You are on page 1of 32

Accepted Manuscript

A new nano lead-doped mesoporous carbon composite as negative electrode


additives for ultralong-cyclability lead-carbon batteries

Leying Wang, Hao Zhang, Wenfeng Zhang, Hao Guo, Gaoping Cao, Hailei
Zhao, Yusheng Yang

PII: S1385-8947(17)32211-8
DOI: https://doi.org/10.1016/j.cej.2017.12.089
Reference: CEJ 18250

To appear in: Chemical Engineering Journal

Received Date: 7 September 2017


Revised Date: 17 December 2017
Accepted Date: 18 December 2017

Please cite this article as: L. Wang, H. Zhang, W. Zhang, H. Guo, G. Cao, H. Zhao, Y. Yang, A new nano lead-
doped mesoporous carbon composite as negative electrode additives for ultralong-cyclability lead-carbon batteries,
Chemical Engineering Journal (2017), doi: https://doi.org/10.1016/j.cej.2017.12.089

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers
we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting proof before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
A new nano lead-doped mesoporous carbon
composite as negative electrode additives for
ultralong-cyclability lead-carbon batteries

Leying Wanga,b, Hao Zhangc,*, Wenfeng Zhangc, Hao Guod, Gaoping Caoc, Hailei

Zhaob, Yusheng Yangb,c

a
School of Materials Science and Engineering, Jingdezhen Ceramic Institute,

Jingdezhen 333403, China

b
School of Materials Science and Engineering, University of Science and Technology

Beijing, Beijing 100083, China

c
Research Institute of Chemical Defense, Beijing 100191, China

d
China Institute of Atomic Energy, P.O. Box 275(30), Beijing 102431, China

*
Corresponding author:

Hao Zhang, Research institute of Chemical Defense, Beijing 100191, China

Tel.: +86-10-66705840; Fax: +86-10-66748572

E-mail: dr.h.zhang@hotmail.com

ABSTRACT

We propose and realize a new nano lead-doped mesoporous carbon composite as the

negative electrode additives, which realize the abundant nano-lead electrodeposition

into carbon pores and the remarkable suppression of irreversible sulfation, to

effectively prolong the cycle life of lead-carbon batteries. We show that through

NaOH activation and followed air oxidation, porous carbon could obtain more

mesopore volume and appropriate acidic groups, which are two critical parameters for

1
effective nano-lead electrodeposition on the internal surface of them. We for the first

time demonstrate that these mesopore system could be loaded much more deposits in

them, and confine the size of lead deposits to be in nano scale by their local effect,

ensuring a much more remarkable suppression of hydrogen evolution and better

reversibility of Pb/PbSO4. In addition, we show that these nano-lead electrodeposits

could stably contribute remarkable pseudocapacitance. All these merits contribute to

the ultralong cyclability achieved by lead-carbon batteries, which will translate into

promising inexpensive systems that could revolutionize the large-scale energy storage

fields.

Keywords: Mesoporous carbon; Nano-lead electrodeposition; Pseudocapacitance;

Irreversible sulfation; Lead-carbon batteries.

1. Introduction
With the great demand for electrochemical energy storage devices to stabilize

renewable energy supply, carbon materials have been widely applied in energy

storage field owing to the unique structure, high electric conductivity, and chemical

stability [1-5]. Lead-carbon batteries, which utilize carbon materials as the negative

electrode additives and have ultralong cycle life under high-rate

partial-state-of-charge (HRPSoC) conditions, is essential to the grid-scale energy

storage and application of renewable green energy [6-9].

Porous carbon is the most critical material for enhancing the performance of

lead-carbon batteries due to the abundant pore structure, high specific surface area

and good energy storage performance of electric double-layer [10-12]. D. Pavlov et al.

[13-14] claimed that in the HRPSoC conditions, small PbSO4 crystals with high

solubility could be formed to sustain a high concentration of Pb2+ ions in the carbon

2
pores after adding a certain amount of active carbon into the negative plate. P.T.

Moseley [15] indicated that the carbon with a high surface area as the negative

electrode additives could not only segregate the lead sulfate crystals to impede their

growth, but also be beneficial for the store of the sulfuric acid electrolyte in the

negative plate to improve the dissolution of PbSO4 through the recharging. With

respect to the study on the electric double-layer property, Xiang et al. [16]

demonstrated that during the high-rate charging, the activated carbon could act as

capacitive buffer accepting excess charge current, which could further supply

electrons for electrode reaction after the charge is completed. However, these aspects

didn’t clearly elucidate the related mechanisms of porous carbon combined with the

specific microscopic internal structure, which would limit the further development of

lead-carbon batteries.

Recently some researchers reported the lead deposits on the surface of porous

carbon additives in lead-carbon battery anodes, which could inhibit the hydrogen

evolution, increase the direction for current distribution, and thus effectively enhance

the reversible reaction of the Pb/PbSO4 [17-19]. Our past work [20-21] verified that

the acidic groups could serve as the active sites of lead electrodeposition, and the

mesopore-dominated porous carbon materials were beneficial for the lead particles

electrodeposited into the nano-sized carbon pores. However, there are still some lead

particles on the external surface that would grow up to form the irreversible sulfation

with the increase of cycles, which influence to further enhance the performance of

lead-carbon batteries. Therefore, it is necessary to make PbSO4 particles

electrodeposited from the external surface to the internal surface of porous carbon as

far as possible. These lead deposits in carbon pores can effectively inhibit the

irreversible sulfation.

3
Based on the data from literatures and our previous works, it is demonstrated that

two critical parameters are highly desired for porous carbon additives, i.e., the acidic

functional groups and the mesopore volume as the active sites and the space for lead

electrodeposition, respectively. The alkaline activation by KOH is an effective method

to prepare the carbon materials with more micropores than that by NaOH [22-23], but

the stronger ability of activating isn’t beneficial to form more mesopores, so it may be

a right way to modify the original micropores into mesopores in carbon materials

through the alkaline activation by NaOH. In addition, the air oxidation could increase

a certain amount of acidic surface functional groups applied in the surface treatment

of carbon materials [24-25]. Combined with the NaOH activation and air oxidation,

carbon materials may obtain enough acidic functional groups and mesopore volume

for the lead electrodeposition in carbon pores.

Here, we design a new nano lead-doped mesoporous carbon composite as the

negative electrode additives. We use the NaOH activation and followed air oxidation

to modify the surface functional groups and porous structure of mesoporous carbon

materials, and then use the modified carbon materials to prepare the new lead-doped

mesoporous carbon composite materials. The new composite materials were treated

by different electrochemical experiments in a Pb2+ containing H2SO4 electrolyte with

a three-electrode device. Under the addition of the new composite, we study the

morphology evolution of lead electrodeposits on the surface of carbon, and further

elucidate the related mechanisms in lead-carbon battery anodes.

2. Experimental

2.1 Preparation of the new lead-doped mesoporous carbon composite

The new lead-doped carbon mesoporous composite were prepared as follows in

Fig. 1. Firstly, mesopore-dominated porous carbon materials (MC) were prepared by

4
using phenolic resin as a carbon precursor and CaCO3 nanoparticles (~50 nm) as a

template [26]. MC with NaOH at a mass ratio of 1:2 were activated at 680 ℃ for 2 h

in N2 atmosphere, and alkalinized-MC (AMC) were obtained after being washed to

neutral and dried at 120 ℃ for 4 h. Secondly, AMC were treated at 300 ℃ for 1 h in air

atmosphere to prepare Modified-MC (MMC). 1 g MMC were impregnated into 150

mL 0.1 mol L1 Pb(NO3)2 under the vacuum, and the PbSO4 were precipitated on the

surface of carbon with the addition of 150 mL 0.1 mol L1 H2SO4 dropwise, finally to

obtain the new lead-doped modified mesoporous carbon composite (Pb@MMC) for

the subsequent experiments. MC were also used to prepare the lead-doped

mesoporous carbon composite (Pb@MC) for reference [21]. The structure parameters

of all the samples were evaluated by N2-adsorption/desorption isotherms measured at

77K (ASAP 2020, Micromeritics). The surface functional groups of the samples were

tested by X-ray photoelectron spectroscopy (XPS, PHI-5300, PEKIN-EIMER) with

Al Ka radiation.

Fig. 1. Schematic of preparation of the new lead-doped mesoporous carbon

composite.

2.2 Electrochemical experiments

5
The following electrochemical experiments were tested with a three-electrode

device that consisted of the Pb@MMC powders as the working electrode, a

Hg/Hg2SO4 (Sat. K2SO4) electrode as reference electrode, and a platinum plate as

counter electrode. The working electrode was composed of a graphite plate as current

collector and Pb@MMC film (preparation in supplementary materials). One side of

the graphite plate was covered with the film (1 cm × 1 cm, 120±10 or 280±10 μm) by

colloidal graphite, other non-testing surfaces of which were sealed with tape. The

Electrochemical Testing Station (Solartron 1280Z) was used for all the

electrochemical experiments at room temperature. Other related powders were

conducted under the same operation and testing.

Pb@MMC films (120±10 and 280±10 μm) were used for the lead

electrodeposition by charging to 1.135 V (vs. Hg/Hg2SO4) at 200 mA g1, and then

potentiostatic charging for 5 h. The high-rate charge-discharge performance of these

samples were evaluated by simulating HRPSoC conditions of lead-acid batteries

according to the following schedule for 600 cycles: charge at 1000 mA g 1 for 45 s

(upper voltage limit of 1.2 V, vs. Hg/Hg2SO4), rest for 5 s; discharge at 1000 mA g1

for 30 s, rest for 5 s. The electrolyte in the above electrochemical testing was 5 mol

L1 H2SO4 aqueous solution with some lead powders (~2g L1).

After the above different electrochemical experiments, the hydrogen evolution

behavior of Pb@MMC films (120±10μm) were tested by cyclic voltammetry (CV)

procedures from 0.7 to 1.5 V (vs. Hg/Hg2SO4) at 1 mV s1 in 5 mol L1 H2SO4. The

capacitive property of the films with the thickness of 280±10 μm were obtained by

6
CV testing from 0 to 1.2 V (vs Hg/Hg2SO4) at 0.7 mV s1 in 5 mol L1 H2SO4.

Through different electrochemical testing, the morphology evolution of nano-lead

on the external surface of MMC were observed using a Field Emission Scanning

electron microscopy (FESEM, Hitachi S4800), and the elementary composition of the

samples were determined by Energy Dispersive Spectrometer (EDS, EMAS350). The

above-mentioned films (120±10μm) were cut into some flakes (~70 nm) by ultrathin

sectioning, and then used to characterize the morphology evolution of nano-lead on

the internal surface of MMC by High Resolution Transmission Electron Microscopy

(HRTEM, Tecnai F20). The Inductively Coupled Plasma-Atomic Emission Spectrum

(ICP-AES, ULTIMA) was used to analyze the content of lead in the above samples.

2.3 Batteries testing


The 2 V simulated lead-acid batteries were assembled with one negative plate (3.8

cm × 6.4 cm × 0.22 cm), two positive plates (3.8 cm × 6.4 cm × 0.25 cm), and a

absorptive glass-microfiber separator with a thickness of 0.17 cm to underway the cell

testing at room temperature. The positive plates and the raw materials of negative

plates were all produced by Zhejiang Narada Power Co., Ltd.

Before the batteries assembled, the 1 wt.% (relative to the lead active materials)

MC, Pb@MC, MMC and Pb@MMC powders were added into the negative plates

during mixing the lead paste (choosing the additive content in supplementary

materials), respectively. The blank negative plate was prepared without carbon

additives for reference. Firstly, all the negative plates were dried at 65℃ for 24 h, and

then the formation process were composed of a charge at 0.22 C 10 for 10 h, a

discharge at 0.1 C10 for 0.5 h and once more charge at 0.2 C10 for 10 h in 1.04 g cm3

H2SO4 aqueous solution. The first step of the HRPSoC condition was that the cells

7
after a full charge were discharged to 60% SoC at 1 C10 rate, and then the cycle

performance of the cells were tested as follows: charge at 2 C10 rate for 90 s (upper

voltage limit of 2.35 V), rest for 10 s; discharge at 2 C10 rate for 60 s, rest for 10 s.

The voltage of the battery at the end of the discharge pulses was recorded on every

cycle, and the cycle testing was stopped when this voltage fell down to 1.70 V. X-ray

diffractometer (XRD, D8 advance) and FESEM were used to characterize the

elementary composition and the morphology of the negative plates of the cells in the

charging state after 20,000 HRPSoC cycles.

3. Results and discussion

3.1 Characterization of MC, AMC, MMC and Pb@MMC

The porous structure parameters of MC, AMC, MMC and Pb@MMC films were

evaluated by N2-adsorption/desorption in Table 1 and Fig. 2. Firstly, through the

NaOH activation process of the mesopore-dominated porous carbon materials (MC),

the total pore volume of alkalinized-MC (AMC) was increased from 0.39 to 0.91 cm3

g1, the mesopore volume changed from 0.16 to 0.27 cm3 g1, and the SBET increased

from 385 to 1414 m2 g1 in Table 1, which were the results of the NaOH activation to

add carbon pores with a range of 0.8~4 and 50~100 nm in Fig. 2. The followed air

oxidation further increased the mesoscale pores (2~10 nm), and changed the

mesopore volume to 0.52 cm3 g1 in Modified-MC (MMC). After the

chemical-precipitation process, the PbSO4 particles could mainly occupy the

mesoscale pores of MMC with a range of 1~10 and 70~100 nm, which decreased the

mesopore volume of the new lead-doped modified mesoporous carbon composite

(Pb@MMC) from 0.52 to 0.26 cm3 g1, the total pore volume from 1.12 to 0.67 cm3

8
g1, and the SBET from 1443 to 639 m2 g1. However, the mesopore volume were just

decreased from 0.16 to 0.09 cm3 g1, when MC were chemical-precipitated to prepare

the lead-doped mesoporous carbon composite (Pb@MC) [21]. The change of porous

structure in carbon films determined that through the NaOH activation and followed

air oxidation, porous carbon could possess more mesopore volume, and then used to

prepare the new lead-carbon composite with more PbSO4 particles deposited in

mesoscale pores.

Table 1. Porous structure parameters of MC, AMC, MMC and Pb@MMC films.

SBET Average pore size Pore volume Mesopore volume


Samples
(m2 g1) (nm) (cm3 g1) (cm3 g1)
MC 385 3.56 0.39 0.16
AMC 1414 3.25 0.91 0.27
MMC 1443 3.10 1.12 0.52
Pb@MMC 639 2.98 0.67 0.26

Fig. 2. Pore-size distribution of MC, AMC, MMC and Pb@MMC films.

9
Fig. 3. XPS spectra of MC, AMC, MMC and Pb@MMC powders.

Fig. 3 shows the XPS spectra of MC, AMC, MMC and Pb@MMC powders. The

sharp peaks at 280~300 eV and 525~545 eV separately correspond to the binding

energy of C1s and O1s. Through the NaOH activation process of MC, the O1s peak of

AMC was obviously weakened due to the reduction of the acidic surface functional

groups. However, the subsequent air oxidation recovered a little oxygen content to

MMC, and Pb@MMC composite made the O1s peak a little weakened because of

PbSO4 particles covering some oxygen-containing functional groups. In order to

further analysis the change of surface functional groups by modification, the C1s peak

of the samples should be fitted into five peaks [27-28], including Peak 1 to graphitic

carbon (C-C, 284.8 eV), Peak 2, peak 3, peak 4 and peak 5 separately to the

phenolic hydroxyl groups (C-O, 286 eV), carbonyl groups (C=O, 287.1 eV),

carboxyl groups (O-C=O, 288.9 eV) and carbonate groups or absorbed carbon

oxides (OCOO, 291.0 eV). The O1s peak of samples should be also deconvoluted to

five peaks as below [29-30]: peak I is attributed to C=O bonds in carbonyl and

carboxylic acids (531.5 eV), peak II to C=O bonds in ester and anhydride groups,

10
OH bonds in alcohols and C-O bonds in ethers (532.5 eV), peak III to C-O bonds

in ester and anhydride groups (533.3 eV), peak IV to C-O bonds in carboxyl groups

(534.8 eV), and peak V to H2O or O2 adsorbed on the surface of carbon (536.2 eV).

Fig. 4. Peak deconvolutions of the XPS C1s spectra. (a) MC, (b) AMC, (c) MMC, (d)

Pb@MMC.

Table 2. Relative contents of functional groups obtained from peak deconvolutions of

the C1s and O1s spectra.

C1s / at.% O1s / at.%


Samples O/C Peak Peak Peak Peak Peak
Peak Peak Peak Peak Peak
1 2 3 4 5
Ⅰ Ⅱ Ⅲ Ⅳ Ⅴ
C-C C-O C=O COO OCOO

MC 0.220 62.17 14.18 8.77 7.58 7.30 18.00 25.45 18.59 35.43 2.52

AMC 0.094 65.66 17.16 5.96 7.23 3.97 16.59 18.63 29.45 35.25 0.07

MMC 0.098 66.07 13.58 9.34 5.85 5.14 19.43 19.54 26.88 29.64 4.51

11
Pb@MMC 0.093 69.04 9.94 6.35 6.39 8.27 17.56 37.48 13.02 31.21 0.72

The deconvolution results of C1s peak could be seen in Fig. 4 and Table 2, and the

half width data of each peak of the samples in Table S1. As mentioned above, when

MC were alkalinized to obtain AMC, some acidic functional groups were reduced to

make the O/C ratio decreased from 0.220 to 0.094, the content of C=O, O-C=O

and COOO groups down to 5.96, 7.23 and 3.97 at.%, respectively, which were in

good agreement with the O1s deconvolution results (in Fig. S1 and Table 2) that

AMC had the lower peak I (16.59 at.%) and peak Ⅱ (18.63 at.%). Through the

following air oxidation , the O/C ratio of MMC got back to 0.098, the content of C=

O and COOO groups were separately increased to 9.34 and 5.14 at.%, which were

well supported by the O1s deconvolution results that MMC had the larger peak I

(19.43 at.%) than that (16.59 at.%) of AMC. After MMC deposited with PbSO4 to

obtain Pb@MMC composite, the content of C=O was reduced to 6.35 at.%, and the

peak I by the O1s results was also decreased to 17.56 at.%, which indicated that

PbSO4 particles covered some C=O groups of MMC.

The content of lead in Pb@MMC film was increased to 7.3 wt.%, higher than 5.7

wt.% in Pb@MC film measured by ICP. Combined with the above characterization

by N2-adsorption/desorption, XPS and ICP, it is fully demonstrated that through the

NaOH activation and followed air oxidation of MC, MMC could obtain more

mesopore volume and appropriate acidic groups. The change of microscopic

properties in porous carbon could deposit more PbSO4 particles in mesoscale pores,

and thus prepare the new lead-doped carbon composite with higher lead content for

12
the following nano-lead electrodeposition.

3.2 Morphology evolution of nano-lead electrodeposits in Pb@MMC

The FESEM images in Fig. 5 present the morphology evolution of lead particles

on the external surface of carbon in Pb@MMC before and after different

electrochemical measures. Through the lead electrodeposition, the PbSO4 particles

with white irregular polyhedron (0.5~2 μm) on the surface of carbon in Fig. 5(a) were

reduced to the Pb particles with petal shaped (~500 nm) in Fig. 5(b), which also

corresponded to the results of EDS testing in Fig. S2. After 600 high-rate cycles’

testing, there were not only some lead particles (1~2 μm) on the external surface of

MMC, which was smaller than that (2~5 μm) on the external surface of MC [21], but

also much gray area that electrodeposited more, smaller lead particles in Fig. 5(c). As

seen from Fig. 5(d,e), many lead particles (200~500 and 50 nm) could be observed on

the surface of carbon from the gray area in magnification. The phenomenon illustrated

that through the modification of porous carbon, more lead particles could be

electrodeposited on the surface of carbon and tend to be embedded into carbon pores

with the increase of cycles, which could inhibit the accumulation of the lead particles

on the external surface of porous carbon, and thus effectively enhance the high-rate

charge-discharge performance.

13
Fig. 5. SEM images of Pb@MMC, (a) Composites, (b) after Pb electrodeposition,

(c~e) after 600 cycles. Inset of (a) and (b) are SEM images of PbSO4 and Pb particles

in scale of 1μm and 500 nm, respectively.

Fig. 6. (a,b) TEM images of Pb@MMC after 600 cycles, (c) pore-size distribution of

MMC film and Pb@MMC films before and after electrochemical measures.

Table 3. Porous structure properties of MMC film and Pb@MMC films before and

after electrochemical measures.


Average
SBET Pore volume Mesopore volume
Samples pore size
(m2 g1) (cm3 g1) (cm3 g1)
(nm)

14
MMC 1443 3.10 1.12 0.52

Pb@MMC 639 2.98 0.67 0.26


Pb@MMC
764 3.12 0.57 0.13
after Pb electrodeposition
Pb@MMC after 600 cycles 594 4.77 0.49 0.23

Through ultrathin sectioning, the internal morphology evolution of pores in

Pb@MMC could be well characterized in Fig. 6(a,b). After 600 high-rate cycles, the

nano-lead particles still existed in the pores with the size of 20~50 nm in Fig. 6(a),

and 5~10 nm in Fig. 6(b). The phenomenon determined that through the long cycling

test, more nano-lead particles on the external surface of MMC were electrodeposited

into carbon pores, and not growing up to form the irreversible sulfation, which could

further increase the surface area and enhance the reactivity of Pb active materials than

Pb@MC [21]. N2-adsorption/desorption were used to further verify the morphology

evolution of lead particles on the surface of MMC in Table 3 and Fig. 6(c). After the

Pb electrodeposition on the surface of MMC, the pore peaks at 1~7 nm and 30~100

nm were weakened in Fig. 6(c), which reduced the total pore volume of Pb@MMC to

0.57 cm3 g1, and the mesopore volume to 0.13 cm3 g1 in Table 3. However, the

nano-lead electrodeposits on the external surface of MMC existed in petal shaped in

Fig. 6(b), so the SBET of Pb@MMC was increased from 639 to 764 m2 g1. Through

600 high-rate cycles, the pore peaks at 1~4 nm were further weakened due to the lead

electrodeposition in microscale and mesoscale pores, which resulted in that the total

pore volume of Pb@MMC fell down to 0.49 cm3 g1, and the SBET to 594 m2 g1 ,but

increased the mesopore volume to 0.23 cm3 g1. The content of lead in Pb@MMC

films before and after Pb electrodeposition, and after 600 cycles’ testing were about

15
7.3, 11.8 and 12.8 wt.% measured by ICP, respectively, compared to 5.7, 8.3 and 8.5

wt.% in Pb@MC [21], which reflected that there were more lead particles

electrodeposited into the pores of MMC than MC through electrochemical

experiments in a Pb2+ containing H2SO4 electrolyte.

The above results determined that through the electrochemical experiments, MMC

with more mesopore volume and appropriate acidic groups, are beneficial for more

nano-lead particles electrodeposited on the external and internal surface of porous

carbon. This is because that through NaOH activation and followed air oxidation,

porous carbon could obtain the appropriate acidic groups as the active sites of lead

electrodeposition, and more mesopore volume to provide enough space for the lead

electrodeposition and confine their nano-size, which finally increase the content of

lead electrodeposits on the internal surface of porous carbon,and thereby reduce the

size of lead particles electrodeposited on the external surface under the long high-rate

cycle testing.

3.3 Electrochemical testing and mechanism research

Different electrochemical experiments were conducted on Pb@MMC films as the

working electrode under a three-electrode system in Fig. 7. These electrochemical

data were compared with that of Pb@MC to study the effect of microscopic

modification on the related mechanisms. The hydrogen evolution behavior of all

samples (120±10μm) after different electrochemical measures were tested in Fig. S3

and Fig. 7(a). The specific current (ih) of CV curves at 1.5 V were used to compare

the hydrogen evolution behavior of different samples. After Pb electrodeposition, the

16
lead deposits could cover the acidic groups that promote hydrogen evolution [20], and

the nano-lead electrodeposits from the external surface to carbon pores could get

higher hydrogen evolution overvoltage [21]. As shown in Fig. S3, through the lead

doping and followed lead electrodeposition, both the ih of MC and MMC at 1.5 V

were obviously decreased, which sufficiently indicated that the nano-lead

electrodeposits could inhibit the hydrogen evolution induced by carbon. As MC were

modified to obtain MMC with the higher specific surface area in Table 1, Pb@MMC

after Pb electrodeposition with the ih of 0.46 A g1 at 1.5 V had the more serious

hydrogen evolution behavior than Pb@MC after Pb electrodeposition with the ih of

0.10 A g in Fig. 7(a). The more cycle testing would result in the more serious
1

hydrogen evolution behavior, so the ih of Pb@MC at 1.5 V after 600 cycles were

increased from 0.10 to 0.60 A g1. However, through the same 600 cycles, the ih of

Pb@MMC at 1.5 V was only changed from 0.46 to 0.59 A g1, which showed that

the hydrogen evolution behavior were more effectively inhibited due to the more

nano-lead electrodeposits in carbon pores than that of Pb@MC.

The capacitive property of Pb@MC and Pb@MMC films (280±10μm) after

different electrochemical measures were compared in Fig. 7(b). The specific

capacitance of all samples could be calculated from the CV curves according to the

following equation.

Cm = SCV/2mυ△V [1]

where Cm is the specific capacitance (F g1), SCV is the area of CV curve, m is the

mass of active materials in the working electrode (g), υ is the scan rate of CV curve

17
(V s1) and △V is the width of the potential window (V). Owing to the higher

specific surface area, the specific capacitance of Pb@MMC after lead

electrodeposition was 242 F g1, higher than 163 F g1 of Pb@MC after Pb

electrodeposition. After 600 cycles, the capacitive property of Pb@MC was decreased

from 163 F g1 to 149 F g1 because of the growth of some PbSO4 particles on the

external surface of porous carbon [21]. However, the capacitive property of

Pb@MMC after 600 cycles was only reduced from 242 F g1 to 239 F g1, which

nearly maintained the excellent capacitive property. It is worth noticing that there was

a mild reduction peak (1.05 ~ 1.15 V) that presented the reaction of Pb2+ reduced to

Pb, and an obvious oxidation peak (1.1 ~ 1.0 V) with Pb oxidized to form PbSO4 in

the CV curves of Pb@MC [31-32]. But there were few obvious redox peaks in the CV

curves of Pb@MMC, which indicated that through electrochemical measures,

Pb@MMC could provide more pseudocapacitance contribution from the more

nano-lead electrodeposits on the surface of porous carbon.

18
Fig. 7. CV curves of samples after electrochemical measures, (a) 0.7~1.5 V, 1 mV

s1, (b) 0~1.2 V, 0.7 mV s1; (c) high-rate cycle performance of samples.

Fig. 7(c) shows the high-rate charge-discharge performance of MC and MMC,

Pb@MC and Pb@MMC films (120±10μm) after Pb electrodeposition under a

three-electrode device. The voltage of every cycle at the end of the discharge pulses

was recorded to characterize the high-rate cycle performance. With the increase of

cycles, the higher discharge voltage presented the charge acceptance of the working

electrode gradually weakened. Through 600 high-rate cycles, Pb@MC and

Pb@MMC had the lower voltage than MC and MMC, respectively. Pb@MMC with

the lower voltage presented the stronger charge acceptance than Pb@MC, which was

due to the more nano-lead electrodeposits into carbon pores that enhanced the

reversibility of Pb/PbSO4. Though the more cycle testing would further promote the

19
hydrogen evolution and decrease the capacitive property of the samples, the portion of

nano-lead electrodeposits on the internal surface of porous carbon could effectively

slow down the degradation. So Pb@MMC with more nano-lead electrodeposits could

obtain the better high-rate cycle performance than Pb@MC.

Fig. 8. Mechanism diagram of the new composite as the negative electrode additives

in lead-carbon battery anodes under HRPSoC conditions.

Through the above experimental results, we characterized the morphology

evolution of lead electrodeposition on the surface of MMC that obtained by the

modification of MC, and demonstrated the related mechanisms through the

electrochemical experiments under the three-electrode device. During the high-rate

cycling test, MMC with more mesopore volume and appropriate acidic groups are

beneficial for getting more nano-lead electrodeposits on the external and internal

surface of carbon than MC. This phenomenon indicates that more nano-lead particles

could independently and effectively proceed the higher-rate charge-discharge testing,

confine their size under the local effect from nano-sized pores, and thereby decrease

20
the size of the lead electrodeposits on the external surface of porous carbon. These

nano-lead electrodeposits could not only effectively inhibit the hydrogen evolution

induced by carbon, but also stably contribute superior pseudocapacitance, and

observably improve the reversibility of Pb/PbSO4 to enhance the charge acceptance of

the electrode. The morphology evolution of lead electrodeposits on the surface of

carbon and the related mechanisms could reasonably apply to the lead-carbon

batteries to explain how the addition of Pb@MMC further restrain the irreversible

sulfation in Fig. 8, and thus improve the following HRPSoC cycle testing.

3.4 HRPSoC cycle testing

Fig. 9 shows the HRPSoC cycle performance of the 2 V simulated lead-carbon

batteries with different samples as negative electrode additives. The blank battery

without carbon additives had completed 11,521 HRPSoC cycles, when its discharge

ending voltage reached 1.7 V. After adding 1wt.% MC, Pb@MC, MMC or Pb@MMC

powders into the negative plates, the HRPSoC cycle life were separately prolonged to

33,770, 44,757, 39,804 and 56547 cycles. The cell added by Pb@MMC had the most

excellent HRPSoC cycle performance, which coincided with the above results tested

under the three-electrode device.

21
Fig. 9. HRPSoC cycle performance of the 2V simulated lead-carbon batteries.

The irreversible sulfation refer to the accumulation of PbSO4 particles that

couldn’t be reduced to Pb through the recharging process, which seriously degrade the

HRPSoC cycle performance of the lead-acid batteries [33-35]. In order to further

confirm the effect of the new composite as the negative electrode additives on

restraining the irreversible sulfation, the morphology and elementary composition of

the negative plates in the above cells were characterized in the recharging state after

20,000 HRPSoC cycles by SEM and XRD. As shown in Fig. S4(a,b), the cell with

1wt.% MMC had more irregular blocky-shaped PbSO4 particles and less spongy lead

bulk than that of the cell with 1wt.% Pb@MMC. Some lead particles (300~400 nm)

could be observed on the surface of carbon in the cell with MMC in magnification in

Fig. S4(c). However, by adding 1wt.% Pb@MMC into the negative plate, more

spongy lead bulk (~200 nm) could be electrodeposited in Fig. S4(d). Through these

SEM images of comparison, the negative plate with 1wt.% Pb@MMC still had more

and smaller lead particles through the recharging after 20,000 cycles, which

22
represented the stronger reversibility Pb/PbSO4 , and thus improved the charge

acceptance of the negative plate to continue the longer HRPSoC cycle life for the cell.

The following XRD testing further supported the SEM results in Fig. S5. After 20,000

HRPSoC cycles, the XRD pattern of the negative plate by adding Pb@MMC

presented the strongest diffraction peak of Pb (JCPDS 04-0686) and weakest PbSO4

phase (JCPDS 36-1461) than that by adding other additives. Through Rietveld

refinement in Fig. S6 and Table S2, the content of Pb phase in the negative plate by

adding Pb@MMC was 22.66 wt.%, higher than that in other negative plates, and the

content of PbSO4 phase was 77.34 wt.%, lower than others.

The SEM and XRD results combined with the above analyses sufficiently indicate

that through much high-rate charge-discharge testing, this new composite could

realize more nano-lead electrodeposits on the surface of carbon, and maintain their

nano-size in the negative plates just like that under the three-electrode device. These

nano-lead particles could improve the reversibility of Pb/PbSO4 to effectively restrain

the irreversible sulfation, and thus contribute to the ultralong cyclability for the

lead-carbon batteries.

4. Conclusions

In this paper, MC were modified to obtain more mesopore volume and appropriate

acidic groups through NaOH activation and followed air oxidation, and then used to

prepare a new nano lead-doped mesoporous carbon composite (Pb@MMC) for

studying the morphology evolution of lead electrodeposition and the related

mechanisms in lead-carbon battery anodes. The results show that this new composite

23
could effectively enhance the nano-lead electrodeposition on the external and internal

surface of porous carbon under the long high-rate charge-discharge testing. These

nano-lead electrodeposits could not only effectively suppress the hydrogen evolution

induced by carbon, but also stably contribute superior pseudocapacitance. In addition,

the enhanced reversibility of Pb/PbSO4 could observably improve the charge

acceptance of the negative plate. By adding 1wt.% Pb@MMC into the negative plates,

more nano-lead particles could be electrodeposited into carbon pores to remarkably

restrain the irreversible sulfation, and thus bring the excellent HRPSoC cycle

performance of lead-acid batteries.

Acknowledgements

The present study was financial supported by the National Key Research and

Development Program of China (2016YFB0901503) and Beijing Nova Program

(Z131109000413060).

References

[1] M.M. Titirici, R.J. White, N. Brun, V.L. Budarin, D.S. Su, F. del Monte, J.H. Clark,

M.J. MacLachlan, Sustainable carbon materials, Chem. Soc. Rev. 44 (1) (2015)

250–290.

[2] J. Yao, T. Mei, Z.Q. Cui, Z.H. Yu, K. Xu, X.B. Wang, Hollow carbon spheres with

TiO2 encapsulated sulfur and polysulfides for long-cycle lithium-sulfur batteries,

Chem. Eng. J. 330 (2017) 644–650.

[3] Z.Q. Ye, F.J. Wang, C. Jia, K.G. Mu, M. Yu, Y.Y. Lv, Z.Q. Shao, Nitrogen and

oxygen-codoped carbon nanospheres for excellent specific capacitance and cyclic

24
stability supercapacitor electrodes, Chem. Eng. J. 330 (2017) 1166–1173.

[4] B.B. Wang, G. Wang, X.M. Cheng, H. Wang, Synthesis and electrochemical

investigation of core-shell ultrathin NiO nanosheets grown on hollow carbon

microspheres composite for high performance lithium and sodium ion batteries,

Chem. Eng. J. 306 (2016) 1193–1202.

[5] L. Wen, F. Li, H.M. Cheng, Carbon Nanotubes and Graphene for Flexible

Electrochemical Energy Storage: from Materials to Devices, Adv. Mater. 28 (22)

(2016) 4306–4337.

[6] H. Ibrahim, A. Ilinca, J. Perron, Energy storage systems—Characteristics and

comparisons, Renew. Sust. Energ. Rev. 12 (2008) 1221–1250

[7] M. Shiomi, T. Funato, K. Nakamura, K. Takahashi, M. Tsubota, Effects of carbon

in negative plates on cycle-life performance of valve-regulated lead/acid batteries,

J. Power Sources 64 (1997) 147–152.

[8] T. Ohmae, T. Hayashi, N. Inoue, Development of 36-V valve-regulated lead-acid

battery, J. Power Sources 116 (1–2) (2003) 105–109.

[9] D. Pavlov, P. Nikolov, T. Rogachev, Influence of carbons on the structure of the

negative active material of lead-acid batteries and on battery performance, J.

Power Sources 196 (11) (2011) 5155–5167.

[10] P.T. Moseley, R.F. Nelson, A.F. Hollenkamp, The role of carbon in

valve-regulated lead–acid battery technology, J. Power Sources 157 (1) (2006)

3–10.

[11] D. Pavlov, P. Nikolov, Capacitive carbon and electrochemical lead electrode

25
systems at the negative plates of lead–acid batteries and elementary processes on

cycling, J. Power Sources 242 (2013) 380–399.

[12] P.T. Moseley, D.A.J. Rand, K. Peters, Enhancing the performance of lead–acid

batteries with carbon – In pursuit of an understanding, J. Power Sources 295

(2015) 268–274.

[13] D. Pavlov, T. Rogachev, P. Nikolov, G. Petkova, Mechanism of action of

electrochemically active carbons on the processes that take place at the negative

plates of lead-acid batteries, J. Power Sources 191 (1) (2009) 58–75.

[14] D. Pavlov, P. Nikolov, T. Rogachev, Influence of expander components on the

processes at the negative plates of lead-acid cells on high-rate

partial-state-of-charge cycling. Part II. Effect of carbon additives on the

processes of charge and discharge of negative plates, J. Power Sources 195 (14)

(2010) 4444–4457.

[15] P.T. Moseley, Consequences of including carbon in the negative plates of

Valve-regulated Lead–Acid batteries exposed to high-rate partial-state-of-charge

operation, J. Power Sources 191 (1) (2009) 134-138.

[16] J. Xiang, P. Ding, H. Zhang, X. Wu, J. Chen, Y. Yang, Beneficial effects of

activated carbon additives on the performance of negative lead-acid battery

electrode for high-rate partial-state-of-charge operation, J. Power Sources 241

(2013) 150–158.

[17] B. Hong, L. Jiang, H. Xue, F. Liu, M. Jia, J. Li, Y. Liu, Characterization of

nano-lead-doped active carbon and its application in lead-acid battery, J. Power

26
Sources 270 (2014) 332–341.

[18] P. Tong, R. Zhao, R. Zhang, F. Yi, G. Shi, A. Li, H. Chen, Characterization of

lead (Ⅱ)-containing activated carbon and its excellent performance of extending

lead-acid battery cycle life for high-rate partial-state-of-charge operation, J.

Power Sources 286 (2015) 91–102.

[19] W. Zhang, H. Lin, H. Lu, D. Liu, J. Yin, Z. Lin, On the electrochemical origin of

the enhanced charge acceptance of the lead–carbon electrode, J. Mater. Chem. A

3 (8) (2015) 4399–4404.

[20] L. Wang, H. Zhang, G. Cao, W. Zhang, H. Zhao, Y. Yang, Effect of activated

carbon surface functional groups on nano-lead electrodeposition and hydrogen

evolution and its applications in lead-carbon batteries, Electrochim. Acta 186

(2015) 654–663.

[21] L. Wang, W. Zhang, L. Gu, Y. Gong, G. Cao, H. Zhao, Y. Yang, H. Zhang,

Tracking the morphology evolution of nano-lead electrodeposits on the internal

surface of porous carbon and its influence on lead-carbon batteries, Electrochim.

Acta 222 (2016) 376–384.

[22] V. Fierro, V. Torné-Fernández, A. Celzard, Methodical study of the chemical

activation of Kraft lignin with KOH and NaOH, Micropor. and Mesopor. Mat.

101 (2007) 419–431.

[23] E. Mora, C. Blanco, J.A. Pajares, R. Santamaria, R. Menendez, Chemical

activation of carbon mesophase pitches, J. Colloid Interf. Sci. 298 (2006)

341–347.

27
[24] C.A. Toles, W.E. Marshall, M.M. Johns, Surface functional groups on

acidactivated nutshell carbons, Carbon 37 (1999) 1207–1214.

[25] L. Geng, S. Wu, Y. Zou, M. Jia, W. Zhang, W. Yan, G. Liu, Correlation between

the microstructures of graphite oxides and their catalytic behaviors in air

oxidation of benzyl alcohol, J. Colloid Interf. Sci. 421 (2014) 71–77.

[26] C. Zhao, W. Wang, Z. Yu, H. Zhang, A. Wang, Y. Yang, Nano-CaCO3 as

Template for Preparation of Disordered Large Mesoporous Carbon with

Hierarchical Porosities, J. Mater. Chem. 20 (2010) 976–980.

[27] A. Rey, J.A. Zazo, J.A. Casas, A. Bahamonde, J.J. Rodriguez, Influence of the

structural and surface characteristics of activated carbon on the catalytic

decomposition of hydrogen peroxide, Appl. Catal. A: Gen. 402 (2011) 146.

[28] L. Meng, S. Park, Influence of carboxyl group formation on ammonia adsorption

of NiO-templated nanoporous carbon surfaces, Mater. Chem. and Phys. 137

(2012) 85.

[29] L. Geng, S. Wu, Y. Zou, M. Jia, W. Zhang, W. Yan, G. Liu, Correlation between

the microstructures of graphite oxides and their catalytic behaviors in air oxidation

of benzyl alcohol, J. Colloid Interf. Sci. 421 (2014) 71.

[30] R. Zhong, Y. Qin, D. Niu, J. Tian, X. Zhang, X. Zhou, S. Sun, W. Yuan, Effect

of carbon nanofiber surface functional groups on oxygen reduction in alkaline

solution, J. Power Sources 225 (2013) 192.

[31] A. Jaiswal, S.C. Chalasani, The role of carbon in the negative plate of the

lead–acid battery, J. Energ. Storage 1 (2015) 15–21.

28
[32] P. Perret, Z. Khani, T. Brousse, D. Bélanger, D. Guay, Carbon/PbO 2 Asymmetric

Electrochemical Capacitor Based On Methanesulfonic Acid Electrolyte,

Electrochim. Acta 56 (2011) 8122–8128.

[33] K. McGregor, Active-material additives for high-rate lead-acid batteries have

there been any positive advances, J. Power Sources 59 (1–2) (1996) 31–43.

[34] A. Kozawa, H. Oho, M. Sano, D. Brodd, R. Brodd, Beneficial effect of

carbon–PVA colloid additives for lead–acid batteries, J. Power Sources 80 (1–2)

(1999) 12–16.

[35] L.A. Yolshina, V.A. Yolshina, A.N. Yolshin, S.V. Plaksin, Novel lead-graphene

and lead-graphite metallic composite materials for possible applications as

positive electrode grid in lead-acid battery, J. Power Sources 278 (2015) 87–97.

29
Highlights

 A new lead-doped mesoporous carbon composite is designed as the negative

additives.

 The composite effectively enhance the nano-lead electrodeposition into carbon

pores.

 The nano-lead deposits stably contribute superior pseudocapacitance.

 The composite significantly prolong the cycle life of lead-carbon battery.

30
Graphical abstract

31

You might also like