Graphene Oxide ... - ACS Publications

Oct 1, 2018 - Zhenqian Cao , Hui Mao* , Xi Guo , Dayin Sun , Zhijia Sun , Baoxin Wang , Yu Zhang , and Xi-Ming Song*. Liaoning Key Laboratory for Gree...
0 downloads 0 Views 3MB Size
Subscriber access provided by UNIV OF NEWCASTLE

Article 2

Hierarchical Ni(OH)/polypyrrole/graphene oxide nanosheets as excellent electrocatalysts for the oxidation of urea Zhenqian Cao, Hui Mao, Xi Guo, Dayin Sun, Zhijia Sun, Baoxin Wang, Yu Zhang, and Xi-Ming Song ACS Sustainable Chem. Eng., Just Accepted Manuscript • DOI: 10.1021/ acssuschemeng.8b04027 • Publication Date (Web): 01 Oct 2018 Downloaded from http://pubs.acs.org on October 3, 2018

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 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 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.

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

ACS Sustainable Chemistry & Engineering

Hierarchical Ni(OH)2/polypyrrole/graphene oxide nanosheets as excellent electrocatalysts for the oxidation of urea Zhenqian Cao, Hui Mao*, Xi Guo, Dayin Sun, Zhijia Sun, Baoxin Wang, Yu Zhang and Xi-Ming Song* Liaoning Key Laboratory for Green Synthesis and Preparative Chemistry of Advanced Materials, College of Chemistry, Liaoning University, Chongshan Middle Road, No. 66, Shenyang 110036, China *

Author to whom any correspondence should be addressed.

Tel.: 86-24-62202378; Fax: 86-24-62202380; E-mail: [email protected] (H . M.); [email protected] (X.M.S)

ABSTRACT: Two kinds of Ni(OH)2 nanostructures, including thin nanosheets (NSs) with the thickness of 15-20 nm and very small nanoparticles (NPs) with the diameter of 2-3 nm, were immobilized on polypyrrole/graphene oxide (PPy/GO) via a facile strategy due to the coordination interaction between Ni2+ and -NH- segments in PPy chains. The obtained Ni(OH)2/PPy/GO presented excellent electrocatalytic activity for urea oxidation reaction (UOR) due to the excellent hierarchical nanostructures and the synergistic effects of lamellar GO, conductive PPy, electrocatalytically active Ni(OH)2. The electrocatalytic mechanism and the role of the each component in the Ni(OH)2/PPy/GO to UOR were also investigated in detail. PPy/GO with good conductivity and large surface area can effectively facilitate the electronic transmission of UOR. Super smaller Ni(OH)2 NPs are electrocatalytic active centers for UOR. Therefore, Ni(OH)2/PPy/GO can serve as a kind of excellent electrocatalysts for UOR in alkaline medium and reveal potential applications to urea-rich wastewater disposal, hydrogen production and direct urea fuel cells (DUFCs).

KEYWORD: nickel hydroxide; polypyrrole/graphene oxide (PPy/GO); urea; electrocatalytic oxidation.

ACS Paragon Plus Environment

1

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

Page 2 of 31

INTRODUCTION In the recent decades, energy crisis and environment pollution have attracted world-wide attention and driven researchers to explore and develop efficient, durable, sustainable and clean energy sources independent of fossil fuels.1, 2 Urea is a kind of widely used nitrogen-release fertilizer in the agricultural industry, but large-scale discharge of urea-rich wastewater which derives from urea synthesis industry and urine waste excreted from human/animal often pollutes the air, the ground and drinking water.3, 4 Recently, urea electrolysis, whose reaction equation is CO(NH2)2 + H2O → N2 + 3H2 + CO2, has attracted tremendous attentions, because it is a broad applied and effective technique approach for purifying urea-rich wastewater, simultaneously producing hydrogen at the cathode,5 which barely need sophisticated and bulky instruments.6 Therefore, direct urea fuel cells (DUFCs) offer a great promising application for energy-sustainable development and mitigating water contamination. The equation of anodic urea-oxidation reaction (UOR) is CO(NH2)2 + 6OH- → N2 + 5H2O + CO2 + 6e-,7, 8 which results in the require of catalysts based on noble metal for UOR, including Pt3 and Ru.9 However, the resource scarcity and the high price of these catalysts hinder the commercialization of the urea electrolysis. Up to now, considerable efforts have been attracted on developing low-cost alternatives with high catalytic performance to noble-metal catalysts.10 For instance, Ni can act as the superior active species for UOR in alkaline medium,3 but the practical application of pure Ni catalysts was hindered by their disadvantages, such as low electroactive sites and easy CO-poisoning behavior usually.2 Recently, inexpensive Ni-based catalysts, including Ni hydroxides/oxides10 or other compounds8 have been developed for UOR. Some researches suggest that NiOOH can be produced during electrocatalytic process of UOR, because that its electroactive sites and anti-poisoning property are highly corresponding to the structural/electronic effects of Ni-based catalysts.11, 12 Therefore, for improving the electroactivity and stability of Ni-based electrocatalysts, their physical/chemical structures would be tuned by preparing supported-type nanocatalysts, which are composed of effective catalytic active sites with fine particle/pore size and appropriate supporters for promoting electronic transmission of UOR.

ACS Paragon Plus Environment

2

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

ACS Sustainable Chemistry & Engineering

Since discovered in 1977,13 conducting polymers possessing high conductivity, redox properties and excellent environmental stability, have attracted many attentions for promising applications in various areas.14,

15

Conducting polymer micro/nanostructures can be selected as good catalyst supports for

preparing supported-type nanocatalysts, because transition metal compounds can be well distributed on their surface because of the coordination interaction between transition metal ions and the heteroatoms in conducting polymer chains. For instance, good electrocatalytic activities can be achieved by PPyNiOx composite film through electrosynthesis in the application of ethanol oxidation in alkaline media;16 a ternary composite composed of transition metal hydroxides, PPy and reduced graphene oxide exhibited excellent bifunctional electrocatalytic activities and promising electrochemical durability for the oxygen reduction and the oxygen evolution reactions in alkaline solution;17 a non-enzymatic glucose sensor can be constructed by the hierarchical nanostructure of PPy nanowires electrode modified with Ni(OH)2 nanoflakes.18 Therefore, the intrinsic electric conduction characteristics of conducting polymer micro/nanostructures19 may enhance the electrocatalytic performance of the obtained supported-type nanocatalysts. In this work, two kinds of Ni(OH)2 nanostructures, including thin nanosheets (NSs) with the thickness of 15-20 nm and very small nanoparticles (NPs) with the diameter of 2-3 nm, were successfully immobilized on polypyrrole/graphene oxide (PPy/GO) by the coordination interaction between Ni2+ and -NH- segments in PPy chains. Through this facile strategy, Ni(OH)2/PPy/GO nanosheets were synthesized. The excellent electrocatalytic activity for UOR was achieved by Ni(OH)2/PPy/GO modified glassy carbon electrode (Ni(OH)2/PPy/GO/GCE) due to the excellent hierarchical nanostructures and the synergistic effects of lamellar GO, conductive PPy, electrocatalytically active Ni(OH)2, especially the good conductivity of PPy can effectively facilitate the electronic transmission of UOR. The electrocatalytic mechanism and the role of the each component in the Ni(OH)2/PPy/GO to UOR were also investigated in detail, indicating that Ni(OH)2/PPy/GO can act as a kind of excellent electrocatalysts for UOR in alkaline media and reveal potential applications to urea-rich wastewater disposal, hydrogen production and direct urea fuel cells (DUFCs). ACS Paragon Plus Environment

3

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

Page 4 of 31

EXPERIMENTAL SECTION Preparation of Ni(OH)2/PPy/GO nanosheets 10 mg of PPy/GO (synthesized as the previous report20) and 14 mL deionized water were under ultrasound for 30 min. Ni(NO3)2 solution (14 mL, 5 mM) and NaSH solution (14 mL, 15 mM) were added into PPy/GO suspension, kept at 60 °C for 6 h. The cooled products were washed by water and ethanol for several times, drying in vacuum at 50 °C for 24 h. The synthetic process of Ni(OH)2/PPy/GO nanosheets is showed in Scheme 1. In addition, pure Ni(OH)2 and Ni(OH)2/GO nanosheets were also fabricated by the methods given in the Supporting Information, whose morphologies were also presented in Fig. S1.

Scheme 1. The synthetic process of Ni(OH)2/PPy/GO nanosheets.

Preparation of Ni(OH)2/PPy/GO/GCE ACS Paragon Plus Environment

4

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

ACS Sustainable Chemistry & Engineering

1 mg of Ni(OH)2/PPy/GO nanosheets were dispersed into 1 mL ethanol for the preparation of Ni(OH)2/PPy/GO suspension. 5 µL suspension were dropped on a GCE (φ = 3 mm) to form a layer and drying in the environment. The working electrode is the above modified GCE.

RESULTS AND DISCUSSION Characterizations of Ni(OH)2/PPy/GO

Fig. 1. XRD patterns of (a) GO, (b) PPy/GO, (c) Ni(OH)2 and (d) Ni(OH)2/PPy/GO.

XRD patterns of GO, PPy/GO, Ni(OH)2 and Ni(OH)2/PPy/GO are shows in Fig. 1. The characteristic peak of GO in Fig. 1(a) was clearly found at 2θ = 11.6° with the (001) interlayer spacing of 0.78 nm, which had been attributed to the existence of hydroxyl, epoxy and carboxyl groups.21 A small characteristic peak at 2θ = 42.4° was obtained from the (001) plane of graphite phases.22 A broad peak center at 2θ = 25° was only observed in the XRD pattern of PPy/GO in Fig. 1(b), which belonged to the ACS Paragon Plus Environment

5

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

Page 6 of 31

characteristic diffraction peak of amorphous PPy.22, 23 The result means that GO nanosheets were coated by PPy nanosheets with electrostatic interactions and π− π stacking between them. The homogeneity and crystallinity of Ni(OH)2 were determined by XRD in Fig. 1(c), and all the diffraction peaks could be matched well with hexagonal β-Ni(OH)2 (JCPDS file number 14-0117, signal



).24 Both the

characteristic peaks which 2θ = 25° of PPy/GO and the characteristic peaks of hexagonal β-Ni(OH)2 could be found in Fig. 1(d), which testified the existence of PPy/GO and Ni(OH)2.

Fig. 2. SEM images of (a) PPy/GO, (b) Ni(OH)2/PPy/GO, TEM image of (c) Ni(OH)2/PPy/GO and HRTEM of (d) Ni(OH)2/PPy/GO.

ACS Paragon Plus Environment

6

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

ACS Sustainable Chemistry & Engineering

SEM and TEM were used for the investigation of the morphology of Ni(OH)2/PPy/GO nanosheets in Fig. 2. Compared to PPy/GO in Fig. 2(a), many smaller NSs and NPs can be clearly observed on Ni(OH)2/PPy/GO in Fig. 2(b), indicating that Ni(OH)2 nanostructures were successfully immobilized on PPy/GO because of the coordination interaction between Ni2+ and -NH- segments in PPy chains. The typical TEM image of Ni(OH)2/PPy/GO nanosheets is presented in Fig. 2(c). It is clearly found that inorganic lamellar structures with the thickness of 15-20 nm and very small NPs with the diameter of 23 nm existed on PPy/GO. A high-resolution TEM (HRTEM) image of Ni(OH)2/PPy/GO clearly presents the lattice fringes corresponded to the highly crystalline nature of Ni(OH)2 in Fig. 2(d). The lattice spacing is measured to 0.231 nm, matching well with (101) interplaner spacing of β-Ni(OH)2 (JCPDS file number 14-0117), which well coincided with Fig. 1(d). In addition, from the XRD data, it is worth to note that compared to the peak at 2θ = 33.2°, the peak at 2θ = 38.4° was obviously widened. According to Scherrer formula,25 the average sizes estimated from the diffraction peaks of (100) and (101) plane were calculated to 18.5 nm and 2.5 nm, respectively, consisting with the average thickness of Ni(OH)2 NSs and the average diameter of Ni(OH)2 NPs observed from TEM image of Ni(OH)2/PPy/GO in Fig. 2(c).

ACS Paragon Plus Environment

7

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

Page 8 of 31

Fig. 3. FTIR spectra of (a) GO, (b) PPy/GO, (c) Ni(OH)2 and (d) Ni(OH)2/PPy/GO.

FTIR spectra are shown in Fig. 3 to confirm the chemical structures of Ni(OH)2/PPy/GO. The characteristic peaks of all samples are identified and assigned are presented in Table 1, confirming that the final product is composed of Ni(OH)2, PPy and GO.

Table 1. vibration modes and band frequencies in GO, PPy/GO, Ni(OH)2 and Ni(OH)2/PPy/GO.

Species

Peak (cm-1)

GO

3421

-OH stretching vibrations

3421

C-H stretching vibrations

Functional group and vibration

References 29, 30, 31, 32

ACS Paragon Plus Environment

8

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

ACS Sustainable Chemistry & Engineering

PPy/GO

1722

carboxyl C=O stretching vibrations

1623

aromatic C=C skeletal vibrations

1046

alkoxy C-O deformation vibrations

1542

C-C antisymmetric vibration in PPy rings

26

1288

=C-N in-plane stretching vibration in PPy rings

27

1175, 789

Ni(OH)2

C-C stretching vibrations

35, 36, 37

900

the doping state of PPy

3642

O-H vibration of the free non-hydrogen bonded hydroxyl group

28

1637

O-H rocking vibration

29

523, 475

the deformation mode of Ni-OH stretching vibrations modes

40, 41

The EDS spectra of Ni(OH)2 and Ni(OH)2/PPy/GO coated on a silicon pellet are showed in Fig. S2, which can further certify the chemical composition of Ni(OH)2/PPy/GO. Only the peaks of Ni and O element (Si pellet as substrate) can be found in Fig. S2(i). The peaks of C and N can be also clearly found in Fig. S2(ii), which further testify that Ni(OH)2 have been immobilized on PPy/GO. However, an additional peak of S element appears in Fig. 4(ii), which may be due to the residual S from the postprocessing.

ACS Paragon Plus Environment

9

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

Page 10 of 31

Fig. 4. XPS spectra of Ni(OH)2/PPy/GO: (i) survey spectra; (ii) C 1s; (iii) N 1s; (iv) O 1s; (v) Ni 2p.

ACS Paragon Plus Environment

10

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

ACS Sustainable Chemistry & Engineering

XPS analysis was performed to further determine the chemical states of Ni(OH)2/PPy/GO, which more clearly demonstrated the evidence for the coordination between Ni2+ and –NH- group. Fig.4(i) shows the characteristic signature for C, O, N and Ni element in the survey scan spectrum of Ni(OH)2/PPy/GO and their core-line spectra are presented in Fig.4(ii)-(v), respectively, and Table 2 displays the corresponding bonding energies and attribution, which can well confirm the existence of Ni(OH)2, PPy and GO in product, as well as the coordination between Ni2+ and –NH- group.

Table 2. binding energies and assignment for the fit of C 1s, N 1s, O 1s and Ni 2p spectra of Ni(OH)2/PPy/GO

Attribution

Element

B. E. (eV)

C 1s

283.6

sp2-hybridized carbon

30

284.8

C-C/C-H,

31

286.2

C-N+ of Py rings

32

287.7

C=N+ of Py rings

32

289.0

O-C=O of Py rings

33, 34

397.9

C=N defects of PPy and the coordination of N-Ni

35, 36

399.1

-NH- group of pyrrole unit

37

400.0

-N+H- group of pyrrole unit

35

530.9

the bound hydroxide groups (OH-)

38

532.9

the oxygen of the carboxylate (O-C=O) in PPy/GO

39

N 1s

O 1s

References

ACS Paragon Plus Environment

11

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

Ni 2p

Page 12 of 31

40

855.8

Ni 2p3/2

873.4

Ni 2p1/2

861.4

the shake-up satellite

879.4

the shake-up satellite

866.2

the second shake-up satellite of Ni 2p3/2

41

42

Electrochemical behavior of Ni(OH)2/PPy/GO/GCE

Fig. 5. Linear sweep voltammograms of the six modified GCEs a-f in 1 M KOH with scanning rate at 10 mV/s. ACS Paragon Plus Environment

12

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

ACS Sustainable Chemistry & Engineering

Fig. 5 presents that the electrochemical behavior of Ni(OH)2/PPy/GO/GCE was investigated by LSV, compared with the other five modified GCEs. Only Ni(OH)2/PPy/GO/GCE exhibits an excellent anodic peak at 0.5 V in Fig. 5(f), implying the higher electrochemical activity because of the oxidation of Ni(OH)2 to the nickel oxyhydroxide (NiOOH).43

Fig. 6. (i) Cyclic voltammograms of Ni(OH)2/PPy/GO/GCE at different scanning rates in 1 M KOH; (ii) plots of jpa vs. scanning rates from (i); (iii) Laviron's plots of (a) anodic and (b) cathodic peak potential vs. ln ν from (i).

ACS Paragon Plus Environment

13

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

Page 14 of 31

Fig. 6(i) present the influence of scanning rates evaluated using CV. A pair of enhanced redox peaks in 1 M KOH solution can be observed, corresponding to Ni(II)/Ni(III) redox couple.44 The oxidation peak current density (jpa) of Ni(II)/Ni(III) redox couple at Ni(OH)2/PPy/GO/GCE increased continuously with the scanning rates. jpa value presented an excellent linear relationship against ν0.5 in the range from 10 to 100 mV/s (R2 = 0.9995) in Fig. 6(ii), indicating that the electrochemical characteristic of the diffusion process of OH- from the solution to Ni(OH)2/PPy/GO/GCE was a typical diffusion-controlled process.45 In addition, by the increasing of scanning rates, the oxidation peak potential (Epa) on Ni(OH)2/PPy/GO/GCE moved in the direction of the positive potential, and the corresponding reduction peak potential (Epc) moved in the direction of negative potential, resulting in the gradual increasing of the difference of anodic peak-to-cathodic peak position (∆Ep). Fig. 6(iii) shows that both of Epa and Epc depended linearly on ln ν (10 to 100 mV/s) with the equations of Epa (V) = 0.5819 + 0.0189 ln ν (V/s) (R2 = 0.9595) and Epc (V) = 0.333 - 0.0069 ln ν (V/s) (R2 = 0.9258), respectively. Hence, according to the Laviron theory for thin-layer quasi-reversible electrochemical process,46 the electrochemical parameters of the electron transfer coefficient α and the apparent charge transfer rate constant ks can be defined by the following equations: Epa = E o +

RT ln ν (1− α )nF

Eq. (1)

Epc = E o −

RT ln ν αnF

Eq. (2)

And ln ks = α ln (1 − α ) + (1 − α ) ln α − ln

RT α (1 − α )nF∆Fp − nFν RT

Eq. (3)

Where n is the number of transferred species, ν is scanning rate and other symbols have their usual meanings. The electron transfer coefficient α was calculated to 0.733 according to Eq. (1) and Eq. (2), and the apparent charge transfer rates constant ks was calculated to 2.21 s−1 according to Eq. (3). Otherwise, according to the previous reports,47,

48

the electroactive surface area (ESA) of the

Ni(OH)2/PPy/GO/GCE could be also estimated by the following equation: ACS Paragon Plus Environment

14

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

ACS Sustainable Chemistry & Engineering

ESA =

Q mq

Eq. (4)

Where Q is the charge for hydrogen desorption (mC/cm2), which could be calculated from CV in Fig. 6(i) of 10 mV/s, m is the quantity of nickel element and q is 246 µC/cm2,49 meanwhile, the proportion of Ni element in Ni(OH)2/PPy/GO was 24.16% measured by ICP measurement, so that the ESA of Ni(OH)2/PPy/GO was calculated to 116 cm2/mg, which was much higher than that of Ni(OH)2 nanoribbon (2.1 cm2/mg),48 Ni-Zn catalysts (67.9 cm2/mg),50 nickel nanowire (79.1 cm2/mg),51 etc., presented in Table 3.

Table 3. ESA for different Ni based electrocatalysts estimated from the CV curves presented in some published papers.

Materials

ESA (cm2/mg)

References

Ni(OH)2 nanoribbon

2.1

48

Ni-Zn catalysts

67.9

50

Nickel nanowire

79.1

51

Ni(OH)2/PPy/GO

116

This work

Electrocatalytic oxidation of urea at Ni(OH)2/PPy/GO/GCE

ACS Paragon Plus Environment

15

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

Page 16 of 31

Fig. 7. (i) Linear sweep voltammograms of Ni(OH)2/PPy/GO/GCE in 1 M KOH (a) in the absence and (b) in the presence of 0.5 M urea with scanning rate at 10 mV/s; (ii) Chronoamperograms of Ni(OH)2/PPy/GO/GCE in 1 M KOH solution in the absence (a) and presence (b) of 0.5 M urea. Applied potential was 0.5 V vs. Hg/HgO.

ACS Paragon Plus Environment

16

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

ACS Sustainable Chemistry & Engineering

Scheme 2. The oxidation process of urea on Ni(OH)2/PPy/GO/GCE in 1 M KOH.

The electrocatalytic activity of Ni(OH)2/PPy/GO/GCE for UOR were investigated by LSV. Ni(OH)2/PPy/GO/GCE appeared a well-defined oxidation peak derived from Ni(OH)2 to NiOOH at 0.5 V in Fig. 7(i)-a. When 0.5 M urea existed, a much stronger oxidation peak appeared at 0.60 V in Fig. 7(i)-b, which was 100 mV more positive than the potential of Ni(II)/Ni(III) conversion at Ni(OH)2/PPy/GO/GCE in the absence of urea. It indicated that the electrooxidation of urea occurred after Ni(II) was oxidized to Ni(III), where Ni(III) was used as an electrocatalytic center for UOR.52 Fig. 7(ii) displays the chronoamperograms of Ni(OH)2/PPy/GO/GCE in the absent and present of urea with the continuous and constant current density. The higher current density can be obtained in the present of urea, revealing that Ni(OH)2/PPy/GO can serve as an excellent and stable electrocatalyst for UOR in alkaline solution. Therefore, the oxidation process of urea on Ni(OH)2/PPy/GO/GCE can be described in ACS Paragon Plus Environment

17

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

Page 18 of 31

Scheme 2, and the potential mechanism for UOR process can be explained as follows: (i) the oxidation process of Ni(II)/Ni(III), the oxidation of Ni(OH)2 to NiOOH occurred fast on Ni(OH)2/PPy/GO/GCE by the power source, as described in Eq. (5); (ii) electrooxidation of urea, urea was oxidized to the product by Ni(III) and this process was much slower than that of Ni(II)/Ni(III) on Ni(OH)2/PPy/GO/GCE,

where Ni(III) was simultaneously reduced to Ni(II),2, 65, 66 as shown in Eq. (6); (iii) total

reaction for UOR, urea was electrooxidation to N2 and CO32- and H2O in alkaline solution, as summarized in Eq. (7).

Fig. 8. Linear sweep voltammograms of the six modified GCEs a-f in 1 M KOH containing 0.5 M urea with scanning rate at 10 mV/s.

The catalytic performance of Ni(OH)2/PPy/GO for UOR was markedly higher than that of the above other five materials modified GCE in Fig. 8. No peaks can be found in Fig. 8(a) obtained by the bare ACS Paragon Plus Environment

18

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

ACS Sustainable Chemistry & Engineering

GCE. It can be obviously found that GO/GCE in Fig. 8(b) showed a weaker anodic peak for UOR than Ni(OH)2/GO/GCE in Fig. 8(e), which well demonstrated that Ni(OH)2 can serve as electrocatalytic active centers for UOR. However, Ni(OH)2/GCE exhibited poor electrocatalytic activity in Fig. 8(d) because of the bad electroconductivity of Ni(OH)2, which impeded the electron transport on the modified electrode. Though GO/GCE (Fig. 8(b)) exhibited higher electrocatalytic activity than PPy/GO/GCE (Fig. 8(c)) because of the existence of abundant oxygen-containing functional groups of GO, the excellent electrocatalytic activity for UOR was still achieved by Ni(OH)2/PPy/GO/GCE (Fig. 8(f)) compared to Ni(OH)2/GO/GCE (Fig. 8(e)), which may be associated with more Ni(OH)2, especially super smaller Ni(OH)2 NPs immobilized on PPy/GO by the coordination interaction between Ni2+ and -NH- segments in PPy chains and the good conductivity of PPy which can effectively facilitate the electronic transmission of UOR. Based on the above discussion, in our opinion, the role of the each component in the Ni(OH)2/PPy/GO to UOR may be explained as follows: lamellar GO provided large surface area; conductive PPy can not only effectively facilitate the electronic transmission of UOR, but also contribute to the immobilization of Ni(OH)2 nanostructures on PPy/GO by the coordination interaction between Ni2+ and -NH- segments in PPy chains, which result in the excellent hierarchical nanostructures of the obtained Ni(OH)2/PPy/GO nanosheets; Ni(OH)2 nanostructures, especially super smaller Ni(OH)2 NPs can serve as electrocatalytic active centers for UOR. Meanwhile, the excellent hierarchical structures between Ni(OH)2 nanostructures (including thin NSs and small NPs) and PPy/GO nanosheets can further effectively improve the electrocatalytic performance of electrocatalysts. Therefore, due to the excellent hierarchical nanostructures and the synergistic effects of lamellar GO, conductive PPy, electrocatalytically active Ni(OH)2, the excellent electrocatalytic activity for UOR was achieved by Ni(OH)2/PPy/GO/GCE.

ACS Paragon Plus Environment

19

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

Page 20 of 31

Fig. 9. (i) Linear sweep voltammograms of Ni(OH)2/PPy/GO/GCE at different scanning rates in 1 M KOH with 0.5 M urea; (ii) plots of potential vs. log ν from (i); (iii) plots of jpa for urea vs. scanning rates from (i) and (iv) Tafel plots of Ni(OH)2/PPy/GO in 1 M KOH with 0.5 M urea.

Fig. 9(i) presents the effect of scanning rates on UOR at Ni(OH)2/PPy/GO/GCE evaluated by LSV. Well-defined oxidation peaks of urea can be obtained at different scanning rates, and jpa increased proportionally. The plots of jpa for urea against the scanning rate shows an excellent linear relationship in Fig. 9(ii), whose linear regression equation is IUOR (mA/cm2) = 0.0686 + 0.4427 ν0.5 (R2 = 0.9976), demonstrating that UOR at Ni(OH)2/PPy/GO/GCE was a typical diffusion-controlled process.67 ACS Paragon Plus Environment

20

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

ACS Sustainable Chemistry & Engineering

Simultaneously, Epa depended linearly on log ν in Fig. 9(iii), which demonstrated a kinetic limitations on UOR.53 The diffusion coefficient of urea (Durea) can be calculated by Eq. (8): Ipa = 2.99 × 105 n[(1 − αUOR )n 0] ACDurea 0.5ν 0.5 0 .5

Eq. (8)

The symbols of above formula are consist with those in the previous report.1 According to the previous report,53 the value of αUOR can be calculated based on Fig. 9(iii) using the linear dependency of Epa with log ν by Eq. (9): Epa = k +

0.03 log ν αUORn 0

Eq. (9)

where k is a constant. Hence, αUOR was calculated to 0.946 by the slope of Fig. 9.(iii) and after substituted into Eq. (e), Durea of 9.2 × 10-10 cm2/s was calculated from Fig. 9(ii) for the electrochemical UOR on Ni(OH)2/PPy/GO/GCE. Fig. 9(iv) depicted the Tafel slope of Ni(OH)2/PPy/GO/GCE for UOR in the kinetics range (0.47-0.50 V) was calculated to 30.9 mV/dec, which was lower than that of rNiMoO4/nickel foam (32 mV/dec),54 mesoporous nickel phosphide nanocatalysts (52 mV/dec),12 smallsized MnO2 nanocrystals (75 mV/dec),55 etc., presented in Table 4.

Table 4. Tafel slopes for various Ni based electrocatalysts estimated from the LSV curves presented in some published papers.

Materials

Scanning rate (mV/s)

Estimated Tafel slopes (mV/dec)

r-NiMoO4/NF a

5

32

54

Mesoporous nickel phosphide nanocrystals

1

52

12

Small-sized MnO2 nanocrystals

-

75

55

Ni(OH)2/NiOOH b

2

44

56

References

ACS Paragon Plus Environment

21

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

Page 22 of 31

Ni8Mo1/G c

1

110

2

Ni(OH)2/PPy/GO

5

30.9

This work

a

Oxygen-vacancies rich NiMoO4 grown on Ni foam substrates

b

Electrodeposition of Ni(OH)2/NiOOH catalyst on GCE

c

Nickel-molybdenum/graphene nanocatalysts

ACS Paragon Plus Environment

22

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

ACS Sustainable Chemistry & Engineering

Fig. 10. (i) Linear sweep voltammograms of Ni(OH)2/PPy/GO/GCE at a scanning rate of 5 mV/s in different [KOH] with 0.5 M urea; (ii) the corresponding variation in jpa of UOR with different [KOH] with 0.5 M urea (data taken from (i)); (iii) Tafel plots at various [KOH] (0.5-4 M) in 0.5 M urea solution at a scanning rate of 5 mV/s; (iv) double logarithmic plot of j as a function of [KOH] at constant electrode potentials: (a) 450, (b) 455, (c) 460 V, conditions as in Fig. 10(iii).

In addition, the effects of the concentration of KOH and urea ([KOH]) and [urea]) on UOR at Ni(OH)2/PPy/GO/GCE

were

also

investigated.

Fig.10(i)

presents

the

LSV

curves

by

Ni(OH)2/PPy/GO/GCE in different [KOH] and the corresponding variation of jpa of UOR is shown in Fig. 10(ii), where an improvement in jpa with increasing OH- concentration can be observed. Along with the increasing of [KOH] in the range of 0.01-3 M, jpa of UOR gradually enhanced, but when [KOH] was higher than 3 M, jpa of UOR remained almost constant. jpa was linearly depended on [KOH] in the range of 0.1-2 M and deviated from linearity up to 2 M KOH, exhibiting a diffusion limitation of the OH- to the electrocatalyst surface on GCE. It may be mainly attributed to the complete coverage of the electrocatalyst surface with OH-, which resulted in impeding the mutual contact between urea molecules and Ni(OH)2/PPy/GO on GCE for further oxidation.1 Fig. 10(iii) presents Tafel plots at various [KOH] (0.5-4 M), revealing that the onset potential can be improved by the increasing OH- concentration.1 Otherwise, the rate of UOR also increased along with the increasing of OH- concentration in the potential region of 0.35-0.5 V, indicating that the high OH- concentration enhanced the rate of UOR.57 Fig. 10(iv) displays the linear relationship between log | j | and log [KOH] at constant electrode potential. The slope of all tested potentials is approximate to 2, suggesting the reaction order is 2 related to OH- ions.1

ACS Paragon Plus Environment

23

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

Page 24 of 31

Fig. 11. (i) Linear sweep voltammograms of Ni(OH)2/PPy/GO/GCE at a scanning rate of 5 mV/s in 1 M KOH with different [urea]; (ii) the change in anodic peak current density at various [urea] from (i).

Fig. 11 presents the effect of [urea] on UOR at Ni(OH)2/PPy/GO/GCE, which was investigated by LSV. jpa of UOR gradually enhanced with the increasing of [urea] in the range of 0.1-1.5 M. It can be found that jpa deviated from linearity at higher [urea] and Epa related to UOR increased with the increasing of [urea] over the range of 0.1-1.5 M, which may be due to the competition between the surface coverage of urea molecules and the adsorbed OH- coverage on the electrocatalyst surface at GCE, consist with the previous report1.

Electrochemical impedance analysis

ACS Paragon Plus Environment

24

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

ACS Sustainable Chemistry & Engineering

Fig. 12. EIS plots of (a) GO/GCE, (b) PPy/GO/GCE, (c) Ni(OH)2/GCE and (d) Ni(OH)2/PPy/GO/GCE in 0.1 M KCl solution containing 2.5 mM K3Fe(CN)6 and 2.5 mM K4Fe(CN)6.

The Nyquist diagrams of GO/GCE, PPy/GO/GCE, Ni(OH)2/GCE and Ni(OH)2/PPy/GO/GCE measured by EIS. Based on the literature,58 a semicircle portion in EIS plots denoted an electro-transfer resistance (Rct), delegating the electron transfer kinetics of the redox probe at the electrode interface. The Niquist diagrams of GO/GCE and PPy/GO/GCE were consistent with previous report.59 The slope of PPy/GO/GCE in Fig. 12(b) showed more dramatic than that of GO/GCE (Fig. 12(a)), indicating the higher electron conduction pathways. The Nyqust diagram of Ni(OH)2/GCE in Fig. 12(c) presents the largest diameter of the impedance arc among all the modified GCEs, which demonstrated Ni(OH)2/GCE possessed the highest interfacial Rct. It is clearly found that the diameter of the impedance arc obtained by Ni(OH)2/PPy/GO/GCE in Fig. 12(d) was distinctly smaller than Ni(OH)2/GCE, implying the lowest Rct. The Rct of the GO/GCE, PPy/GO/GCE, Ni(OH)2/GCE and Ni(OH)2/PPy/GO/GCE is 1185, 831, ACS Paragon Plus Environment

25

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

Page 26 of 31

4597 and 1989 Ω, respectively. It indicated that the excellent hierarchical Ni(OH)2/PPy/GO and good conductive PPy/GO were conducive to promote the electrons transfer of Ni(II)/Ni(III) redox conversion on the electrode interface, which further demonstrated the excellent electrocatalytic activity of Ni(OH)2/PPy/GO/GCE for UOR.

CONCLUSIONS In general, novel hierarchical Ni(OH)2/PPy/GO nanosheets were prepared via a facile strategy. The Ni(OH)2/PPy/GO/GCE presented the excellent electrocatalytically activity for UO and Tafel slope was calculated to 30.9 mV/dec, because of the excellent hierarchical nanostructures and the synergistic effects of the lamellar GO, conductive PPy, electrocatalytically active Ni(OH)2. The electrocatalytic mechanism of UOR on Ni(OH)2/PPy/GO/GCE have been investigated in detail, where urea were oxidized by Ni(III) on the electrode surface and PPy/GO with good conductivity and large surface area could effectively facilitate the electronic transmission of UOR. Therefore, the Ni(OH)2/PPy/GO nanosheets could serve as good electrocatalysts for UOR in alkaline medium and reveal promising applications to urea-rich wastewater disposal, hydrogen production and DUFCs.

Supporting Information SEM images of (a) Ni(OH)2 and (b) Ni(OH)2/GO EDS spectra of (i) Ni(OH)2 and (ii) Ni(OH)2/PPy/GO

Acknowledgements The financial supports from the National Natural Science Foundation of China (No. 51203072, 51773085 and 21203082) are greatly appreciated.

ACS Paragon Plus Environment

26

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

ACS Sustainable Chemistry & Engineering

References 1. Vedharathinam, V.; Botte, G. G., Understanding the electro-catalytic oxidation mechanism of urea on nickel electrodes in alkaline medium. Electrochimica Acta 2012, 81 (11), 292-300, DOI 10.1016/j.electacta.2012.07.007. 2. Shi, W.; Ding, R.; Li, X.; Xu, Q.; Liu, E., Enhanced performance and electrocatalytic kinetics of Ni-Mo/graphene nanocatalysts towards alkaline urea oxidation reaction. Electrochimica Acta 2017, 242, 247-259, DOI 10.1016/j.electacta.2017.05.002. 3. Boggs, B. K.; King, R. L.; Botte, G. G., Urea electrolysis: direct hydrogen production from urine. Chemical Communications 2009, 32 (32), 4859-4861, DOI 10.1039/B905974A. 4. Wang, D.; Yan, W.; Botte, G. G., Exfoliated nickel hydroxide nanosheets for urea electrolysis. Electrochemistry Communications 2011, 13 (10), 1135-1138, DOI 10.1016/j.elecom.2011.07.016. 5. Simka, W.; Piotrowski, J.; Robak, A.; Nawrat, G., Electrochemical treatment of aqueous solutions containing urea. Journal of Applied Electrochemistry 2009, 39 (7), 1137-1143, DOI 10.1007/s10800-008-9771-4. 6. Simka, W.; Piotrowski, J.; Nawrat, G., Influence of anode material on electrochemical decomposition of urea. Electrochimica Acta 2007, 52 (18), 5696-5703, DOI 10.1016/j.electacta.2006.12.017. 7. Ji, R. Y.; Chan, D. S.; Jow, J. J.; Wu, M. S., Formation of open-ended nickel hydroxide nanotubes on three-dimensional nickel framework for enhanced urea electrolysis. Electrochemistry Communications 2013, 29 (10), 21-24, DOI 10.1016/j.elecom.2013.01.006. 8. Liang, Y.; Liu, Q.; Asiri, A. M.; Sun, X., Enhanced electrooxidation of urea using NiMoO4 ·xH2O nanosheet arrays on Ni foam as anode. Electrochimica Acta 2015, 153, 456-460, DOI 10.1016/j.electacta.2014.11.193. 9. Miller, A. T.; Hassler, B. L.; Botte, G. G., Rhodium electrodeposition on nickel electrodes used for urea electrolysis. Journal of Applied Electrochemistry 2012, 42 (11), 925-934, DOI 10.1007/s10800012-0478-1. 10. Xu, W.; Wu, Z.; Tao, S., Urea‐ Based Fuel Cells and Electrocatalysts for Urea Oxidation. Energy Technology 2016, 4 (11), 1329-1337, DOI 10.1002/ente.201600185 11. Wang, L.; Li, M.; Huang, Z.; Li, Y.; Qi, S.; Yi, C.; Yang, B., Ni–WC/C nanocluster catalysts for urea electrooxidation. Journal of Power Sources 2014, 264, 282-289, DOI 10.1016/j.jpowsour.2014.04.104. 12. Ding, R.; Li, X.; Shi, W.; Xu, Q.; Wang, L.; Jiang, H.; Yang, Z.; Liu, E., Mesoporous Ni-P nanocatalysts for alkaline urea electrooxidation. Electrochimica Acta 2016, 222, 455-462, DOI 10.1016/j.electacta.2016.10.198. 13. Shirakawa, H.; Louis, E. J.; Macdiarmid, A. G.; Chiang, C. K.; Heeger, A. J., Synthesis of electrically conducting organic polymers: Halogen derivatives of polyacetylene, (CH)x. Journal of the Chemical Society, Chemical Communications 1977, 16 (16), 578-580, DOI 10.1039/C39770000578. 14. Xiao, R.; Cho, S. I.; Liu, R.; Lee, S. B., Controlled electrochemical synthesis of conductive polymer nanotube structures. Journal of the American Chemical Society 2007, 129 (14), 4483-4489, DOI 10.1021/ja068924v. 15. Abidian, M. R.; Kim, D. H.; Martin, D. C., Conducting-Polymer Nanotubes for Controlled Drug Release. Advanced Materials 2010, 18 (4), 405-409, DOI 10.1002/adma.200501726 16. Ehsani, A.; Mahjani, M. G.; Jafarian, M.; Naeemy, A., Electrosynthesis of polypyrrole composite film and electrocatalytic oxidation of ethanol. Electrochimica Acta 2012, 71 (3), 128-133, DOI 10.1016/j.electacta.2012.03.107. 17. Jia, X.; Gao, S.; Liu, T.; Li, D.; Tang, P.; Feng, Y., Fabrication and Bifunctional Electrocatalytic Performance of Ternary CoNiMn Layered Double Hydroxides/Polypyrrole/Reduced Graphene Oxide ACS Paragon Plus Environment

27

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

Page 28 of 31

Composite for Oxygen Reduction and Evolution Reactions. Electrochimica Acta 2017, DOI 10.1016/j.electacta.2017.05.120. 18. Yang, J.; Cho, M.; Pang, C.; Lee, Y., Highly sensitive non-enzymatic glucose sensor based on over-oxidized polypyrrole nanowires modified with Ni(OH)2 nanoflakes. Sensors & Actuators B Chemical 2015, 211, 93-101, DOI 10.1016/j.snb.2015.01.045. 19. Chang, H. H.; Chang, C. K.; Tsai, Y. C.; Liao, C. S., Electrochemically synthesized graphene/polypyrrole composites and their use in supercapacitor. Carbon 2012, 50 (6), 2331-2336, DOI 10.1016/j.carbon.2012.01.056. 20. Wang, P.; Zheng, Y.; Li, B., Preparation and electrochemical properties of polypyrrole/graphite oxide composites with various feed ratios of pyrrole to graphite oxide. Synthetic Metals 2013, 166 (1), 33-39, DOI 10.1016/j.synthmet.2013.01.002. 21. Liu, J.; Bai, H.; Wang, Y.; Liu, Z.; Zhang, X.; Sun, D. D., Self-Assembling TiO2 Nanorods on Large Graphene Oxide Sheets at a Two-Phase Interface and Their Anti-Recombination in Photocatalytic Applications. Advanced Functional Materials 2010, 20 (23), 4175–4181, DOI 10.1002/adfm.201001391. 22. Shayeh, J. S.; Ehsani, A.; Ganjali, M. R.; Norouzi, P.; Jaleh, B., Conductive polymer/reduced graphene oxide/Au nano particles as efficient composite materials in electrochemical supercapacitors. Appl. Surf. Sci. 2015, 353, 594-599, DOI 10.1016/j.apsusc.2015.06.066. 23. Fan, L. Q.; Liu, G. J.; Wu, J. H.; Liu, L.; Lin, J. M.; Wei, Y. L., Asymmetric supercapacitor based on graphene oxide/polypyrrole composite and activated carbon electrodes. Electrochim. Acta 2014, 137 (8), 26-33, DOI 10.1016/j.electacta.2014.05.137. 24. Cai, F. S.; Zhang, G. Y.; Chen, J.; Gou, X. L.; Liu, H. K.; Dou, S. X., Ni(OH)2 tubes with mesoscale dimensions as positive-electrode materials of alkaline rechargeable batteries. Angewandte Chemie 2004, 116 (32), 4308-4312, DOI 10.1002/ange.200460053. 25. Wang, J.; Teschner, D.; Yao, Y.; Huang, X.; Willinger, M. G.; Shao, L. D.; Schlogl, R., Fabrication of Nanoscale Ni/NiO Heterostructures as Electrocatalyst for Efficient Methanol Oxidation. Journal of Materials Chemistry A 2017, 5 (20), 9946-9951, DOI 10.1039/C7TA01982C. 26. Vivekchand, S. R. C.; Kam, K. C.; Gundiah, G.; Govindaraj, A.; Cheetham, A. K.; Rao, C. N. R., Electrical properties of inorganic nanowire-polymer composites. Journal of Materials Chemistry 2005, 15 (46), 4922-4927, DOI 10.1039/B511429B. 27. Costa, M. B. G.; Juárez, J. M.; Martínez, M. L.; Beltramone, A. R.; Cussa, J.; Anunziata, O. A., Synthesis and characterization of conducting polypyrrole/SBA-3 and polypyrrole/Na–AlSBA-3 composites. Materials Research Bulletin 2013, 48 (2), 661-667, DOI 10.1016/j.materresbull.2012.11.030. 28. Zhan, F.; Geng, B.; Guo, Y.; Wang, L., One-step synthesis of hierarchical carnation-like NiS superstructures via a surfactant-free aqueous solution route. Journal of Alloys & Compounds 2009, 482 (1–2), L1-L5, DOI 10.1016/j.jallcom.2009.03.155. 29. Patil, A. M.; Lokhande, A. C.; Shinde, P. A.; Kim, J. H.; Lokhande, C. D., Vertically aligned NiS nano-flakes derived from hydrothermally prepared Ni(OH)2 for high performance supercapacitor. Journal of Energy Chemistry 2018, 27 (3), 791-800, DOI 10.1016/j.jechem.2017.05.005. 30. Oprea, R.; Peteu, S. F.; Subramanian, P.; Qi, W.; Pichonat, E.; Happy, H.; Bayachou, M.; Boukherroub, R.; Szunerits, S., Peroxynitrite activity of hemin-functionalized reduced graphene oxide. Analyst 2013, 138 (15), 4345-4352, DOI 10.1039/C3AN00678F. 31. Carrillo, I.; Blanca, E. S. D. L.; Fierro, J. L. G.; Raso, M. A.; Acción, F.; Enciso, E.; Redondo, M. I., Conductivity and long term stability of polypyrrole poly(styrene- co -methacrylic acid) core–shell particles at different polypyrrole loadings. Thin Solid Films 2013, 539 (4), 154-160, DOI 10.1016/j.tsf.2013.05.086. 32. Qiao, Y.; Shen, L.; Wu, M.; Guo, Y.; Meng, S., A novel chemical synthesis of bowl-shaped polypyrrole particles. Materials Letters 2014, 126, 185-188, DOI 10.1016/j.matlet.2014.04.041. 33. Chebil, S.; Monod, M. O.; Fisicaro, P., Direct electrochemical synthesis and characterization of polypyrrole nano- and micro-snails. Electrochimica Acta 2014, 123 (4), 527-534, DOI 10.1016/j.electacta.2014.01.058. ACS Paragon Plus Environment

28

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

ACS Sustainable Chemistry & Engineering

34. Carquigny, S.; Sanchez, J. B.; Berger, F.; Lakard, B.; Lallemand, F., Ammonia gas sensor based on electrosynthesized polypyrrole films. Talanta 2009, 78 (1), 199-206, DOI 10.1016/j.talanta.2008.10.056. 35. Xin, S.; Yang, N.; Gao, F.; Zhao, J.; Li, L.; Teng, C., Free-standing and flexible polypyrrole nanotube/reduced graphene oxide hybrid film with promising thermoelectric performance. Materials Chemistry and Physics 2018, 212, 440-445, DOI 10.1016/j.matchemphys.2018.03.025. 36. Wang, Y. P.; Ding, Z. J.; Liu, Q. X.; Liu, W. J.; Ding, S. J.; Zhang, D. W., Plasma-assisted atomic layer deposition and post-annealing enhancement of low resistivity and oxygen-free nickel nanofilms using nickelocene and ammonia precursors. Journal of Materials Chemistry C 2016, 4 (47), 11059-11066, DOI 10.1039/C6TC03606F. 37. Wang, Y.; Rouabhia, M.; Zhang, Z., PPy-coated PET fabrics and electric pulse-stimulated fibroblasts. Journal of Materials Chemistry B 2013, 1 (31), 3789-3796, DOI 10.1039/C3TB20257G. 38. Li, L.; Xu, J.; Lei, J.; Zhang, J.; Mclarnon, F.; Wei, Z.; Li, N.; Pan, F., A one-step, cost-effective green method to in situ fabricate Ni(OH)2 hexagonal platelets on Ni foam as binder-free supercapacitor electrode materials. Journal of Materials Chemistry A 2015, 3 (5), 1953-1960, DOI 10.1039/C4TA05156D. 39. Abdul Bashid, H. A.; Lim, H. N.; Kamaruzaman, S.; Abdul, R. S.; Yunus, R.; Huang, N. M.; Yin, C. Y.; Rahman, M. M.; Altarawneh, M.; Jiang, Z. T., Electrodeposition of Polypyrrole and Reduced Graphene Oxide onto Carbon Bundle Fibre as Electrode for Supercapacitor. Nanoscale research letters 2017, 12 (1), 246, DOI 10.1186/s11671-017-2010-3. 40. Zhu, J.; Huang, L.; Xiao, Y.; Shen, L.; Chen, Q.; Shi, W., Hydrogenated CoOx nanowire@Ni(OH)2 nanosheet core-shell nanostructures for high-performance asymmetric supercapacitors. Nanoscale 2014, 6 (12), 6772-6781, DOI 10.1039/C4NR00771A. 41. Bag, S.; Raj, C. R., Layered inorganic-organic hybrid material based on reduced graphene oxide and α-Ni(OH)2 for high performance supercapacitor electrodes. Journal of Materials Chemistry A 2014, 2 (42), 17848-17856, DOI 10.1039/C4TA02937B. 42. Marchetti, L.; Miserque, F.; Perrin, P.; Pijolat, M., XPS study of Ni‐ base alloys oxide films formed in primary conditions of pressurized water reactor. Surface and Interface Analysis 2015, 47 (5), 632-642, DOI 10.1002/sia.5757 43. Ding, R.; Qi, L.; Jia, M.; Wang, H., Facile synthesis of mesoporous spinel NiCo₂ O₂ nanostructures as highly efficient electrocatalysts for urea electro-oxidation. Nanoscale 2014, 6 (3), 1369-1376, DOI 10.1039/C3NR05359H. 44. Prathap, M. U. A.; Satpati, B.; Srivastava, R., Facile preparation of β-Ni(OH)2-NiCo2O4 hybrid nanostructure and its application in the electro-catalytic oxidation of methanol. Electrochimica Acta 2014, 130, 368-380, DOI 10.1016/j.electacta.2014.03.043. 45. Qian, Q.; Hu, Q.; Li, L.; Shi, P.; Zhou, J.; Kong, J.; Zhang, X.; Sun, G.; Huang, W., Sensitive Fiber Microelectrode Made of Nickel Hydroxide Nanosheets Embedded in Highly-Aligned Carbon Nanotube Scaffold for Nonenzymatic Glucose Determination. Sensors & Actuators B Chemical 2018, 257, 23-28, DOI 10.1016/j.snb.2017.10.110. 46. Laviron, E., General expression of the linear potential sweep voltammogram in the case of diffusionless electrochemical systems. Journal of Electroanalytical Chemistry 1979, 101 (1), 19-28, DOI 10.1016/S0022-0728(79)80075-3. 47. Xia, B. Y.; Wang, J. N.; Wang, X. X., Synthesis and Application of Pt Nanocrystals with Controlled Crystallographic Planes. Scandinavian Journal of Trauma, Resuscitation and Emergency Medicine 2009, 17 (3), 356-356, DOI 10.1021/jp906734d. 48. Wang, D.; Yan, W.; Vijapur, S. H.; Botte, G. G., Enhanced electrocatalytic oxidation of urea based on nickel hydroxide nanoribbons. J. Power Sources 2012, 217, 498-502, 49. Hahn, F.; Beden, B.; Croissant, M. J.; Lamy, C., In situ uv visible reflectance spectroscopic investigation of the nickel electrode-alkaline solution interface. Electrochimica Acta 1986, 17 (24), 335342, DOI 10.1016/0013-4686(86)80087-1. ACS Paragon Plus Environment

29

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

Page 30 of 31

50. Yan, W.; Wang, D.; Botte, G. G., Electrochemical decomposition of urea with Ni-based catalysts. Applied Catalysis B Environmental 2012, 127 (3), 221-226, DOI 10.1016/j.apcatb.2012.08.022. 51. Wei, Y.; Dan, W.; Diaz, L. A.; Botte, G. G., Nickel nanowires as effective catalysts for urea electro-oxidation. Electrochimica Acta 2014, 134 (21), 266-271, DOI 10.1016/j.electacta.2014.03.134. 52. Zhu, X.; Dou, X.; Dai, J.; An, X.; Guo, Y.; Zhang, L.; Tao, S.; Zhao, J.; Chu, W.; Zeng, X. C., Metallic Nickel Hydroxide Nanosheets Give Superior Electrocatalytic Oxidation of Urea for Fuel Cells. Angewandte Chemie 2016, 128 (40), 12653-12657, DOI 10.1002/anie.201606313. 53. Kakati, N.; Maiti, J.; Kang, S. L.; Viswanathan, B.; Yoon, Y. S., Hollow Sodium Nickel Fluoride Nanocubes Deposited MWCNT as An Efficient Electrocatalyst for Urea Oxidation. Electrochimica Acta 2017, 240, 175-185, DOI 10.1016/j.electacta.2017.04.055. 54. Tong, Y.; Chen, P.; Zhang, M.; Zhou, T.; Zhang, L.; Chu, W.; Wu, C.; Xie, Y., Oxygen Vacancies Confined in Nickel Molybdenum Oxide Porous Nanosheets for Promoted Electrocatalytic Urea Oxidation. Acs Catalysis 2017, 8 (1), 1-7, DOI 10.1021/acscatal.7b03177. 55. Chen, S.; Duan, J.; Vasileff, A.; Qiao, S. Z., Size Fractionation of Two-Dimensional SubNanometer Thin Manganese Dioxide Crystals towards Superior Urea Electrocatalytic Conversion. Angew. Chem. Int. Ed. 2016, 55 (11), 3804-3808, DOI 10.1002/anie.201600387. 56. Xiao, M.; Tian, Y.; Yan, Y.; Feng, K.; Miao, Y., Electrodeposition of Ni(OH)2/NiOOH in the Presence of Urea for the Improved Oxygen Evolution. Electrochimica Acta 2015, 164, 196-202, DOI 10.1016/j.electacta.2015.02.205. 57. Liu, S. J., Kinetics of methanol oxidation on poly(Ni-tetramethyldibenzotetraaza[14] annulene)modified electrodes. Electrochimica Acta 2004, 49 (19), 3235-3241, DOI 10.1016/j.electacta.2004.02.038. 58. Shao, M.; Xu, X.; Han, J.; Zhao, J.; Shi, W.; Kong, X.; Wei, M.; Evans, D. G.; Duan, X., Magnetic-field-assisted assembly of layered double hydroxide/metal porphyrin ultrathin films and their application for glucose sensors. Langmuir 2011, 27 (13), 8233-8240, DOI 10.1021/la201521w. 59. Mao, H.; Ji, C.; Liu, M.; Sun, Y.; Liu, D.; Wu, S.; Zhang, Y.; Song, X. M., Hydrophilic polymers/polypyrrole/graphene oxide nanosheets with different performence in electrocatalytic application to simultaneous determination of dopamine and ascorbic acid. Rsc Advances 2016, 6 (113), 111632-111639, DOI 10.1039/C6RA23341D.

ACS Paragon Plus Environment

30

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

ACS Sustainable Chemistry & Engineering

Hierarchical Ni(OH)2/PPy/GO nanosheets exhibited the excellent electrocatalytic activity towards urea oxidation reaction. 82x31mm (300 x 300 DPI)

ACS Paragon Plus Environment