sustainable atmospheric ammonia synthesis and nitrogen

155
Sustainable atmospheric ammonia synthesis and nitrogen fixation using non-thermal plasma (NTP) A DISSERTATION SUBMITTED TO THE FACULTY OF THE GRADUATE SCHOOL OF THE UNIVERSITY OF MINNESOTA BY Peng Peng IN PARTIAL FULFILLMENT OF THE REQUIREMENTS FOR THE DEGREE OF DOCTOR OF PHILOSOPHY Advisor: Dr. R. Roger Ruan Oct, 2018

Upload: others

Post on 08-Apr-2022

8 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Sustainable atmospheric ammonia synthesis and nitrogen

Sustainable atmospheric ammonia synthesis and nitrogen

fixation using non-thermal plasma (NTP)

A DISSERTATION SUBMITTED TO THE FACULTY OF THE GRADUATE

SCHOOL OF THE UNIVERSITY OF MINNESOTA BY

Peng Peng

IN PARTIAL FULFILLMENT OF THE REQUIREMENTS FOR THE DEGREE OF

DOCTOR OF PHILOSOPHY

Advisor: Dr. R. Roger Ruan

Oct, 2018

Page 2: Sustainable atmospheric ammonia synthesis and nitrogen

© Copyright by Peng Peng 2018

Page 3: Sustainable atmospheric ammonia synthesis and nitrogen

i

Acknowledgements

Mentorship is what I value the most during my research career. So first of all, I

would like to express my deepest gratitude towards my Ph.D. advisor, Prof. Roger

Ruan. I am very grateful that I have had the opportunity to receive the mentorship

from many excellent individuals throughout my career. I am greatly indebted to

Prof. Paul Chen, Prof. Tom Halbach, and Prof. Brock Lundberg to be the

committee members of my defense and helping me revise my thesis. I want to

thank Dr. Sharon Marsh (Land O’ Lakes Inc.) and Eric Schroeder (Great Plains

Institute) for being my industry mentor, Prof. Friedrich Srienc for giving me the

opportunity to conduct my first research project, and Dr. John Barrett for teaching

me my very first set of academic research skills.

I also would like to thank everyone in Prof. Ruan’s research group for offering

help and assistance whenever I encountered difficulties. I would have never

completed this thesis without them.

Last but not least, I owe my family everything for their unconditional love and

support. They are my lifetime mentors.

Page 4: Sustainable atmospheric ammonia synthesis and nitrogen

ii

Dedication

This thesis is dedicated to my beloved family, friends, mentors, and collegues.

Page 5: Sustainable atmospheric ammonia synthesis and nitrogen

iii

Abstract

Ammonia has recently been intensively studied as a clean, sustainable fuel

source and an efficient energy storage medium due to its effectiveness as a

hydrogen carrier molecule. However, the current method of ammonia synthesis,

known as the Haber-Bosch process, requires a large fossil fuel input, high

temperatures and pressures, as well as a significant capital investment. These

volatile conditions and high operating costs prevent decentralized and small-scale

ammonia production at the level of small farms and local communities.

Non-thermal plasma technology represents a promising alternative method of

clean ammonia synthesis, as it circumvents the volatile operating conditions,

fossil fuel use, and high capital costs of the Haber-Bosch process.

In this thesis research, this emerging technology was realized at the bench

scale and was optimized by various efforts, including catalyst improvement,

system development, and absorption enhancement. For the catalyst improvement,

a multi-functional catalyst was introduced and deposited onto various supporting

materials. For absorption enhancement, MgCl2 was implemented for the

absorption enhancement, taking advantage of its capability of forming Mg3N2 and

Mg(NH3)6Cl2 during the process. Meanwhile, the pulse density modulation (PDM)

was introduced to improve the performance of system. The series of efforts

improved the energy efficiency of the system by approximately one fold compared

with previous studies, and achieved the highest value of 20 g/kwh. Furthermore,

an innovative plasma gas-liquid ammonia synthesis approach was explored. It

was found that this approach could produce other nitrogen compounds (nitrate

Page 6: Sustainable atmospheric ammonia synthesis and nitrogen

iv

and nitrite acids) while generating ammonium, which could potentially add value

to the products of the plasma-assisted ammonia synthesis process.

Based on the results, the challenges and future opportunities of the

plasma-assisted ammonia synthesis approach were discussed. Lastly,

recommendations were made on how this technology could be beneficial to the

ammonia industry, through its potential to promote localized and environmentally

friendly energy production and storage.

Page 7: Sustainable atmospheric ammonia synthesis and nitrogen

v

Table of Contents List of tables ........................................................................................................ vii

List of figures ...................................................................................................... viii

CHAPTER 1 Introduction ...................................................................................... 1

CHAPTER 2 Literature review .............................................................................. 6

2.1 Introduction ................................................................................................. 6

2.2 Review methodology ................................................................................... 8

2.2.1 General methodology ........................................................................... 8

2.2.2 Initial Screening .................................................................................... 8

2.2.3 Categorization of literature ................................................................... 9

2.2.4 Data extraction, analysis, and reporting ............................................... 9

2.3 Reaction principles and kinetics ................................................................ 12

2.4 Reactor development ................................................................................ 18

2.5 Catalyst development................................................................................ 29

2.6 Reactant type, composition, and feed rate ................................................ 36

2.6.1 The effect of N2 and H2 composition and feed rate ........................... 36

2.6.2 Reactants other than N2 and H2 ......................................................... 37

2.7 Conclusions .............................................................................................. 40

CHAPTER 3 Atmospheric plasma-assisted ammonia synthesis using Ru-based multifunctional catalyst supported by MgO ......................................................... 42

3.1 Introduction ............................................................................................... 42

3.2 Materials and methods .............................................................................. 45

3.3 Results and discussions ............................................................................ 49

3.3.1 Synergy between NTP, catalysts, and promoters ............................... 49

3.3.2 Feeding gas ratio & flow rate effects .................................................. 50

3.3.4 Frequency & voltage effects ............................................................... 52

3.4 Conclusions .............................................................................................. 54

CHAPTER 4 Atmospheric plasma-assisted ammonia synthesis using Ru-based multifunctional catalyst supported by mesoporous silica MCM-41 ...................... 55

4.1 Introduction ............................................................................................... 55

4.2 Materials and methods .............................................................................. 56

Page 8: Sustainable atmospheric ammonia synthesis and nitrogen

vi

4.2.1 Improvements ..................................................................................... 56

4.2.2 Catalyst preparation ........................................................................... 56

4.3 Results and discussions ............................................................................ 59

4.3.1 Characteristics of the Catalysts .......................................................... 59

4.3.2 Catalytic synthesis results .................................................................. 67

4.3.3 Synergistic effects between the plasma and catalyst ......................... 71

4.4 Conclusions .............................................................................................. 75

CHAPTER 5 Absorption-enhanced NTP ammonia synthesis with the assistance of magnesium chloride ....................................................................................... 76

5.1 Introduction ............................................................................................... 76

5.2 Materials and methods .............................................................................. 78

5.2.1 System improvements ........................................................................ 78

5.2.2 Absorbent preparation and characterization ....................................... 78

5.3 Results and discussions ............................................................................ 80

5.4 Conclusions .............................................................................................. 93

CHAPTER 6 Atmospheric plasma-assisted nitrogen fixation using water and nitrogen jet plasma ............................................................................................. 94

6.1 Introduction ............................................................................................... 94

6.2 Materials and Methods .............................................................................. 97

6.2.1 Improvements of the system .............................................................. 97

6.2.2 Description of apparatus ..................................................................... 97

6.3 Results and discussion ........................................................................... 100

6.3.1 Synthesis results .............................................................................. 100

6.3.2 Gas phase reactions ......................................................................... 103

6.3.3 Liquid phase reactions ...................................................................... 104

6.4 Conclusions ............................................................................................ 112

CHAPTER 7 Challenges and Opportunities ..................................................... 113

7.1 Challenges .............................................................................................. 113

7.2 Opportunities ........................................................................................... 115

CHAPTER 8 Conclusions and future remarks .................................................. 123

Bibliography ...................................................................................................... 125

Page 9: Sustainable atmospheric ammonia synthesis and nitrogen

vii

List of tables

Table 1 A summary of the reactor development for NTP ammonia synthesis ..... 23

Table 2 A summary of the catalyst development for NTP ammonia synthesis .... 33

Table 3 A summative comparison between the conventional Haber-Bosch and NTP ammonia synthesis ................................................................................... 120

Page 10: Sustainable atmospheric ammonia synthesis and nitrogen

viii

List of figures

Figure 1 A flow diagram of the non-thermal plasma ammonia synthesis process ............ 4

Figure 2 A process flow diagram of the systematic and integrated literature review process used in this review. ........................................................................................... 11

Figure 3 A schematic diagram of Dielectric Barrier Discharge plasma generation .......... 19

Figure 4 Dissociation of nitrogen under NTP ................................................................. 44

Figure 5 a) Process flow diagram of the NTP ammonia synthesis approach. b) A simplified diagram of the wiring connection between the invertor and transformer ......... 47

Figure 6 Ammonia synthesis under different conditions. ................................................ 50

Figure 7 Effect of N2 composition on ammonia synthesis efficiency ............................... 51

Figure 8 Effect of gas flow rate on ammonia synthesis efficiency .................................. 52

Figure 9 Effect of discharge frequency on ammonia synthesis efficiency ....................... 53

Figure 10 Effect of applied voltage on ammonia synthesis efficiency ............................. 53

Figure 11 FTIR spectra of the Ru/SI-MCM-41 catalyst ................................................... 60

Figure 12 a) SEM, b) TEM, c) AFM topography, and d) PiFM images of the Ru-promoter/Si-MCM-41 catalyst .................................................................................. 61

Figure 13 EDS results of the Ru-promoter/Si-MCM-41 catalyst ..................................... 62

Figure 14 XPS spectrum of the Ru-promoter/Si-MCM-41 catalyst. a) survey scan. b) high resolution scan for Ru3dC1s. c) high-resolution scan for the Cs .................................... 64

Figure 15 XRD spectrum of the calcined and reduced Ru-promoter/Si-MCM-41 catalyst ...................................................................................................................................... 65

Figure 16 Synthesis efficiency results with different catalyst components ...................... 68

Figure 17 Effects of the N2 feed concentration on the ammonia synthesis effeciency under different voltage conditions .................................................................................. 69

Figure 18 Effects of the inlet gas flow rate on the ammonia synthesis effeciency under different voltage conditions ............................................................................................ 69

Figure 19 Effects of the applied voltage on the ammonia synthesis efficiency and outlet concentration ................................................................................................................. 71

Figure 20 Effects of the applied frequency on the ammonia synthesis efficiency and outlet concentration ................................................................................................................. 71

Figure 21 The proposed mechanism of NTP ammonia synthesis using Ru-promoter/Si-MCM-41 catalyst .................................................................................. 73

Figure 22 Schematic diagrams of the MgCl2 locations relative to the plasma discharge region ............................................................................................................................ 81

Figure 23 Ammonia synthesis results with MgCl2 packed at different locations relative to the plasma discharge region .......................................................................................... 81

Figure 24 Ammonia synthesis results with MgCl2 under different voltage conditions ...... 84

Figure 25 Ammonia synthesis results with MgCl2 under different frequency conditions .. 84

Figure 26 Ammonia synthesis results with MgCl2 under different nitrogen compositions 85

Figure 27 Ammonia synthesis results with MgCl2 under different gas flow rates ............. 85

Figure 28 XRD patterns of the MgCl2 samples under different conditions ...................... 89

Figure 29 SEM image and EDS analysis results of the untreated MgCl2 sample ........... 90

Figure 30 SEM image and EDS analysis results of the MgCl2 sample treated in the plasma discharge region ............................................................................................... 91

Figure 31 SEM image and EDS analysis results of the MgCl2 samples treated below the

Page 11: Sustainable atmospheric ammonia synthesis and nitrogen

ix

plasma discharge region ............................................................................................... 92

Figure 32 Process flow diagram of the in-situ atmospheric nitrogen fixation process ..... 99

Figure 33 Concentrations of nitrate, nitrite and ammonium at different experimental conditions at 30 C ........................................................................................................ 102

Figure 34 Concentrations of nitrate, nitrite and ammonium of the in-situ reactor at different temperatures ................................................................................................. 106

Figure 35 Synthesis rates of nitrate, nitrite and ammonium at different experimental conditions at 30℃ ........................................................................................................ 108

Figure 36 Synthesis rates at nitrate, nitrite and ammonium of the in-situ reactor at different temperatures ................................................................................................. 109

Figure 37 The propose mechanism for the In-situ plasma jet nitrogen fixation ............. 111

Figure 38 An illustrative diagram of the renewable-to-ammonia approach ................... 118

Page 12: Sustainable atmospheric ammonia synthesis and nitrogen

1

CHAPTER 1 Introduction

Since the 1st half of the 20th century, the synthesis of ammonia was publicly

recognized due to the efforts of Fritz Haber and Carl Bosch, two Nobel Prize

recipients, who developed and industrialized the Haber-Bosch process for

ammonia production. The Haber-Bosch process has shaped our world and

resulted in humanity becoming heavily dependent on ammonia (Erisman, Sutton,

Galloway, Klimont, & Winiwarter, 2008). The most well-known applications of

ammonia include fertilizer, explosives during the 1st world war, and pesticides, etc.

(Galloway et al., 2008; Smil, 1997). In recent years, innovative applications of

ammonia have been intensively studied, including refrigeration, fermentation, and

energy carrier potential (Von Blottnitz & Curran, 2007). For example, ammonia

has been investigated for its potential as an energy source for fuel cells (Cox &

Treyer, 2015; Maffei, Pelletier, Charland, & McFarlan, 2007), transportation (Miura

& Tezuka, 2014), and other off-grid power applications (Davis et al., 2009). Due to

its high hydrogen content, ammonia has been viewed as one of the most efficient

and economical hydrogen carriers (Christensen, Johannessen, Sørensen, &

Nørskov, 2006; Cumaranatunge et al., 2007), as well as derivative species like

ammonia bromine (Smythe & Gordon, 2010). In terms of energy content,

ammonia has a heat of combustion of around 22 MJ/kg and a low heating value

comparable to diesel fuels (Zamfirescu & Dincer, 2009). Furthermore, the high

boiling point of ammonia makes it an ideal material for indirect hydrogen storage

(Demirci & Miele, 2009; Lan, Irvine, & Tao, 2012). Most importantly, the complete

Page 13: Sustainable atmospheric ammonia synthesis and nitrogen

2

combustion of ammonia is a sustainable process that does not emit any

greenhouse gases. Therefore, ammonia, or ammonia bromide has also been

introduced with petroleum-based fuels to vehicle engines for cleaner emissions

(Stephens, Baker, Matus, Grant, & Dixon, 2007; Zamfirescu & Dincer, 2009).

Although the Haber-Bosch process is responsible for providing over 130

million tons of ammonia annually to support approximately 40% of the world’s

population, it is also responsible for up to 2% of the global energy consumption (B.

Patil, Hessel, Lang, & Wang, 2016). The reaction conditions of the Haber Bosch

process lie in the range of 200 to 400 atm and 400 to 600 °C, respectively

(Hargreaves, 2014). These intense temperature and pressure conditions are the

main disadvantages of the Haber-Bosch process, as they prevent the possibility

of lowering capital costs (Gilland, 2014; Razon, 2014). Additionally, the high

pressure required for the traditional Haber-Bosch Process is also a limiting factor

in reducing the economies of scale of localized production facilities due to the high

energy (and cost) requirements of compression (Razon, 2014). Life cycle

assessment (LCA) results have shown that the ammonia generation pathways

can greatly impact the for environmental performance of the ammonia-based

power applications (Cox & Treyer, 2015). Therefore, researchers are seeking new

methods of ammonia synthesis, which occur under more moderate conditions

(Vojvodic et al., 2014; Zhang, Chen, & Wen, 2012), require less carbon input

(Gilbert, Alexander, Thornley, & Brammer, 2014), or can be powered by

renewable energy sources (Bardi, El Asmar, & Lavacchi, 2013; H. Liu, 2014).

Since the start of late 1990s, many studies were carried out on non-thermal

Page 14: Sustainable atmospheric ammonia synthesis and nitrogen

3

plasma (NTP) based processes for pollution control (H. H. Kim, 2004). Studies

have shown that NTP is capable for the destruction and removal of sulfur dioxide

(H. Ma, Chen, Zhang, Lin, & Ruan, 2002), hydrogen sulfide (H. Ma, Chen, & Ruan,

2001), odors (R. R. Ruan et al., 1999), and other volatile organic compounds (R.

R. Ruan et al., 2000; R. R. Ruan et al., 1999). In recent years, the results from

these studies have been applied to decompose pollutant gases from animal

production to food processing facilities (Schiavon et al., 2015; Schiavon, Schiorlin,

Torretta, Brandenburg, & Ragazzi, 2017). Results show that NTP is more efficient

and less energy intensive than most of the traditional gas treatment technologies

(Stasiulaitiene et al., 2016). (H. Ma et al., 2002) discovered that by injecting SO2

into the odor stream passing through an NTP reactor, solid particles were

produced at the exit of the reactor. These particles were then confirmed to be

ammonia sulfate ((NH4)2SO4). Researchers believed that SO2 was ionized and

reacted under the NTP conditions, resulting in (NH4)2SO4. These studies

confirmed that NTP is capable of ionizing gaseous compounds and may lead to

innovative methods of chemical synthesis reactions. Since then, NTP has been

intensely investigated in the fields of flow control in the past twenty years (Nie et

al., 2013; B. M. Penetrante & Schultheis, 2013).

During this period, NTP was found to be a potential alternative to the high

temperature and pressure method for the synthesis of many chemicals (Petitpas

et al., 2007; R. Ruan et al., 2014), such as benzene and isooctane (Rahemi,

Haghighi, Babaluo, Jafari, & Estifaee, 2013). Additionally the NTP-assisted

nitrogen fixation method has been viewed as an attractive alternatives to the

Page 15: Sustainable atmospheric ammonia synthesis and nitrogen

4

Haber-Bosch process (B. Patil, Cherkasov, et al., 2016). In the early years of NTP

nitrogen fixation, emphasis was placed on the synthesis of nitric oxide compounds

instead of ammonia (B. Patil, Cherkasov, et al., 2016). NOx synthesis was favored

over ammonia production because it offered a thermodynamically favorable

method of nitrogen fixation, requiring lower energy input (B. Patil, Wang, Hessel,

& Lang, 2015; W. Wang, Patil, Heijkers, Hessel, & Bogaerts, 2017).

Figure 1 A flow diagram of the non-thermal plasma ammonia synthesis process

A flow diagram of the NTP ammonia synthesis approach is shown in Figure 1.

This process relies on the plasma discharge to dissociate the reactants and form

ammonia with the assistance of catalysts. Various sources such as microwave

and dielectric barrier discharge can be used to generate the plasma required for

the synthesis. Depending on the individual study, the products then go through a

separation process to collect the ammonia product and recycle the unreacted

reactants. The first attempt at using NTP to synthesize ammonia was made at the

Page 16: Sustainable atmospheric ammonia synthesis and nitrogen

5

end of 1980s, where the authors tried to synthesize ammonia in high-frequency

discharges (Sugiyama et al., 1986). (Uyama, Nakamura, Tanaka, & Matsumoto,

1993) attempted the NTP synthesis of ammonia using microwave and radio

frequency (RF) plasma. The process was operated at a pressure much lower than

atmospheric pressure, approximately 700 Pa. After the concept of using plasma

to synthesize ammonia was proved in these studies, researchers have started

conducting more thorough investigations into the NTP-assisted ammonia

synthesis process. Further research indicated that the non-equilibrium plasma,

generated under non-thermal conditions, could dissociate diatomic nitrogen and

hydrogen at atmospheric pressure with the assistance of select catalysts. The low

operating pressure of NTP offers the opportunity to circumvent the compression

processes by flowing hydrogen gas directly into the reactor from other renewable

sources, such as gasification of biomass or electrolysis of water. Since

pressurizing the system is not required, the NTP system can be operated

continuously or incorporated directly into a syn-gas production stream. A recent

study proposed that the moderate process conditions could potentially allow for

the economically viable construction and operation of widely distributed NTP

based fertilizer production systems on farms and renewable hydrogen production

sites (Rune Ingels & David B. Graves, 2015).

Page 17: Sustainable atmospheric ammonia synthesis and nitrogen

6

CHAPTER 2 Literature review

2.1 Introduction

Along with the NTP methods, there are several recently developed synthesis

approaches that could produce ammonia under low temperature and pressure,

namely the biological and electro-chemical cell synthesis. There have been

review articles published regarding the biological (Hinnemann and Nørskov, 2006)

and electro-chemical cell synthesis (Guo et al., 2017; Kyriakou et al., 2017) of

ammonia. Thus the first goal of this section is to provide a comprehensive review

with an emphasis on the plasma-assisted ammonia synthesis. In this dissertation,

the three crucial factors to the successful development of non-thermal plasma

ammonia synthesis technology were identified as: understanding of relevant

reaction kinetics, reactor development, and catalyst development. The

development of these aspects will significantly improve the energy efficiency and

environmental sustainability of this technology. Additionally, this paper analyzes

the effects of reactant type, composition, and feed rate on ammonia synthesis

efficiency, reactant conversion. Most importantly, this paper recommended how

the improvements of these factors could result in a cleaner ammonia production

process. A cradle to factory gate life cycle assessment shows that nitrogen

fixation by plasma can potentially enhance the global warming potential as

compared to the conventional fixation pathways by 19% (Anastasopoulou et al.,

Page 18: Sustainable atmospheric ammonia synthesis and nitrogen

7

2016). However, the published life cycle assessment on nitrogen fixation

technologies focuses primarily on the production of nitric acid using plasma.

Therefore, another goal of this study is to fill in the knowledge gap of between the

published literature and a complete life cycle assessment that targets specifically

on the NTP synthesis of ammonia.

Page 19: Sustainable atmospheric ammonia synthesis and nitrogen

8

2.2 Review methodology

2.2.1 General methodology

In order to effectively select the literature that could represent the current

status and help identify the challenges and opportunities of the NTP technology,

this paper incorporates a systematic and integrated literature review methodology

adapted from (Tranfield, Denyer, & Smart, 2003; Whittemore & Knafl, 2005). As

indicated in (Whittemore & Knafl, 2005), the general steps of an integrated

literature review are topic identification, literature search, data evaluation, data

analysis and presentation, as indicated in Figure 2.

2.2.2 Initial Screening

During the literature search stage, the methodology from (Tranfield et al., 2003)

is implemented. The literature search method includes: the identification of

literature, selection of literature, and literature quality assessment. First,

approximately 300 pieces of literature were selected regarding the synthesis of

ammonia via an intensive search from the research databases including: Elsevier,

Web Of Science, and Wiley Online Library, etc. At this stage, the problems of the

conventional ammonia synthesis, such as the high temperature/pressure, and the

large consumption of fossil fuels, were identified. Second, qualities such as the

validity, reliability, and trustworthiness of the studies are used to determine

Page 20: Sustainable atmospheric ammonia synthesis and nitrogen

9

whether they should be analyzed after the initial literature screening (Elo et al.,

2014). In this review, a high-quality study should have a detailed methodology,

complete list of parameters, and a comprehensive and thorough discussion of the

results, etc.

2.2.3 Categorization of literature

Third, the pieces of literature are categorized into different sections. Studies

that involve the life cycle assessment and the environmental issues of the

conventional ammonia production indicate the need of developing an alternative

ammonia synthesis process. These studies also point out the necessity to

improve the accessibility of ammonia production to small industries and farms.

Next, the studies about using NTP to synthesis ammonia under low temperature

and pressure conditions are identified, which function as the building blocks (or

“foundation”) of this review paper. Finally, studies in the related area are selected

to provide potential insight to resolve the challenges of the NTP ammonia

synthesis. Examples of these studies include the plasma nitrogen fixation to

produce nitric oxides and catalyst development of high temperature and pressure

ammonia synthesis.

2.2.4 Data extraction, analysis, and reporting

For the literature included in this review, only the data that effectively

Page 21: Sustainable atmospheric ammonia synthesis and nitrogen

10

represents the consequence of the study and addresses the trend of the studies is

extracted, both qualitatively and quantitatively. The emphasis of data extraction

also depends on the categorization of literature. Representative data such as the

energy efficiency, catalyst information, and conversion are selected as analysis to

the literature on NTP ammonia synthesis. During the data analysis stage, all the

units of the synthesis efficiency and conversion rate are converted for comparison

and analysis. During the reporting stage, two types of comparison are made, one

between the conventional and NTP ammonia synthesis technologies, and the

other within the NTP ammonia synthesis studies. During the reporting phase,

challenges of the NTP ammonia synthesis technology are discussed and the

recommendations to overcome the challenges are proposed.

Page 22: Sustainable atmospheric ammonia synthesis and nitrogen

11

Figure 2 A process flow diagram of the systematic and integrated literature review process used in this

review.

Page 23: Sustainable atmospheric ammonia synthesis and nitrogen

12

2.3 Reaction principles and kinetics

Since the development of the ammonia production process, intensive

investigations have been made to understand the kinetics of ammonia synthesis

reactions. However, less effort has been demonstrated to study the kinetics of

ammonia synthesis under NTP conditions, compared with the traditional

Haber-Bosch process. The kinetics provides a foundation for manipulating and

modeling the reaction to increase the rate, conversion, and efficiency of an

ammonia synthesis reaction, and is also what guides the design of the NTP

reactors. Overall, the NTP has two major functions for the catalytic conversion of

nitrogen and hydrogen to ammonia. First, it directly causes N2 and H2 to

dissociate and form NH3 with or without catalyst. Second, it provides the ionization

energy necessary to produce electrons for the catalysis system to function (Neyts,

Ostrikov, Sunkara, & Bogaerts, 2015). Therefore, modeling of the synthetic

process under NTP conditions will not only allow more efficient reaction, but will

benefit significantly to the reactor and process design, which will further contribute

to a cleaner production of ammonia using this technology. On a macroscopic level,

the ammonia synthesis using N2 and H2 as the reactants were centered on

equation (2.1), and its kinetics could be described via the following power

equation (2.2) (Jennings, 2013).

Page 24: Sustainable atmospheric ammonia synthesis and nitrogen

13

N2 (g) + 3H2 (g) ↔ 2NH3 (g); ∆H= -92.44 kJ/mol (2.1)

𝑟 = 𝑁2𝛼𝐻2

𝛽 (2.2)

The reported 𝛼 and 𝛽 values for the power kinetic equation in the NTP

ammonia synthesis were 𝛼 = 0.77 and 𝛽 = 1.16 for a wool-like copper catalyst

(Aihara et al., 2016). Comparing with the reported coefficients under high

temperature and pressure conditions using transitional metal based catalysts (0.8

to1.2 for 𝛼 and 1.5 to 2.2 for 𝛽), the 𝛼 and 𝛽 values from the NTP synthesis

were lower, which corresponded to the slower synthesis rate of NTP over the high

temperature and pressure process (Hagen et al., 2003). Furthermore, it was

reasonable to propose that the kinetics of the NTP ammonia synthesis is limited

more by the dissociation of N2 than H2 since the 𝛼 value is lower than the 𝛽

value. However, due to the complex and non-equilibrium nature of the NTP, a

more detailed kinetics analysis must be carried out to provide a better

understanding of the intermediate species and chemical pathways in the NTP

plasma synthesis (Hong et al., 2017).

Previous studies hypothesized that the synthesis of ammonia from nitrogen

and hydrogen gas involved the following three mechanisms: decomposition,

structural rearrangement, and fragment elimination (B. Penetrante et al., 1997;

Van Durme, Dewulf, Leys, & Van Langenhove, 2008). It has been widely

understood that the first steps of plasma ammonia synthesis are the dissociation

and ionization of the inlet nitrogen and hydrogen gases (Mindong Bai, Zhang, Bai,

Page 25: Sustainable atmospheric ammonia synthesis and nitrogen

14

Bai, & Ning, 2003; Matsumoto, 1981). Furthermore, the formation of radical

species by the elemental impacts in the plasma region is the essential step in the

reaction (Mizushima, Matsumoto, Sugoh, Ohkita, & Kakuta, 2004). Previous

studies proposed a detailed mechanism that can be described by the following

reaction, with the dissociation of the injected N2 gas being the rate-limiting step

(Baldur Eliasson & Kogelschatz, 1991).

N2 → 2N∗ (2.3)

H2 → 2H∗ (2.4)

N∗ + H∗ → NH (2.5)

NH + H∗ → NH2∗ (2.6)

NH2∗ + H∗ → NH3 (2.7)

Amongst the reactive plasma species shown above, it is believed that the NH

radical is the precursor of NH3 formation and its concentration has a strong effect

on the ammonia synthesis rate (Uyama & Matsumoto, 1989). In a recent study

performed by (Hong et al., 2017), the authors analyzed the kinetics of the NTP

ammonia synthesis by taking into account both electron and vibrational kinetics

and surface reactions of the adsorbed species. The authors confirmed that the

reactions between radicals and vibrationally-excited molecules had a more

significant affect on the reaction kinetics of ammonia synthesis than the ion-ion

reactions. Additionally, the authors identified several other important factors in the

NTP-mechanism including: electron excitation and vibrational excitation reactions

Page 26: Sustainable atmospheric ammonia synthesis and nitrogen

15

in the gas phase, and surface adsorbed radicals (important in the later stages of

forming NH3). The general mechanisms described by equations (2.3) to (2.7) can

then be broken down into more detailed gas-phase (Carrasco, Jiménez-Redondo,

Tanarro, & Herrero, 2011) and surface adsorption (Hong et al., 2017) reactions.

Some examples of the reactions in the gas phase and surface adsorption phase

are specified in equations (2.8) to (2.9) and (2.10) to (2.11), respectively. In these

equations, surf stands for the catalyst adsorption surface; (s) stands for the

surface-adsorbed species; and (υ) represents the vibrationally-excited molecules.

Examples of the synergistic effect between the gas phase and the surface

adsorbed species are represented by equations (2.12) and (2.13).

e + N2 → e + 2N (2.8)

N + H2(υ) → H + NH (2.9)

N + surf → N (s) (2.10)

H + surf → H (s) (2.11)

NH + H(s) → NH2(s) (2.12)

NH2(s) + H(s) → NH3 (2.13)

In the gas-phase, to weaken and break the triple bond of diatomic nitrogen,

electrons must pass into the anti-bonding orbital of N2 through the d orbital

(Henrici‐Olivé & Olive, 1969). The energy provided by the NTP could then break

the weakened nitrogen triple bond. However, NTP alone could not provide

sufficient energy for efficient nitrogen fixation. Based on the specific type of

Page 27: Sustainable atmospheric ammonia synthesis and nitrogen

16

discharge, the energy provided by NTP ranges from 3 to 6eV, which is insufficient

to ionize the Ru catalyst (ionization energy of 7.36eV) (Alexander Fridman &

Kennedy, 2004; Kerpal, Harding, Lyon, Meijer, & Fielicke, 2013). For the surface

adsorption reactions, they are the primary mechanism for NH3 (or NOx) synthesis

since the previously discussed gas-phase reactions are capable of providing the

excited molecules necessary for surface reactions. Under the NTP discharge

conditions, the nitrogen and hydrogen gas form N, H, and NHx radicals. Within the

discharge volume, these radicals will react with the surface of the catalysts via a

direct adsorption process and stick to the surface of the catalysts (Carrasco et al.,

2011). In the gas phase, the atomic nitrogen formed by the electron-excited

diatomic N2 reacts with the vibrationally-excited hydrogen molecules to form NH

radicals. These NH radicals can then form NH3 via reaction with the atomic

nitrogen molecules in the gas-phase or the ones adsorbed by the catalyst surface

(Hong et al., 2017). This general mechanism agrees with the previously discussed

importance of specific energy input and active surface area (i.e. specific energy

input provides high-energy electrons and active surface area dictates surface

reactions). Therefore, whenever a ruthenium-based catalyst is selected as an

ammonia synthesis catalyst without the presence of intense temperature and

pressure, promoters must be present to facilitate the electron transfer and

enhance the catalytic activity. In summary, it is the interactive reactions between

these two phases that form the synergistic synthesis of NH3 by non-thermal

plasma (Neyts et al., 2015). Therefore, the development of the catalyst is crucial

in two aspects, to promote the dissociation of diatomic nitrogen in the gas-phase

Page 28: Sustainable atmospheric ammonia synthesis and nitrogen

17

and to facilitate the surface adsorption reactions.

Page 29: Sustainable atmospheric ammonia synthesis and nitrogen

18

2.4 Reactor development

As mentioned in section 1, one of the advantages of the NTP synthesis of

ammonia is that it could be operated under low temperature and pressure. The

proper reactor design could significantly improve the efficiency of the ammonia

production system and lead to a higher production to energy input ratio, which

further represents a cleaner production of ammonia. Previous studies have

explored a wide array of NTP reactors as means of ammonia synthesis. When the

technology first started to emerge, Uyama et al. (Uyama et al., 1993) attempted

the NTP synthesis of ammonia using microwave and RF plasma. These

processes were achieved at approximately 700 Pa, much lower than atmospheric

pressure. Due to the vacuum condition required to carry out the reactions, the

experiments were carried out in a batch reactor. To eliminate the vacuum

requirements and operates the system in a continuous fashion, the studies

published in recent years have been largely focused on the synthesis of ammonia

using dielectric barrier discharge (DBD) plasma except those conducted by

(Nakajima & Sekiguchi, 2008).

Figure 3 shows a schematic diagram of a typical DBD plasma reactor, also

known as a silent discharge reactor. The structure of the DBD reactor is

composed of two electrodes and a dielectric barrier. As shown in the figure below,

the dielectric barrier is placed between the high voltage electrode and the ground

electrode. For DBD to occur, a strong electric field is required. The distance

Page 30: Sustainable atmospheric ammonia synthesis and nitrogen

19

between the electrodes and the applied voltage dictate the strength of the electric

field. The optimum voltage frequency to generate DBD is between 1 kHz and 10

MHz (Kogelschatz, 2003).

The discharge occurs and the plasma is generated in the small gap between

the two electrodes. Among the plasma treatment technologies, DBD plasma

offers the advantages of low operational/maintenance cost (Conrads & Schmidt,

2000; C. Liu, Brown, & Meenan, 2006), and has the broadest range of operational

pressures, ranging from 5 to 105 Pa (Schütze et al., 1998; Q. Wang, Chen, Jia,

Chen, & Li, 2011). For example, a recent comparison between DBD plasma and

microwave-induced argon plasma showed that DBD plasma generates higher

electron density and atomic oxygen concentration with less temperature increase

(Florian, Merbahi, Wattieaux, Plewa, & Yousfi, 2015).

Figure 3 A schematic diagram of Dielectric Barrier Discharge plasma generation

Page 31: Sustainable atmospheric ammonia synthesis and nitrogen

20

By using DBD plasma, most studies on NTP ammonia synthesis after the 20th

century were able to achieve continuous operation under atmospheric pressure.

(Mingdong Bai, Bai, Zhang, & Bai, 2000) used a planner structured DBD reactor

to synthesize ammonia. The study at atmospheric pressure enhanced the

conversion of ammonia and most importantly, greatly shortened the gas residence

time from the scale of hours to the scale of milliseconds. The authors further

made an improvement to the planner structured system by reducing the discharge

gap to the level of 10-1 millimeters (Mindong Bai, Zhang, Bai, Bai, & Gao, 2008). In

two more recent studies, Gómez et al. (2015, 2016) introduced a ferroelectric

packed bed reactor for the NTP synthesis of ammonia. The reactor consisted of

two circular plain electrodes with catalysts packed between them. The authors

pointed out that by using this design; the effect of the electrode gap or the size

and shape of the inner-electrode could be investigated more flexibly than the

planner or tubular electrodes. The results from this study showed a conversion of

nitrogen to ammonia at approximately 8%. Note that in this study, the ferroelectric

material acts both as the catalyst and electrode, which is later implemented in a

more recent study that uses a wool-like copper wires as both the catalyst and

electrode in a tubular NTP reactor (Aihara et al., 2016). Since the current N2

conversion of the continuous NTP ammonia synthesis is low (less than 10% of N2

conversion), Due to the low conversion rate of most NTP ammonia synthesis

technologies, a method of recycling the gas streams could contribute to material

saving and lowering the capital cost of this technology, which further promotes the

clean production of ammonia.

Page 32: Sustainable atmospheric ammonia synthesis and nitrogen

21

Another significant part of the reactor development was the selection and

design of the electrical discharge systems and operating conditions. Proper

matching between the electrical discharge power supply and the reactor was

essential to reduce the energy input of the system by lowering the specific energy

input to reach the same plasma intensity. For the electrical discharge conditions,

the alternate current (AC) and direct current (DC) discharges have been used for

the NTP synthesis of ammonia. In general, AC was used more commonly for DBD

plasma, as DC cannot pass the insulated dielectric layer (Kogelschatz, 2003).

However, it was worthwhile to point out that the studies carried out by (Mindong

Bai et al., 2008; Mindong Bai et al., 2003) used DC as their power supply. Table 2

further illustrated AC power’s superiority to DC power as the greatest energy

efficiencies and conversions were achieved by studies carried out using AC

power as opposed to DC power. Previous research indicated that the current of

the DC DBD plasma spikes when the voltage reached a certain value, usually 5kV

to 8kV depending on the specific dielectric layer (Ishida, Hirasawa, Dozen, & Tada,

2011). The dramatic increase in the operating current led to a much greater

energy consumption. For this reason, the NTP ammonia synthesis studies that

used AC power were conducted at higher voltages than their DC counterparts.

Besides the energy efficiency, the different voltage ranges used in NTP studies

are also associated with the plasma intensity, electron density, and ionization

potentials (Keller, Rajasekaran, Bibinov, & Awakowicz, 2012). For DBD plasmas,

the main collision of the ionized gas is with neutrals (Paliwoda, 2016). The

collision intensity and frequency of the plasma was closely related to the electron

Page 33: Sustainable atmospheric ammonia synthesis and nitrogen

22

temperature, which was proportional with the driving voltage (Naz, Ghaffar,

Rehman, Shahid, & Shukrullah, 2012). Research also showed that for AC power

at higher frequencies (over 10 kHz), the N2 breakdown voltage and the power

consumed per cycle of the DBD plasma was significantly lower than at low

frequency regions (less than 100 Hz) (Khatun, Sharma, & Barhai, 2010). The

plasma intensity was directly related to the effects of the feeding gas

compositions in the DBD ammonia synthesis reactor, which would be discussed in

detail in Section 6.

Page 34: Sustainable atmospheric ammonia synthesis and nitrogen

23

Table 1 A summary of the reactor development for NTP ammonia synthesis

Plasma

type

Plasma

parameter

Pressure Temperatur

e

Reactor

time

Greatest

reported

energy

efficiency

Greatest

Conversion

Comments Referenc

e

Glow

dischar

ge

DC, 6 mA 5-10 torr Room

temperature*

Batch N/A 77.4% for

MgO & CaO;

no ammonia

detected for

Al2O3,

WO3, and

SiO2-Al2O3

Batch

process, 30

min

(Sugiyam

a et al.,

1986)

Microw

ave

plasma

13.56 MHz to

2540 MHz,

150W

5 Torr 600-700 K Batch 0.078

g/kWh**

N/A Batch

process. 2

hrs

(Tanaka,

Uyama, &

Matsumot

o, 1994;

Page 35: Sustainable atmospheric ammonia synthesis and nitrogen

24

Uyama et

al., 1993)

DBD

dischar

ge

DC,

140kV/cm,

1us pulse

width,

5000Hz

750 Torr Room

temperature*

Continuous N/A 0.5 % Continuous

process,

residence

time 1.90s

(Mingdon

g Bai et

al., 2000)

DBD

dischar

ge

DC, 80us

pulse width,

1800V to

2150V,

10kHz

750 Torr

78 to 155 C Continuous 1.83 g/kWh N/A Continuous

process,

residence

time 0.20s

(Mindong

Bai et al.,

2003)

DBD

dischar

ge

AC, 2.5 to

4.5kV,

21.5kHz

750 Torr Room

temperature*

Continuous 0.34 g/kWh**

(5.5E-9

mol/J)

2.3 % Continuous

process,

residence

time 0.95s to

(Mizushim

a et al.,

2004)

Page 36: Sustainable atmospheric ammonia synthesis and nitrogen

25

3.82s

DBD

dischar

ge

DC, 240 W,

10 kHz; the

pulse width,

40 μs.

750 Torr 293K to

473K

Continuous N/A 9.1 % Continuous

process,

residence

time 1.6 s

(Mindong

Bai et al.,

2008)

Microw

ave

plasma

2.45GHz,

1.3kW

750 Torr 790K to

1240K

Continuous 0.78 g/kWh** 2.5E-2% Continuous

process,

residence

time 0.0065s

to 0.27s

(Nakajima

&

Sekiguchi

, 2008)

DBD

plasma

jet

DC

pulse,16kHz,

300W

750 Torr Room

temperature*

Simi-continu

ous

(ammonia

absorbed

after

1.16 g/kWh** 1.7E-6% at

45 min**

Semi-continu

ous

absorption

process,

Residence

(Kubota,

Masatomi

, & Tamio,

2010)

Page 37: Sustainable atmospheric ammonia synthesis and nitrogen

26

production) time 45 min

DBD

dischar

ge

AC, 1kHz, 11

to 13kV

750 Torr 450K to

1750K

Continuous N/A 4.2% Continuous

process,

residence

time 27s

(Hong,

Prawer, &

Murphy,

2014)

DBD

dischar

ge

AC, 0 to

7.5kV,

750 Torr 323K Continuous 0.9 g/kWh 2.7% Continuous

process,

residence

time 8.8 to

17.5s

(Gómez-

Ramírez,

Cotrino,

Lambert,

&

González-

Elipe,

2015)

DBD

dischar

ge

AC, 50kHz,

5kV

750 Torr Room

temperature*

Continuous 3.3 g/kWh

(H2:N2=0.5:

1)

3.5%

(H2:N2=3:1)

Continuous

process,

residence

(Aihara et

al., 2016)

Page 38: Sustainable atmospheric ammonia synthesis and nitrogen

27

time 9.7s

DBD

dischar

ge

AC, 500 to

5000Hz, 2.5

to 5.5kV,

750 Torr 333K Continuous 0.7 g/kWh 7% Continuous

process,

residence

time 9 to

121s

(Gómez‐

Ramírez,

Montoro

‐Damas,

Cotrino,

Lambert,

& Gonzá

lez‐

Elipe,

2016)

DBD

dischar

ge

AC, 500 and

1kHz, 12 to

17.5kV

750 Torr 387 K Continuous 0.16 g/kWh 2.3% Continuous

process,

residence

time 50s

(Hong et

al., 2016)

DBD AC, 7 to 14 750 Torr Room Continuous 0.68 g/kWh N/A Continuous (Xie et al.,

Page 39: Sustainable atmospheric ammonia synthesis and nitrogen

28

dischar

ge

kV, 11 to

14.5 kHZ

temperature process,

residence

time

2016)

DBD

dischar

ge

AC, 20 kV,

20 kHz

750 Torr 413 K Continuous N/A 12% Continuous

process,

residence

time 4 min

(Akay &

Zhang,

2017)

DBD

dischar

ge

AC, 5 kV, 50

kHz

750 Torr Room

temperature

Continuous N/A N/A Continuous

process,

residence

time 9.7 s,

(M.

Iwamoto,

Akiyama,

Aihara, &

Deguchi,

2017)

* Only the gas temperature before the plasma treatment is reported.

** Unit conversion is required from the original result to g/kWh

Page 40: Sustainable atmospheric ammonia synthesis and nitrogen

29

2.5 Catalyst development

Catalyst development is another critical factor to improve the NTP ammonia

synthesis process and achieve a cleaner production. To facilitate the NTP

ammonia synthesis reaction and enhance its efficiency, various efforts have been

made to the catalyst development for this technology. The first catalysts used to

synthesize ammonia under plasma conditions were iron and molybdenum wires

(Tanaka et al., 1994; Uyama et al., 1993). Other studies have examined the

performance of catalysts loaded on metal oxide supports, Al2O3 (Mizushima et

al., 2004)and MgO (Mingdong Bai et al., 2000; Mindong Bai et al., 2003), in

non-thermal plasma reactors for ammonia synthesis. Table 3 summarizes the

catalyst used for the NTP synthesis and their maximum energy efficiency or

conversion rate reached. It can be seen from the table that there are several key

factors for the catalyst selection of the NTP synthesis of ammonia.

The first factor is the presence of catalyst. Although many would believe that

catalyst is a critical factor in the NTP synthesis of ammonia, there are three

studies that successfully generated ammonia without the presence of any catalyst.

For the first study without catalyst, (Nakajima & Sekiguchi, 2008) relied on using

the quenching gas including Ar and He. This study was able to generate ammonia

at an energy efficiency of 0.78 g/kWh (best reported value). The authors also

stated the importance of using quenching gas to improve the synthesis of

ammonia. For the second study, (Mindong Bai et al., 2008) successfully

Page 41: Sustainable atmospheric ammonia synthesis and nitrogen

30

synthesized ammonia at a relatively high conversion rate (9% H2 equivalent)

using CH4 and N2. For these two studies, the reactant species and reactant

composition played another key role and the details of the gas reactants will be

discussed in the later section. For the other study without catalyst, the authors

used a liquid solution containing water and ethanol to immediately absorb the

ammonia after reaction in the plasma jet and achieved a relatively comparable

energy efficiency (1.16 g/kWh) (Kubota, Masatomi, et al., 2010).

It has been understood that the decomposition of ammonia, which is very

likely to happen under NTP conditions, is one of the limiting factors of the NTP

synthesis (Lu et al., 2014). Therefore, by injecting the reactant and product

mixture right into the solution immediately after the reduction, the authors were

able to separate and collect ammonia and prevent it from decomposing. If this

idea is adopted in future studies with more effective catalysts, it is reasonable to

achieve a breakthrough on the synthesis efficiency.

The second factor is the type of catalysts. So far, for all the studies of plasma

synthesis of ammonia, metallic catalysts have been used except several studies

that did not introduce any catalyst. The metallic catalyst can be divided into

several types as well. The first is a mono catalyst system, where only one type of

metal is used, which in these cases are MgO (Mingdong Bai et al., 2000; Mindong

Bai et al., 2003), Cu wires (Aihara et al., 2016; Tanaka et al., 1994), and Lead

zironate titanate (Gómez-Ramírez et al., 2015; Gómez‐Ramírez et al., 2016).

Among the studies that used mono catalysts, the copper wires with wool-like

shape presented the greatest energy efficiency (Aihara et al., 2016). Note that in

Page 42: Sustainable atmospheric ammonia synthesis and nitrogen

31

this study, the copper wools act not only as the catalyst, but an electrode of the

NTP discharge system as well. The authors further performed a comparison

analysis between different wool-like metallic catalysts including Au, Pt, Pd, etc.

This study shows that for wool-like metallic catalysts, Au has the highest activity

(M. Iwamoto et al., 2017). The second type is the catalyst-support system, which

coat catalyst (ruthenium and carbon) on to supporting materials including MgO

and Al2O3 (Hong et al., 2016; Mizushima et al., 2004; Xie et al., 2016). Another

study performed by (Akay & Zhang, 2017) uses microporous silica to support

nickel catalyst to enhance the surface area. For this catalytic system, promoters

are used to enhance the synthesis efficiency and conversion rate.

The third factor is the shape and structure of the catalyst. As discussed in the

previous section, the reactors used in the plasma synthesis of ammonia were

fixed-bed reactors with gas-phase reactions. Therefore, the shape and structure

of the catalyst was crucial as they could influence the flow channels, reaction time,

surface area, number of active sites, etc., which were all critical parameters for

the plasma synthesis of ammonia. More specifically for DBD plasmas, shielding

effects and discharge with between the catalyst and electrodes should also be

taken into consideration. From table 3, it can be seen that the shape of the

catalyst used in the previous studies were powders, wires, and pellets. Prior to

2014, the study on NTP synthesis of ammonia mainly used powder and wire

catalysts. Pellet-shaped catalyst were introduced and used in most studies after

2014. As (H. H. Kim, Teramoto, Ogata, Takagi, & Nanba, 2016) mentioned,

powder catalysts are not suitable for plasma process because they tend to spread

Page 43: Sustainable atmospheric ammonia synthesis and nitrogen

32

under plasma conditions. According to other literature, the spreading of the

powdered catalyst could be due to the charging and static electricity during the

plasma treatment (Mazumder et al., 2006). In a later study, carried out by (Aihara

et al., 2016) mentioned in the previous paragraph, the authors claimed that the

wool-like copper catalyst could increase the surface area of the electrode,

allowing for more discharge, and enabling more effective ammonia synthesis over

the electrode discharge.

Page 44: Sustainable atmospheric ammonia synthesis and nitrogen

33

Table 2 A summary of the catalyst development for NTP ammonia synthesis

Catalyst type Catalyst Catalyst shape Reference

Mono catalyst MgO &

CaO Al2O3;

WO3, and

SiO2-Al2O3

Disks (Sugiyama et al.,

1986)

Mono catalyst Iron and

molybdenum

Wires (Tanaka et al., 1994;

Uyama et al., 1993)

Mono catalyst MgO Powders (Mingdong Bai et al.,

2000)

Mono catalyst MgO Powders (Mindong Bai et al.,

2003)

Catalyst-support Ru/ Alumina Powders (Mizushima et al.,

2004)

N/A No catalyst N/A (Mindong Bai et al.,

Page 45: Sustainable atmospheric ammonia synthesis and nitrogen

34

2008)

N/A No catalyst N/A (Nakajima &

Sekiguchi, 2008)

N/A No catalyst N/A (Kubota, Masatomi,

et al., 2010)

Mono catalyst MgO and glass

spheres

Pallets (Hong et al., 2014)

Mono catalyst Lead zirconate

titanate, BaTiO3

Pellets (Gómez-Ramírez et

al., 2015)

Electrode catalyst Copper Small wool-like wires (Aihara et al., 2016)

Mono catalyst Lead zirconate

titanate (pallets)

Pallets (Gómez‐Ramírez

et al., 2016)

Catalyst-support Nanodiamonds and

diamond-like carbon

coated Al2O3

Sphere powders (Hong et al., 2016)

Catalyst-support Ru on Al2O3 Powders (Xie et al., 2016)

Page 46: Sustainable atmospheric ammonia synthesis and nitrogen

35

Catalyst-support Ni on microporous

silica

Pallets (Akay & Zhang,

2017)

Mono catalyst Au, Pt, Pd, Ag, or Cu Small wool-like wires (M. Iwamoto et al.,

2017)

Page 47: Sustainable atmospheric ammonia synthesis and nitrogen

36

2.6 Reactant type, composition, and feed rate

2.6.1 The effect of N2 and H2 composition and feed rate

From table 2, it can be seen that the studies with the highest conversion do

not correlate with the highest energy efficiencies. Several studies demonstrated

that the greatest ammonia synthesis occurs with a feed-gas composition equal to

(Mindong Bai et al., 2003; Gómez‐Ramírez et al., 2016), or nearly equal to, the

stoichiometric ratio in equation (2.1) (Gómez-Ramírez et al., 2015; Mizushima et

al., 2004). The increased hydrogen content promotes an increase in both electron

density and temperature, which leads to a positive shift, or increase, in the mean

electron energy distribution. This shift favors the formation of the NH species in

equations 2.3 to 2.7. For these studies, the plasma discharge conditions were

relatively less intense, utilizing lower voltages and frequencies (below 5 kHz and

500 kV/m). However, there are several studies that reported increased ammonia

synthesis under high N2 compositions (Aihara et al., 2016; Mingdong Bai et al.,

2000; Hong et al., 2016; Nakajima & Sekiguchi, 2008). For these studies, the

gas-phase discharge conditions are relatively high, above 500 kV/m and 10 kHz.

According to the recent kinetic modeling study, the effects of the N2 and H2

composition on the continuous NTP synthesis of ammonia depend on several

factors, including the electron temperature and density (Hong et al., 2017). It is

mentioned in this study that N2 compositions higher than the stoichiometric value

in equation (2.1) are favored under high-density plasma conditions with

Page 48: Sustainable atmospheric ammonia synthesis and nitrogen

37

high-energy electrons. The same trend was observed in the study using thermal

plasma to synthesize ammonia as well (Van Helden et al., 2007).

The second factor affected by the reactant feed rate is the residence time.

Residence time also has a significant impact on ammonia synthesis, especially for

the continuous NTP packed-bed reactors. From table 2, it can be observed that

the range of residence time in the NTP ammonia synthesis studies varied from the

scale of milliseconds to a minute. The study from (Gómez-Ramírez et al., 2015)

indicated that longer residence time would result in less ammonia production. This

concept is supported by several other studies that specifically explored the

dissociation of ammonia under plasma conditions (Lu et al., 2014; L. Wang, Zhao,

Liu, Gong, & Guo, 2013). These studies revealed that non-thermal plasma could

also promote the decomposition of ammonia back to hydrogen and nitrogen gas,

especially with the presence of iron-based catalysts (L. Wang et al., 2013).

Therefore, decreasing the residence time of the feed-gas within the NTP region

helps prevent efficiency diminishing back reactions, such as ammonia

decomposition.

2.6.2 Reactants other than N2 and H2

Similar to the conventional Haber-Bosch process, most of the NTP

ammonia synthesis studies used N2 and H2 gas as the reactants. However, there

are two exceptions. Overall, the attempts of investigating in reactants other than

N2 and H2 is to explore the possibilities of improving the efficiency for the NTP

ammonia synthesis approach and enhance its environmental sustainability.

Page 49: Sustainable atmospheric ammonia synthesis and nitrogen

38

However, there is one study that has the potential of leading to a promising

alternative to combine the NTP process with other waste treatment technologies

for a cleaner production of ammonia. The study carried out by (Mindong Bai et al.,

2008) used CH4 and N2 as feed gas for NTP ammonia synthesis. Although there

were no catalysts used in this study, the authors managed to produce ammonia

with a maximum conversion of 0.8%. The authors did not report the synthesis

efficiency data. But it is notable that in this study, many CxHy species were formed

during the reaction from the dissociation of CH4, such as C2H6, C2H4, C3H8, etc.

Although this study is only at preliminary stage, its results provide a potential

alternative for the clean production of ammonia. As CH4 is a greenhouse gas

emission during anaerobic digestion, used at many waste water treatment

facilities (Z.-Y. Ma et al., 2015), its use for synthesizing ammonia could

significantly improve the sustainability of the waste treatment industry. The

second study used N2 gas with water and ethanol solution. Note that in this study,

the water and ethanol solution used act not only as the reactants, but also

absorbents for the synthesized ammonia. Therefore, the ammonia produced in

this study was in the form of aqueous NH4+. Like the previous study that used N2

and CH4 as the reactants, this study also led to the formation of side products,

which were aqueous NO2- and NO3

-.

Although N2 and H2 were used as the reactants, the two studies carried out

by (Nakajima & Sekiguchi, 2008) and (Hong et al., 2014) incorporated the idea of

introducing a “quenching gas” into the system. The authors suggested that due to

the exothermic nature of ammonia synthesis, injection of a quenching gas could

Page 50: Sustainable atmospheric ammonia synthesis and nitrogen

39

effectively lower the temperature of the reactor, which could further enhance the

ammonia synthesis by preventing the decomposition of ammonia. This study

demonstrated an increase in ammonia production with the introduction of a

quenching gas. As a result, the study by (Hong et al., 2014) managed to achieve a

4.2% conversion to ammonia, which is one of the highest conversions yet

reported amongst NTP ammonia synthesis studies. With the exception of the

three studies just discussed, a review of literature demonstrated an almost

unanimous use of N2 and H2 as reactant feedstock. However, despite the

continuity in choice of feedstock, these studies exhibited remarkably different

results when testing gas feed rates. Of note, an additional study that utilized

expanding thermal plasma (ETP) in the synthesis of ammonia from N2 and H2

also confirmed the positive effects of introducing Ar as a quenching gas (Van

Helden et al., 2007).

Page 51: Sustainable atmospheric ammonia synthesis and nitrogen

40

2.7 Conclusions

In summary, the current development of this technology is mainly focused on

the kinetic study, reactor/catalyst development, and reactant composition studies.

For the kinetics studies, most of the efforts have been made trying to identify

the rate-limiting step of the plasma ammonia synthesis and classify the complex

plasma reactions. For reactor development, reactors from rectangular,

ferroelectric, to tubular DBD discharge reactors, etc. have been explored. System

parameters such as discharge conditions and residence time are the crucial factor

to develop and design efficient reactors for this technology and realize cleaner

production of ammonia with less energy input. For catalyst development,

non-catalyst, mono catalyst, catalyst-support, and catalyst-support with promoters

were investigated. During the investigation of the reactant composition, the effects

of changing the N2 to H2 ratio were studied. However, several preliminary studies

revealed the possibility of using CH4 or a liquid-gas reaction to synthesis ammonia

under NTP conditions. These studies provided a potential avenue of combining

the NTP technology with other waste treatment processes for a cleaner and more

environmental sustainable ammonia production industry.

From the results obtained in previous studies, the main barriers faced by this

technology were identified as the fixation of nitrogen gas and prevention of

unwanted back reactions (ammonia decomposition). To overcome these

obstacles, reactors with instantaneous product absorption and separation are

Page 52: Sustainable atmospheric ammonia synthesis and nitrogen

41

recommended, along with catalysts of stronger plasma synergistic activities.

Furthermore, development of a catalyst with higher activities and introduction of a

fast-separation reactor, using solid or liquid sorbents, could potentially solve the

challenges and ensure these technologies competitiveness. Despite these

challenges, this technology’s remarkably low theoretical energy floor provides an

enticing opportunity for future use in industrial process. Its other advantages, such

as low capital cost, less land use, small-scale accessibility, offer this technology

many opportunities to reshape the ammonia production industry towards a

greener and cleaner process.

Page 53: Sustainable atmospheric ammonia synthesis and nitrogen

42

CHAPTER 3 Atmospheric plasma-assisted

ammonia synthesis using Ru-based

multifunctional catalyst supported by MgO

3.1 Introduction

The concept of synthesizing ammonia at atmospheric pressure through the

non-thermal plasma (NTP) assisted catalytic reactions was proven (Mingdong Bai

et al., 2000; Mindong Bai et al., 2003; Lan, Irvine, & Tao, 2013; Mizushima et al.,

2004). NTP is electrically energized matter in a gaseous state, which is not in

thermodynamic equilibrium, and can be generated through electric discharge in a

gaseous volume. The NH radicals formed in the nitrogen-hydrogen plasma and

the NH radicals would be quenched by hydrogen molecules to form ammonia

(Chauvet, Thérèse, Caillier, & Guillot, 2014; H. Ma et al., 2002). The

non-thermal plasma species have an energy level in the range of 2 to 10 eV at

temperature close to ambient condition. It is known that these kinds of species

can alter organic or inorganic compounds through at least three mechanisms: (1)

decomposition, (2) structural rearrangement, and (3) fragment elimination (B.

Penetrante et al., 1997; Van Durme et al., 2008). One or all of these

mechanisms may be responsible for promoting the disassociation of hydrogen

and nitrogen, which is necessary for the synthesis of ammonia through catalytic

Page 54: Sustainable atmospheric ammonia synthesis and nitrogen

43

reactions. Ru based catalysts have been widely used and proved efficient in

conventional ammonia synthesis, especially under the presence of MgO or

carbon nanotube support and Cs promoter (Larichev, 2010; Y. Yan et al., 2015).

Figure 4 provides a general illustration of an electron passing onto the

anti-bonding orbital of N2 through the d orbital of Ru in order to weaken the triple

bond of nitrogen. The weakened triple bond can then be broken with additional

energy, in this case, energy provided by NTP. However, the energy provided by

NTP is approximately 3 to 6eV according to the specific type of discharge, which

is insufficient to ionize Ru, which has ionization energy of 7.36eV (Alexander

Fridman & Kennedy, 2004; Kerpal et al., 2013; B. Penetrante et al., 1996; B. M.

Penetrante, Hsiao, Merritt, Vogtlin, & Wallman, 1995). The situation changes

when promoter Cs is attached to Ru. Cs can easily be ionized, producing

electrons that are passed onto Ru, and then to nitrogen (J. Iwamoto, Itoh, Kajita,

Saito, & Machida, 2007; Kitano et al., 2012). Therefore NTP on one hand directly

causes N2 and H2 to dissociate and form NH3 with or without catalyst, and on the

other hand, provides ionization energy necessary to produce electrons for the

catalysis system to function (Neyts et al., 2015). The promoters used in these

studies are alkaline metals, typically Cs, Ba, and K. Studies have shown that for

these promoters are able to increase the catalytic activity of the catalytic system

by effectively decreasing the energy barrier for the dissociation of N2, and facilitate

the destabilization of the NHx species in equations generated during the plasma

synthesis of NH3 (Aika, 2017; Z. Ma, Zhao, Pei, Xiong, & Hu, 2017). Attaching

promoters such as Cs and Ba to Ru can enhance the catalytic conversion process.

Page 55: Sustainable atmospheric ammonia synthesis and nitrogen

44

Since Cs can easily be ionized, it can pass electrons onto Ru and nitrogen, which

will aid the dissociation of nitrogen gas (J. Iwamoto et al., 2007). The high applied

frequency and the metallic cation promoters added could enhance the electron

transfer within the catalyst systems to aid the dissociation and ionization of

nitrogen, which requires the highest ionization potential among the reactions

(Mindong Bai et al., 2003; Larichev, 2010). The fact that the Ru catalysts with Cs

promoters has much higher dissociative catalytic activity on N2 than catalysts

without C was proved by a study that used an isotropic analysis on the kinetic of

the ammonia synthesis (Hinrichsen, Rosowski, Hornung, Muhler, & Ertl, 1997). A

relatively more recent kinetic study pointed out that in this catalyst-promoter

ammonia synthesis system, the cesium and barium each acted as an electronic

promoter and structural promoter that controls the concentration of active sites

(Szmigiel et al., 2002).

Figure 4 Dissociation of nitrogen under NTP

Page 56: Sustainable atmospheric ammonia synthesis and nitrogen

45

3.2 Materials and methods

The catalysts used for synthesis study were loaded with Ru (10 wt%) as

the main catalyst. Cs, K, and Ba were used as promoters. After impregnation, the

catalysts were dried under infrared light for 4 hr and then calcined at 500 °C in a

muffle furnace for 4 h. Then the catalysts were reduced at 500 °C using hydrogen

gas for 4 h prior to application. The NTP synthesis reactor consisted of two

co-axial quartz tubes and a stainless steel tube, which act as the dielectric barrier

and the discharging electrodes, respectively. Figure 5 shows the process flow

diagram of the NTP ammonia synthesis system used in this study. The AC high

voltage to the system was supplied using an SSD-110 inverter and a coupled

transformer manufactured by Plasma Technics Inc. (Wisconsin, USA). The space

between the dielectric barrier and the electrode was 1.5 mm and the length of the

reactor is 30 cm. Catalysts were packed to the same length matched with the

length of the electrode between the two dielectric barriers (20 mm). When a

certain voltage is applied to the electrode, electrical discharge takes place

between the dielectric barriers and on the surface of the catalyst particles. The

voltage and frequency of the power supply could be adjusted during synthesis

process and was measured using a Tektronix oscilloscope (Oregon, USA)

connected to the high voltage electrode of the reactor. The temperatures were

measured using an inferred thermometer (Fluke, USA). The synthesis efficiency

was determined by dividing the ammonia synthesis rate (g/hr) over the power

consumption (kW) at the DBD region. The power consumption was calculated by

Page 57: Sustainable atmospheric ammonia synthesis and nitrogen

46

multiplying the voltage and current at the discharge region. The current was

determined using coil coefficient provided of the transformer (provided by the

manufacture), and the current from the inverter side, which was measured using a

Sperry DSA-540A current meter (USA). Nitrogen and hydrogen gases were fed

from compressed gas cylinders at approximately 26 C and pressure 60 psi. The

gas flow rate is controlled using BROOKS Smart Mass Flow gas flow meters. The

ammonia was qualitatively detected using FTIR. The product gas was collected

using dilute sulfuric acid solution and the quantitative measurements were carried

out using a DR 5000™ UV-Vis Spectrophotometer (Hach Technology Inc. USA).

Page 58: Sustainable atmospheric ammonia synthesis and nitrogen

47

Figure 5 a) Process flow diagram of the NTP ammonia synthesis approach. b) A simplified diagram of the wiring connection between the invertor and transformer

The energy efficiencies listed in Chapter 2 was referring to using the “plasma

power” measured at the point of plasma discharge. The most common

measurements of “plasma power” were the analytical approach, current and

Page 59: Sustainable atmospheric ammonia synthesis and nitrogen

48

voltage based methods, the Lissajous approach, and a comparative approach.

The descriptions of these methods was reported in previous literature (Hołub,

2012). The method used in this thesis research to characterize energy efficiency

was a derivative of the current and voltage based methods. Figure 5 b) showed a

simplified wiring diagram of between the invertor and the transformer.

The measured parameters used to determine the power consumption were

the voltage on the secondary side and current on the primary side of the invertor.

The secondary voltage, V2 was measured using the Trek high voltage probe and

the primary current, A1 was measured using the Sperry DSA-540A current meter.

The energy efficiency in g/kWh was estimated using the following equation

𝜂(𝑒𝑓𝑓𝑖𝑐𝑖𝑒𝑛𝑐𝑦) = �̇�∙�̅�

𝐴1𝑅𝑀𝑆∙𝑁1𝑁2∙𝑉2𝑅𝑀𝑆

(3.1)

Where �̇� was the synthesis rate in mol/hr, �̅� was the molecular mass in g/mol,

N1 and N2 were the coil numbers on the primary and secondary sides of the

transformer, and RMS was the room mean square value of the current or the

voltage.

Page 60: Sustainable atmospheric ammonia synthesis and nitrogen

49

3.3 Results and discussions

3.3.1 Synergy between NTP, catalysts, and promoters

Figure 6 shows the ammonia yields using different catalysts under NTP

environment. It is indicated that NTP alone was able to promote ammonia

synthesis, suggesting that NTP species provided energy to dissociate N2 and H2

and allowed the formation of NH3 under atmospheric conditions. Additionally,

ammonia yield is improved with the presence of the MgO support and increased

dramatically when both promoter Cs and MgO support are introduced. Therefore,

it is reasonable to conclude that NTP, Ru/MgO and Cs promoters are critical

components in this NTP reaction system. Previous studies shows that the

presence of MgO particles can act as supporting material for Ru based catalysts

and is able to promote intensive surface discharges, which are expected to favor

dissociation of N2 and H2 and promote ammonia formation (Mingdong Bai et al.,

2000; Sugiyama et al., 1986). However, the catalyst packing affects the

characteristics of the discharge. Since the high dielectric constant of MgO could

intensify the discharge, it is possible that the yield increase with Ru/MgO is

caused by the change within the plasma discharge. Similarly for the yield increase

when the system includes both MgO support and Cs promoter under the same Ru

load. Since the addition of Cs promoter causes the formation of an electric double

layer between Cs+ cations and Ru atoms, the work function of ruthenium is

decrease and discharge conditions could be altered (Larichev, 2010; Larichev et

Page 61: Sustainable atmospheric ammonia synthesis and nitrogen

50

al., 2007).

Figure 6 Ammonia synthesis under different conditions.

3.3.2 Feeding gas ratio & flow rate effects

N2/H2 ratio of 5:1, 4:1, 3:1, 2:1, 1:1, 1:2, 1:3 were tested on the pilot system

for ammonia synthesis under NTP conditions. Results show that N2/H2 ratio=3:1

produces the highest amount of ammonia. This confirms that ammonia synthesis

has higher efficiency under a N2 rich environment. This also indicates

residual/unreacted nitrogen after the gas exited the reaction volume. This

indicates that reaction under the condition of higher N2/H2 ratio was more

desirable for ammonia concentration. The possible reason is that nitrogen has a

higher chemical bond energy that requires higher energy to break the bond, and

the high density of nitrogen gas in the plasma system could lead to higher density

of active nitrogen molecular. The possibility of active nitrogen molecule reacting

0

0.5

1

1.5

2

2.5

3

3.5

4

NTP NTP + MgO NTP + Ru /MgO NTP + Cs-Ru

/MgO

Am

mo

nia

yie

ld (

%)

Page 62: Sustainable atmospheric ammonia synthesis and nitrogen

51

with hydrogen would be increased. Results in Figure 8 show that higher flow rate

produced ammonia much more efficiently in the current NTP plasma catalytic

synthesis system. Tests on the pilot system show that higher gas flow rate is able

to improve energy efficiency. Note that when the synthesis efficiency reaches a

plateau at gas flow rate of 5 to 8 liters/min.

Figure 7 Effect of N2 composition on ammonia synthesis efficiency

0

0.3

0.6

0.9

1.2

0

20

40

60

80

0.25 0.33 0.50 0.67 0.75 0.80 0.83

Eff

icie

ncy

(g/

kW

h)

Rat

e (μ

mo

l∙m

in-1

)

N2 composition

Rate

Efficiency

Page 63: Sustainable atmospheric ammonia synthesis and nitrogen

52

Figure 8 Effect of gas flow rate on ammonia synthesis efficiency

3.3.4 Frequency & voltage effects

The maximum frequency and voltage that can be achieved by the power

supply are 25,000 Hz and 30,000 V respectively. Tests show that higher applied

voltage resulted in a higher ammonia concentration, which is reasonable greater

plasma energy favors the ionization of Ru catalyst, which further leads to

breakage of N2 bond. However, the trade off for voltage increase is energy

consumption. The increase in applied power requires larger energy input from the

power supply. By calculating the energy consumption associated with ammonia

synthesis rate under different voltages, results in figure 10 conclude that the

overall ammonia synthesis efficiency decreases with increasing voltage. On the

0.00

0.30

0.60

0.90

1.20

0

20

40

60

80

1 2 3 4 5

Eff

icie

ncy

(g/

kW

h)

Rat

e (μ

mo

l∙m

in-1

)

Gas flow rate (L/min)

Rate

Efficiency

Page 64: Sustainable atmospheric ammonia synthesis and nitrogen

53

other hand, frequency shows less significant effect on ammonia synthesis

efficiency. Based on experimental results of voltage and frequency analysis, we

determined that 10,000Hz and 6000v are found to provide the highest energy

efficiency for NPT ammonia synthesis.

Figure 9 Effect of discharge frequency on ammonia synthesis efficiency

Figure 10 Effect of applied voltage on ammonia synthesis efficiency

-0.10

0.10

0.30

0.50

0.70

0.90

1.10

0

20

40

60

80

100

8000 10,000 12,000 14,000 16,000

Eff

icie

ncy

(g/

kW

h)

Rat

e (μ

mo

l∙m

in-1

)

Frequency (Hz)

Rate

Efficiency

0.00

0.30

0.60

0.90

1.20

0

20

40

60

80

100

120

5000 6000 7000 8000 9000

Eff

icie

ncy

(g/

kW

h)

Rat

e (u

mo

l/m

in)

Voltage (V)

Rate

Efficiency

Page 65: Sustainable atmospheric ammonia synthesis and nitrogen

54

3.4 Conclusions

This study proved the feasibility of the synthesis of ammonia using

non-thermal plasma above micro-gap discharge at low temperature. Results also

show that N2/H2 ratio=3:1 produced the highest amount of ammonia, which further

confirms that ammonia synthesis has higher efficiency under a N2 rich

environment due to the relative difficulty in disassociation of nitrogen. Achieve

further enhancement of ammonia generation efficiency under NTP condition is

needed to match the industrial requirements of ammonia production, which could

be potentially by improving reactor configuration and new catalyst preparation.

Page 66: Sustainable atmospheric ammonia synthesis and nitrogen

55

CHAPTER 4 Atmospheric plasma-assisted

ammonia synthesis using Ru-based

multifunctional catalyst supported by

mesoporous silica MCM-41

4.1 Introduction

To enhance the ammonia synthesis efficiency of the NTP approach, an

effective catalytic-promoter system and mesoporous supporting material were

used. In this study, a heterogeneous multifunctional catalytic system is used,

which consists of different solid surface catalytic sites (Dixit, Mishra, Joshi, &

Shah, 2013). Ruthenium, one of the catalysts with the highest catalytic activity for

ammonia synthesis, is selected to be the main catalyst. Additionally, the Cs and

Ba promoters were also used, which facilitates the NH3 synthesis by being ionized

under relatively low energy and transporting electrons onto Ru and N. The Ru

catalyst and promoters were deposited onto a Si-MCM-41 support to form the

Ru-promoter/Si-MCM-41 catalyst system.

Page 67: Sustainable atmospheric ammonia synthesis and nitrogen

56

4.2 Materials and methods

4.2.1 Improvements

The mesoporous-structured Si-MCM-41 was chosen in this study for

several reasons. First, mesoporous materials are notable for having a large

surface area to volume ratio when compared with traditional metal oxide supports

(Erjavec, Kaplan, Djinović, & Pintar, 2013). Second, the highly ordered pores with

diameters from 10 nm to 50 nm make the Si-MCM-41 material an ideal host for

the Ru and promoters (Gallezot, Chaumet, Perrard, & Isnard, 1997; Mondal,

Mondal, & Roy, 2016). Third, the inter-connected pores can increase the

residence time of the reactant gases in the plasma region and enhance the

plasma catalytic interactions (Mondal et al., 2016). Lastly, due to its low

conductivity, Si-MCM-41 can maintain a stable discharge environment for the

dielectric barrier discharge (DBD) plasma. On the other hand, mesoporous

catalysts, such as iron or carbon-based or nickel-based materials; tend to form

micro discharges within the pores, or between the surface of the support and

electrodes, which will lower the conversion and energy efficiency of the ammonia

synthesis (Akay & Zhang, 2016).

4.2.2 Catalyst preparation

As mentioned previously, the catalytic system used in this study was

ruthenium-based catalysts with the Si-MCM-41 supports. Liquid impregnation was

Page 68: Sustainable atmospheric ammonia synthesis and nitrogen

57

used to deposit the catalysts and promoters onto the support. The primary recipe

and preparation steps were determined from previous literature, which was also

described in the previous chapter (Kowalczyk, Krukowski, Raróg-Pilecka,

Szmigiel, & Zielinski, 2003; Saadatjou, Jafari, & Sahebdelfar, 2015). All the

chemicals used in this study were purchased at Sigma Aldrich (USA). For each

gram of Si-MCM-41 support, 0.17g of Ru(NO)(NO3)3, 0.12g of CsNO3, 0.05g of

KNO3, and 0.036g of Ba(NO3)2 were used. During liquid impregnation, 10 minutes

of ultrasulnication was used followed by 20 minutes of stirring. After impregnation,

the catalysts were dried under infrared light for 4 hr with constant stirring and then

calcined at 500°C in a muffle furnace for 4 hr. Before being used for the ammonia

synthesis reactions, the catalysts were reduced at 500°C for 4 hr under hydrogen

environment.

The calcined catalysts in this study were characterized using Fourier

transform infrared spectroscopy (FTIR), Transmission electron microscopy (TEM),

scanning electron microscope (SEM), Atomic-force microscopy (AFM), Infrared

Photo induced Force Microscopy (IR PIFM), Energy-dispersive X-ray

spectroscopy (EDS), X-ray photoelectron spectroscopy (XPS), and X-ray

diffraction (XRD). The reduced catalyst was also compared analyzed using XRD.

The FTIR spectrum was obtained under the Attenuated Total Reflection (ATR)

mode of a Nicolet™ iS™ 50 FT-IR Spectrometer. The TEM and SEM image are

taken using FEI Tecnai T12 and Ultra-high Resolution SEM SU8200 (Hitachi,

Japan). AFM and PiFM images were taken using the Vistascope Platform from

Molecular Vista (California, USA). X-ray diffraction (XRD) patterns were obtained

Page 69: Sustainable atmospheric ammonia synthesis and nitrogen

58

on a Siemens D5005 X-ray Diffractometer with sealed Cu source and the XPS

analysis was performed using the SSX-100 XPS device (Surface Science

Instruments, USA).

Page 70: Sustainable atmospheric ammonia synthesis and nitrogen

59

4.3 Results and discussions

4.3.1 Characteristics of the Catalysts

Figure 11 shows the FTIR spectra of the Ru-promoter/Si-MCM41 catalyst.

Three high intensity peaks are observed at 1100 cm-1, 460 cm-1, and 750 cm-1.

The Si-O-Si peak is observed at wavenumber 1100 cm-1 (Chaudhary, Ghatak,

Bhatta, & Khushalani, 2006). The other peak that denotes the Si-MCM-41 support

is the one at approximately 460 cm-1 (Nyquist & Kagel, 2012). The band at 600

cm-1 to 890 cm-1 represents the Ruthenium complex (Nakamoto, 1986). Figure 12

shows the results from the microscopy analysis. The mesoporous structure of the

silica Si-MCM-41 can be revealed from the SEM image. Although the catalytic

system is presented in an agglomerated form, pores are still visible in the SEM

image. As shown in the image, the micro and meso pores in the material are

mainly from 10 to 50 nm. The results found from the TEM image agree with

previous findings for the Ru-MCM complex with mesoporous structures (Vanama,

Kumar, Ginjupalli, & Komandur, 2015). The size of the ruthenium particles ranges

from approximately 10 to 20 nm, with agglomeration shown in the surface of the

Si-MCM-41 support. From the size distribution of the ruthenium particles (2 to 6

mm) (Joo et al., 2010) and the pores of the support, it could be found that the

Ruthenium particles could be deposited both inside and outside of the pores of

the Si-MCM-41 support. In fact, some of the Ru nanoparticles are present on the

outside of pores as well. These dispersion results partially indicate the strong

Page 71: Sustainable atmospheric ammonia synthesis and nitrogen

60

interactions between the Ru nanoparticles and the support, which agree with

previous findings (Makgwane & Ray, 2013; Vanama et al., 2015). As observed in

previous studies, the Si-MCM-41 structured mesoporous materials often exist in

the form of agglomerated spherical particles (Costa et al., 2014; B. Yan, Li, &

Zhou, 2009). Therefore, the AFM and PiFM images were taken, focusing on the

one agglomerated spherical particle. From both the AFM topography and PiFM

image, the particle size is approximately 500 nm in diameter, which also agree

with previous findings (Costa et al., 2014). Since the wavenumber of the PiFM is

selected based on the FTIR spectra, the different color bands in the PiFM image

correspond to different atomic interactions in the catalyst system. As observed in

the TEM image, the PiFM image further indicates the non-uniformed distribution

of the ruthenium particles. However, the microscopy images are not capable of

determining the detailed distribution of each catalytic compound. Therefore, an

elemental mapping was performed to further study the deposition pattern of the

catalyst.

Figure 11 FTIR spectra of the Ru/SI-MCM-41 catalyst

0

0.04

0.08

0.12

0.16

0.2

250 550 850 1150 1450 1750

Lo

g (

1/T

)

Wavenumber (cm-1)

Page 72: Sustainable atmospheric ammonia synthesis and nitrogen

61

Figure 12 a) SEM, b) TEM, c) AFM topography, and d) PiFM images of the Ru-promoter/Si-MCM-41 catalyst

Page 73: Sustainable atmospheric ammonia synthesis and nitrogen

62

Figure 13 EDS results of the Ru-promoter/Si-MCM-41 catalyst

The elemental mapping results represented in Figure 13 show that the Si, Ru

elements are deposited less evenly than the Cs and K elements. This difference in

the distribution is caused by how the species were differently deposited onto the

supports during the liquid impregnation process. Note that due to its small fraction,

Page 74: Sustainable atmospheric ammonia synthesis and nitrogen

63

the Ba element was not detected during several repeated analysis. This could also

be due to the fact that it has been deposited into the porous structures for the

cluster. For the other inorganic salts of KNO3 and CsNO3, they were easily

dissolved into the liquids and were deposited onto the supports during the drying

and calcine stage. Therefore, they are more likely to be deposited on the support

area in the form of the oxidized state. On the other hand, the Si-MCM-41 support

and Ru(NO)(NO3)3 forms dispersion instead of dissolving in the solution during the

liquid impregnation process. Since dispersion was formed instead of a stable

solution, the Ruthenium particles are more likely to agglomerate than the

promoters, which leads to a less uniform distribution of the element. This specific

distribution of catalysts and promoters could induce localized promotion when the

catalyst system is in service.

0

500

1000

1500

2000

2500

3000

05001000

Co

un

ts

Binding Energy (eV)

Page 75: Sustainable atmospheric ammonia synthesis and nitrogen

64

Figure 14 XPS spectrum of the Ru-promoter/Si-MCM-41 catalyst. a) survey scan. b) high resolution scan for

Ru3dC1s. c) high-resolution scan for the Cs

The XPS data shows the oxidation states of different elements within the

calcined catalyst-promoter system. Results of the survey scan and the

high-resolution scan are shown in Figure 14. The binding energy for Si2p is shown

at 90 eV. The two bands between 265 eV and 277 eV correspond to the band for

Ru3d, formed by the split spin-orbit components of Ru (K. Kim & Winograd, 1974).

Furthermore, the binding energy of the However, the two peaks of the Ru3d shown

at 267 eV and 271 eV should have an area ratio of approximately 2 to 3, while the

peaks determined by the high-resolution scan fail to match with the theoretical

values (Rochefort, Dabo, Guay, & Sherwood, 2003). The slightly lower binding

energy of Ru3d in this study could be due to the influence of the promoters on the

Ru oxide clusters, which have been reported in previous literature (Larichev, 2010).

Furthermore, one peak is found at approximately 284 eV, which corresponds to

the peak of C1s. From the peaks fitting data shown in the high-resolution scan in

Figure 14, some of the Ru3d peaks overlap with the C1s peak are observed

100

250

400

260270280290

Co

un

ts

Binding Energy (eV)

200

350

500

650

703713723733

Co

un

ts

Binding Energy (eV)

Page 76: Sustainable atmospheric ammonia synthesis and nitrogen

65

(Berthoud et al., 2008). This overlap is formed due to the carbon contamination

from the carbonation effect generated from the excessive heating rate during the

calcination process in the Si-MCM-41 structure (Dong, Fei, Zhang, & Yu, 2015;

Mun, Ehrhardt, Lambert, & Madic, 2007). Lastly, the binding energy of the Si2p

peak at approximately 100 eV is lower than the reported values (Vanama et al.,

2015; Wagner & Muilenberg, 1979). This could happen due to the fact that the

originally surrounded oxygen atoms are replaced by nitrogen, which are more

electronegative (Xia & Mokaya, 2004). Due to the high nitrogen contents of the

compounds used in the liquid impregnation process, the SiN4 environment is

more dominant compared with other Si-MCM-41 environments reported in

previous literature.

Figure 15 XRD spectrum of the calcined and reduced Ru-promoter/Si-MCM-41 catalyst

degrees 2-theta2 4 6 8 10

inte

nsity (

a.u

.)

0

50

100

150

200X R D calcined

degrees 2-theta10 20 30 40 50 60

inte

nsity (

a.u

.)

0

20

40

60

80

100

120

degrees 2-theta2 4 6 8 10

inte

nsity (

a.u

.)

0

50

100

150

200X R D reduced

degrees 2-theta10 20 30 40 50 60

inte

nsity (

a.u

.)

0

20

40

60

80

100

120

Page 77: Sustainable atmospheric ammonia synthesis and nitrogen

66

The XRD patterns shown in Figure 15 indicate the overall amorphous nature

of the catalyst system after loading Ru and the other promoters. The sharp Bragg

peak at 2.2° shows the hexagonal pore structure of the unloaded Si-MCM-41

support (Beck et al., 1992; W. Zhu et al., 2013). The results shows a slight

decrease in intensity in this peak of calcined and reduced catalysts compared with

the previously reported raw Si-MCM-41 (Gokulakrishnan, Pandurangan,

Somanathan, & Sinha, 2010). This result further indicates that although slightly

destroyed and deformed from the Ru complex, the pore shape within the

Si-MCM-41 structure is largely retained after calcination and H2 reduction

(Deshmukh, Kinage, Kumar, & Meijboom, 2010). The broad peak shown at

around 2θ of 23° corresponds to the amorphous Si-MCM-41 structure

(Dündar-Tekkaya & Yürüm, 2015). Furthermore, the peak of Ru in the reduced

catalyst at 2θ of approximately 43° replaces the peak for Ru2O3 in the calcined

catalyst. This new peak indicates the appearance of a complex after the H2

reduction process. Since the Ru-Si complex is a fairly stable complex that could

be formed under high temperature conditions, it is likely to be formed during the

reduction process (Gasser, Ruiz, Kolawa, & Nicolet, 1999; Y. Wang, Zhang, Cao,

Liu, & Shao, 2011). The peak corresponding to Ru appears after the reduction.

Interestingly, the appearance of the band at 2θ of 43° indicates the formation of

RuSi complex after the H2 reduction process (Perring, Bussy, Gachon, &

Feschotte, 1999).

Page 78: Sustainable atmospheric ammonia synthesis and nitrogen

67

4.3.2 Catalytic synthesis results

The improvement on the ammonia synthesis efficiency is shown in Figure 16

when different components of the catalytic system are introduced. The results

indicate that using Ru catalyst and promoters can lead to approximately 3 times

increase of the synthesis efficiency to the process. Figure 17 shows the effect of

N2 concentration in the feed gas to the NH3 synthesis efficiency. For Si-MCM-41

catalysts, the greatest synthesis efficiency occurs at around 50 vol% of N2 feed

concentration. This result holds for all the conditions under the applied voltages

from 5 kV to 7 kV. The temperature range of the experiments performed in section

3.2 were between 100 °C to 150 °C, which was far less than the temperature

required for the Haber Bosch process. Previous literature results show that the

temperature of the electrode increases linearly with the discharge voltage and

frequency, and agrees with the range of the temperature measurements in this

study (Jidenko, Bourgeois, & Borra, 2010). Meanwhile, the narrow temperature

range in this study further indicated the dominance of catalytic effects of the

plasma and catalysts used in this study. The optimum feeding concentration

obtained for the Si-MCM-41 catalysts differs from the 33 vol% obtained from

previous chapter, which used MgO supports with the Cs promoters but operated

under frequencies lower than 1.6 kHz. Due to the higher surface area of the

mesoporous structure, the residence time and interaction mechanism of the two

reactants will increase in the catalytic system, which could allow more NH radicals

formed in the nitrogen-hydrogen plasma. The results presented in this section are

Page 79: Sustainable atmospheric ammonia synthesis and nitrogen

68

obtained using the catalyst prepared according to the steps described in section

2.1.

Figure 16 Synthesis efficiency results with different catalyst components

Results from Figure 17 also suggest that the NH3 synthesis efficiency is less

dependent on the N2 feed concentration at greater voltage conditions. The output

concentration and concentration of NH3 increases with the applied voltage, the

synthesis efficiency is lower at higher voltage conditions. The effect of gas flow

rate at the optimum feeding concentration of 50 vol% N2 at different voltage

conditions is shown in Figure 18. The ammonia synthesis efficiency increases with

the flow rate. But the trend becomes less obvious after the flow rate is above 4

L/min for the three voltage conditions. Since the plasma synthesis reaction is a

rapid process, it is reasonable that the higher flow rate leads to greater amount of

ammonia produced, and therefore greater synthesis efficiency. Results shown in

Figures 17 and 18 indicate a decrease pattern of NH3 synthesis efficiency with

voltage, and are further confirmed in Figure 19.

0.30

0.50

0.70

0.90

NTP only SupportwithoutRu and

promoters

Supportwithout

promoters

With Ruand

promoters

g N

H3

/kW

h)

Page 80: Sustainable atmospheric ammonia synthesis and nitrogen

69

Figure 17 Effects of the N2 feed concentration on the ammonia synthesis effeciency under different voltage

conditions

Figure 18 Effects of the inlet gas flow rate on the ammonia synthesis effeciency under different voltage

conditions

Figure 19 shows that although the concentration of NH3 increases with

voltage, the synthesis efficiency decreases. The results further indicate that the

reaction equilibrium shift towards the NH3 synthesis at higher voltage. However,

the excess energy from the dielectric discharge at higher voltage leads to the

lower efficiency. As mentioned in section 2.2, the synthesis efficiency is

determined using the ammonia synthesis rate divided by the power. Therefore,

0.20

0.40

0.60

0.80

1.00

0.2 0.4 0.6 0.8

g N

H3

/kW

h

N2 feed concentration

5 kV

6 kV

7 kV

0

0.2

0.4

0.6

0.8

1

0.5 1.5 2.5 3.5 4.5 5.5 6.5

g N

H3/k

Wh

Gas flow rate (L/min)

5 kV

6 kV

7 kV

Page 81: Sustainable atmospheric ammonia synthesis and nitrogen

70

the compensation between the increase in ammonia production and the decrease

in energy efficiency at higher voltage conditions could be the reason for the

synthesis efficiency plateau observed at intervals 5.5-6 kV and 6.5-7kV in Figure

19. On the other hand, in the interval of 6-6.5 kV, the increased power dissipated

into the synthesis reaction might not be enough to increase the synthesis rate of

ammonia, which causes the plateau in ammonia ppm and a relatively dramatic

decrease in synthesis efficiency. Figure 20 shows the effect of the applied

frequency on the synthesis rate and synthesis efficiency at 50 vol% N2 feed

concentration and 5 kV of the applied voltage. Results show that the synthesis

rate of NH3 is less dependent on the frequency than the synthesis efficiency. The

rate reaches the maximum at around 22000 Hz while the synthesis efficiency

increases more dramatically. The greatest synthesis efficiency achieved was 1.7

g/kWh of ammonia at the frequency of 2.6 kHz. It is expected that the resonance

effect of the dielectric barrier discharge can contribute to the homogeneity of the

discharge, which can further increase synthesis efficiency at higher frequency

conditions (B Eliasson & Gellert, 1990; A Fridman, Chirokov, & Gutsol, 2005).

Page 82: Sustainable atmospheric ammonia synthesis and nitrogen

71

Figure 19 Effects of the applied voltage on the ammonia synthesis efficiency and outlet concentration

Figure 20 Effects of the applied frequency on the ammonia synthesis efficiency and outlet concentration

4.3.3 Synergistic effects between the plasma and catalyst

Although the ammonia synthesis process can overall be described by the

general reaction shown in equation (2.1) in section 2, the detailed reaction

0.00

0.30

0.60

0.90

1.20

0.00

20.00

40.00

60.00

80.00

100.00

5 5.5 6 6.6 7

Eff

icie

ncy

(g/

kW

h)

Rat

e (u

mo

l/m

in)

Voltage (kV)

Rate

Efficiency

0.00

0.40

0.80

1.20

1.60

2.00

0.00

10.00

20.00

30.00

40.00

50.00

60.00

20000 22000 24000 26000

Eff

icie

ncy

(g/

kW

h)

Rat

e (μ

mo

l∙m

in-1

)

Frequency (Hz)

Rate

Efficiency

Page 83: Sustainable atmospheric ammonia synthesis and nitrogen

72

mechanism under the plasma conditions needs further explanation. It has been

understood that the synthesis of ammonia from nitrogen and hydrogen gas

involves the following three mechanisms: decomposition, structural

rearrangement, and fragment elimination (B. Penetrante et al., 1997; Van Durme

et al., 2008). For the plasma synthesis of ammonia, the first decomposition step

includes the dissociation and ionization of the inlet nitrogen and hydrogen gases

(Mindong Bai et al., 2003; Matsumoto, 1981). Furthermore, the formation of radical

species generated by the elemental impacts in the plasma region is the essential

part of the reaction (Mizushima et al., 2004).

Within the reactive plasma species shown above, it is believed that NH radical

is the precursor of NH3 formation and its concentration has a strong effect on the

synthesis rate of ammonia (Uyama & Matsumoto, 1989). It is believed that in the

last step of the mechanism, the NH radicals formed in the nitrogen-hydrogen

plasma and the H radicals would be combined with the hydrogen molecules and

produce ammonia (Chauvet et al., 2014; H. Ma et al., 2002).

The high applied frequency combined with a metallic cation promoter could

enhance the electron transfer within the catalyst systems to aid the dissociation

and ionization of nitrogen, which requires highest ionization potential among the

reactions (Mindong Bai et al., 2003; Larichev, 2010). Some most recent

literature have reported that the Cs, K, and Ba are the exclusive promoters that

can increase the catalytic activity of the catalytic system, effectively lower the N2

dissociation barrier, and facilitate the destabilization of the NHx species in

equations (2.4) to (2.6) (Aika, 2017; Z. Ma et al., 2017). Therefore, the synthesis

Page 84: Sustainable atmospheric ammonia synthesis and nitrogen

73

efficiency in this shows improvement compared with previous reported results,

which did not exceed 1.3 g/kWh (Gómez‐Ramírez et al., 2016). However, the

efficiency still requires significant enhancement for this process to be comparable

with the high-temperature and high-pressure synthesis.

Figure 21 The proposed mechanism of NTP ammonia synthesis using Ru-promoter/Si-MCM-41 catalyst

To explain the lack of conversion rate and synthesis efficiency, discussions on the

synergistic effect and interdependence between the plasma and catalyst are

required. Figure 21 shows the proposed mechanism of the NTP ammonia

synthesis using Ru-promoter/Si-MCM-41 catalyst. With the use of multifunctional

catalyst, the transformation of the multi-step reactions could be realized within this

one catalytic system (Dixit et al., 2013). In this process, the plasma formed from

the dielectric discharges result in the chemical and electronic changes in the

catalyst to aid the formation of ammonia (Mizushima, Matsumoto, Ohkita, &

Kakuta, 2007). Electron transfer occurs between the Ru catalyst and the

promoters once the plasma ionizes the catalytic system. This process provides

N2H2

Plasma

NH*

HH

H2

e-PromoterNH2*

H*

NH3

NH3

NH*

Ru Cs

Cs

Si-MCM-41 support

Page 85: Sustainable atmospheric ammonia synthesis and nitrogen

74

dissociation energy to the inlet gases and forms the precursor for a series of

reactions described above. Therefore, ammonia is predominantly formed due to

the increase in concentration of the NH* precursor (Neyts et al., 2015). Although

the large surface area is introduced using Si-MCM-41 as a catalyst support,

catalyst-shielding effects remain in this process due to the surface-treatment

nature of plasma (Amama, Ogebule, Maschmann, Sands, & Fisher, 2006; Bao et

al., 2014). Experimental results indicate that the high frequency of dielectric

barrier discharge promotes the synergistic effects between the plasma and

catalysts. Furthermore, due to the high porosity of the Si-MCM-41 structure

comparing with metal oxide catalysts, less H2 is required in this study compared to

MgO (Mindong Bai et al., 2003). Results have reported that the surface area of Ru

deposited Si-MCM-41 are in the range of 600 to 1000 m2/g, while the Ru

deposited MgO catalyst are approximately 100 m2/g. The high porosity of the

MCM41 supports allows more frequent attachment of the hydrogen radical on the

catalyst surface, therefore the faster mechanism shown in Figure 21 is more likely

to happen. While the shielding effects could be one of the limitations for ammonia

synthesis efficiency, the other limitations are the dissociation of product and

reverse reaction under plasma conditions. Studies have reported the likelihood of

ammonia dissociation when subjected to dielectric plasma discharge. Therefore,

future improvements of the plasma catalytic ammonia synthesis approach could

be made targeting to solve these two specific problems.

Page 86: Sustainable atmospheric ammonia synthesis and nitrogen

75

4.4 Conclusions

In this study, the non-thermal synthesis of ammonia using the Ru-based

multifunctional catalyst was studied, along with the effects of different

experimental parameters of the non-thermal plasma. It is found that the high

voltage inhibits while the high frequency favors the synthesis of ammonia. The

most efficient synthesis achieved was 1.7 g/kWh, which occurred under the

condition of 5000V and 26,000 Hz. The high surface area, low conductivity, the

structured and interconnected pores of the Si-MCM-41 structure enhanced the

efficiency of ammonia synthesis comparing with metal oxide supports. Lastly, a

new concept that includes a two-step method, separating dissociation and

synthesis parts of the process designs is proposed for future investigations. Future

studies could be made on scale-up simulations and to explore other catalyst and

plasma generation options, which could potentially lower the capital cost and

promote the commercialization of this technology.

Page 87: Sustainable atmospheric ammonia synthesis and nitrogen

76

CHAPTER 5 Absorption-enhanced NTP ammonia

synthesis with the assistance of magnesium

chloride

5.1 Introduction

It is mentioned in section 4.3 that one of the most critical challenges of using

NTP to synthesize ammonia was the dissociation of the product within the plasma

region. Therefore, the work in the chapter aims to separate the produced

ammonia from the NTP process immediately after it is synthesized. Various

sorbents have been investigated for the separation of ammonia from gas streams,

such as MgCl2, CaCl2, and activated carbon, etc (Huang, Li, & Chen, 2008;

Sharonov & Aristov, 2005b; Sharonov, Veselovskaya, & Aristov, 2006; H. Zhu, Gu,

Yao, Gao, & Chen, 2009).

In several recent studies, magnesium chloride has been used as an efficient

absorbent that could lower the operating pressure and enhance the performance

of the Haber-Bosch process (Malmali, Wei, McCormick, & Cussler, 2016). The

absorption of ammonia onto MgCl2 usually relies on the formation of Mg(NH3)6Cl2,

which has a strong ammonia-holding capacity (H. Zhu et al., 2009). Other

advantages of MgCl2 are its compatible cost and easy regeneration.

Furthermore, a recent study showed that magnesium-based catalysts can

Page 88: Sustainable atmospheric ammonia synthesis and nitrogen

77

promote the ammonia synthesis under plasma conditions by forming the

intermediate compound of Magnesium nitride (Mg3N2) (Zen, Abe, & Teramoto,

2018). The authors propose Mg3N2 to be an ideal solid-state ammonia carrier

under atmospheric conditions. This study also shows that the Mg3N2 intermediate

could be used to form additional NH3 molecules by reacting with water after the

plasma treatment under nitrogen environments. Compared with MgO, MgCl2 is

easier to react with N2 to form Mg2N3 due to its lower lattice energy. Since the

lattice energy of MgCl2 is significantly lower than MgO, it is reasonable that the

formation of Mg2N3 from MgCl2 much easier than MgO.Due to all the advantages

mentioned above, MgCl2 is identified as the absorbent used for the study in this

chapter.

Page 89: Sustainable atmospheric ammonia synthesis and nitrogen

78

5.2 Materials and methods

5.2.1 System improvements

The experimental apparatus, setup, and procedures were the same as the

previous chapter with the addition of a pulse density modulation (PDM) added to

the SSD 110 inverter. Introducing the PDM to the inverter is critical to the system

improvement in various ways. Recent literature showed that adding the PDM

control of high-frequency inverters could lead to high efficiency (by greatly

reducing the switching loss), small size, and the ability of self power factor

correction (Sandali & Chériti, 2017). In other words, the PDM not only allowed a

linear control of plasma according to the command signals, but also capable of

reducing the power delivered to output of the inverter, which was the load, while

maintaining the high frequency and voltage. The PDM was installed to the system

by connecting it to the first, third, and sixth nodes on the invertors control panel.

The nodes, along the horizontal position of the PDM was adjusted until a stable

waveform on the oscilloscope was formed before starting the system.

5.2.2 Absorbent preparation and characterization

In this chapter, MgCl2 acted as one of the reactants, and also the absorbent

for the process. Anhydrous MgCl2 powder was purchased from Sigma-Aldrich

(USA). The absorbents were packed in the “catalyst region” showed in Figure 5.

Page 90: Sustainable atmospheric ammonia synthesis and nitrogen

79

The system was dried in nitrogen environment under 130 ℃ before starting each

run. The unreacted and reacted MgCl2 samples were characterized using X-ray

diffraction (XRD). Data analysis was performed with the assistance of the

Materials Data Incorporated's JADE 8.0 software package with whole pattern

fitting.

Page 91: Sustainable atmospheric ammonia synthesis and nitrogen

80

5.3 Results and discussions

Figure 22 showed a schematic diagram of how the MgCl2 was placed at

different positions relative to the plasma discharge region. The corresponding

ammonia synthesis rate and energy efficiency of the different placements were

reported in Figure 23. The results showed that the ammonia synthesis rate and

synthesis efficiency both increased when MgCl2 was placed in the plasma region

and was the highest among the different placements. This showed that the

formation of nitrogen-based intermediate could be a critical factor to the

enhancement caused by MgCl2. The results also showed around 6% of ammonia

was absorbed by MgCl2 and was not dissociated in the plasma region. For the

other two circumstances, when MgCl2 was placed immediately below the plasma

discharge region or 5 cm below the discharge region, the total ammonia synthesis

rates were increased compared with synthesis without MgCl2 as well. However,

with a closer look to the rate composition, the enhancement of these two

displacements were caused mainly by the absorption, as the synthesis rates

(shown in red) were approximately the same as the plasma only synthesis. Due to

the short residence time of the reactant and product gases in the reactor, part of

the synthesized ammonia was absorbed by MgCl2 immediately after passing the

plasma discharge region, before forming back to N2 and H2.

Page 92: Sustainable atmospheric ammonia synthesis and nitrogen

81

Figure 22 Schematic diagrams of the MgCl2 locations relative to the plasma discharge region

Figure 23 Ammonia synthesis results with MgCl2 packed at different locations relative to the plasma

discharge region

The dielectric effect was what contributed to the catalytic functions of the

0

2

4

6

8

10

0.00

10.00

20.00

30.00

40.00

50.00

Plasmaonly

Indischarge

Belowdischarge

5 cm belowdischarge

Eff

icie

ncy

(g/

kW

h)

Rat

e (μ

mo

l∙m

in-1

)

Total rate

Rate absorbed

Efficiency

Page 93: Sustainable atmospheric ammonia synthesis and nitrogen

82

MgCl2. Within the plasma discharge region, the magnesium is capable of bonding

at its surface with the free radicals (N*, H*) formed by the electron excitation and

vibrational excitation in the gas phase, providing additional reaction sites with the

NHx* free radicals to form ammonia 8, 14, 27. To further study how the change in

discharge characteristics due to dielectric effect may contribute to improvement of

efficiency, Figures 24 to 27 showed the effects of different plasma parameters to

the NH3 synthesis rate. absorption rate, and energy efficiency. From the results,

the effects of the different parameters could be divided into two categories. For

the first category, the changes in the parameters, which were the discharge

voltage and N2 composition, affected both the synthesis rate and the efficiency.

Figure 24 showed both the maximum NH3 synthesis rate and the energy

efficiency occurred at 6.44 kV. For the synthesis rate, the absorbed rate and the

total rate shared the same relation to the discharge voltage. Both rates increased

with voltage until 6.44 kV and decreased between 6.44 kV to 7.14 kV. This trend

agreed with the findings in the previous chapters that the increase in the

discharge voltage could promote the forward reaction to form NH3, until the

dissociation affect started to dominate. This “double-sided sword” effect was more

obvious in the study using MgCl2, as the absorbed rate started to decrease after

the voltage increased beyond 6.44 kV. Similarly for the N2 composition, at lower

N2 contents, NH3 was synthesized less efficiently and at lower rates. As the N2

composition started to increase within the reactant stream, both the rate and the

efficiency started to rise. However, the performance was less ideal during the N2

content of 0.75. The most efficient synthesis was achieved at the N2 composition

Page 94: Sustainable atmospheric ammonia synthesis and nitrogen

83

of 0.5, which represented a volumetric flow rate ratio of 1:1 for N2 and H2. On the

other hand, for parameters including the discharge frequency and flow rate, the

affects of them to the synthesis rate was smaller than to the efficiency. The

influence of these two parameters on the discharge current from the invertor to

the transformer was obvious. When the discharge frequency was adjusted to 23

kHz, the current was significantly smaller than other conditions. For the gas flow

rates, the high gas flow rates contributed to the increase in energy efficiency and

the maximum was reached at 4 L/min, although the influence on the synthesis

rate was not as obvious. However, the rates for ammonia absorption by the MgCl2

decreased at higher flow rates due to the decrease in the residence time. As

discussed in Chapter 4, due to the fast and rapid nature of plasma-related

reactions, higher flow rates could lead to more products being formed in the region,

which could further cause greater synthesis efficiency.

Page 95: Sustainable atmospheric ammonia synthesis and nitrogen

84

Figure 24 Ammonia synthesis results with MgCl2 under different voltage conditions

Figure 25 Ammonia synthesis results with MgCl2 under different frequency conditions

0

2

4

6

8

10

0

10

20

30

40

50

5.04 5.74 6.44 7.14

Eff

icie

ncy (

g/k

Wh

)

Ra

te (

μm

ol∙m

in-1

)

Voltage (kV)

Total rate

Rate absorbed

Efficiency

0

4

8

12

16

20

0

10

20

30

40

50

60

20 21.5 23 24.5 26

Eff

icie

ncy (

g/k

Wh)

Ra

te (

μm

ol∙m

in-1

)

Frequency (kHz)

Page 96: Sustainable atmospheric ammonia synthesis and nitrogen

85

Figure 26 Ammonia synthesis results with MgCl2 under different nitrogen compositions

Figure 27 Ammonia synthesis results with MgCl2 under different gas flow rates

0

5

10

15

20

0

10

20

30

40

50

60

70

0.25 0.375 0.5 0.625 0.75

Eff

icie

ncy (

g/k

Wh

)

Ra

te (

μm

ol∙m

in-1

)

N2 composition

Total rate

Rate absorbed

Efficiency

0

5

10

15

20

0

10

20

30

40

50

60

70

1 2 3 4 5

Eff

icie

ncy (

g/k

Wh

)

Ra

te (

μm

ol∙m

in-1

)

Reactant flow rate (L/min)

Total rate

Rate absorbed

Efficiency

Page 97: Sustainable atmospheric ammonia synthesis and nitrogen

86

The XRD results of the untreated MgCl2 and the treated MgCl2 in and below

the discharge region were shown in Figure 28. Based on the proposed

mechanisms of this process, three compounds were identified from the XRD

patterns, which were MgCl2, Mg3N2, and Mg(NH3)6Cl2. The peaks associated with

each compound were labeled, along with their intensities. To improve the clarity

and readability of the figure, the peaks associated with MgCl2 were labeled on the

untreated pattern and the peaks for Mg3N2 and Mg(NH3)6Cl2 were labeled on the

in discharge and below discharge patterns, respectively. The XRD pattern for the

untreated sample matched closely with the ones for MgCl2 in the Jade software

database, and with the findings reported in previous literature (Stavrou et al.,

2016). The diffraction peaks at 21,32, and 34 2-degree theta were associated to

Mg3N2 (Luo, Kang, Fang, & Wang, 2011). And the peaks at 16 and 22 2-degree

theta corresponded to Mg(NH3)6Cl2 (Guangming, Peihua, Zhiming, Mingzhen, &

Minxong, 2004; Y. Liu et al., 2011). The results from the figure showed that both

Mg3N2 and Mg(NH3)6Cl2 were identified in the in-plasma and below-plasma

treated samples. Intensity wise, the intensities of the MgCl2 associated peaks

decreased after being treated by the plasma, and further decreased when placed

below the plasma discharge region. This result corresponds with the findings from

Figure 28. When MgCl2 was placed below the plasma discharge region, the ratio

of the absorbed ammonia was higher than when it was placed in the plasma

region. Therefore, it is reasonable that the concentration of pure MgCl2 would

decrease compared with the untreated and in-discharge samples.

Page 98: Sustainable atmospheric ammonia synthesis and nitrogen

87

The SEM images from Figures 29 to 31 indicated that the effects of the

plasma treatment on the shape and size of the MgCl2 were negligible. However,

the EDS analysis results demonstrated difference between the MgCl2 samples

treated under different conditions. For the MgCl2 samples treated inside and

below the plasma discharge regions, nitrogen was identified during the EDS

analysis. The nitrogen concentration in the treated sample reported by the EDS

analysis was 0.16 wt%, which showed relatively good correspondence with the

theoretical nitrogen concentration based on the amount of ammonia absorbed

during the process.

From the analysis above, two critical reaction mechanisms of how MgCl2

promoted the ammonia synthesis could be developed. The first mechanism

supported the hypothesis from in Section 5.1. Similar to other Mg-based catalysts,

such as MgO, the MgCl2 could form the critical intermediate compound Mg3N2

(magnesium nitride) during the plasma synthesis via the nitridation mechanism

(Zen et al., 2018). Compared with MgO, MgCl2 had smaller lattice energy so that it

was easier to carry out the nitridation mechanism for MgCl2 than MgO. As a result,

the ammonia synthesis rate in this chapter was higher than those achieved for

MgO (Zen et al., 2018). The second mechanism was the rapid absorption of the

produced ammonia by MgCl2 in the plasma region. The fast absorption was

achieved via the formation of Mg(NH3)6Cl2. Past literature repeatedly reported the

strong ability of MgCl2 for absorbing and holding ammonia from this reaction

(Sharonov & Aristov, 2005a; H. Zhu et al., 2009). Compared with the absorption

result from these studies, which was about 33 mg NH3/g MgCl2 at 100 ℃, the

Page 99: Sustainable atmospheric ammonia synthesis and nitrogen

88

amount absorbed by MgCl2 was larger (over 100 mg NH3/g MgCl2) due to the first

absorption mechanism that involved the formation of Mg3N2. This could indicate

within the synergistic absorption effects, the formation of Mg3N2 was favored over

the Mg(NH3)6Cl2. Another possibility was that Mg(NH3)6Cl2 could be dissociated

under the treatment of plasma. While the ability of retaining the Mg3N2 compound

within plasma was reported in previous literature, no research managed to

determine the fate of Mg(NH3)6Cl2 under plasma treatment. Therefore, this could

be a direction for future studies on exploring ammonia holding capabilities of

different absorbents under plasma conditions. Past literature reported the ability

of MgCl2 for absorbing and holding ammonia of forming Mg(NH3)6Cl2 (Sharonov &

Aristov, 2005a; H. Zhu et al., 2009), which was about 33 mg NH3/g MgCl2 at 100

℃. The fact that the amount absorbed in the solid-phase in this study was larger

(over 100 mg NH3/g MgCl2) indicated that within the synergistic absorption effects,

the formation of Mg3N2 was favored over the Mg(NH3)6Cl2.

Page 100: Sustainable atmospheric ammonia synthesis and nitrogen

89

Figure 28 XRD patterns of the MgCl2 samples under different conditions

Page 101: Sustainable atmospheric ammonia synthesis and nitrogen

90

Figure 29 SEM image and EDS analysis results of the untreated MgCl2 sample

Page 102: Sustainable atmospheric ammonia synthesis and nitrogen

91

Figure 30 SEM image and EDS analysis results of the MgCl2 sample treated in the plasma discharge region

Page 103: Sustainable atmospheric ammonia synthesis and nitrogen

92

Figure 31 SEM image and EDS analysis results of the MgCl2 samples treated below the plasma discharge

region

Page 104: Sustainable atmospheric ammonia synthesis and nitrogen

93

5.4 Conclusions

In conclusion, two major improvements were carried out in this chapter. First,

the pulse density modulation (PDM) added to the SSD 110 inverter improved the

efficiency of the invertor-transformer system by reducing the current and switching

energy loss. Second, magnesium chloride used in this chapter could enhance the

synthesis rate and efficiency through forming two critical compounds for the

process, Mg3N2 (reaction with N2 in the plasma) and Mg(NH3)6Cl2 (absorption of

the NH3 product). These improvements led to a greatest energy efficiency of 20.5

g/kwh, with the corresponding synthesis rate being 50 mmol/min.

Page 105: Sustainable atmospheric ammonia synthesis and nitrogen

94

CHAPTER 6 Atmospheric plasma-assisted nitrogen

fixation using water and nitrogen jet plasma

6.1 Introduction

The fixation of nitrogen gas, one of the most abundant resources in the world,

has been historically affecting the human society for over centuries.

Nitrogen-derived chemicals such as nitric acid and ammonium nitrate has been

used largely in the field of agriculture, paints and dyes, and explosives. In recent

years, ammonia has been intensively investigated as a green fuel and sustainable

hydrogen storage material. However, the conventional industrial nitrogen fixation

process that the Haber-Bosch method to generate ammonia, which is further used

to synthesize other nitrogen-based chemicals, is an energy-intensive process that

consumes a large amount of fossil fuels. This process itself is responsible for

approximately 1 to 2% of the global electrical usage and emits significant

greenhouse gases (Galloway et al., 2008; Razon, 2014). Furthermore, the high

temperature (400 to 600 C) and pressure requirements (400 to 600 atm) of this

process make the synthesis of ammonia centralized and prevent local farms and

small industries to perform on-site ammonia generation (Hargreaves, 2014;

Tallaksen, Bauer, Hulteberg, Reese, & Ahlgren, 2015).

Plasma-based nitrogen fixation provides an alternative nitrogen fixation

process that is sustainable, accessible to distributed production, and most

Page 106: Sustainable atmospheric ammonia synthesis and nitrogen

95

importantly under atmospheric temperature and pressure. The NOx – compounds

formed during the fixation process using nitrogen plasma and water have been

reported responsible for bacterial inactivation (Shen et al., 2016), and can

promote the increase plants’ growth rates(Park et al., 2013). There have been

several attempts on the synthesis of NOx – compounds using plasma-assisted

processes. In these studies, different plasma discharge methods have been used

to fix nitrogen gas into water and produced nitric and nitrous acids, including

gliding arc discharge plasma (Bian, Shi, & Yin, 2009; Burlica, Kirkpatrick, & Locke,

2006; Park et al., 2013; Yang, Li, Zhong, Guan, & Hu, 2016), nozzle-type plasma

jet (Shen et al., 2016), and dielectric barrier discharge plasma (Kovačević et al.,

2017). While the emphasis of these studies has been focusing on the production

of NOx – in the liquid phase, the generation of NOx gas during the process is rarely

considered and the ammonium content in the liquid phase has not been reported.

On the other hand, a majority of the studies on the plasma-assisted synthesis

of ammonia and NOx gas under atmospheric temperature and pressure have

been using catalytic plasma discharge of nitrogen and other related gases, such

as hydrogen (Gómez‐Ramírez, Montoro‐Damas, Cotrino, Lambert, & Gonzá

lez‐Elipe, 2017) and oxygen (B. Patil, Cherkasov, et al., 2016; B. S. Patil et al.,

2017). The process that uses nitrogen plasma and water circumvents the use of

hydrogen gas, which can reduce the capital cost and safety issues of the process.

However, this is a novel process that has been sparsely investigated. The idea

was first introduced by Kubota et al., which used a nozzle-type plasma jet for the

Page 107: Sustainable atmospheric ammonia synthesis and nitrogen

96

production of ammonia in water (Kubota, Koga, Ohno, & Hara, 2010). Later, two

most recent studies reported the synthesis of ammonia by purging the nitrogen

plasma gas into water under the irradiation of UV light (Haruyama et al., 2016;

Sakakura et al., 2017a). The authors pointed out that the UV irradiation could

improve the synthesis of ammonia by promoting the surface dissociation of water

molecules as hydrogen donors in the reaction. The findings from the literature

also indicated that the synthesis of ammonia is a reaction that primarily happens

at the liquid surface and is highly dependent on the reaction area (Haruyama et al.,

2016). Therefore, increasing the surface area and reducing the distance between

the plasma generation and the water surface are critical to improve the

performance of the process, Furthermore, since previous literature focus on the

separate synthesis of ammonium, nitrite, and nitrate, it would be ideal to have a

process that aims to co-synthesis all three compounds during the fixation.

Page 108: Sustainable atmospheric ammonia synthesis and nitrogen

97

6.2 Materials and Methods

6.2.1 Improvements of the system

In this process, two significant improvements are introduced to the

process to enhance the performance of the process. First, a spray-type plasma jet

is used, which greatly increases the reaction area at the plasma-water interface.

Second, this study uses an innovative in-situ synthesis approach to facilitate the

plasma-liquid interface reactions. This approach largely shortens the distance

between the discharge region and the liquid interface, therefore significantly

reduced the back reaction of the nitrogen plasma species. Furthermore,

understanding towards all of the nitrogen-based products formed during the

process is critical for improving and optimization the plasma-based nitrogen

fixation technology. Thus, this paper first takes into consideration of both the

ammonium and NOx – generated in the liquid phase, as well as the NOx produced

in the gas phase. In this paper, a comprehensive reaction mechanism of forming

the different nitrogen fixation products is proposed, including the synergistic

integrations between the gas phase and liquid surface products.

6.2.2 Description of apparatus

The in-situ synthesis of ammonia was carried out in the 10-liter constant

stirring tank reactor shown in Figure 32 filled with 2 L of distilled water. The

plasma discharge head was installed inside the reactor. The distance between the

Page 109: Sustainable atmospheric ammonia synthesis and nitrogen

98

discharge head and the liquid surface was 3 to 4 cm. The plasma was formed and

dispersed onto the liquid surface by an atmospheric corona discharge plasma jet

generator (Enercon Industries, USA). The discharge head consisted of two

electrodes separated by a distance of 3 cm. Nitrogen gas was supplied to the

system at a flow rate of 57 standard liter per minute (SLPM). The high flow rate of

the gas allowed the plasma to spread out through the nozzle and disperse onto

the liquid surface. A well-insulated electric cord connected the high voltage

electrode to a step-up transformer located in the main console outside of the

reactor, which amplified the 120 V, 60 Hz utilities alternate current (AC) power

supply to 4.4 kV. The UV irradiation was produced using a 15 W UV lamp supplied

by Technical Precision Inc. (USA). Ammonia in the gas outlet was detected using

a colorimetric gas detection system from RAE Systems, Inc. (USA). The NOx

concentration was determined by an IMR gas analyzer (IMR Inc. USA) and the

concentration of N2 was verified using a Varian CP-4900 gas chromatography

(Agilent, USA). The liquid samples collected from the reactor were analyzed using

a DR 5000™ UV-Vis Spectrophotometer (Hach Technology Inc. USA).

Page 110: Sustainable atmospheric ammonia synthesis and nitrogen

99

Figure 32 Process flow diagram of the in-situ atmospheric nitrogen fixation process

Page 111: Sustainable atmospheric ammonia synthesis and nitrogen

100

6.3 Results and discussion

6.3.1 Synthesis results

During the experiments, the concentration of the products in the liquid phase

built up as time increased, as shown in Figures 33 and 34. On the other hand, the

NOx concentration built up instantly to the steady state value shown in Figures 35

d) and 36 d), and remained throughout each run. In general, the concentrations

shown in Figures 33 and 34 increased at approximately constant rates with time at

various experimental conditions. Due to the short life spam of the active plasma

species, the distance between the plasma discharge and the liquid surface is an

important factor in this experiment. The in-situ setup led to twice as much as

products than the ex-situ setup, where the plasma discharge was placed outside of

the reactor. From Figure 32, it was obvious that the in-situ operation, and the UV

irradiation improved the nitrogen fixation significantly, as well as the final

concentrations of nitrate, nitrite and ammonium at the end of the 20-minute

reaction. Furthermore, based on the results in Figure 35, the total nitrogen fixation

rate increased from 22.3 μmol min-1 (ex situ) to 51.1 μmol min-1 (in situ), which

also showed improvements compared with other reported ex-situ synthesis with a

similar reaction area (Haruyama et al., 2016; Sakakura et al., 2017b). For each

nitrogen fixation product, the increase in both final concentration and synthesis

rate was also dramatic (greater than 2 times increase). In general, one of the main

challenges for the plasma-assisted ammonia synthesis was the tendency for the

Page 112: Sustainable atmospheric ammonia synthesis and nitrogen

101

dissociated nitrogen plasma species to reform the diatomic nitrogen. The in-situ

plasma generation (shown in Figure 32) greatly shortened the distance between

the plasma discharge point and the liquid surface. Therefore, a region with

concentrated nitrogen plasma species was formed, which enabled the dissociated

nitrogen to react directly with the hydroxyl or hydrogen radicals to form NOx or

ammonia. This increased the possibility for the formation of ammonia at the liquid

gas interface. Furthermore, the spray-type plasma jet used in this study increased

the surface area of the reaction. The reaction surface area of the spray-type

plasma jet was approximately 5 cm2, which was significantly larger compared with

a nozzle type plasma jet of similar power (usually less than 1 cm2). Second, since

high flow rate is used, the plasma species.

Page 113: Sustainable atmospheric ammonia synthesis and nitrogen

102

Figure 33 Concentrations of nitrate, nitrite and ammonium at different experimental conditions at 30 C

0

40

80

120

160

0 5 10 15 20

μm

ol∙

L-1

Time (min)

a) NO3-

Ex-situ w/ UV

In-situ w/ UV

In-situ w/o UV

0

5

10

15

20

25

0 5 10 15 20

μm

ol∙

L-1

Time (min)

b) NH4+

Ex-situ w/ UV

In-situ w/ UV

In-situ w/o UV

0

100

200

300

400

0 5 10 15 20

μm

ol∙

L-1

Time (min)

c) NO2-

Ex-situ w/ UV

In-situ w/ UV

In-situ w/o UV

Page 114: Sustainable atmospheric ammonia synthesis and nitrogen

103

6.3.2 Gas phase reactions

The co-synthesis of the nitrate, nitrite, ammonium, and NOx demonstrated in

this study was the result from both the gas phase and liquid phase reactions, as

well as the interactions between these two phases. Figure 37 described the

proposed synergistic mechanism involved in this study. In the gas phase, the

electron excitation of the nitrogen gas led to the formation of free nitrogen radicals

in the gas phase (equation 6.1) (Carrasco et al., 2011). At the mean time, the

hydroxyl radicals were formed by the electron excitation of the vaporized water via

the mechanism shown in equation (6.2) (Itikawa & Mason, 2005; D.-X. Liu,

Bruggeman, Iza, Rong, & Kong, 2010). The excited free nitrogen and hydrogen

radicals formed the NH radical, which acted as the main intermediate to form

ammonia. As pointed out by Sakakura et al. (Sakakura et al., 2017a), there should

also be a negligible amount of hydrogen gas formed by the hydrogen radicals in

the gas phase (equation 6.3), which would immediately react with the hydroxyl

radicals in the liquid phase and could promote the formation of NH4+ in the solution.

However, without the presence of catalysts, this series of reactions in the gas

phase (equations (6.4) to (6.6)) was minor. And due to the close distance between

the discharge head and the liquid surface, the NH3 particles produced in the gas

phase were immediately absorbed in the liquid. Therefore, the ammonia

concentration in the outlet gas was below 1-ppm detection limit. Under the effect

of UV irradiation, the reaction of the nitrogen and hydroxyl radicals represented by

Page 115: Sustainable atmospheric ammonia synthesis and nitrogen

104

equations (6.7) and (6.8) would occur, leading to the production of the NOx

species in the gas phase (Sharma, Hosoi, Mori, & Kuroda, 2013).

N2(𝑣) + e → 2N(𝑣) + e (6.1)

H2O (𝑣 ) + e → OH (𝑣) + H(𝑣) + e (6.2)

2H(𝑣) → H2(𝑣) (6.3)

N(𝑣) + H(𝑣) → NH(𝑣) (6.4)

NH(𝑣) + H(𝑣) → NH2(𝑣) (6.5)

NH2(𝑣) + H(𝑣) → NH3(𝑣) (6.6)

N(𝑣) + OH(𝑣) → NO(𝑣) + H(𝑣) (6.7)

NO(𝑣) + OH(𝑣) → NO2(𝑣) + H (𝑣) (6.8)

6.3.3 Liquid phase reactions

In the liquid phase, a two-step mechanism was demonstrated in previous

literature for the ammonia synthesis (Sakakura et al., 2017b). At the liquid-gas

interface, the free radicals formed by the electron-excited nitrogen gas, and its

related radicals, reacted with the water molecules excited by UV light (equation 6.9)

to form ammonia at the surface of the liquid, which was then dissolved into the

solution. On the other hand, the main nitrogen-based products in the aqueous

Page 116: Sustainable atmospheric ammonia synthesis and nitrogen

105

phase formed during the experiments were the NO3 −

and NO2 −, with NO2

– being

the predominant compound. In the liquid phase, the plasma species in the gas

phase formed by the dissociated nitrogen gas have the tendency to form more

stable compounds in the liquid, which were the nitrate and nitrite products

according to reactions mechanisms and demonstrated in an earlier publications

(Szili, Hong, Oh, Gaur, & Short, 2017), (R Atkinson et al., 1989).

H2O (𝑙 )UV irradiation→ OH (𝑙) + H(𝑙) (6.9)

H2(𝑣) + OH(𝑙) → H2O + H (𝑙) (6.10)

N(𝑣) + H(𝑙) → NH(𝑙) (6.11)

NH(𝑣 𝑜𝑟 𝑙) + H(𝑙) → NH2(𝑙) (6.12)

NH2(𝑣 𝑜𝑟 𝑙) + H(𝑙) → NH3(𝑙) (6.13)

NO(𝑣) + OH(𝑙) → HNO2(𝑙) (6.14)

NO2(𝑣) + OH(𝑙) → HNO3(𝑙) (6.15)

Page 117: Sustainable atmospheric ammonia synthesis and nitrogen

106

Figure 34 Concentrations of nitrate, nitrite and ammonium of the in-situ reactor at different temperatures

Literature documented the formation of the hydroxyl and hydrogen radicals

0

40

80

120

160

0 5 10 15 20

μm

ol∙

L-1

Time (min)

a) NO3-

40C

30C

50C

70C

60C

0

10

20

30

40

0 5 10 15 20

μm

ol∙

L-1

Time (min)

b) NH4+

40C

30C

70C

50C

60C

0

100

200

300

400

0 5 10 15 20

μm

ol∙

L-1

Time (min)

c) NO2-

40C

30C

50C

70C

60C

Page 118: Sustainable atmospheric ammonia synthesis and nitrogen

107

from water with the assistance of UV light, which was represented by equation

(6.9) (Chai, Zheng, Zhao, & Pollack, 2008). This reaction could further react with

the plasma-excited nitrogen radicals at the liquid-gas interface, in either the gas or

liquid phase, to promote the formation of nitrite acid, ammonium, and nitrate

(equations 6.10 to 6.15). The promotional effect of UV irradiation towards the

liquid surface dissociation plays a critical role in nitrogen fixation. Although the

promotion effect of the UV light on plasma-assisted ammonia synthesis was

reported (Haruyama et al., 2016; Matsuo et al., 2015), results from this study

showed that the promotion of UV irradiation could promote the formation of other

nitrogen species in the liquid as well. More specifically, the promotional effect of

UV irradiation was selectively and more effective towards the NO2 – and NH4

+,

respectively. As shown in Figure 35, introducing the UV light into the system led to

approximately 3 and 2 times increase in the synthesis rate of NO2 – and NH4

+,

whereas the rate for NO2 – increased only around 1.3 times. To explain this, it was

noteworthy to point out that the rate constant of reaction (6.7) is the orders of 10-11

cm6 ∙ molecule−1s−1, while the rate constant for reaction (6.8) is in the order of 10-32

cm6 ∙ molecule−1s−1, which explained the predominant improvement of the UV

irradiation on the synthesis rate of NO2 –

in the aqueous phase (R. Atkinson et al.,

2004). On the other hand, UV irradiation had minor effect on the production of

NOx.

Page 119: Sustainable atmospheric ammonia synthesis and nitrogen

108

Figure 35 Synthesis rates of nitrate, nitrite and ammonium at different experimental conditions at 30℃

0

2

4

6

8

10

12

14

16

18

Ex-situ w/

UV

In-situ w/o

UV

In-situ w/

UV

Ra

te (

μm

ol∙

min

-1)

a) NO3−

0

0.5

1

1.5

2

2.5

Ex-situ w/ UV In-situ w/o UV In-situ w/ UV

Ra

te (

μm

ol∙

min

-1)

b) NH4+

0

5

10

15

20

25

30

35

40

Ex-situ w/

UV

In-situ w/o

UV

In-situ w/

UV

Ra

te (

μm

ol∙

min

-1)

c) NO2−

0

20

40

60

Ex-situ w/

UV 30C

In-situ w/o

UV

In-situ w/

UV

pp

m

d) NOx

Page 120: Sustainable atmospheric ammonia synthesis and nitrogen

109

Figure 36 Synthesis rates at nitrate, nitrite and ammonium of the in-situ reactor at different temperatures

Overall, the temperature had minor effects on the production of the

ammonium and the other nitrogen-based products. The influences of temperature

on the synthesis of nitrate and nitrate were more dramatic than ammonium.

Previous literature showed that the hydroxyl radicals formed by the dissociation of

water under UV irradiation was the highest at around 40 °C (Matsuo et al., 2015),

which corresponded with the finding from this study that the highest synthesis

rates for NO3 – (15.1 μmol∙min-1) and NO2

– (40.3 μmol∙min-1) was detected at

0

4

8

12

16

20

20 40 60 80

Rate

mol∙

min

-1)

Temperature (℃)

a) NO3−

0

0.6

1.2

1.8

2.4

3

20 40 60 80

Ra

te (

μm

ol∙

min

-1)

Temperature (℃)

b) NH4+

0

15

30

45

60

20 40 60 80

Ra

te (

μm

ol∙

min

-1)

Temperature (℃)

c) NO2−

0

30

60

90

120

150

20 40 60 80

pp

m

Temperature (℃)

d) NOx

Page 121: Sustainable atmospheric ammonia synthesis and nitrogen

110

around 40 °C, shown in Figure 36. On the other hand, the NOx concentration in

the gas phase increased with temperature. As mentioned previously, the NOx in

the outlet gas was formed by reactions between the vaporized H2O and the

nitrogen plasma. At higher temperatures, the vapor content of water in the gas

phase increased dramatically, therefore favored the formation of NOx gases

represented by equations (6.7) and (6.8). As the byproducts of this process, the

exhaust NOx could be absorbed to form nitrate and nitrite solution in the

downstream steps. Based on the experiments from this study, the optimum

operation temperature was determined to be around 40 °C, where the nitrogen

fixation rate was the highest and the formation of NOx byproducts was low.

Page 122: Sustainable atmospheric ammonia synthesis and nitrogen

111

Figure 37 The propose mechanism for the In-situ plasma jet nitrogen fixation

Page 123: Sustainable atmospheric ammonia synthesis and nitrogen

112

6.4 Conclusions

In conclusion, the co-synthesis of nitrate, nitrite, and ammonium during a

plasma-assisted nitrogen fixation process from water and nitrogen was

investigated. The in-situ spray-type plasma used in this study could significantly

enhance the nitrogen fixation rate. The UV irradiation also promoted the fixation

via the excitation of vapor and liquid water molecules. The fixation mechanism

determined in this study was a synergistic interaction between the gas and liquid

phase. The highest nitrate, nitrite, and ammonium synthesis rate achieved in this

study are 15.1 μmol∙min-1, 40.3 μmol∙min-1, and 2.5 μmol∙min-1, respectively.

Page 124: Sustainable atmospheric ammonia synthesis and nitrogen

113

CHAPTER 7 Challenges and Opportunities

7.1 Challenges

Based on the relevant studies discussed in the previous chapters, the main

challenges of the NTP synthesis of ammonia are two-fold. The first challenge is

the nitrogen fixation. The energy required to break the nitrogen triple bond is

approximately 945 kJ/mol (9.79 eV), which is the largest activation energy

required in both conventional and NTP ammonia synthesis (Misra, Schlüter, &

Cullen, 2016). The specific energy plays a critical role in facilitating the reaction to

synthesize ammonia. Unfortunately, the effects of increased specific energy input

are a bit of a double-edged sword with regards to ammonia synthesis. Increased

specific energy input leads to an increase in the number of high-energy electrons

and important reaction intermediates (such as NH*), which are reportedly

responsible for both product formation and undesired back reactions (the

decomposition of the product) (Van Helden et al., 2007). This leads to the second

challenge of the NTP ammonia synthesis, which is the immediate dissociation of

ammonia after it is produced. The decomposition effect of NTP on ammonia was

first revealed by studies related to air pollution and off-gas treatment (Oda, 2003;

Urashima & Chang, 2000). (L. Wang et al., 2013) demonstrated that NTP could

decompose ammonia with the assistance of Fe catalysts and avoid nitrogen

poisoning during the process. Furthermore, other studies also indicate that

Page 125: Sustainable atmospheric ammonia synthesis and nitrogen

114

transitional metals including Fe, Co, Cu and Ni demonstrated relatively high

activities during plasma-assisted ammonia decomposition (Ji et al., 2013; L.

Wang et al., 2015). Therefore, to minimize the dissociation of ammonia during its

plasma-assisted synthesis process, it is encouraged that transition metals should

be avoided during the catalyst selection.

Page 126: Sustainable atmospheric ammonia synthesis and nitrogen

115

7.2 Opportunities

This section is a discussion of some of the opportunities to overcome several

of the major difficulties in developing an efficient method NTP ammonia synthesis,

as well as future opportunities in the application of this technology as a

competitive and environmentally friendly alternative to the traditional method of

ammonia synthesis. Since NTP ammonia synthesis is still considered a

cutting-edge research topic, it is very possible that the opportunities for

breakthroughs could be inspired from the most recent and innovative ammonia

synthesis studies under intense conditions. Some studies that synthesize

ammonia under high temperature and pressures have been depositing ruthenium

onto carbon nanotubes, or developed innovative structures to enhance the

catalytic binding efficiencies (Wu et al., 2003). Another creative catalyst structure

that could potentially benefit NTP ammonia synthesis is the use of electrode

support within the catalyst matrix. Several studies, although performed under high

temperature and pressure, used a new catalyst structure that loads ruthenium

onto a new electrode support 12CaO∙7Al2O3 to prepare a unique catalytic

structure of Ru/C12A7:e- (Kanbara et al., 2015; Kitano et al., 2012; Kitano et al.,

2015). The three studies pointed out that the stable electrode material can

perform both as electron donor and as a reversible hydrogen storage site, while

simultaneously enhancing ammonia synthesis by shifting the rate limiting reaction

from nitrogen dissociation to hydrogen bonding. With regards to the previously

discussed reactor development, the rapid absorption of ammonia after its

Page 127: Sustainable atmospheric ammonia synthesis and nitrogen

116

synthesis is critically important to preventing back reactions. A recent study

pointed out a potential opportunity to overcome this obstacle, showing that the

introduction of Lewis acidic and redox-active sites to porous sorbents could

enhance the ammonia absorption (Tan et al., 2015), which could be another

worthwhile topic of investigation related to NTP ammonia synthesis. Since

ammonia can be easily decomposed by NTP, it is greatly beneficial if the

ammonia can be absorbed and removed from the plasma discharge region

immediately after it is synthesized. In this way, this technology can be significantly

more environmental sustainable due to the large increase in conversion and

efficiency.

Despite the current challenges, NTP is still considered a promising technology

for the future of ammonia production. The largest opportunity of this technology is

the elimination of fossil fuels from the ammonia synthesis process. Due to its

“green” nature, this technology is recognized as one potential method of realizing

sustainable agriculture (Pfromm, 2017). Furthermore, this technology could act as

an essential part of the wind-to-ammonia concept, where the hydrogen from

renewable sources can be fed the NTP reactor to sustainably generate ammonia

to be used as a fertilizer and clean fuel (Tallaksen et al., 2015). The new version

of this renewable concept with the implementation of NTP ammonia synthesis

technology is demonstrated in Figure 38. On farm sites or renewable energy

plants, the renewable energy sources such as wind, solar or hydro can be used to

provide energy for the electrolysis of water, which would provide the hydrogen gas

required for ammonia synthesis. The hydrogen gas, or other hydrogen-rich gases

Page 128: Sustainable atmospheric ammonia synthesis and nitrogen

117

such as methane, can also be converted from biomass agricultural waste through

gasification or anaerobic digestion. Lastly, the NTP plasma synthesis can be

powered by the renewable energy technologies. This renewable energy solution

provides a promising alternative for farm sites and small industries to become

accessible to sustainable energy and fertilizer via a renewable pathway.

Furthermore, this pathway has promising opportunities due to the fact that it fully

utilizes the advantages of the NTP ammonia synthesis such as less land use, low

capital cost, and less intense reaction conditions, and incorporates these

advantages with other renewable technology processes. Due to ammonia’s

excellent performance as an energy carrier and storage material, one of the

advantages of using wind energy to power the NTP ammonia synthesis is that the

ammonia could be considered as an excellent off-peak energy storage material

(Matzen, Alhajji, & Demirel, 2015). Furthermore, the combination of using wind

and electrolyzes with the NTP ammonia synthesis will reduce fuel transportation

costs and enable the synthesis fuel and fertilizers on site. For small industries and

local farms, this is critical since the distributed energy and fertilizer synthesis will

free them from market price fluctuations (Morgan, Manwell, & McGowan, 2014).

Additionally, using gasification to convert biomass into hydrogen gas can utilize

the waste produced by farms and generate sustainable fuel and energy (Lozano

& Lozano, 2017). However, the main challenges for the wind and biomass

gasification technologies are their inconsistency (Karltorp, 2016). For wind energy,

the power produced for the plasma is highly dependent on the wind resource at

the specific time frame. On the other hand, the large variance of the biomass

Page 129: Sustainable atmospheric ammonia synthesis and nitrogen

118

would result in the change of syn-gas components from the gasification process

(Sansaniwal, Rosen, & Tyagi, 2017). Therefore, the hydrogen separation process

needs to be adjusted according to the syngas composition, which would introduce

additional undesired cost.

Figure 38 An illustrative diagram of the renewable-to-ammonia approach

In terms of energy conversion, non-thermal plasma is a technology that could

theoretically fix nitrogen more efficiently than the conventional thermal fixation

methods (B. Patil, Hessel, et al., 2016). (Cherkasov, Ibhadon, & Fitzpatrick, 2015)

Pressure swing adsorption (PSA)

Renewable energy source (wind, solar, hydro etc.)

Electrolysis of water

Biomass conversion

Hydrogen NTP synthesis

Ammonia Nitrogen

Page 130: Sustainable atmospheric ammonia synthesis and nitrogen

119

reported that NTP N-fixation has a theoretical efficiency floor of 0.2MJ/mol, which

is more efficient than the Haber-Bosch method of 0.48 MJ/mol. Additionally,

(Azizov et al., 1980) allegedly reported an energy consumption of 0.29 MJ/mol

through the oxidation of diatomic Nitrogen in a microwave-induced plasma

generator. Other factors that make NTP one of the most attractive alternatives to

the Haber-Bosch method are the low land requirement and mobile capability.

Being able to produce ammonia in a continuous manner at a scale accessible to

the small industries and local farms, the NTP ammonia synthesis is well suitable

for the world’s trend of distributed and renewable energy production (Rune Ingels

& David B Graves, 2015).

A summative comparison between the NTP synthesis and the conventional

Haber-Bosch process is shown in Table 1. To achieve this, discussions and

analysis needs to surpass the summarization of published literature on process

developments. Therefore, this paper addresses the correlations between the

state-of-art of this technology with cleaner productions and how its development

could improve the sustainability in the ammonia production industry.

Page 131: Sustainable atmospheric ammonia synthesis and nitrogen

120

Table 3 A summative comparison between the conventional Haber-Bosch and NTP ammonia synthesis

Advantages Drawbacks

Conventional

Haber-Bosch

synthesis

Large scale available;

High production rate

Large energy input;

High temperature and

pressure Requires fossil

fuels;

Emission of green house

gases;

Centralized production;

High capital cost

Non-thermal

plasma

synthesis

Clean, carbon-free

production;

Low temperature and

pressure;

Accessible to small-scale,

on-site production;

Conversion and efficiency

need to be improved;

Available only at small scale

As mentioned in Section 1, the two other relatively new technologies that

synthesis ammonia under low temperature and pressure conditions are the

biological and electro-chemical synthesis. Among the three processes, NTP is

reported to have greater future improvement prospects of a variety of N-fixation

Page 132: Sustainable atmospheric ammonia synthesis and nitrogen

121

methods including Haber-Bosch, Biological, etc. (Cherkasov et al., 2015). The

biological ammonia synthesis utilizes nitrogenase enzyme in microorganisms to

fix the nitrogen from nitrogen gas or nitrate/nitrite solutions and convert them into

ammonia (Hinnemann & Nørskov, 2006). For this technology, the growth media

provides the energy sources of the microorganisms. Therefore, electricity, or other

energy input is not required. However, the drawbacks of this technology are that

the system could only be operated in the liquid phase with slow production rates,

and produces byproducts such as nitrate and nitrite species. Furthermore, like

other biological conversion processes, the improvements of this technology relies

heavily on the engineering of complex enzymes, which is a more challenging

obstacle compared with the development of inorganic catalysts. Instability in the

market of enzyme producers and lack of fixed cost is also a large hindrance to

expanding enzyme-based technologies. For the electro-chemical synthesis

approach, it uses electrolytes to facilitate hydrogen transfer between the reactants

and produce ammonia (Kyriakou, Garagounis, Vasileiou, Vourros, & Stoukides,

2017). The advantages of this technology is that it can be operated with both

liquid and gas reactants. However, the major limitations of this technology are the

high cost for the coated mesoporous electrolytes (Montoya, Tsai, Vojvodic, &

Nørskov, 2015). Furthermore, the production rate and the efficiency of this

technology could only be enhanced via the development of new electrolytes and

coated catalysts, which restricts its room for improvements. On the other hand,

the wide selection of plasma discharge types offers more opportunities for the

NTP synthesis of ammonia. According to the plasma type, the operational

Page 133: Sustainable atmospheric ammonia synthesis and nitrogen

122

pressure could be from vacuum to atmospheric pressure. Furthermore, the

different discharge type, as listed in table 2, has different discharge principles and

leads to a larger room of improvement for this technology beyond the catalyst

development.

Page 134: Sustainable atmospheric ammonia synthesis and nitrogen

123

CHAPTER 8 Conclusions and future remarks

In conclusion, the NTP synthesis of ammonia is a potential alternative

technology to the Haber-Bosch ammonia synthesis process that could lead to

cleaner ammonia production. It is a promising technology for the future of

ammonia production and is suitable for distributed hydrogen energy production

and storage. Overall, this technology could be beneficial to the ammonia industry,

through its potential to promote localized and environmentally friendly energy

production and storage. Throughout this research, the energy efficiency of the

plasma-assisted approach was increased to approximately one-fold. Although the

efficiency for the plasma approach (20 g/kWh achieved in this research) still

requires improvement to be competitive compared with the Haber-Bosch

approach. Future researcher could work to further develop catalysts with stronger

plasma synergistic activities, and optimizing the reactor to be more compatible

with the electric circuit within invertor and transformer. Additionally, exploring

absorbents with strong ammonia-holding capacity under plasma region could be

another direction to improve the energy efficiency of this process. Finally, the

opportunities of the non-thermal plasma technology lie with providing an avenue

towards a cleaner ammonia industry, including a renewable pathway that

incorporates this technology with other renewable energy approaches. So future

research regarding the compatibility of the plasma ammonia synthesis process

Page 135: Sustainable atmospheric ammonia synthesis and nitrogen

124

with renewable energy sources are significantly beneficial to its industrial

adoption.

Page 136: Sustainable atmospheric ammonia synthesis and nitrogen

125

Bibliography

Aihara, K., Akiyama, M., Deguchi, T., Tanaka, M., Hagiwara, R., & Iwamoto, M. (2016). Remarkable catalysis of a wool-like copper electrode for NH 3 synthesis from N 2 and H 2 in non-thermal atmospheric plasma. Chemical communications, 52(93), 13560-13563.

Aika, K.-i. (2017). Role of alkali promoter in ammonia synthesis over ruthenium catalysts—Effect on reaction mechanism. Catalysis Today, 286, 14-20.

Akay, G., & Zhang, K. (2016). Process Intensification in Ammonia Synthesis Using Novel Coassembled Supported Microporous Catalysts Promoted by Nonthermal Plasma. Industrial & Engineering Chemistry Research. doi:10.1021/acs.iecr.6b02053

Akay, G., & Zhang, K. (2017). Process Intensification in Ammonia Synthesis Using Novel Coassembled Supported Microporous Catalysts Promoted by Nonthermal Plasma. Industrial & Engineering Chemistry Research, 56(2), 457-468.

Amama, P. B., Ogebule, O., Maschmann, M. R., Sands, T. D., & Fisher, T. S. (2006). Dendrimer-assisted low-temperature growth of carbon nanotubes by plasma-enhanced chemical vapor deposition. Chemical communications(27), 2899-2901.

Atkinson, R., Baulch, D., Cox, R., Hampson Jr, R., Kerr, J., & Troe, J. (1989). Evaluated kinetic and photochemical data for atmospheric chemistry: supplement III. IUPAC subcommittee on gas kinetic data evaluation for atmospheric chemistry. Journal of Physical and Chemical reference data, 18(2), 881-1097.

Atkinson, R., Baulch, D. L., Cox, R. A., Crowley, J. N., Hampson, R. F., Hynes, R. G., . . . Troe, J. (2004). Evaluated kinetic and photochemical data for atmospheric chemistry: Volume I - gas phase reactions of Ox, HO, NOx and SOx species. Atmos. Chem. Phys., 4(6), 1461-1738. doi:10.5194/acp-4-1461-2004

Azizov, R., Zhivotov, V., Krotov, M., Rusanov, V., Tarasov, Y., & Fridma, A. (1980). Synthesis of nitrogen oxides in a nonequilibrium microwave discharge under electron-cyclotron resonance conditions, Khimiya Vysok. Energii, 14, 366-368.

Page 137: Sustainable atmospheric ammonia synthesis and nitrogen

126

Bai, M., Bai, X., Zhang, Z., & Bai, M. (2000). Synthesis of ammonia in a strong electric field discharge at ambient pressure. Plasma Chemistry and Plasma Processing, 20(4), 511-520.

Bai, M., Zhang, Z., Bai, M., Bai, X., & Gao, H. (2008). Synthesis of ammonia using CH4/N2 plasmas based on micro-gap discharge under environmentally friendly condition. Plasma Chemistry and Plasma Processing, 28(4), 405-414.

Bai, M., Zhang, Z., Bai, X., Bai, M., & Ning, W. (2003). Plasma synthesis of ammonia with a microgap dielectric barrier discharge at ambient pressure. Plasma Science, IEEE Transactions on, 31(6), 1285-1291.

Bao, X., Malik, M. A., Norton, D. G., Neculaes, V. B., Schoenbach, K. H., Heller, R., . . . Inzinna, L. P. (2014). Shielded sliding discharge-assisted hydrocarbon selective catalytic reduction of NOx over Ag/Al2O3 catalysts using diesel as a reductant. Plasma Chemistry and Plasma Processing, 34(4), 825-836.

Bardi, U., El Asmar, T., & Lavacchi, A. (2013). Turning electricity into food: the role of renewable energy in the future of agriculture. Journal of Cleaner Production, 53, 224-231.

Beck, J., Vartuli, J., Roth, W. J., Leonowicz, M., Kresge, C., Schmitt, K., . . . McCullen, S. (1992). A new family of mesoporous molecular sieves prepared with liquid crystal templates. Journal of the American Chemical Society, 114(27), 10834-10843.

Berthoud, R., Délichère, P., Gajan, D., Lukens, W., Pelzer, K., Basset, J.-M., . . . Copéret, C. (2008). Hydrogen and oxygen adsorption stoichiometries on silica supported ruthenium nanoparticles. Journal of Catalysis, 260(2), 387-391.

Bian, W., Shi, J., & Yin, X. (2009). Nitrogen fixation into water by pulsed high-voltage discharge. IEEE transactions on plasma science, 37(1), 211-218.

Burlica, R., Kirkpatrick, M. J., & Locke, B. R. (2006). Formation of reactive species in gliding arc discharges with liquid water. Journal of Electrostatics, 64(1), 35-43.

Carrasco, E., Jiménez-Redondo, M., Tanarro, I., & Herrero, V. J. (2011). Neutral and ion chemistry in low pressure dc plasmas of H 2/N 2 mixtures: routes for the efficient production of NH3 and NH 4+. Physical Chemistry Chemical Physics, 13(43), 19561-19572.

Page 138: Sustainable atmospheric ammonia synthesis and nitrogen

127

Chai, B.-h., Zheng, J.-m., Zhao, Q., & Pollack, G. H. (2008). Spectroscopic studies of solutes in aqueous solution. The Journal of Physical Chemistry A, 112(11), 2242-2247.

Chaudhary, Y. S., Ghatak, J., Bhatta, U. M., & Khushalani, D. (2006). One-step method for the self-assembly of metal nanoparticles onto facetted hollow silica tubes. Journal of Materials Chemistry, 16(36), 3619-3623.

Chauvet, L., Thérèse, L., Caillier, B., & Guillot, P. (2014). Characterization of an asymmetric DBD plasma jet source at atmospheric pressure. Journal of Analytical Atomic Spectrometry, 29(11), 2050-2057.

Cherkasov, N., Ibhadon, A., & Fitzpatrick, P. (2015). A review of the existing and alternative methods for greener nitrogen fixation. Chemical Engineering and Processing: Process Intensification, 90, 24-33.

Christensen, C. H., Johannessen, T., Sørensen, R. Z., & Nørskov, J. K. (2006). Towards an ammonia-mediated hydrogen economy? Catalysis Today, 111(1), 140-144.

Conrads, H., & Schmidt, M. (2000). Plasma generation and plasma sources. Plasma Sources Science and Technology, 9(4), 441.

Costa, J. A., Garcia, A. C., Santos, D. O., Sarmento, V. H., Porto, A. L., Mesquita, M. E. d., & Romão, L. P. (2014). A new functionalized MCM-41 mesoporous material for use in environmental applications. Journal of the Brazilian Chemical Society, 25(2), 197-207.

Cox, B., & Treyer, K. (2015). Environmental and economic assessment of a cracked ammonia fuelled alkaline fuel cell for off-grid power applications. Journal of Power Sources, 275, 322-335.

Cumaranatunge, L., Mulla, S., Yezerets, A., Currier, N., Delgass, W., & Ribeiro, F. (2007). Ammonia is a hydrogen carrier in the regeneration of Pt/BaO/Al 2 O 3 NOx traps with H 2. Journal of Catalysis, 246(1), 29-34.

Dündar-Tekkaya, E., & Yürüm, Y. (2015). Effect of loading bimetallic mixture of Ni and Pd on hydrogen storage capacity of MCM-41. international journal of hydrogen energy, 40(24), 7636-7643. doi:http://dx.doi.org/10.1016/j.ijhydene.2015.02.108

Davis, B. L., Dixon, D. A., Garner, E. B., Gordon, J. C., Matus, M. H., Scott, B., & Stephens, F. H. (2009). Efficient regeneration of partially spent ammonia borane fuel. Angewandte Chemie International Edition, 48(37), 6812-6816.

Demirci, U. B., & Miele, P. (2009). Sodium borohydride versus ammonia borane,

Page 139: Sustainable atmospheric ammonia synthesis and nitrogen

128

in hydrogen storage and direct fuel cell applications. Energy & Environmental Science, 2(6), 627-637.

Deshmukh, A., Kinage, A., Kumar, R., & Meijboom, R. (2010). Heterogenized Ru (II) phenanthroline complex for chemoselective hydrogenation of diketones under biphasic aqueous medium. Journal of Molecular Catalysis A: Chemical, 333(1), 114-120.

Dixit, M., Mishra, M., Joshi, P., & Shah, D. (2013). Study on the catalytic properties of silica supported copper catalysts. Procedia Engineering, 51, 467-472.

Dong, Y., Fei, X., Zhang, H., & Yu, L. (2015). Effects of Calcination Process on Photocatalytic Activity of TiO2/MCM-41 Photocatalyst. Journal of Advanced Oxidation Technologies, 18(2), 322-330.

Eliasson, B., & Gellert, B. (1990). Investigation of resonance and excimer radiation from a dielectric barrier discharge in mixtures of mercury and the rare gases. Journal of Applied Physics, 68(5), 2026-2037.

Eliasson, B., & Kogelschatz, U. (1991). Nonequilibrium volume plasma chemical processing. Plasma Science, IEEE Transactions on, 19(6), 1063-1077.

Elo, S., Kääriäinen, M., Kanste, O., Pölkki, T., Utriainen, K., & Kyngäs, H. (2014). Qualitative content analysis: A focus on trustworthiness. Sage Open, 4(1), 2158244014522633.

Erisman, J. W., Sutton, M. A., Galloway, J., Klimont, Z., & Winiwarter, W. (2008). How a century of ammonia synthesis changed the world. Nature Geoscience, 1(10), 636-639.

Erjavec, B., Kaplan, R., Djinović, P., & Pintar, A. (2013). Catalytic wet air oxidation of bisphenol A model solution in a trickle-bed reactor over titanate nanotube-based catalysts. Applied Catalysis B: Environmental, 132, 342-352.

Florian, J., Merbahi, N., Wattieaux, G., Plewa, J.-M., & Yousfi, M. (2015). Comparative Studies of Double Dielectric Barrier Discharge and Microwave Argon Plasma Jets at Atmospheric Pressure for Biomedical Applications.

Fridman, A., Chirokov, A., & Gutsol, A. (2005). Non-thermal atmospheric pressure discharges. Journal of Physics D: Applied Physics, 38(2), R1.

Fridman, A., & Kennedy, L. A. (2004). Plasma physics and engineering: CRC press.

Page 140: Sustainable atmospheric ammonia synthesis and nitrogen

129

Gómez-Ramírez, A., Cotrino, J., Lambert, R., & González-Elipe, A. (2015). Efficient synthesis of ammonia from N2 and H2 alone in a ferroelectric packed-bed DBD reactor. Plasma Sources Science and Technology, 24(6), 065011.

Gómez‐Ramírez, A., Montoro‐Damas, A. M., Cotrino, J., Lambert, R. M., &

González‐Elipe, A. R. (2016). About the enhancement of chemical yield

during the atmospheric plasma synthesis of ammonia in a ferroelectric packed bed reactor. Plasma Processes and Polymers.

Gómez‐Ramírez, A., Montoro‐Damas, A. M., Cotrino, J., Lambert, R. M., &

González‐Elipe, A. R. (2017). About the enhancement of chemical yield

during the atmospheric plasma synthesis of ammonia in a ferroelectric packed bed reactor. Plasma Processes and Polymers, 14(6).

Gallezot, P., Chaumet, S., Perrard, A., & Isnard, P. (1997). Catalytic wet air oxidation of acetic acid on carbon-supported ruthenium catalysts. Journal of Catalysis, 168(1), 104-109.

Galloway, J. N., Townsend, A. R., Erisman, J. W., Bekunda, M., Cai, Z., Freney, J. R., . . . Sutton, M. A. (2008). Transformation of the nitrogen cycle: recent trends, questions, and potential solutions. Science, 320(5878), 889-892.

Gasser, S., Ruiz, R., Kolawa, E., & Nicolet, M. A. (1999). Instability of Amorphous

Ru‐Si‐O Thin Films under Thermal Oxidation. Journal of The

Electrochemical Society, 146(4), 1546-1548.

Gilbert, P., Alexander, S., Thornley, P., & Brammer, J. (2014). Assessing economically viable carbon reductions for the production of ammonia from biomass gasification. Journal of Cleaner Production, 64, 581-589.

Gilland, B. (2014). Is a Haber-Bosch World Sustainable? Population, Nutrition, Cereals, Nitrogen and Environment. The Journal of Social, Political, and Economic Studies, 39(2), 166.

Gokulakrishnan, N., Pandurangan, A., Somanathan, T., & Sinha, P. (2010). Uptake of decontaminating agent from aqueous solution: a study on adsorption behaviour of oxalic acid over Al-MCM-41 adsorbents. Journal of Porous Materials, 17(6), 763-771.

Guangming, L., Peihua, M., Zhiming, W., Mingzhen, L., & Minxong, C. (2004). Investigation of thermal decomposition of MgCl2 hexammoniate and MgCl2 biglycollate biammoniate by DTA–TG, XRD and chemical analysis. Thermochimica acta, 412(1), 149-153. doi:https://doi.org/10.1016/j.tca.2003.09.018

Page 141: Sustainable atmospheric ammonia synthesis and nitrogen

130

Hagen, S., Barfod, R., Fehrmann, R., Jacobsen, C. J., Teunissen, H. T., & Chorkendorff, I. (2003). Ammonia synthesis with barium-promoted iron–cobalt alloys supported on carbon. Journal of Catalysis, 214(2), 327-335.

Hargreaves, J. (2014). Nitrides as ammonia synthesis catalysts and as potential nitrogen transfer reagents. Applied Petrochemical Research, 4(1), 3-10.

Haruyama, T., Namise, T., Shimoshimizu, N., Uemura, S., Takatsuji, Y., Hino, M., . . . Kohno, M. (2016). Non-catalyzed one-step synthesis of ammonia from atmospheric air and water. Green Chemistry, 18(16), 4536-4541.

Henrici‐Olivé, G., & Olive, S. (1969). Non‐Enzymatic Activation of Molecular

Nitrogen. Angewandte Chemie International Edition in English, 8(9), 650-659.

Hinnemann, B., & Nørskov, J. K. (2006). Catalysis by Enzymes: The Biological Ammonia Synthesis. Topics in Catalysis, 37(1), 55-70. doi:10.1007/s11244-006-0002-0

Hinrichsen, O., Rosowski, F., Hornung, A., Muhler, M., & Ertl, G. (1997). The kinetics of ammonia synthesis over Ru-based catalysts: 1. The dissociative chemisorption and associative desorption of N2. Journal of Catalysis, 165(1), 33-44.

Hołub, M. (2012). On the measurement of plasma power in atmospheric pressure DBD plasma reactors. International Journal of Applied Electromagnetics and Mechanics, 39(1-4), 81-87.

Hong, J., Aramesh, M., Shimoni, O., Seo, D. H., Yick, S., Greig, A., . . . Murphy, A. B. (2016). Plasma Catalytic Synthesis of Ammonia Using Functionalized-Carbon Coatings in an Atmospheric-Pressure Non-equilibrium Discharge. Plasma Chemistry and Plasma Processing, 36(4), 917-940.

Hong, J., Pancheshnyi, S., Tam, E., Lowke, J. J., Prawer, S., & Murphy, A. B. (2017). Kinetic modelling of NH3 production in N2–H2 non-equilibrium atmospheric-pressure plasma catalysis. Journal of Physics D: Applied Physics, 50(15), 154005.

Hong, J., Prawer, S., & Murphy, A. B. (2014). Production of Ammonia by Heterogeneous Catalysis in a Packed-Bed Dielectric-Barrier Discharge: Influence of Argon Addition and Voltage. IEEE Transactions on Plasma Science, 42(10), 2338-2339.

Huang, C.-C., Li, H.-S., & Chen, C.-H. (2008). Effect of surface acidic oxides of activated carbon on adsorption of ammonia. Journal of Hazardous

Page 142: Sustainable atmospheric ammonia synthesis and nitrogen

131

Materials, 159(2-3), 523-527.

Ingels, R., & Graves, D. B. (2015). Improving the Efficiency of Organic Fertilizer and Nitrogen Use via Air Plasma and Distributed Renewable Energy. 5(2-4), 257-270. doi:10.1615/PlasmaMed.2016015763

Ingels, R., & Graves, D. B. (2015). Improving the Efficiency of Organic Fertilizer and Nitrogen Use via Air Plasma and Distributed Renewable Energy. Plasma Medicine, 5(2-4).

Ishida, T., Hirasawa, K., Dozen, M., & Tada, Y. (2011). Appearance of DC dielectric barrier discharge. Paper presented at the Electrical Insulating Materials (ISEIM), Proceedings of 2011 International Conference on.

Itikawa, Y., & Mason, N. (2005). Cross sections for electron collisions with water molecules. Journal of Physical and Chemical reference data, 34(1), 1-22.

Iwamoto, J., Itoh, M., Kajita, Y., Saito, M., & Machida, K.-i. (2007). Ammonia synthesis on magnesia supported ruthenium catalysts with mesoporous structure. Catalysis Communications, 8(6), 941-944.

Iwamoto, M., Akiyama, M., Aihara, K., & Deguchi, T. (2017). Ammonia Synthesis on Wool-Like Au, Pt, Pd, Ag, or Cu Electrode Catalysts in Nonthermal Atmospheric-Pressure Plasma of N2 and H2. ACS Catalysis, 7(10), 6924-6929.

Jennings, J. R. (2013). Catalytic ammonia synthesis: fundamentals and practice: Springer Science & Business Media.

Ji, J., Duan, X., Gong, X., Qian, G., Zhou, X., Chen, D., & Yuan, W. (2013). Promotional effect of carbon on Fe catalysts for ammonia decomposition: a density functional theory study. Industrial & Engineering Chemistry Research, 52(48), 17151-17155.

Jidenko, N., Bourgeois, E., & Borra, J. (2010). Temperature profiles in filamentary dielectric barrier discharges at atmospheric pressure. Journal of Physics D: Applied Physics, 43(29), 295203.

Joo, S. H., Park, J. Y., Renzas, J. R., Butcher, D. R., Huang, W., & Somorjai, G. A. (2010). Size effect of ruthenium nanoparticles in catalytic carbon monoxide oxidation. Nano letters, 10(7), 2709-2713.

Kanbara, S., Kitano, M., Inoue, Y., Yokoyama, T., Hara, M., & Hosono, H. (2015). Mechanism Switching of Ammonia Synthesis Over Ru-Loaded Electride Catalyst at Metal–Insulator Transition. J. Am. Chem. Soc, 137(45), 14517-14524.

Page 143: Sustainable atmospheric ammonia synthesis and nitrogen

132

Karltorp, K. (2016). Challenges in mobilising financial resources for renewable energy—The cases of biomass gasification and offshore wind power. Environmental Innovation and Societal Transitions, 19, 96-110.

Keller, S., Rajasekaran, P., Bibinov, N., & Awakowicz, P. (2012). Characterization of transient discharges under atmospheric-pressure conditions applying nitrogen photoemission and current measurements. Journal of Physics D: Applied Physics, 45(12), 125202.

Kerpal, C., Harding, D. J., Lyon, J. T., Meijer, G., & Fielicke, A. (2013). N2 activation by neutral ruthenium clusters. The Journal of Physical Chemistry C, 117(23), 12153-12158.

Khatun, H., Sharma, A., & Barhai, P. (2010). Experimental study of low-pressure nitrogen dielectric barrier discharge. Brazilian Journal of Physics, 40(4), 450-453.

Kim, H. H. (2004). Nonthermal plasma processing for air‐pollution control: a

historical review, current issues, and future prospects. Plasma Processes and Polymers, 1(2), 91-110.

Kim, H. H., Teramoto, Y., Ogata, A., Takagi, H., & Nanba, T. (2016).

Atmospheric‐pressure nonthermal plasma synthesis of ammonia over

ruthenium catalysts. Plasma Processes and Polymers.

Kim, K., & Winograd, N. (1974). X-ray photoelectron spectroscopic studies of ruthenium-oxygen surfaces. Journal of Catalysis, 35(1), 66-72.

Kitano, M., Inoue, Y., Yamazaki, Y., Hayashi, F., Kanbara, S., Matsuishi, S., . . . Hosono, H. (2012). Ammonia synthesis using a stable electride as an electron donor and reversible hydrogen store. Nature chemistry, 4(11), 934-940.

Kitano, M., Kanbara, S., Inoue, Y., Kuganathan, N., Sushko, P. V., Yokoyama, T., . . . Hosono, H. (2015). Electride support boosts nitrogen dissociation over ruthenium catalyst and shifts the bottleneck in ammonia synthesis. Nature communications, 6.

Kogelschatz, U. (2003). Dielectric-barrier discharges: their history, discharge physics, and industrial applications. Plasma chemistry and plasma processing, 23(1), 1-46.

Kovačević, V. V., Dojčinović, B. P., Jović, M., Roglić, G. M., Obradović, B. M., & Kuraica, M. M. (2017). Measurement of reactive species generated by dielectric barrier discharge in direct contact with water in different atmospheres. Journal of Physics D: Applied Physics, 50(15), 155205.

Page 144: Sustainable atmospheric ammonia synthesis and nitrogen

133

Kowalczyk, Z., Krukowski, M., Raróg-Pilecka, W., Szmigiel, D., & Zielinski, J. (2003). Carbon-based ruthenium catalyst for ammonia synthesis: Role of the barium and caesium promoters and carbon support. Applied Catalysis A: General, 248(1–2), 67-73. doi:https://doi.org/10.1016/S0926-860X(03)00150-9

Kubota, Y., Koga, K., Ohno, M., & Hara, T. (2010). Synthesis of Ammonia through Direct Chemical Reactions between an Atmospheric Nitrogen Plasma Jet and a Liquid. Plasma and Fusion Research, 5, 042-042.

Kubota, Y., Masatomi, O., & Tamio, H. (2010). Synthesis of Ammonia through Direct Chemical Reactions between an Atmospheric Nitrogen Plasma Jet and a Liquid. Plasma and Fusion Research, 5, 042-042.

Kyriakou, V., Garagounis, I., Vasileiou, E., Vourros, A., & Stoukides, M. (2017). Progress in the Electrochemical Synthesis of Ammonia. Catalysis Today, 286(Supplement C), 2-13. doi:https://doi.org/10.1016/j.cattod.2016.06.014

Lan, R., Irvine, J. T., & Tao, S. (2012). Ammonia and related chemicals as potential indirect hydrogen storage materials. international journal of hydrogen energy, 37(2), 1482-1494.

Lan, R., Irvine, J. T., & Tao, S. (2013). Synthesis of ammonia directly from air and water at ambient temperature and pressure. Scientific reports, 3.

Larichev, Y. V. (2010). Effect of Cs+ Promoter in Ru/MgO Catalysts. The Journal of Physical Chemistry C, 115(3), 631-635.

Larichev, Y. V., Moroz, B. L., Zaikovskii, V. I., Yunusov, S. M., Kalyuzhnaya, E. S., Shur, V. B., & Bukhtiyarov, V. I. (2007). XPS and TEM studies on the role of the support and alkali promoter in Ru/MgO and Ru-Cs+/MgO catalysts for ammonia synthesis. The Journal of Physical Chemistry C, 111(26), 9427-9436.

Liu, C., Brown, N. M., & Meenan, B. J. (2006). Uniformity analysis of dielectric barrier discharge (DBD) processed polyethylene terephthalate (PET) surface. Applied surface science, 252(6), 2297-2310.

Liu, D.-X., Bruggeman, P., Iza, F., Rong, M.-Z., & Kong, M. G. (2010). Global model of low-temperature atmospheric-pressure He+ H2O plasmas. Plasma Sources Science and Technology, 19(2), 025018.

Liu, H. (2014). Ammonia synthesis catalyst 100 years: Practice, enlightenment and challenge. Chinese Journal of Catalysis, 35(10), 1619-1640. doi:http://dx.doi.org/10.1016/S1872-2067(14)60118-2

Page 145: Sustainable atmospheric ammonia synthesis and nitrogen

134

Liu, Y., Ma, R., Luo, R., Luo, K., Gao, M., & Pan, H. (2011). Hydrogen Storage Properties of the Mg (NH3) 6Cl2-LiH Combined System. Materials transactions, 52(4), 627-634.

Lozano, F. J., & Lozano, R. (2017). Assessing the potential sustainability benefits of agricultural residues: Biomass conversion to syngas for energy generation or to chemicals production. Journal of Cleaner Production.

Lu, S., Chen, L., Huang, Q., Yang, L., Du, C., Li, X., & Yan, J. (2014). Decomposition of ammonia and hydrogen sulfide in simulated sludge drying waste gas by a novel non-thermal plasma. Chemosphere, 117, 781-785. doi:http://dx.doi.org/10.1016/j.chemosphere.2014.10.036

Luo, J., Kang, X., Fang, Z., & Wang, P. (2011). Promotion of hydrogen release from ammonia borane with magnesium nitride. Dalton Transactions, 40(24), 6469-6474.

Ma, H., Chen, P., & Ruan, R. (2001). H2S and NH3 removal by silent discharge plasma and ozone combo-system. Plasma Chemistry and Plasma Processing, 21(4), 611-624.

Ma, H., Chen, P., Zhang, M., Lin, X., & Ruan, R. (2002). Study of SO2 removal using non-thermal plasma induced by dielectric barrier discharge (DBD). Plasma Chemistry and Plasma Processing, 22(2), 239-254.

Ma, Z.-Y., Feng, P., Gao, Q.-X., Lu, Y.-N., Liu, J.-R., & Li, W.-T. (2015). CH4 emissions and reduction potential in wastewater treatment in China. Advances in Climate Change Research, 6(3), 216-224. doi:https://doi.org/10.1016/j.accre.2015.11.006

Ma, Z., Zhao, S., Pei, X., Xiong, X., & Hu, B. (2017). New insights into the support morphology-dependent ammonia synthesis activity of Ru/CeO 2 catalysts. Catalysis Science & Technology, 7(1), 191-199.

Maffei, N., Pelletier, L., Charland, J., & McFarlan, A. (2007). A direct ammonia fuel cell using barium cerate proton conducting electrolyte doped with gadolinium and praseodymium. Fuel Cells, 7(4), 323-328.

Makgwane, P. R., & Ray, S. S. (2013). Nanosized ruthenium particles decorated carbon nanofibers as active catalysts for the oxidation of p-cymene by molecular oxygen. Journal of Molecular Catalysis A: Chemical, 373, 1-11. doi:http://dx.doi.org/10.1016/j.molcata.2013.02.017

Malmali, M., Wei, Y., McCormick, A., & Cussler, E. L. (2016). Ammonia synthesis at reduced pressure via reactive separation. Industrial & Engineering Chemistry Research, 55(33), 8922-8932.

Page 146: Sustainable atmospheric ammonia synthesis and nitrogen

135

Matsumoto, O. (1981). II. Plasma production and plasma chemical reaction. J. Electron Mater.(Japan), 20(3), 130-135.

Matsuo, K., Takatsuji, Y., Kohno, M., Kamachi, T., Nakada, H., & Haruyama, T. (2015). Dispersed-phase Interfaces between Mist Water Particles and Oxygen Plasma Efficiently Produce Singlet Oxygen (1O2) and Hydroxyl Radical (• OH). Electrochemistry, 83(9), 721-724.

Matzen, M., Alhajji, M., & Demirel, Y. (2015). Technoeconomics and sustainability of renewable methanol and ammonia productions using wind power-based hydrogen. J Adv Chem Eng, 5(128), 2.

Mazumder, M., Sims, R., Biris, A., Srirama, P., Saini, D., Yurteri, C., . . . Sharma, R. (2006). Twenty-first century research needs in electrostatic processes applied to industry and medicine. Chemical Engineering Science, 61(7), 2192-2211.

Misra, N. N., Schlüter, O., & Cullen, P. J. (2016). Chapter 1 - Plasma in Food and Agriculture Cold Plasma in Food and Agriculture (pp. 1-16). San Diego: Academic Press.

Miura, D., & Tezuka, T. (2014). A comparative study of ammonia energy systems as a future energy carrier, with particular reference to vehicle use in Japan. Energy, 68, 428-436.

Mizushima, T., Matsumoto, K., Ohkita, H., & Kakuta, N. (2007). Catalytic effects of metal-loaded membrane-like alumina tubes on ammonia synthesis in atmospheric pressure plasma by dielectric barrier discharge. Plasma Chemistry and Plasma Processing, 27(1), 1-11.

Mizushima, T., Matsumoto, K., Sugoh, J.-i., Ohkita, H., & Kakuta, N. (2004). Tubular membrane-like catalyst for reactor with dielectric-barrier-discharge plasma and its performance in ammonia synthesis. Applied Catalysis A: General, 265(1), 53-59.

Mondal, D. K., Mondal, C., & Roy, S. (2016). Catalytic wet air oxidation of aqueous solution of phenol in a fixed bed reactor over Ru catalysts supported on ceria promoted MCM-41. RSC Advances, 6(115), 114383-114395.

Montoya, J. H., Tsai, C., Vojvodic, A., & Nørskov, J. K. (2015). The Challenge of Electrochemical Ammonia Synthesis: A New Perspective on the Role of Nitrogen Scaling Relations. ChemSusChem, 8(13), 2180-2186. doi:10.1002/cssc.201500322

Morgan, E., Manwell, J., & McGowan, J. (2014). Wind-powered ammonia fuel

Page 147: Sustainable atmospheric ammonia synthesis and nitrogen

136

production for remote islands: A case study. Renewable Energy, 72(Supplement C), 51-61. doi:https://doi.org/10.1016/j.renene.2014.06.034

Mun, C., Ehrhardt, J. J., Lambert, J., & Madic, C. (2007). XPS investigations of ruthenium deposited onto representative inner surfaces of nuclear reactor containment buildings. Applied Surface Science, 253(18), 7613-7621. doi:http://dx.doi.org/10.1016/j.apsusc.2007.03.071

Nakajima, J., & Sekiguchi, H. (2008). Synthesis of ammonia using microwave discharge at atmospheric pressure. Thin Solid Films, 516(13), 4446-4451.

Nakamoto, K. (1986). Infrared and Raman spectra of inorganic and coordination compounds: Wiley Online Library.

Naz, M., Ghaffar, A., Rehman, N., Shahid, S., & Shukrullah, S. (2012). Characterization of an In-house Built 50 Hz Single Dielectric Barrier Discharge System Having Asymmetric Electrodes. International Journal of Engineering & Technology IJET-IJENS, 12(05), 53-60.

Neyts, E. C., Ostrikov, K., Sunkara, M. K., & Bogaerts, A. (2015). Plasma catalysis: synergistic effects at the nanoscale. Chemical reviews, 115(24), 13408-13446.

Nie, Y., Zheng, Q., Liang, X., Gu, D., Lu, M., Min, M., & Ji, J. (2013). Decomposition treatment of SO2F2 using packed bed DBD plasma followed by chemical absorption. Environmental science & technology, 47(14), 7934-7939.

Nyquist, R. A., & Kagel, R. O. (2012). Handbook of infrared and raman spectra of inorganic compounds and organic salts: infrared spectra of inorganic compounds (Vol. 4): Academic press.

Oda, T. (2003). Non-thermal plasma processing for environmental protection: decomposition of dilute VOCs in air. Journal of electrostatics, 57(3), 293-311.

Paliwoda, M. C. (2016). Intensity control of dielectric barrier discharge filaments.

Park, D. P., Davis, K., Gilani, S., Alonzo, C.-A., Dobrynin, D., Friedman, G., . . . Fridman, G. (2013). Reactive nitrogen species produced in water by non-equilibrium plasma increase plant growth rate and nutritional yield. Current Applied Physics, 13, S19-S29.

Patil, B., Cherkasov, N., Lang, J., Ibhadon, A., Hessel, V., & Wang, Q. (2016). Low temperature plasma-catalytic NO x synthesis in a packed DBD reactor: Effect of support materials and supported active metal oxides. Applied

Page 148: Sustainable atmospheric ammonia synthesis and nitrogen

137

Catalysis B: Environmental, 194, 123-133.

Patil, B., Hessel, V., Lang, J., & Wang, Q. (2016). Plasma-Assisted Nitrogen Fixation Reactions Alternative Energy Sources for Green Chemistry (pp. 296-338).

Patil, B., Wang, Q., Hessel, V., & Lang, J. (2015). Plasma N 2-fixation: 1900–2014. Catalysis today, 256, 49-66.

Patil, B. S., Peeters, F., van Rooij, G. J., Medrano, J., Gallucci, F., Lang, J., . . . Hessel, V. (2017). Plasma assisted nitrogen oxide production from air: Using pulsed powered gliding arc reactor for a containerized plant. AIChE Journal.

Penetrante, B., Hsiao, M., Bardsley, J., Merritt, B., Vogtlin, G., Kuthi, A., . . . Bayless, J. (1997). Identification of mechanisms for decomposition of air pollutants by non-thermal plasma processing. Plasma sources science and technology, 6(3), 251.

Penetrante, B., Hsiao, M., Merritt, B., Vogtlin, G., Wallman, P., Neiger, M., . . .

Broer, S. (1996). Pulsed corona and dielectric‐barrier discharge

processing of NO in N2. Applied Physics Letters, 68(26), 3719-3721.

Penetrante, B. M., Hsiao, M. C., Merritt, B. T., Vogtlin, G. E., & Wallman, H. P. (1995). Comparison of electrical discharge techniques for nonthermal plasma processing of NO in N 2. Plasma Science, IEEE Transactions on, 23(4), 679-687.

Penetrante, B. M., & Schultheis, S. E. (2013). Non-Thermal Plasma Techniques for Pollution Control: Part B: Electron Beam and Electrical Discharge Processing (Vol. 34): Springer Science & Business Media.

Perring, L., Bussy, F., Gachon, J., & Feschotte, P. (1999). The Ruthenium–Silicon system. Journal of alloys and compounds, 284(1), 198-205.

Petitpas, G., Rollier, J.-D., Darmon, A., Gonzalez-Aguilar, J., Metkemeijer, R., & Fulcheri, L. (2007). A comparative study of non-thermal plasma assisted reforming technologies. international journal of hydrogen energy, 32(14), 2848-2867.

Pfromm, P. H. (2017). Towards sustainable agriculture: Fossil-free ammonia. Journal of Renewable and Sustainable Energy, 9(3), 034702.

Rahemi, N., Haghighi, M., Babaluo, A. A., Jafari, M. F., & Estifaee, P. (2013). Synthesis and physicochemical characterizations of Ni/Al 2 O 3–ZrO 2 nanocatalyst prepared via impregnation method and treated with

Page 149: Sustainable atmospheric ammonia synthesis and nitrogen

138

non-thermal plasma for CO 2 reforming of CH 4. Journal of Industrial and Engineering Chemistry, 19(5), 1566-1576.

Razon, L. F. (2014). Life cycle analysis of an alternative to the haber‐bosch

process: Non‐renewable energy usage and global warming potential of

liquid ammonia from cyanobacteria. Environmental Progress & Sustainable Energy, 33(2), 618-624.

Rochefort, D., Dabo, P., Guay, D., & Sherwood, P. (2003). XPS investigations of thermally prepared RuO 2 electrodes in reductive conditions. Electrochimica Acta, 48(28), 4245-4252.

Ruan, R., Deng, S., Le, Z., Cheng, Y., Lin, X., & Chen, P. (2014). Non-thermal plasma synthesis with carbon component: U.S. Patent 8,641,872, Washington, DC: U.S. Patent and Trademark Office.

Ruan, R. R., Chen, P. L., Ning, A., Bogaard, R. L., Robinson, D. G., Deng, S., . . . Bie, C. (2000). Dielectric barrier discharge system and method for decomposing hazardous compounds in fluids: Google Patents.

Ruan, R. R., Han, W., Ning, A., Chen, P. L., Goodrich, P. R., & Zhang, R. (1999). Treatment of odorous and hazardous gases using non-thermal plasma. Journal of Advanced Oxidation Technologies, 4(3), 328-332.

Saadatjou, N., Jafari, A., & Sahebdelfar, S. (2015). Ruthenium nanocatalysts for ammonia synthesis: a Review. Chemical Engineering Communications, 202(4), 420-448.

Sakakura, T., Uemura, S., Hino, M., Kiyomatsu, S., Takatsuji, Y., Yamasaki, R., . . . Haruyama, T. (2017a). Excitation of H2O at plasma/water interface by UV irradiation for elevation of ammonia production. Green Chemistry. doi:10.1039/C7GC03007J

Sakakura, T., Uemura, S., Hino, M., Kiyomatsu, S., Takatsuji, Y., Yamasaki, R., . . . Haruyama, T. (2017b). Excitation of H 2 O at plasma/water interface by UV irradiation for elevation of ammonia production. Green Chemistry.

Sandali, A., & Chériti, A. (2017). Pulse Density Modulation Applied to Series

Resonant Inverter and Ac‐Ac Conversion Recent Developments on

Power Inverters: InTech.

Sansaniwal, S., Rosen, M., & Tyagi, S. (2017). Global challenges in the sustainable development of biomass gasification: An overview. Renewable and Sustainable Energy Reviews, 80, 23-43.

Schütze, A., Jeong, J. Y., Babayan, S. E., Park, J., Selwyn, G. S., & Hicks, R. F.

Page 150: Sustainable atmospheric ammonia synthesis and nitrogen

139

(1998). The atmospheric-pressure plasma jet: a review and comparison to other plasma sources. Plasma Science, IEEE Transactions on, 26(6), 1685-1694.

Schiavon, M., Scapinello, M., Tosi, P., Ragazzi, M., Torretta, V., & Rada, E. C. (2015). Potential of non-thermal plasmas for helping the biodegradation of volatile organic compounds (VOCs) released by waste management plants. Journal of Cleaner Production, 104, 211-219.

Schiavon, M., Schiorlin, M., Torretta, V., Brandenburg, R., & Ragazzi, M. (2017). Non-thermal plasma assisting the biofiltration of volatile organic compounds. Journal of Cleaner Production, 148, 498-508.

Sharma, V., Hosoi, K., Mori, T., & Kuroda, S. (2013). Electrical and Optical Characterization of Cold Atmospheric Pressure Plasma Jet and the Effects of N2 Gas on Argon Plasma Discharge. Paper presented at the Applied Mechanics and Materials.

Sharonov, V. E., & Aristov, Y. I. (2005a). Ammonia adsorption by MgCl2, CaCl2 and BaCl2 confined to porous alumina: the fixed bed adsorber. Reaction Kinetics and Catalysis Letters, 85(1), 183-188.

Sharonov, V. E., & Aristov, Y. I. (2005b). Ammonia adsorption by MgCl 2, CaCl 2 and BaCl 2 confined to porous alumina: the fixed bed adsorber. Reaction Kinetics and Catalysis Letters, 85(1), 183-188.

Sharonov, V. E., Veselovskaya, J. V., & Aristov, Y. I. (2006). Ammonia sorption on composites ‘CaCl2 in inorganic host matrix’: isosteric chart and its performance. International Journal of Low-Carbon Technologies, 1(3), 191-200.

Shen, J., Tian, Y., Li, Y., Ma, R., Zhang, Q., Zhang, J., & Fang, J. (2016). Bactericidal Effects against S. aureus and Physicochemical Properties of Plasma Activated Water stored at different temperatures. Scientific reports, 6.

Smil, V. (1997). Global population and the nitrogen cycle. Scientific American, 277(1), 76-81.

Smythe, N. C., & Gordon, J. C. (2010). Ammonia borane as a hydrogen carrier: dehydrogenation and regeneration. European Journal of Inorganic Chemistry, 2010(4), 509-521.

Stasiulaitiene, I., Martuzevicius, D., Abromaitis, V., Tichonovas, M., Baltrusaitis, J., Brandenburg, R., . . . Schwock, A. (2016). Comparative life cycle assessment of plasma-based and traditional exhaust gas treatment

Page 151: Sustainable atmospheric ammonia synthesis and nitrogen

140

technologies. Journal of Cleaner Production, 112, 1804-1812.

Stavrou, E., Yao, Y., Zaug, J. M., Bastea, S., Kalkan, B., Konôpková, Z., & Kunz, M. (2016). High-pressure X-ray diffraction, Raman, and computational studies of MgCl 2 up to 1 Mbar: Extensive pressure stability of the β-MgCl 2 layered structure. Scientific reports, 6, 30631.

Stephens, F. H., Baker, R. T., Matus, M. H., Grant, D. J., & Dixon, D. A. (2007). Acid initiation of ammonia–borane dehydrogenation for hydrogen storage. Angewandte Chemie International Edition, 46(5), 746-749.

Sugiyama, K., Akazawa, K., Oshima, M., Miura, H., Matsuda, T., & Nomura, O. (1986). Ammonia synthesis by means of plasma over MgO catalyst. Plasma chemistry and plasma processing, 6(2), 179-193.

Szili, E. J., Hong, S.-H., Oh, J.-S., Gaur, N., & Short, R. D. (2017). Tracking the Penetration of Plasma Reactive Species in Tissue Models. Trends in Biotechnology. doi:https://doi.org/10.1016/j.tibtech.2017.07.012

Szmigiel, D., Bielawa, H., Kurtz, M., Hinrichsen, O., Muhler, M., Raróg, W., . . . Zieliński, J. (2002). The kinetics of ammonia synthesis over ruthenium-based catalysts: The role of barium and cesium. Journal of Catalysis, 205(1), 205-212.

Tallaksen, J., Bauer, F., Hulteberg, C., Reese, M., & Ahlgren, S. (2015). Nitrogen fertilizers manufactured using wind power: greenhouse gas and energy balance of community-scale ammonia production. Journal of Cleaner Production, 107, 626-635. doi:http://dx.doi.org/10.1016/j.jclepro.2015.05.130

Tan, B., Chen, C., Cai, L.-X., Zhang, Y.-J., Huang, X.-Y., & Zhang, J. (2015). Introduction of lewis acidic and redox-active sites into a porous framework for ammonia capture with visual color response. Inorganic chemistry, 54(7), 3456-3461.

Tanaka, S., Uyama, H., & Matsumoto, O. (1994). Synergistic effects of catalysts and plasmas on the synthesis of ammonia and hydrazine. Plasma Chemistry and Plasma Processing, 14(4), 491-504.

Tranfield, D., Denyer, D., & Smart, P. (2003). Towards a methodology for

developing evidence‐informed management knowledge by means of

systematic review. British journal of management, 14(3), 207-222.

Urashima, K., & Chang, J.-S. (2000). Removal of volatile organic compounds from air streams and industrial flue gases by non-thermal plasma technology. IEEE Transactions on Dielectrics and Electrical Insulation, 7(5), 602-614.

Page 152: Sustainable atmospheric ammonia synthesis and nitrogen

141

Uyama, H., & Matsumoto, O. (1989). Synthesis of ammonia in high-frequency discharges. Plasma Chemistry and Plasma Processing, 9(1), 13-24.

Uyama, H., Nakamura, T., Tanaka, S., & Matsumoto, O. (1993). Catalytic effect of iron wires on the syntheses of ammonia and hydrazine in a radio-frequency discharge. Plasma Chemistry and Plasma Processing, 13(1), 117-131.

Van Durme, J., Dewulf, J., Leys, C., & Van Langenhove, H. (2008). Combining non-thermal plasma with heterogeneous catalysis in waste gas treatment: a review. Applied Catalysis B: Environmental, 78(3), 324-333.

Van Helden, J., Wagemans, W., Yagci, G., Zijlmans, R., Schram, D., Engeln, R., . . . Röpcke, J. (2007). Detailed study of the plasma-activated catalytic generation of ammonia in N 2-H 2 plasmas. Journal of Applied Physics, 101(4), 043305.

Vanama, P. K., Kumar, A., Ginjupalli, S. R., & Komandur, V. R. C. (2015). Vapor-phase hydrogenolysis of glycerol over nanostructured Ru/MCM-41 catalysts. Catalysis Today, 250, 226-238. doi:http://dx.doi.org/10.1016/j.cattod.2014.03.036

Vojvodic, A., Medford, A. J., Studt, F., Abild-Pedersen, F., Khan, T. S., Bligaard, T., & Nørskov, J. (2014). Exploring the limits: A low-pressure, low-temperature Haber–Bosch process. Chemical Physics Letters, 598, 108-112.

Von Blottnitz, H., & Curran, M. A. (2007). A review of assessments conducted on bio-ethanol as a transportation fuel from a net energy, greenhouse gas, and environmental life cycle perspective. Journal of Cleaner Production, 15(7), 607-619.

Wagner, C. D., & Muilenberg, G. (1979). Handbook of X-ray photoelectron spectroscopy: Perkin-Elmer.

Wang, L., Yi, Y., Zhao, Y., Zhang, R., Zhang, J., & Guo, H. (2015). NH3 Decomposition for H2 Generation: Effects of Cheap Metals and Supports on Plasma–Catalyst Synergy. ACS Catalysis, 5(7), 4167-4174.

Wang, L., Zhao, Y., Liu, C., Gong, W., & Guo, H. (2013). Plasma driven ammonia decomposition on a Fe-catalyst: eliminating surface nitrogen poisoning. Chemical Communications, 49(36), 3787-3789.

Wang, Q., Chen, P., Jia, C., Chen, M., & Li, B. (2011). Effects of air dielectric barrier discharge plasma treatment time on surface properties of PBO fiber. Applied Surface Science, 258(1), 513-520.

Wang, W., Patil, B., Heijkers, S., Hessel, V., & Bogaerts, A. (2017). Nitrogen

Page 153: Sustainable atmospheric ammonia synthesis and nitrogen

142

Fixation by Gliding Arc Plasma: Better Insight by Chemical Kinetics Modelling. ChemSusChem, 10(10), 2145-2157.

Wang, Y., Zhang, M.-l., Cao, F., Liu, Y.-t., & Shao, L. (2011). Interficial stability of Cu/Cu (Ru)/Si contact system for barrier-free copper metallization. Journal of alloys and compounds, 509(19), L180-L182.

Whittemore, R., & Knafl, K. (2005). The integrative review: updated methodology. Journal of advanced nursing, 52(5), 546-553.

Wu, S., Chen, J., Zheng, X., Zeng, H., Zheng, C., & Guan, N. (2003). Novel preparation of nanocrystalline magnesia-supported caesium-promoted ruthenium catalyst with high activity for ammonia synthesis. Chemical Communications(19), 2488-2489.

Xia, Y., & Mokaya, R. (2004). Ordered mesoporous MCM-41 silicon oxynitride solid base materials with high nitrogen content: synthesis, characterisation and catalytic evaluation. Journal of Materials Chemistry, 14(16), 2507-2515.

Xie, D., Sun, Y., Zhu, T., Fan, X., Hong, X., & Yang, W. (2016). Ammonia synthesis and by-product formation from H 2 O, H 2 and N 2 by dielectric barrier discharge combined with an Ru/Al 2 O 3 catalyst. RSC Advances, 6(107), 105338-105346.

Yan, B., Li, Y., & Zhou, B. (2009). Covalently bonding assembly and photophysical properties of luminescent molecular hybrids Eu–TTA–Si and Eu–TTASi–MCM-41 by modified thenoyltrifluoroacetone. Microporous and Mesoporous Materials, 120(3), 317-324.

Yan, Y., Miao, J., Yang, Z., Xiao, F.-X., Yang, H. B., Liu, B., & Yang, Y. (2015). Carbon nanotube catalysts: recent advances in synthesis, characterization and applications. Chemical Society Reviews, 44(10), 3295-3346.

Yang, J., Li, T., Zhong, C., Guan, X., & Hu, C. (2016). Nitrogen Fixation in Water Using Air Phase Gliding Arc Plasma. Journal of The Electrochemical Society, 163(10), E288-E292.

Zamfirescu, C., & Dincer, I. (2009). Ammonia as a green fuel and hydrogen source for vehicular applications. Fuel processing technology, 90(5), 729-737.

Zen, S., Abe, T., & Teramoto, Y. (2018). Indirect Synthesis System for Ammonia from Nitrogen and Water Using Nonthermal Plasma Under Ambient Conditions. Plasma Chemistry and Plasma Processing, 38(2), 347-354.

Page 154: Sustainable atmospheric ammonia synthesis and nitrogen

143

Zhang, C., Chen, J., & Wen, Z. (2012). Assessment of policy alternatives and key technologies for energy conservation and water pollution reduction in China’s synthetic ammonia industry. Journal of Cleaner Production, 25, 96-105.

Zhu, H., Gu, X., Yao, K., Gao, L., & Chen, J. (2009). Large-scale synthesis of MgCl2· 6NH3 as an ammonia storage material. Industrial & Engineering Chemistry Research, 48(11), 5317-5320.

Zhu, W., zhou, Y., Ma, W., Li, M., Yu, J., & Xie, K. (2013). Using silica fume as silica source for synthesizing spherical ordered mesoporous silica. Materials Letters, 92, 129-131. doi:http://dx.doi.org/10.1016/j.matlet.2012.10.044

Page 155: Sustainable atmospheric ammonia synthesis and nitrogen

144

Acknowledgements and references to reproduced/adapted journal publications

The author has published parts of this dissertation in the peer-reviewed journals listed below. Parts of the published contents are adapted/reproduced in this dissertation, in chapters included in parentheses, with permissions from the following publishers.

1. Peng, P., Chen, P., Schiappacasse, C., Zhou, N., Anderson, E., Chen, D., ... & Ruan, R. (2018). A review on the non-thermal plasma-assisted ammonia synthesis technologies. Journal of Cleaner Production, 177, 597-609. DOI: 10.1016/j.jclepro.2017.12.229. Reprinted/adapted in parts with permission from Elsevier. (Chapters 1, 2 and 7)

2. Reprinted/adapted by permission from Copy Right Clearance Center Inc. Springer Nature, Plasma Chemistry and Plasma Processing. Peng, P., Li, Y., Cheng, Y., Deng, S., Chen, P., & Ruan, R. (2016). Atmospheric pressure ammonia synthesis using non-thermal plasma assisted catalysis. Plasma Chemistry and Plasma Processing, 36(5), 1201-1210. DOI: 10.1007/s11090-016-9713-6. Copyright 2016. (Chapter 3, except for the contents listed in source 4)

3. Peng, P., Cheng, Y., Hatzenbeller, R., Addy, M., Zhou, N., Schiappacasse, C., ... & Ruan, R. (2017). Ru-based multifunctional mesoporous catalyst for low-pressure and non-thermal plasma synthesis of ammonia. International Journal of Hydrogen Energy, 42(30), 19056-19066. DOI: 10.1016/j.ijhydene.2017.06.118. Reprinted/adapted with permission from Hydrogen Energy Publications and Elsevier. (Chapter 4)

4. Reprinted/adapted in parts with permission from Peng, P., Chen, P., Addy, M., Cheng, Y., Anderson, E., Zhou, N., ... & Ruan, R. (2018). Atmospheric plasma-assisted ammonia synthesis enhanced via synergistic catalytic absorption. ACS Sustainable Chemistry & Engineering. In press. DOI: 10.1021/acssuschemeng.8b03887. Copyright (2018) American Chemical Society. (Figure 5, page 48 and Chapter 5)

5. Peng, P., Chen, P., Addy, M., Cheng, Y., Zhang, Y., Anderson, E., ... & Ruan, R. (2018). In situ plasma-assisted atmospheric nitrogen fixation using water and spray-type jet plasma. Chemical Communications, 54(23), 2886-2889. DOI: 10.1039/c8cc00697k. Reprinted/adapted with permission from The Royal Society of Chemistry. (Chapter 6)

Erratum In Chapter 5.4, page 93, the corresponding synthesis rate should be 50 µmol/min.