic reactions are as follows:
4
5 588 CO3 2− + 4e− + 3BO2 − → C + 3BO3 3− 7
6
7
8 589 CO3 2− + 2e− + 2BO2 − → CO + 2BO3 3− 8
9
10 590
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
44 591
45
46 592 Fig. 11 Cathodic reactions and characterizations. (a) Standard theoretical potentials of possible
47
48 593 cathodic reactions in molten Na2CO3–K2CO3 with and without borates, (b) CO2 absorption
49
50
51 594 kinetics in molten Na2CO3–K2CO3 and NK–Na2B4O7–0.05 under various CO2 partial
52
53 595 pressures, cyclic voltammograms on Ni working electrodes (c) in bare Na2CO3–K2CO3 and
54
55
56 596 Na2CO3–K2CO3–Na2B4O7 (NK–Na2B4O7-0.05) and (d) in molten Na2CO3–K2CO3 containing
57
58
59
60
61
62
63
48
64
65
597 different Na2B4O7 contents at 750 °C under an Ar atmosphere. Scan rate: 100 mV s−1.
1
2 598 Reproduced with permission from ref. [97]; copyright 2023, RSC.
3
4
5 599
6
7 600 Fig. 11a reveals that the theoretical potentials of the borate-assisted CO2RR (equations
8
9 601 (7) and (8)) are much more positive than that of alkali metal deposition, suggesting that CO2RR
10
11
12 602 could be achieved in Na–K carbonate owing to the spontaneous transformation effect of borates
13
14 603 toward O2− (equation ( 5 )) [97]. More interestingly, CO2RR precursors (CO32−) and
15
16
17 604 O2− acceptor (BO2−) can be regenerated by borate-involved CO2 absorption as given
18
19 605 by equation (6), which could additionally create a new synergistic “capture–electroreduction”
20
21
22
606 loop mediated by borates at the cathode.
23
24 607 To elucidate the CO2 absorption capability in NK–Na2B4O7–0.05, the CO2 capture
25
26 608 kinetics was investigated, and the results clearly showed that the CO2 absorption capacity in
27
28
29 609 NK–Na2B4O7-0.05 was over 40 times higher than that in borate-free Na–K carbonate (Fig. 11b)
30
31 610 [97]. Cyclic voltammetry was conducted in Na2CO3–K2CO3 with and without Na2B4O7 to
32
33
34 611 verify that borate could change the thermodynamic pathway of CO2RR. As shown in Fig. 11c,
35
36 612 a broad extra cathodic peak (c1) as the peak of carbon deposition was observed in Na2CO3–
37
38
39 613 K2CO3–Na2B4O7, while only the cathodic limit of alkali metal deposition (c0) could be found
40
41 614 in Na2CO3–K2CO3. Furthermore, the current density of the cathodic peak c1 increased with an
42
43
615 increase in the concentration of Na2B4O7 (Fig. 11d), suggesting that the cathodic process was
44
45
46 616 associated with borates [97].
47
48 617
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
49
64
65
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
33 618
34
35 619 Fig. 12 Anodic reactions and product distributions. (a) Theoretical oxygen evolution reaction
36
37
38
620 (OER) potentials of typical borate species. (b) Anodic linear sweep voltammetry on a Pt
39
40 621 working electrode in molten Na2CO3–K2CO3 containing different Na2B4O7 contents under an
41
42 622 Ar atmosphere at 750 °C. Scan rate: 5 mV s−1. (c) Anodic potential variation in different molten
43
44
45 623 electrolytes during galvanostatic electrolysis under an Ar atmosphere and (d) galvanostatic
46
47 624 electrolysis in different molten electrolytes at a cathodic current density of 100 mA cm −2.
48
49
50 625 Reproduced with permission from ref. [97]; copyright 2023, RSC.
51
52 626
53
54 627 As mentioned above, there is a large amount of oxo-basic BO33− derived from
55 628 Na2B4O7, which is capable of absorbing CO2 (equation (6)) in the bulk electrolyte.
56
57 629 Furthermore, BO33− can be deemed an O2− carrier. Therefore, borates could contribute to the
58 630 oxygen evolution reaction (OER) similar to the effect of O2− but providing more advantages
59 631 in leveraging reaction kinetics because of the versatile oxo-structures and higher saturated
60
61
62
63
50
64
65
632 solubility (at least 10 times higher) compared to O2− [101]. The thermodynamic calculation of
1 633 theoretical OER potentials originating from borate species is presented in
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
634
37 635 Fig. 12a. Interestingly, the OER potentials of oxo-basic borates, such as B2O54−, BO33−,
38
39
40 636 and B4O72− (which hardly exist in carbonate) with relatively high water solubility, to generate
41
42 637 BO2− (equations (9)–(11)) that can transform O2− at the cathode are more preferential than that
43
44
45 638 of CO32–.
46
47 639 2BO3 3− − 4e → 2BO2 − + O2 9
48
49
50 640 4BO3 3− − 4e → 2B2 O5 4− + O2 10
51
52 641 2B2 O5 4− − 4e → 4BO2 − + O2 11
53
54
55
56
57
58
59
60
61
62
63
51
64
65
642 Linear sweep voltammetry was conducted on a Pt working electrode to investigate the
1 643 anodic reactions occurring in Na–K carbonate with Na2B4O7.
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
644
37 645 Fig. 12b shows that the negative shift onset potential of the polarization curve in the
38
39
40 646 borate-containing electrolyte is increasingly obvious with the increasing concentration of
41
42 647 Na2B4O7, suggesting that the anodic current is also associated with borates.
43
44
45 648 To verify the advantages of borate-involved OER, galvanostatic electrolysis was
46 649 conducted in borate-containing, O2−-saturated, and bare Na–K carbonate, during which the
47 650 variations of the corresponding anodic potential and O2 content were recorded in real time.
48
49 651 As shown
50
51
52
53
54
55
56
57
58
59
60
61
62
63
52
64
65
652 in
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
34 653
35
36 654 Fig. 12c, the corresponding anodic potential in borate-containing electrolyte was highly
37
38 655 stable over a wide anodic current density range [97]. Consequently, the overpotential of borate-
39
40
41 656 involved OER was minimal compared to those of CO32−- and O2−-involved OER at the
42
43 657 universal current density.
44
45
46 658 The borate-containing electrolyte exhibited better CO2RR performance because of its
47 659 tunable cathodic product selectivity and the pure oxygen at the anode as well as the lower
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
53
64
65
660 applied cell voltages (
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
34 661
35
36 662 Fig. 12d) and corresponding more positive cathodic potentials [97], suggesting the
37
38 663 beneficial effect of borates in reducing both cathodic and anodic polarizations, with a
39
40
41 664 considerable decrease in the applied cell voltage. Accordingly, the energy efficiency of CO2RR
42
43 665 in borate-containing electrolytes can reach as high as 60% or more [97].
44
45
46 666
47
48 667 6. Commercial CNTs
49
50
51 668 CNTs can be described as carbon sheets comprising 60 carbon atoms rolled into a
52
53 669 cylinder [102]. CNTs are classified into two types according to the number of carbon layers
54
55
56
670 they contain. SWCNTs comprise a single graphene layer with diameter varying between 0.4
57
58 671 and 3 nm and usually occur as hexagonal-packed bundles. MWCNTs consist of two or more
59
60 672 cylinders, each made up of graphene sheets. Their diameter varies from 10 to 20 nm.
61
62
63
54
64
65
673 SWCNTs are believed to possess robust mechanical properties due to the strong
1
2 674 covalent bond among the C atoms. They have a Young’s modulus of 1 TPa, which is five-time
3
4
5 675 that of steel [103,104]. They are highly flexible, regain their original shape after buckling and
6
7 676 bending, and have a low density, approximately one-sixth of the density of steel [103,105].
8
9
10 677 Moreover, the properties of CNTs as an insulator, semiconductor, or conductor depend on the
11
12 678 chirality, i.e., the way in which carbon atoms are arranged. The electrical properties of
13
14
15 679 SWCNTs and MWCNTs are different because SWCNT graphene sheets can be rolled up with
16
17 680 varying degrees of twist, and the resulting SWCNTs may have various chiral structures,
18
19
681 affording different materials for nanoelectronics.
20
21
22 682 Kobashi et al. studied the correlation between the structures and properties of CNTs,
23
24 683 and three general categories emerged: (1) low crystallinity and low specific surface area (SSA)
25
26
27 684 (large-diameter MWCNTs), (2) low crystallinity and higher SSA (larger-diameter SWCNTs
28
29 685 and double-walled CNTs), and (3) high crystallinity and moderate SSA (small-diameter
30
31
32 686 SWCNTs) [106]. SWCNTs with a smaller diameter have high crystallinity and high SSA with
33
34 687 high conductivity and mechanical strength. They are suitable for fibers and transparent
35
36
37
688 conducting films. CNTs with a high surface area and lower crystallinity are appropriate for
38
39 689 supercapacitors [107], elastomers, and actuators because in these applications, surface area is
40
41 690 the key parameter. Further, MWCNTs with low crystallinity and low SSA are the easiest to
42
43
44 691 synthesize, available on a large scale and at the lowest cost. Accordingly, they are heavily used
45
46 692 in Li-ion battery electrodes. As another evidence, few-walled CNTs (FWCNTs) were reported
47
48
49 693 to comprise 2–6 graphene layers, having a length of tens of micrometers and diameter range of
50
51 694 2–8 nm [77]. Some applications of FWCNTs have been found related to field emission, and
52
53
54 695 they could become the next generation of CNT-based reinforced composite materials [108].
55
56 696
57
58
59
60
61
62
63
55
64
65
697 Table 4 Recent technologies for the synthesis of commercial CNTs. Reproduced with
1
2 698 permission from ref. [109]; copyright 2019, American Chemical Society.
3
4
5 Manufacturer CNT Synthesis Layer Diameter Wall Length BET G/D
6 grade method number SSA ratio
7 (CVD) (m2/g)
8
9 Showa Denko VGCF Floating MW 128 (150) 8 10 5.4
10 catalyst (13)
11 Cnano FloTube Fluidized MW 9.7 (11) 8.9 10 210 0.7
12 9000 bed (≥200)
13
14
Nanocyl NC7000 Catalytic MW 8.7 (9.5) 6.7 1.5 280 0.6
15 (250–
16 300)
17 Kumho K-Nanos- Continuous MW 9.1 (8– 6.2 Average 270 0.8
18
19
Petrochemical I00p fluidized 15) bundle (0.7–
20 bed length of 1.0
21 26
22 JEIO JC142 Continuous MW 4.3 2.4 Tens to 640 0.5
23 hundreds
24
25 by bundle
26 length
27 Zeon SG-CNT Fixed bed MW 3.7 (1–5) 1.2 100–400 1080 4.6
28 HT (pilot by forest (800)
29
30
plant) height
31 Meijo Nano eDIPS Floating SW 2.2 (2 ± 1.4 Millimeter 480 58.8
32 Carbon EC2.0 catalyst 0.8 scale by
33 bundle
34 length
35
36 OCSiAl TUBALL Floating SW 2.1 (1.8 ± 1.0 Hundreds 390 25.6
37 catalyst 0.4) by bundle (~400) (~50)
38 length
39 699
40
41
42 700 Currently, CNTs have been commercialized. Several methods are available for the
43
44 701 synthesis of nanotubes. The arc-discharge method was the first method employed to produce
45
46
47 702 nanotubes by the arc vaporization of two graphite rods placed in an inert gas for the reaction.
48
49 703 Carbon rods are evaporated by direct current, which creates a high-temperature discharge
50
51
52 704 between two electrodes, which causes the evaporation of the anode and deposition of nanotubes
53
54 705 on the cathode [110]. The second method is laser ablation, in which a piece of graphite is
55
56
706 vaporized by laser irradiation under an inert atmosphere [110]. This process produces soot
57
58
59 707 containing CNTs, which are cooled at the walls of a quartz tube. This method is very expensive;
60
61
62
63
56
64
65
708 hence, it is mainly used for producing SWCNTs. High-purity SWCNTs have been produced
1
2 709 with less defects and contaminants using this method. The third method is CVD. In this method,
3
4
5 710 the nanotubes are produced by the thermal decomposition of hydrocarbons in the presence of
6
7 711 metal catalysts. There are several techniques under this method, including floating catalyst,
8
9
10 712 fixed bed, fluidized bed, and rotary kiln [111]. These different techniques yield different
11
12 713 properties of CNTs. Table 4 summarizes the CNT properties depending on the production
13
14
15 714 methods. Fluidized-bed CVD is an efficient technique to deposit on, functionalize, or coat each
16
17 715 individual particle of a powder from gaseous species. This technique combines two processes
18
19
716 simultaneously: the first is the deposition process and the second is resisting the particles in the
20
21
22 717 deposition area by introducing a gas upward through the powder.
23
24 718 For the wide application of CNTs, the most efficient method of synthesis, which can be
25
26
27 719 scaled up to commercial production, is required. Presently, CNTs are being manufactured by
28
29 720 CVD using hydrocarbon precursors via a gas-phase reaction. These methods produce both
30
31
32 721 MWCNTs and SWCNTs depending on the operating conditions (Table 5).
33
34 722
35
36
37
723 Table 5 Summary of reaction conditions for CNT synthesis via the CVD process. Reproduced
38
39 724 with permission from Ref. [112]; copyright 2022, Sage Journals.
40
41
Type of Catalyst Reaction Precursor Reaction Product Reference
42
43 CVD material temperature gas time
44 used (°C)
45 Thermal Si (100) – Al Ferrocene CNTs [113]
46 CVD and Fe
47
48 Thermal Copper foil 900 CH4 and 30 min CNTs [114]
49 CVD using a Ni acetylene
50 thin film
51 Thermal Silicon wafer 1000 Acetylene 6–14 min CNTs [115]
52
53
CVD with
54 predeposited
55 2-nm-thick
56 Fe
57
Thermal Rice straw, 800 Camphor 120 min MWCNTs [116]
58
59 CVD Fe–Ni
60
61
62
63
57
64
65
Floating Monometallic 1050 Ethylene 10 s SWCNTs [117]
1 catalyst (Fe, Co, Ni),
2
CVD Bimetallic
3
4 (Co–Ni, Co–
5 Fe)
6 Thermal Fe and Ni 750 N2/CH4 MWCNTs [118]
7 CVD and
8
9 H2/CH4
10 Thermal Ni–Mo and 700 Natural 60 min MWCNTs [119]
11 CVD Co–Mo gas
12 supported on
13
14
Al2O3
15 Thermal Cu and Au 900 CH4 30 min MWCNTs [120]
16 CVD
17 725
18
19
20 726
21
22 727 7. CVD
23
24
25 728 CVD involves separating a gas containing carbon atoms that continuously flows
26
27 729 through a catalytic nanoparticle to produce carbon atoms, which then produce CNTs on the
28
29 730 surface of the catalyst or substrate. The synthesis process involves allowing a catalyst to
30
31
32 731 decompose a carbon source (usually a gaseous hydrocarbon) at sufficiently high temperatures
33
34 732 in a tubular reactor. In the synthesis of CNTs by CVD, four parameters play a vital role in
35
36
37 733 determining the structure and morphology of the CNTs: catalyst, carbon source, temperature
38
39 734 control, and substrate [121].
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55 735
56
57
58
736 Fig. 13 (a) SWCNT and FWCNT growth scenario from small catalyst particles. (b) MWCNT
59
60 737 growth scenario from large catalyst particles. Reproduced with permission from ref. [122];
61
62
63
58
64
65
738 copyright 2008, Elsevier. (c) Schematic diagram of a CVD setup. Reproduced with permission
1
2 739 from ref. [123]; copyright 2011, Intech Open.
3
4
5 740
6
7 741 Catalysts play a very important role in determining CNT structure and morphology.
8
9 742 Catalyst nanoparticles act as a nucleation site for CNT growth. Generally, nanometer-sized
10
11
12 743 particles are required to synthesize SWCNTs, whereas MWCNTs can be produced without a
13
14 744 catalyst. In terms of the nucleation site, the size of a catalyst determines the diameter of CNTs.
15
16
17 745 When the size of the catalyst is a few nanometers, SWCNTs are formed, whereas particles that
18
19 746 are a few tens of nanometer wide favor MWCNT formation. This leads to two types of growth
20
21
22
747 mechanisms of CNTs: base/root growth and tip growth. In root growth, the catalyst particle
23
24 748 stays pinned at the support surface and the support-particle interaction must be considered. By
25
26 749 contrast, the tip-growth mechanism occurs when the catalyst is lifted off from the support
27
28
29 750 surface during CNT growth owing to weak support–catalyst interaction.
30
31 751 Small catalyst particles contain more reactive metal particles, which implies a
32
33 752 stronger interaction with the carbon patch [122]. Hence, after dehydrogenation of the first
34 753 carbon source molecule, a graphene cap forms and cuts off the carbon source on the particles
35 754 (
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50 755
51
52 756 Fig. 13a – steps 2 and 3). Therefore, the carbon flow can only be supplied to the tight
53 757 interface between the catalyst and substrate. In the case of SWCNT growth, carbon is
54
55
56
57
58
59
60
61
62
63
59
64
65
758 incorporated into the edges of the graphene cap, raising the catalyst surface (
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15 759
16
17 760 Fig. 13a – step 4). Thus, small catalyst particles follow the root growth mechanism
18
19 761 [122].
20
21
22 762 Additional pathways need to be considered to describe FWCNT growth. In this case,
23 763 the carbon diffusion within the catalyst itself and its extrusion under the first graphene cap
24 764 could be considered (
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39 765
40
41 766 Fig. 13a – step 4’). The number of additional caps may directly depend on the catalyst
42 767 size. When no further hemispherical structures can be formed, the more favorable path for
43
44 768 atomic carbon incorporation is at the edges of the FWCNT nuclei, leading to stable FWCNT
45 769 growth (
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 770
61
62
63
60
64
65
771 Fig. 13a – step 5’) [122].
1
2 772 In larger catalyst particles, the carbon sheets have sufficient time to diffuse before
3
4 773 forming a complete hemispherical cap. Thus, in the early stage of the CNT nucleation, the
5 774 first graphitic sections or crosslinked carbon chains formed on the surface of the large
6 775 catalyst surface can quickly diffuse toward the catalyst/substrate interface and stabilize it (
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21 776
22
23 777 Fig. 13b − step 2) [122]. This step leaves the top surface of the nanoparticle exposed
24
25
778 for subsequent carbon absorption. As observed by in situ TEM, this nucleation step leads to
26 779 particle elongation (
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41 780
42
43 781 Fig. 13b − step 3) and finally drives the particle lift-off from the substrate (
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58 782
59 783 Fig. 13b − step 4). Thus, large catalyst particles follow the tip-growth mechanism [122].
60
61
62
63
61
64
65
784 TM (Fe, Co, Ni, Pd, Pt, Au, Mn, W, Ti) nanoparticles are considered the most effective
1
2 785 catalysts owing to their high rate of carbon diffusion. Furthermore, these metals were found to
3
4
5 786 have stronger adhesion to the growing CNTs; thus, they were more effective in forming small-
6
7 787 diameter CNTs such as SWCNTs [59]. Shin et al. compared the catalytic efficiency of Ni and
8
9
10 788 Fe. They reported that at the same temperature, synthesizing CNTs from Ni-based catalytic
11
12 789 systems is more difficult than from Fe-based catalytic systems [88].
13
14
15 790 In the CVD method to produce CNTs, the catalyst is deposited on a substrate (
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30 791
31 792 Fig. 13c), and the nucleation of the catalyst is realized via thermal annealing. Thus,
32
33
34 793 temperature also plays an important role in the CNT formation. SWCNTs have a higher energy
35
36 794 of formation owing to their small diameters, high curvature, and high strain energy [121].
37
38
39 795 Therefore, a higher temperature is required to produce SWCNTs compared to MWCNTs. As
40
41 796 summarized in Table 5, SWCNTs can be successfully synthesized at a reaction temperature of
42
43
44 797 1050 °C, whereas MWCNTs can be synthesized at lower reaction temperatures.
45
46 798 In addition to temperature, the catalyst needs appropriate carbon sources and substrate
47
48 799 materials. To produce SWCNTs, the carbon source must be highly stable at high temperatures,
49
50
51 800 e.g., methane, carbon monoxide, carbon dioxide, etc. Furthermore, the surface morphology and
52
53 801 texture properties of the substrate affect the yield and quality of the grown catalyst. The
54
55
56 802 interaction between the catalyst and substrate affects the CNT growth environment. When the
57
58 803 catalyst–substrate interaction is strong (the metal makes an obtuse contact angle with the
59
60
61
62
63
62
64
65
804 substrate), initial hydrocarbon decomposition and carbon diffusion occur similarly to that in
1
2 805 the tip-growth case. However, the CNT precipitation fails to push the metal particle up;
3
4
5 806 therefore, the precipitation is compelled to emerge out of the metal’s apex (farthest from the
6
7 807 substrate, having minimum interaction with the substrate) [123]. First, carbon crystallizes out
8
9
10 808 as a hemispherical dome (the most favorable closed-carbon network on a spherical
11
12 809 nanoparticle), which then extends up in the form of a seamless graphitic cylinder. Subsequent
13
14
15 810 hydrocarbon decomposition occurs on the lower peripheral surface of the metal, and as-
16
17 811 dissolved carbon diffuses upward. Thus, the CNT grows up with the catalyst particle rooted to
18
19
812 its base; hence, this is known as the “base-growth model.”
20
21
22 813 When the catalyst–substrate interaction is weak (the metal makes an acute contact angle
23
24 814 with the substrate), the hydrocarbon decomposes on the top surface of the metal, carbon
25
26
27 815 diffuses down through the metal, and CNTs precipitate out across the metal bottom, pushing
28
29 816 the entire metal particle off the substrate [123]. As long as the metal’s top is open for fresh
30
31
32 817 hydrocarbon decomposition, a concentration gradient exists in the metal, allowing carbon
33
34 818 diffusion, and CNTs continue to grow increasingly longer. Once the metal is fully covered with
35
36
37
819 excess carbon, its catalytic activity ceases, and the CNT growth stops. This is known as the
38
39 820 “tip-growth model” [123].
40
41 821 Various substrates for CNT growth have been reported, such as alumina [124], graphite
42
43
44 822 [125], silica [126], titania [127], and zeolite [128]. Alumina materials are reportedly better
45
46 823 catalyst supports than silica owing to stronger metal–support interaction, which allows high
47
48
49 824 metal dispersion and thus a high density of catalytic sites [126]. Such interactions prevent metal
50
51 825 species from aggregating and forming unwanted large clusters that lead to graphite particles or
52
53
54 826 defective MWCNTs.
55
56 827
57
58
59
828 8. Conclusion and Perspectives
60
61
62
63
63
64
65
829 In this paper, we review the recent progress in the electrochemical conversion of CO2
1
2 830 into CNT. The growth of CNT is successfully regulated by the operational conditions, including
3
4
5 831 the cathode, anode, composition of molten salt, temperature, and electrolysis potential.
6
7 832 However, SWCNT production has not been achieved yet. Nevertheless, this method is still
8
9
10 833 promising to develop as a CNT production method owing to its higher selectivity and high
11
12 834 current efficiency to produce valuable CNTs. This method is also believed to become a solution
13
14
15 835 to the environmental problems due to the excessive concentration of CO2 emissions.
16
17 836 To produce carbon nanostructure morphology deposited at the cathode, various TMs,
18
19
837 particularly “goldilocks zone” TMs, can be used as catalysts for CNT growth. Because of the
20
21
22 838 strong TM–carbon bond adhesion energy, TMs possibly act as a nucleation site for CNT growth
23
24 839 through the base or tip mechanism. TM alloy-based anodes can also be utilized in this system.
25
26
27 840 In terms of stability, Ni–Fe-based alloys are better as anodes and potential catalysts for CNT
28
29 841 growth, as metal erosion occurs.
30
31
32 842 Moreover, electrolyte composition plays crucial roles owing to its effect on CO2
33
34 843 solubility and electrode reactions leading to final carbon product properties. Borate ion has
35
36
37
844 excellent performance as an additive in molten salt-based CO2 electrolysis systems. Borate-
38
39 845 containing electrolytes successfully tune cathodic product selectivity by shifting the cathodic
40
41 846 potential to be more positive. By using borate additives, Li-free molten salts can be used in
42
43
44 847 such electrolysis systems. The overpotential of borate-involved OER has also been successfully
45
46 848 minimized, with a considerable decrease in the applied cell voltage. The high energy efficiency
47
48
49 849 of CO2RR can be reached in borate-containing electrolytes.
50
51 850 For SWCNT production, an ideal catalyst environment must be designed using some
52
53
54 851 CVD-related principles. Considering key CVD strategies in producing SWCNTs, the process
55
56 852 should be controlled so that the reaction leads to the base-growth mechanism. This mechanism
57
58
853 can be achieved by controlling the substrate–catalyst interaction to be strong enough. Alumina,
59
60
61
62
63
64
64
65
854 graphite, silica, titania, and zeolite are reportedly good materials to support the catalysis of
1
2 855 SWCNT growth. However, to use these supporting materials at a suitable operating
3
4
5 856 temperature, further in-depth studies are required.
6
7 857
8
9 858 Conflicts of interest
10
11
12 859 The authors declare that they have no known competing financial interests or personal
13
14
15 860 relationships that could have influenced the work reported in this paper.
16
17
18 861
19
20 862 Acknowledgments
21
22 863 This work was supported by the National Research Foundation of Korea (NRF) Grant funded
23
24
25 864 by the Korean Government (MSIT) (RS-2023-00210114, NRF- 2022R1A4A1019296) and by
26
27 865 the KIST Institutional Program (Project No.2V09760-23-021).
28
29
866
30
31
32
33
867 References
34
35
36 868 [1] R. Lindsey, E. Dlugokencky, Climate Change: Atmospheric Carbon Dioxide (Accessed
37 869 2023), Climate.Gov. (2023).
38
39 870 [2] Z. Lou, Q. Chen, Y. Zhang, W. Wang, Y. Qian, Diamond formation by reduction of
40 871 carbon dioxide at low temperatures, J Am Chem Soc. 125 (2003).
41 872 https://doi.org/10.1021/ja035177i.
42
43 873 [3] A. Chakrabarti, J. Lu, J.C. Skrabutenas, T. Xu, Z. Xiao, J.A. Maguire, N.S. Hosmane,
44 874 Conversion of carbon dioxide to few-layer graphene, J Mater Chem. 21 (2011) 9491.
45
875 https://doi.org/10.1039/c1jm11227a.
46
47 876 [4] C. Chen, Z. Lou, Formation of C60 by reduction of CO2, J Supercrit Fluids. 50 (2009)
48
49 877 42–45. https://doi.org/10.1016/j.supflu.2009.04.008.
50
878 [5] Z. Li, D. Yuan, H. Wu, W. Li, D. Gu, A novel route to synthesize carbon spheres and
51
52 879 carbon nanotubes from carbon dioxide in a molten carbonate electrolyzer, Inorg Chem
53 880 Front. 5 (2018). https://doi.org/10.1039/c7qi00479f.
54
55 881 [6] B. Deng, J. Tang, M. Gao, X. Mao, H. Zhu, W. Xiao, D. Wang, Electrolytic synthesis of
56 882 carbon from the captured CO2 in molten LiCl–KCl–CaCO3: Critical roles of electrode
57 883 potential and temperature for hollow structure and lithium storage performance,
58 884 Electrochim Acta. 259 (2018). https://doi.org/10.1016/j.electacta.2017.11.025.
59
60
61
62
63
65
64
65
885 [7] S. Vadukumpully, J. Paul, N. Mahanta, S. Valiyaveettil, Flexible conductive
1 886 graphene/poly(vinyl chloride) composite thin films with high mechanical strength and
2 887 thermal stability, Carbon N Y. 49 (2011). https://doi.org/10.1016/j.carbon.2010.09.004.
3
4 888 [8] H. Dai, A.G. Rinzler, P. Nikolaev, A. Thess, D.T. Colbert, R.E. Smalley, Single-wall
5
889 nanotubes produced by metal-catalyzed disproportionation of carbon monoxide,
6
7 890 Chem Phys Lett. 260 (1996). https://doi.org/10.1016/0009-2614(96)00862-7.
8
9 891 [9] H. V. Ijije, R.C. Lawrence, N.J. Siambun, S.M. Jeong, D.A. Jewell, D. Hu, G.Z. Chen,
10 892 Electro-deposition and re-oxidation of carbon in carbonate-containing molten salts,
11 893 Faraday Discuss. 172 (2014) 105–116. https://doi.org/10.1039/C4FD00046C.
12
13 894 [10] S. Arcaro, F.A. Berutti, A.K. Alves, C.P. Bergmann, MWCNTs produced by electrolysis
14 895 of molten carbonate: Characteristics of the cathodic products grown on galvanized
15 896 steel and nickel chrome electrodes, Appl Surf Sci. 466 (2019) 367–374.
16
897 https://doi.org/10.1016/j.apsusc.2018.10.055.
17
18 898 [11] A. Douglas, R. Carter, M. Li, C.L. Pint, Toward Small-Diameter Carbon Nanotubes
19
20 899 Synthesized from Captured Carbon Dioxide: Critical Role of Catalyst Coarsening,
21 900 ACS Appl Mater Interfaces. 10 (2018) 19010–19018.
22 901 https://doi.org/10.1021/acsami.8b02834.
23
24 902 [12] S. Licht, A. Douglas, J. Ren, R. Carter, M. Lefler, C.L. Pint, Carbon nanotubes
25 903 produced from ambient carbon dioxide for environmentally sustainable lithium-ion and
26 904 sodium-ion battery anodes, ACS Cent Sci. 2 (2016) 162–168.
27
28
905 https://doi.org/10.1021/acscentsci.5b00400.
29
906 [13] M.F. Yu, B.S. Files, S. Arepalli, R.S. Ruoff, Tensile loading of ropes of single wall
30
31 907 carbon nanotubes and their mechanical properties, Phys Rev Lett. 84 (2000).
32 908 https://doi.org/10.1103/PhysRevLett.84.5552.
33
34 909 [14] Y. Chen, M. Wang, J. Zhang, J. Tu, J. Ge, S. Jiao, Green and sustainable molten salt
35 910 electrochemistry for the conversion of secondary carbon pollutants to advanced
36 911 carbon materials, J Mater Chem A Mater. 9 (2021) 14119–14146.
37 912 https://doi.org/10.1039/D1TA03263A.
38
39 913 [15] Y. Tsuru, M. Nomura, F.R. Foulkes, Effects of boric acid on hydrogen evolution and
40
41
914 internal stress in films deposited from a nickel sulfamate bath, J Appl Electrochem. 32
42 915 (2002). https://doi.org/10.1023/A:1020130205866.
43
44 916 [16] T.J. Melton, J. Joyce, J.T. Maloy, J.A. Boon, J.S. Wilkes, Electrochemical Studies of
45 917 Sodium Chloride as a Lewis Buffer for Room Temperature Chloroaluminate Molten
46 918 Salts, J Electrochem Soc. 137 (1990). https://doi.org/10.1149/1.2086315.
47
48 919 [17] V. Kaplan, E. Wachtel, K. Gartsman, Y. Feldman, I. Lubomirsky, Conversion of CO[sub
49 920 2] to CO by Electrolysis of Molten Lithium Carbonate, J Electrochem Soc. 157 (2010)
50 921 B552. https://doi.org/10.1149/1.3308596.
51
52 922 [18] D. Chery, V. Lair, M. Cassir, CO2 electrochemical reduction into CO or C in molten
53
923 carbonates: A thermodynamic point of view, Electrochim Acta. 160 (2015).
54
55 924 https://doi.org/10.1016/j.electacta.2015.01.216.
56
57 925 [19] L. Hu, B. Deng, Z. Yang, D. Wang, Buffering electrolyte alkalinity for highly selective
58 926 and energy-efficient transformation of CO2 to CO, Electrochem Commun. 121 (2020)
59 927 106864. https://doi.org/10.1016/j.elecom.2020.106864.
60
61
62
63
66
64
65
928 [20] Z. Kang, J. Xu, H. Liu, Y. Lin, X. Liu, M. He, Thermophysical properties of LiF–LiCl–
1 929 Li2CO3 eutectic mixture/multi-walled carbon nanotubes for thermal energy storage,
2 930 Solar Energy Materials and Solar Cells. 241 (2022) 111744.
3
4
931 https://doi.org/10.1016/j.solmat.2022.111744.
5
932 [21] T. Delise, A.C. Tizzoni, M. Ferrara, M. Telling, L. Turchetti, N. Corsaro, S. Sau, S.
6
7 933 Licoccia, Phase Diagram Predictive Model for a Ternary Mixture of Calcium, Sodium,
8 934 and Potassium Nitrate, ACS Sustain Chem Eng. 8 (2020).
9 935 https://doi.org/10.1021/acssuschemeng.9b04472.
10
11 936 [22] G. Mohan, M. Venkataraman, J. Gomez-Vidal, J. Coventry, Assessment of a novel
12 937 ternary eutectic chloride salt for next generation high-temperature sensible heat
13 938 storage, Energy Convers Manag. 167 (2018).
14
15 939 https://doi.org/10.1016/j.enconman.2018.04.100.
16
940 [23] R.D. Rogers, K.R. Seddon, Ionic Liquids - Solvents of the Future?, Science (1979).
17
18 941 302 (2003). https://doi.org/10.1126/science.1090313.
19
20 942 [24] C.M. Gordon, J.D. Holbrey, A.R. Kennedy, K.R. Seddon, Ionic liquid crystals:
21 943 Hexafluorophosphate salts, J Mater Chem. 8 (1998).
22 944 https://doi.org/10.1039/a806169f.
23
24 945 [25] W. Li, X. Wu, C. Qi, H. Rong, L. Gong, Study on the relationship between the
25 946 interaction energy and the melting point of amino acid cation based ionic liquids,
26 947 Journal of Molecular Structure: THEOCHEM. 942 (2010).
27
28
948 https://doi.org/10.1016/j.theochem.2009.11.027.
29
949 [26] R. Jiang, M. Gao, X. Mao, D. Wang, Advancements and potentials of molten salt CO2
30
31 950 capture and electrochemical transformation (MSCC-ET) process, Curr Opin
32 951 Electrochem. 17 (2019) 38–46. https://doi.org/10.1016/j.coelec.2019.04.011.
33
34 952 [27] H. Kawamura, Y. Ito, Electrodeposition of cohesive carbon films on aluminum in a
35 953 LiCl-KCl-K2CO3 melt, J Appl Electrochem. 30 (2000).
36 954 https://doi.org/10.1023/A:1003927100308.
37
38 955 [28] B. Kaplan, H. Groult, S. Komaba, N. Kumagai, F. Lantelme, Synthesis of
39 956 Nanostructured Carbon Material by Electroreduction in Fused Alkali Carbonates,
40
41
957 Chem Lett. 30 (2001) 714–715. https://doi.org/10.1246/cl.2001.714.
42 958 [29] I.A. Novoselova, N.F. Oliinyk, S.V. Volkov, A.A. Konchits, I.B. Yanchuk, V.S. Yefanov,
43
44 959 S.P. Kolesnik, M.V. Karpets, Electrolytic synthesis of carbon nanotubes from carbon
45 960 dioxide in molten salts and their characterization, Physica E Low Dimens Syst
46 961 Nanostruct. 40 (2008) 2231–2237. https://doi.org/10.1016/j.physe.2007.10.069.
47
48 962 [30] S. Licht, B. Wang, S. Ghosh, H. Ayub, D. Jiang, J. Ganley, A new solar carbon capture
49 963 process: Solar thermal electrochemical photo (STEP) carbon capture, Journal of
50 964 Physical Chemistry Letters. 1 (2010). https://doi.org/10.1021/jz100829s.
51
52 965 [31] K. Otake, H. Kinoshita, T. Kikuchi, R.O. Suzuki, CO2 gas decomposition to carbon by
53
966 electro-reduction in molten salts, Electrochim Acta. 100 (2013).
54
55 967 https://doi.org/10.1016/j.electacta.2013.02.076.
56
57 968 [32] H. Yin, L. Gao, H. Zhu, X. Mao, F. Gan, D. Wang, On the development of metallic inert
58 969 anode for molten CaCl2–CaO System, Electrochim Acta. 56 (2011) 3296–3302.
59 970 https://doi.org/10.1016/j.electacta.2011.01.026.
60
61
62
63
67
64
65
971 [33] Q. Song, Q. Xu, X. Shang, Z. Ning, Y. Qi, K. Yu, Electrochemical Preparation of a
1 972 Carbon/Cr-O-C Bilayer Film on Stainless Steel in Molten LiCl-KCl-K 2 CO 3 , J
2 973 Electrochem Soc. 162 (2015). https://doi.org/10.1149/2.1101501jes.
3
4 974 [34] L. Li, Z. Shi, B. Gao, J. Xu, X. Hu, Z. Wang, Electrochemical behavior of carbonate ion
5
975 in the LiF-NaF-Li2CO3 system, Electrochemistry. 82 (2014).
6
7 976 https://doi.org/10.5796/electrochemistry.82.1072.
8
9 977 [35] H. V. Ijije, C. Sun, G.Z. Chen, Indirect electrochemical reduction of carbon dioxide to
10 978 carbon nanopowders in molten alkali carbonates: Process variables and product
11 979 properties, Carbon N Y. 73 (2014). https://doi.org/10.1016/j.carbon.2014.02.052.
12
13 980 [36] J. Ge, L. Hu, W. Wang, H. Jiao, S. Jiao, Electrochemical Conversion of CO2 into
14 981 Negative Electrode Materials for Li-Ion Batteries, ChemElectroChem. 2 (2015).
15 982 https://doi.org/10.1002/celc.201402297.
16
17 983 [37] L. Hu, Y. Song, J. Ge, J. Zhu, S. Jiao, Capture and electrochemical conversion of CO
18 984
2 to ultrathin graphite sheets in CaCl 2 -based melts, J Mater Chem A Mater. 3 (2015)
19
20 985 21211–21218. https://doi.org/10.1039/C5TA05127D.
21
22
986 [38] T. Kikuchi, R. Ishida, S. Natsui, T. Kumagai, I. Ogino, N. Sakaguchi, M. Ueda, R.O.
23 987 Suzuki, Carbon nanotube synthesis via the calciothermic reduction of carbon dioxide
24 988 with iron additives, ECS Solid State Letters. 4 (2015).
25 989 https://doi.org/10.1149/2.0031509ssl.
26
27 990 [39] F. Matsuura, T. Wakamatsu, S. Natsui, T. Kikuchi, R.O. Suzuki, CO gas production by
28 991 molten salt electrolysis from CO2 gas, ISIJ International. 55 (2015).
29
992 https://doi.org/10.2355/isijinternational.55.404.
30
31 993 [40] J. Ren, F.-F. Li, J. Lau, L. González-Urbina, S. Licht, One-Pot Synthesis of Carbon
32
33 994 Nanofibers from CO 2, Nano Lett. 15 (2015) 6142–6148.
34 995 https://doi.org/10.1021/acs.nanolett.5b02427.
35
36 996 [41] H. Wu, Z. Li, D. Ji, Y. Liu, L. Li, D. Yuan, Z. Zhang, J. Ren, M. Lefler, B. Wang, S.
37 997 Licht, One-pot synthesis of nanostructured carbon materials from carbon dioxide via
38 998 electrolysis in molten carbonate salts, Carbon N Y. 106 (2016) 208–217.
39 999 https://doi.org/10.1016/j.carbon.2016.05.031.
40
411000 [42] L. Li, Z. Shi, B. Gao, X. Hu, Z. Wang, Electrochemical conversion of CO2 to carbon
421001 and oxygen in LiCl-Li2O melts, Electrochim Acta. 190 (2016).
43
441002 https://doi.org/10.1016/j.electacta.2015.12.202.
45
46
1003 [43] Y. Chen, Q. Xu, Q. Song, H. Li, Z. Ning, X. Lu, D.J. Fray, Electrochemistry of acetylide
471004 anion and anodic formation of carbon films in a LiCl–KCl–CaCl2–CaC2 melt,
481005 Electrochem Commun. 64 (2016) 1–4. https://doi.org/10.1016/j.elecom.2015.12.015.
49
501006 [44] J. Ge, J. Wang, J. Cheng, S. Jiao, Electrochemical Conversion of CO 2 in Molten
511007 CaCl 2 -LiCl-CaO Utilizing a Low-Cost (1- x )CaTiO 3 - x Ni Inert Anode , J
521008 Electrochem Soc. 163 (2016). https://doi.org/10.1149/2.1361608jes.
53
541009 [45] J. Ge, S. Wang, L. Hu, J. Zhu, S. Jiao, Electrochemical deposition of carbon in LiCl–
551010 NaCl–Na2CO3 melts, Carbon N Y. 98 (2016) 649–657.
56
571011 https://doi.org/10.1016/j.carbon.2015.11.065.
58
59
60
61
62
63
68
64
65
1012 [46] L. Hu, Y. Song, S. Jiao, Y. Liu, J. Ge, H. Jiao, J. Zhu, J. Wang, H. Zhu, D.J. Fray,
11013 Direct Conversion of Greenhouse Gas CO2 into Graphene via Molten Salts
21014 Electrolysis, ChemSusChem. 9 (2016). https://doi.org/10.1002/cssc.201501591.
3
41015 [47] B. Deng, Z. Chen, M. Gao, Y. Song, K. Zheng, J. Tang, W. Xiao, X. Mao, D. Wang,
5
1016 Molten salt CO 2 capture and electro-transformation (MSCC-ET) into capacitive
6
71017 carbon at medium temperature: effect of the electrolyte composition, Faraday
81018 Discuss. 190 (2016) 241–258. https://doi.org/10.1039/C5FD00234F.
9
101019 [48] H. Wu, Z. Li, D. Ji, Y. Liu, G. Yi, D. Yuan, B. Wang, Z. Zhang, Effect of molten
111020 carbonate composition on the generation of carbon material, RSC Adv. 7 (2017)
121021 8467–8473. https://doi.org/10.1039/C6RA25229J.
13
141022 [49] M.A. Hughes, J.A. Allen, S.W. Donne, The properties and performance of carbon
151023 produced through the electrochemical reduction of molten carbonate: A study based
16
1024 on step potential electrochemical spectroscopy, Electrochim Acta. 278 (2018) 340–
17
181025 351. https://doi.org/10.1016/j.electacta.2018.05.045.
19
201026 [50] L. Hu, B. Deng, K. Du, R. Jiang, Y. Dou, D. Wang, Tunable Selectivity and High
211027 Efficiency of CO2 Electroreduction via Borate-Enhanced Molten Salt Electrolysis,
221028 IScience. 23 (2020) 101607. https://doi.org/10.1016/j.isci.2020.101607.
23
241029 [51] W. Weng, B. Jiang, Z. Wang, W. Xiao, In situ electrochemical conversion of CO 2 in
251030 molten salts to advanced energy materials with reduced carbon emissions, Sci Adv. 6
261031 (2020). https://doi.org/10.1126/sciadv.aay9278.
27
281032 [52] J. Cao, S. Jing, H. Wang, W. Xu, M. Zhang, J. Xiao, Y. Peng, X. Ning, Z. Wang, W.
29
1033 Xiao, Pure and Metal‐confining Carbon Nanotubes through Electrochemical
30
311034 Reduction of Carbon Dioxide in Ca‐based Molten Salts, Angewandte Chemie. 135
321035 (2023). https://doi.org/10.1002/ange.202306877.
33
341036 [53] X. Liu, G. Licht, X. Wang, S. Licht, Controlled Transition Metal Nucleated Growth of
351037 Carbon Nanotubes by Molten Electrolysis of CO2, Catalysts. 12 (2022) 137.
361038 https://doi.org/10.3390/catal12020137.
37
381039 [54] Z. Yang, H. Yin, B. Deng, D. Wang, Electrolyte engineering for efficient molten-
391040 carbonate electrolysis of CO2, Chemical Engineering Journal. 473 (2023) 145146.
40
411041 https://doi.org/10.1016/j.cej.2023.145146.
42
1042 [55] H. Shi, M. Cai, W. Li, X. Chen, K. Du, L. Guo, P. Wang, P. Li, B. Deng, H. Yin, D.
43
441043 Wang, Direct molten-salt electro-reduction of CO2 in porous electrodes, Chemical
451044 Engineering Journal. 462 (2023) 142240. https://doi.org/10.1016/j.cej.2023.142240.
46
471045 [56] G. Dey, J. Ren, T. El-Ghazawi, S. Licht, How does an amalgamated Ni cathode affect
481046 carbon nanotube growth? A density functional theory study, RSC Adv. 6 (2016)
491047 27191–27196. https://doi.org/10.1039/C6RA03460H.
50
511048 [57] F. Silvearv, P. Larsson, Sarah.L.T. Jones, R. Ahuja, J.A. Larsson, Establishing the
521049 most favorable metal–carbon bond strength for carbon nanotube catalysts, J Mater
53
54
1050 Chem C Mater. 3 (2015) 3422–3427. https://doi.org/10.1039/C5TC00143A.
55
1051 [58] C. Tabtimsai, V. Ruangpornvisuti, B. Wanno, Density functional theory investigation of
56
571052 the VIIIB transition metal atoms deposited on (5,5) single-walled carbon nanotubes,
581053 Physica E Low Dimens Syst Nanostruct. 49 (2013) 61–67.
591054 https://doi.org/10.1016/j.physe.2013.01.019.
60
61
62
63
69
64
65
1055 [59] F. Ding, P. Larsson, J.A. Larsson, R. Ahuja, H. Duan, A. Rosén, K. Bolton, The
11056 Importance of Strong Carbon−Metal Adhesion for Catalytic Nucleation of Single-
21057 Walled Carbon Nanotubes, Nano Lett. 8 (2008) 463–468.
3
4
1058 https://doi.org/10.1021/nl072431m.
5
1059 [60] P. Larsson, J.A. Larsson, R. Ahuja, F. Ding, B.I. Yakobson, H. Duan, A. Rosén, K.
6
71060 Bolton, Calculating carbon nanotube-catalyst adhesion strengths, Phys Rev B
81061 Condens Matter Mater Phys. 75 (2007). https://doi.org/10.1103/PhysRevB.75.115419.
9
101062 [61] P. Atkins, J. de Paula, J. Keeler, Atkins’ Physical Chemistry, 2022.
111063 https://doi.org/10.1093/hesc/9780198847816.001.0001.
12
131064 [62] M. Johnson, J. Ren, M. Lefler, G. Licht, J. Vicini, S. Licht, Data on SEM, TEM and
141065 Raman Spectra of doped, and wool carbon nanotubes made directly from CO 2 by
151066 molten electrolysis, Data Brief. 14 (2017) 592–606.
16
1067 https://doi.org/10.1016/j.dib.2017.08.013.
17
181068 [63] M. Johnson, J. Ren, M. Lefler, G. Licht, J. Vicini, X. Liu, S. Licht, Carbon nanotube
19
201069 wools made directly from CO2 by molten electrolysis: Value driven pathways to
211070 carbon dioxide greenhouse gas mitigation, Mater Today Energy. 5 (2017).
221071 https://doi.org/10.1016/j.mtener.2017.07.003.
23
241072 [64] A.G.M. Da Silva, T.S. Rodrigues, S.J. Haigh, P.H.C. Camargo, Galvanic replacement
251073 reaction: Recent developments for engineering metal nanostructures towards catalytic
261074 applications, Chemical Communications. 53 (2017).
27
28
1075 https://doi.org/10.1039/c7cc02352a.
29
1076 [65] D.M. Tang, C. Liu, W.J. Yu, L.L. Zhang, P.X. Hou, J.C. Li, F. Li, Y. Bando, D. Golberg,
30
311077 H.M. Cheng, Structural changes in iron oxide and gold catalysts during nucleation of
321078 carbon nanotubes studied by in situ transmission electron microscopy, ACS Nano. 8
331079 (2014). https://doi.org/10.1021/nn403927y.
34
351080 [66] T.G. Mattos, F.D.A. Aarão Reis, Effects of diffusion and particle size in a kinetic model
361081 of catalyzed reactions, J Catal. 263 (2009). https://doi.org/10.1016/j.jcat.2009.01.011.
37
381082 [67] L. Zhu, C.W. Brian, S.F. Swallen, P.T. Straus, M.D. Ediger, L. Yu, Surface self-diffusion
391083 of an organic glass, Phys Rev Lett. 106 (2011).
40
41
1084 https://doi.org/10.1103/PhysRevLett.106.256103.
421085 [68] C. Huang, S. Ruan, T. Cai, L. Yu, Fast Surface Diffusion and Crystallization of
43
441086 Amorphous Griseofulvin, Journal of Physical Chemistry B. 121 (2017).
451087 https://doi.org/10.1021/acs.jpcb.7b07319.
46
471088 [69] H. Rabiee, L. Ge, X. Zhang, S. Hu, M. Li, Z. Yuan, Gas diffusion electrodes (GDEs)
481089 for electrochemical reduction of carbon dioxide, carbon monoxide, and dinitrogen to
491090 value-added products: A review, Energy Environ Sci. 14 (2021).
501091 https://doi.org/10.1039/d0ee03756g.
51
521092 [70] S. Garg, M. Li, A.Z. Weber, L. Ge, L. Li, V. Rudolph, G. Wang, T.E. Rufford, Advances
53
1093 and challenges in electrochemical CO 2 reduction processes: an engineering and
54
551094 design perspective looking beyond new catalyst materials, J Mater Chem A Mater. 8
561095 (2020) 1511–1544. https://doi.org/10.1039/C9TA13298H.
57
581096 [71] J. Ge, X. Zou, S. Almassi, L. Ji, B.P. Chaplin, A.J. Bard, Electrochemical Production of
591097 Si without Generation of CO2 Based on the Use of a Dimensionally Stable Anode in
60
61
62
63
70
64
65
1098 Molten CaCl2, Angewandte Chemie - International Edition. 58 (2019).
11099 https://doi.org/10.1002/anie.201905991.
2
31100 [72] P.C. Pistorius, D.J. Frayt, Formation of silicon by electrodeoxidation, and implications
41101 for titanium metal production, J South Afr Inst Min Metall. 106 (2006).
5
61102 [73] D.R. Sadoway, Inert anodes for the Hall-Héroult cell: The ultimate materials challenge,
71103 JOM. 53 (2001). https://doi.org/10.1007/s11837-001-0206-5.
8
91104 [74] S.M. Jeong, H.S. Shin, S.H. Cho, J.M. Hur, H.S. Lee, Electrochemical behavior of a
10
1105 platinum anode for reduction of uranium oxide in a LiCl molten salt, Electrochim Acta.
11
121106 54 (2009). https://doi.org/10.1016/j.electacta.2009.05.080.
13
141107 [75] K. Du, R. Yu, M. Gao, Z. Chen, X. Mao, H. Zhu, D. Wang, Durability of platinum
151108 coating anode in molten carbonate electrolysis cell, Corros Sci. 153 (2019) 12–18.
161109 https://doi.org/10.1016/j.corsci.2019.03.028.
17
181110 [76] H. Yin, X. Mao, D. Tang, W. Xiao, L. Xing, H. Zhu, D. Wang, D.R. Sadoway, Capture
191111 and electrochemical conversion of CO2 to value-added carbon and oxygen by molten
201112 salt electrolysis, Energy Environ Sci. 6 (2013) 1538.
21
22
1113 https://doi.org/10.1039/c3ee24132g.
231114 [77] C. Qian, H. Qi, B. Gao, Y. Cheng, Q. Qiu, L.-C. Qin, O. Zhou, J. Liu, Fabrication of
24
251115 Small Diameter Few-Walled Carbon Nanotubes with Enhanced Field Emission
261116 Property, J Nanosci Nanotechnol. 6 (2006) 1346–1349.
271117 https://doi.org/10.1166/jnn.2006.140.
28
291118 [78] K. Du, K. Zheng, Z. Chen, H. Zhu, F. Gan, D. Wang, Unusual temperature effect on
301119 the stability of nickel anodes in molten carbonates, Electrochim Acta. 245 (2017).
311120 https://doi.org/10.1016/j.electacta.2017.05.149.
32
331121 [79] S. Jiao, D.J. Fray, Development of an inert anode for electrowinning in calcium
34
1122 chloride-calcium oxide melts, Metallurgical and Materials Transactions B: Process
35
361123 Metallurgy and Materials Processing Science. 41 (2010).
371124 https://doi.org/10.1007/s11663-009-9281-8.
38
391125 [80] S. Zuca, M. Terzi, M. Zaharescu, K. Matiasovsky, Contribution to the study of SnO2-
401126 based ceramics, J Mater Sci. 26 (1991). https://doi.org/10.1007/bf00544681.
41
421127 [81] L. Hu, Y. Song, J. Ge, S. Jiao, J. Cheng, Electrochemical Metallurgy in CaCl 2 -CaO
431128 Melts on the Basis of TiO 2 ·RuO 2 Inert Anode , J Electrochem Soc. 163 (2016).
441129 https://doi.org/10.1149/2.0131603jes.
45
461130 [82] X. Cheng, H. Yin, D. Wang, Rearrangement of oxide scale on Ni-11Fe-10Cu alloy
47
1131 under anodic polarization in molten Na2CO3-K2CO3, Corros Sci. 141 (2018).
48
491132 https://doi.org/10.1016/j.corsci.2018.07.014.
50
511133 [83] E. Gavrilova, G. Goupil, B. Davis, D. Guay, L. Roué, On the key role of Cu on the
521134 oxidation behavior of Cu-Ni-Fe based anodes for Al electrolysis, Corros Sci. 101
531135 (2015). https://doi.org/10.1016/j.corsci.2015.09.006.
54
551136 [84] P. Guan, A. Liu, Z. Shi, X. Hu, Z. Wang, Corrosion behavior of Fe-Ni-Al alloy inert
561137 anode in cryolite melts, Metals (Basel). 9 (2019). https://doi.org/10.3390/met9040399.
57
581138 [85] F.-F. Li, S. Liu, B. Cui, J. Lau, J. Stuart, B. Wang, S. Licht, Solar Fuels: A One-Pot
591139 Synthesis of Hydrogen and Carbon Fuels from Water and Carbon Dioxide (Adv.
60
61
62
63
71
64
65
1140 Energy Mater. 7/2015), Adv Energy Mater. 5 (2015).
11141 https://doi.org/10.1002/aenm.201570039.
2
31142 [86] S. Licht, H. Wu, STEP Iron, a Chemistry of Iron Formation without CO2 Emission:
41143 Molten Carbonate Solubility and Electrochemistry of Iron Ore Impurities, The Journal
5
1144 of Physical Chemistry C. 115 (2011) 25138–25147. https://doi.org/10.1021/jp2078715.
6
71145 [87] S. Mahammadunnisa, E.L. Reddy, D. Ray, C. Subrahmanyam, J.C. Whitehead, CO2
8
91146 reduction to syngas and carbon nanofibres by plasma-assisted in situ decomposition
101147 of water, International Journal of Greenhouse Gas Control. 16 (2013).
111148 https://doi.org/10.1016/j.ijggc.2013.04.008.
12
131149 [88] K.Y. Shin, H.C. Su, C.H. Tsai, In situ growth of single-walled carbon nanotubes by
141150 bimetallic technique with/without dielectric support for nanodevice applications,
151151 Journal of Vacuum Science & Technology B: Microelectronics and Nanometer
16
1152 Structures. 24 (2006). https://doi.org/10.1116/1.2151223.
17
181153 [89] E.S. Rubin, J.E. Davison, H.J. Herzog, The cost of CO2 capture and storage,
19
201154 International Journal of Greenhouse Gas Control. 40 (2015).
211155 https://doi.org/10.1016/j.ijggc.2015.05.018.
22
231156 [90] D. Pletcher, The cathodic reduction of carbon dioxide - What can it realistically
241157 achieve? A mini review, Electrochem Commun. 61 (2015).
251158 https://doi.org/10.1016/j.elecom.2015.10.006.
26
271159 [91] B. Deng, Z. Chen, M. Gao, Y. Song, K. Zheng, J. Tang, W. Xiao, X. Mao, D. Wang,
281160 Molten salt CO2 capture and electro-transformation (MSCC-ET) into capacitive
29
1161 carbon at medium temperature: Effect of the electrolyte composition, Faraday
30
311162 Discuss. 190 (2016). https://doi.org/10.1039/c5fd00234f.
32
331163 [92] W. Xiao, D. Wang, The electrochemical reduction processes of solid compounds in
341164 high temperature molten salts, Chem Soc Rev. 43 (2014).
351165 https://doi.org/10.1039/c3cs60327j.
36
371166 [93] T. Usami, M. Kurata, T. Inoue, H.E. Sims, S.A. Beetham, J.A. Jenkins, Pyrochemical
381167 reduction of uranium dioxide and plutonium dioxide by lithium metal, Journal of
391168 Nuclear Materials. 300 (2002) 15–26. https://doi.org/10.1016/S0022-3115(01)00703-6.
40
411169 [94] H. FLOOD, T. FORLAND, The acidic and basic properties of oxides, Acta Chem
421170 Scand. 1 (1947). https://doi.org/10.3891/acta.chem.scand.01-0592.
43
441171 [95] C. Wang, H. Yu, H. Liu, Z. Jin, Thermodynamic optimization of the Na2O-B2O3
45
46
1172 pseudo-binary system, Journal of Phase Equilibria. 24 (2003).
471173 https://doi.org/10.1361/105497103770330965.
48
491174 [96] C. Halliday, T. Harada, T. Alan Hatton, Toward a Mechanistic Understanding and
501175 Optimization of Molten Alkali Metal Borates (AxB1-xO1.5-x) for Higherature CO2
511176 Capture, Chemistry of Materials. 32 (2020).
521177 https://doi.org/10.1021/acs.chemmater.9b03876.
53
541178 [97] Z. Yang, B. Deng, L. Hu, K. Du, H. Yin, D. Wang, Selective CO 2 Electroreduction with
551179 Enhanced Oxygen Evolution Efficiency in Affordable Borate-Mediated Molten
56
571180 Electrolyte, ACS Energy Lett. 8 (2023) 1762–1771.
581181 https://doi.org/10.1021/acsenergylett.3c00082.
59
60
61
62
63
72
64
65
1182 [98] E.I. Kamitsos, M.A. Karakassides, Structural studies of binary and pseudo binary
11183 sodium borate glasses of high sodium content, Physics and Chemistry of Glasses. 30
21184 (1989).
3
41185 [99] E.I. Kamitsos, G.D. Chryssikos, Borate glass structure by Raman and infrared
5
1186 spectroscopies, J Mol Struct. 247 (1991). https://doi.org/10.1016/0022-
6
71187 2860(91)87058-P.
8
91188 [100] D.K. Lindberg, R. V. Backman, Effect of Temperature and Boron Contents on the
101189 Autocausticizing Reactions in Sodium Carbonate/Borate Mixtures, Ind Eng Chem
111190 Res. 43 (2004) 6285–6291. https://doi.org/10.1021/ie040016o.
12
131191 [101] M. Cassir, G. Moutiers, J. Devynck, Stability and Characterization of Oxygen Species
141192 in Alkali Molten Carbonate: A Thermodynamic and Electrochemical Approach, J
151193 Electrochem Soc. 140 (1993). https://doi.org/10.1149/1.2220995.
16
171194 [102] S. Polizu, O. Savadogo, P. Poulin, L. Yahia, Applications of carbon nanotubes-based
181195 biomaterials in biomedical nanotechnology, J Nanosci Nanotechnol. 6 (2006).
19
201196 https://doi.org/10.1166/jnn.2006.197.
21
22
1197 [103] K. Donaldson, R. Aitken, L. Tran, V. Stone, R. Duffin, G. Forrest, A. Alexander, Carbon
231198 nanotubes: A review of their properties in relation to pulmonary toxicology and
241199 workplace safety, Toxicological Sciences. 92 (2006).
251200 https://doi.org/10.1093/toxsci/kfj130.
26
271201 [104] S. Karthikeyan, P. Mahalingam, M. Karthik, Large scale synthesis of carbon
281202 nanotubes, E-Journal of Chemistry. 6 (2009). https://doi.org/10.1155/2009/756410.
29
301203 [105] M.F.L. De Volder, S.H. Tawfick, R.H. Baughman, A.J. Hart, Carbon nanotubes:
311204 Present and future commercial applications, Science (1979). 339 (2013).
32
331205 https://doi.org/10.1126/science.1222453.
34
1206 [106] K. Kobashi, S. Ata, T. Yamada, D.N. Futaba, T. Okazaki, K. Hata, Classification of
35
361207 Commercialized Carbon Nanotubes into Three General Categories as a Guide for
371208 Applications, ACS Appl Nano Mater. 2 (2019) 4043–4047.
381209 https://doi.org/10.1021/acsanm.9b00941.
39
401210 [107] Y. Zheng, Y. Tian, S. Liu, X. Tan, S. Wang, Q. Guo, J. Luo, Z. Li, One-step microwave
411211 synthesis of NiO/NiS@CNT nanocomposites for high-cycling-stability supercapacitors,
421212 J Alloys Compd. 806 (2019) 170–179. https://doi.org/10.1016/j.jallcom.2019.07.213.
43
441213 [108] Y. Hou, J. Tang, H. Zhang, C. Qian, Y. Feng, J. Liu, Functionalized few-walled carbon
45
46
1214 nanotubes for mechanical reinforcement of polymeric composites, ACS Nano. 3
471215 (2009). https://doi.org/10.1021/nn9000512.
48
491216 [109] K. Kobashi, S. Ata, T. Yamada, D.N. Futaba, T. Okazaki, K. Hata, Classification of
501217 Commercialized Carbon Nanotubes into Three General Categories as a Guide for
511218 Applications, ACS Appl Nano Mater. 2 (2019) 4043–4047.
521219 https://doi.org/10.1021/acsanm.9b00941.
53
541220 [110] V. Shanov, Y.-H. Yun, M.J. Schulz, Synthesis and Characterization of Carbon
551221 Nanotube Materials (Review), Journal of the University of Chemical Technology and
56
571222 Metallurgy. 41 (2006).
58
59
60
61
62
63
73
64
65
1223 [111] N. Anzar, R. Hasan, M. Tyagi, N. Yadav, J. Narang, Carbon nanotube - A review on
11224 Synthesis, Properties and plethora of applications in the field of biomedical science,
21225 Sensors International. 1 (2020) 100003. https://doi.org/10.1016/j.sintl.2020.100003.
3
41226 [112] V. Sivamaran, V. Balasubramanian, M. Gopalakrishnan, V. Viswabaskaran, A. Gourav
5
1227 Rao, S.T. Selvamani, Carbon nanotubes, nanorings, and nanospheres: Synthesis and
6
71228 fabrication via chemical vapor deposition—a review, Nanomaterials and
81229 Nanotechnology. 12 (2022). https://doi.org/10.1177/18479804221079495.
9
101230 [113] N.C. Das, K. Yang, Y. Liu, P.E. Sokol, Z. Wang, H. Wang, Quantitative characterization
111231 of vertically aligned multi-walled carbon nanotube arrays using small angle x-ray
121232 scattering, in: J Nanosci Nanotechnol, 2011. https://doi.org/10.1166/jnn.2011.4110.
13
141233 [114] G. Atthipalli, R. Epur, P.N. Kumta, B.L. Allen, Y. Tang, A. Star, J.L. Gray, The effect of
151234 temperature on the growth of carbon nanotubes on copper foil using a nickel thin film
16
1235 as catalyst, Thin Solid Films. 519 (2011). https://doi.org/10.1016/j.tsf.2011.02.046.
17
181236 [115] N. Tripathi, P. Mishra, H. Harsh, S.S. Islam, Fine-tuning control on CNT diameter
19
201237 distribution, length and density using thermal CVD growth at atmospheric pressure:
211238 an in-depth analysis on the role of flow rate and flow duration of acetylene (C2H2)
221239 gas, Applied Nanoscience (Switzerland). 5 (2015). https://doi.org/10.1007/s13204-
231240 013-0288-8.
24
251241 [116] N.A. Fathy, Carbon nanotubes synthesis using carbonization of pretreated rice straw
261242 through chemical vapor deposition of camphor, RSC Adv. 7 (2017).
27
28
1243 https://doi.org/10.1039/c7ra04882c.
29
1244 [117] S. Ahmad, Y. Liao, A. Hussain, Q. Zhang, E.X. Ding, H. Jiang, E.I. Kauppinen,
30
311245 Systematic investigation of the catalyst composition effects on single-walled carbon
321246 nanotubes synthesis in floating-catalyst CVD, Carbon N Y. 149 (2019).
331247 https://doi.org/10.1016/j.carbon.2019.04.026.
34
351248 [118] V.M. Sivakumar, A.Z. Abdullah, A.R. Mohamed, S.P. Chai, Optimized parameters for
361249 carbon nanotubes synthesis over Fe and Ni catalysts VIA methane CVD, Reviews on
371250 Advanced Materials Science. 27 (2011).
38
391251 [119] A.E. Awadallah, S.M. Abdel-Hamid, D.S. El-Desouki, A.A. Aboul-Enein, A.K. Aboul-
40
41
1252 Gheit, Synthesis of carbon nanotubes by CCVD of natural gas using hydrotreating
421253 catalysts, Egyptian Journal of Petroleum. 21 (2012).
431254 https://doi.org/10.1016/j.ejpe.2012.11.005.
44
451255 [120] P.K. Tyagi, I. Janowska, O. Cretu, C. Pham-Huu, F. Banhart, Catalytic action of gold
461256 and copper crystals in the growth of carbon nanotubes, J Nanosci Nanotechnol. 11
471257 (2011). https://doi.org/10.1166/jnn.2011.3740.
48
491258 [121] K.A. Shah, B.A. Tali, Synthesis of carbon nanotubes by catalytic chemical vapour
501259 deposition: A review on carbon sources, catalysts and substrates, Mater Sci
51
521260 Semicond Process. 41 (2016). https://doi.org/10.1016/j.mssp.2015.08.013.
53
1261 [122] A. Gohier, C.P. Ewels, T.M. Minea, M.A. Djouadi, Carbon nanotube growth
54
551262 mechanism switches from tip- to base-growth with decreasing catalyst particle size,
561263 Carbon N Y. 46 (2008) 1331–1338. https://doi.org/10.1016/j.carbon.2008.05.016.
57
581264 [123] M. Kumar, Carbon nanotube synthesis and growth mechanism, Nanotechnol Percept.
591265 6 (2010) 7–28. https://doi.org/10.4024/N02KU10A.ntp.06.01.
60
61
62
63
74
64
65
1266 [124] H. Hongo, M. Yudasaka, T. Ichihashi, F. Nihey, S. Iijima, Chemical vapor deposition of
11267 single-wall carbon nanotubes on iron-film-coated sapphire substrates, Chem Phys
21268 Lett. 361 (2002). https://doi.org/10.1016/S0009-2614(02)00963-6.
3
41269 [125] B. Kitiyanan, W.E. Alvarez, J.H. Harwell, D.E. Resasco, Controlled production of
5
1270 single-wall carbon nanotubes by catalytic decomposition of CO on bimetallic Co-Mo
6
71271 catalysts, Chem Phys Lett. 317 (2000). https://doi.org/10.1016/S0009-2614(99)01379-
81272 2.
9
101273 [126] C. Mattevi, C.T. Wirth, S. Hofmann, R. Blume, M. Cantoro, C. Ducati, C. Cepek, A.
111274 Knop-Gericke, S. Milne, C. Castellarin-Cudia, S. Dolafi, A. Goldoni, R. Schloegl, J.
121275 Robertson, In-situ X-ray photoelectron spectroscopy study of catalyst-support
131276 interactions and growth of carbon nanotube forests, Journal of Physical Chemistry C.
14
151277 112 (2008). https://doi.org/10.1021/jp802474g.
16
1278 [127] S.P. Chai, S.H.S. Zein, A.R. Mohamed, Preparation of carbon nanotubes over cobalt-
17
181279 containing catalysts via catalytic decomposition of methane, Chem Phys Lett. 426
191280 (2006). https://doi.org/10.1016/j.cplett.2006.05.026.
20
211281 [128] M. Daraktchiev, B. Van De Moortèle, R. Schaller, E. Couteau, L. Forró, Effects of
221282 carbon nanotubes on grain boundary sliding in zirconia polycrystals, Advanced
231283 Materials. 17 (2005). https://doi.org/10.1002/adma.200400598.
24
251284
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
61
62
63
75
64
65
Conflict of Interest

Conflicts of interest

The authors declare that they have no known competing financial interests or personal

relationships that could have influenced the work reported in this paper.

You might also like