You are on page 1of 40

Subscriber access provided by UNIV OF REGINA

Article
Cadmium Removal from Aqueous Solution by
Deionization Supercapacitor with Birnessite Electrode
Qichuan Peng, Lihu Liu, Yao Luo, Yashan Zhang, Wen-Feng Tan, Fan Liu, Steven L. Suib, and Guohong Qiu
ACS Appl. Mater. Interfaces, Just Accepted Manuscript DOI: 10.1021/acsami.6b12224 Publication Date (Web): 30 Nov 2016
Downloaded from http://pubs.acs.org on December 6, 2016

Just Accepted

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

ACS Applied Materials & Interfaces is published by the American Chemical Society.
1155 Sixteenth Street N.W., Washington, DC 20036
Published by American Chemical Society. Copyright American Chemical Society.
However, no copyright claim is made to original U.S. Government works, or works
produced by employees of any Commonwealth realm Crown government in the course
of their duties.
Page 1 of 39 ACS Applied Materials & Interfaces

1
2
3 Cadmium Removal from Aqueous Solution by Deionization Supercapacitor with
4
5 Birnessite Electrode
6
7
8 Qichuan Peng, Lihu Liu, Yao Luo, Yashan Zhang, Wenfeng Tan, Fan Liu, Steven L. Suib,
9
10
Guohong Qiu*,
11
12

13 Key Laboratory of Arable Land Conservation (Middle and Lower Reaches of Yangtse River), Ministry
14
15
16 of Agriculture, College of Resources and Environment, Huazhong Agricultural University, Wuhan
17
18 430070, China
19
20

21 Department of Chemistry, University of Connecticut, 55 North Eagleville Road, Storrs, Connecticut,
22
23
24 06269-3060, USA
25
26
27
28
29 ABSTRACT: Birnessite is widely used as an excellent adsorbent for heavy metal ions and as active
30
31
electrode materials for supercapacitors. The occurrence of redox reactions of manganese oxides is
32
33
34 usually accompanied by the intercalation-deintercalation of cations during the charge-discharge
35
36
37 processes of supercapacitors. In this study, based on the charge-discharge principle of
38
39 supercapacitor and excellent adsorption properties of birnessite, birnessite-based electrode was used
40
41
42 to remove Cd2+ from aqueous solutions. The Cd2+ removal mechanism and the influences of
43
44
45 birnessite loading and pH on the removal performance were investigated. The results showed that
46
47 Cd2+ was adsorbed on the surfaces and interlayers of birnessite, and the maximum electrosorption
48
49
50 capacity of birnessite for Cd2+ was about 900.7 mg g-1 (8.01 mmol g-1), which was significantly
51
52
higher than the adsorption isotherm capacity of birnessite (125.8 mg g-1). The electrosorption
53
54
55 specific capacity of birnessite for Cd2+ increased with an increase in initial Cd2+ concentration, and
56
57
58 decreased with an increase in the loading of active birnessite. In the pH range of 3.06.0, the
59 1
60
ACS Paragon Plus Environment
ACS Applied Materials & Interfaces Page 2 of 39

1
2
3
electrosorption capacity increased at first with an increase in pH and then reached equilibrium
4
5 above pH 4.0. This work provides a new method for the highly efficient adsorption of Cd2+ from
6
7
8 polluted wastewater.
9
10
Keywords: Electrochemical removal; Deionization supercapacitor; Birnessite; Galvanostatic
11
12
13 charge-discharge; Cd2+
14
15
16
17
18 1. INTRODUCTION
19
20
21 In recent years, the shortage of clean drinking water sources has become a global issue mostly
22
23
24 due to water pollution caused by intensive human activities. Heavy metals ions, as major
25
26 components of water pollutants, accumulated in vivo and difficult to degrade, cause increasing risks
27
28
29 to the ecological environment and human health.1,2 In particular, cadmium, released from industrial
30
31
processes such as battery production, dyes and metallurgy,3 if enriched in human body, can cause
32
33
34 skeletal deformities, chronic pulmonary problems, renal degradation, muscular cramps, diarrhea and
35
36
37 erythrocyte destruction.4,5 The guideline value for the discharge of cadmium in water recommended
38
39 by World Health Organization (WHO) is 0.005 mg L-1.5 Hence, it is highly necessary to strictly
40
41
42 control the concentration of cadmium in discharged wastewater.
43
44
45 Currently, the common methods for Cd2+ removal from water bodies include chemical
46
47 precipitation, membrane filtration, ion-exchange, and adsorption etc.5,6 For instance, Ca(OH)2,
48
49
50 which is used as a precipitant, can remove 99.67% of Cd2+ with an initial Cd2+ concentration of 150
51
52
mg L-1.7 Under high pressure conditions, polyamide membrane can reduce the concentration of
53
54
55 Cd2+ from 200 mg L-1 to 2 mg L-1 by reverse osmosis technology.8 Synthetic zeolites can remove
56
57
58 100% of Cd2+ with an initial Cd2+ concentration of 100 mg L-1, but appropriate pretreatment
59 2
60
ACS Paragon Plus Environment
Page 3 of 39 ACS Applied Materials & Interfaces

1
2
3
systems such as the removal of suspended solids from the wastewater are required before ion
4
5 exchange.9 The adsorption capacity of activated carbon cloth treated by HNO3 for Cd2+ is about 146
6
7
8 mg g-1 (1.30 mmol g-1).10 However, the application of most of these methods is limited due to their
9
10
high cost, complex operation or secondary pollution.
11
12
13 As a salt ion removal technology, capacitive deionization has attracted enormous attention in
14
15
16 recent years due to the advantages of low energy consumption, simple operation and environmental
17
18 friendliness.1114 The extensively used electrode for deionization capacitor are carbon materials
19
20
21 including porous carbon,13 grapheme,15 carbon aerogels,16 and nitrogen-doped carbon materials.17,18
22
23
24 Under a direct voltage, an electric double layer can be formed on the surface of the electrodes to
25
26 store oppositely charged ions.1420 This technique is also used to remove heavy metal ions. For
27
28
29 example, the removal efficiencies could reach 81%, 78%, and 42% for 0.05 mmol L-1 Pb2+, Cr3+,
30
31
and Cd2+, respectively, using active carbon electrode under a cell voltage of 1.2 V.21 MnO2/carbon
32
33
34 fiber composite was prepared as an electrosorptive electrode with Cu2+ eletrodeposition capacity of
35
36
37 172.88 mg g-1 (2.72 mmol g-1).22 The removal efficiencies reached 67% and 58% for As(III) and
38
39 As(V), respectively, using flexible carbon fabric supported magnetite multiwalled carbon nanotube
40
41
42 nanocomposite by cyclic voltammetry.23 The principle of capacitive deionization is electrostatic
43
44
45 adsorption, which is similar to that of electric double-layer capacitors. The energy storage of
46
47 pseudocapacitors relies on redox reactions accompanied by the intercalation-deintercalation of
48
49
50 alkali metal ions and H+ on electrode materials.24 As for electrochemical supercapacitors, besides
51
52
double-layer capacitance, pseudocapacitance also contributes much to the specific capacitance
53
54
55 during the charge-discharge processes.24 These processes can also facilitate the adsorption of other
56
57
58
59 3
60
ACS Paragon Plus Environment
ACS Applied Materials & Interfaces Page 4 of 39

1
2
3
metal ions, such as heavy metal ions, on electrode surface, which has received rather limited
4
5 research attention.
6
7
8 In the field of electrochemical energy storage, some transition metal compounds, such as Fe3O4,
9
10
MnO2, V2O5, SnO2, and MoS2, have been investigated as common pseudocapacitive electrode
11
12
13 materials.2527 Manganese oxide has attracted enormous attention due to its characteristics of high
14
15
16 specific capacitance, abundant resources, low cost and environmental friendliness. Birnessite, the
17
18 most common manganese oxide mineral, is composed of layers of edge-sharing MnO6 octahedra
19
20
21 separated by layers of water molecules and alkali metal ions.28 A previous research has indicated
22
23
24 that Cd2+ adsorbed on birnessite occupies sites the above and below octahedral vacancies at the
25
26 birnessite interlayers, and the adsorption capacity of Cd2+ is 148 mg g-1.29 Manganese dioxide
27
28
29 nanosheets were prepared from tetramethylammonium hydroxide, H2O2, and MnCl24H2O, and
30
31
have superior adsorption capacity for Cd2+ (about 348.1 mg g-1).30
32
33
34 In this study, based on its characteristics of excellent capacitance and adsorption capacity for
35
36
37 Cd2+, birnessite was used as a pseudocapacitive electrode materials to remove Cd2+ from aqueous
38
39 solutions by the galvanostatic charge-discharge method. The electrochemical performance of
40
41
42 birnessite was evaluated by cyclic voltammetry (CV). The crystal structure, morphology and
43
44
45 relative contents of Mn with different valence states in birnessite after charge-discharge tests were
46
47 characterized by X-ray diffraction (XRD), field emission scanning electron microscope (FESEM)
48
49
50 and X-ray absorption spectroscopy (XPS), respectively. The mechanism of electrochemical sorption
51
52
of Cd2+ was further explored and discussed. The influence of initial pH of electrolyte and the
53
54
55 loading of electrode materials on the electrosorption capacity for Cd2+ was clarified. This work
56
57
58 provides a new method for the efficient adsorption of Cd2+ from wastewater.
59 4
60
ACS Paragon Plus Environment
Page 5 of 39 ACS Applied Materials & Interfaces

1
2
3
4
5 2. MATERIALS AND METHODS
6
7
8 2.1. Preparation of Birnessite. Birnessite was prepared from the reduction of boiling solution of
9
10
potassium permanganate by concentrated hydrochloric acid.31 A 300 mL solution of 0.667 mol L-1
11
12
13 potassium permanganate was stirred and heated at 100 oC, and then 45 mL 6 mol L-1 hydrochloric
14
15
16 acid was added dropwise to the boiling solution with vigorous stirring at a speed of 0.7 mL min-1
17
18 with a constant flow pump. After 30 min reaction, the precipitate was aged for 12 h at 60 oC. The
19
20
21 mineral was washed with deionized water until the conductivity of the filtrate was less than 20 S
22
23
24 cm-1, and subsequently washed three times with ethanol. The product was dried in an oven at 40 oC.
25
26 The synthetic mineral was ground in an agate mortar and sieved with 100-mesh sieve.
27
28
29
30
31
2.2. Electrochemical Removal of Cd2+. Electrochemical experiments were carried out at room
32
33
34 temperature in a three-electrode system, using a saturated calomel electrode as reference electrode,
35
36
37 1.53.0 cm2 carbon fabric as counter electrode. The volume of electrolyte was 30 mL, and 0.1 mol
38
39 L-1 Na2SO4 was used as background electrolyte. The working electrodes were prepared by mixing
40
41
42 75 wt% of the as-prepared birnessite, 15 wt% of acetylene black (AB), and 10 wt% of
43
44
45 polyvinylidene fluoride (PVDF) slurry. The mixtures of the as-prepared birnessite and AB were
46
47 ground for 30 min, and then transferred to a 5 mL centrifuge tube in which PVDF with
48
49
50 N-methyl-2-pyrrolidone (NMP) as dispersant was added. After ultrasonic dispersion, the slurry was
51
52
uniformly coated on carbon fabric and dried in a vacuum drying oven at 40 oC for 12 h. The loading
53
54
55 of electrode materials (birnessite, AB, and PVDF) was controlled to 5 mg. CdSO4 solutions with the
56
57
58 initial concentrations of 0, 200, 400, 600, 800, 1000, 1400, 1800, and 2200 mg L-1 were
59 5
60
ACS Paragon Plus Environment
ACS Applied Materials & Interfaces Page 6 of 39

1
2
3
respectively used as electrolytes and the pH of the electrolyte was adjusted to 6.0 with 0.1 mol L-1
4
5 H2SO4 or NaOH. 50 cycles of galvanostatic charge-discharge tests (battery testing system,
6
7
8 Shenzhen Neware Electronic Ltd., China) were carried out in electrolytes with different initial
9
10
concentrations of Cd2+ to study the electrosorption capacity of the as-prepared birnessite for Cd2+.
11
12
13 To simply compare the Cd2+ ion adsorption rate, the 1, 4, 7, 10, 20, 30, 40, and 50 cycles of
14
15
16 galvanostatic charge-discharge tests were conducted and the corresponding electrosorption
17
18 capacities were recorded in 200 and 2200 mg L-1 Cd2+ electrolytes. The potential window was set as
19
20
21 00.9 V (vs SCE) and the current density was 0.1 A g-1. The electrosorption capacity was calculated
22
23
24 using formula (1):
25
26
27
28 (C0 Ce )V
29 Qe = (1)
30
m
31
32
33
34 where Qe (mg g-1) is the electrosorption capacity, C0 (mg L-1) is the initial concentration of Cd2+, Ce
35
36
37 (mg L-1) is the concentration of Cd2+ after charge-discharge test, V (mL) is the volume of electrolyte
38
39 and m (mg) is the mass of active birnessite on the electrode. Specific capacitance C (F g-1) of
40
41
42 birnessite was calculated using formula (2):
43
44
45
46
47
It
C= (2)
48 m V
49
50
51
52
where I (A) is the discharge current, t (s) is the discharge time, m (mg) is the mass of birnessite on
53
54
55 the electrode and V (V) is the potential window.
56
57
58
59 6
60
ACS Paragon Plus Environment
Page 7 of 39 ACS Applied Materials & Interfaces

1
2
3
In order to study the effect of loading of active birnessite on electrosorption capacity for Cd2+, 50
4
5 cycles of galvanostatic charge-discharge tests were carried out with different masses of electrode
6
7
8 materials: 0, 5, 10, 15, and 20 mg in 200 mg L-1 Cd2+ solution with an initial pH of 6.0. The effects
9
10
of initial pH values of electrolyte on electrosorption capacity for Cd2+ were also investigated.
11
12
13 Galvanostatic charge-discharge tests of 50 cycles were carried out with 5 mg electrode materials in
14
15
16 200 mg L-1 Cd2+ solution with different values of initial pH. The initial pH values of electrolyte
17
18 were respectively adjusted to 3.0, 3.5, 4.0, 4.5, 5.0, 5.5, and 6.0 with 0.1 mol L-1 H2SO4 or NaOH.
19
20
21 The above experiments were performed three times and the mean was recorded.
22
23
24 The cyclic electrochemical adsorption experiments of birnessite electrode were conducted at pH
25
26 6.0. After 50 cycles of charge-discharge tests in the electrolytes containing different concentrations
27
28
29 of Cd2+ (200 and 1000 mg L-1), the working electrodes were respectively transferred into 0.1 mol
30
31
L-1 Na2SO4 electrolyte for 50 cycles of charge-discharge activation. After activation and
32
33
34 regeneration, the corresponding birnessite electrode was then transferred into the mixed solutions of
35
36
37 0.1 mol L-1 Na2SO4 and Cd2+ with the same concentrations (200 and 1000 mg L-1) to test the
38
39 electrochemical adsorption performance for the second time.
40
41
42 To compare the electrochemical adsorption capacity, constant cell voltages deionization was also
43
44
45 performed at 0, 0.3, 0.6, and 0.9 V in the electrolyte containing 200 mg L-1 Cd2+ for 12 h. Carbon
46
47 fabric and the as-prepared birnessite were used as positive and negative electrodes, respectively.
48
49
50 The potential of birnessite electrode was determined using a saturated calomel electrode during the
51
52
electrochemical processes.
53
54
55 In order to explore the adsorption sites of Cd2+ on birnessite, the sorption isotherms experiments
56
57
58 for Cd2+ were performed. The synthetic birnessite was suspended in 0.1 mol L-1 Na2SO4 solution to
59 7
60
ACS Paragon Plus Environment
ACS Applied Materials & Interfaces Page 8 of 39

1
2
3
obtain 5 g L-1 suspension. The suspension was equilibrated for a few days, during which the pH of
4
5 the suspension was adjusted to 6.00.05 with 0.1 mol L-1 H2SO4 or NaOH. Cd2+ solutions were
6
7
8 prepared using 15 mmol L-1 CdSO4 with 0.1 mol L-1 Na2SO4. Sorption isotherm experiments were
9
10
performed by adding 08 mL Cd2+ solution to 5 mL mineral suspension in a series of 50 mL
11
12
13 polypropylene bottles, respectively. The total volume of solution in each bottle was made up to 15
14
15
16 mL by adding 0.1 mol L-1 Na2SO4. The final suspensions were shaken at 251 oC and 200 r min-1.
17
18 The pH of the suspensions was adjusted to 6.00.05 during the equilibration period. After 24 h of
19
20
21 equilibration, the suspensions were filtered by 0.22-m microporous membrane. The sorption
22
23
24 isotherm experiment was performed three times and the mean was recorded.
25
26 All of the above chemical reagents were of analytical grade and were purchased from China
27
28
29 National Medicine Group Shanghai Chemical Reagent Company. All aqueous solutions at different
30
31
concentrations were freshly prepared with ultrapure water (18 M cm resistivity). The carbon
32
33
34 fabric (HCP020) was supplied by Shanghai Hesen Electric Co., Ltd.
35
36
37
38
39 2.3. Characterization and Analysis. The crystal structure of the as-prepared birnessite was
40
41
42 characterized by X-ray powder diffraction (XRD, Bruker D8 Advance diffractometer with Cu K)
43
44
45 at a tube voltage of 40 kV, tube current of 40 mA, scanning rate of 1 min-1 and step intervals of
46
47 0.02. Field emission scanning electron microscope (FESEM, Hitachi, SU8000) was used to
48
49
50 observe the morphology of the product. The functional groups of synthetic birnessite were
51
52
characterized by Fourier transform infrared spectroscopy (FTIR, Nicolet 8700). The average
53
54
55 oxidation state of manganese in birnessite was determined by the oxalic acid-permanganate
56
57
58 back-titration method.32 The BET specific surface area and pore size distribution of the as-prepared
59 8
60
ACS Paragon Plus Environment
Page 9 of 39 ACS Applied Materials & Interfaces

1
2
3
samples were measured with a fully automatic surface area analyzer (Micromeritics ASAP2020)
4
5 using nitrogen isothermal adsorption-desorption method. About 0.2 g of sample was degassed at
6
7
8 110 C for 12 h. X-ray photoelectron spectroscopy (XPS, Multilab 2000) with a monochromatized
9
10
Mg K X-ray source (1253.6 eV) was used to analyze the relative contents of Mn with different
11
12
13 valence states in the as-prepared birnessite, using C1s peak (284.6 eV) as charge referencing.
14
15
16 The chemical composition of the product was determined by the following method: 0.01 g
17
18 sample and 0.1 g hydroxylamine hydrochloride were accurately weighed, and after dissolving in
19
20
21 deionized water, the volume of the solution was fixed to 250 mL with deionized water. The contents
22
23
24 of Mn and K in the dilute solution were measured by atomic absorption spectroscopy (AAS, Varina
25
26 AAS240FS) and flame photometer (Sherwood Model 410). The crystal water content of the sample
27
28
29 was calculated by thermal gravimetric analysis (TG, Diamond TG, PerkinElmer Instruments) at a
30
31 o
heating rate of 10 C min-1 under nitrogen atmosphere. Before and after 50 cycles of
32
33
34 charge-discharge tests and constant cell voltages deionization experiment, the concentrations of
35
36
37 Cd2+ and released Mn2+ in liquid phase were measured by AAS.
38
39 The cyclic voltammetry (CV) tests of the as-prepared birnessite in Cd2+ electrolytes were carried
40
41
42 out on the electrochemical workstation (CHI660E, Shanghai Chenhua Instrument Co. Ltd., China)
43
44
45 at a scan rate of 0.5 mV s-1 in a potential window of 00.9 V (vs. SCE). The influences of
46
47 concentration, pH, and the mass of electrode materials on the electrochemical behaviors of Cd2+
48
49
50 were further investigated.
51
52
53
54
55 3. RESULTS
56
57
58
59 9
60
ACS Paragon Plus Environment
ACS Applied Materials & Interfaces Page 10 of 39

1
2
3
3.1. Preparation and Characterization of Birnessite. Figure 1a shows the XRD pattern of the
4
5 as-prepared product, and the six main peaks, which respectively occurred at 0.713, 0.355, 0.246,
6
7
8 0.177, 0.142, and 0.122 nm, corresponded to the (001), (002), (100), (004), (110), and (202) crystal
9
10
planes of birnessite (JCPDS No. 86-0666). According to the d value of (001) crystal plane (d001), the
11
12
13 average crystal size was 7.55 nm as calculated by the Scherrer formula.33 There was no other
14
15
16 detectable peak in the XRD patterns, indicating that single-phase birnessite was obtained. The
17
18 FESEM image of the synthetic birnessite indicated that uniform flower-like microspheres (with
19
20
21 diameters of about 300 nm) composed of lamellar plates were formed (Figure 1b), which was
22
23
24 consistent with the typical morphology of hexagonal birnessite.34
25
26 The product was further characterized by FTIR spectroscopy (Figure S1). The two bands at 446
27
28
29 and 518 cm-1 could be attributed to the Mn-O lattice vibrations of birnessite, and the dominant
30
31
absorption peaks at 1645 and 3391 cm-1 were attributed to the stretching and bending vibrations of
32
33
34 adsorbed water and crystalline water.35 The results of FTIR further showed that the synthesized
35
36
37 product was single-phase birnessite. The manganese average oxidation state (AOS) was determined
38
39 to be 3.79 by the oxalic acid-permanganate back-titration method. The total Mn and K contents of
40
41
42 the synthetic birnessite were 53.72% and 9.92%, respectively. The content of crystal water in the
43
44
45 product was 13.22% as measured by TGA (Figure S2). According to above analyses, the chemical
46
47 composition of the birnessite was calculated to be K0.26MnO2.020.75H2O.
48
49
50 Figure S3 shows the N2 adsorption-desorption isotherms and pore size distributions of the
51
52
birnessite. As can be observed, the isotherms followed typical IUPAC type-V adsorption isotherm
53
54
55 patterns with the presence of a H3-type hysteresis loop, indicating that the birnessite had
56
57
58 mesoporous structures.36 The BET specific surface area and pore volume of the sample were 14.2
59 10
60
ACS Paragon Plus Environment
Page 11 of 39 ACS Applied Materials & Interfaces

1
2
3
m2 g-1 and 0.067 cm3 g-1, respectively. The average pore diameter was calculated to be 3.585 nm by
4
5 the Barrett-Joyner-Halenda (BJH) method (Figure S3b).
6
7
8
9
10
3.2. Electrochemical Removal of Cd2+. Charge-discharge tests were carried out with different
11
12
13 initial concentrations of Cd2+ in electrolyte ranging from 0 to 2200 mg L-1 with 5 mg electrode
14
15
16 material. As shown in Figure 2a, when the concentration of Cd2+ was 0 mg L-1, the specific
17
18 capacitance was kept at about 167.9 F g-1. This result indicated that the capacitive property of the
19
20
21 birnessite was stable in 0.1 mol L-1 Na2SO4 solution, which was consistent with the results of
22
23
24 previous studies.37,38 After Cd2+ was added to the electrolyte, the initial specific capacitance was
25
26 obviously higher than that in 0.1 mol L-1 Na2SO4 solution and gradually increased with the
27
28
29 increasing concentration of Cd2+. Meanwhile, with the increase of charge-discharge cycles, decay of
30
31
specific capacitance was observed, and the higher the Cd2+ concentration was, the more significant
32
33
34 the decay of specific capacitance would be. The results suggested that the electrochemical stability
35
36
37 of birnessite decreased in the electrolyte containing Cd2+.
38
39 The electrochemical behaviors of Cd2+ on the birnessite electrode surface were confirmed by CV.
40
41
42 Figure 2b shows the CV curves with 5 mg electrode materials in different concentrations of Cd2+ at
43
44
45 a scan rate of 0.5 mV s-1. When the concentration of Cd2+ was 0 mg L-1 in the electrolyte, a couple
46
47 of symmetrical redox peaks were observed in the CV curves. The anodic current peak at 0.55 V
48
49
50 corresponded to the oxidation of Mn(III) to Mn(IV) while the cathodic peak at 0.35 V corresponded
51
52
to the reduction of Mn(IV) to Mn(III).39 In addition, a rectangular shape of the CV curve was
53
54
55 observed, suggesting excellent reversibility and capacitance properties of the birnessite electrode in
56
57
58 0.1 mol L-1 Na2SO4 solution. When there was Cd2+ in the electrolyte, the closed area of CV curves
59 11
60
ACS Paragon Plus Environment
ACS Applied Materials & Interfaces Page 12 of 39

1
2
3
increased with the increasing concentration of Cd2+, indicating that the intercalation of Cd2+ could
4
5 increase the specific capacitance of birnessite, which was consistent with the results of galvanostatic
6
7
8 charge-discharge tests (Figure 2a). However, the CV curves deviated from a regular rectangular
9
10
shape in the Cd2+ electrolyte compared with single Na2SO4 electrolyte, and the degree of deviation
11
12
13 increased with the increase of Cd2+ concentration. These results confirmed that the intercalation of
14
15
16 Cd2+ in birnessite reduced the reversibility of the electrode.
17
18 In order to determine the electrosorption capacity of birnessite for Cd2+, the concentration of Cd2+
19
20
21 in the bulk electrolyte was further measured. After 50 cycles of charge-discharge tests in the
22
23
24 electrolyte with different initial concentrations of Cd2+, the electrosorption capacity of Cd2+ was
25
26 measured as shown in Figure 3a. When the concentration of Cd2+ was 200 mg L-1, the
27
28
29 electrosorption capacity of Cd2+ was 196.5 mg g-1. The electrosorption capacity increased at first
30
31
with the increase of initial Cd2+ concentrations and then reached equilibrium in 1400 mg L-1 Cd2+.
32
33
34 The maximum electrosorption capacity reached as high as 900.7 mg g-1 when Cd2+ concentration
35
36
37 was increased to 2200 mg L-1. However, the removal efficiency decreased with the increase of Cd2+
38
39 concentration in the electrolyte (Figure S4). The removal efficiency reached the maximum of
40
41
42 12.28% with the initial concentration of Cd2+ at 200 mg L-1, and it decreased to 5.16% at the Cd2+
43
44
45 concentration of 2200 mg L-1. In the initial stage, the adsorption capacity increased with an increase
46
47 in the cycles of charge-discharge. Low concentrations of Cd2+ facilitated the rapid electrochemical
48
49
50 adsorption. For example, the electrosorption capacity reached equilibrium at 10 and 20 cycles for
51
52
the electrolyte at the Cd2+ concentrations of 200 and 2200 mg L-1, respectively (Figure 3b).
53
54
55 In order to analyze the effect of Cd2+ enrichment on the crystal structure of birnessite, the wet
56
57
58 working electrode was characterized using XRD (Figure 4). The birnessite was not converted into
59 12
60
ACS Paragon Plus Environment
Page 13 of 39 ACS Applied Materials & Interfaces

1
2
3
other crystal phases. The d001 after charge-discharge tests gradually decreased with increasing Cd2+
4
5 concentration in the electrolyte, indicating a decrease of crystallinity and partial dissolution of
6
7
8 birnessite. The d110 value of initial birnessite was 0.1419 nm and the values after charge-discharge
9
10
tests in 200, 600, and 1000 mg L-1 Cd2+ solutions increased to 0.1428, 0.1424, and 0.1429 nm,
11
12
13 respectively, suggesting that the Mn AOS of the birnessite decreased after charge-discharge tests.40
14
15
16 The chemical components of the pristine birnessite and birnessite-based electrodes after
17
18 charge-discharge tests in 1000 mg L-1 Cd2+ electrolyte were further characterized by XPS. XPS
19
20
21 broad scans of the initial birnessite sample showed that Mn, O, and K were present in the birnessite
22
23
24 structure (Figure 5a). Cd was observed in the XPS broad scans and the Cd3d spectrum (Figure 5a, b)
25
26 after charge-discharge tests, indicating the presence of Cd on the surface of electrode materials. In
27
28
29 order to infer the relative contents of the three species Mn(II), Mn(III), and Mn(IV), the multiplet
30
31
fitting of the Mn 2p3/2 peaks was conducted as shown in Figure 5c, d. After charge-discharge tests
32
33
34 in 1000 mg L-1 Cd2+ electrolyte, the relative contents of Mn(II), Mn(III), and Mn(IV) varied from
35
36
37 6.62% to 4.17%, from 13.10% to 16.34%, and from 82.73% to 77.04%, respectively. The
38
39 corresponding Mn AOS decreased from 3.78 to 3.70, which was consistent with the result of XRD
40
41
42 (Figure 4).
43
44
45 SEM was used to analyze the effect of Cd on the morphology of birnessite. As shown in Figure
46
47 6a, the flower-like microspheres of birnessite disappeared due to dissolution-recrystallization
48
49
50 process during the charge-discharge tests in Na2SO4 solution.37 After charge-discharge tests in 200
51
52
mg L-1 Cd2+ electrolyte, the birnessite retained the morphology of flower-like microspheres and was
53
54
55 partially dissolved (Figure 6b). However, the flower-like microspheres disappeared in 600 and 1000
56
57
58 mg L-1 Cd2+ electrolytes (Figure 6c, d), and the diameter of birnessite particles decreased with
59 13
60
ACS Paragon Plus Environment
ACS Applied Materials & Interfaces Page 14 of 39

1
2
3
increasing the concentration of Cd2+, suggesting that the dissolution of the birnessite was increased.
4
5 This was consistent with the results of XRD experiments (Figure 4).
6
7
8 The cyclic electrochemical adsorption performance of the birnessite electrode was studied at pH
9
10
6.0. Compared with the capacitance performance of fresh birnessite electrode in 0.1 mol L-1 Na2SO4,
11
12
13 no significant difference in specific capacitance was observed when the birnessite electrode was
14
15
16 transferred from the electrolyte containing 200 mg L-1 Cd2+ to background electrolyte of 0.1 mol L-1
17
18 Na2SO4; however, when the initial concentration of Cd2+ was controlled at 1000 mg L-1, the specific
19
20
21 capacitance of the birnessite electrode significantly decreased after transferring to 0.1 mol L-1
22
23
24 Na2SO4 solution (Figure 2a and Figure S5). After 50 cycles of charge-discharge activation and
25
26 regeneration in 0.1 mol L-1 Na2SO4 solution, the release ratios of the electrochemically adsorbed
27
28
29 Cd2+ ions were about 49.77% and 11.97% for the birnessite electrodes previously treated in 200 and
30
31
1000 mg L-1 Cd2+ solutions, respectively. When the activated birnessite electrodes were used to
32
33
34 remove Cd2+ ions for the second time, the specific adsorption capacity reached 90.2 and 229.5 mg
35
36
37 g-1 when the initial Cd2+ concentration was controlled at 200 and 1000 mg L-1, respectively.
38
39
40
41
42 3.3. Effect of Initial pH and Loading of Active Birnessite. The pH could affect the capacitance
43
44
45 characteristics of birnessite41 and thereby affect the electrochemical removal of Cd2+. To investigate
46
47 the effect of initial pH, 50 cycles of galvanostatic charge-discharge tests were carried out in 200 mg
48
49
50 L-1 Cd2+ electrolyte with pH ranging from 3.0 to 6.0 using 5 mg electrode material. As shown in
51
52
Figure S6a, the initial specific capacitance decreased with increasing pH, which was consistent with
53
54
55 the results of CV (Figure S6b). Decay of specific capacitance was observed with increasing cycles
56
57
58 and became more obvious with decreasing pH. As shown in Figure S7, the XRD patterns of the wet
59 14
60
ACS Paragon Plus Environment
Page 15 of 39 ACS Applied Materials & Interfaces

1
2
3
working electrode indicated that the birnessite was not converted into other crystal phases and the
4
5 relative intensity of diffraction peak gradually decreased with decreasing pH, indicating that some
6
7
8 of the birnessite was dissolved at low pH values. Figure 7a shows that the electrosorption capacity
9
10
of Cd2+ was 54.78 mg g-1 at pH 3.0 and rose at first with increasing pH and then reached
11
12
13 equilibrium at pH 4.0. After 50 cycles of charge-discharge tests, the release of Mn2+ decreased with
14
15
16 the increase of pH (Figure S8), which agreed with the fact that high pH facilitated the chemical
17
18 stability of birnessite (Figure S7).
19
20
21 In order to study the effect of birnessite loading on electrosorption capacity of Cd2+, the loading
22
23
24 of electrode materials was controlled at 0, 5, 10, 15, and 20 mg. As shown in Figure S9a, the
25
26 specific capacitance of pure carbon fabric was about 0 F g-1, confirming that the carbon fabric
27
28
29 almost had no electrochemical activity. The initial specific capacitance decreased with increasing
30
31
loading of birnessite on the electrode, which was consistent with the results of CV (Figure S9b).
32
33
34 The decay of specific capacitance was observed with increasing cycles and became more significant
35
36
37 with decreasing loading. The XRD patterns (Figure S10) of the wet working electrode showed that
38
39 the birnessite was not converted into other crystal phases. The electrosorption capacity of Cd2+
40
41
42 using 0 mg electrode material was 0 mg g-1, which was consistent with the result of galvanostatic
43
44
45 charge-discharge tests (Figure S9a). Hence, the Cd2+ removal was not affected by carbon fabric.
46
47 The electrosorption capacity of Cd2+ using 5 mg electrode material was 196.5 mg g-1 and decreased
48
49
50 with increasing loading of electrode materials (Figure 7b).
51
52
53
54
55 4. DISCUSSION
56
57
58
59 15
60
ACS Paragon Plus Environment
ACS Applied Materials & Interfaces Page 16 of 39

1
2
3
4.1. Electrochemical Removal Mechanism. In the charge-discharge process of electric
4
5 double-layer capacitors, the charged ions in the electrolyte could move towards the oppositely
6
7
8 charged electrodes and accumulate on the electrode surface due to electrostatic forces. This
9
10
principle is used for desalination.1120 The charge storage mechanism of pseudocapacitance for
11
12
13 manganese oxides could be described as MnO2 + A+ + e- MnO2A, where A+ is alkali metal ion or
14
15
16 H+.24,39 During the charge-discharge processes of birnessite in Na2SO4 solution, the redox reaction
17
18 between Mn(IV) and Mn(III) is accompanied by the intercalation-deintercalation of Na+ and H+,
19
20
21 and this process is preferably reversible.37,39 In this work, the symmetrical rectangular shape of CV
22
23
24 curve and the stable crystal structure (Figure 2b and Figure 4) also confirmed that the birnessite in
25
26 Na2SO4 solution had good pseudocapacitive performance. In this work, the specific capacitance of
27
28
29 birnessite increased with increasing Cd2+ concentration due to the increase of conductance and ion
30
31
diffusion rate. However, the increase of Cd2+ concentration improved the specific capacitance of
32
33
34 birnessite only at the initial stage, which was accompanied by the decrease of cyclic stability
35
36
37 (Figure 2a), and the room for the increase of capacitance was limited, which could also be indicated
38
39 by cyclic voltammograms (Figure 2b). The ions with large hydrated radius lead to the decrease in
40
41
42 the reversibility of manganese oxide electrodes. In the charge-discharge process of batteries,
43
44
45 incompletely reversible intercalation-deintercalation of Zn2+ and Mg2+ causes Mn2+ release and the
46
47 decline of specific capacitance of birnessite and -MnO2.42,43 In this work, the (001) layer spacing
48
49
50 of the as-prepared birnessite was 0.71 nm. When the electrolyte contained Cd2+, the large hydrated
51
52
radius of Cd2+ (0.426 nm)44 led to poorly reversible intercalation-deintercalation. The Cd2+ adsorbed
53
54
55 on electrode was incompletely deintercalated in the charge process, and after several cycles of
56
57
58 charge-discharge, the Cd2+ was accumulated on the electrode. The adsorption capacity increased
59 16
60
ACS Paragon Plus Environment
Page 17 of 39 ACS Applied Materials & Interfaces

1
2
3
and then reached equilibrium with the increase of the cycles of charge-discharge, and high initial
4
5 concentration facilitated the irreversible enrichment of Cd2+ on the birnessite electrode (Figure 3b).
6
7
8 The Mn2+ release of MnO2 electrode also suggested the incompletely reversible electrochemical
9
10
processes. During the discharge process, Mn(IV) is reduced to Mn(III), which can be
11
12
13 disproportionated into Mn2+ ions and Mn(IV) oxides due to the Jahn-Teller effect.43 During the
14
15
16 discharge process, approximately one-third of the total Mn in cryptomelane was released into
17
18 ZnSO4 electrolyte, which caused a decrease of specific capacitance.43,45 Here, the content of
19
20
21 released Mn2+ was shown in Figure S11 after charge-discharge tests with different initial
22
23
24 concentrations of Cd2+. When there was Cd2+ in the electrolyte, the maximum Mn2+ release was
25
26 38.02% of the total Mn in the birnessite, which was significantly higher than the Mn2+ release in the
27
28
29 absence of Cd2+ (about 3.40%). This result suggests that the Mn2+ released into the electrolyte is
30
31
derived from the reduction of Mn(IV). The incomplete deintercalation of Cd2+ adsorbed on the
32
33
34 electrode in the charge process results in the incomplete oxidation of Mn(III) to Mn(IV), and the
35
36
37 unoxidized Mn(III) disproportionates into Mn(IV) oxide and Mn2+ ions, which are then released
38
39 into the electrolyte. At the same time, due to the Mn2+ release, the content of active Mn in the
40
41
42 electrode decreased, which caused the decay of specific capacitance of the birnessite. The release of
43
44
45 Mn2+ increased with increasing Cd2+ concentration (Figure S11), leading to a decrease of Mn AOS
46
47 and XRD diffraction peak (Figure 4) after charge-discharge.
48
49
50 Besides the release of activated Mn, the undesorpted Cd2+ ions in the interlayer of birnessite
51
52
also significantly decreased the adsorption capacity during the regeneration process. However, a
53
54
55 high electrochemical adsorption capacity of 229.5 mg g-1 suggested that the birnessite electrodes
56
57
58 could be potentially re-used for many times.
59 17
60
ACS Paragon Plus Environment
ACS Applied Materials & Interfaces Page 18 of 39

1
2
3
When ordered mesoporous carbon/carbon nanotube composite electrodes were used to remove
4
5 NaCl, the ions could be adsorbed onto the electrode surface or into the mesoporous by electrostatic
6
7
8 forces.11 In sorption isotherms of Cd2+ on birnessite, the Cd2+ adsorbed onto birnessite was found to
9
10
occupy the sites above and below MnO6 octahedral vacancies.29 Therefore, in the discharge process,
11
12
13 Cd2+ moves towards the working electrode due to electrostatic forces. After reaching the electrode
14
15
16 surface, a part of Cd2+ ions occupy MnO6 octahedral vacancies or are inserted into the interlayer
17
18 and the other part are adsorbed on the electrode surface or into the mesoporous. Obviously, the
19
20
21 mesoporous structure is favorable for the electrochemical adsorption of Cd2+ by birnessite.
22
23
24 In order to further analyze the sorption site of Cd2+ on birnessite, the sorption isotherms of Cd2+
25
26 on synthetic birnessite were carried out. As shown in Figure 8, the adsorption capacity sharply
27
28
29 increased at low concentrations and then the increase rate slowed down, and the sorption isotherm
30
31
curve was conformed to L-type.30 The data were fitted by the Langmuir equation and the maximum
32
33
34 adsorption capacity was 125.8 mg g-1 (1.12 mmol g-1), which is consistent with the results of
35
36
37 previous studies.29 The maximum Mn2+ release was calculated to be 2.14% of the total Mn in the
38
39 birnessite, and in addition, the crystal structure of birnessite was not obviously changed by the
40
41
42 adsorption of Cd2+ as seen from the XRD patterns (Figure S12). Mn2+ in the solution was derived
43
44
45 from the Mn(II) which was exchanged by Cd2+ from the sites above and below octahedral vacancies.
46
47 The above analyses demonstrate that Cd2+ is mainly adsorbed on the electrode surface or into the
48
49
50 mesoporous channel.
51
52
Higher adsorption capacity of Cd2+ was obtained owing to the electrochemical processes. A series
53
54
55 of Cd2+ adsorption capacities of different adsorbents were compared (Table S1). As for isothermal
56
57
58 adsorption, the capacity for activated carbon derived from bagasse, synthetic zeolites, activated
59 18
60
ACS Paragon Plus Environment
Page 19 of 39 ACS Applied Materials & Interfaces

1
2
3
carbon cloth and manganese dioxide nanosheets reached as high as 49.07, 50.8, 146, 348.1 mg g-1,
4
5 respectively.5,9,10,30 The inorganic adsorption capacity of as-obtained birnessite for Cd2+ was 125.8
6
7
8 mg g-1, which approached the previously reported value of 148 mg g-1.29 During the
9
10
charge-discharge processes, dissolution-recrystallization decreased the particle size and increased
11
12
13 the specific surface area of the birnessite electrode (Figure 6). Larger surface area and the
14
15
16 incompletely reversible intercalation-deintercalation reaction remarkably enhanced the adsorption
17
18 specific capacity during the charge-discharge processes.
19
20
21 In this study, multi-cycle charge-discharge processes contributed to the larger surface area owing
22
23
24 to the dissolution-recrystallization. Capacitive deionization was also conducted to compare the
25
26 removal capacity under a constant voltage. When cell voltages were set at 0, 0.3, 0.6, and 0.9 V for
27
28
29 12 h, the adsorption capacities were 59.3, 64.9, 105.6, and 161.8 mg g-1, the contents of released
30
31
Mn2+ were about 5.6, 4.8, 4.6, and 4.7 mg g-1, and the removal efficiencies were 3.71, 4.05, 6.60,
32
33
34 and 10.11%, respectively (Figure S13). The potential of birnessite electrode was respectively
35
36
37 determined to be 0.33, 0.30, 0.23, and 0.03 V (vs SCE) in the above systems (Figure S14).
38
39 Electrostatic interaction was possibly the main mechanism for the removal of Cd2+, as indicated by
40
41
42 the excellent chemical stability in crystal structure (Figure S15) and low release of Mn2+ during the
43
44
45 electrochemical adsorption processes. The potential for the electrochemical intercalation of Cd2+
46
47 into birnessite was lower than 0.3 V, as shown in the cyclic voltammograms (Figure. 2b). In the
48
49
50 present work, when the cell voltage of 0.9 V was applied, the potential of the birnessite electrode
51
52
reached 0.03 V, which facilitated the intercalation of Cd2+ into birnessite with the highest removal
53
54
55 capacity. These results further indicate that the pseudocapacitive property of birnessite contributies
56
57
58 greatly to the higher electrochemical Cd2+adsorption capacity.
59 19
60
ACS Paragon Plus Environment
ACS Applied Materials & Interfaces Page 20 of 39

1
2
3
4
5 4.2. Effect of Initial pH and Loading of Active Birnessite. In Na2SO4 electrolyte, the
6
7
8 concentration of H3O+ had a more significant effect on the specific capacitance of birnessite than
9
10
that of Na+.41 The mobility of H+ is considerably higher than that of Na+, which reduces the ion
11
12
13 transfer resistance between electrolyte and electrode interface.46 In this experiment, a lower pH of
14
15
16 the electrolyte usually resulted in higher initial specific capacitance. Decay of specific capacitance
17
18 was observed with increasing cycles and became more obvious with decreasing pH (Figure S6a),
19
20
21 which is mainly due to the dissolution of Mn and the consumption of H+. A previous study showed
22
23
24 that MnO2 would be reduced to Mn2+ and dissolved during charge-discharge in low pH
25
26 electrolytes.47 After 50 cycles of charge-discharge tests in 200 mg L-1 Cd2+ solution with different
27
28
29 pH values, the XRD patterns and the release of Mn2+ also showed that the birnessite was partially
30
31
dissolved (Figure S7, S8). The dissolution of Mn caused a decrease of active birnessite content on
32
33
34 the electrode and the consumption of H+, which led to the increase of ion transfer resistance
35
36
37 between electrolyte and the electrode interface. After charge-discharge tests in electrolyte with
38
39 initial pH of 3.0, 3.5, 4.0, 4.5, 5.0, 5.5, and 6.0, the pH values were changed to 3.510.05,
40
41
42 5.840.06, 6.540.07, 6.620.01, 6.840.04, 6.880.08, and 6.880.05, respectively. These results
43
44
45 further suggest that the release of Mn2+ led to the consumption of H+ and thereby the decrease of
46
47 specific capacitance. The release of Mn2+ also caused the reduction of active birnessite mass on the
48
49
50 electrode and thus the decrease of electrosorption capacity of Cd2+.
51
52
High concentration of H+ accelerates the disproportionation reaction and the release of Mn2+,
53
54
55 resulting in the decrease of the crystallinity of the birnessite and the capacity fade of supercapacitor
56
57
58
59 20
60
ACS Paragon Plus Environment
Page 21 of 39 ACS Applied Materials & Interfaces

1
2
3
during the charge-discharge processes. Therefore, the effect of pH on the dissolution of birnessite
4
5 should be attributed to the electrochemical processes rather than physical reaction.
6
7
8 The change of pH will affect the surface charge of birnessite in inorganic adsorption processes.
9
10
High pH will increase the negative charge on the surface of birnessite with a subsequent increase in
11
12
13 the adsorption capacity for heavy metals. In this work, the adsorption of Cd2+ occurred on the
14
15
16 birnessite electrode, and surface charge changed during the charge-discharge processes. The
17
18 electrosorption capacity of Cd2+ was 54.8 mg g-1 at pH 3.0 and rose at first with increasing pH and
19
20
21 then reached equilibrium at pH 4.0 (Figure 7a). There was no obvious effect of pH on the
22
23
24 electrosorption capacity of Cd2+ at pH 4.06.0. That is to say, the decrease of the competitive
25
26 adsorption of H+ at higher pH did not enhance the electrochemical adsorption capacity. These
27
28
29 results further confirm that electrostatic adsorption possibly plays a minor role in Cd2+
30
31
electrochemical adsorption. The finding that pseudocapacitive property of birnessite contributes
32
33
34 much to the electrochemical adsorption of Cd2+ could be further confirmed by other results. For
35
36
37 example, the adsorption capacity increased with the increase in the cycles of charge-discharge due
38
39 to the incompletely reversible intercalation-deintercalation reaction. The electrochemical stability
40
41
42 significantly decreased with the increase of Cd2+ concentration in the electrolyte owing to the redox
43
44
45 reaction accompanied by the release of active Mn from birnessite (Figure 2a). During the activation
46
47 and regeneration processes, the release ratio of electrochemically adsorbed Cd2+ reached as high as
48
49
50 49.77%, and these Cd2+ ions should be from the surface and the interlayers of birnessite.
51
52
In the charge-discharge process of MnO2-based supercapacitors, redox reactions occur on the
53
54
55 electrode surface.41,48 In this work, the electrode thickness increased with increasing loading of
56
57
58 electrode materials. During the charge-discharge process, the cations in the electrolyte could not
59 21
60
ACS Paragon Plus Environment
ACS Applied Materials & Interfaces Page 22 of 39

1
2
3
easily reach the inside of electrode materials. Therefore, the specific capacitance and the
4
5 electrosorption capacity of Cd2+ decreased with increasing the mass of birnessite on the electrode.
6
7
8 In this work, though the electrosorption capacity of birnessite for Cd2+ could reach up to 900.7
9
10
mg g-1, the Mn2+ release was evaluated to be about 204.3 mg g-1. In future studies, we will improve
11
12
13 the experimental devices and electrochemical methods to improve the stability of birnessite
14
15
16 electrode and reduce the release amount of Mn2+. Moreover, this method can also be applied to
17
18 remove other heavy metal ions from contaminated waters.
19
20
21
22
23
24 5. CONCLUSIONS
25
26 In summary, birnessite could be used as electrode materials to electrochemically remove Cd2+
27
28
29 from aqueous solutions based on the principle of supercapacitors. In the charge-discharge processes
30
31
accompanied by cation intercalation-deintercalation, Cd2+ moves towards the working electrode due
32
33
34 to electrostatic forces and incompletely reversible insertion reaction. The adsorption capacity
35
36
37 increased and then reached equilibrium with the increase in charge-discharge cycles. The maximum
38
39 electrosorption capacity was evaluated to be about 900.7 mg g-1 (8.01 mmol g-1), which is
40
41
42 significantly higher than the adsorption isotherm capacity (125.8 mg g-1). Upon reaching the
43
44
45 electrode surface, a small amount of Cd2+ occupies MnO6 octahedral vacancies and most of the
46
47 Cd2+ ions are adsorbed on the electrode surfaces or into the mesoporous channels of birnessite. In
48
49
50 low pH electrolyte, the partial dissolution of the birnessite would result in the decrease of Cd2+
51
52
removal efficiency. The electrosorption capacity of Cd2+ increased with increasing pH and reached
53
54
55 equilibrium in the pH range of 4.06.0. Excess birnessite on the electrode is not in full contact with
56
57
58 the electrolyte, causing a decrease of Cd2+ electrosorption capacity. This work provides a new
59 22
60
ACS Paragon Plus Environment
Page 23 of 39 ACS Applied Materials & Interfaces

1
2
3
method for the efficient adsorption of high concentration Cd2+ and other heavy metal ions from
4
5 wastewater.
6
7
8
9
10
ASSOCIATED CONTENTS
11
12
13 Supporting Information
14
15
16 This includes the conventional characterization of as-prepared birnessite, electrochemical
17
18 properties for birnessite electrodes, XRD patterns of birnessite after charge-discharge test, and
19
20
21 adsorption isotherm, released Mn2+ concentration and electrosorption capacity for Cd2+ after
22
23
24 charge-discharge tests and capacitive deionization experiments under constant voltages. This
25
26 material is available free of charge via the Internet at http://pubs.acs.org.
27
28
29
30
31
AUTHOR INFORMATION
32
33
34 Corresponding Author
35
36
37 *Qiu GH, E-mail: qiugh@mail.hzau.edu.cn.
38
39 Notes
40
41
42 The authors declare no competing financial interest.
43
44
45
46
47 ACKNOWLEDGMENTS
48
49
50 The authors thank the National Natural Science Foundation of China (Grant No. 41571228,
51
52
41330852), the Fok Ying-Tong Education Foundation (Grant No. 141024), the Natural Science
53
54
55 Foundation of Hubei Province of China (Grant No. 2014CFA016), and the Fundamental Research
56
57
58 Funds for the Central Universities (Program No. 2662015JQ002) for financial support. Steven L.
59 23
60
ACS Paragon Plus Environment
ACS Applied Materials & Interfaces Page 24 of 39

1
2
3
Suib thanks support of the US Department of Energy, Office of Basic Energy Sciences, Division of
4
5 Chemical, Biological and Geological Sciences under grant DE-FG02-86ER13622.A000.
6
7
8
9
10
REFERENCES
11
12
13 (1) Qu, X.; Alvarez, P. J. J.; Li, Q. Applications of Nanotechnology in Water and Wastewater
14
15
16 Treatment. Water Res. 2013, 47, 39313946.
17
18 (2) Nriagu, J. O.; Pacyna, J. M. Quantitative Assessment of Worldwide Contamination of Air,
19
20
21 Water and Soils by Trace Metals. Nature 1988, 333, 134139.
22
23
24 (3) Lv, M.; Wang, X.; Li, J.; Yang, X.; Zhang, C. A.; Yang, J.; Hu, H. Cyclodextrin-Reduced
25
26 Graphene Oxide Hybrid Nanosheets for the Simultaneous Determination of Lead(II) and
27
28
29 Cadmium(II) Using Square Wave Anodic Stripping Voltammetry. Electrochim. Acta 2013, 108,
30
31
412420.
32
33
34 (4) Riederer, A. M.; Belova, A.; George, B. J.; Anastas, P. T. Urinary Cadmium in the 19992008
35
36
37 U.S. National Health and Nutrition Examination Survey (NHANES). Environ. Sci. Technol. 2013,
38
39 47, 11371147.
40
41
42 (5) Mohan, D.; Singh, K. P. Single- and Multi-Component Adsorption of Cadmium and Zinc
43
44
45 Using Activated Carbon Derived from Bagasse-an Agricultural Waste. Water Res. 2002, 36,
46
47 23042318.
48
49
50 (6) Kurniawan, T. A.; Chan, G. Y. S.; Lo, W. H.; Babel, S. Physico-Chemical Treatment Techniques
51
52
for Wastewater Laden with Heavy Metals. Chem. Eng. J. 2006, 118, 8398.
53
54
55 (7) Charerntanyarak, L. Heavy Metals Removal by Chemical Coagulation and Precipitation. Water
56
57
58 Sci. Technol. 1999, 39, 135138.
59 24
60
ACS Paragon Plus Environment
Page 25 of 39 ACS Applied Materials & Interfaces

1
2
3
(8) Qdais, H. A.; Moussa, H. Removal of Heavy Metals from Wastewater by Membrane Processes:
4
5 a Comparative Study. Desalination 2004, 164, 105110.
6
7
8 (9) lvarez-Ayuso, E.; Garca-Snchez, A.; Querol, X. Purification of Metal Electroplating Waste
9
10
Waters Using Zeolites. Water Res. 2003, 37, 48554862.
11
12
13 (10) Rangel-Mendez, J. R.; Streat, M. Adsorption of Cadmium by Activated Carbon Cloth:
14
15
16 Influence of Surface Oxidation and Solution pH. Water Res. 2002, 36, 12441252.
17
18 (11) Peng, Z.; Zhang, D.; Shi, L.; Yan, T. High Performance Ordered Mesoporous Carbon/Carbon
19
20
21 nanotube Composite Electrodes for Capacitive Deionization. J. Mater. Chem. 2012, 22, 66036612.
22
23
24 (12) Jeon, S. I.; Park, H. R.; Yeo, J. G.; Yang, S.; Cho, C. H.; Han, M. H.; Kim, D. K. Desalination
25
26 via a New Membrane Capacitive Deionization Process Utilizing Flow-Electrodes. Energy Environ.
27
28
29 Sci. 2013, 6, 14711475.
30
31
(13) Porada, S.; Borchardt, L.; Oschatz, M.; Bryjak, M.; Atchison, J. S.; Keesman, K. J.; Kaskel,
32
33
34 S.; Biesheuvel, P. M.; Presser, V. Direct Prediction of the Desalination Performance of Porous
35
36
37 Carbon Electrodes for Capacitive Deionization. Energy Environ. Sci. 2013, 6, 37003712.
38
39 (14) Yoon, H.; Lee, J.; Kim, S. R.; Kang, J.; Kim, S.; Kim, C.; Yoon, J. Capacitive Deionization
40
41
42 with Ca-Alginate Coated-Carbon Electrode for Hardness Control. Desalination 2016, 392, 4653.
43
44
45 (15) Wang, H.; Yan, T.; Liu, P.; Chen, G.; Shi, L.; Zhang, J.; Zhong, Q.; Zhang D. In Situ Creating
46
47 Interconnected Pores across 3D Graphene Architectures and Their Application as High Performance
48
49
50 Electrodes for Flow-through Deionization Capacitors. J. Mater. Chem. A 2016, 4, 49084919.
51
52
(16) Macas, C.; Rasines, G.; Lavela, P.; Zafra, M. C.; Tirado, J. L.; Ania, C. O. Mn-Containing
53
54
55 N-Doped Monolithic Carbon Aerogels with Enhanced Macroporosity as Electrodes for Capacitive
56
57
58 Deionization. ACS Sustainable Chem. Eng. 2016, 4, 24872494.
59 25
60
ACS Paragon Plus Environment
ACS Applied Materials & Interfaces Page 26 of 39

1
2
3
(17) Wei, Y.; Huo, Y.; Tian, G.; Meng, Q.; Cao, B. Nitrogen-Doped Functional Graphene
4
5 Nanocomposites for Capacitive Deionization of NaCl Aqueous Solutions. J. Solid State
6
7
8 Electrochem. 2016, 20, 23512362.
9
10
(18) Wang, Z.; Yan, T.; Fang, J.; Shi, L.; Zhang, D. Nitrogen-Doped Porous Carbon Derived from
11
12
13 a Bimetallic MetalOrganic Framework as Highly Efficient Electrodes for Flow-through
14
15
16 Deionization Capacitors. J. Mater. Chem. A 2016, 4, 1085810868.
17
18 (19) Zhao, S.; Yan, T.; Wang, H.; Zhang, J.; Shi, L.; Zhang, D. Creating 3D Hierarchical Carbon
19
20
21 Architectures with Micro-, Meso-, and Macropores via a Simple Self-Blowing Strategy for a
22
23
24 Flow-through Deionization Capacitor. ACS Appl. Mater. Interfaces 2016, 8, 1802718035.
25
26 (20) Yin, H.; Zhao, S.; Wan, J.; Tang, H.; Chang, L.; He, L.; Zhao, H.; Gao, Y.; Tang, Z.
27
28
29 Three-Dimensional Graphene/Metal Oxide Nanoparticle Hybrids for High-Performance Capacitive
30
31
Deionization of Saline Water. Adv. Mater. 2013, 25, 62706276.
32
33
34 (21) Huang, Z.; Lu, L.; Cai, Z.; Ren, Z. J. Individual and Competitive Removal of Heavy Metals
35
36
37 Using Capacitive Deionization. J. Hazard. Mater. 2016, 302, 323331.
38
39 (22) Hu, C.; Liu, F.; Lan, H.; Liu, H.; Qu, J. Preparation of a Manganese Dioxide/Carbon Fiber
40
41
42 Electrode for Electrosorptive Removal of Copper Ions from Water. J. Colloid Interface Sci. 2015,
43
44
45 446, 359365.
46
47 (23) Mishra, A. K.; Ramaprabhu, S. Magnetite Decorated Multiwalled Carbon Nanotube Based
48
49
50 Supercapacitor for Arsenic Removal and Desalination of Seawater. J. Phys. Chem. C 2010, 114,
51
52
25832590.
53
54
55 (24) Wei, W.; Cui, X.; Chen, W.; Ivey, D. G. Manganese Oxide-Based Materials as
56
57
58 Electrochemical Supercapacitor Electrodes. Chem. Soc. Rev. 2011, 40, 16971721.
59 26
60
ACS Paragon Plus Environment
Page 27 of 39 ACS Applied Materials & Interfaces

1
2
3
(25) Cottineau, T.; Toupin, M.; Delahaye, T.; Brousse, T.; Blanger, D. Nanostructured Transition
4
5 Metal Oxides for Aqueous Hybrid Electrochemical Supercapacitors. Appl. Phys. A 2006, 82,
6
7
8 599606.
9
10
(26) Tang, H.; Wang, J.; Yin, H.; Zhao, H.; Wang, D.; Tang, Z. Growth of Polypyrrole Ultrathin
11
12
13 Films on MoS2 Monolayers as High-Performance Supercapacitor Electrodes. Adv. Mater. 2015, 27,
14
15
16 11171123.
17
18 (27) Yu, N.; Yin, H.; Zhang, W.; Liu, Y.; Tang, Z.; Zhu, M. Q. High-Performance Fiber-Shaped
19
20
21 All-Solid-State Asymmetric Supercapacitors Based on Ultrathin MnO2 Nanosheet/Carbon Fiber
22
23
24 Cathodes for Wearable Electronics. Adv. Energy Mater. 2016, 6, 1501458.
25
26 (28) Feng, X. H.; Zhai, L. M.; Tan, W. F.; Liu, F.; He, J. Z. Adsorption and Redox Reactions of
27
28
29 Heavy Metals on Synthesized Mn Oxide Minerals. Environ. Pollut. 2007, 147, 366373.
30
31
(29) Wang, Y.; Feng, X.; Villalobos, M.; Tan, W.; Liu, F. Sorption Behavior of Heavy Metals on
32
33
34 Birnessite: Relationship with its Mn Average Oxidation State and Implications for Types of
35
36
37 Sorption Sites. Chem. Geol. 2012, 292293, 2534.
38
39 (30) Peng, L.; Zeng, Q.; Tie, B.; Lei, M.; Yang, J.; Luo, S.; Song, Z. Manganese Dioxide
40
41
42 Nanosheet Suspension: A Novel Absorbent for Cadmium(II) Contamination in Waterbody. J.
43
44
45 Colloid Interface Sci. 2015, 456, 108115.
46
47 (31) Mckenzie, R. M. The Synthesis of Birnessite, Cryptomelane, and Some Other Oxides and
48
49
50 Hydroxides of Manganese. Mineral. Mag. 1971, 38, 493502.
51
52
(32) Kijima, N.; Yasuda, H.; Sato, T.; Yoshimura, Y. Preparation and Characterization of Open
53
54
55 Tunnel Oxide -MnO2 Precipitated by Ozone Oxidation. J. Solid State Chem. 2001, 159, 94102.
56
57
58 (33) Patterson, A. L. The Scherrer Formula for X-ray Particle Size Determination. Phys. Rev. 1939,
59 27
60
ACS Paragon Plus Environment
ACS Applied Materials & Interfaces Page 28 of 39

1
2
3
56, 978982.
4
5 (34) Liu, L.; Min, M.; Liu, F.; Yin, H.; Zhang, Y.; Qiu, G. Influence of Vanadium Doping on the
6
7
8 Supercapacitance Performance of Hexagonal Birnessite. J. Power Sources 2015, 277, 2635.
9
10
(35) Zhao, W.; Liu, F.; Feng, X.; Tan, W.; Qiu, G.; Chen, X. Fourier Transform Infrared
11
12
13 Spectroscopy Study of Acid Birnessites before and after Pb2+ Adsorption. Clay Miner. 2012, 47,
14
15
16 191204.
17
18 (36) Kruk, M.; Jaroniec, M. Gas Adsorption Characterization of Ordered Organic-Inorganic
19
20
21 Nanocomposite Materials. Chem. Mater. 2001, 13, 31693183.
22
23
24 (37) Athoul, L.; Moser, F.; Dugas, R.; Crosnier, O.; Blanger, D.; Brousse, T. Variation of the
25
26 MnO2 Birnessite Structure upon Charge/Discharge in an Electrochemical Supercapacitor Electrode
27
28
29 in Aqueous Na2SO4 Electrolyte. J. Phys. Chem. C 2008, 112, 72707277.
30
31
(38) Ghodbane, O.; Pascal, J. L.; Favier, F. Microstructural Effects on Charge-Storage Properties
32
33
34 in MnO2-Based Electrochemical Supercapacitors. ACS Appl. Mater. Interfaces 2009, 1, 11301139.
35
36
37 (39) Kanoh, H.; Tang, W.; Makita, Y.; Ooi, K. Electrochemical Intercalation of Alkali-Metal Ions
38
39 into Birnessite-Type Manganese Oxide in Aqueous Solution. Langmuir 1997, 13, 68456849.
40
41
42 (40) Zhao, W.; Wang, Q. Q.; Liu, F.; Qiu, G. H.; Tan, W. F.; Feng, X. H. Pb2+ Adsorption on
43
44
45 Birnessite Affected by Zn2+ and Mn2+ Pretreatments. J. Soils Sediments 2010, 10, 870878.
46
47 (41) Zhang, S. W.; Chen, G. Z. Manganese Oxide Based Materials for Supercapacitors. Energy
48
49
50 Mater. 2008, 3, 186200.
51
52
(42) Rasul, S.; Suzuki, S.; Yamaguchi, S.; Miyayama, M. High Capacity Positive Electrodes for
53
54
55 Secondary Mg-Ion Batteries. Electrochim. Acta 2012, 82, 243249.
56
57
58 (43) Lee, B.; Yoon, C. S.; Lee, H. R.; Chung, K. Y.; Cho, B. W.; Oh, S. H.
59 28
60
ACS Paragon Plus Environment
Page 29 of 39 ACS Applied Materials & Interfaces

1
2
3
Electrochemically-Induced Reversible Transition from the Tunneled to Layered Polymorphs of
4
5 Manganese Dioxide. Sci. Rep. 2014, 4, 60666074.
6
7
8 (44) Volkov, A. G.; Paula, S.; Deamer, D. W. Two Mechanisms of Permeation of Small Neutral
9
10
Molecules and Hydrated Ions Across Phospholipid Bilayers. Bioelectrochem. Bioenerg. 1997, 42,
11
12
13 153160.
14
15
16 (45) Pan, H.; Shao, Y.; Yan, P.; Cheng, Y.; Han, K. S.; Nie, Z.; Wang, C.; Yang, J.; Li, X.;
17
18 Bhattacharya, P.; Mueller, K. T.; Liu, J. Reversible Aqueous Zinc/Manganese Oxide Energy
19
20
21 Storage from Conversion Reactions. Nat. Energy 2016, 1, 16039.
22
23
24 (46) Ghaemi, M.; Ataherian, F.; Zolfaghari, A.; Jafari, S. M. Charge Storage Mechanism of
25
26 Sonochemically Prepared MnO2 as Supercapacitor Electrode: Effects of Physisorbed Water and
27
28
29 Proton Conduction. Electrochim. Acta 2008, 53, 46074614.
30
31
(47) Long, J. W.; Rhodes, C. P.; Young, A. L.; Rolison, D. R. Ultrathin, Protective Coatings of
32
33
34 Poly(o-phenylenediamine) as Electrochemical Proton Gates: Making Mesoporous MnO2
35
36
37 Nanoarchitectures Stable in Acid Electrolytes. Nano Lett. 2003, 3, 11551161.
38
39 (48) Pang, S. C.; Anderson, M. A.; Chapman, T. W. Novel Electrode Materials for Thin-Film
40
41
42 Ultracapacitors: Comparison of Electrochemical Properties of Sol-Gel-Derived and
43
44
45 Electrodeposited Manganese Dioxide. J. Electrochem. Soc. 2000, 147, 444450.
46
47
48
49
50
51
52
53
54
55
56
57
58
59 29
60
ACS Paragon Plus Environment
ACS Applied Materials & Interfaces Page 30 of 39

1
2
3 Figure Captions:
4
5 Figure 1. XRD patterns (a) and SEM images (b) of the as-prepared birnessite.
6
7
8 Figure 2. Cyclic capacitance (a) and cyclic voltammograms (b) for birnessite electrodes in
9
10
11 electrolyte with different initial concentrations of Cd2+.
12
13 Figure 3. Electrosorption capacities of birnessite electrodes for Cd2+ after 50 cycles of
14
15
16 charge-discharge tests in electrolyte with different initial concentrations of Cd2+ (a) and in 0.1 mol
17
18
19
L-1 Na2SO4 with Cd2+ concentrations of 200 and 2200 mg L-1 at different cycles in charge-discharge
20
21 tests (b).
22
23
24 Figure 4. XRD patterns of birnessite electrodes after 50 cycles of charge-discharge tests in
25
26 electrolytes with different initial concentrations of Cd2+: (a) 0 mg L-1, (b) 200 mg L-1, (c) 600 mg
27
28
29 L-1, (d) 1000 mg L-1, (e) 1400 mg L-1, (f) 1800 mg L-1, and (g) 2200 mg L-1.
30
31
32 Figure 5. XPS broad scans (a) and Cd3d spectra (b) of pristine birnessite and electrodes after 50
33
34 cycles of charge-discharge tests in 1000 mg L-1 Cd2+ electrolyte, and Mn2p3/2 spectra of pristine
35
36
37 birnessite (c) and electrodes (d) after 50 cycles of charge-discharge tests in 1000 mg L-1 Cd2+
38
39
40
electrolyte.
41
42 Figure 6. SEM images of birnessite electrodes after 50 cycles of charge-discharge tests with
43
44
45 different initial concentrations of Cd2+: (a) 0 mg L-1, (b) 200 mg L-1, (c) 600 mg L-1, and (d) 1000
46
47 mg L-1.
48
49
50 Figure 7. Electrosorption Cd2+ capacity of birnessite electrodes after 50 cycles of charge-discharge
51
52
53 tests in 200 mg L-1 Cd2+ solution with (a) different pH values and (b) masses of electrode materials.
54
55 Figure 8. Isotherm for Cd2+ adsorption and released Mn2+ at 25 oC and pH 6.0.
56
57
58
59 30
60
ACS Paragon Plus Environment
Page 31 of 39 ACS Applied Materials & Interfaces

1
2
3 Figures
4
5 a b
6
7
200 nm
8
9
10
11
12
13
14
15
1.0 m
16
17
18 Figure 1. XRD patterns (a) and SEM images (b) of the as-prepared birnessite.
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 31
60
ACS Paragon Plus Environment
ACS Applied Materials & Interfaces Page 32 of 39

1
2
3
4
5
6
a b
7
8
9
10
11
12
13
14
15
16
17
18
19 Figure 2. Cyclic capacitance (a) and cyclic voltammograms (b) for birnessite electrodes in
20
21 electrolyte with different initial concentrations of Cd2+.
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 32
60
ACS Paragon Plus Environment
Page 33 of 39 ACS Applied Materials & Interfaces

1
2
3
4
5 a b
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20 Figure 3. Electrosorption capacities of birnessite electrodes for Cd2+ after 50 cycles of
21
22 charge-discharge tests in electrolyte with different initial concentrations of Cd2+ (a) and in 0.1 mol
23
24
25 L-1 Na2SO4 with Cd2+ concentrations of 200 and 2200 mg L-1 at different cycles in charge-discharge
26
27
28
tests (b).
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 33
60
ACS Paragon Plus Environment
ACS Applied Materials & Interfaces Page 34 of 39

1
2
3
4
5
6
7
8
9
10 g
11 f
12 e
13
d
14
15 c
16 b
17 a
18
19
20
21
22 Figure 4. XRD patterns of birnessite electrodes after 50 cycles of charge-discharge tests in
23
24 electrolytes with different initial concentrations of Cd2+: (a) 0 mg L-1, (b) 200 mg L-1, (c) 600 mg
25
26 L-1, (d) 1000 mg L-1, (e) 1400 mg L-1, (f) 1800 mg L-1 and (g) 2200 mg L-1.
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 34
60
ACS Paragon Plus Environment
Page 35 of 39 ACS Applied Materials & Interfaces

1
2
3
4
5 a b
6
7
8
9
10
11
12
13
14
15
16
17
18
c d
19
20
21
22
23
24
25
26
27
28
29
30
31 Figure 5. XPS broad scans (a) and Cd3d spectra (b) of pristine birnessite and electrodes after 50
32
33 cycles of charge-discharge tests in 1000 mg L-1 Cd2+ electrolyte, and Mn2p3/2 spectra of pristine
34
35 birnessite (c) and electrodes (d) after 50 cycles of charge-discharge tests in 1000 mg L-1 Cd2+
36
37 electrolyte.
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59 35
60
ACS Paragon Plus Environment
ACS Applied Materials & Interfaces Page 36 of 39

1
2
3
4
5 a b
6
7
8 100 nm
9
10
11
12
13
14
15 1.0 m 1.0 m
16
17 c d
18
19
20 50 nm 50 nm
21
22
23
24
25
26
27 1.0 m 1.0 m
28
29
30 Figure 6. SEM images of birnessite electrodes after 50 cycles of charge-discharge tests with
31
32 different initial concentrations of Cd2+: (a) 0 mg L-1, (b) 200 mg L-1, (c) 600 mg L-1, and (d) 1000
33
34 mg L-1.
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 36
60
ACS Paragon Plus Environment
Page 37 of 39 ACS Applied Materials & Interfaces

1
2
3
4
5 a b
6
7
8
9
10
11
12
13
14
15
16
17
18 Figure 7. Electrosorption Cd2+ capacity of birnessite electrodes after 50 cycles of charge-discharge
19
20 tests in 200 mg L-1 Cd2+ solution with (a) different pH values and (b) masses of electrode materials.
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 37
60
ACS Paragon Plus Environment
ACS Applied Materials & Interfaces Page 38 of 39

1
2
3
4
5
6
7
8
9
10
11
12 1 1
13 Y= + R 2 =0.9544
14 125.8 1.264Ce
15
16
17
18
19
20
21
22 Figure 8. Isotherm for Cd2+ adsorption and released Mn2+ at 25 oC and pH 6.0.
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 38
60
ACS Paragon Plus Environment
Page 39 of 39 ACS Applied Materials & Interfaces

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

You might also like