radiofrequency plasma synthesis of ammonia over ni-mof-74

20
1 Radiofrequency Plasma Synthesis of Ammonia over Ni-MOF-74 Javishk Shah , Ting Wu , Jolie Lucero , Moises A. Carreon and Maria L. Carreon, , * Russell School of Chemical Engineering, The University of Tulsa, 800 S Tucker Dr, Tulsa, OK 74104, US Department of Chemical and Biochemical Engineering, Colorado School of Mines, 1500 Illinois St, Golden, CO 80401, US *[email protected] Keywords: Plasma Catalysis, Ni-MOF-74, Ammonia Synthesis, Metal-organic Frameworks, Non-Thermal Plasma. Herein, we demonstrate a synergistic approach consisting on radiofrequency plasma to synthesize ammonia in the presence of Ni-MOF-74 as catalyst. The Ni-MOF displayed higher ammonia yields as com-pared to the pure Ni metal. Specifically, ammonia yields as high as 0.23 g-NH3 (kWh-g-catalyst) -1 and energy cost of 265 MJ mol -1 over Ni-MOF were observed. The enhanced catalytic activity of the Ni-MOF in the presence of plasma was attributed to the presence of pores that improved mass transfer of guest and product molecules during reaction, the presence of open Ni metal sites, and lower surface hydrogen re-combination. Furthermore, the ammonia energy

Upload: others

Post on 08-Apr-2022

4 views

Category:

Documents


0 download

TRANSCRIPT

1

Radiofrequency Plasma Synthesis of Ammonia over

Ni-MOF-74

Javishk Shah†, Ting Wu‡, Jolie Lucero‡, Moises A. Carreon‡ and Maria L. Carreon, †,*

†Russell School of Chemical Engineering, The University of Tulsa, 800 S Tucker Dr, Tulsa, OK

74104, US

‡Department of Chemical and Biochemical Engineering, Colorado School of Mines, 1500

Illinois St, Golden, CO 80401, US

*[email protected]

Keywords: Plasma Catalysis, Ni-MOF-74, Ammonia Synthesis, Metal-organic Frameworks,

Non-Thermal Plasma.

Herein, we demonstrate a synergistic approach consisting on radiofrequency plasma to synthesize

ammonia in the presence of Ni-MOF-74 as catalyst. The Ni-MOF displayed higher ammonia

yields as com-pared to the pure Ni metal. Specifically, ammonia yields as high as 0.23 g-NH3

(kWh-g-catalyst)-1 and energy cost of 265 MJ mol-1 over Ni-MOF were observed. The enhanced

catalytic activity of the Ni-MOF in the presence of plasma was attributed to the presence of pores

that improved mass transfer of guest and product molecules during reaction, the presence of open

Ni metal sites, and lower surface hydrogen re-combination. Furthermore, the ammonia energy

2

yield of our plasma-Ni MOF catalyst is superior to those of the state-of-the-art RF plasma catalytic

systems.

INTRODUCTION

Ammonia, a chemical discovered over 100 years ago has found its way into every aspect of

human life. At some stage in the food, fertilizer, textile, cosmetic and chemical industry ammonia

has been used. Besides such heavy use in various industries, ammonia is found in regular

household as a water treatment and cleaning agent. Ammonium nitrate, ammonium phosphate,

pigments, urea, industrial solvents and coolants are a few compounds manufactured from

ammonia. Recently, ammonia has gained attention as fuel and a safer hydrogen storage alternative

due to its high hydrogen density as compared to metal hydrides.1

The Haber-Bosch process is the benchmark technology for ammonia production.2 Typically, this

is an energy intensive process which conventionally operates at a pressure of 150-250 bar and

temperature of 400-500 °C in the presence of an iron-based catalyst.3 The anticipated 2018

ammonia production is projected to be 186 million metric tonnes.4 A chemical used at such large

scale also accounts to a share of more than 1% of global electricity consumption5 as well as 2-3%

of total greenhouse gas emissions.6 Most of the energy in the Haber-Bosch process is spent in

excitation and breakage of the triple nitrogen bond, the rate limiting step.7 Due to the

thermodynamic feasibility of the process, a maximum of 15 % yield can be obtained. A constant

recycle of the unreacted product is required to make the process economical.8 In principle, this

process can be made less energy intensive by (a) employing a highly active catalyst and/or (b)

aiding the process at low temperature.

3

Plasma, the fourth state of matter where the molecules or atoms exist as excited species and ions

can help to dissociate the strong triple nitrogen bond. In the past few decades, non-thermal plasma

technology has been used for various gas-phase chemical synthesis such as ozone,9 VOC

removal,10 NOx treatment,11 among others. Furthermore, various plasmas such as dielectric

barrier,12-17 cascading arc,18 microwave19-20 and radiofrequency discharge21 have been used for

ammonia synthesis. Most of these reports are either non-catalytic or employ a metal loaded on a

porous support such as alumina or silica.22 The metal acts as the active phase, while the support

material is usually a dielectric, not having much effect on the yield of ammonia. In particular,

Nickel has been used as an active metal for the synthesis of ammonia. For instance, Mizushima et

al. employed a Ni-loaded alumina membrane for ammonia synthesis in a Dielectric Barrier

Discharge (DBD) reactor.23 In general, the activity of Ni was better than Fe which is a traditional

catalyst for ammonia synthesis. In this report, it was proposed that Ni adsorbed the hydrogen which

then reacted with nitrogen on the alumina support membrane.23 Akay and Zhang employed

Ni/SiO2 filled in BaTiO3 cavities as catalytic promoters for ammonia synthesis in a DBD reactor.

The energy cost was decreased about 20% with this catalyst. However, if the mass of catalyst was

increased above 5 wt% the energy cost increased.24 This can be due to decrease in the mass or void

volume of dielectric material required to sustain the plasma. A potential way to alleviate this issue

is employing high surface area catalysts. In this respect, Metal organic frameworks (MOFs)

combine highly desirable properties that make them ideal candidates for diverse catalytic

applications,25 such as uniform crystallinity, microporosity high surface area,26 and high thermal

and chemical stability.27-28 In principle, the open and unsaturated metal cations or metal-based-

clusters can act as the catalytic active species, and the ordered microporous crystalline structure

can serve as the pathway for guest and product molecules to effectively diffuse with enhanced

4

mass transfer.29-31 Besides plasma catalysis, electrochemical reduction has also gained attention

for the reduction of nitrogen to ammonia at ambient conditions. Liu et al. in their review lists the

use of Co and Mo catalysts, which are shown to have high potential for ammonia synthesis.32 The

catalysts mediate the reaction by breaking the nitrogen triple bond.33 These results were also

corroborated computationally by Martirez et al.33 and experimentally by Tusji et al.34 Yan’s group

currently working on electrochemical synthesis of ammonia employed nanomaterials such as

PdCu nanocluster on Graphene,35 ceria-reduced graphene oxide hybrid,36 Au nanoclusters on

titania37 and Au nanorods as catalysts.38 Ammonia as well as hydrazine evolved as products. The

peak ammonia yield of 21.3 µg-NH3 h-1 mgcat

-1 was obtained with a Faradaic efficiency of 10%.37

The alternative pathway to electrocatlaytic reduction was elucidated using DFT while Au

nanoclusters were used as catalysts. Interestingly, the pathway changed with the plane orientations

of the Au catalysts. Au(310) catalyzed reaction was reported to have lower activation energy as

compared to Au(210) catalyzed nitrogen reduction reaction.38

Herein we demonstrate the use of a Nickel based MOF (Ni-MOF-74) as catalyst for the synthesis

of ammonia via non thermal plasma catalysis. In principle, the excited formed plasma species may

interact with the Ni open metal sites, and MOF pores may be beneficial for enhanced mass transfer

leading to improved ammonia yields.

EXPERIMENTAL SECTION

Synthesis of Ni-MOF-74

Ni-MOF-74 consists of nickel (II) oxide chains connected by 2,5-dihydroxyterephthalte linkers,

displaying one-dimensional parallel hexagonal channels with pore size area of 1.03×0.55 nm2.39-

42 The nickel oxide chains form a three-dimensional hexagonal packing of helical O5Ni, where 5

5

oxygen atoms are from the terephthalic and hydroxy group on the ligand and one free-site atom

that was originally occupied by a solvent or water molecule, resulting in an activated stable

structure with open metal sites providing binding sites for guest gas molecules.40-43

The synthesis approach to prepare Ni-MOF-74 was modified from a previous report described

elsewhere.39 In a typical synthesis, a 1:1:1 (v/v/v) mixture of 10.5 mL dimethylformamide (DMF,

Acros, 99.9%), 10.5 mL ethanol (Fisher, 99.99%), and 10.5 mL DI water was placed in one beaker

to form the mixed solvent solution. 1.503 g of nickel (II) nitrate hexahydrate (Sigma-Aldrich,

≥97%) and 0.502 g of 2,5-dihydroxyterephthalic acid (H4DHTP, Sigma-Aldrich, 98%) were

placed in another beaker. Then the nickel source and the ligand H4DHTP were added to the solvent

mixture. The solution was dissolved using a magnetic stirrer until the ligand and the nickel source

were completely dissolved. The clear solution was transferred to a 45 mL Teflon lined autoclave

(General Acid Digestion Vessel 4744, Parr Instrument). The Ni-MOF-74 crystals synthesis was

carried out in a conventional oven at 150 °C for 24 h with heating and cooling rate of 10 °C min-1

and 0.1°C min-1 respectively. After the autoclave cooled down to room temperature, the clear

portion of the solvent was decanted from the solution, replaced by methanol, and centrifuged at

4000 rpm for 5 minutes so that the crystals could be collected. Then the collected crystals were

washed with methanol and centrifuged and collected two or three additional times. The resulting

crystals were dried under vacuum at 250 °C for 5 h.

Catalyst Characterization

Ni-MOF-74 crystals were characterized with X-ray diffraction (PXRD) using a Siemens

Kristalloflex 810 diffractometer at 30 kV and 25 mA with Cu Kα1 radiation (λ=1.54059 Å). A

Micromeritics ASAP 2020 automated system was used to measure the sample’s Brunauer–

Emmett–Teller surface areas (BET) by N2 physisorption. Scanning electron microscope (SEM,

6

JEOL JSM-7000F) was used to inspect the morphology of the crystals. The Temperature

Programmed Desorption experiments were performed on a Perkin Elmer Pyris I

Thermogravimetric Analyzer. The sample was degassed at 150 °C overnight to remove any

molecules trapped on the pores. Ammonia was adsorbed on Ni-MOF-74 after degassing. The

sample was then heated from 50 °C to 500 °C with a heating ramp of 5 °C/min.

Plasma Ammonia Synthesis Experiments

The experiments were performed in an in-house built plasma reactor (Figure S1). The reaction

was conducted by introducing nitrogen (Praxair, 99%) and hydrogen (Praxair, 99.99%) at a 1:4

N2:H2 ratio to the reaction chamber using mass flow controllers. The nitrogen and hydrogen flow

rates were 4 and 16 sccm, respectively. The plasma was ignited using an RF Power Supply with a

Matching Network from Seren IPS, Inc. The vacuum pressure in the chamber is 0.3 torr with N2

and H2 flow. Ni-MOF and for comparison Ni metal were used as catalysts. The catalysts were

loaded into soda-lime glass pipette plugged with glass wool (Figure S2 and S3). The mass of each

catalyst was 0.2 g. The reactor was uniquely designed for ammonia synthesis by adding an on-line

Agilent 7820A gas chromatograph (GC), equipped with a gas sampling valve and HP-PlotQ

column (30 m x 0.32 mm x 20 µm). The gases were analyzed every 3 minutes for 60 minutes using

the GC. The experiments were performed at input powers of 100 W, 150 W, 200 W and 300W.

The light emitted from the discharge was lead through an optical system and the emission spectra

of the glow region were measured on the loaded catalyst. The measurements were recorded using

a dual channel UV-VIS-NIR spectrophotometer in scope mode (Avantes Inc., USB2000 Series).

The spectral range was from 200-1100 nm, using a line grating of 600 lines mm-1 and resolution

of 0.4 nm. A bifurcated fiber optic cable with 400 μm was employed.

RESULTS AND DISCUSSION

7

Figure 1a shows the XRD pattern of Ni-MOF-74 collected from the autoclave after synthesis.

For comparison, the simulated pattern of this MOF is shown. The XRD pattern of the synthesized

crystals match well the simulated XRD pattern39, 44, confirming the formation of Ni-MOF-74.

Figure 1b shows a representative SEM picture of Ni-MOF-74. Elongated ~ 2 x 8-10 micron

crystals are observed. Ni-MOF-74 crystals displayed BET surface area of ~723 m2 g-1, in well

agreement with previous reports.39, 43

Figure 1. (a) XRD pattern and (b) representative SEM image of synthesized Ni-MOF-74 crystals

employed as catalyst in this study (inset – SEM image with higher magnification).

8

The radiofrequeancy plasma-catalysis experiments were carried out at constant flow rate and gas

ratio at different power. The temperature increased with power. The highest temperature reached

was 82.3 ℃ at 300 W. To understand the catalytic effect of Ni-MOF-74, the catalytic activity was

compared with catalytic activity of glass wool bed (no catalyst) and Ni pieces on glass wool bed.

The catalytic activity results as a function of plasma power and catalyst are summarized in Figure

2.

Irrespective of the presence or absence of the catalyst, the ammonia yield increased with plasma

power. The trend reached almost a steady value in case of glass wool and Ni catalyst after 15

minutes. When Ni-MOF-74 was employed as a catalyst, the ammonia yield follows an

Figure 2. Ammonia yield (%) vs. time (min) as a function of plasma power for all studied catalysts,

a) 100 W, b) 150 W, c) 200 W and d) 300 W.

9

asymmetrical bell type behavior. The ammonia yield reached a maximum between 9 to 15 minutes,

then started to decrease steadily. The highest observed ammonia yield achieved was ~10.3%

corresponding to the Ni-MOF-74 catalyst at 300 W. The standard deviation was less than 10% in

all cases after 20 minutes of reaction time except for recycled Ni-MOF-74 at 300 W.

The general trend of catalytic activity was Ni-MOF (Fresh) > Ni-MOF (Recycle) > Ni > Glass

Wool except at 300 W. It is evident from these results that the MOF is more active than the pure

Ni metal. This can be (at least in part) attributed to the presence of pores that can improve mass

transfer of guest and product molecules during reaction. The presence of pores has been

demonstrated to be beneficial to increase the catalytic activity in plasma systems. For instance, it

has been reported that for several zeolite (microporous crystalline materials) catalysts, when

exposed to plasma there is an enhancement on the catalytic activity.45-49 It has been postulated that

the presence of pores tend to form micro-arc discharges in plasma reactors leading to an improved

catalytic activity.50-51

At 100 W, the catalytic activity of Ni-MOF-74 and Ni is comparable but is much better than

only glass wool. This can be attributed to the limited excited species available to interact with the

catalyst at this low plasma power. Therefore, the catalyst does not have a major impact at this

power. At 150 and 200 W, the glass wool and Ni have similar activity but Ni-MOF-74 exhibited

and enhanced behavior, almost 1.5 times better. At 300 W, the order of the catalytic activity

changes. The fresh Ni-MOF displays the highest ammonia yields at this power but there is steep

decrease in the activity of recycled Ni-MOF.

The turnover frequency (TOF) vs. time has been plotted for various powers in Figure S4. The

TOF did vary for various powers. The catalyst reaches its peak TOF around 12-15 minutes for all

plasma powers and the starts to deactivate. The TOF for fresh and recycled MOFs are quite similar

10

for low powers (100 W, 150 W and 200 W) while at 300 W there’s a significant difference. As it

is seen from Figure 2 that the reaction achieves steady state after 60 minutes we compared the

TOF of fresh and recycled catalyst at 100 W, 150 W and 200 W. The catalyst at 300 W is not

considered as the recycled MOF loses the crystallinity mid-way of the reaction. The TOF at 60

minutes for pure Ni at 100, 150 and 200 W is 0.07, 0.077 and 0.83 s-1 whereas the TOF for Ni-

MOF are 0.2, 0.24 and 0.25 s-1 at 100 W, 150 W and 200 W, respectively. The molar ratio of Ni

in Ni:Ni-MOF is 3:1 (approx.). There seems to be deactivation after the average turnover number

of 812.92, 1023.79 and 1033.88. The catalysts produced a total of 5.69, 6.92 and 7.55 µmol-NH3

before deactivating. The percentage deactivation of catalyst (Table S1) for a 60 minute run is

24.5%, 27%, 28.8% and 40.42% for fresh Ni-MOF-74 at plasma powers of 100 W, 150 W, 200 W

and 300 W, respectively. The number is bit under-determined at the ammonia absorbed in the

MOF is not added to it. As per the TPD experiments (Figure S5), a maximum ammonia loading

of 3.14 mmol-NH3/g-MOF is obtained on a dry basis which is in accordance with Glover et al.’s

reported value of 2.9 mmol-NH3/g-MOF.52 A total of 0.624 mmol-NH3 can be absorbed in the 0.2

g of loaded catalyst. The deactivation can merely be an absorption of the ammonia in the MOF.

The catalyst might be retaining its original activity but the decrease in ammonia signal on the GC

can be due to the trapping of ammonia in the pores.

Figure 3a-d shows representative SEM images of the recycled catalysts. These images suggest

that the shape and size of Ni-MOF-74 are preserved after plasma treatment, irrespective of the

plasma power. However, XRD patterns of the recycled MOF samples (Figure 3e) indicate that

there is a slight XRD peak displacement as a function of plasma power that may be related to small

changes in the unit cell volume of the framework. Specifically, there is a displacement to higher 2

theta angles indicating a decrease in the interplanar spacing. This decrease in interplanar spacing

11

suggests framework contraction, and therefore a higher degree of structural local disorder within

the porous framework. The degree of structural local disorder increased with plasma power. This

is supported by the fact that for low plasma powers (100, 150, and 200 W), two well defined sharp

XRD peaks are evident. For higher plasma powers (300 W) only one broad peak is observed. The

XRD peak displacement may be caused by internal stress within the porous framework developed

by plasma irradiation. The presence of one broader peak for the 300 W sample suggests onset of

framework amorphization. Therefore, the limited stability of the Ni-MOF catalysts at 300 W

(Figure 2d) may be related to higher degree of structural local disorder, and amorphization. The

local structural distortion in the Ni-MOF may occur due to the partial damage occurring from

plasma species such as excited electrons and non-reactive species. As the energy from the species

is transferred to a bond or atom in the structure due to collision, the molecular structure undergo

minor transformation forming defects as reported in other common catalysts.53-54

For all studied plasma powers, Ni-MOF catalysts displayed a decrease in ammonia yield as a

function of time. This behavior can be explained by the high ammonia sorption capacity over Ni

Figure 3. SEM images of spent catalyst at various plasma powers a) 100 W, b) 150 W, c) 200 W

and d) 300 W, e) XRD pattern of spent catalyst.

12

based MOFs. Recently, Dinca’s group studied the ammonia sorption on several MOFs containing

triazolate linker. The Ni based MOFs displayed the highest ammonia sorption capacity without

disrupting the structure. A pore filling behavior of ammonia was also established and a high

ammonia sorption capacity of 35 mmol/g was achieved.55 In our study, as the reaction proceeded,

the ammonia molecules start to fill the pores of the MOF which likely reduced the ammonia

detection in the exhaust gases greatly. The ammonia sorption is a quasi-reversible process between

25-100 °C which makes release of ammonia trapping easier by physical treatment such as thermal

heating.55 Yaghi’s group confirmed the retention of ammonia in Ni-MOF-74.52, 56 We run a TPD

study of ammonia (Figure S5) which showed that ammonia started releasing at 60 °C and stopped

at 210 °C while catalyst decomposition is seen at 300 °C. The total ammonia loading was found

to be 3.14 mmol/g-MOF whereas the Glover et al. reports a value of 2.9 mmol/g-MOF.52 The slight

reduction in the catalytic activity after overnight thermal treatment at 100 °C confirms the finding.

To achieve the complete activation, the catalyst should be heated at 200 °C for 2 hours. The

decrease in activity can be a direct result of ammonia adsorption in the MOF. We found that the

peak and average energy yield for ammonia formation was 0.26 g-NH3 (g-catalyst*kWh)-1 and

0.23 g-NH3 (g-catalyst*kWh)-1 at 50 W when Ni-MOF-74 was used as a catalyst. The energy

yields decreased with increase in plasma power. Table 1 compares the catalytic performance of

existing Ni based catalysts in the presence of plasma for ammonia synthesis. The Ni-MOF catalysts

displayed the highest reported energy yields, and ammonia synthesis rate which is comparable to

previous reported Ni based catalysts. Furthermore, the energy cost associated for the Ni-MOF

catalyst is lower than most of the Ni-plasma catalytic systems. Plasma-catalytic ammonia energy

yields range from 0.025-4.45 g-NH3 kWh-1 and energy costs range from 32-850 MJ/mol.12-13, 16-17,

23-24, 57-60 Finally, it is important to highlight that our radiofrequency plasma Ni-MOF system

13

displays energy yields which are one order of magnitude higher than the only previous reported

work on the synthesis of ammonia using RF plasma-catalyst system.20 Specifically the energy

yield for our system was as high as 0.23 g-NH3 kWh-1, as compared to the state of the art literature

for RF plasma-Fe catalyst of 0.012 g-NH3 kWh-1.59

Table 1. Ammonia yield (%) vs. time (min) as a function of plasma power for all studied catalysts,

a) 100 W, b) 150 W, c) 200 W and d) 300 W.

Plasma Catalyst Ammonia

Synthesis Rate

(µmol min-1)

Energy

Yield

(g-NH3

kWh-1)

Energy Cost

(MJ mol-1)

DBD23-24 Ni/ Alumina

Membrane

27 0.178 344

4.5 wt% Catalyst

(Ni/SiO2) + 94.5 wt%

BaTiO3

- - 81

Non-thermal

Atmospheric

Pressure

Plasma17

Ni Mesh 32 0.018 400

RF Plasma (This

Study)

Ni 16.67 0.17 359

Ni-MOF-74 21.17 0.23 265

To learn about the effect of the formed plasma species on the ammonia yield and plasma-

catalytic interaction, we collected the emission spectra of the plasmas as a function of time through

operando UV-Vis. Not only the ammonia interaction with the Ni-MOF but the plasma species and

catalyst interactions affect the ammonia yield. The emission spectra revealed the presence of seven

different species. The nitrogen derived species detected were N2 second positive system(𝐶3𝛱𝑢 →

14

𝐵3𝛱𝑔), N2+ first negative system(𝐵2𝛴𝑢+ → 𝑋2𝛴𝑔

+), N2 first positive system(𝐵3𝛱𝑔 → 𝐴3𝛴𝑢+) and

atomic N(2𝑝23𝑝 → 2𝑝23𝑠) with characteristic peaks at 337.1 nm, 391.4 nm, 662.3 nm and 746.8

nm, respectively. The hydrogen species present were Hα (656.3 nm) and Hβ (486.1 nm). The NH

peak comes up as a pre-shoulder at 336 nm to N2 SPS characteristic peak.61 The Hβ peak was

observed only at higher powers. Figure S6 shows the emission spectra at plasma-catalyst interface

as a function of plasma power. When comparing the spectrums for Ni versus Ni-MOF-74 it is clear

that the H peak intensity is much higher in the metal than in the Ni-MOF. For several metals

including Ni, the peak intensity is directly attributed to the hydrogen recombination.62 Therefore,

higher intensity of Hα peak increases the probability of surface hydrogen recombination with the

gas phase atomic hydrogen in turn reducing the total surface hydrogen available to react and lead

to ammonia. In addition, the surface hydrogen plays a major role in formation of NH3 rather than

the gas phase radicals and gas phase reactions. This may explains in part the higher ammonia yields

obtained when using the Ni-MOF-74 as compared to Ni metal.

Conclusions

In summary, we demonstrate a radiofrequency plasma-catalyst approach to synthesize ammonia.

Specifically, a Ni based metal organic framework was employed as catalyst. As compared to pure

Ni, the Ni-MOF displayed higher ammonia yields. The structural stability of the Ni-MOF catalyst

depended on the plasma power. The Ni-MOF catalyst was stable at low to moderate plasma powers

(100-200 W). High plasma powers (300 W) caused internal stress within the porous framework

leading to amorphization, and therefore compromising the catalyst performance. Ammonia yields

as high as 0.23 g-NH3 (kWh-g-catalyst)-1 and energy cost of 265 MJ mol-1 over Ni-MOF were

observed. The enhanced catalytic activity of the Ni-MOF in the presence of radiofrequency plasma

was attributed to the porous structure which improved mass transfer of guest and product

15

molecules during reaction, the presence of unsaturated Ni metal sites, and lower surface hydrogen

recombination.

ASSOCIATED CONTENT

Supporting Information.

The following files are available free of charge.

Emission Spectra at different powers and catalyst, Schematic of Experimental Setup, Pictures of

Catalyst loading, formula for calculating energy cost and energy yield, details of GC calibration.

AUTHOR INFORMATION

Corresponding Author

*[email protected]

Author Contributions

The manuscript was written through contributions of all authors

Funding Sources

Maria L. Carreon would like to thank the University of Tulsa for financial support as Faculty

Startup funds. Moises A. Carreon thanks National Science Foundation NSF-CBET Award #

1705675 for financial support.

REFERENCES

1. Yin, S.; Xu, B.; Zhou, X.; Au, C., A mini-review on ammonia decomposition catalysts

for on-site generation of hydrogen for fuel cell applications. Appl. Catal., A 2004, 277 (1-2), 1-9.

2. Erisman, J. W.; Sutton, M. A.; Galloway, J.; Klimont, Z.; Winiwarter, W., How a century

of ammonia synthesis changed the world. Nat. Geosci. 2008, 1 (10), 636.

3. Schlögl, R., Ammonia synthesis. Wiley Online Library: 2008.

16

4. Paolucci, C.; Khurana, I.; Parekh, A. A.; Li, S.; Shih, A. J.; Li, H.; Di Iorio, J. R.;

Albarracin-Caballero, J. D.; Yezerets, A.; Miller, J. T., Dynamic multinuclear sites formed by

mobilized copper ions in NOx selective catalytic reduction. Sci. 2017, 357 (6354), 898-903.

5. Kitano, M.; Inoue, Y.; Yamazaki, Y.; Hayashi, F.; Kanbara, S.; Matsuishi, S.; Yokoyama,

T.; Kim, S.-W.; Hara, M.; Hosono, H., Ammonia synthesis using a stable electride as an electron

donor and reversible hydrogen store. Nat. Chem. 2012, 4 (11), 934.

6. Tock, L.; Maréchal, F.; Perrenoud, M., Thermo‐environomic evaluation of the ammonia

production. Can. J. Chem. Eng. 2015, 93 (2), 356-362.

7. Honkala, K.; Hellman, A.; Remediakis, I.; Logadottir, A.; Carlsson, A.; Dahl, S.;

Christensen, C. H.; Nørskov, J. K., Ammonia synthesis from first-principles calculations. Sci.

2005, 307 (5709), 555-558.

8. Kirova-Yordanova, Z., Exergy analysis of industrial ammonia synthesis. Energy 2004, 29

(12-15), 2373-2384.

9. Eliasson, B.; Hirth, M.; Kogelschatz, U., Ozone synthesis from oxygen in dielectric

barrier discharges. J. Phys. D: Appl. Phys. 1987, 20 (11), 1421.

10. Kim, H. H., Nonthermal plasma processing for air‐pollution control: a historical review,

current issues, and future prospects. Plasma Processes Polym. 2004, 1 (2), 91-110.

11. Patil, B.; Cherkasov, N.; Lang, J.; Ibhadon, A.; Hessel, V.; Wang, Q., Low temperature

plasma-catalytic NOx synthesis in a packed DBD reactor: Effect of support materials and

supported active metal oxides. Appl. Catal., B 2016, 194, 123-133.

12. Bai, M.; Zhang, Z.; Bai, M.; Bai, X.; Gao, H., Synthesis of ammonia using CH 4/N 2

plasmas based on micro-gap discharge under environmentally friendly condition. Plasma Chem.

Plasma Process. 2008, 28 (4), 405-414.

13. Bai, M.; Zhang, Z.; Bai, X.; Bai, M.; Ning, W., Plasma synthesis of ammonia with a

microgap dielectric barrier discharge at ambient pressure. IEEE Trans. Plasma Sci. 2003, 31 (6),

1285-1291.

14. Wang, L.; Zhao, Y.; Liu, C.; Gong, W.; Guo, H., Plasma driven ammonia decomposition

on a Fe-catalyst: eliminating surface nitrogen poisoning. Chem. Commun. 2013, 49 (36), 3787-

3789.

15. Wang, W.; Patil, B.; Heijkers, S.; Hessel, V.; Bogaerts, A., Nitrogen fixation by gliding

arc plasma: better insight by chemical kinetics modelling. ChemSusChem 2017, 10 (10), 2145-

2157.

16. Gómez-Ramírez, A.; Cotrino, J.; Lambert, R.; González-Elipe, A., Efficient synthesis of

ammonia from N2 and H2 alone in a ferroelectric packed-bed DBD reactor. Plasma Sources Sci.

Technol. 2015, 24 (6), 065011.

17. Iwamoto, M.; Akiyama, M.; Aihara, K.; Deguchi, T., Ammonia Synthesis on Wool-Like

Au, Pt, Pd, Ag, or Cu Electrode Catalysts in Nonthermal Atmospheric-Pressure Plasma of N2

and H2. ACS Catal. 2017, 7 (10), 6924-6929.

18. Van Helden, J.; Wagemans, W.; Yagci, G.; Zijlmans, R.; Schram, D.; Engeln, R.;

Lombardi, G.; Stancu, G.; Röpcke, J., Detailed study of the plasma-activated catalytic generation

of ammonia in N 2-H 2 plasmas. J. Appl. Phys. 2007, 101 (4), 043305.

19. Uyama, H.; Matsumoto, O., Synthesis of ammonia in high-frequency discharges. II.

Synthesis of ammonia in a microwave discharge under various conditions. Plasma Chem.

Plasma Process. 1989, 9 (3), 421-432.

20. Uyama, H.; Matsumoto, O., Synthesis of ammonia in high-frequency discharges. Plasma

Chem. Plasma Process. 1989, 9 (1), 13-24.

17

21. Tanaka, S.; Uyama, H.; Matsumoto, O., Synergistic effects of catalysts and plasmas on

the synthesis of ammonia and hydrazine. Plasma Chem. Plasma Process. 1994, 14 (4), 491-504.

22. Bogaerts, A.; Neyts, E. C., Plasma Technology: An Emerging Technology for Energy

Storage. ACS Energy Lett. 2018, 3 (4), 1013-1027.

23. Mizushima, T.; Matsumoto, K.; Ohkita, H.; Kakuta, N., Catalytic effects of metal-loaded

membrane-like alumina tubes on ammonia synthesis in atmospheric pressure plasma by

dielectric barrier discharge. Plasma Chem. Plasma Process. 2007, 27 (1), 1-11.

24. Akay, G.; Zhang, K., Process intensification in ammonia synthesis using novel

coassembled supported microporous catalysts promoted by nonthermal plasma. Ind. Eng. Chem.

Res. 2017, 56 (2), 457-468.

25. Gascon, J.; Corma, A.; Kapteijn, F.; Llabres i Xamena, F. X., Metal organic framework

catalysis: Quo vadis? ACS Catal. 2013, 4 (2), 361-378.

26. Farha, O. K.; Eryazici, I.; Jeong, N. C.; Hauser, B. G.; Wilmer, C. E.; Sarjeant, A. A.;

Snurr, R. Q.; Nguyen, S. T.; Yazaydın, A. O. z. r.; Hupp, J. T., Metal–organic framework

materials with ultrahigh surface areas: is the sky the limit? J. Am. Chem. Soc. 2012, 134 (36),

15016-15021.

27. Cavka, J. H.; Jakobsen, S.; Olsbye, U.; Guillou, N.; Lamberti, C.; Bordiga, S.; Lillerud,

K. P., A new zirconium inorganic building brick forming metal organic frameworks with

exceptional stability. J. Am. Chem. Soc. 2008, 130 (42), 13850-13851.

28. Mondloch, J. E.; Bury, W.; Fairen-Jimenez, D.; Kwon, S.; DeMarco, E. J.; Weston, M.

H.; Sarjeant, A. A.; Nguyen, S. T.; Stair, P. C.; Snurr, R. Q., Vapor-phase metalation by atomic

layer deposition in a metal–organic framework. J. Am. Chem. Soc. 2013, 135 (28), 10294-10297.

29. Park, K. S.; Ni, Z.; Côté, A. P.; Choi, J. Y.; Huang, R.; Uribe-Romo, F. J.; Chae, H. K.;

O’Keeffe, M.; Yaghi, O. M., Exceptional chemical and thermal stability of zeolitic imidazolate

frameworks. Proc. Natl. Acad. Sci. U. S. A. 2006, 103 (27), 10186-10191.

30. Férey, G., Hybrid porous solids: past, present, future. Chem. Soc. Rev. 2008, 37 (1), 191-

214.

31. MacGillivray, L. R., Metal-organic frameworks: design and application. John Wiley &

Sons: 2010.

32. Liu, K.-H.; Zhong, H.-X.; Li, S.-J.; Duan, Y.-X.; Shi, M.-M.; Zhang, X.-B.; Yan, J.-M.;

Jiang, Q., Advanced catalysts for sustainable hydrogen generation and storage via hydrogen

evolution and carbon dioxide/nitrogen reduction reactions. Progress in Materials Science 2017.

33. Martirez, J. M. P.; Carter, E. A., Prediction of a low-temperature N2 dissociation catalyst

exploiting near-IR–to–visible light nanoplasmonics. Science advances 2017, 3 (12), eaao4710.

34. Tsuji, Y.; Kitano, M.; Kishida, K.; Sasase, M.; Yokoyama, T.; Hara, M.; Hosono, H.,

Ammonia synthesis over Co–Mo alloy nanoparticle catalyst prepared via sodium naphthalenide-

driven reduction. Chem. Commun. 2016, 52 (100), 14369-14372.

35. Shi, M. M.; Bao, D.; Li, S. J.; Wulan, B. R.; Yan, J. M.; Jiang, Q., Anchoring PdCu

Amorphous Nanocluster on Graphene for Electrochemical Reduction of N2 to NH3 under

Ambient Conditions in Aqueous Solution. Advanced Energy Materials 2018, 1800124.

36. Li, S. J.; Bao, D.; Shi, M. M.; Wulan, B. R.; Yan, J. M.; Jiang, Q., Amorphizing of Au

nanoparticles by CeOx–RGO hybrid support towards highly efficient electrocatalyst for N2

reduction under ambient conditions. Advanced Materials 2017, 29 (33), 1700001.

37. Shi, M. M.; Bao, D.; Wulan, B. R.; Li, Y. H.; Zhang, Y. F.; Yan, J. M.; Jiang, Q., Au

Sub‐Nanoclusters on TiO2 toward Highly Efficient and Selective Electrocatalyst for N2

Conversion to NH3 at Ambient Conditions. Advanced Materials 2017, 29 (17), 1606550.

18

38. Bao, D.; Zhang, Q.; Meng, F. L.; Zhong, H. X.; Shi, M. M.; Zhang, Y.; Yan, J. M.; Jiang,

Q.; Zhang, X. B., Electrochemical reduction of N2 under ambient conditions for artificial N2

fixation and renewable energy storage using N2/NH3 cycle. Advanced Materials 2017, 29 (3),

1604799.

39. Wu, X.; Bao, Z.; Yuan, B.; Wang, J.; Sun, Y.; Luo, H.; Deng, S., Microwave synthesis

and characterization of MOF-74 (M= Ni, Mg) for gas separation. Microporous Mesoporous

Mater. 2013, 180, 114-122.

40. Dietzel, P. D.; Panella, B.; Hirscher, M.; Blom, R.; Fjellvåg, H., Hydrogen adsorption in

a nickel based coordination polymer with open metal sites in the cylindrical cavities of the

desolvated framework. Chem. Commun. 2006, (9), 959-961.

41. Millward, A. R.; Yaghi, O. M., Metal− organic frameworks with exceptionally high

capacity for storage of carbon dioxide at room temperature. J. Am. Chem. Soc. 2005, 127 (51),

17998-17999.

42. Valenzano, L.; Civalleri, B.; Chavan, S.; Palomino, G. T.; Areán, C. O.; Bordiga, S.,

Computational and Experimental Studies on the Adsorption of CO, N2, and CO2 on Mg-MOF-

74. J. Phys. Chem. C 2010, 114 (25), 11185-11191.

43. Yang, D.-A.; Cho, H.-Y.; Kim, J.; Yang, S.-T.; Ahn, W.-S., CO 2 capture and conversion

using Mg-MOF-74 prepared by a sonochemical method. Energy Environ. Sci. 2012, 5 (4), 6465-

6473.

44. Crystallographic Open Database. http://www.crystallography.net/cod/ (accessed May,

2018).

45. Van Durme, J.; Dewulf, J.; Leys, C.; Van Langenhove, H., Combining non-thermal

plasma with heterogeneous catalysis in waste gas treatment: A review. Appl. Catal., B 2008, 78

(3-4), 324-333.

46. Liu, C.-j.; Yu, K.; Zhang, Y.-p.; Zhu, X.; He, F.; Eliasson, B., Characterization of plasma

treated Pd/HZSM-5 catalyst for methane combustion. Appl. Catal., B 2004, 47 (2), 95-100.

47. Yoon, S.; Panov, A. G.; Tonkyn, R. G.; Ebeling, A. C.; Barlow, S. E.; Balmer, M. L., An

examination of the role of plasma treatment for lean NOx reduction over sodium zeolite Y and

gamma alumina: Part 1. Plasma assisted NOx reduction over NaY and Al2O3. Catal. Today

2002, 72 (3-4), 243-250.

48. Zhang, Y.-P.; Ma, P.-S.; Zhu, X.; Liu, C.-J.; Shen, Y., A novel plasma-treated

Pt/NaZSM-5 catalyst for NO reduction by methane. Catal. Commun. 2004, 5 (1), 35-39.

49. Zhu, X.; Huo, P.; Zhang, Y.-p.; Cheng, D.-g.; Liu, C.-j., Structure and reactivity of

plasma treated Ni/Al2O3 catalyst for CO2 reforming of methane. Appl. Catal., B 2008, 81 (1-2),

132-140.

50. Pârvulescu, V. I.; Magureanu, M.; Lukes, P., Plasma chemistry and catalysis in gases

and liquids. John Wiley & Sons: 2012.

51. Aresta, M., Carbon dioxide as chemical feedstock. John Wiley & Sons: 2010.

52. Glover, T. G.; Peterson, G. W.; Schindler, B. J.; Britt, D.; Yaghi, O., MOF-74 building

unit has a direct impact on toxic gas adsorption. Chemical Engineering Science 2011, 66 (2),

163-170.

53. Park, J.; Park, H.; Hahn, Y.; Yi, G.-C.; Yoshikawa, A., Dry etching of ZnO films and

plasma-induced damage to optical properties. J. Vac. Sci. Technol., B: Microelectron. Nanometer

Struct.--Process., Meas., Phenom. 2003, 21 (2), 800-803.

54. Cardinaud, C.; Peignon, M.-C.; Tessier, P.-Y., Plasma etching: principles, mechanisms,

application to micro-and nano-technologies. Appl. Surf. Sci. 2000, 164 (1-4), 72-83.

19

55. Rieth, A. J.; Dinca, M., Controlled Gas Uptake in Metal–Organic Frameworks with

Record Ammonia Sorption. J. Am. Chem. Soc. 2018, 140 (9), 3461-3466.

56. Britt, D.; Tranchemontagne, D.; Yaghi, O. M., Metal-organic frameworks with high

capacity and selectivity for harmful gases. Proc. Natl. Acad. Sci. U. S. A. 2008, 105 (33), 11623-

11627.

57. Nakajima, J.; Sekiguchi, H., Synthesis of ammonia using microwave discharge at

atmospheric pressure. Thin Solid Films 2008, 516 (13), 4446-4451.

58. Gómez‐Ramírez, A.; Montoro‐Damas, A. M.; Cotrino, J.; Lambert, R. M.; González‐Elipe, A. R., About the enhancement of chemical yield during the atmospheric plasma synthesis

of ammonia in a ferroelectric packed bed reactor. Plasma Processes Polym. 2017, 14 (6).

59. Matsumoto, O., Plasma catalytic reaction in ammonia synthesis in the microwave

discharge. J. Phys. IV France 1998, 8 (PR7), Pr7-411-Pr7-420.

60. Peng, P.; Cheng, Y.; Hatzenbeller, R.; Addy, M.; Zhou, N.; Schiappacasse, C.; Chen, D.;

Zhang, Y.; Anderson, E.; Liu, Y., Ru-based multifunctional mesoporous catalyst for low-

pressure and non-thermal plasma synthesis of ammonia. Int. J. Hydrogen Energy 2017, 42 (30),

19056-19066.

61. Zhu, X.-M.; Pu, Y.-K., Optical emission spectroscopy in low-temperature plasmas

containing argon and nitrogen: determination of the electron temperature and density by the line-

ratio method. J. Phys. D: Appl. Phys. 2010, 43 (40), 403001.

62. Shah, J.; Wang, W.; Bogaerts, A.; Carreon, M. L., Ammonia synthesis by radio

frequency plasma catalysis: revealing the underlying mechanisms. ACS Applied Energy

Materials 2018.

20

BRIEFS. Alternate routes for ammonia synthesis have gained attention due to the energy intensive

process, the letter describes the use of a molecular sieve for catalysis which opens a whole new

class of material that can used for catalysis as well as ammonia storage in one-pot synthesis.

SYNOPSIS. A novel framework, Ni-MOF-74 for the synthesis of ammonia via plasma catalysis.

The MOF exhibits a better catalytic activity than pure metal as well as goes through structural

changes.