You are on page 1of 43

Author’s Accepted Manuscript

Metal-organic framework-derived nitrogen-doped


highly disordered carbon for electrochemical
ammonia synthesis using N2 and H2O in alkaline
electrolytes

Shreya Mukherjee, David A. Cullen, Stavros


Karakalos, Kexi Liu, Hao Zhang, Shuai Zhao, Hui
www.elsevier.com/locate/nanoenergy
Xu, Karren L. More, Guofeng Wang, Gang Wu

PII: S2211-2855(18)30199-X
DOI: https://doi.org/10.1016/j.nanoen.2018.03.059
Reference: NANOEN2607
To appear in: Nano Energy
Received date: 24 February 2018
Revised date: 16 March 2018
Accepted date: 22 March 2018
Cite this article as: Shreya Mukherjee, David A. Cullen, Stavros Karakalos, Kexi
Liu, Hao Zhang, Shuai Zhao, Hui Xu, Karren L. More, Guofeng Wang and Gang
Wu, Metal-organic framework-derived nitrogen-doped highly disordered carbon
for electrochemical ammonia synthesis using N2 and H2O in alkaline
electrolytes, Nano Energy, https://doi.org/10.1016/j.nanoen.2018.03.059
This is a PDF file of an unedited manuscript that has been accepted for
publication. As a service to our customers we are providing this early version of
the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting galley proof before it is published in its final citable form.
Please note that during the production process errors may be discovered which
could affect the content, and all legal disclaimers that apply to the journal pertain.
Metal-organic framework-derived nitrogen-doped highly disordered carbon for

electrochemical ammonia synthesis using N2 and H2O in alkaline electrolytes

Shreya Mukherjeea, David A. Cullenb, Stavros Karakalosc, Kexi Liud, Hao Zhanga, Shuai Zhaoe,

Hui Xue, Karren L. Moref, Guofeng Wang*,d, Gang Wu*,a

a
Department of Chemical and Biological Engineering, University at Buffalo, The State University

of New York, Buffalo, New York 14260, United States.

b
Materials Science and Technology Division, Oak Ridge National Laboratory, Oak Ridge,

Tennessee 37831, United States

c
Department of Chemical Engineering, University of South Carolina, Columbia, South Carolina

29208, United States

d
Department of Mechanical Engineering and Materials Science, University of Pittsburgh,

Pittsburgh, Pennsylvania 15261, United States

e
Giner Inc., Newton, Massachusetts 02466, United States

f
Center for Nanophase Materials Sciences, Oak Ridge National Laboratory, Oak Ridge, Tennessee

37831, United States

*
Corresponding Authors

gangwu@buffalo.edu

guw8@pitt.edu

1
Abstract: Ammonia (NH3) is considered an important chemical for both agriculture fertilizer and

renewable energy. The conventional Haber-Bosh process to produce NH3 is energy intensive and

leads to significant CO2 emission. Alternatively, electrochemical synthesis of ammonia (ESA)

through the nitrogen reduction reaction (NRR) by using renewable electricity has recently

attracted significant attention. Herein, we report a metal-organic framework-derived

nitrogen-doped nanoporous carbon as electrocatalysts for the NRR. It exhibits a remarkable

production rate of NH3 up to 3.4×10-6 mol cm-2 h-1 with a Faradaic efficiency (FE) of 10.2 % at

-0.3 V vs RHE under room temperature and ambient pressure using aqueous 0.1 M KOH

electrolyte. Increasing the temperature to 60oC was found to further improve production rates to

7.3×10-6 mol cm-2 h-1. The stability of the nitrogen-doped electrocatalyst was demonstrated during

an 18-hour continuous test with constant production rates. First principles calculations were used

to elucidate the possible active sites and reaction pathway. The moiety, which consists of three

pyridinic N atoms adjacent with one carbon vacancy embedded in a carbon layer, is able to

strongly adsorb N2 and further realize NN triple bond dissociation for the subsequent protonation

process. The rate-determining step of the NRR is predicted to be the adsorption and bond

activation of the N2 molecule. Increasing overpotentials are favorable for the protonation process

during NH3 generation. Further doping Fe into the nitrogen-doped carbon likely blocks the N3

active sites and facilitates the hydrogen evolution reaction, a strong competitor to the NRR, thus

yielding negative effect on ammonia production. This work provides a new insight into the rational

design and synthesis of nitrogen-doped and defect-rich carbon as efficient NRR catalysts for NH3

synthesis at ambient conditions.

2
Graphical Abstract:

Keywords: Electrocatalysis; NH3 synthesis; MOFs; nanocarbon; first principle calculation

3
1. Introduction

Apart from being produced for fertilizers, ammonia is also being considered as a carbon neutral

liquid fuel because of its high energy density and suitability, and is projected to play a significant

role in the future hydrogen economy [1-3]. The highly energy-intensive traditional Haber–Bosch

process for ammonia production (400-500oC and 150–250 atm) leads to substantial greenhouse

gas emission due to the use of fossil fuels [4, 5]. There is a growing need for developing alternative

and sustainable approaches for ammonia synthesis. To this end, environmentally benign processes

such as photocatalysis and electrocatalysis via the nitrogen reduction reaction (NRR) at ambient

conditions without using hydrogen as a reactant is a rapidly expanding field of research [6-12].

The electrochemical synthesis of ammonia (ESA) was discovered by Davy et al. in 1807 by using

only pure water and dissolved air between two gold electrodes [13]. The main challenge of ESA is

the slow kinetics of N2 adsorption and subsequent NN triple bond cleavage. In addition, because

the standard reduction potential for the hydrogen evolution reaction (HER) at 0 V is comparable to

that for the NRR (0.057 V vs. SHE), the competing HER significantly reduces the Faradaic

efficiency (FE) of the ESA. The cell reactors are categorized as either alkaline or acidic systems. In

a typical proton conducting electrolyte cell, reduction reaction of N2 gas occurs at the cathode, and

oxidation reaction of H2 takes place at the anode. The produced protons H+ at the anode diffuse to

the cathode where they combine with dissociated N to form NH3 [14]. However, there are two

apparent limitations for ESA in the acidic media. The HER is more dominant in acids as evidenced

by the two orders of magnitude higher current density in acidic medium than that in alkaline
4
electrolytes [15]. Utilization of H2 during the ESA will encounter grand challenges associated with

H2 production, transportation, and storage, which also contradicts with the original intention of

using the NH3 as the alternative H2 carrier. Therefore, we develop efficient ESA in alkaline media

by directly using N2 and H2O.

A key consideration to the design of effective NRR catalysts is the mitigation of the HER. A

series of precious metals (e.g., Pt, Au, and Ru) and nonprecious metals (e.g., Fe, Mo, Ni) were

studied, but all of them suffer from poor FEs due to their high activity for the HER over the NRR

[6, 16-20]. Recently, density functional theory (DFT) calculations have predicted that transition

metal nitrides such as VN, CrN, ZrN, NbN, and MnN could be active for the NRR [21]. The

theoretical calculations also show that the nitride-driven formation of NH3 follows the Mars-van

Krevelen mechanism, in which one surface N atom of the nitride is converted to form NH3 and the

nitride catalyst is later regenerated with the dissociated N2 [21, 22]. DFT studies further predicted

that some oxides might be more active for nitrogen binding than adsorption of protons [23]. The

mechanism of nitrogen binding could be associative, where N2 tends to break simultaneously with

hydrogen addition, or where nitrogen dissociates and attaches on the active site and then is

protonated. Skulson et al. theoretically predicted that several metals oxides including Nb, Re, Ta,

and Ir have lower or similar binding energy of N-NH, which is an intermediate formed after the

first protonation of the adsorbed nitrogen molecule, compared to binding energy of protons, thus,

promising to be more active for the NRR over the HER [23]. However, these theoretical

predictions have yet to be experimentally verified. Development of highly efficient catalysts


5
exclusively active for the NRR instead of HER still remains a significant challenge for practical

ESA applications. The catalysts studied to date are included in Table S1 and are sorted by their

NH3 production rates and FE.

Compared to transition metal-based catalysts, “metal-free” carbon catalysts may provide a new

opportunity to develop effective NRR catalysts while mitigating the competing HER. Among

studied carbon catalysts, metal-organic framework (MOF)-derived carbons have high surface

areas, easily controllable functionalization, tailorable morphologies, and tunable porosity [24-27].

Some of them have exhibited encourgaing catalytic selectivity for the CO2 reduction reactions

(CO2RR) over HER [28-32]. Due to the similarity between the NRR and CO2RR, such

MOF-derived carbons could be explored for the NRR catalysts during the ESA with alleviated

HER interference [33]. Besides, compared to transition metals such as Fe, Co or Ni doping [34,

35], “metal-free” MOF-derived N-doped carbons have suppressed activity for the HER. Thus, we

report a nitrogen-doped and defect-rich nanoporous carbon catalyst derived from zinc-based

zeolitic imidazolate frameworks (ZIF-8) via one-step thermal activation. The resulting carbon is

nearly “metal-free”, because heating treatments at high temperatures are able to complete remove

Zn metal due to the low melting point of Zn at 907oC. The 3D frame work structures of ZIFs could

be nearly maintained during the carbonization, giving high surface areas and tunable porosity as

well as excellent thermal and chemical stability. Importantly, the nitrogen doping and the degree

of graphitization could be controlled by tuning the heating conditions, which modify the electronic

and geometric structures of carbon, thus favoring for the NRR electrocatalysis [36-39]. In this
6
work, the resulting carbon catalyst with optimal nitrogen doping and carbon microstructure

exhibited very encouraging NRR activity, yielding a NH3 production rate of 3.4×10-6 mol cm-2 h-1

and an FE as high as 10.2 % at -0.3 V vs. RHE at room temperature and ambient pressure in 0.1 M

KOH electrolyte. The production rate can be further increased to 7.3×10-6 mol cm-2 h-1 at 60oC.

The existence of possible FeN4 moiety in carbon was also investigated, but it plays a negative role

in catalyzing the NRR likely due to reduced carbon/nitrogen vacancies and enhanced HER. DFT

calculations further elucidate that a carbon moiety containing three pyridinic nitrogen atoms (N3)

is able to provide strong adsorption of N2 and favorable for subsequent NN cleavage during the

NRR.

2. Experimental details

2.1 Catalyst synthesis

To prepare ZIF-8 nanocrystal precursors, Zn(NO3)2 6H2O and 2-methylimidazole were separately

dissolved in equal volumes of methanol solution (e.g., 300 ml) at room temperature. A molar ratio

of 1:4 for Zn(NO3)2 and 2-methylimidazole were accurately controlled. Then, the solution of

2-methylimidazole was gradually added to the solution of Zn(NO3)2 6H2O followed by an aging

process at 60oC for 24 hours. Then, the solid precipitate was collected from the colloidal solution

by centrifugation at 11000 rpm for 15 min and washed with methanol at least three times to

remove impurities. Next, the products were dried at 60oC for 6 h. The prepared ZIF-8 nanocrystal

precursors were further subjected to high temperature activation (700-1100oC) under constant N2

7
flow in a tube furnace. The duration of heating treatment was also varied to adjust the degree of

graphitization and the N content of the resulting carbon catalysts. The prepared samples are

labeled as C-ZIF-T-t, where T represents the carbonization temperature (e.g., 700, 900, 1100, and

1200oC) and t represents the duration (30 min,1 h, and 3 h) for the heating treatment. To study the

effect of Fe doping on the NRR, as control samples, Fe-doped ZIF-8 was also prepared under the

identical conditions except for adding Fe(NO3)2.6H2O along with Zn(NO3)2 6H2Oduring the

synthesis of ZIF-8. They are denoted as xFe-C-ZIF-T-t, where x represents the doping content of

Fe up to 20 at.%.

2.2 Electrochemical measurements

A rotating glassy-carbon-disk electrode (GC-RDE) coated with catalyst layers was used as the

working electrode. When preparing the catalyst ink, 10 mg of the catalyst and 30 µL of Nafion

were dispersed in 1.0 ml isopropyl alcohol and ultrasonicated for 30 min to obtain a homogenous

suspension. 20 µL of the ink was dropped onto the GC-RDE with a surface area of 0.245 cm2.

NRR activity of catalyst for NH3 electrochemical synthesis was measured in a three-electrode

electrochemical cell with 150 mL N2 saturated 0.1 M KOH. A graphite rod and an Hg/HgO

electrode were used as the counter and the reference electrodes, respectively. The reference

electrode was calibrated with respect to the reversible hydrogen electrode (RHE) by bubbling pure

H2 at 1.0 atm into a calibration tube fitted with a Pt wire coated with Pt black. A catalyst loading of

0.8 mg cm–2 was used for all of tests. Another electrochemical cell with Ar-saturated 0.1 M KOH

8
solution was used separately as a control to not only determine the blank content of NH3, but also

study the catalyst activity for the HER. Ar or N2 were purged into the corresponding

electrochemical cell at the rate of 200 mL min-1 for 30 minutes before all of tests to eliminate the

dissolved air in KOH solutions, and a constant flow rate of 20 mL min-1 was maintained during the

test. The electrochemical cells were completely sealed to prevent solvation of oxygen from air.

Cyclic voltammetry (CV) and linear sweep voltammetry (LSV) were performed at a scan rate of

50 and 5 mV s-1, respectively, to study the possible redox and current density associated with both

the NRR and HER. Potentiostatic tests were used to continuously generate NH3 during the NRR.

To determine the production rate of NH3, 5 mL of electrolyte was collected after the potentiostatic

tests for different durations, which was then used for indophenol test and spectrophotometric

analysis to determine the NH3 concentration of the electrolytes [40, 41] . The UV-Vis diffuse

reflectance spectra (UV-vis DRS) of the collected electrolytes were recorded on a Shimadzu

UVProbe 2600 Series UV-vis- spectrophotometer equipped with an integrating sphere at

wavelengths from 500 to 725 nm. The peak for ammonia was observed at 650 nm. The detailed

experiments are included in the supporting information (Figure S1). The concentration (mol L-1)

of ammonia was calculated from a calibration curve prepared by using standard ammonium

chloride solutions (Figure S2). The ammonia production rates (R) are calculated by using equation

(1):

R[NH3] = ([NH3] ×V)/(t ×A) (1)

9
Where [NH3] is the concentration (mol L-1) of ammonia, V the volume of electrolyte, t is the

reaction time, and A is the surface area of the working electrode. Assuming three electrons are

consumed to produce one NH3 molecule, the FE was calculated using equation (2):

FE =3F×[NH3]×V/(17×Q), (2)

Where F is the Faraday constant and Q is the total charge passed through the electrodes during the

reaction duration according to the total current density.

2.3 Material characterization

The X-ray diffraction (XRD) patterns of the samples were recorded on a Rigaku Ultima IV

diffractometer with Cu Kα X-rays operating at 200 mA and 40 kV, using Cu Kα as the radiation

source (λ=0.15418 nm). A JEOL JEM-2100F field emission electron microscope operating at 200

kV with an EDAX energy disperse X-ray spectroometer (EDS) and Nion UltraSTEM U100

operated at 60kV were used for electron microscopy studies at Oak Ridge National Laboratory

(ORNL). The scanning electron microscopy (SEM) images of the samples were taken on a Hitachi

SU 70 SEM at a working voltage of 5 kV. Raman spectroscopy was performed using a Renishaw

Raman system at 514 nm excitation. Powder samples were deposited on a standard microscope

glass slide. X-ray photoelectron spectroscopy (XPS) was conducted using a Kratos AXIS Ultra

DLD XPS system equipped with a hemispherical energy analyzer and a monochromatic Al Kα

source operated at 15 keV and 150 W; the pass energy was fixed at 40 eV for the high-resolution

scans. Before analysis, the samples were placed in vacuum at 80°C overnight to remove any

10
adsorbed substances. The specific surface area (SBET) measurements were carried out using N2

adsorption/desorption at 77 K on a Micromeritics TriStar II instrument. Samples were degassed at

130 °C for 5 h under vacuum prior to nitrogen physisorption measurements.

2.4 First principles calculations

Spin-polarized DFT calculations were performed using the Vienna ab initio simulation package

(VASP) code [42, 43]. Projector augmented wave (PAW) pseudopotential was used to describe the

core and valence electrons. A plane wave basis set with a kinetic energy cutoff of 400 eV was used

to expand the wave functions [44]. Electronic exchange and correlation were described within the

framework of generalized gradient approximation (GGA) of the revised Perdew, Burke and

Ernzernhof (RPBE) functionals [45]. The pyridinic N3 moiety embedded in the graphene layer was

investigated as the model of active sites. The Brillouin zone was sampled using Monkhorst-Pack 4

× 4 × 1 k-point grids. The atomic positions were optimized until the forces were below 0.01 eV/Å

during structural optimization. The computational hydrogen electrode method developed by

Nørskov was used to calculate the free energy of each intermediate state involved in

electrochemical NRR [46].

3. Results and discussion

3.1 Catalyst structures and morphologies

ZIF-8 nanocrystals, which are made from Zn ions with 2-methylimidazole, consist of Zn-N4

complex connected by hydrocarbon networks. They are ideal 3D framework precursors and

provide both N and C sources simultaneously. One-step thermal activation can directly carbonize
11
these well-defined nanocrystals and form nitrogen-doped nanoporous carbon. When the heating

temperature is above the boiling point of Zn (907 oC), the metallic Zn species evaporates, leaving

an open porous structure. Unlike other transition metals (e.g., Fe, Co, Ni, and Mn), Zn doesn’t play

a catalytic role in the formation of graphitized carbon. Thus, highly disordered and porous

N-doped carbon is obtained after the thermal activation of ZIF-8. In this work, as shown in Figure

1a-1e, the ZIF-8 precursor nanocrystals exhibited a polyhedral structure and uniform dispersion

with a size around 50 nm. HR-TEM images further indicate an apparent micropore feature.

Atomically-dispersed zinc sites were clearly observed in high-angle annular dark-field scanning

transmission electron microscopy (HAADF-STEM) images (Figure 1e), though it should be noted

the crystalline structure has likely collapsed due to the sensitivity of ZIF-8 to the electron beam

[47]. After a typical heating treatment at 1100oC in N2 for one hour, the ZIFs have completely

converted to carbon as evidenced by XRD patterns, showing low degree of graphitization (Figure

S3). During the carbonization, the well-dispersed ZIF nanocrystals are fused together and lose

their original polyhedral structures (Figure 1f-g). HR-TEM images (Figure 1i and Figure S4)

present highly disordered 3D carbon structures, indicative of graphitic-type carbon with high

vacancy and dislocation densities. As expected, most of atomic Zn disappears from carbon phases

after the heating treatment at 1100oC (Figure 1j).

12
Figure 1. Comparison of the morphology and microstructure between (a–e) ZIF-8 nanocrystal
precursors and (f–j) the corresponding carbon catalysts generated through one-step thermal
activation at 1100°C under a N2 atmosphere for 1 hour.

To provide insightful understanding of the correlation between disorder degree in the carbon

structure and NRR activity, Raman spectra were recorded as a function of heating temperature and

time during the synthesis. As shown in Figure 2(a), Raman spectra for all of catalysts can be fitted

into four carbon species. Two characteristic carbon resonances around 1600 cm−1 (G band) and

1350 cm−1 (D band) are dominant, which correspond to the planar motion of sp2-hybridized carbon

atoms in an ideal carbon plane and from carbon atoms at the edge, respectively [48-51]. The ratios

of these two peaks could be an indication of the overall degree of graphitization in carbon. In

addition, two broad signals at ca. 1200 and 1510 cm−1 are also convolved in these Raman spectra,

which are correspondingly associated with the carbon atoms outside of a perfectly planar sp2

carbon network (such as aliphatic or amorphous structures) and integrated five-member rings or

heteroatoms in carbon layers [48]. Based on above analysis, the peak around 1510 cm-1 becomes

obviously smaller with an increase in heating temperatures, suggesting a reduction of nitrogen


13
doping and other irregular carbon structures, i.e. five-side ring. D peaks become more dominant

when heating temperature is above 900 oC, indicating an abundance of carbon edge sites. This is in

good agreement with the observation of HR-TEM images showing highly disordered carbon rich

in vacancies and dislocations. Meanwhile, G peaks are narrower at higher temperatures, likely due

to decreased portion of ideal carbon planes. Therefore, higher heating temperatures during catalyst

synthesis yield more disordered carbon structures, which seems to be more favorable for the NRR.

It should be noted that additional time heating at high temperature (such as 3 h at 1100oC) did not

lead to significant change in the ratio of the G and D bands, and hence the decreased NH3

production rates with extended annealing time cannot be explained purely through carbon

structural changes.

XPS analysis was employed to elucidate the effects of nitrogen doping content and position on

NH3 production rates in these carbon catalysts. N 1s spectra are compared as a function of heating

temperature and duration in Figure 2b. The elemental quantification determined by XPS is also

listed in Table S2. Overall, pyridinic, pyrrolic, and graphitic nitrogen atoms can be doped into

carbon structures, which can be identified by their differences in binding energies at ca. 398.6,

400.2, and 401.1 eV. Pyridinic and graphitic nitrogen atoms are at the edge and interior of the

carbon planes, respectively, while pyrrolic N is in a five-side carbon ring. With an increase of

heating temperatures up to 900oC, pyridinic and pyrrolic N are initially generated. However,

pyrrolic N is not thermally stable and become insignificant when the heating temperature is above

1000oC. As for the best performing C-ZIF-1100-1h catalyst prepared from 1100oC, there are only
14
dominant pyridinic and graphitic N doped into carbon. With an increase of heating temperature,

the total nitrogen content is found continuously dropped, likely due to the loss of pyrrolic N. In

addition, the level of graphitic N is relatively increased compared to the pyridinic N edge sites. The

similar trend was also observed for ZIF catalysts containing transition metals such as Co or Fe

used for the oxygen reduction reaction [52, 53]. Especially up to 1100oC, two types of doped

nitrogen atoms become comparable. Thus, the formation of more stable graphitic N seems to

promote the NRR. When the heating duration is extended from one to three hours, the carbon

structures nearly remain the same (C1s spectra in Figure S3 and Table S4). However, the relative

content of pyridinic nitrogen is significantly decreased, which corresponds to the reduced NRR

activity for the ESA. This may suggest that pyridinic N is also crucial for participating into active

sites for the NRR. It also should be noted that increased heating temperature leads to reduced Zn

content in catalysts. This could be beneficial for the NRR by removing zinc sites and leaving

pyridinic N available. More discussion will be provided in the modeling section to elucidate the

contribution of pyridinic and graphitic N sites to the NRR.

15
(a) N doping/ (b)
700-1h D peak C5-ring 700-1h Pyridinic N
G peak

sp2 carbon Pyrrolic N


outside plane

800-1h 800-1h

900-1h

Intensity (a. u.)


900-1h
Intensity (a. u.)

1100-1h 1100-1h
Graphitic N

1100-3h 1100-3h

1000 1200 1400 1600 1800 404 402 400 398 396
Raman Shift (cm-1) Binding Energy (eV)

Figure 2. Raman spectra (a) and XPS N1s spectra (b) for the ZIF-8-derived carbon catalysts
synthesized at various heating temperatures (700 to 1100oC) and duration times (1 vs 3 hours).

BET surface areas and pore size distribution of these ZIF-derived carbon catalysts were also

studied as a function of heating temperatures (Figure S6 and Table S4). The as-prepared ZIF-8

precursors with an average size of 50 nm show apparent features of the micropores and a high

surface area up to 1100 m2/g. However, surface areas of the heat-treated carbon catalysts were
16
found largely dependent on heating temperatures. Due to the formation of metallic zinc, low

temperatures (700-900oC) yield relatively low surface areas, which is in good agreement with

other ZIF-derived catalysts [52]. However, increasing temperatures up to 1100oC favors the

evaporation of zinc and generates a high surface area up to 780 m2/g. The pore size distribution

indicates that, in addition to micropore, significant meso- and macro-pores are dominant in the

carbonized ZIF carbons, generating a maximum pore volume at 1100oC. The favorable

morphology obtained at 1100oC may partially contribute to the highest catalytic activity for the

NRR.

3.2 Nitrogen reduction for electrochemical ammonia synthesis

Effects of carbonization conditions during catalyst preparation. Carbonization conditions for

ZIF-8 nanocrystals proved effective towards tuning the level of nitrogen doping and the degree of

graphitization in carbon structures [54-56]. We therefore studied the effects of heating temperature

and duration on catalyst NRR activity toward the ESA. Within the temperature range from 600 to

1100oC, the catalyst prepared from 1100oC (i.e., C-ZIF-1100) exhibited the highest NRR activity

in terms of maximum production rate for the ESA, i.e., 3.4×10-6 mol cm-2 h-1 and an FE as high as

10.2 % at -0.3 V vs. RHE at room temperature and ambient pressure in 0.1 M KOH electrolyte.

This is in good agreement with the largest difference in current density measured from N2 and Ar

saturated electrolytes during the LSV tests (Figure S7). In addition, we tested NRR activity with

the best performing C-ZIF-1100 catalysts synthesized for three different heating durations (i.e., 30

17
min, 1 h, and 3 h). One-hour duration was found to be optimal for the highest production rate of

ESA. The absorbance spectra derived from UV-Vis spectroscopy after each potentiostatic test and

corresponding production rates are compared in Figure 3a and 3b, respectively. It should be noted

that the electrolyte itself was tested for ammonia before potentiostatic test to ensure there was no

contamination of ammonia from the environment.

Both the NRR and the HER are largely dependent on applied reduction potentials. Although

increasing overpotentials can accelerate the NRR, it would significantly boost the competing HER

as well to decrease FE for the ESA. Thus, given the balance between increased electrochemical

activation for the NRR and maximum inhibition of the HER, we also studied the potential

dependence of the ESA on the best performing C-ZIF-1100-1h catalyst. The measured current

densities in both N2 and Ar saturated electrolytes at various potentials were compared in Figure

3c. The UV-Vis absorbance spectra of the electrolytes after one-hour potentiostatic tests at

different reduction potentials are compared in Figure S8. The corresponding NH3 production rates

and FEs calculated from these spectra are compared in Figure 3d. The current densities measured

with N2 and Ar increase with decreasing potentials, i.e., more negative. In the relatively low

overpotential range (i.e., E > - 0.2 V), although the current density in Ar, which is directly related

to HER, is low, the corresponding EAS is also insignificant due to insufficient electrochemical

overpotential for the NRR. On further reducing the potential, EAS and FE reach the maximum at

-0.3 V, suggesting a tradeoff between the NRR and the HER at the given testing conditions. As the

potential is further incrementally decreased to -0.5 V, both ESA and FE are decreased because of
18
the more prevailing HER, as evidenced by the significantly increased current density measured

with Ar in this potential range. The comparison of production rates and the difference in current

densities in N2 and Ar for all samples is tabulated in Table S5.

We also tested un-pyrolyzed ZIF-8 precursors for its NRR activity. Ammonia is detected from

both Ar and N2 saturated electrolytes during the NRR, suggesting the instability of nitrogen

species in the ZIF-8 precursors. The nitrogen species also appeared to be instable for the samples

heated at or below 900oC. If the Ar saturated electrolyte was used, about 0.005 ppm ammonia was

detected for C-ZIF-700-1h (Figure S9) and C-ZIF-900-1h samples (Figure S10) in Ar

electrolytes, but there is no NH3 observed with the C-ZIF-1100-1h sample. Thus, higher heating

temperature is helpful to stabilize N species in the carbon catalysts during the NRR. Furthermore,

the production rate of the C-ZIF-1100 -1h catalyst remained nearly constant during a continuous

NRR test up to 18 hours as shown in Figure 3e and 3f. Oppositely, no ammonia is detected in Ar

atmosphere during the stability test, indicating excellent stability of the doped nitrogen in carbon

within the reductive potentials for the NRR.

19
Figure 3. (a) Comparison of UV-Vis absorbance spectra after potentiostatic tests in 0.1 M KOH at
25oC for the NRR on various C-ZIF catalysts synthesized from different heating temperatures and
duration. (b) Corresponding NH3 production rate and FE. (c) Comparison of current density in N2
(black dots) and Ar (red dots) as function of applied reduction potentials for the best performing
C-ZIF-1100-1h catalyst. (d) Corresponding NH3 production rate and FE as function of potential
during the NRR. (e) Stability test in both N2 and Ar saturated electrolyte, and (f) the constant

20
production rate with reaction time indicating good stability of the C-ZIF-8-1100-1h catalyst for the
NRR.

In addition, we also studied the NRR activity on other N-doped carbon catalysts derived from

different nitrogen/carbon precursors (Figure S11). For example, we synthesized N-doped carbon

from polyaniline (PANI) under the identical heating treatment conditions to the ZIF-derived

carbon catalyst. Using the same testing and measurement protocols, though the PANI-derived

carbon catalyst shows NRR activity, the production rate is one order of magnitude lower relative

to the ZIF-derived one. In addition, N-doped carbon tubes derived from dicyandiamide (DCDA)

was also studied and exhibited much lower catalytic activity as well. A carbon black sample (e.g.,

commercial Ketjenblack EC-600J) free of nitrogen doping only generates negligible NRR

activity, suggesting the importance of doped nitrogen atoms in carbon. Therefore, the major

advantage of ZIF-derived carbon is due to the formation of highly disordered amorphous carbon

with stable and appropriate nitrogen doping. Oppositely, other N-doped carbon materials derived

from PANI and DCDA have higher degree of graphitization [57-61], which is lack of necessary

carbon defects. The 3D carbon frameworks derived from ZIF-8 free from active transition metals

such as Fe, Co, or Ni can suppress the HER, thus benefiting the enhancement of the NRR for

ammonia production.

Negative effect of Fe doping on ammonia synthesis. As for the ORR and the CO2RR, the

co-doping of transition metals including Fe, Co, or Ni with N into carbon can dramatically increase

21
catalyst activity [52, 53, 62, 63]. Here, we also studied the effect of Fe doping on the NRR activity

toward the ESA. The doping content of Fe into ZIFs was increased from 1 to 20 at. % against the

total metal including Zn and Fe. According to our previous work, the FeN4 moiety is very likely

embedded into the carbon lattice [52]. To investigate whether Fe-N4 could be an active site for the

NRR, we studied the ESA on the Fe-ZIF catalyst as a function of Fe content. Figure S12a-c shows

the LSV for the Fe-ZIF catalysts. As the Fe content is gradually increased from 1 to 20 at %, the

current density during the LSV measurement for Ar saturated electrolyte gradually increased and

became higher than that in N2, suggesting enhanced HER activity. The current densities against

time at -0.3 V vs RHE for 20Fe-ZIF-1100-1h and “Fe-free” ZIF-1100-1h were further compared in

Figure S12d and 12e. Unlike the C-ZIF-1100-1h catalyst showing higher current density in N2

electrolyte, the current density in Ar for 20Fe-ZIF-1100-1h catalysts is much higher than that in

N2. Accordingly, ammonia production rates determined with these Fe-doped catalysts is nearly

two orders of magnitude lower relative to “Fe-free” C-ZIF catalysts. Also, increasing Fe content

leads to a reduction of NH3 production rates. The negative effect from Fe doping probably is due to

the enhancement of the HER, as evidenced by the increased current density in Ar-saturated

electrolytes. The proton is likely more easily absorbed than N2 on the FeN4 sites. Another possible

reason for the reduced NH3 production rates is the poisoning of Fe sites by the NH-like

intermediates. From the XPS analysis of Fe-MOFs that we reported earlier [52], the content of

pyridinic nitrogen is three times lesser than graphitic nitrogen. Thus, Fe ions may coordinate with

pyridinic N in carbon defects, thereby blocking the potential active sites.


22
Ammonia synthesis in different electrolytes. It was reported that alkali metals play a

promotional role in enhancing the NRR activity [9, 64]. Here we compared two cations in alkaline

solutions, 0.1 M KOH and 0.1 M NaOH, by using the best performing C-ZIF-1000-1h catalyst.

Though the current density in Ar is similar in both electrolytes, the current density in N2 is higher

for KOH than that in NaOH. After one-hour potentiostatic tests under identical conditions, a much

higher absorbance is observed in 0.1 M KOH, indicating a higher production rate of NH3 (Figure

4a). Similar difference in production rates between the two electrolytes is also observed for other

catalysts such as Fe-doped ZIF-1100-1h. Regardless of the type of catalysts or concentration of

alkaline electrolytes, the overall production rate can be up to one order of magnitude higher in

KOH as compared to NaOH shown in Figure 4(b). This indicates that KOH is a more favorable

electrolyte to generate NH3 relative to NaOH, suggesting the promotion role of K+. Detailed

discussion will be provided later in the modeling section.

23
Figure 4. (a) Comparison of absorbance spectra after potentiostatic tests for the CF-ZIF-1100-1h
catalyst in 0.1 M KOH and 0.1 M NaOH (b) Comparison of ammonia production rates for different
catalysts in various KOH and NaOH electrolytes. (c) The UV-vis spectra after potentiostatic tests
in N2 and Ar for CF-ZIF-1100-1h at -0.3 V vs RHE at 20, 40, and 60 oC in 0.1 M KOH at -0.3 V vs
RHE. (d) Comparison of ammonia production rates in N2 and Ar for CF-ZIF-1100-1h at -0.3 V vs
RHE at 20, 40 and 60 oC in 0.1 M KOH and 0.1 M HCl electrolytes.

Though our focus was to develop catalysts for room temperature and ambient pressure

conditions, we studied the effect of temperature on the ESA production rates based on aqueous 0.1

M KOH electrolyte. In principle, nitrogen reduction is favored at higher temperatures and

24
pressures. To test this, we varied the testing temperatures from 25 to 60oC. As the temperature

increased, the difference of current density between N2 and Ar is also increased, corresponding to

increased NH3 production rates from 3.4×10-6 at 25oC to 7.3×10-6 mol cm-2 h-1 at 60oC. The

corresponding UV-vis absorbance is shown in Figure 4(c). According to Arrhenius equation,

activation energy of the best performing C-ZIF-1000-1h catalyst was determined to be 7.4 kJ/mol

(Figure S13). Further, NH3 was not detected in the catalyst tested in Ar at higher temperatures,

thus implying the excellent stability of the catalyst. Though ESA production rates are increased

with electrolyte temperature in 0.1 M KOH, the trend is exactly opposite in 0.1 M HCl (Figure 4d)

showing a lower production rate at 60oC. This confirms the prediction that HER rates are much

faster in acidic electrolytes relative to alkaline environment. Higher temperatures yielded

enhanced HER in acidic electrolytes, which is in good agreement with the increased current

density in Ar compared to N2. As a result, the production rate in 0.1 M HCl is nearly one order of

magnitude lower than that in 0.1 M KOH at -0.3 V vs RHE. A constant NH3 production rate is

retained during 10-hour stability tests in acid (Figure S14). Also, no NH3 was detected during the

stability test in the Ar-saturated 0.1 M HCl electrolyte. These results indicate good stability of the

C-ZIF-1100-1h catalyst during the NRR even in more corrosive acidic media. It should be noted

that, using alkaline electrolytes for studying the NRR, some of generated NH3 might be evaporated

from the electrolyte. Thus the measured production rates of NH3 on the ZIF-derived carbon

catalysts are likely underestimated.

25
3.3 DFT calculation to elucidate possible active sties and reaction pathway

Pyrolysis of ZIF-8 plays a key role in introducing nitrogen doping and creating carbon defects in

catalysts. It was observed by XPS that pyridinic nitrogen content decreases in the carbon

framework with an increase in heating temperature, whereas the graphitic nitrogen content

increases. We hypothesize that the possible active site for the NRR could be associated with such

doped nitrogen atoms and interstitial carbon vacancies forming an active moiety, which is able to

adsorb the N2 molecule for subsequent NN bond dissociation. To examine our hypothesis, we

performed DFT calculations to acquire a fundamental understanding of the chemical nature of

possible active sites and the possible reaction pathway on the modelled active site as shown in

Figure 5.

26
Figure 5. (a) Atomistic structure scheme showing the reaction pathway of the N2 reduction on the
N3 site. In (a), the blue, gray, and white balls represent N, C, and H atoms, respectively. (b)
Predicted free energy evolution of N2 reduction on an N3 site in N-doped carbon under electrode
potentials of 0 and -0.3 V.

As the first step, we conducted an extensive search for the possible active sites related to

pyridinic, pyrrolic, and graphitic nitrogen embedded in a graphene layer using the DFT

calculations (Figure S15). However, in this study, no other site was found, for which the

adsorption of the N2 molecule was energetically favourable except for the structure shown in
27
Figure 5. The only possible active site is derived from the pyridinic N3 moiety embedded in a

graphitic layer, which contains one protonated pyridinic nitrogen and one adjacent atomic

vacancy. The atomic vacancy might be the result of removal of one pyridinic nitrogen caused by

the heat treatment at higher temperature, which is in good agreement with reduced pyridinic N

content at 1100 oC. The DFT calculations also suggest that a N2 molecule could be favourably

attracted by the N vacancy (a defect) and thus initiate the N2 reduction reaction to form NH3. From

the predicted free energy diagram for the N2 reduction cycle under electrode potential of 0 V

(Figure 5, blue line) the free energy change for the N2 trapping step is 1.44 eV. Our calculation

shows that the trapped N2 forms a ring with the carbon matrix. The N-N bond length is predicted to

be 1.37 Å which is larger than 1.10 Å of the N-N bond length in N2 molecule. The elongation of the

N-N bond indicates that the N-N triple bond is significantly weakened in this step thus favouring

the breaking of the strong N-N bond. The subsequent protonation reactions are quite easy with

production of an NH3 molecule until the N4 moiety is regenerated. The N4 moiety is quite stable

such that the further protonation of the pyridinic N in the N4 moiety has a large free energy

increment of 1.75 eV. This free energy change decreases as the electrode potential decreases as

shown in Figure 5 (red line). With an applied potential of -0.3 V vs RHE, the free energy

requirement decreases to 1.45 eV. Here, the alkali metal such as K+ used in the electrolyte may

play a role in donating electrons, thereby driving the evolution of the second NH3 molecule

without adding even higher potential. Though we have not included the role of potassium in the

DFT calculations, its effect on favouring ammonia synthesis is evident from our experiments. As
28
significant graphitic N atoms exist in the ZIF-derived carbon catalysts, we also examined the

graphitic N and its adjacent C as possible active sites for N2 reduction using the DFT

computational method. The results indicate that N2 molecule cannot be adsorbed on either the

graphitic N or its adjacent C atoms. Favorable adsorption of N2 molecule is a necessary condition

for these sites to be active for N2 reduction. Consequently, we predict that the isolated graphitic N

embedded in a carbon layer does not contribute to the observed catalytic activity for N2 reduction.

However, it should be noted that graphitic N doping likely tunes electronic structure of carbon and

affects the adsorption energy between active sties and reactants. Deeper understanding on the role

of nitrogen doping and carbon in defects in promoting the NRR is one of our future focus.

Conclusions

In summary, a ZIF-derived carbon catalyst has exhibited very encouraging activity and stability

for the NRR during the NH3 electrosynthesis in alkaline KOH media. The crucial nitrogen doping

and carbon structures can be tuned by varying the pyrolysis conditions including temperature and

duration. Unlike other reactions such as ORR and CO2RR, doping of Fe into carbon in the form of

FeN4 cannot boost NRR activity in this work, likely due to blockage of N vacancy active sites and

promotion of the competitive HER activity. Thus, the best NRR activity is obtained for “Fe-free”

carbon catalysts, which were prepared from ZIF-8 precursors pyrolyzed at 1100 oC under N2 flow

for one hour. Using 0.1 M KOH as electrolyte at ambient conditions, the catalyst exhibited a

respectable production rate of NH3 up to 3.4×10-6 mol cm-2 h-1 with a Faradaic efficiency of 10.2 %

29
at -0.3 V vs RHE. The initial stability was also studied during an 18-hour continuous measurement

retaining constant NH3 production rates. Also, NH3 production rates is monotonically increased

with the temperature of 0.1 M KOH electrolyte, achieving 7.3×10-6 mol cm-2 h-1 at 60oC.

Accordingly, an activation energy for the NRR on the ZIF-derived catalyst is determined to be 7.4

kJ/mol. However, the temperature dependence in 0.1 M HCl electrolyte is opposite due to

significantly enhanced HER in acids at higher temperature. Our DFT calculations further elucidate

that the moiety containing carbon vacancies and three pyridinic N atoms embedded into carbon

planes could act as the active sites. The rate-determining step is the adsorption of N2. Reduced

reaction potential is favorable for further reducing the activation energy for the protonation

process for releasing the second NH3 molecule. The alkali metal ions such as K+ also found

playing a promotional role in donating electrons during the NRR. Unlike conventional NRR

catalysts, which use nitrides and oxides, the carbon catalysts derived from well-defined ZIF

precursors with optimal nitrogen doping and carbon defects can provide new insight to design

effective catalysts for efficient electrochemical synthesis of ammonia.

Acknowledgements

This work was financially supported by start-up funding from the University at Buffalo, SUNY

along with U.S. Department of Energy’s Advanced Research Projects Agency-Energy (ARPA-E)

office’s REFUL program. Electron microscopy conducted at Oak Ridge National Laboratory's

Center for Nanophase Materials Sciences, which is a U.S. Department of Energy, Office of

30
Science User Facility. The computation were carried out on the computer facility at Center for

Simulation and Modeling of the University of Pittsburgh and at the Extreme Science and

Engineering Discovery Environment (XSEDE), which is supported by National Science

Foundation grant number ACI-1053575.

References

[1] M.D. Fryzuk, Nature, 427 (2004) 498.

[2] L. Green, International Journal of Hydrogen Energy, 7 (1982) 355-359.

[3] M. Jewess, R.H. Crabtree, ACS Sustainable Chemistry & Engineering, 4 (2016) 5855-5858.

[4] S. Giddey, S.P.S. Badwal, C. Munnings, M. Dolan, ACS Sustainable Chemistry &

Engineering, 5 (2017) 10231-10239.

[5] Y. Bicer, I. Dincer, G. Vezina, F. Raso, Environmental management, 59 (2017) 842-855.

[6] S.-J. Li, D. Bao, M.-M. Shi, B.-R. Wulan, J.-M. Yan, Q. Jiang, Advanced Materials, 29 (2017)

1700001-n/a.

[7] G.-F. Chen, X. Cao, S. Wu, X. Zeng, L.-X. Ding, M. Zhu, H. Wang, Journal of the American

Chemical Society, 139 (2017) 9771-9774.

[8] S. Chen, S. Perathoner, C. Ampelli, C. Mebrahtu, D. Su, G. Centi, Angewandte Chemie, 129

(2017) 2743-2747.

[9] S. Chen, S. Perathoner, C. Ampelli, C. Mebrahtu, D. Su, G. Centi, ACS Sustainable Chemistry

& Engineering, 5 (2017) 7393-7400.

31
[10] L.M. Duman, W.S. Farrell, P.Y. Zavalij, L.R. Sita, Journal of the American Chemical

Society, 138 (2016) 14856-14859.

[11] M. Lashgari, P. Zeinalkhani, Applied Catalysis A: General, 529 (2017) 91-97.

[12] M. Iwamoto, M. Akiyama, K. Aihara, T. Deguchi, ACS Catalysis, 7 (2017) 6924-6929.

[13] H. Davy, Philosophical Transactions of the Royal Society of London, 97 (1807) 1-56.

[14] I.A. Amar, C.T. Petit, L. Zhang, R. Lan, P.J. Skabara, S. Tao, Solid State Ionics, 201 (2011)

94-100.

[15] W. Sheng, Z. Zhuang, M. Gao, J. Zheng, J.G. Chen, Y. Yan, Nature Communications, 6

(2015) 5848.

[16] S. Back, Y. Jung, Physical Chemistry Chemical Physics, 18 (2016) 9161-9166.

[17] D. Bao, Q. Zhang, F.-L. Meng, H.-X. Zhong, M.-M. Shi, Y. Zhang, J.-M. Yan, Q. Jiang, X.-B.

Zhang, Advanced Materials, 29 (2017) 1604799-n/a.

[18] B.M. Lindley, Q.J. Bruch, P.S. White, F. Hasanayn, A.J. Miller, Journal of the American

Chemical Society, 139 (2017) 5305-5308.

[19] C.J. Pickett, J. Talarmin, Nature, 317 (1985) 652.

[20] M.-M. Shi, D. Bao, B.-R. Wulan, Y.-H. Li, Y.-F. Zhang, J.-M. Yan, Q. Jiang, Advanced

Materials, 29 (2017) 1606550-n/a.

[21] Y. Abghoui, A.L. Garden, J.G. Howalt, T. Vegge, E. Skulason, ACS Catalysis, 6 (2015)

635-646.

[22] Y. Abghoui, E. Skúlason, The Journal of Physical Chemistry C, 121 (2017) 6141-6151.

32
[23] Á.B. Höskuldsson, Y. Abghoui, A.B. Gunnarsdóttir, E. Skúlason, ACS Sustainable

Chemistry & Engineering, 5 (2017) 10327-10333.

[24] H.-L. Jiang, B. Liu, Y.-Q. Lan, K. Kuratani, T. Akita, H. Shioyama, F. Zong, Q. Xu, Journal

of the American Chemical Society, 133 (2011) 11854-11857.

[25] K. Shen, X. Chen, J. Chen, Y. Li, ACS Catalysis, 6 (2016) 5887-5903.

[26] G. Wu, A. Santandreu, W. Kellogg, S. Gupta, O. Ogoke, H. Zhang, H.-L. Wang, L. Dai, Nano

Energy, 29 (2016) 83-110.

[27] H. Zhang, H. Osgood, X. Xie, Y. Shao, G. Wu, Nano Energy, 31 (2017) 331-350.

[28] X. Mao, T.A. Hatton, Industrial & Engineering Chemistry Research, 54 (2015) 4033-4042.

[29] T. Ma, Q. Fan, H.C. Tao, Z.S. Han, M.W. Jia, Y.N. Gao, W.J. Ma, Z.Y. Sun, Nanotechnology,

28 (2017).

[30] A. Vasileff, Y. Zheng, S.Z. Qiao, Advanced Energy Materials, 7 (2017)

10.1002/aenm.201700759.

[31] T. Su, Q. Shao, Z. Qin, Z. Guo, Z. Wu, ACS Catalysis, 8 (2018) 2253-2276.

[32] T. Wang, H. Xie, M. Chen, A. D'Aloia, J. Cho, G. Wu, Q. Li, Nano Energy, 42 (2017) 69-89.

[33] Y. Liu, Y. Su, X. Quan, X. Fan, S. Chen, H. Yu, H. Zhao, Y. Zhang, J. Zhao, ACS Catalysis,

8 (2018) 1186-1191.

[34] L. Zhang, W. Liu, Y. Dou, Z. Du, M. Shao, The Journal of Physical Chemistry C, 120 (2016)

29047-29053.

[35] L. Zhang, J. Xiao, H. Wang, M. Shao, ACS Catalysis, 7 (2017) 7855-7865.

33
[36] Y. Song, W. Chen, C. Zhao, S. Li, W. Wei, Y. Sun, Angewandte Chemie International

Edition, 56 (2017) 10840-10844.

[37] R. Li, Z. Wei, X. Gou, ACS Catalysis, 5 (2015) 4133-4142.

[38] S.N. Zhao, X.Z. Song, S.Y. Song, H.J. Zhang, Coordination Chemistry Reviews, 337 (2017)

80-96.

[39] X. Zhao, F. Yin, N. Liu, G. Li, T. Fan, B. Chen, Journal of Materials Science, 52 (2017)

10175-10185.

[40] H. Verdouw, C. Van Echteld, E. Dekkers, Water Research, 12 (1978) 399-402.

[41] W. Bolleter, C. Bushman, P.W. Tidwell, Analytical Chemistry, 33 (1961) 592-594.

[42] G. Kresse, J. Furthmuller, Comput. Mater. Sci., 6 (1996) 15-50.

[43] J. Zhao, L. Wu, C. Zhan, Q. Shao, Z. Guo, L. Zhang, Polymer, (2017).

[44] G. Kresse, D. Joubert, Phys. Rev. B, 59 (1999) 1758-1775.

[45] B. Hammer, L.B. Hansen, J.K. Nørskov, Phys. Rev. B, 59 (1999) 7413-7421.

[46] J.K. Nørskov, J. Rossmeisl, A. Logadottir, L. Lindqvist, J.R. Kitchin, T. Bligaard, H. Jonsson,

Journal of Physical Chemistry B, 108 (2004) 17886-17892.

[47] Y. Zhu, J. Ciston, B. Zheng, X. Miao, C. Czarnik, Y. Pan, R. Sougrat, Z. Lai, C.-E. Hsiung, K.

Yao, I. Pinnau, M. Pan, Y. Han, Nature Materials, 16 (2017) 532.

[48] G. Wu, C.M. Johnston, N.H. Mack, K. Artyushkova, M. Ferrandon, M. Nelson, J.S.

Lezama-Pacheco, S.D. Conradson, K.L. More, D.J. Myers, P. Zelenay, Journal of Materials

Chemistry, 21 (2011) 11392-11405.

34
[49] G. Wu, N.H. Mack, W. Gao, S. Ma, R. Zhong, J. Han, J.K. Baldwin, P. Zelenay, ACS Nano,

6 (2012) 9764–9776.

[50] S. Gupta, S. Zhao, X.X. Wang, S. Hwang, S. Karakalos, D. Su, H. Xu, G. Wu, ACS Catalysis,

7 (2017) 8386–8393.

[51] C. Cheng, R. Fan, Z. Wang, Q. Shao, X. Guo, P. Xie, Y. Yin, Y. Zhang, L. An, Y. Lei,

Carbon, 125 (2017) 103-112.

[52] H. Zhang, S. Hwang, M. Wang, Z. Feng, S. Karakalos, L. Luo, Z. Qiao, X. Xie, C. Wang, D.

Su, Y. Shao, G. Wu, Journal of the American Chemical Society, 139 (2017) 14143-14149.

[53] F. Pan, H. Zhang, K. Liu, D. Cullen, K. More, M. Wang, Z. Feng, G. Wang, G. Wu, Y. Li,

ACS Catalysis, (2018) 3116-3122.

[54] W. Xia, A. Mahmood, R. Zou, Q. Xu, Energy & Environmental Science, 8 (2015) 1837-1866.

[55] W. Xia, J. Zhu, W. Guo, L. An, D. Xia, R. Zou, Journal of Materials Chemistry A, 2 (2014)

11606-11613.

[56] Q.L. Zhu, W. Xia, T. Akita, R. Zou, Q. Xu, Advanced Materials, 28 (2016) 6391-6398.

[57] S. Gupta, L. Qiao, S. Zhao, H. Xu, Y. Lin, S.V. Devaguptapu, X. Wang, M.T. Swihart, G. Wu,

Advanced Energy Materials, 6 (2016) 1601198.

[58] H. Sheng, M. Wei, A. D’Aloia, G. Wu, ACS Applied Materials & Interfaces, 8 (2016)

30212-30224.

[59] X. Wang, Q. Li, H. Pan, Y. Lin, Y. Ke, H. Sheng, M.T. Swihart, G. Wu, Nanoscale, 7 (2015)

20290-20298.

35
[60] Q. Li, P. Xu, W. Gao, S. Ma, G. Zhang, R. Cao, J. Cho, H.-L. Wang, G. Wu, Advanced

Materials, 26 (2014) 1378-1386.

[61] Q. Li, H. Pan, D. Higgins, R. Cao, G. Zhang, H. Lv, K. Wu, J. Cho, G. Wu, Small, 11 (2015)

1443-1452.

[62] X.X. Wang, D.A. Cullen, Y.-T. Pan, S. Hwang, M. Wang, Z. Feng, J. Wang, M.H. Engelhard,

H. Zhang, Y. He, Y. Shao, D. Su, K.L. More, J.S. Spendelow, G. Wu, Advanced Materials, 30

(2018) 1706758.

[63] Z. Qiao, H. Zhang, S. Karakalos, S. Hwang, J. Xue, M. Chen, D. Su, G. Wu, Applied

Catalysis B: Environmental, 219 (2017) 629-639.

[64] J. Kong, A. Lim, C. Yoon, J.H. Jang, H.C. Ham, J. Han, S. Nam, D. Kim, Y.-E. Sung, J. Choi,

H.S. Park, ACS Sustainable Chemistry & Engineering, 5 (2017) 10986-10995.

36
Shreya Mukherjee received her M.S. degree from Inidan Institute of Technology, Guwahati in
2016. She joined Ph.D. program under the supervision of Dr. Gang Wu in the department of
Chemical and Biological Engineering at University of Buffalo in fall 2016. Her current research is
the development of high-efficient and low cost electrocatalysts for sustainable synthesis of
ammonia.

Dr. David A. Cullen is the R&D Staff Scientist and Team Leader in the Electron Microscopy
Group of the Materials Science and Technology Division at Oak Ridge National Laboratory. He
is also a joint faculty member in the Bredesen Center for Interdisciplinary Research and Graduate
Education at the University of Tennessee, Knoxville. He received a B.S. degree in Applied
Physics from Brigham Young University, and a Ph.D. in Materials Science and Engineering from
Arizona State University. He joined ORNL in 2010 as an Alvin M. Weinberg Fellow. His current
research interests are focused on the advanced characterization of catalysts by electron
microscopy.

Stavros Karakalos received his MS and PhD degrees from University of Patras (2009). He has
worked as a Postdoctoral Associate at FORTH/ICE-HT, University of California-Riverside,

37
Harvard University, and as a visiting scientist at Fermi Elettra Synchrotron, Italy and at the
Advanced Light Source, Berkeley. He joined the University of South Carolina (USC) in 2016,
where he presently is a Research Assistant Professor in Chemical Engineering and the Director of
the USC XPS Facility. His research interests are in the field of surface science and heterogeneous
catalysis.

Kexi Liu received his Ph.D. in Materials Science from the University of Pittsburgh under the
supervision of Prof. Guofeng Wang in 2018. His research focuses on understanding and predicting
the catalytic activity of electrocatalysts for energy conversion using the first-principles density
functional theory calculation methods.

Hao Zhang received the B.S. degree (2014) in Chemical Engineering at China University of
Petroleum (East China). He obtained the M.S. degree (2017) at Tianjin University with a major
research on defect-engineered materials for photocatalytic applications. He is now a Ph.D. student
in Dr. Gang Wu’s group at University at Buffalo, SUNY. His research focuses on the
electrocatalytic and photocatalytic NH3 synthesis at ambient conditions.

38
Shuai Zhao, received her Ph.D. in Chemical Engineering from the University of Connecticut in
2016. She then joined Giner, Inc as a project scientist. Her research focuses on electrochemical
catalyst development for acidic and alkaline fuel cells and electrolyzers, Li-ion batteries, ammonia
synthesis, etc., as well as system optimization, hydrogen storage.

Hui Xu is Technical Director of Energy Conversion Materials at Giner Inc. He earned his Ph.D. in
chemical engineering from University of Connecticut, where he studied oxygen reduction
mechanisms of platinum and other metal alloy catalysts. He subsequently pursued his postdoctoral
studies at Los Alamos National Laboratory, working on advanced electrode design and structure
characterization for PEM fuel cells. His research scope ranges from fuel cells, electrolyzes,
batteries to electrochemical capacitors. Dr. Xu is the principal investigator of multiple DOE, DOD,
USDA, and industrial projects with a total funding more than 15 million dollars.

39
Dr. Karren L. More is the Group Leader for the Electron and Atom Probe Microscopy Group in
the Center for Nanophase Materials Sciences (CNMS) at Oak Ridge National Laboratory (ORNL).
She received her Ph.D. from North Carolina State University and has been a research staff member
at ORNL since 1988, first in the Materials Science and Technology Division, and more recently
(since 2013) at the CNMS. Her research interests are focused on using advanced electron
microscopy towards understanding the structure and chemistry of nanomaterials, especially
related to performance, stability, and durability.

Guofeng Wang is currently an associate professor in the Department of Mechanical Engineering


and Materials Science at the University of Pittsburgh. He receives his Ph.D. in Materials Science
from California Institute of Technology. His research focuses on development and application of
advanced computational methods for rational design of high-performance electrocatalysts with use
in energy conversion and energy storage technologies. His research has been supported by
National Science Foundation (NSF) and Department of Energy (DOE). He has published more
than 80 journal articles, book chapters, and refereed conference proceedings.

40
Gang Wu is an assistant professor in the Department of Chemical and Biological Engineering
at the University at Buffalo (SUNY-Buffalo). He completed his Ph.D. studies at the Harbin
Institute of Technology in 2004 followed by extensive postdoctoral trainings at Tsinghua
University (2004-2006), the University of South Carolina (2006-2008), and Los Alamos
National Laboratory (LANL) (2008-2010). Then he was promoted to be a staff scientist at
LANL until he joined SUNY-Buffalo in 2014. His research focuses on functional materials
and catalysts for electrochemical energy storage and conversion. To date, he has written more
than 170 papers with a citation of >12,500 times.

41
Highlights

 A highly disordered carbon exhibits a remarkable production rate of NH3


 DFT elucidates the possible N3 active sites consisting of three pyridinc N
 Introduction of Fe leads to negative effect on NH3 synthesis
 K+ ions have promotional role during the NH3 synthesis

42

You might also like