plasma dissociation of hydrogen sulfide

57
Plasma Dissociation of Hydrogen Sulfide A Thesis Submitted to the Faculty of Drexel University by Kirill A. Gutsol in partial fulfillment of the requirements for the degree of Master of Science in Mechanical Engineering September 2012

Upload: hadien

Post on 01-Jan-2017

255 views

Category:

Documents


3 download

TRANSCRIPT

Page 1: Plasma dissociation of hydrogen sulfide

Plasma Dissociation of Hydrogen Sulfide

A Thesis

Submitted to the Faculty

of

Drexel University

by

Kirill A. Gutsol

in partial fulfillment of the

requirements for the degree

of

Master of Science in Mechanical Engineering

September 2012

Page 2: Plasma dissociation of hydrogen sulfide

ii

ACKNOWLEDGMENTS

This work would not have been possible without the guidance of many people who

in one way or another contributed to the completion of my Master of Science degree.

First and foremost, I would like to express my deepest gratitude to my supervisor

and most knowledgeable teacher, Dr. Alexander Fridman for his guidance and support

throughout my studies. It was a great experience working with him and learning in the

process. I am grateful to Drs. Alexander Rabinovich and Andrey Starikovskiy for introducing

me and guiding me through the immense forest of experimental plasma science.

I would like to thank Chevron Energy Technology Company for continued support of

our project on Plasma Dissociation of Hydrogen Sulfide. I would like to acknowledge Drexel

University Machine Shop and Gary Nirenberg for help with manufacturing numerous parts

and for providing guidance.

With great pleasure I would like to acknowledge colleagues from Fuel Conversion

Lab of Drexel Plasma Institute, who worked with me in the lab: Michael Gallagher, Fela

Odeyemi, Robert Geiger, and Ryan Robinson. And special thanks to Thomas Nunnally, who

worked with me on Hydrogen Sulfide Dissociation project.

Next, I extend my gratitude to my family Natalia, Alexander, and Ksenia Gutsol for

their love, encouragement and invaluable support throughout my life. In addition, I want to

thank my mom for her endless patience and my dad for making science an exciting part of

my life.

Finally, I am grateful to my loving wife Krishna Priya Arjunan for being there for me

through successes and failures. This thesis would not have been possible without her help,

understanding, and support.

Page 3: Plasma dissociation of hydrogen sulfide

iii

TABLE OF CONTENTS

List of Tables ............................................................................................................................................................... v

List of Figures ........................................................................................................................................................... vi

Abstract ..................................................................................................................................................................... viii

Chapter 1: Introduction ......................................................................................................................................... 1

1.1 Background ......................................................................................................................................................... 1

1.2. Evaluation and comparison parameters................................................................................................ 3

1.2.1 Specific energy requirement (SER) ....................................................................................................... 3

1.2.2 Specific energy input (SEI) ....................................................................................................................... 4

1.2.3 SEI and SER term applied to plasma processes ............................................................................... 4

1.2.4 SEI and SER sensitivity ............................................................................................................................... 5

Chapter 2: Hydrogen sulfide dissociation modeling ................................................................................. 8

2.1 Motivation and terminology ........................................................................................................................ 8

2.1.1 Quenching ........................................................................................................................................................ 8

2.1.2 Absolute quenching ..................................................................................................................................... 9

2.1.3 Ideal quenching ............................................................................................................................................. 9

2.2 Results of the thermodynamic equilibrium modeling .................................................................... 10

2.2.1 Ideal vs. absolute quenching .................................................................................................................. 10

2.2.2 Effect of reduced pressure ...................................................................................................................... 12

2.3 Chemical kinetics of hydrogen sulfide dissociation ......................................................................... 14

2.3.1 Mechanism I .................................................................................................................................................. 14

2.3.2 Mechanism II ................................................................................................................................................ 17

2.4 Modeling summary ........................................................................................................................................ 18

Chapter 3: Experimental model verification .............................................................................................. 20

3.1 Motivation ......................................................................................................................................................... 20

3.2 Non-thermal plasma discharges .............................................................................................................. 20

Page 4: Plasma dissociation of hydrogen sulfide

iv

3.3 Materials and methods ................................................................................................................................. 21

3.3.1 Measurements .............................................................................................................................................. 22

3.3.2 Plasma discharge configuration ........................................................................................................... 24

3.4 Results ................................................................................................................................................................. 26

3.4.1 AC Corona Discharge ................................................................................................................................. 27

3.4.2 Dielectric Barrier Discharge ................................................................................................................... 30

3.4.3 Streamer Discharge ................................................................................................................................... 32

3.4.4 Contracted Normal Glow Discharge .................................................................................................... 34

3.5 Discussion .......................................................................................................................................................... 38

3.5.1 Corona Discharge ........................................................................................................................................ 38

3.5.2 Dielectric Barrier Discharge ................................................................................................................... 39

3.5.3 Streamer Discharge ................................................................................................................................... 39

3.5.4 Contracted Normal Glow Discharge .................................................................................................... 41

3.5.5 Summary of the Results ........................................................................................................................... 43

Chapter 4: Conclusions ........................................................................................................................................ 45

List of References ................................................................................................................................................... 47

Page 5: Plasma dissociation of hydrogen sulfide

v

LIST OF TABLES

Table 1: Mechanism I for chemical kinetic modeling of H2S dissociation. ..................................... 15

Table 2: The reactions amended to Mechanism I to form Mechanism II. ....................................... 18

Table 3: Different mechanisms for electronic impact initiated dissociation of hydrogen sulfide. ........................................................................................................................................................................ 40

Page 6: Plasma dissociation of hydrogen sulfide

vi

LIST OF FIGURES

Figure 1: SEI and SER sensitivity example. ................................................................................................... 5

Figure 2: Results of the modeling at atmospheric pressure, comparing ideal and absolute quenching concepts. ............................................................................................................................................. 11

Figure 3: Results of the modeling at atmospheric pressure, comparing SER for ideal and absolute quenching concepts. ........................................................................................................................... 12

Figure 4: Final hydrogen mole fraction, α as function of temperature and pressure (under assumption of ideal quenching). ..................................................................................................................... 13

Figure 5: SER as function of SEI and pressure (under assumption of ideal quenching). ......... 14

Figure 6: Comparison of SER of the kinetic (Mechanism I) and thermodynamic modeling. .. 16

Figure 7: Residence time required for kinetics model to reach equilibrium as function of reactor temperature. ............................................................................................................................................ 17

Figure 8: Simplified schematic of the experimental setup (NV –needle valve, BV – ball valve, PG – pressure gauge, and 3WV – 3-way ball valve). ................................................................................ 22

Figure 9: Schematic of the corona reactor (1 – power supply, 2 – silicone stopper, 3 – SS inlet tube, 4 – grounded furnace walls, 5 – molybdenum wire, 6 – SS spring and hook, and 7 – quartz tube). ......................................................................................................................................................... 24

Figure 10: Schematic of the DBD reactor (1 – power supply, 2 – silicone stopper, 3 – SS inlet tube, 4 – furnace walls, 5 – molybdenum wire, 6 – SS spring and hook, 7 – quartz tube, and 8 – DBD ground). ....................................................................................................................................................... 25

Figure 11: Schematic of the streamer and contracted normal glow discharge reactors (1 – power supply, 2 – silicone stopper, 3 – electrodes, 4 – quartz tube, and 5 – Teflon® nozzle). ....................................................................................................................................................................................... 26

Figure 12: Comparison of measured hydrogen mole fraction in experiment and in thermodynamic equilibrium modeling as function of temperature. ................................................ 28

Figure 13: Full and relative SER of hydrogen production (b) as a function of gas temperature. The modeling was performed for the pressure of 0.25 atm under assumption of absolute quenching. ......................................................................................................................................... 29

Page 7: Plasma dissociation of hydrogen sulfide

vii

Figure 14: Hydrogen production in DBD as a function of gas temperature and comparison with results of thermodynamic equilibrium modeling under the assumption of absolute quenching. The modeling was performed for the pressure of 0.25 atm. ........................................ 31

Figure 15: Average power of the streamer discharge as a function of the discharge frequency. ................................................................................................................................................................. 33

Figure 16: Measured hydrogen concentration in the streamer discharge as a function of discharge frequency. ............................................................................................................................................ 34

Figure 17: V-I characteristic of the contracted glow discharge. ......................................................... 35

Figure 18: Hydrogen concentration as function of discharge power in the contracted glow discharge reactor. .................................................................................................................................................. 36

Figure 19: Contracted normal glow discharge reactor in operation (less then 1 min after start). ........................................................................................................................................................................... 37

Figure 20: Contracted normal glow discharge reactor in operation (15 min after start). ...... 37

Figure 21: SER as function of SEI for all the discharge configurations (including MW [3, 4, 30] and GAT [6]) and thermodynamic equilibrium model under ideal quenching assumption. .............................................................................................................................................................. 43

Page 8: Plasma dissociation of hydrogen sulfide

viii

ABSTRACT

Plasma Dissociation of Hydrogen Sulfide

Kirill A. Gutsol

Supervisor: Alexander Fridman, Ph.D.

Plasma dissociation of hydrogen sulfide has been studied by kinetic and

thermodynamic modeling and experimentally in four different discharges: AC corona,

dielectric barrier, streamer, and contracted glow discharge. The performance of corona

discharge and DBD was studied in the initial gas temperature range of 300 – 1200 K. A

robust reactor was created to perform the above studies in pure H2S with high degree of

reproducibly. A specific energy requirement (SER) was calculated as function of energy

input for each type of the discharge and compared with earlier experimental results and

modeling. The results showed that discharges at high 𝐸/𝑛 and low specific energy release

(corona, dielectric barrier, and streamer) perform significantly worse than contracted glow

discharge at low 𝐸/𝑛 where specific energy release was high and gas temperature was

elevated. The SER for non-thermal dissociation was 12 – 14 eV/molec and went down to 2.4

eV/molec in the case of the glow discharge with elevated temperature. Experiments

confirmed that an electrical discharge in H2S allows obtain SER close to the predicted by

thermodynamic equilibrium modeling if the discharge has gas temperature that is high

enough. Further reduction of SER in a plug flow reactor is not possible because of the

absence of chemical chain processes and energy recovery mechanisms.

Page 9: Plasma dissociation of hydrogen sulfide

1

CHAPTER 1: INTRODUCTION

1.1 Background

Hydrogen sulfide (H2S) processing is a necessity for the oil refining and gas

industries. Currently H2S, which is a byproduct of hydro-desulfurization and often

comprises a significant portion of natural gas deposits, is treated via the Claus process. A

gross reaction of the Claus process is partial oxidation of hydrogen sulfide to sulfur and

water. This process, even though well established, has the major drawback of oxidizing

valuable hydrogen, which could be used for hydro-desulfurization and, of course, for many

other applications including “green” energy generation. Hydrogen sulfide can be a good

source for hydrogen production since its dissociation enthalpy to hydrogen and solid sulfur

is only 0.2 eV/molec (compare with the dissociation enthalpy of liquid water that is 2.96

eV/molec). Unfortunately, a process that could accomplish the dissociation and is

economically feasible has not yet been implemented in a large scale.

Much research work has been done on hydrogen sulfide dissociation. Plasma

dissociation became very popular after the experiments in Kurchatov Institute in Soviet

Union demonstrated that dissociation could be achieved with the Specific Energy

Requirement (SER) less than 1 eV/molec. Review papers cover the research done in this

field in great detail [1, 2], and they discuss different methods for H2S dissociation including

plasmas.

Researchers from Kurchatov Institute indicated that their very promising results

were obtained under the conditions of thermal dissociation with additional energy recovery

inside the reactor due to the opposite flows of condensing sulfur clusters and incoming

hydrogen sulfide in a strong centrifugal field [3]. Another possible explanation they

Page 10: Plasma dissociation of hydrogen sulfide

2

provided is based on the dissociation process equilibrium shift due to the high diffusion

coefficient of hydrogen [4]. Thermodynamic equilibrium modeling shows that, at

atmospheric pressure dissociation of H2S to gaseous products (H2, HS, HS2, S2, …) has a

thermodynamic minimum of SER of 2.1 eV per molecule of H2 produced (assuming absolute

quenching), and 2.05 eV per H2 molecule produced (assuming ideal quenching when SH and

H radicals are also converted to H2 molecules). This 0-dimentional time-independent model

is a good first approximation of a plug flow reactor at high temperature, where all reactions

are fast enough to establish equilibrium, while all the reverse reactions are stopped by

quenching.

The best experimentally obtained SER is 0.76 eV/molec, which was reported for a

low pressure microwave plasma reactor [5]. Recently reported results of experiments in

Gliding Arc Tornado (GAT) reactor and Gliding Arc Plasmatron (GAP) at atmospheric

pressure show SER of 1.2 eV/molec [6]. As these results are significantly better than those

of thermodynamic equilibrium modeling, a possibility exists that the process is defined by

some non-equilibrium effects (for example on periphery of thermal plasma discharge) that

were not taken into consideration.

This uncertainty was the primary motivation for our research because it is

necessary to understand all the steps of the process to be able to design large (industrial)

scale installations. This work investigates the H2S dissociation process in different plasmas

and compares the obtained results to each other as well as to those of other researchers.

The experiments were designed so that as many parameters remained constant as possible

in between the different discharges such as reactor geometry, measurement and control

instrumentation.

Page 11: Plasma dissociation of hydrogen sulfide

3

Another motivation is that many researchers have performed experiments with

hydrogen sulfide in a number of different discharges, but almost all of these experiments

were done in mixtures where hydrogen sulfide was significantly diluted by a carrier gas.

The dilution of hydrogen sulfide by a carrier gas can make radical changes in the discharge

properties, which could, quite possibly, change the results of an otherwise very well

prepared and executed experiment.

1.2. Evaluation and comparison parameters

In the following sections many experimental and computational results will be

evaluated in terms of the specific energy requirement (SER), which is a measure of

effectiveness of a (plasma) system in production or destruction of a specific chemical

substance or a group of substances. Specifically most of the processes and systems

described in this article are evaluated by their SER of hydrogen production or hydrogen

sulfide dissociation.

1.2.1 Specific energy requirement (SER)

Specific energy requirement is an effective tool for the comparison of the systems,

which have very different underlying principles. Using SER, it is possible to compare

systems, which use thermal or non-thermal plasma, and even systems that do not use

plasma at all.

The common units used in the literature for SER are kilojoule per mole (kJ/mol),

kilowatt hour per normal cubic meter (kWh/Nm3), and electron volt per molecule

(eV/molec). The conversion factor is such, that

1𝑘𝑊ℎ𝑚3 = 0.84

𝑒𝑉𝑚𝑜𝑙𝑒𝑐

= 80.64𝑘𝐽𝑚𝑜𝑙

. (1)

Page 12: Plasma dissociation of hydrogen sulfide

4

The eV/molec unit is used exclusively throughout this thesis.

SER is not measured directly but calculated from the measured or computed

parameters. In general,

ε = 𝑊/𝑄𝑖 (2)

where 𝜀 is SER, W is total power input, and 𝑄𝑖 is a flow rate of the substance of interest in

the product mixture. When SER of destruction of a substance is sought then 𝑄𝑖 is the

difference of initial and final flow rate of the substance.

1.2.2 Specific energy input (SEI)

Another important parameter for comparison of different systems is specific energy

input (SEI), J, which is a measure of total power input to the source stream

𝐽 =𝑊𝑄0

(3)

and 𝑄0 is initial substance flow rate. In this research, we are focusing on hydrogen sulfide

dissociation, and thus, unless stated otherwise, SEI is a measure of energy input per

molecule of H2S, and it is commonly measured using the same units as SER. Calculation of

SEI is different for each system and model and will be explained where necessary.

1.2.3 SEI and SER term applied to plasma processes

It is obvious that SER and SEI are very important parameters from the application

standpoint; however SER has also very important physicochemical sense. Though SER is a

“macro”-parameter, usually it provides very significant information about the mechanism of

a plasma chemical process, especially if kinetic models are available and there is some

understanding about the plasma system applied (cold, thermal, or intermediate – “warm”).

For example, if SER in cold plasma is comparable with the energy necessary to produce one

electron-ion pair (more than 10 eV/molec) it means usually that the plasma chemical

Page 13: Plasma dissociation of hydrogen sulfide

5

process is controlled by a reaction that involves recombination of charged particles. If SER

in warm or thermal plasma is close to that predicted by thermodynamic equilibrium

modeling, it usually means that the process is controlled by temperature alone and there is

an effective quenching mechanism that suppresses the reverse reactions. If SER in cold

plasma is close to the reaction enthalpy, it means that some very special non-equilibrium

conditions are realized, like selective vibrational excitation (e.g. CO2 dissociation in

microwave plasma at low pressure [7]) or a chain process that allows multiple use of a

single charged or electronically excited particle for the chemical process propagation (for

example see the idea of non-equilibrium water dissociation [7]).

1.2.4 SEI and SER sensitivity

It is important, that SER is not very sensitive to system non-uniformity. To illustrate

this, let’s consider a rather extreme case. Let us assume that we have a gas (e.g. H2S) flow

through a very non-uniform discharge system (Figure 1).

Figure 1: SEI and SER sensitivity example.

In this system 16 of the flow, 𝑄, is passing throw a Zone #1 with optimal conditions

where SER has the lowest value, 𝜀1, and conversion degree is 𝛼1. In this case the power

released in the Zone #1 is

Page 14: Plasma dissociation of hydrogen sulfide

6

𝑊1 =16𝑄𝛼1𝜀1 (4)

Next, let us assume that another 16 of the flow 𝑄 is passing through Zone #2 with the

same power as Zone #1, but unfavorable conditions where SER has rather high value,

𝜀2 = 3𝜀1. Lastly let us assume that 23 of the flow 𝑄 is passing throw Zone #3, where there is

no power release, and therefore, there is no conversion (𝛼3 = 0). It is easy to see that in this

extreme case 𝛼2 = 13𝛼1, and total flow rate of the product (H2 for H2S dissociation with

sulfur condensation) is

𝑄𝑝 =16𝑄𝛼1 +

16𝑄𝛼2 =

29𝑄𝛼1, (5)

the total power is

𝑊 = 𝑊1 + 𝑊2 = 2𝑊1 =13𝑄𝛼1𝜀1. (6)

Also, in this case, the average SER is

ε =𝑊𝑄𝑝

=32𝜀1, (7)

which is rather close to the lowest value and is absolutely insensitive to the flow through

the plasma-free zone. SEI, on the other hand, is very sensitive to the flow through the

plasma-free zone. Thus, for the example above,

𝐽1 = 𝛼1𝜀1 = 𝐽2; 𝐽3 = 0, (8)

and the average SEI is

𝐽 =𝑊𝑄

=13𝛼1𝜀1. (9)

Page 15: Plasma dissociation of hydrogen sulfide

7

Though SEI has much less sense for process characterization than SER, presentation

of SER as a function of SEI is very convenient and will be used in this thesis because it

provides information about the conversion degree α = 𝐽/ε. Also, it is rather obvious, that

SEI has significant meaning in 0-D and 1-D kinetic and thermodynamic simulations. For

example, in these cases, process temperature is completely defined by SEI.

Page 16: Plasma dissociation of hydrogen sulfide

8

CHAPTER 2: HYDROGEN SULFIDE DISSOCIATION MODELING

2.1 Motivation and terminology

A number of research group have attempted research on plasma dissociation of

hydrogen sulfide. Usually each group is focusing on one specific plasma discharge with a

relatively fixed set of parameters. In our research we attempt to breach this limitation, and

we using modeling as starting point that will allow us to obtain a solid hypothesis for future

experimental studies.

2.1.1 Quenching

In the process of modeling we compute species concentration at certain

temperature, which is not the same that can be effectively reached in experimental and

production conditions. Most of the time the end products are collected at ambient

temperature, which in the lab environment corresponds to 300 K. So, we know

concentrations at elevated temperature (suitable for high degree of dissociation or low

SER), but we also need know how the drop in temperature to ambient will affect the

calculated concentrations. Thus, the resulting concentration depends on the cooling or

quenching speed of the process.

Let’s take hydrogen sulfide dissociation as an example. The dissociation causes H2S

to split into a number of species, which may include some of the following or even more

species depending on the temperature: H, H2, HS, H2S, S, and S2. Here S and S2 is a gaseous

because H2S dissociates at temperatures above 750K, and at these temperatures sulfur is

already evaporated and found in the forms from S2 to S8 (with S8 being present at lower

temperatures). At the room temperature conditions (300 K) only the following species

Page 17: Plasma dissociation of hydrogen sulfide

9

remain: H2, H2S, and S (solid sulfur). In the worst case scenario the quenching is so slow that

all the products recombine back into H2S.

Let’s say the main product of interest is hydrogen, and let’s say final hydrogen

concentration is [H2]f. Without performing detailed quenching kinetics modeling we can still

make meaningful estimations for the value of [H2]f, using absolute and ideal quenching

concepts [8].

2.1.2 Absolute quenching

One of the ways to estimate [H2]f is using the concept of absolute quenching.

Absolute quenching concept assumes that all stable products are preserved during

quenching, and the rest of the radicals convert to the initial substance. So, in our case:

𝑀H2𝑓 = 𝑀H2 , 𝑀SS

𝑓 = 2𝑀S2 (10)

where 𝑀𝑋 is the number of moles of specie X. Now taking to account that SS is solid and this

doesn’t contribute to gaseous concentration of the products then

[H2S]𝑓 = 1 − [H2]𝑓 . (11)

2.1.3 Ideal quenching

The concept of ideal quenching assumes that all the radicals are converted into

products rather than initial substance. So, in the case of hydrogen sulfide dissociation:

𝑀H2𝑓 = 𝑀H2 +

12𝑀H +

12𝑀HS, 𝑀SS

𝑓 = 2𝑀S2 + 𝑀S + 𝑀HS, 𝑀H2S𝑓 = 𝑀H2S (12)

The final concentration of hydrogen ([H2]𝑓) is calculated, similar to hydrogen sulfide final

concentration the case of absolute quenching, using the following equation:

Page 18: Plasma dissociation of hydrogen sulfide

10

[H2]𝑓 = 1 − [H2S]𝑓 . (13)

2.2 Results of the thermodynamic equilibrium modeling

All thermodynamic modeling was performed in CHEMKIN® 4.1.1 software, which

uses STANJAN libraries to determine equilibrium mixture composition. In general, the

method for determining equilibrium is based on minimizing Gibbs free energy.

The enthalpy of the mixture and mixture composition are calculated for each set

temperature in the model. This data is used to calculate corresponding SEI and SER of

species:

𝐽 = 𝐻𝑇 − 𝐻0 (14)

𝜀 =𝐽𝛼

(15)

where 𝐻𝑇 is enthalpy of the mixture at temperature 𝑇, 𝐻0 is the enthalpy of the H2S at room

temperature and α is the conversion degree.

2.2.1 Ideal vs. absolute quenching

Earlier given definitions of absolute and ideal quenching can be very well illustrated

with thermodynamic model of hydrogen sulfide dissociation. The results produced by such

modeling are shown in Figure 2 and Figure 3. Note the difference in both hydrogen

concentrations and SER of two different quenching mechanisms.

Page 19: Plasma dissociation of hydrogen sulfide

11

Figure 2: Results of the modeling at atmospheric pressure, comparing ideal and absolute quenching

concepts.

The Figure 3 clearly shows that the process following ideal quenching has lower SER

(2.03 eV/molec vs. 2.07 eV/molec), and therefore, it is of greater interest. Moreover, ideal

quenching result becomes even better at higher values of SEI. So, the following modeling

results are provided for ideal quenching conditions and following experimental sections

compare results with ideal quenching modeling results.

0.00

0.10

0.20

0.30

0.40

0.50

0.60

0.70

0.80

0.90

1.00

750 1250 1750 2250 2750

H 2 Mol

e Fr

actio

n, α

Temperature, T (K)

Absolute

Ideal

Page 20: Plasma dissociation of hydrogen sulfide

12

Figure 3: Results of the modeling at atmospheric pressure, comparing SER for ideal and absolute

quenching concepts.

2.2.2 Effect of reduced pressure

In agreement with Le Chatelier's principle, the reduction in pressure causes greater

dissociation and respectively lowers specific energy requirements. The results of the

modeling are show in Figure 4 and Figure 5.

0.00

0.50

1.00

1.50

2.00

2.50

3.00

3.50

4.00

4.50

5.00

0.00 0.50 1.00 1.50 2.00 2.50 3.00

SER,

ε(e

V/m

olec

)

SEI, J (eV/molec)

Absolute

Ideal

Page 21: Plasma dissociation of hydrogen sulfide

13

Figure 4: Final hydrogen mole fraction, α as function of temperature and pressure (under assumption

of ideal quenching).

We can observe that thermodynamic equilibrium modeling even in ideal quenching

assumptions cannot explain the best results of experiments with microwave, radio

frequency, and gliding arc plasma discharges. Even at pressures of 0.001 atm (0.76 Torr)

the best modeling results for SER are 1.46 eV/molec. This observation leads to conclusion

that the earlier mentioned plasma systems must have noticeable degree of non-equilibrium,

which helps to reduce SER. If the system is non-equilibrium then it may be possible to

observe it in chemical kinetics modeling, which we discuss in greater detail in the following

sections.

0.00

0.10

0.20

0.30

0.40

0.50

0.60

0.70

0.80

0.90

1.00

750 1250 1750 2250 2750

H 2 Mol

e fr

actio

n, α

Temperature, T (K)

1.0 atm

0.1 atm

0.01 atm

0.001 atm

Page 22: Plasma dissociation of hydrogen sulfide

14

Figure 5: SER as function of SEI and pressure (under assumption of ideal quenching).

2.3 Chemical kinetics of hydrogen sulfide dissociation

Chemical kinetics modeling was performed using CHEMKIN ® 4.1.1 Software, using

plug-flow reactor approximation model. For each set of calculations the following

parameters were kept constant: reactor cross-section area, volumetric flow rate (1 L/min),

and pressure (1 atm). Dependence of dissociation effectiveness on temperature was

investigated by changing constant temperature of isothermal reactor in the increments of

100 K. SEI and SER of the model was calculated in the same way as it was done for

thermodynamic equilibrium models (see equations (14 and (15)

2.3.1 Mechanism I

The first kinetic mechanism that we considered was introduced by researcher from

the Kurchatov Institute in Moscow [3, 5, 9]. The complete list of reaction of this mechanism

0.00

0.50

1.00

1.50

2.00

2.50

3.00

3.50

4.00

4.50

5.00

0.00 0.50 1.00 1.50 2.00 2.50 3.00

SER,

ε (e

V/m

olec

)

SEI, J (eV/molec)

1.0 atm

0.1 atm

0.01 atm

0.001 atm

Page 23: Plasma dissociation of hydrogen sulfide

15

is presented in Table 1. The mechanism is characterized by high activation energy of the

two initiating reactions (1 and 2 in Table 1). As a results despite of the fact that some small

dissociation is possible thermodynamically under 1000 K, but due the kinetic limitations

the process is very slow in this temperature range and thus, the same concentration of

hydrogen cannot be obtained keeping residence time within several seconds (technological

limitations for flow systems).

Table 1: Mechanism I for chemical kinetic modeling of H2S dissociation.

# Reaction 𝐴,𝑐𝑚3

𝑚𝑜𝑙𝑒𝑐 ⋅ 𝑠 n 𝐸𝑎 ,

𝑘𝑐𝑎𝑙𝑚𝑜𝑙𝑒

1 H2S + M ↔ SH + H + M 2.92E-08 0.00 66.21 2 H2S ↔ H2 + S 3.16E-10 0.00 65.49 3 H2S + H ↔ H2 + SH 2.31E-07 1.94 0.90 4 SH + S ↔ H + S2 4.00E-11 0.00 0.00 5 SH + H ↔ H2 + S 3.01E-11 0.00 0.00 6 SH + SH ↔ H2 + S2 1.00E-14 0.00 0.00 7 SH + SH ↔ H2S + S 1.50E-11 0.00 0.00 8 S2 + M ↔ S + S + M 7.95E-11 0.00 76.96

10 H2 ↔ H + H 3.70E-10 0.00 96.08

The comparison of this kinetic model results with those of thermodynamic

equilibrium are shown in Figure 6. We can observe that kinetic modeling curve matches

thermodynamic equilibrium, but this is only true if either make the flow extremely slow or

reactor length extremely long. The time that it takes for kinetic system to reach equilibrium

is shown as function of reactor temperature in Figure 7.

Page 24: Plasma dissociation of hydrogen sulfide

16

Figure 6: Comparison of SER of the kinetic (Mechanism I) and thermodynamic modeling.

In general, industry acceptable time is no more than a few seconds. This time limit

comes from huge flow rates of gas that needs to be processed yet maintain reasonable size

of the installation. From the Figure 7 we can see that this time corresponds isothermic

reactor temperature of 1200 - 1300 K.

0.00

0.50

1.00

1.50

2.00

2.50

3.00

3.50

4.00

4.50

5.00

0 0.5 1 1.5 2 2.5 3

SER,

ε (e

V/m

olec

)

SEI, J (eV/molec)

Thermodynamic

Kinetic

Page 25: Plasma dissociation of hydrogen sulfide

17

Figure 7: Residence time required for kinetics model to reach equilibrium as function of reactor temperature.

2.3.2 Mechanism II

The later research by Binoist, et al. introduced and experimentally confirmed

mechanism that included additional species, H2S2 and HSS. These species are radicals and

not found in products, but they provide additional pathway for dissociation. The Mechanism

I was amended by the reactions involving H2S2 and HSS. Reactions amended to Mechanism I

are shown in Table 2.

1.0E-06

1.0E-05

1.0E-04

1.0E-03

1.0E-02

1.0E-01

1.0E+00

1.0E+01

1.0E+02

1.0E+03

1.0E+04

900 1400 1900 2400 2900

Tim

e, t

(s)

Temperature, T (K)

Feas

ible

tim

e fr

ame

Page 26: Plasma dissociation of hydrogen sulfide

18

Table 2: The reactions amended to Mechanism I to form Mechanism II.

# Reaction 𝐴,cm3

molec ⋅ s n 𝐸𝑎 ,

kcalmole

1 SH + H2S ↔ H2S2 + H 3.32E-10 0.50 27.00 2 H2S2 + M ↔ SH + SH + M 3.43E-07 1.00 57.12 3 HSS + HSS ↔ H2S2 + S2 3.46E-15 2.37 -1.67 4 HS + HSS ↔ H2S + S2 3.66E-13 3.05 -1.10 5 H + HSS ↔ S + H2S 7.32E-11 0.00 6.32 6 H + HSS ↔ H2 + S2 2.51E-12 1.65 -1.10 7 S + HSS ↔ HS + S2 2.00E-02 2.20 -0.60

Addition of the seven reactions from Table 2 does not change the outcome of the

calculated SER values. Still, close inspection of time resolved solutions and their comparison

with thermodynamic modeling of the mixture including H2S2 and HSS showed that there is

significant accumulation (~2%) of H2S2 at lower temperatures 900-1200 K.

Literature research into properties of H2S2 revealed that H2S2 belongs to a group

sulfur compounds known as sulfanes. So, H2S2 is not a radical it is a stable compound that

can be present as a product even at room temperature [10].

2.4 Modeling summary

Thermodynamic equilibrium modeling of hydrogen sulfide dissociation showed that

• Earlier reported minimum for SER of hydrogen production under

assumption of ideal quenching (2.03 eV/molec) remains valid after adding

new species (H2S2 and HSS) to list.

• The thermodynamic SER minimum is achieved at the temperature of 1875 K

at atmospheric pressure. The respective temperatures of SER minimum for

the pressures of 0.1 atm, 0.01 atm, and 0.001 atm are: 1625 K, 1425 K, and

1250 K.

Page 27: Plasma dissociation of hydrogen sulfide

19

• There is rather strong dependance of SER minimum on pressure, which

favors low pressure plasma system like microwave system used in

Kurchatove institute [5].

Chemical kinetic modeling of hydrogen sulfide dissociation sowed that

• Given sufficent time, both kinetic mechanism I and II produce SER of

hydrogen sulfide dissociation the same as the thermodynamic model.

• The time required for kinetic models to reach equilibrium is an exponential

function of reactor temperature. And even though, at SER favorable

temperatures of 1700 K or higher the equilibration time is less than 0.01 s,

the time is too long for any temperature lower than 1300 K.

In general, modeling shows that high-temperature low-pressure conditions are

favorable for efficient H2S dissociation. Nonetheless, none of the models (even with ideal

quenching assumption) were able to match the efficiency reported from the experiments

with gliding arc, microwave, and radio frequency plasma.

Page 28: Plasma dissociation of hydrogen sulfide

20

CHAPTER 3: EXPERIMENTAL MODEL VERIFICATION

3.1 Motivation

Earlier described modeling of hydrogen sulfide dissociation allowed us to formulate

a hypothesis for experimental part of our study. The hypothesis can be stated as follows:

there exists a plasma assisted mechanism of low temperature hydrogen sulfide dissociation, in

which non-equilibrium dissociation takes place. This allows for high concentration of

hydrogen to be produced at the temperatures as low as 800-1200 K. As a result of low energy

input and high conversion, the specific energy requirement of the dissociation process is

significantly lower than SER of equilibrium processes with ideal quenching (2.1 eV/molec).

The hypothesis clearly indicates that experiments should be performed at elevated

temperatures, but these temperatures should not be sufficient for thermal decomposition of

H2S. Thus, there is additional limitation for our experimental study – we must use non-

thermal plasma discharges to catalyze the dissociation.

3.2 Non-thermal plasma discharges

All plasma discharges are commonly divided into thermal and non-thermal. The

former ones are also known as quasi-equilibrium, and the later are non-equilibrium. The

primary distinction in between these two groups of discharges is that unlike thermal

plasma discharges the non-thermal plasmas have significant differences in translational,

rotational, vibrational, and electronic temperatures. The detailed discussion on all the non-

thermal plasmas is outside of the scope of this work, but it is covered in a number of books

on plasma discharge physics, chemistry, and applications [8, 11, 12]. The non-thermal

Page 29: Plasma dissociation of hydrogen sulfide

21

discharges that we chose for the experimental portion of this study are corona, dielectric

barrier (DBD), streamer, and glow.

3.3 Materials and methods

Throughout all the experiments, a common reactor design was used to minimize

potential errors and uncertainties related to changes in geometry of the reactor and

instrumentation. The body of the reactor was made out of a quartz tube with length of 0.6

or 1.2 m, internal diameter of 47 mm, and wall thickness of 1.5 mm. Reactor ends were

closed with high purity silicone stoppers, which are heat resistant up to 250 °C. All the

metal parts and connections, unless especially mentioned, were made out of type 316

stainless steel (SS). Gas transport outside of the high temperature zone was provided by

clear PVC tubing with 3.2 mm (inlet) and 9.5 mm (exhaust) internal diameter. Exhaust

mixture was filtered from sulfur via 10 μm polyester filter and scrubbed off hydrogen

sulfide via 3 L washing bottle of 20% water solution of sodium hydroxide (NaOH). A double-

stage diaphragm vacuum pump was used to maintain low pressure (usually below 200

Torr) in the discharge zone. Gas flow rate was measured and controlled with rotameters

(gas pressure in the flow rate measurement point was close to 1 atm).

There was no carrier gas used, and all experiments were performed with 100%

research grade hydrogen sulfide. Ultra-high purity nitrogen was used to purge H2S out of

the reactor at the end of each experiment. A simplified schematic of the installation is show

in Figure 8.

Page 30: Plasma dissociation of hydrogen sulfide

22

Figure 8: Simplified schematic of the experimental setup (NV –needle valve, BV – ball valve, PG –

pressure gauge, and 3WV – 3-way ball valve).

3.3.1 Measurements

Exhaust gas mixture was sampled with vacuum tight disposable 60 mL syringe. The

samples were taken from a low pressure stage of the experimental system, and

consequentially, the size of the sample at normal conditions was 6 – 15 mL depending on

the system pressure. The samples were analyzed with a gas chromatograph (Agilent Micro

GC 3000). The mole fractions of H2, H2S, N2, and O2 were measured. Measures were taken to

ensure that small amount of nitrogen and oxygen were the result of a sampling procedure

only and were not a part of the exhaust gas. This parasitic nitrogen and oxygen were always

present in the same relative concentrations as in ambient air and the overall molar fraction

did not exceed 4%.

The mole fractions of H2S and H2 were normalized to 100%. This is acceptable under

assumption that the only gaseous product that is produced in H2S dissociation is hydrogen

or, in other words, if

Page 31: Plasma dissociation of hydrogen sulfide

23

H2S → αH2 + (1 − α)H2S + Ssolid (16)

where α is molar fraction of H2 in the product stream that is equal to the conversion degree

for this particular process. In this case the SER of H2 production and SER of H2S dissociation

are the same.

Plasma power input was calculated from measurements of voltage and current

obtained from an oscilloscope (Tektronics TDS5052B), using a high voltage probe and

either probe or shunt for current measurement depending on the type of the discharge.

Voltage-current product was integrated over time to calculate average power.

The flow rate of the hydrogen sulfide before the reactor was measured using

rotameters with 2% full scale accuracy. The rotameters were calibrated for air and the

conversion factor 𝑓𝑚 was used to adjust for H2S:

𝑄H2S = 𝑓𝑚𝑄𝐴𝑖𝑟 (17)

𝑓𝑚 = �𝑀Air

𝑀H2S (18)

where M is molecular weight. The use of such a conversion factor makes flow rate

measurement precision rather low with error on the level of 10%. In the general case,

additional corrections for pressure and temperature would have to be made; however, the

inlet gas pressure was maintained at 1 atm and temperature at 25 °C throughout the

experiments, thus eliminating the need for such correction.

SEI was calculated from H2S flow rate 𝑄, set at the inlet of the reactor, and from

voltage and current measurements as follows:

𝐽 =∫ 𝑈𝐼d𝑡𝜏0𝑄τ

(19)

Page 32: Plasma dissociation of hydrogen sulfide

24

where J is specific energy input, U and I are voltage and current respectively, and τ is a

length of a single period for high frequency corona and DBD discharges. In the case of a DC

discharge the above formula reduces to 𝐽 = 𝑈𝐼/𝑄. And, SER of hydrogen production is

ε =𝐽α

(20)

where α is the conversion degree (equal to the mole fraction of hydrogen in the exhaust

mixture after cooling and sulfur separation).

3.3.2 Plasma discharge configuration

Four different types of discharges were used in the experiments: AC corona,

dielectric barrier discharge (DBD), streamer discharge, and contracted normal glow

discharge.

The AC corona discharge was arranged along the reactor axis with 0.25 mm

molybdenum wire serving as a high voltage electrode and furnace casing serving as a

ground. A custom-built high-voltage high-frequency AC power supply was used. The

schematic of the corona reactor is shown in Figure 9.

Figure 9: Schematic of the corona reactor (1 – power supply, 2 – silicone stopper, 3 – SS inlet tube, 4 –

grounded furnace walls, 5 – molybdenum wire, 6 – SS spring and hook, and 7 – quartz tube).

Page 33: Plasma dissociation of hydrogen sulfide

25

The DBD was arranged similar to the corona but with the ground electrode made

out of a thin (~0.5 𝑚𝑚) aluminum 20 mm wide stripe wrapped tightly around the quartz

tube. Two series of experiments were performed, where DBD was placed before and after

the furnace. The power supply that was used in AC corona discharge experiments was used

for DBD experiments as well. The schematic of the DBD reactor is shown in Figure 10.The

length of the furnace and flow rate of the H2S was chosen in corona and DBD experiments so

that the gas would heat up to the set furnace temperature. The estimation showed that the

expected residence time within the furnace can be up to 1 minute, which would guarantee

gas preheating.

Figure 10: Schematic of the DBD reactor (1 – power supply, 2 – silicone stopper, 3 – SS inlet tube, 4 – furnace walls, 5 – molybdenum wire, 6 – SS spring and hook, 7 – quartz tube, and 8 – DBD ground).

Both streamer and contracted normal glow discharges were organized between two

pointed electrodes at 20 cm distance from the inlet. The furnace was not used for the

streamer discharge and contracted normal glow discharge experiments. A high voltage

nanosecond power supply (FID FPG 20-01NK10) was used as a power source for the

streamer discharge.

The contracted normal glow discharge geometry was similar to the streamer

geometry with the addition of a nozzle located 2 cm upstream from the tips of the

Page 34: Plasma dissociation of hydrogen sulfide

26

electrodes. The nozzle concentrated the H2S flow and provided better electrode cooling. The

glow discharge was powered by a high-voltage DC power supply (Universal Voltronics BRC

10,000). The schematic of streamer and glow discharge reactors is shown in Figure 11.

Figure 11: Schematic of the streamer and contracted normal glow discharge reactors (1 – power

supply, 2 – silicone stopper, 3 – electrodes, 4 – quartz tube, and 5 – Teflon® nozzle).

3.4 Results

The experimental results were obtained at varying flow rate, pressure, and

temperature conditions in the dissociation reactor. The experiments with corona and DBD

were focused on identifying and quantifying the effect that low power discharges with high

E/n have on H2S dissociation when these discharges are combined with elevated gas

temperatures (as usually in discharge physics, E/n is the ratio of electric field to gas

density). These experiments were designed to perform in such a way that the discharges by

themselves would cause measurable dissociation, but would not increase the gas

temperature considerably. The streamer and contracted normal glow discharge

experiments were focused on obtaining SER of the process where plasma is the only energy

source, and a wide range of power and flow rates were used to obtain a correspondingly

wide range of SEI. An attempt to find the optimal conditions for SER minimization was

made.

Page 35: Plasma dissociation of hydrogen sulfide

27

3.4.1 AC Corona Discharge

The experiments with corona discharge were performed at constant flow rate,

pressure, and power supply settings. The flow rate was set to 0.22 L/min (here as well as in

the following sections L/min refer to normal liters per minute at temperature 298 K and

pressure 1 atm). The power supply was set to its maximum frequency and voltage settings,

which corresponded to the discharge power of 2.1 W. The pressure was set to 265 Torr.

This regime was chosen in order to obtain noticeable and reproducible H2S dissociation of

0.5% at room temperature. The corresponding SEI was equal to 0.13 eV/molec, while SER

was 26.6 eV/molec.

The corona discharge operating in this regime did not provide noticeable gas

heating. Consequently, the furnace was used to heat the gas in the reactor to the desired

temperature. The furnace (Carbolite STF 16/610) allowed temperature variation in the

range 300 – 1200 K. Temperatures beyond 1200 K compromised reactor integrity by

melting the silicone stoppers and plastic tubing. The furnace temperature controller

provided stable operation at high temperatures (above 700 K) where the error would only

be on the level of 10 K or about 1%. At the lower temperatures the error was noticeably

higher, 25 – 50 K or up to 10%. Dependence of hydrogen concentration and SER of

hydrogen production on temperature in corona discharge and results of thermodynamic

equilibrium modeling for H2 concentration are shown on Figure 12 and Figure 13.

Page 36: Plasma dissociation of hydrogen sulfide

28

Figure 12: Comparison of measured hydrogen mole fraction in experiment and in thermodynamic

equilibrium modeling as function of temperature.

0

2

4

6

8

10

12

14

16

18

20

800 850 900 950 1000 1050 1100 1150 1200

H 2 Con

cent

ratio

n, [H

2] (m

ole

%)

Temperature, T (K)

w/ Corona

w/out Corona

Modeling

Page 37: Plasma dissociation of hydrogen sulfide

29

Figure 13: Full and relative SER of hydrogen production (b) as a function of gas temperature. The modeling was performed for the pressure of 0.25 atm under assumption of absolute quenching.

In the case of the corona experiment SER was calculated in two ways: relative

hydrogen production between the experiments with and without plasma, and full SER that

includes the heating energy of the furnace.

So, for each temperature setting relative SER, 𝜀𝑟:

And full SER, εf:

0

10

20

30

40

50

60

70

80

250 350 450 550 650 750 850 950 1050 1150

SER

(eV/

mol

ec)

Temperature (K)

Relative

Full

εr =𝐽

𝛼𝑝 − 𝛼𝑡 (21)

εf =Δ𝐻 + 𝐽𝛼𝑝

(22)

Page 38: Plasma dissociation of hydrogen sulfide

30

Here α is hydrogen concentration in a product flow for the case with plasma (αp) or

with the furnace only (αt), and Δ𝐻 is difference between gas enthalpies at the furnace and

room temperatures.

Summarizing results for AC corona discharge, at high temperatures hydrogen

production rate was below thermodynamic equilibrium; however, the experimental data

followed the thermodynamic equilibrium model under assumption of absolute quenching.

Hydrogen concentration difference between the experiments with corona and without

varied in the range from 0.2% to 1.7%, and corresponding SER of relative hydrogen

production in AC corona discharge also varied in a wide range 7.9 – 68.9 eV/molec at the

constant SEI of 0.13 eV/molec. A significant minimum of SER for the relative hydrogen

production is observed at the temperatures close to 1000 K. At this temperature range

corona accelerates the approach to chemical equilibrium, which may indicate a plasma

catalytic effect. It is worthwhile to study this effect in more detail in the future.

3.4.2 Dielectric Barrier Discharge

The experiments with DBD were performed at constant pressure, flow rate, and

power supply settings. The pressure was 187 Torr, and flow rate was 0.09 L/min. A

noticeable difference was observed in the results of DBD and corona experiments. The

hydrogen production in the DBD reactor was much higher at room temperature (18.5%).

The discharge filaments were few and probably hot. The exact temperature was not

measured, but the discharge heated up the surface of the reactor, which at that time was not

heated by the furnace, to approximately 60 – 70 °C, and inside the reactor within the area

covered by the outer electrode, the sulfur remained liquid (sulfur melting temperature is

388 K and boiling temperature is 717 K). The observed heating corresponded to the

significant increase of SEI (in comparison with corona), which was found to be 1.68

Page 39: Plasma dissociation of hydrogen sulfide

31

eV/molec. Dependence of hydrogen production in DBD on temperature and its comparison

with results of thermodynamic equilibrium modeling are shown on Figure 14. The

experiments with DBD positioned after the furnace did not differ significantly from those

obtained with DBD positioned in front of the furnace, and therefore, only the data from the

DBD positioned in front is presented.

Figure 14: Hydrogen production in DBD as a function of gas temperature and comparison with

results of thermodynamic equilibrium modeling under the assumption of absolute quenching. The modeling was performed for the pressure of 0.25 atm.

Summarizing DBD experimental results, dissociation of 18.5% of H2S was observed

at the room temperature. This level of dissociation, which is significantly above the

equilibrium value, remained virtually constant in the temperature range 300 – 700 K. A

reduction of dissociation was observed at temperatures above 700 K, which was likely

caused by the accelerated reverse reactions at higher temperatures. The SER of hydrogen

0

5

10

15

20

250 500 750 1000

H 2 Con

cent

ratio

n (m

ole

%)

Temperature (K)

Experiment

Modeling

Page 40: Plasma dissociation of hydrogen sulfide

32

production without additional gas heating was found to be 12 eV/molec at SEI of 1.68

eV/molec.

3.4.3 Streamer Discharge

Corona and DBD discharge, though different in their organization and SEI value,

have an important common feature that the following experiment with nanosecond

streamer discharge had to highlight. This common feature is low current and high 𝐸/𝑛 in

combination with overall low gas heating.

The streamer discharge was generated at different frequencies from 10 to 160 Hz

with 10 ns pulse of voltage (30 kV). The experiment was performed at constant flow rate of

0.15 L/min and constant pressure of 150 Torr. It was found that hydrogen production and

discharge power increased proportionally with the frequency (Figure 15) because energy of

each streamer was approximately constant, and as a result the SER remained virtually

constant and equal to 14 eV/molec.

Page 41: Plasma dissociation of hydrogen sulfide

33

Figure 15: Average power of the streamer discharge as a function of the discharge frequency.

0

2

4

6

8

10

0 50 100 150

Pow

er (W

)

Frequency (Hz)

Page 42: Plasma dissociation of hydrogen sulfide

34

Figure 16: Measured hydrogen concentration in the streamer discharge as a function of discharge

frequency.

3.4.4 Contracted Normal Glow Discharge

Completion of the experiments with DBD, corona, and streamer showed that the use

of discharges with high 𝐸/𝑛 and low (room) gas temperature is not an efficient approach for

hydrogen sulfide dissociation. The following experiments were designed to show how a

warm discharge [13-15] would perform at this task. This kind of discharge is much closer in

its properties to those that showed high efficiency of H2S dissociation [3, 6]. A warm

discharge is a non-equilibrium discharge with stepwise ionization, in which stable

concentration of electronically excited molecules is supported not only by electrons, but

also by gas temperature and/or by chemical processes. On the other hand, the gas

temperature of warm discharge is below the threshold of thermal ionization, where the role

of electric field in ionization becomes insignificant. Ionization in warm discharges has

0

1

2

3

4

5

6

0 50 100 150

H 2 Con

cent

ratio

n (m

ole

%)

Frequency (Hz)

Page 43: Plasma dissociation of hydrogen sulfide

35

comparable sensitivity to gas temperature as well as to 𝐸/𝑛 parameter. For H2S, the warm

discharge temperature range is between 1000 K and 3700 K.

The experiments with contracted normal glow discharge [13] were performed at

two flow rates of 1 and 2 L/min and constant pressure of 150 Torr. SEI of the process was

varied by varying the discharge current from 40 to 200 mA. The current remained stable

only above 40 mA. At the current of 200 mA and above the electrodes deteriorated quickly

and reactor parts overheated. The dependence of discharge voltage on current is presented

on Figure 17.

Figure 17: V-I characteristic of the contracted glow discharge.

Growing voltage-current characteristic is not typical for contracted glow discharges

[13], and in this particular case, it can be related to the growth of hydrogen concentration

and gas thermal conductivity. The gas temperature in the discharge channel was not

measured, and as it is one of the most important parameters, which would allow direct

500

600

700

800

0.04 0.05 0.06 0.07 0.08 0.09

Volta

ge (V

)

Current (A)

Page 44: Plasma dissociation of hydrogen sulfide

36

comparison with modeling results, it becomes a high priority objective for future

experiments. Taking into account discharge current and pressure, and comparing this

discharge with contracted glow discharges in different gases [15], we made preliminary

estimation that the gas temperature in the discharge should be in the range 2000 – 4000 K.

The elevated temperature in the discharge causes significant thermal dissociation,

but in this case hydrogen production is limited by slow quenching. The inefficient

quenching is illustrated on the Figure 18. The sufficiently fast quenching would maintain

the linearity of the hydrogen production curve, but instead it is leaning towards saturation.

Figure 18: Hydrogen concentration as function of discharge power in the contracted glow discharge

reactor.

The contracted glow discharge reactor in operation is shown in Figure 19 and

Figure 20. The figures illustrate intensive deposition of sulfur on cold walls of the reactor,

right outside of the discharge zone.

0

5

10

15

20

0 10 20 30 40 50 60

H 2 Con

cent

ratio

n (m

ole

%)

Power (W)

Page 45: Plasma dissociation of hydrogen sulfide

37

Figure 19: Contracted normal glow discharge reactor in operation (less then 1 min after start).

Figure 20: Contracted normal glow discharge reactor in operation (15 min after start).

A number of setting configurations were tested in experiments, and the

corresponding optimum setting was found at flow rate of 1 L/min and current of 50 mA.

The dependence of hydrogen production on discharge power at 1 L/min flow rate is shown

on Figure 18. The minimum SER corresponding to 50 mA current was 2.35 eV/molec and

the hydrogen production was 15.4%.

An attempt to increase H2S conversion was made and the reactor was modified to

include a quartz tunnel enclosing the plasma so that almost all hydrogen sulfide flow came

in contact with the discharge. Conversion rates of up to 60% were obtained but the SEI in

Page 46: Plasma dissociation of hydrogen sulfide

38

this system was rather high and the corresponding SER was not lower than in the system

without the tunnel.

3.5 Discussion

The experiments presented in this paper were designed to compare the influence of

different types of plasma discharges on dissociation of hydrogen sulfide, and they were not

intended to obtain the absolute minimum SER of hydrogen sulfide dissociation process. The

research goal was to obtain reproducible SEI and SER measurements using a variety of

plasmas in highly controllable conditions close to those in a plug flow reactor.

3.5.1 Corona Discharge

Literature research showed that there were no experiments done in the past with

AC corona discharge. Some work was done with pulsed corona, but if implemented

correctly, this discharge is the same as multiple streamer discharges, and therefore, these

works are compared with our streamer discharge experiments.

The dependence of H2 concentration on temperature, which is presented in Figure

12, can be explained, by low thermal dissociation kinetics at relatively low temperatures

(below 1100 K). However, in the case when the gas is heated to high temperature (close to

1200 K), it does not cool fast enough to prevent reverse reactions. An interesting

observation can be made from the Figure 13. At gas temperature of 970 K, SER of relative

hydrogen production is several times lower than at any other point. This result indicates

that corona discharge may have a catalytic effect on thermal dissociation, which is expected

in this temperature range. More experimental data is necessary to test this hypothesis and

to quantify this effect. It is also necessary to look at the chemical kinetics at this

temperature, and it would be advantageous to use some of the kinetic mechanisms

Page 47: Plasma dissociation of hydrogen sulfide

39

developed recently that have good agreement with experimental data in a close

temperature range [16-18].

3.5.2 Dielectric Barrier Discharge

There are few publications reporting the use of dielectric barrier discharge for H2S

dissociation, but nonetheless, one reported experiment had conditions that are very similar

to those presented in this paper [19]. The results that were reported in the literature are all

in the higher range of SEI 1 – 15 eV/molec, and corresponding SER was also quite high 37 –

106 eV/molec. The experimental results presented here were also obtained at rather high

SEI of 1.68 eV/molec, but the corresponding SER of 12 eV/molec was much lower than in

the experiments of Traus and Suhr [19]. This can be explained by the difference in pressure,

temperature and discharge frequency conditions in the experiments presented here and in

those presented by Traus and Suhr. The discharge of Traus and Suhr was arranged in a high

temperature zone with a discharge gap of only 1.5 mm, and in that case a high temperature

wall could catalyze the reverse reactions (compare with Figure 10).

The overall performance of the DBD discharge does not seem to be sufficient for

economically reasonable hydrogen sulfide dissociation, and the following should be noted:

DBD as well as corona has high E/n and low gas temperature. Even though it may be

possible to reduce SER by reducing SEI in the DBD reactor it still seems unlikely that this

discharge configuration could perform as well as higher temperature plasma discharges.

3.5.3 Streamer Discharge

Experiments with the streamer discharge were designed to test if cold discharges

with high E/n in a wide range of SEI will still have high SER until they transform into the

discharges with high gas temperature. It can be observed from Figure 16 that hydrogen

production grows linearly with specific energy input in the streamer discharge. This

Page 48: Plasma dissociation of hydrogen sulfide

40

indicates that electronic impact is the primary source of dissociation. A number of

mechanisms for hydrogen sulfide dissociation initiated by electronic impact have been

proposed, and some of them are presented in Table 1.

Table 3: Different mechanisms for electronic impact initiated dissociation of hydrogen sulfide.

Mechanism Source

Direct ionization of H2S followed by dissociative neutralization:

e + H2S → H2S+ + 2e H2S+ + e → HS + H

Ma et al., 2001 Helfritch et al., 1993

Ionization of the balance gas (M), leading to charge transfer reaction, and subsequent dissociative neutralization:

e + M → M+ + 2e M+ + H2S → H2S+ + M H2S+ + e → HS + H

Abolentsev et al., 1995

Direct electron collision dissociation of H2S:

e + H2S → HS + H + e

Electron collision dissociation or excitation of the balance gas, which produces active species that contribute to H2S dissociation:

e + M → M∗ + e M∗ + H2S → H + HS + M

Zhao et al., 2007

In general, the results of our experiments with the streamer discharge are

comparable with the SER of pulsed corona discharge reported by other research groups [20,

21]. No other research works were found discussing a single streamer for H2S dissociation,

which is mostly due to the fact that a single streamer system is not practical for any

applications. Only recent reports showed some pulsed corona results that are significantly

better than any other previously reported results for corona as well as many other

discharges at pressures of 1 atm and higher [22]. It is important to note that high

Page 49: Plasma dissociation of hydrogen sulfide

41

frequencies used in the experiments by John et al. [22] (when the best results were

obtained) may have led to formation of sparks or just heating of the discharge channels.

These assumptions are based on the fact that the discharge was visually much brighter than

more powerful corona units operating at atmospheric pressure [23]. The authors reported

that the discharge was not uniformly distributed along the high voltage wire anode, but was

concentrated in different places depending on the gas nature. Existence of these

“preferable” places means that the conductance did not drop completely between voltage

pulses, and this makes this discharge closer to a filamentary DBD or spark discharge. The

reported energy input could heat the gas in the reactor by 450 K, thus making it 750 K total.

This temperature alone cannot be responsible for the SER of 4.9 eV/molec (the best results

reported in [22]), however, if the energy input was not uniform, local temperature increase

could be much higher and in that case the reported low SER can be a result of local thermal

dissociation.

3.5.4 Contracted Normal Glow Discharge

The switch from streamer to a contracted glow discharge was performed without

changing reactor geometry or significant changes in reactor pressure, but the results were

very different. The reason for this is a change in gas temperature. Comparing the results

with thermodynamic expectations (see section 1.1), it can be concluded that in the hot

plasma channel hydrogen sulfide dissociates to the equilibrium concentration

corresponding to 2000 – 3000 K, and then once the mixture leaves the plasma zone, it starts

to recombine to hydrogen sulfide. There was no special cooling used to increase quenching

speed, and therefore the resulting concentration of hydrogen is noticeably lower than

expected in the case of absolute quenching. It is important to note that this experiment was

performed in the simplest geometry, which can be viewed as a plug flow reactor where the

Page 50: Plasma dissociation of hydrogen sulfide

42

heat is released in very short zone across the flow. This zone also does not cover the reactor

cross section fully but approximately 30% of it.

There was not much work done on H2S dissociation in glow discharges and the

results that were reported [24] are not particularly impressive. Traus et al. reported SER of

25 – 42 eV/molec with 23 – 32% conversion of H2S at low pressures and SER 25 – 36

eV/molec with 25 – 35% conversion of H2S atmospheric pressure. What is important is that

these reported values correspond to a very high SEI of more than 7 eV/molec. Most likely,

there were very high discharge energy losses to the reactor wall because of very short

discharge length.

A discharge that usually can be considered “warm,” and therefore is comparable

with the contracted normal glow discharge, is gliding arc. This type of discharge has been

successfully used for different kinds of fuel conversion and waste treatment [25, 26], and

has been used for H2S conversion as well. The SER results that were reported for gliding

arcs range from 500 eV/molec to 1.2 eV/molec [6, 27-29]. Gliding arc discharge is sensitive

to flow organization, and a review of the literature shows that the researchers often use

very different gas-dynamic approaches.

Assuming that the main dissociation path in warm discharges is thermal, then, the

SER of hydrogen production will largely depend on reactor configuration. Reactor

configuration can, in general, allow SER reduction because of energy recovery or

equilibrium shift due to high diffusion coefficient of hydrogen. It is necessary to remember

that our system was designed to imitate a plug flow reactor, which allows to separate the

effect of the process itself and possible effects related to gas-dynamic or diffusion effects.

Page 51: Plasma dissociation of hydrogen sulfide

43

3.5.5 Summary of the Results

The overview of the experimental results presented in this paper is summarized in

the form of SER as function of SEI (Figure 21).

Figure 21: SER as function of SEI for all the discharge configurations (including MW [3, 4, 30] and

GAT [6]) and thermodynamic equilibrium model under ideal quenching assumption.

The results are compared with thermodynamic equilibrium modeling and two of the

best sets of results reported earlier [5, 6]. Since all experiments were performed at low

pressure in the range from 150 Torr (contracted glow) to 265 Torr (corona), both GAT and

0.1

1

10

100

0 0.5 1 1.5 2 2.5

SER,

ε (e

V/m

olec

)

SEI, J (eV/molec)

Model @ 1 atm

Model @ 0.25 atm

Model @ 0.25 atm, 0.3 SEI

GAT @ 0.3 atm

MW @ 0.1 atm

Corona

Streamer

DBD

Glow

Page 52: Plasma dissociation of hydrogen sulfide

44

Microwave results were also reported for low pressure, the modeling curves are given for

low pressure as well. The second modeling curve corresponds to the low pressure and a

case when only 30% of the gas is treated by the plasma, which is probably a better

approximation of the experimental conditions in the contracted glow discharge.

The experimental points for corona, dielectric barrier, streamer, and glow

discharges all have 10% error for both SEI and SER. The major contribution to this error is

flow rate measurements. This error is avoidable and will be minimized in the future work.

Analysis showed that error of GAT experiments was larger than it was reported earlier [6],

and the lowest SER of hydrogen production in GAT reactor was 1.5 ± 0.3 eV/molec. The

error of the experiments with microwave discharge is unknown, but more information

about the experiments can found in papers [3, 4, 30, 31]. Flow organizations in the GAT and

microwave plasma reactors were rather complex. These flow organizations could result in

partial thermal energy recovery in the systems, which results in SER reduction below the

thermodynamic equilibrium level.

Page 53: Plasma dissociation of hydrogen sulfide

45

CHAPTER 4: CONCLUSIONS

The experiments on dissociation of hydrogen sulfide in a plug flow reactor with

assistance of corona, dielectric barrier, and streamer discharges showed that low

temperature discharges with high 𝐸/𝑛 and low gas temperature produce results

characterized by a common SER higher than 10 eV/molec. This corresponds to the reaction

mechanisms that are initiated by direct electronic impact. Significant reduction of SER at

low temperature conditions does not look possible because of the absence of chemical chain

processes.

The results obtained in contracted glow discharge at lower 𝐸/𝑛 have the lowest SER

of 2.35 eV/molec, which is close to the thermodynamic equilibrium value in the assumption

of absolute quenching. The SER is slightly higher than the equilibrium one and this can be

explained by multiple factors like discharge non-uniformity (see section 1.1), non-optimal

temperature, and reverse reactions taking place downstream of the discharge zone.

Experiments confirmed that an electrical discharge in H2S allows obtaining SER

close to that predicted by thermodynamic equilibrium modeling if the discharge has gas

temperature that is high enough for thermal dissociation. Further reduction of SER in a plug

flow reactor with thermal dissociation is not possible because of the absence of energy

recovery mechanisms.

Contracted glow discharge experiments led us to the conclusion that thermal

dissociation is the key factor in obtaining reasonably low SER for hydrogen production from

hydrogen sulfide. Electrical discharges, operating at gas temperature of 2000 – 4000 K,

which is optimal for hydrogen sulfide pyrolysis, create a preferred medium for plasma

dissociation of hydrogen sulfide. These conditions can be produced in the following

Page 54: Plasma dissociation of hydrogen sulfide

46

discharges: contracted normal glow, gliding arc, microwave, and low pressure radio

frequency discharge (RF) [30].

Page 55: Plasma dissociation of hydrogen sulfide

47

LIST OF REFERENCES

[1] Luinstra EA. Hydrogen from H2S: Technologies and Economics. Calgary: Sulfotech Research; 1995.

[2] Zaman J, Chakma A. Production of hydrogen and sulfur from hydrogen sulfide. Fuel Processing Technology. 1995;41:159-98.

[3] Bagautdinov AZ, Jivotov VK, Eremenko JI, Kalachev IA, Musinov SA, Potapkin BV, et al. Plasma chemical production of hydrogen from H2S-containing gases in MCW discharge. International Journal of Hydrogen Energy. 1995;20:193-5.

[4] Kasharin AV, Potapkin BV, Rusanov VD, Fridman AA. Energetics of plasma-chemical systems in selective transfer processes. Journal of Engineering Physics and Thermophysics. 1989;57:1335-43.

[5] Balebanov AV. Dissociation of Hydrogen Sulfide in a Plasma. Doklady Physical Chemistry, Proceedings of the Academy of Sciences of the USSR. 1986;283:709-12.

[6] Nunnally T, Gutsol KA, Rabinovich A, Fridman AA, Starikovsky A, Gutsol AF, et al. Dissociation of H2S in non-equilibrium gliding arc "tornado" discharge. International Journal of Hydrogen Energy. 2009;34:7618-25.

[7] Asisov RI, Vakar AK, Gutsol AF, Zhivotov VK, Krasheninnikov EG, Rusanov VD, et al. Plasma chemical methods of energy carrier production. International Journal of Hydrogen Energy. 1985;10:475-7.

[8] Fridman AA. Plasma chemistry: Cambridge University Press; 2008.

[9] Kasharin AB, Potapkin BV, Rusanov VD, Fridman AA. Obout energetics of of plasmachemical systems in selective nature of transport process. 1988.

Page 56: Plasma dissociation of hydrogen sulfide

48

[10] Steudel R. Topics in current chemistry. In: Steudel R, editor. Elemental sulfur and sulfur-rich compounds. Berlin: Springer-Verlag; 2003.

[11] Fridman AA, Kennedy LA. Plasma physics and engineering. New York, London: Taylor and Francis; 2004.

[12] Raizer YP. Gas discharge physics. Berlin Heidelberg: Springer-Verlag; 1991.

[13] Staack D, Farouk B, Gutsol AF, Fridman AA. Characterization of a DC atmospheric pressure normal glow discharge. Plasma Sources Science and Technology. 2005;14:700.

[14] Staack D, Farouk B, Gutsol AF, Fridman AA. Spectroscopic studies and rotational and vibrational temperature measurements of atmospheric pressure normal glow plasma discharges in air. Plasma sources science and technology. 2006;15:818–27.

[15] Staack D, Farouk B, Gutsol AF, Fridman AA. DC normal glow discharges in atmospheric pressure atomic and molecular gases. Plasma Sources Science and Technology. 2008;17:025013.

[16] Binoist M, Labégorre B, Monnet F, Clark PD, Dowling NI, Huang M, et al. Kinetic Study of the Pyrolysis of H2S. Industrial & Engineering Chemistry Research. 2003;42:3943-51.

[17] Shiina H, Oya M, Yamashita K, Miyoshi A, Matsui H. Kinetic Studies on the Pyrolysis of H2S. The Journal of Physical Chemistry. 1996;100:2136-40.

[18] Woiki D, Roth P. Kinetics of the High-Temperature H2S Decomposition. The Journal of Physical Chemistry. 1994;98:12958-63.

[19] Traus I, Suhr H. Hydrogen sulfide dissociation in ozonizer discharges and operation of ozonizers at elevated temperatures. Plasma Chemistry and Plasma Processing. 1992;12:275-85.

[20] Helfritch DJ. Pulsed corona discharge for hydrogen sulfide decomposition. Industry Applications, IEEE Transactions on. 1993;29:882-6.

[21] Zhao GB, John S, Zhang JJ, Hamann JC, Muknahallipatna SS, Legowski S, et al. Production of hydrogen and sulfur from hydrogen sulfide in a nonthermal-plasma pulsed corona discharge reactor. Chemical Engineering Science. 2007;62:2216-27.

Page 57: Plasma dissociation of hydrogen sulfide

49

[22] John S, Hamann JC, Muknahallipatna SS, Legowski S, Ackerman JF, Argyle MD. Energy efficiency of hydrogen sulfide decomposition in a pulsed corona discharge reactor. Chemical Engineering Science. 2009;64:4826-34.

[23] Tak G, Gutsol AF, Fridman AA. Pulsed Corona Plasma Pilot Plant for VOC Abatement in Industrial Streams as Mobile and Educational Laboratory. 17th International Symposium on Plasma Chemistry. Toronto, Canada2005. p. 1128-9.

[24] Traus I, Suhr H, Harry JE, Evans DR. Application of a rotating high-pressure glow discharge for the dissociation of hydrogen sulfide. Plasma Chemistry and Plasma Processing. 1993;13:77-91.

[25] Gallagher MJ, Geiger R, Polevich A, Rabinovich A, Gutsol A, Fridman A. On-board plasma-assisted conversion of heavy hydrocarbons into synthesis gas. Fuel. 2010;89:1187-92.

[26] Kalra CS, Gutsol AF, Fridman AA. Gliding arc discharges as a source of intermediate plasma for methane partial oxidation. IEEE transactions on plasma science. 2005;33.

[27] Czernichowski A. Gliding arc. Applications to engineering and environment control. 11th International Symposium on Plasma Chemistry. Loughborough, UK: Pure and Applied Chemistry; 1994. p. 1301-10.

[28] Czernichowski A, Czernichowski M, Wesolowska K. Glidarc-Plasma Reactors for Hydrogen Recovery from Waste H2S. HYPOTHESIS V. Porto Conte, Italy2003.

[29] Dalaine V, Cormier JM, Lefaucheux P. A gliding discharge applied to H2S destruction. Journal of Applied Physics. 1998;83:2435-41.

[30] Krasheninnikov EG, Rusanov VD, Sanyuk SV, Fridman AA. Dissociation of hydrogen sulfide in an RF discharge. Soviet Physics - Technical Physics. 1986;31:645-8.

[31] Bagautdinov AZ, Jivotov VK, Eremenko JI, Kalachev IA, Kozbagarov AI, Konstantinov EI, et al. Plasmachemical Conversion of Hydrogen Sulfide. 2nd International Symposium on Heat and Mass Transfer under Plasma Conditions. Antalya, Turkey1999.