You are on page 1of 12

DOI: 10.1002/slct.

201702061 Full Papers

1
2 z Sustainable Chemistry
3
4
5
Sulfur-Immobilized Nitrogen and Oxygen Co–Doped
6 Hierarchically Porous Biomass Carbon for Lithium-Sulfur
7
8 Batteries: Influence of Sulfur Content and Distribution on
9
10
Its Performance
11 Reshma Chulliyote,[a] Haritha Hareendrakrishnakumar,[a] Murugan Raja,[b]
12
Joseph Mary Gladis,*[a] and Arul Manuel Stephan[b]
13
14
15
Hierarchically porous carbon with inherently doped heteroa- investigated as cathode materials for Li S battery. PC with 54%
16
toms and the quantity of active material (sulfur) confined sulfur displayed better performance with an initial discharge
17
within this carbon matrix play a major role for the high capacity of 1606 mA h g 1 and a cycling stability of
18
performance of Li S batteries. Herein, we discuss the influence 1269 mA h g 1 at 0.1C rate after 100 cycles due to better
19
of sulfur content and distribution onto the N and O co-doped dispersion of sulfur in the porous architecture. The higher
20
hierarchically porous biomass carbon matrix (PC) to achieve cycling stability of PCSC (54%) is due to the N and O co-doped
21
high specific capacitance and cycling stability. Sulfur encapsu- hierarchical porous carbon layers, enhancing the sulfur utiliza-
22
lated PC was prepared from an eco-friendly source with a high tion ratio and mitigating the polysulfide shuttle during the
23
surface area of 2065 m2 g 1 and a pore volume of 1.5 cm3 g 1. cycling process.
24
PC with 54, 68 & 73% of sulfur content (PCSCs) have been
25
26
27 Introduction
During this process, soluble sulfides reduced with lithium are
28
Rechargeable lithium battery plays a vital role in day-to-day life deposited on the anode surface, as the lower order polysulfides
29
due to its tremendous applications in smart phones, digital are insoluble in the electrolyte. This contributes to the
30
cameras, laptops, watches and so on. It is a great challenge to deprivation of active material, capacity waning and increase in
31
develop lithium batteries with high energy density for the wide cell resistance.[1–8]
32
variety of applications from micro to macro level. The lithium- Attempts have been made by various researchers to
33
sulfur (Li S) battery has a high theoretical energy density of improve the conductivity of sulfur by incorporating or wrap-
34
2567 W h kg 1 in which lithium metal, as the anode has a ping additives such as carbon nanotube, graphene, graphene
35
theoretical capacity of 3860 mA h g 1 and sulfur as cathode has oxide, carbon nanofiber,[9–12] polyindole,[13] polyaniline, polypyr-
36
a theoretical capacity of 1675 mA h g 1. Sulfur is identified as a role, polythiophene,[14] and porous carbons.[15–24] These reports
37
cathode material for Li S batteries due to its unique properties demonstrated improvements in cycling stability and active
38
such as low cost, non- toxic, abundance on earth and high material utilization. However, desired results are far from being
39
theoretical capacity.[1,2] In spite of these merits, the insulating achieved. Among the various materials investigated, porous
40
nature of sulfur and the end discharge product, Li2S limits its carbon proved to be highly effective to improve the utilization
41
discharging capacity.[3] On discharge, sulfur undergoes reduc- of active material and Li + ion transport due to its high surface
42
tion to form polysulfides of lithium Li2Sx (2 < x < 8). High order area, good conductivity and porous structure.[2] In addition,
43
polysulfides are generally soluble in the electrolyte and create a porous carbonaceous materials suppress polysulfide shuttling
44
redox shuttling between higher and lower order polysulfides. to some extent due to the very weak interactions with polar
45
polysulfides hence, capacity fading is observed for long term
46
cycling. This demands host materials with functional groups
47
[a] R. Chulliyote, H. Hareendrakrishnakumar, Dr. J. M. Gladis exhibiting high adsorption/binding capacity for polysulfides to
48 Department of Chemistry confine the species physically/chemically onto the porous
49 Indian Institute of Space Science and Technology
carbon architecture. Recently, there has been more progress to
50 Thiruvananthapuram, 695547 (India)
E-mail: marygladis@iist.ac.in develop porous carbon materials capable of confining poly-
51
jmargladis@gmail.com sulfides physically/chemically.[25] Sulfur is embedded in the
52 [b] Dr. M. Raja, Dr. A. M. Stephan pores of carbon in a highly dispersed state. In addition to the
53 Electrochemical Power Systems Division
high surface area and pore size, binding groups in the porous
54 CSIR-Central Electrochemical Research Institute
Karaikudi, 630006 (India) carbon, improve its interaction with sulfur to reduce the loss of
55
Supporting information for this article is available on the WWW under active materials. Most of the methods involved sophisticated
56
https://doi.org/10.1002/slct.201702061 techniques and are economically unfavourable, which also lack
57

ChemistrySelect 2017, 2, 10484 – 10495 10484  2017 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Wiley VCH
1732 / 102613
Montag, 13.11.2017
[S. 10484/10495] 1
Full Papers
consistency. Hence, it is essential to identify safe, low cost and Results and Discussion
1
environmentally benign material for large scale applications of
2 Material Characterization
energy storage technologies such as electric and hybrid
3
vehicles. A wide spectrum of porous carbons derived from The preparation of PC derived from pinecone at 800 8C using
4
plant and animal origin has been studied as cathode materials ZnCl2 as activating agent is illustrated in Scheme 1. ZnCl2 is
5
for Li S batteries.[26–43] Conventional porous carbon materials
6
are hydrophobic in nature, have weak interactions with
7
polysulfides.[43] Doping of hetero atoms (Nitrogen (N)/ Sulfur
8
(S)/ Oxygen (O)/ Boron (B)/ Phosphorous (P)) in the carbon
9
matrix reduces the charge transfer resistance, improves the
10
electronic conductivity, and enhances the active sites to
11
immobilize the sulfur/polysulfides.[44–47] Co-doping is a more
12
effective strategy to improve the physical adsorption and
13
chemical bonding to trap lithium polysulphides.[48] Codoping of
14
N and O intrinsically/extrinsically into the carbon framework
15
has pyridinic N, Pyrrolic N, graphitic N, C O and C=O etc.[49]
16
These functional groups with lone pair of electrons help to
17
improve the electronic conductivity and wettability of the
18
carbon matrices, which facilitate the confinement of polysul-
19
fides. Thus, dual doping of N and O atoms with high surface
20
area is expected to enhance the performance of cathode
21
material synergistically.[45–49] The porous structure with oxygen
22
and nitrogen group species help to retain the active material
23
and relieve the volumetric expansion during cycling.[48] Most of
24
the reported carbon materials with dopants were adopted
25
tedious multistep procedures and chemicals for synthesis. Thus
26 Scheme 1. Schematic illustration of the preparation of PC and PCSC.
a simple, scalable and a low cost method are worthy to
27
consider for realizing Li S batteries. Herein, we selected pine-
28
cone as the carbon source to obtain high surface area carbon
29
with hetero atoms (nitrogen and oxygen) for hosting sulfur due comparatively less toxic and non-corrosive. The amount of
30
to the widespread occurrence of pinecone all over the world. activated carbon obtained by ZnCl2 activation is more than
31
Pinecone or activated carbon from pinecone is used to adsorb KOH activation. It promotes the dehydration of the biomass
32
metal ions and organics from aqueous solution,[50] CO2 precursor at low temperatures and aromatic condensation
33
adsorption[51] and electrode material for supercapacitors.[52,53] reactions at high temperatures.[54,55] The nitrogen adsorption-
34
However, pinecone derived porous carbon is firstly reported as desorption isotherms and the corresponding DFT pore size
35
nitrogen and oxygen doped cathode material for Li S batteries. distribution curves of PC are shown in Figure 1. According to
36
The high surface area along with inherently doped N and O on
37
hierarchically porous carbon matrix dramatically increases the
38 Table 1. Surface area and pore volume for PC and PCSCs
performance of the cathode material. In order to understand
39
the role of sulfur distribution within the porous structure for PC PCSC PCSC PCSC
40 (54%) (68%) (73%)
effective sulfur utilization and capacity retention during cycling,
41
hierachically porous carbon (PC) with different sulfur composi- BET surface area (m2 g 1) 2065 439 105 10
42
tions (54, 68 & 73%) was prepared and characterized structur- Total pore volume 1.50 0.41 0.15 0.04
43 (cm3 g 1)
ally (SEM & XPS) and electrochemically. The results show that
44
the hierarchical porous carbon-sulfur composite (PCSC) with
45
54% of sulfur exhibits a high specific capacity of 1606 mA h g 1
46
with excellent capacity retention of 79% after 100 cycles with IUPAC, the adsorption curve of PC is a type IV isotherm with
47
0.29% average fading per cycle at 0.1C rate and a good rate adsorption hysteresis loop indicating the characteristic range of
48
capability (923 mA h g 1 at 2C) as a cathode material for Li S pores from meso to micropores. The rapid adsorption at low
49
battery. The physical adsorption of polysulfides onto the porous relative pressure suggests the existence of micropores and the
50
carbon and chemical binding of polysulfides through O and N hysteresis loop in Figure 1a exemplify mesopores. The exten-
51
dopants increase the electrochemical performance and dimin- sion in the curve at a relative pressure greater than 0.9 is due
52
ish the active material loss during the cycling process. The to the existence of macropores.[27–31] This is evidenced by
53
excellent electrochemical performance is due to the synergic Figure 1b depicts the pore size distribution of PC and PCSCs. It
54
effect of the hierarchically porous features of the activated has a small fraction of macropores along with mesopores and
55
carbon with oxygen and nitrogen binding sites to immobilise micropores distributed in the PC revealing a hierarchical porous
56
the active polysulfide species. structure. The high surface area (2065 m2 g 1) and pore volume
57

ChemistrySelect 2017, 2, 10484 – 10495 10485  2017 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Wiley VCH
1732 / 102613
Montag, 13.11.2017
[S. 10485/10495] 1
Full Papers

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18 Figure 1. (a) Nitrogen adsorption/desorption isotherm (b) Pore size distribution curve of PC [inset: enlarged at 0–10 nm].
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38 Figure 2. (a) X-ray diffractograms of sulfur, PC, and PCSCs (b) Raman spectra for PC and PCSCs.
39
40
41
(1.5 cm3 g 1) reveals the well-developed porous structure. The no. 00–001-0478). The XRD pattern of PC exhibits two broad
42
BET surface area of PC, 2065 m2 g 1 is reduced to 410 m2 g 1 peaks at 2q = 268 and 438 corresponding to (002) and (100)
43
(54%), 100 m2 g 1 (68%) and 10 m2 g 1 (73%) after sulfur reflections of amorphous carbon and disordered graphitic
44
loading. The total pore volume of PC is 1.5 cm3 g 1 and PCSCs structure respectively. During the preparation process of PCSCs,
45
are 0.406 cm3 g 1 (54%), 0.154 cm3 g 1 (68%) and 0.037 cm3 g 1 sulfur melts and diffuse into the pores of the carbon and is
46
(73%). It is observed that the surface area and pore volume are trapped inside the network due to the strong adsorbing
47
decreased when the sulfur content is increased on PC indicates capacity of a PC. The PCSCs exhibits less intense diffraction
48
that the pores are embedded with sulfur so that PC act as an peaks of crystalline sulfur indicating the presence of trace
49
ideal host for sulfur. When the sulfur content is 54%, a fraction quantities of bulk crystalline sulfur on the surface, in addition
50
of pores are either partially filled or unfilled in the composite to the well-adsorbed sulfur inside the porous network. The
51
enhances the accessibility of electrolyte, hence good Li-ion structure of the PC and PCSCs are further confirmed by Raman
52
transport during cycling. spectrum as shown in Figure 2b. The peaks obtained at
53
XRD patterns of PC and PCSCs at room temperature are 1336 cm 1 (D band) and 1595 cm 1 (G band) in the Raman
54
shown in Figure 2a. The diffraction peaks for the elemental spectrum of PC and PCSCs are attributed to the disordered
55
sulfur in the wide range of 2q = 158 - 608 match well with the amorphous carbon and the crystalline graphite respectively.[56–
56 58]
characteristic diffraction peaks of orthorhombic sulfur (JCPSD The broadening and the intensity ratio ID/IG, provides
57

ChemistrySelect 2017, 2, 10484 – 10495 10486  2017 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Wiley VCH
1732 / 102613
Montag, 13.11.2017
[S. 10486/10495] 1
Full Papers

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18 Figure 3. (a)Thermogram of sulfur, PC and PCSCs (b) FTIR spectrum of PC and PCSCs.
19
20
21
evidence for the degree of defects present in carbon materials. 1148 cm 1, 1042 cm 1 is due to stretching vibrations of C O,
22
Broadening of peaks may be due to the deviation in the C N and N O stretching vibrations of the PC and PCSCs. These
23
interplanar bond lengths and angles. The ID/IGratio for PC is results prove that, the presence of oxygen and nitrogen
24
0.808 which indicates the disordered crystalline carbon struc- functionalities on the surface of PC and PCSCs. After incorpo-
25
ture of the material. However, the ID/IG ratio of PCSCs is ration of sulfur into the PC, we have observed that the new
26
relatively more than that of PC which indicates that the defects peaks emerged at 1100 cm 1, 918 cm 1 and 618 cm 1 is due to
27
get more pronounced with the increase in sulfur impregnation. stretching vibration of S O, N S and C S bonds respectively.
28
Crystalline sulfur exhibits three peaks at 148 cm 1, 214 cm 1 Herein, we hypothesised that the heteroatoms with lone pairs
29
and 470 cm 1 (Figure S1a).[56–59] The observed sulfur peaks in may interact with polysulfides which further leads to the
30
PCSCs indicating that sulfur is incorporated into the carbon anchoring of polysulfides on cathode during the charge-
31
matrix successfully by melt diffusion process. discharge process.[60,61]
32
The TGA curve of PC in the presence of air demonstrates CHNS analysis demonstrated the presence of oxygen (14%)
33
99.09% purity of the carbon material (Figure S1b). The fraction and nitrogen (2%) in the PC (Table S1). The nature of surface
34
of sulfur present in PCSCs is measured using thermogravimetric functional groups in the PC is investigated in detail by XPS
35
analysis (Figure 3a). TGA curve of PC is quite stable; decom- analysis. The XPS survey scan (Figure S1c) reveals the presence
36
position below 100 8C is due to the elimination of water during of C1s, O1s and N1s in the porous carbon at 283.7 eV, 400 eV
37
TGA analysis. TGA curves of PCSCs show major weight loss in and 533.1 eV respectively. The nitrogen (2.22%) and oxygen
38
the temperature range 200–400 8C. The sulfur content is (14.67%) content obtained from XPS analysis (Table S2) are very
39
calculated as 54%, 68% and 73% based on the weight loss of close to the CHNS results. The high resolution C1s (Figure 4a)
40
sulfur from 100 to 400 8C. The weight loss in PCSCs starts at a peak divided into 5 components, C C bond (284.7 eV) of SP2,
41
higher temperature than the elemental sulfur due to the well C N (285.4), C O bond (286.6 eV), C=N bond (287.3 eV) and C=
42
dispersed sulfur within the micro and mesopores of PC. The O (288.6 eV).[46–47] The high resolution O1s spectrum (Figure 4b)
43
weight loss in PCSCs (68% and 73%) commences at a lower is subdivided into three components: N O bond (530.5), C O
44
temperature (166 8C) than PCSC (54%), indicating the presence bond (531.7 eV), and C=O bond (533.1 eV). 27 at. % of O
45
of sulfur at the surface of the porous carbon. The second species in PC is N O, 33 at. % C O and 40 at. % C=O which are
46
weight loss (~ 250 8C) is attributed to the confined sulfur within believed to promote the chemical adsorption of sulfur. XPS
47
the porous network. TGA results are well agreed with XRD and results reveal that PC matrix is rich with oxygen functional
48
Raman analysis. The presence of functional groups on the groups.[44,45] The N1s spectrum (Figure 4c) contains 4 contribu-
49
surface of porous carbon and composites is determined by tions at 398.2 eV, 400 eV, 401.7 eV and 403.7 eV which are
50
FTIR spectroscopy. Figure 3b shows the IR spectrum of PC and assigned to the pyridinic, pyrrolic, graphitic and nitrogen oxide
51
PCSCs. The band at 3455 cm 1and 3267 cm 1 is due to respectively, which includes 27 at. % pyridinic nitrogen, 34 at.%
52
stretching vibrations of O H and N H bonds respectively. The pyrrolic nitrogen, 25 at.% graphitic nitrogen and 14 at.% N O
53
band at 2939 cm 1 and 2854 cm 1 are due to symmetric and species.[46,47–59] These functional groups are consistent with the
54
asymmetric stretching vibrations of C H bonds. The band at FTIR spectral results. The above XPS results confirm the intrinsic
55
1725 cm 1 and 1648 cm 1 are due to stretching vibrations of doping of nitrogen and oxygen onto porous carbon matrix
56
C=O and C=N in the PC and PCSCs. The band at 1247 cm 1,
57

ChemistrySelect 2017, 2, 10484 – 10495 10487  2017 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Wiley VCH
1732 / 102613
Montag, 13.11.2017
[S. 10487/10495] 1
Full Papers

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
Figure 4. XPS for PC (a) C1s (b) O1s (c) N1s and (d) Schematic illustration of nitrogen and oxygen containing porous carbon.
35
36
37
during pyrolysis. A schematic representation of PC is exhibited oxide are 8, 34, 25, 21 and 10 respectively. New S O and N S
38
in Figure 4d based on the above observations.[62,63] bonds indicates that the nitrogen and oxygen groups interact
39
XPS studies of PCSC (54%) was carried out to explore the with sulfur in the composite.[46,47–49] Figure.5d displays the
40
surface chemistry of PC after sulfur incorporation and to study deconvoulted S2p spectrum of PCSC (54%) has S2p3/2 and
41
the interaction between heteroatoms and sulfur. The survey 2p1/2 with an energy separation of 1.2 eV and an intensity ratio
42
scans of the composite (Figure S1c) reveal the presence of S2s of 2:1. Pristine sulfur shows S2p doublet with a binding energy
43
and S2p at 240 eV and 163.4 eV respectively. The C1s (Figure. of 164.3-165.5 eV.[65] After impregnation of sulfur onto PC, S S
44
5a) is split into 6 components, C C bond (284.7 eV) of SP2, C N bond is slightly shifted to 163.6-164.8 compared to sulfur
45
(285.4), C O bond (286.6 eV), C=N bond (287.3 eV) and C=O powder may be due to the changes in the electronic
46
(288.6 eV). Also, the existence of a new C S bond at 285.2 eV distribution on the sulfur after melt diffusion which indicates
47
reveals the effective doping of sulfur on to the carbon the possibility of C S bond.[50,61,64,66,67] Further, sulfur is bonded
48
matrix.[44,45,64] The high resolution O1s spectrum (Figure. 5b) is to both nitrogen and oxygen [S N (163.8-165) and S O (164.1-
49
subdivided into four components: C O bond (530.5 eV), N O 165.3)] in the carbon matrix.[13,66] It is expected that the
50
bond (530.5), C=O bond (533.1 eV) and a new S O bond interaction of sulfur with carbon matrix through S O and S N
51
(531.1 eV), 17 at.% of O species in PC are S O, 19 at.% is N O, bonds effectively confine polysulfides during charge discharge
52
24 at.% is C O and 40 at.% C=O. The N1s spectrum (Figure. 5c) process.[68] These major changes in the XPS spectra before and
53
contains 5 contributions at 398.2 eV, 400 eV, 401.7 eV and after melt diffusion, indicates the interaction of sulfur with
54
403.7 eV which are assigned to the pyridinic, pyrrolic, graphitic heteroatoms and modifies the electronic distribution around
55
and nitrogen oxide respectively and a new N S at 396.2 eV. The atoms that is reflected in the binding energy shift. The
56
at. % of pyridinic, pyrrolic, graphitic, N S and pyridinic nitrogen heteroatoms with lone pairs act as lewis base and it interacts
57

ChemistrySelect 2017, 2, 10484 – 10495 10488  2017 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Wiley VCH
1732 / 102613
Montag, 13.11.2017
[S. 10488/10495] 1
Full Papers

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 Figure 5. XPS for PCSC (a) C1s (b) O1s (c) N1s and (d) S2p.
36
37
38
with Li (Lewis acid) during charge discharge process that helps carbon matrix, SAED pattern (inset) displays the crystalline
39
to anchor polysulfides. The O and N functional groups nature of the PCSCs which confirms the presence of sulfur in
40
synergically contribute to improve the performance of Li S the system. The hierarchical porous structure of carbon matrix
41
battery along with the high surface area of the carbon is helpful during volume changes in lithium-sulfur batteries.The
42
material.[69] The morphology and microstructure of the PC and presence of nitrogen and oxygen in the PC as well as nitrogen,
43
PCSCs were investigated in detail by SEM and TEM which is oxygen and sulfur in the PCSCs are further confirmed by
44
presented in Figure 6 and 7 respectively. SEM of PC (Figure 6a- elemental mapping analysis (Figure S2-S5).
45
c) reveals that the prepared carbon materials have a hierarch-
46
ical morphology with interconnected pores. The TEM image of
47 Electrochemical performance of the composite
a PC (Figure 6d) displays a large number of nano pores, which
48
affirms the high surface area and porosity of the material. SAED The electrochemical properties of PCSCs have been evaluated
49
pattern (inset) exhibits the amorphous nature of the PC. It is by assembling CR2032 type coin cell with metallic lithium as
50
obvious from the morphological changes after infiltrating sulfur the counter electrode. The cyclic voltammograms of PCSCs in a
51
into the micro, meso and macro pores of PC, suggesting potential range of 1.6–3 V at a scan rate 0.1 mV s 1 are shown
52
uniform diffusion of sulfur into the pores (Figure 7a-c). Also, in Figure 8a and Figure S6. The two reduction and one
53
bulk crystalline sulfur is observed on the surface of PCSCs (68% oxidation peaks are observed due to the multistep reaction
54
and 73%) in addition to the diffused sulfur. These results between sulfur and lithium. In the cathodic scan process, the
55
corroborate with TGA and XRD analysis. TEM image (Figure 7d) peak at 2.36 V is due to the reduction of elemental sulfur to
56
of PCSC (54%) shows no agglomerated sulfur particle onto the long chain lithium polysulfides (Li2Sx, 4  x  8). The second
57

ChemistrySelect 2017, 2, 10484 – 10495 10489  2017 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Wiley VCH
1732 / 102613
Montag, 13.11.2017
[S. 10489/10495] 1
Full Papers

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
Figure 6. SEM of (a) PC (b) and (c) are enlarged portion of (a) and (d) TEM image of PC (inset: SAED pattern).
38
39
40
peak at 2.01 is due to the conversion of these long chain upto 100th charge-discharge profiles of PCSCs at 0.1C rate. The
41
polysulfides to short chain polysulfides (Li2Sx, 2  x  4). In the charge-discharge profiles of PCSCs exhibit two plateaux (~
42
subsequent anodic scan process, a peak at 2.39 V is observed, 2.36 V and 2.01 V) in the discharge curve and one plateau
43
which corresponds to the conversion of these polysulfides to (2.39 V) at the charging curve aligning with the two reduction
44
sulfur.[30–33] In the successive cycles, the peak positions, the peaks and one oxidation peak of CV. The second plateau is due
45
anodic and cathodic current are almost similar ascribing the to the formation short chain lithium sulfides, exhibiting stable
46
reactive reversibility and cycling stability of the developed voltage during cycling indicates better retention in the electro-
47
electrode material. The anodic and cathodic peak positions of chemical performance. This may be ascribed to the N and O
48
PCSCs with 54, 68 & 73% sulfur content are same. However, a containing hierarchically porous carbon host with micro and
49
decrease in the current density is observed when the sulfur mesopores enhance the electronic and ionic transport at the
50
content in the composite increases (Figure S6). This may be cathode. An impressive high discharge capacity of
51
due to the poor conductivity of sulfur crystals at the surface of 1606 mA h g 1 is observed for PCSC (54%) (Figure 8b), com-
52
PCSC 68% & 73% owing to the high sulfur content as indicated pared to similar material reported elsewhere (Table S3). PCSC
53
by N2 adsorption studies and SEM.[70] (68%) and PCSC (73%) display an initial discharge capacity of
54
The electrochemical performance of the sulfur-carbon 1370 mA h g 1 and 1198 mA h g 1 respectively (Figure 8c and d).
55
composite is further tested by galvanostatic charge-discharge The amount of active material utilization is decreased as the
56
experiments. Figure 8 (b, c and d) shows the 1, 2, 10, 20, 30, sulfur content increases. The sulfur utilization is 95.9%, 81.8%
57

ChemistrySelect 2017, 2, 10484 – 10495 10490  2017 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Wiley VCH
1732 / 102613
Montag, 13.11.2017
[S. 10490/10495] 1
Full Papers

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 Figure 7. SEM of (a) PCSC (54%) (b) PCSC (68%) (c) PCSC (73%) and (d) TEM image of PCSC (54%) (Inset: SAED pattern).
38
39
40
and 71.5% for PCSC (54%), PCSC (68%) and PCSC (73%) Cycling performance and coulombic efficiency of PCSCs at
41
respectively. The reason could be the uniform dispersion of 0.1C rate is shown in Figure 9a. It exhibits good electrochemical
42
sulfur within the pores is feasible in the PCSC (54%) which performance and coulombic efficiency of more than 99% up to
43
improves the ionic transport. In the second cycle its discharge 100 cycles, due to the reduced shuttling effect. The rate
44
capacities become 1531 mA h g 1, 1370 mA h g 1 and capability of PCSCs at various current densities from 0.1C to 2C
45
1198 mA h g 1 with 4.6%, 8.9% and 11.1% capacity degradation is shown in Figure 9b. The discharge capacity of PCSC (54%,
46
for PCSC (54%), PCSC (68%) and PCSC (73%) respectively. After 68% and 73%) at various current densities of 0.1C, 0.2C, 0.5C,
47
100 cycles, PCSC (54%) shows a reversible capacity of 1C and 2C are 1606 mA h g 1, 1435 mA h g 1,1265 mA h g 1,
48
1269 mA h g 1 which is higher than that of sulfur rich PCSCs 1027 mA h g 1 & 923 mA h g 1 (54%), 1370 mA h g 1,
49
(1028 mA h g 1 and 891 mA h g 1) with 79% capacity retention 1 1
1172 mA h g , 1026 mA h g , 819 mA h g 1
& 730 mA h g 1
50
compared to initial discharge capacity. These results are (68%) and 1198 mA h g , 1002 mA h g , 840 mA h g 1,
1 1
51
attributed to the heteroatoms induced chemical adsorption of 606 mA h g 1 & 513 mA h g 1 (73%) respectively. There is no
52
polysulfides in the hierarchically porous carbon matrix.[25] The specific capacity loss is observed while discharging at different
53
specific capacity may decrease if the capacity was normalized current densities. When the current density, reduced back to
54
by the weight of the whole electrode. Since the active material 0.2C the discharge capacity is recovered, indicating an excellent
55
in the electrode is less as sulfur loading is decreased. rate performance of PCSC cathodes. These higher specific
56
capacities at different current densities are appreciable to
57

ChemistrySelect 2017, 2, 10484 – 10495 10491  2017 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Wiley VCH
1732 / 102613
Montag, 13.11.2017
[S. 10491/10495] 1
Full Papers

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 Figure 8. (a) Cyclic voltammograms of the PCSCs 54% and Galvanostatic charge-discharge curves of PCSCs (b) 54% (c) 68% (d) 73%.
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51 Figure 9. Cyclic stability performances and coulombic efficiency of PCSCs (a) PCSCs and (b) rate capability studies of the PCSCs at different C rate.
52
53
54
obtain high energy density and power density. The outstanding atoms with lone pairs which facilitate chemical adsorption of
55
rate capabilities may be due to the good conductivity, surface polysulfides and Li + diffusion during charge-discharge proc-
56
area, and pore volume of the carbon material and the hetero esses.
57

ChemistrySelect 2017, 2, 10484 – 10495 10492  2017 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Wiley VCH
1732 / 102613
Montag, 13.11.2017
[S. 10492/10495] 1
Full Papers
The adsorption of polysulphides by PC was investigated These results are in good agreement with excellent cycling
1
using the UV visible absorption spectra of polysulfide solution stability and rate capability of the cathode material.
2
with PC (Figure 10). Inset shows the digital photographs of the EIS was carried out to understand the electrochemical
3
reaction between lithium and sulfur when PCSCs are used as
4
cathode material for lithium sulfur battery. EIS before and after
5
100 cycles at the OCV voltage is shown in Figure 11. The
6
impedance spectra consist of depressed semicircle at high
7
frequency region and a sloping line in the low frequency
8
region. Semicircle in the high frequency region relates to
9
charge transfer resistance between electrode/electrolyte inter-
10
faces. A slope in the low frequency region is called as Warburg
11
resistance attributed to the Li + ion diffusion within the
12
cathode. The Resistance of the electrolyte (Re) determined from
13
the intersection of the front end of the semicircle on the Z’ axis,
14
Re is comparatively more for PCSCs with higher sulfur content
15
which can be correlated to the less active material utilization
16
due to the accumulation of sulfur on the surface of carbon
17
matrix (Figure 11 a and b). PCSCs (54%, 68% and 73%) exhibit
18
two depressed semicircles at high and medium frequency
19
region and a slope at low frequency region. It indicates two
20
types of resistances offered by the system due to charge
21 Figure 10. UV-vis absorption spectra of polysulfide solution with PC. Inset is
transfer reaction between electrode-electrolyte interfaces of
22 the digital images of the polysulfide adsorption test [(a) lithium polysulfide
solution (b) lithium polysulfide mixed with PC after 0 min. (c) lithium PCSCs.[71,72] Nyquist impedance plot is evaluated by a simple
23
polysulfide mixed with PC after 10 min.]. circuit model which is shown in the inset of Figure 11 &
24
Figure S7. The partial semi-circle at high frequency region is
25
related to the resistance offered for the charge transfer
26
lithium polysulfide solution and the variation in the colour of between the porous carbon material and electrolyte i. e.,
27
polysulfide solution after adding PC. UV visible spectra show electrical double layer resistance.[71,73] The semi-circle at the
28
the broad absorption in between 200–600 nm for various medium frequency region is related to the charge transfer
29
sulfide species in the case of control polysulfide solution. resistance offered due to the interaction between sulfur and
30
Initially it is a homogenous mixture of PC and polysulfide electrolyte; it is due to faradaic process. It is observed that the
31
solution, after 10 minutes the solution become colourless diameter of small semicircle i. e. charge transfer resistance is
32
indicating a strong interaction between PC and polysulfides. A very small for PCSC (54%) when compared to PCSC (68% and
33
dramatic reduction in the peak intensity is observed in the UV- 73%). It may be due to the low density sulfur in the meso/
34
visible spectra of the solution after 10 minutes demonstrate the micropores favor the diffusion of electrolyte hence the trans-
35
excellent adsorption of lithium polysulfides onto PC.[43,45,71] portation of lithium ion. Figure 11 b represents an EIS analysis
36
after 100 cycles, It is clear that charge transfer resistance is
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56 Figure 11. Nyquist plot of the electrodes (a) before discharge and (b) after 100 cycles at the OCV [inset: fitting equivalent circuit diagram].
57

ChemistrySelect 2017, 2, 10484 – 10495 10493  2017 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Wiley VCH
1732 / 102613
Montag, 13.11.2017
[S. 10493/10495] 1
Full Papers
decreased as a result of enhanced electronic conductivity upon [11] D.-W. Wang, Q. Zeng, G. Zhou, L. Yin, F. Li, H.-M. Cheng, I. R. Gentle,
1 G. Q. M. Lu, J. Mater. Chem. A 2013, 1, 9382–9394.
cycling, may be due to the well disseminated electrolyte in the
2 [12] W. Wang, Y. Wang, Y. Huang, C. Huang, Z. Yu, H. Zhang, A. Wang, K.
cathode.[74] The disappearance of partial semicircle after 100 Yuan, J. Appl. Electrochem. 2010, 40, 321–325.
3
cycles attributed to the non-availability of free active carbon [13] R. Chulliyote, H. Hareendrakrishnakumar, M. Raja, J. M. Gladis, A. M.
4
sites due to the physical /chemical adsorption of active material Stephan, Sustainable Energy Fuels 2017, 1, 1774–1781.
5 [14] H. Cheng, S. Wang, J. Mater. Chem. A 2014, 2, 13783–13794.
during the charge-discharge process. These results further
6 [15] N. Jayaprakash, J. Shen, S. S. Moganty, A. Corona, L. A. Archer, Angew.
confirm the ability of N and O containing porous network to Chem. Int. Ed. 2011, 50, 5904–5908.
7
confine polysulfides and enhances cycling performance of [16] J. Schuster, G. He, B. Mandlmeier, T. Yim, K. T. Lee, T. Bein, L. F. Nazar,
8
hierarchical porous carbon sulfur composites. Angew. Chem. Int. Ed. 2012, 51, 3591–3595.
9 [17] Y.-S. Su, A. Manthiram, Nat. Commun. 2012, 3, 1166.
10 [18] C. Liang, N. J. Dudney, J. Y. Howe, Chem. Mater. 2009, 21, 4724–4730.
11 Conclusion [19] C. Zhao, L. Liu, H. Zhao, A. Krall, Z. Wen, J. Chen, P. Hurley, J. Jiang, Y. Li ,
Nanoscale 2014, 6, 882–888.
12
In summary, the hierarchically porous, N and O inherently co- [20] Y. Qu, Z. Zhang, X. Zhang, G. Ren, X. Wang, Y. Lai, Y. Liu, J. Li, Electrochim.
13 Acta 2014, 137, 439–446.
doped carbon with high surface area from a renewable raw
14 [21] M. Gao, X. Xiong, W. Wang, S. Zhao, C. Li, H. Zhang, Z. Yu, Y. Huang, J.
material have been explored to develop a high performance Power Sources 2014, 248, 1149–1155.
15
cathode material for Li S batteries. The cyclability and the [22] D. Wang, A. Fu, H. Li, Y. Wang, P. Guo, J. Liu, X. S. Zhao, J. Power Sources
16
specific capacity of the system is influenced by the sulfur 2015, 285, 469–477.
17 [23] J. Yang, S. Wang, Z. Ma, Z. Du, C. Li, J. Song, G. Wang, G. Shao,
content and distribution in the porous matrix with high surface
18 Electrochim. Acta 2015, 159, 8–15.
area 2065 m2 g 1 and the dopants (N&O). PCSC (54%) cathode [24] J. Wang, Y. Wu, Z. Shi, C. Wu, Electrochim. Acta 2014, 144, 307–314.
19
demonstrates 1606 mAhg 1 initial capacity and sustains a [25] S. S. Zhang, Inorg. Chem. Front. 2015, 2, 1059–1069.
20
capacity of 1269 mA h g 1 after 100 cycles at 0.1C rate and a [26] S. Zhang, M. Zheng, Z. Lin, N. Li, Y. Liu, B. Zhao, H. Pang, J. Cao, P. He, Y.
21 Shi, J. Mater. Chem. A 2014, 2, 15889–15896.
good rate capability (923 mA h g 1 at 2C). The sulfur utilization
22 [27] S. Zhao, C. Li, W. Wang, H. Zhang, M. Gao, X. Xiong, A. Wang, K. Yuan, Y.
is decreased with increase of sulfur content on PCSCs due to Huang, F. Wang, J. Mater. Chem. A 2013, 1, 3334–3339.
23
the accumulation of sulfur on the surface of porous carbon [28] S. Wei, H. Zhang, Y. Huang, W. Wang, Y. Xia, Z. Yu, Energy Environ Sci.
24
matrix. PCSC (54%) had also been firstly explored as a cathode 2011, 4, 736–740.
25 [29] K. Yang, Q. Gao, Y. Tan, W. Tian, L. Zhu, C. Yang, Microporous Mesoporous
material displays 79% retention with 0.29% capacity fading per
26 Mater. 2015, 204, 235–241.
cycle. Furthermore, being a low cost, easily available and facile [30] J. Xu, K. Zhou, F. Chen, W. Chen, X. Wei, X.-W. Liu, J. Liu, ACS Sustain.
27
N and O co-doped hierarchically porous carbon is an ideal host Chem. Eng. 2016, 4, 666–670.
28
for sulfur hence; PCSCs may be a promising cathode material [31] M. Xu, M. Jia, C. Mao, S. Liu, S. Bao, J. Jiang, Y. Liu, Z. Lu, Sci. Rep. 2016, 6,
29 18739.
for Li S batteries.
30 [32] Y. Qu, Z. Zhang, X. Zhang, G. Ren, Y. Lai, Y. Liu, J. Li, Carbon 2015, 84,
31 399–408.
[33] S.-H. Chung, C.-H. Chang, A. Manthiram, ACS Nano 2016, 10, 10462–
32 Acknowledgment 10470.
33 [34] G. Ren, S. Li, Z.-X. Fan, J. Warzywoda, Z. Fan, J. Mater. Chem. A 2016, 4,
Financial support from the Indian Institute of Space Science and
34 16507–16515.
Technology (IIST), Trivandrum is greatly acknowledged. [35] J. Guo, J. Zhang, F. Jiang, S. Zhao, Q. Su, G. Du, Electrochim. Acta 2015,
35
176, 853–860.
36
[36] J. J. Cheng, Y. Pan, J. A. Pan, H. J. Song, Z. S. Ma, RSC Adv. 2015, 5, 68–74.
37 Conflict of Interest [37] X. Gu, Y. Wang, C. Lai, J. Qiu, S. Li, Y. Hou, W. Martens, N. Mahmood, S.
38 Zhang, Nano Res. 2015, 8, 129–139.
The authors declare no conflict of interest. [38] K. Yang, Q. Gao, Y. Tan, W. Tian, W. Qian, L. Zhu, C. Yang, Chem. Eur. J.
39
2016, 22, 3239–3244.
40
[39] M. Raja, N. Angulakshmi, A. M. Stephan, RSC Adv. 2016, 6, 13772–13779.
41 Keywords: Biomass · Carbon-sulfur composite · Cathode
[40] N. Moreno, A. Caballero, L. Hernn, J. Morales, Carbon 2014, 70, 241–
42 material · Hierarchical porous carbon · Lithium-sulfur batteries 248.
43 [41] H. Yao, G. Zheng, W. Li, M. T. McDowell, Z. Seh, N. Liu, Z. Lu, Y. Cui, Nano
Lett. 2013, 13, 3385–3390.
44
[1] O. Ogoke, G. Wu, X. Wang, A. Casimir, L. Ma, T. Wu, J. Lu, J. Mater. Chem. [42] S. Lu, Y. Chen, J. Zhou, Z. Wang, X. Wu, J. Gu, X. Zhang, A. Pang, Z. Jiao,
45 A 2017, 5, 448–469. L. Jiang, Sci. Rep. 2016, 6, 20445.
46 [2] M. Liu, X. Qin, Y.-B. He, B. Li, F. Kang, J. Mater. Chem. A 2017, 5, 5222– [43] Y. Xia, R. Fang, Z. Xiao, H. Huang, Y. Gan, R. Yan, X. Lu, C. Liang, J. Zhang,
47 5234. X. Tao, W. Zhang, ACS Appl. Mater. Interfaces 2017, 9, 23782–23791.
[3] Q. Pang, X. Liang, C. Y. Kwok, L. F. Nazar, Nat. Energy 2016, 1, 16132. [44] M. Chen, S. Jiang, C. Huang, X. Wang, S. Cai, K. Xiang, Y. Zhang, J. Xue,
48
[4] W. Kang, N. Deng, J. Ju, Q. Li, D. Wu, X. Ma, L. Li, M. Naebe, B. Cheng, ChemSusChem 2017, 10, 1803–1812.
49 Nanoscale 2016, 8, 16541–16588. [45] C. Jin, W. Zhang, Z. Zhuang, J. Wang, H. Huang, Y. Gan, Y. Xia, C. Liang, J.
50 [5] Z. Lin, C. Liang, J. Mater. Chem. A 2015, 3, 936–958. Zhang, X. Tao, J. Mater. Chem. A 2017, 5, 632–640.
51 [6] A. Manthiram, Y. Fu, S.-H. Chung, C. Zu, Y.-S. Su, Chem. Rev. 2014, 114, [46] G. Zhou, E. Paek, G. S. Hwang, A. Manthiram, Nat. Commun. 2015, 6,
52 11751–11787. 7760.
[7] M. Li, B. Yang, Z. Zhang, L. Wang, Y. Zhang, J. Appl. Electrochem. 2013, [47] X. Yuan, B. Liu, H. Hou, K. Zeinu, Y. He, X. Yang, W. Xue, X. He, L. Huang,
53 43, 515–521. X. Zhu, L. Wu, J. Hu, J. Yang, J. Xie, RSC Adv. 2017, 7, 22567–22577.
54 [8] S. Rehman, K. Khan, Y. Zhao, Y. Hou, J. Mater. Chem. A 2017, 5, 3014– [48] Q. Pang, X. Liang, C. Y. Kwok, L. F. Nazar, J. Electrochem. Soc. 2015, 162,
55 3038. A2567-A2576.
56 [9] L. Chen, L. L. Shaw, J. Power Sources 2014, 267, 770–783. [49] Z. Wang, Y. Dong, H. Li, Z. Zhao, H. Bin Wu, C. Hao, S. Liu, J. Qiu, X. W.
[10] R. Chen, T. Zhao, F. Wu, Chem. Commun. 2015, 51, 18–33. Lou, Nat. Commun. 2014, 5, 5002.
57

ChemistrySelect 2017, 2, 10484 – 10495 10494  2017 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Wiley VCH
1732 / 102613
Montag, 13.11.2017
[S. 10494/10495] 1
Full Papers
[50] M. Momčilović, M. Purenović, A. Bojić, A. Zarubica, M. Rand̄elović, [65] R. Demir-Cakan, M. Morcrette, F. Nouar, C. Davoisne, T. Devic, D.
1 Desalination 2011, 276, 53–59. Gonbeau, R. Dominko, C. Serre, G. Frey, J.-M. Tarascon, J. Am. Chem.
2 [51] K. Li, S. Tian, J. Jiang, J. Wang, X. Chen, F. Yan, J. Mater. Chem. A 2016, 4, Soc. 2011, 133, 16154–16160.
3 5223–5234. [66] L. Zhang, L. Ji, P.-A. Glans, Y. Zhang, J. Zhu, J. Guo, Phys. Chem. Chem.
[52] A. Bello, N. Manyala, F. Barzegar, A. A. Khaleed, D. Y. Momodu, J. K. Phys. 2012, 14, 13670–13675.
4
Dangbegnon, RSC Adv. 2016, 6, 1800–1809. [67] G. Li, J. Sun, W. Hou, S. Jiang, Y. Huang, J. Geng, Nat. Commun. 2016, 7,
5 [53] K. Karthikeyan, S. Amaresh, S. N. Lee, X. Sun, V. Aravindan, Y.-G. Lee, Y. S. 10601.
6 Lee, ChemSusChem 2014, 7, 1435–1442. [68] S. Niu, W. Lv, C. Zhang, Y. Shi, J. Zhao, B. Li, Q.-H. Yang, F. Kang, J. Power
7 [54] J. I. Hayashi, A. Kazehaya, K. Muroyama, A. P. Watkinson, Carbon 2000, Sources 2015, 295, 182–189.
38, 1873–1878. [69] Y. Peng, Y. Zhang, J. Huang, Y. Wang, H. Li, B. J. Hwang, J. Zhao, Carbon
8
[55] H. Zhang, Y. Yan, L. Yang, Adsorption 2010, 16, 161–166. 2017, 124, 23–33.
9 [56] A. T. Ward, J. Phys. Chem. A 1968, 72, 4133–4139. [70] W. G. Wang, X. Wang, L. Y. Tian, Y. L. Wang, S. H. Ye, J. Mater. Chem. A
10 [57] M. S. Dresselhaus, G. Dresselhaus, A. Jorio, J. Phys. Chem. C 2007, 111, 2014, 2, 4316–4323.
11 17887–17893. [71] Y. Tan, Z. Zheng, S. Huang, Y. Wang, Z. Cui, J. Liu, X. Guo, J. Mater. Chem.
[58] X. Tao, J. Zhang, Y. Xia, H. Huang, J. Du, H. Xiao, W. Zhang, Y. Gan, J. A 2017, 5, 8360–8366.
12
Mater. Chem. A 2014, 2, 2290–2296. [72] F. Qin, K. Zhang, J. Fang, Y. Lai, Q. Li, Z. Zhang, J. Li, New J. Chem. 2014,
13 [59] Y. Fu, Y.-S. Su, A. Manthiram, J. Electrochem. Soc. 2012, 159, A1420- 38, 4549–4554.
14 A1424. [73] S. K. Meher, P. Justin, G. R. Rao, Electrochim. Acta 2010, 55, 8388–8396.
15 [60] Y. Z. Zhang, S. Liu, G. C. Li, G. R. Li, X. P. Gao, J. Mater. Chem. A 2014, 2, [74] Y. Fu, Y.-S. Su, A. Manthiram, ACS Appl. Mater. Interfaces 2012, 4, 6046–
4652–4659. 6052.
16
[61] H. Ding, J.-S. Wei, H.-M. Xiong, Nanoscale 2014, 6, 13817–13823.
17 [62] J. Xu, D. Su, W. Zhang, W. Bao, G. Wang, J. Mater. Chem. A 2016, 4,
18 17381–17393.
[63] Y. Hao, Z. Shi, J. Wang, Q. Xu, RSC Adv. 2015, 5, 31629–31636. Submitted: September 5, 2017
19
[64] J. Yang, F. Chen, C. Li, T. Bai, B. Long, X. Zhou, J. Mater. Chem. A 2016, 4, Revised: November 1, 2017
20
14324–14333. Accepted: November 3, 2017
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

ChemistrySelect 2017, 2, 10484 – 10495 10495  2017 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Wiley VCH
1732 / 102613
Montag, 13.11.2017
[S. 10495/10495] 1

You might also like