You are on page 1of 132

REMOVAL OF HYDROGEN SULFIDE FROM BIOGAS

USING COW-MANURE COMPOST

A Thesis

Presented to the Faculty of the Graduate School

of Cornell University

in Partial Fulfillment of the Requirements for the Degree of

Master of Science

by

Steven McKinsey Zicari

January 2003
© 2003 Steven McKinsey Zicari
ABSTRACT

The two-part objective of this study was to determine currently available H2S

removal technologies suitable for use with farm biogas systems, and to test the

feasibility of utilizing on-farm cow-manure compost as an H2S adsorption medium.

Integrated farm energy systems utilize anaerobic digesters to provide a waste

treatment solution, improved nutrient recovery, and energy generation potential in the

form of biogas, which consists mostly of methane and carbon dioxide, with smaller

amounts of water vapor, and trace amounts of H2S and other impurities. Hydrogen

sulfide usually must be removed before the gas can be used for generation of

electricity or heat. Biogas has remained a virtually untapped resource in the United

States due to many factors, including relatively high gas processing costs.

There are many chemical, physical, and biological methods currently available

for removal of H2S from an energy gas stream. Dry based chemical processes have

traditionally been used for biogas applications and remain competitive based on

capital and media costs. Iron Sponge, Media-G2®, and potassium-hydroxide-

impregnated activated-carbon systems are the most attractive, with estimated capital

costs of $10,000-$50,000 and media costs of $0.35-$3.00/kg H2S removed. These

processes are simple and effective, but also incur relatively high labor costs for

materials handling and disposal. Other significant drawbacks include a continually

produced solid waste stream and growing environmental concern over appropriate

disposal methods. Additions of air (2-6%) to the digester headspace, or iron

compounds introduced directly in the digester, show promise as partial H2S abatement
methods, but have limited and inconsistent operational histories. Liquid based and

membrane processes require significantly higher capital, energy and media costs, and
do not appear economically competitive for selective H2S removal from biogas at this
time. Commercial biological processes for H2S removal are available (Biopuric and

Thiopaq) that boast reduced operating, chemical, and energy costs, but require higher

capital costs than dry based processes.

Initial testing of cow-manure compost indicates that it has potential as an

effective and economic matrix for H2S removal. Polyvinyl-chloride (PVC) test

columns were constructed and a 2:1 biogas-to-air mixture passed through the columns

containing anaerobically digested cow-manure compost. The most significant trials

ran for 1057 hours with an empty-bed gas residence time near 100 seconds and inlet

H2S concentrations averaging 1500 ppm, as measured by an electrochemical sensor

with 40:1 sample dilution.

Removal efficiencies over 80% were observed for a majority of the run time.

Elimination capacities recorded were between 16 – 118 g H2S/m3 bed/hr. This is

significant considering only minimal moisture, and no temperature or pH controls

were implemented. Temperature in the bed varied from 19-43°C and the moisture

contents in the spent column ranged from 41-70%, with pH values from 4.6 to 6.9. It

is not clear whether the major mechanism for sulfur removal from the gas stream was

biological, chemical or physical, but it is known that the sulfur content in the packing

increased by over 1400%, verifying sequestration of sulfur in the compost.

These initial results indicate that future work is warranted for examining the

suitability of cow-manure compost as a biofiltration medium for use with biogas.


BIOGRAPHICAL SKETCH

Steven McKinsey Zicari was born in Rochester, New York, to Richard E. and

D. June Zicari. He grew up with his older brother, Zev, and attended West

Irondequoit public schools through the 12th grade. Steven enrolled at Cornell

University in the fall of 1995, and focused his studies on biological engineering. He

graduated with a Bachelor of Science degree in Agricultural and Biological

Engineering in May, 1999, Cum Laude. As an undergraduate, he also participated in

the alpine ski team, symphonic band, and the engineering co-op program. His co-op

experiences were with Genencor International in Rochester, NY, and Nestle R&D in

New Milford, CT.

After working briefly for the New York State Department of Agriculture and

Markets as a farm products inspector, and also as a ski instructor in Vail, Colorado,

Steven decided to return to Cornell for graduate school in the Fall of 2000. He has

held teaching and research assistant positions in the Department of Biological and

Environmental Engineering and his current research interests include sustainable

development, alternative and renewable energy systems, and biological processes.

iii
To my family

iv
ACKNOWLEDGMENTS

I would like to thank my major advisor, Dr. Norman Scott, for his guidance,

creativity, and tireless effort in researching sustainable development. I have learned a

great deal by working closely with him. I am also grateful to my minor advisor, Dr.

A. Brad Anton, for his helpful insights, positive encouragement, and superb technical

competence. I also extend thanks to committee member Dr. Anthony Hay for his

continual enthusiasm, willingness to help, and for sharing his exceptional

understanding of biological systems.

I acknowledge the Biological and Environmental Engineering department,

especially Dr. Dan Aneshansley and Dr. Michael Walter, for supporting me with

teaching opportunities and sound advice during my studies here. Also the knowledge

and cooperation of Dr. Larry Walker, Doug Caveny and Peter Wright are greatly

valued. Additionally, I greatly appreciate the cooperation of Robert, Wayne, and

Aaron Aman for allowing me to perform tests at AA Dairy.

Special thanks are also given to fellow graduate student John Poe Tyler. His

expert mechanical and engineering skills, as well as determination, were invaluable. I

also thank Tina Jeoh for her constant motivation, encouragement and assistance.

The support of my fellow research-group members, officemates and fellow

graduate students are also greatly appreciated, especially Kristy Graf, Jianguo Ma,

Stefan Minott, Scott Pryor, and Kelly Saikkonen.

Lastly, I would like to thank all of my family and friends for their support,

without which, this would not have been possible.

v
TABLE OF CONTENTS

BIOGRAPHICAL SKETCH.........................................................................................iii
ACKNOWLEDGMENTS.............................................................................................. v

1. INTRODUCTION ...................................................................................................... 1

2. BACKGROUND........................................................................................................ 3
2.1. INTEGRATED FARM ENERGY SYSTEMS ....................................... 3
2.1.1. AA Dairy .......................................................................................... 4
2.2. ANAEROBIC DIGESTION ................................................................... 6
2.3. BIOGAS COMPOSITION...................................................................... 9
2.4. QUALITY REQUIREMENTS FOR BIOGAS UTILIZATION .......... 10
2.5. TRADITIONAL H2S GAS-PHASE REMOVAL METHODS ............ 12
2.5.1. Dry H2S Removal Processes .......................................................... 13
2.5.1.1. Iron Oxides .............................................................................. 14
2.5.1.2. Zinc Oxides ............................................................................. 22
2.5.1.3. Alkaline Solids ........................................................................ 24
2.5.1.4. Adsorbents............................................................................... 24
2.5.2. Liquid H2S Removal Processes ...................................................... 30
2.5.2.1. Liquid-Phase Oxidation Processes .......................................... 31
2.5.2.2. Alkaline Salt Solutions ............................................................ 35
2.5.2.3. Amine Solutions ...................................................................... 36
2.5.3. Physical Solvents............................................................................ 38
2.5.3.1. Water Washing ........................................................................ 39
2.5.3.2. Other Physical Solvents........................................................... 39
2.5.4. Membrane Processes ...................................................................... 40
2.6. ALTERNATIVE H2S CONTROL METHODS.................................... 41
2.6.1. In-Situ (Digester) Sulfide Abatement............................................. 41
2.6.2. Dietary Adjustment ........................................................................ 42
2.6.3. Aeration .......................................................................................... 43
2.7. BIOLOGICAL H2S REMOVAL METHODS ...................................... 43
2.7.1. History and Development............................................................... 43
2.7.2. Biological Sulfur Cycles................................................................. 45
2.7.3. Example Applications .................................................................... 50
2.8. RESEARCH STATEMENT ................................................................. 54

3. MATERIALS AND METHODS ............................................................................. 55


3.1. REACTORS .......................................................................................... 55
3.1.1. Small Reactors................................................................................ 55
3.1.2. Large Reactors................................................................................ 58
3.2. EXPERIMENTAL SETUP ON SITE ................................................... 60
3.3. GAS SAMPLING AND MEASUREMENT......................................... 64
3.3.1. Electrochemical Sensor .................................................................. 64
3.3.2. Gas Sampling Tubes....................................................................... 66

vi
3.3.3. Gas Chromatography...................................................................... 67
3.4. TEMPERATURE MEASUREMENT................................................... 67
3.5. PRESSURE MEASUREMENT............................................................ 68
3.6. COMPOST CHARACTERIZATION................................................... 68
3.6.1. Moisture Content ............................................................................ 68
3.6.2. Void Fraction:................................................................................. 69
3.6.3. Bulk Density................................................................................... 69
3.6.4. Particle Size Distribution................................................................ 69
3.6.5. pH ................................................................................................... 70
3.6.6. Trace Element Analysis.................................................................. 70
3.6.7. Sulfate Content ............................................................................... 70
3.7. OPERATIONAL PROCEDURES ........................................................ 70

4. RESULTS AND DISCUSSION............................................................................... 72


4.1. ORIGINAL COMPOST CHARACTERISTICS .................................. 72
4.2. OPERATIONAL SUMMARY ............................................................. 74
4.3. PRESSURE MEASUREMENTS.......................................................... 75
4.4. TEMPERATURE MEASUREMENTS ................................................ 77
4.5. HYDROGEN SULFIDE MEASUREMENTS...................................... 84
4.5.1. Electrochemical Sensor .................................................................. 84
4.5.2. Gas Detector Tubes ........................................................................ 90
4.5.3. Gas Chromatography...................................................................... 91
4.6. BIOGAS-EXPOSED-COMPOST ASSESSMENT .............................. 93
4.6.1. Moisture.......................................................................................... 93
4.6.2. pH ................................................................................................... 95
4.6.3. Trace Element Analysis.................................................................. 95
4.7. DISCUSSION........................................................................................ 97
4.8. SCALE-UP CONSIDERATIONS ........................................................ 98

5. SUMMARY AND CONCLUSIONS..................................................................... 102


5.1. CURRENTLY AVAILABLE H2S REMOVAL METHODS............. 102
5.1.1. Dry-Based Processes .................................................................... 102
5.1.2. Liquid-Based Chemical and Physical Processes .......................... 105
5.1.3. Membrane Separation................................................................... 105
5.1.4. In-Situ Digester Sulfide Control................................................... 105
5.1.5. Biogas Aeration ............................................................................ 106
5.1.6. Biological Removal Techniques................................................... 106
5.1.7. Comparison of Characteristics of H2S Removal Processes.......... 106
5.2. TESTING OF COW-MANURE COMPOST ..................................... 108

6. FUTURE WORK AND RECOMMENDATIONS ................................................ 109

APPENDIX A: H2S Scavenger Media Disposal ........................................................ 111

REFERENCES ........................................................................................................... 112

vii
LIST OF TABLES

Table 2.1: Characteristics of Typical Agricultural Anaerobic Digesters ....................... 7


Table 2.2: Physical, Chemical and Safety Characteristics of Hydrogen Sulfide ......... 10
Table 2.3: Biogas Utilization Technologies and Gas Processing Requirements.......... 11
Table 2.4: Principal Gas Phase Impurities ................................................................... 12
Table 2.5: Assumed Biogas Characteristics for Process Comparisons ........................ 13
Table 2.6: Typical Specifications for 15-lb Iron Sponge ............................................. 15
Table 2.7: Iron Sponge Design Parameter Guidelines ................................................. 17
Table 2.8: System Characteristics of 15-lb Iron Sponge Design at AA Dairy............. 18
Table 2.9: System Characteristics of SulfaTreat® Design at AA Dairy ....................... 20
Table 2.10: System Characteristics of Sulfur-Rite® Design at AA Dairy.................... 21
Table 2.11: System Characteristics of Media-G2® Design at AA Dairy ..................... 22
Table 2.12: Processes for Adsorbent Regeneration...................................................... 25
Table 2.13: Basic Types of Commercial Molecular Sieves ......................................... 26
Table 2.14: Summary of 5A Molecular Sieve Design at AA Dairy............................. 28
Table 2.15: System Characteristics for KOH-Impregnated Activated Carbon at AA
Dairy ..................................................................................................................... 29
Table 2.16: Henry’s Law Constants at 25° C and 1-Atmosphere ................................ 39
Table 2.17: Specific Microorganisms Studied for Biofiltration of H2S ....................... 49
Table 2.18: Media Tested for Biofiltration of Hydrogen Sulfide................................. 50
Table 3.1: Cross Sensitivity Data for Electrochemical H2S Sensor ............................. 65
Table 3.2: Summary of Experimental Trial Conditions ............................................... 71
Table 4.1: Cow-Manure Compost Characterization..................................................... 73
Table 4.2: Summary of Temperature Extremes for Trials 3-6 ..................................... 81
Table 4.3: H2S Gas Detector Tube Readings for AA Dairy Raw Digester Gas........... 90
Table 4.4: GC-MS Results for AA Dairy Digester Gas ............................................... 91
Table 4.5: Moisture Contents Along Bed Depth .......................................................... 94
Table 4.6: pH Levels Along Bed Depth ....................................................................... 95
Table 4.7: Elemental Analysis of Raw and Tested Compost ....................................... 96
Table 4.8: Estimated Comparison of Cow-Manure Compost and Iron-Sponge H2S-
Removal Systems at AA Dairy........................................................................... 101
Table 5.1: Summary Table Comparing Dry-Based H2S Removal Processes for Farm
Biogas ................................................................................................................. 103
Table 5.2: Summary Table Comparing Dry-Based H2S Removal Processes for AA
Dairy ................................................................................................................... 104
Table 5.3: Summary of H2S Removal Process Characteristics .................................. 107
Table A.1. Approximate Media Change-out and Disposal Costs (1996 est.) ............ 111

viii
LIST OF FIGURES

Figure 2.1: Schematic of AA Dairy Integrated Farm Energy System............................ 5


Figure 2.2: Anaerobic Digestion Process ....................................................................... 8
Figure 2.3: Equilibrium Constant for the Reaction ZnO + H2S = ZnS + H2O............. 23
Figure 2.4: Adsorption Zones in a Molecular Sieve Bed, Adsorbing Both Water Vapor
and Mercaptans from Natural Gas........................................................................ 27
Figure 2.5: Generic Absorber/Stripper Schematic ....................................................... 30
Figure 2.6: Reduction-Oxidation Cycle of Quinones................................................... 32
Figure 2.7: Conventional Flow Diagram for LO-CAT® Process ................................. 33
Figure 2.8: Flow Scheme for Alkanolamine Acid-gas Removal Processes................. 38
Figure 2.9: Biofiltration System Schematic ................................................................. 45
Figure 2.10: The Global Sulfur Cycle. ......................................................................... 46
Figure 2.11: Biological Redox Cycle for Sulfur .......................................................... 47
Figure 2.12: Steps in the Oxidation of Sulfur Compounds by Thiobacillus Species. .. 48
Figure 3.1: Schematic of Small Columns..................................................................... 56
Figure 3.2: Schematic of Small Columns with Leachate Recycle ............................... 57
Figure 3.3: Schematic of Large Columns..................................................................... 59
Figure 3.4: Experimental Setup at AA Dairy ............................................................... 61
Figure 3.5: Humidifier and Air/Biogas Mixing Vessel ................................................ 63
Figure 4.1: AA-Dairy “Field of Dreams” Cow-Manure Compost ............................... 72
Figure 4.2: Pressure Drop Across Bed for Trials 3-6................................................... 76
Figure 4.3: Temperatures (15-Minute Average) for Column 3 .................................... 78
Figure 4.4: Temperatures (15-Minute Average) for Column 4 .................................... 78
Figure 4.5: Temperatures (15-Minute Average) for Column 5 .................................... 79
Figure 4.6: Temperatures (15-Minute Average) for Column 6 .................................... 79
Figure 4.7: Temperature Difference Between Bed and Inlet-gas for Columns 3-6 ..... 80
Figure 4.8: H2S Concentrations for Trial 3 .................................................................. 85
Figure 4.9: H2S Removal Efficiency During Trial 3.................................................... 86
Figure 4.10: H2S Concentrations for Trial 4 ................................................................ 86
Figure 4.11: H2S Removal Efficiency During Trial 4.................................................. 87
Figure 4.12: H2S Concentrations and Removal Efficiency for Column 5 ................... 89
Figure 4.13: H2S Concentrations and Removal Efficiency for Column 6 ................... 89
Figure 4.14: GC-MS Results for AA Dairy Digester Gas............................................ 92
Figure 4.15: Pictures of Columns after Exposure to Biogas for 1057 hours................ 93

ix
CHAPTER

1. INTRODUCTION

Anaerobic digestion (AD) of agricultural waste has been practiced for many

years and provides a waste treatment solution, improved nutrient recovery, and energy

generation potential. Because of growing environmental constraints, an increase in the

average dairy farm herd size, and rising energy costs from increased demand, dairy

farmers are looking to AD coupled with on-site cogeneration of heat and power in

response to these pressures. However, there are hurdles to implementation of these

systems, including high capital costs, availability of economic and environmentally

acceptable methods of gas processing, and economic means for biogas utilization.

Because of these limitations, agricultural biogas production has remained a virtually

untapped resource in the United States.

Biogas consists mainly of methane (CH4) and carbon dioxide (CO2), with

smaller amounts of water vapor and trace amounts of hydrogen sulfide (H2S), and

other impurities. Various degrees of gas processing are necessary depending on the

desired gas utilization process. Hydrogen sulfide is typically the most problematic

contaminant because it is toxic and corrosive to most equipment. Additionally,

combustion of H2S leads to sulfur dioxide emissions, which have harmful


environmental effects. Removing H2S as soon as possible is recommended to protect

downstream equipment, increase safety, and enable possible utilization of more

efficient technologies such as microturbines and fuel cells.

1
2

The most commonly used method for H2S removal from biogas involves
adsorption onto chemically active solid media. Though this process is effective, it is

labor intensive and generates a waste stream that poses environmental disposal risks.

These factors led to the identification of an opportunity for testing on-farm

manure compost as the adsorption medium. A similar process, known as biofiltration,

has shown its ability to remove H2S through the microbial action of naturally

occurring bacteria. Biofilters show promise as environmentally friendly, alternative

air pollution control technologies with lower capital, labor, and disposal costs.

The following objectives were specifically addressed in this study:

1) Survey currently available chemical, physical, and biological methods of

H2S removal from agricultural digester biogas.

2) Test the feasibility of utilizing on-farm cow-manure compost for H2S

removal.

AA Dairy farm in Candor, NY, which has produced biogas since 1998,

served as the site for experimental testing. Although removal of water vapor, carbon

dioxide, and other contaminants is also desirable, assessing all of the technologies

required for removal of these compounds is beyond the scope of this project.

It is hoped that this research not only benefits farmers who are looking to

install integrated farm energy systems, but also designers and operators of other

agricultural facilities, landfills, wastewater treatment plants, food-processing facilities,

and pulp and paper mills, where renewable, bio-based energy can be produced.
CHAPTER

2. BACKGROUND

2.1. INTEGRATED FARM ENERGY SYSTEMS

The concept of an integrated farm energy (IFE) system is an embodiment of

principals of industrial ecology, which attempt to improve on the efficiency and

sustainability of a system by optimizing energy and material usage while minimizing

pollution and waste. IFE systems, as referred to here, use anaerobic digestion (AD) to

treat the volatile organic fraction of animal manures, thereby producing biogas and an

improved waste stream. The biogas is then used for on-site heat and/or power

generation, and the digested material is either applied back to the land or processed

further into value-added compost. In 1995, a study estimated that three to five

thousand such systems could be economically installed throughout the next decade in

the U.S. (Lusk 1996). In 1999, there were only 34 operating farm-digester sites

registered with the EPA’s AgSTAR program, though this number has since grown

(Roos and Moser 2000). According to one estimate, if all of the dairy manure biomass

in New York state could be collected and processed using anaerobic digestion and

diesel engine generation, an annual energy potential of 280 GWh, enough to support

the electricity demand of about 47,000 households, would be produced in addition to

providing all of the electricity for the farms (Ma 2002).

Extensive research on these integrated systems was done during the 1970’s and

1980’s by Cornell University researchers, and further information on their

3
4

development and design can be found in Jewell, et al. (1978), Walker, et al. (1985),

and Pellerin, et al. (1988). The integrated farm energy system used in this study is

operating at AA Dairy in Candor, NY.

2.1.1. AA Dairy

AA Dairy is a 2,200-acre, 500-head milking facility, which installed an

anaerobic digester in 1997. Resource recovery is achieved in part through use of an

anaerobic digester, diesel-engine cogenerator, liquid-solid separator, liquid-waste

storage lagoon, composting process and land application of effluent, as depicted in

Figure 2.1. Most of the cows are housed in a free-stall facility equipped with alley

scrapers for manure collection. A 1330 m3 concrete plug-flow digester, designed by


Resource Conservation Management, operates with approximately a 40-day retention

time and 1300 m3 per day of biogas produced on average. The digested solids are

separated and composted aerobically for a period of 60 days and sold to consumers as

a specialty organic fertilizer. The liquid portion is stored in a lagoon until land

application for nutrient value and water are needed. The biogas is combusted in a

converted Caterpillar 3306B diesel engine, which powers a generator continuously

producing 65 to 75 kW. Electricity unused by the farm (~535 kWh/day average) is

then sold to the local utility (NYSEG). Heat from the engine is currently used to

maintain the digester in its desired mesophillc operating range. The current method

for dealing with biogas impurities, such as hydrogen sulfide, is to perform a 70-quart

oil change weekly. No other gas processing technologies are employed, and the

annual operating cost for the resource recovery system, including labor and engine

maintenance, is estimate as $17,500 (Minott 2002).


LIVESTOCK
ENGINE/GENERATOR
Electricity
Biogas
(500 cow
(1100-1400 m3/day)
free-stall (60-80 kW
milking continuously)
facility) Heat
(~60% CH4,
~40% CO2
Manure <4000ppm H2S) (Used to maintain
digester temperature)
LIQUID/SOLID
SEPARATOR

ANAEROBIC DIGESTER
(1330 m3 plug flow
Liquids Solids

Compost

Irrigated Nutrients

LIQUID STORAGE LAGOON COMPOSTING FACILITY


(9000 m3 capacity) (Open windrow system)
Figure 2.1: Schematic of AA Dairy Integrated Farm Energy System
6

There are many benefits to such farm systems, which include (RDA 2000):

• Waste treatment benefits: Natural waste treatment process that requires less

land than composting, reduces solid waste volume and weight, and reduces

contaminant runoff.

• Energy benefits: Generates a high-quality renewable fuel, which has numerous

end-use applications.

• Environmental benefits: Potential to reduce carbon dioxide and methane

emissions, eliminates odors, produces a bio-available nutrient stream, and

maximizes recycling benefits. Reduces dependence on fossil fuels.

• Economic benefits: More cost effective than many other treatment options

when viewed from a life-cycle analysis. Typical payback periods of 4-8 years.

2.2. ANAEROBIC DIGESTION

Six to eight million family-sized low-technology digesters are used in China

and India to provide biogas for cooking and lighting. Also, there are over 800 farm-

based digesters operating in Europe and North America (Wellinger and Linberg 2000).

Farm-based anaerobic digestion in the U.S. has mainly focused on manures from dairy

and swine operations because they are often liquid or slurry based. Systems have been

designed for poultry manures, but the higher solids content results in precipitation of

solids unless constantly mixed. There are many types of anaerobic digestion systems

for manure, which include batch, mixed-tank, plug-flow, fixed-film, and lagoon

digesters. Table 2.1 describes the different characteristics of 3 typical farm digesters.
7

Table 2.1: Characteristics of Typical Agricultural Anaerobic Digesters


Covered Lagoon Complete Mix Plug Flow
Round/Square In ground
Vessel Deep lagoon
In/Above ground Tank rectangular tank
Level of
Low Medium Low
Technology
Additional Heat No Yes Yes
Total Solids 0.5-1.5% 3-11% <11%
HRT (days) 40-60 15+ 15+
Farm Type Dairy/Hog Dairy/Hog Dairy only
Optimum Climate Temperate/Warm All All
Source: Roos and Moser (2000), AgSTAR Handbook, pgs.1-2

There are two ways to derive methane from biomass, thermally and

biologically. Although thermal conversion is rapid and complete, it is limited in its

application to materials of low water content. This is because large amounts of energy

are needed to vaporize water before reaching substrate-gasification temperature.

Biological conversions, utilizing methanogenic bacteria, are advantageous because

they require less energy and can be applied to wet or dry feedstock on a variety of

scales. Unfortunately, anaerobic digestion is often slow, requires precise solids

loading and an anoxic environment, and is only about 50% effective in its conversion

of organic matter.

The microbial process of anaerobic digestion and methane production occurs

through the complex action of interdependent microbial communities as depicted in

Figure 2.2. The first step involves the hydrolysis of organic compounds, including

carbohydrates, proteins, and lipids, via hydrolytic bacteria. Here, the substrate is

broken down into usable-sized molecules such as organic acids, alcohols, neutral

compounds, hydrogen and carbon dioxide. The second stage, carried out by

transitional bacteria, consists of converting soluble organic matter into methanogenic

substrates such as hydrogen, carbon dioxide and acetate. Lastly, methanogenic


8

bacteria utilize these intermediates for conversion to methane and carbon dioxide

(Chynoweth and Issacson 1987).

Complex
Organic
Carbons HYDROLYTIC
BACTERIA
Organic
Acids,
Neutral
Compounds
TRANSITIONAL Hydrogen Producing
BACTERIA Acetogenic Bacteria

Homoacetogenic
Bacteria
H2, CO2, One- Acetic Acid
Carbon
Compounds

METHANOGENIC
BACTERIA

CH4 + CO2

Figure 2.2: Anaerobic Digestion Process


Source: Chynoweth and Issacson (1987), pg. 3

There are a number of factors which influence the digestion process, including,

temperature, bacterial consortium, nutrient composition, moisture content, pH, and

residence time.

Sulfur is an essential nutrient for methanogens but sulfur levels too high may

limit methanogenesis. Sulfur can enter the digester in the feedstock itself or from

chemicals used in an agricultural environment, such as copper and zinc sulfate

solutions that are used to prevent dairy cow foot-rot, and are inadvertently washed into
9

the digester. Farm animals consume sulfur either in their food source, mostly in the

form of sulfur-containing amino acids such as cystine and methionine, or from their

drinking water source, which may contain significant amounts of sulfate. Sulfur that

is not used by the animal for nutrition is excreted in the manure.

Sulfate-reducing bacteria actually can out-compete methanogens during the

anaerobic digestion process. Therefore, sulfide production generally proceeds to

completion before methanogenesis occurs. The energetics of sulfate reduction with H2


is favorable to the reduction of CO2 with H2, forming either CH4 or acetate (Madigan,

et al. 2000).

The toxic level of total dissolved sulfide in anaerobic digestion is reported as

200-300 mg/l. Also, a head gas concentration of 6% H2S is the upper limit for

methanogenesis, while 0.5% H2S (11.5 mg/l) is optimum (Chynoweth and Issacson

1987).

2.3. BIOGAS COMPOSITION

Biogas composition depends heavily on the feedstock, but mainly consists of

methane and carbon dioxide, with smaller amounts (ppm) of hydrogen sulfide and

ammonia. Trace amounts of organic sulfur compounds, halogenated hydrocarbons,

hydrogen, nitrogen, carbon monoxide, and oxygen are also occasionally present.

Usually, the mixed gas is saturated with water vapor and may contain dust particles

and siloxanes (Wellinger and Linberg 2000). Water-saturated biogas from dairy-

manure digesters consists primarily of 50-60% methane, 40-50% carbon dioxide, and

less than 1% sulfur impurities, of which the majority exists as hydrogen sulfide

(Pellerin, et al. 1987).


10

Hydrogen sulfide is poisonous, odorous, and highly corrosive. Some

characteristics of H2S are described in Table 2.2. Because of these characteristics,


hydrogen sulfide removal is usually performed directly at the gas-production site.

Table 2.2: Physical, Chemical and Safety Characteristics of Hydrogen Sulfide


Molecular Weight 34.08
Specific Gravity (relative to air) 1.192
Auto Ignition Temperature 250° C
Explosive Range in Air 4.5 to 45.5 %
Odor Threshold 0.47 ppb
8-hour time weighted average (TWA) (OSHA) 10 ppm
15-minute short term exposure limit (STEL) (OSHA) 15 ppm
Immediately Dangerous to Life of Health (IDLH) (OSHA) 300 ppm
Source: OSHA (2002), Occupational Safety and Health Administration, www.OSHA.gov

The actual amount of water vapor entrained in the gas depends on the gas

composition, pressure, and temperature. Approximately 25 kg of water is present in

1400 m3 of saturated natural gas at 21° C and atmospheric pressure (Kohl and Neilsen

1997).

2.4. QUALITY REQUIREMENTS FOR BIOGAS UTILIZATION

Biogas can be used for all applications designed for natural gas, assuming

sufficient purification. On-site, stationary biogas applications generally have fewer

gas processing requirements. A summary of potential biogas utilization technologies

and their gas processing requirements are given in Table 2.3.


11

Table 2.3: Biogas Utilization Technologies and Gas Processing Requirements


Technology Recommended Gas Processing Requirements
Heating H2S < 1000 ppm, 0.8-2.5 kPa pressure, remove condensate
(Boilers)1 (kitchen stoves: H2S < 10 ppm)
Internal H2S < 100 ppm, 0.8-2.5 kPar pressure, remove condensate,
Combustion remove siloxanes (Otto cycle engines more susceptible
Engines1 to H2S than diesel engines)
H2S tolerant to 70,000 ppm, > 350 BTU/scf, 520 kPa pressure,
Microturbines2
remove condensate, remove siloxanes
PEM: CO < 10 ppm, remove H2S
PAFC: H2S < 20 ppm, CO < 10 ppm, Halogens < 4 ppm
Fuel Cells3 MCFC: H2S < 10 ppm in fuel (H2S < 0.5 ppm to stack),
Halogens < 1 ppm
SOFC: H2S < 1 ppm, Halogens < 1 ppm
Stirling Engines4 Similar to boilers for H2S, 1-14 kPa pressure
H2S < 4 ppm, CH4 > 95%, CO2 < 2 % volume, H2O < (1*10-4)
Natural Gas
kg/MMscf, remove siloxanes and particulates, > 3000
Upgrade1,5
kPa pressure
Sources: 1 Wellinger and Linberg (2000)
2
Capstone Turbine Corp.(2002)
3
XENERGY (2002)
4
STM Power (2002)
5
Kohl and Neilsen (1997)

Technologies such as boilers and Stirling engines have the least stringent gas

processing requirements because of their external combustion configurations. Internal

combustion engines and microturbines are the next most tolerant to contaminants.

Fuel cells are generally less tolerant to contaminants due to the potential for catalytic

poisoning. Upgrading to natural-gas quality usually requires expensive and complex

processing and must be done when injection into a natural-gas pipeline or production

of vehicle fuel is desired.

Although not covered in this study, techniques for removal of CO2 may also

simultaneously reduce H2S levels. Many facilities in Europe have utilized water
scrubbing, polyethylene glycol scrubbing, carbon molecular-sieves or membranes for

upgrading of biogas to natural gas or vehicle fuel. Readers are directed to the
12

following references for more information on these systems: Kohl and Neilsen (1997),

Wellinger and Linberg (2000), CADDET (2001), Eriksen, et al. (1999), (Schomaker,

et al. (2000), and Jensen and Jensen (2000).

2.5. TRADITIONAL H2S GAS-PHASE REMOVAL METHODS

Since biogas is similar in composition to raw natural gas, purification

techniques developed and used in the natural-gas industry can be evaluated for their

suitability with biogas systems. The ultimate process chosen is dependent on the gas

use, composition, physical characteristics, energy and resources available, byproducts

generated, and the volume of gas to be treated.

Principal gas phase impurities that may be present are listed in Table 2.4

below. Other constituents that may be problematic include water or other

condensates, and particulate matter. Hydrocarbons, such as methane, are the desired

product gases.

Table 2.4: Principal Gas Phase Impurities


1. Hydrogen sulfide
2. Carbon dioxide
3. Water vapor
4. Sulfur dioxide
5. Nitrogen oxides
6. Volatile organic compounds (VOC’s)
7. Volatile chlorine compounds (e.g., HCl, Cl2)
8. Volatile fluorine compounds (e.g., HF, SiF4)
9. Basic nitrogen compounds
10. Carbon monoxide
11. Carbonyl sulfide
12. Carbon disulfide
13. Organic sulfur compounds
14. Hydrogen cyanide
Source:Kohl and Neilsen (1997), pg 3
13

Gas purification processes generally fall into one of the following five

categories: 1) Absorption into a liquid; 2) Adsorption on a solid; 3) Permeation

through a membrane; 4) Chemical conversion to another compound; or 5)

Condensation (Kohl and Neilsen 1997).

For the purposes of process comparison, gas characteristics similar to those at

AA Dairy, which are typical for a farm digester treating waste from around 500 dairy

cows, will be assumed and summarized as shown in Table 2.5.

Table 2.5: Assumed Biogas Characteristics for Process Comparisons


Gas composition: ~60% CH4
~40% CO2
1000-4000 ppm H2S
Gas flow rate: ~1400 m3/day
Gas pressure: < 2 kPa
Gas temperature: ~ 25°C
Water saturated: Yes

With the flow rate and sulfur levels above, 1.9 – 7.7 kg of H2S are present in

the gas stream daily, or 690 – 2,815 kg yearly. Desirable attributes for a gas

purification system include low capital and operating costs, ease of operation and

media disposal, and minimal material and energy inputs. H2S removal processes will
be divided into dry-based, liquid-based, physical-solvent, membrane, alternative, and

biological processes for this summary. Media disposal costs are not discussed here

but very well may be the most significant costs for a project. For a further discussion

of this point, see Appendix A.

2.5.1. Dry H2S Removal Processes

Dry H2S removal techniques have historically been used at facilities with less

than 200kg S/day in the U.S. All of the dry sorption processes discussed here are
14

configured with the dry media in box or tower type vessels where gas can flow

upwards or downwards through the media. Since all of the dry-sorption media to be

discussed eventually becomes saturated with contaminant and inactive, it is common

to have two vessels operated in parallel so one vessel can remain in service while the

other is offline for media replacement.

2.5.1.1. Iron Oxides

As one of the oldest methods still in practice, iron oxides remove sulfur by

forming insoluble iron sulfides. It is possible to extend bed life by admitting air,

thereby forming elemental sulfur and regenerating the iron oxide, but eventually the

media becomes clogged with elemental sulfur and must be replaced. The most well-

known iron oxide product is called “iron sponge.” Recently, proprietary iron-oxide

media such as SulfaTreat®, Sulfur-Rite®, and Media-G2® have been offered as


improved alternatives to iron sponge.

Iron Sponge

Iron-oxide-impregnated wood-chips (generally pine) are used to selectively

interact with H2S and mercaptans. The primary active ingredients are hydrated iron-

oxides (Fe2O3) of alpha and gamma crystalline structures. Lesser amounts of Fe3O4

(Fe2O3.FeO) also contribute to the activity (Anerousis and Whitman 1985). Typical

specifications for iron sponge are listed below in Table 2.6. Grades of iron sponge

with 100, 140, 190, 240 and 320 kg Fe2O3/m3 are traditionally available, with the 190

kg Fe2O3/m3 (15-lb/bushel) grade being the most common. Bulk density for this grade

is consistently around 800 kg/m3 (50 lb/ft3) in place (Revell 2001).


15

Table 2.6: Typical Specifications for 15-lb Iron Sponge


Source: Kohl and Neilsen,(1997), pg. 1302

The chemical reactions involved are shown in Equations 2.1-2.2:(Crynes 1978)

Fe2O3 + 3H2S Æ Fe2S3 + 3H2O ∆H= -22 kJ/g-mol H2S (2.1)

2Fe2S3 + O2 Æ 2Fe2O3 + 3S2 ∆H= -198 kJ/g-mol H2S (2.2)

As seen from Equation 2.1, one kg of Fe2O3 stochiometrically removes 0.64 kg

of H2S. Equation 2.2 represents the highly exothermic regeneration of iron oxide and
16

formation of elemental sulfur upon exposure to air. Iron sponge is also capable of

removing mercaptans via Equation 2.3: (Zapffe 1963)

Fe203 + 6RSH = 2Fe(RS)3 + 3H20 (2.3)

Iron sponge can be operated in batch mode with separate regeneration, or with

a small flow of air in the gas stream for continuous revification. In batch mode,

operational experience indicates that only about 85% (0.56 kg H2S/ kg Fe2O3) of the

theoretical efficiency can be achieved (Taylor 1956).

Spent iron sponge can be regenerated in place by recirculation of the gas in the

vessel adjusted to 8% O2 concentration and 0.3-0.6 m3/m3bed/min space velocity

(Taylor 1956). Alternatively, the sponge can be removed, spread out into a layer 0.15-

m thick, and kept continually wetted for 10 days. It is imperative to manage the heat

buildup in the sponge during regeneration to maintain activity and prevent combustion

(Revell 1997). Due to buildup of elemental sulfur and loss of hydration water, iron

sponge activity is reduced by 1/3 after every regeneration. Therefore, regeneration is

only practical once or twice before new iron sponge is needed.

Removal rates as high as 2.5 kg H2S/ kg Fe2O3 have been reported in

continuous-revivification mode with a feed-gas stream containing only a few tenths of

a percent of oxygen (Taylor 1956). Equation 2.4 can be used to calculate percent air

recirculation necessary for optimum performance, dependent on inlet H2S


concentration in the gas (Vetter et al. 1990).

% Air recirculation required = 1.90 +(mg/m3 H2S measured)/3024 (2.4)


17

At Huntington’s farm in Cooperstown, N.Y., a removal level of 1.84 kg

H2S/kg Fe2O3 was reported using 140 kg Fe2O3/m3 (12 lb/bushel)-grade sponge and
continuous revivification with 2.29% air recirculation (Vetter et al. 1990).

Because iron sponge is a mature technology, there are design parameter

guidelines that have been determined for optimum operation. Table 2.7, below, is a

comprehensive collection of published design criteria for iron sponge systems.

Table 2.7: Iron Sponge Design Parameter Guidelines


Vessels: Stainless-steel box or tower geometries are recommended for ease of
handling and to prevent corrosion. Two vessels, arranged in series are
suggested to ensure sufficient bed length and ease of handling (Lead/Lag).
Gas Flow: Down-flow of gas is recommended for maintaining bed moisture. Gas
should flow through the most fouled bed first.
Gas Residence Time: A residence time of greater than 60 seconds, calculated using
the empty bed volume and total gas flow, is recommended.1
Temperature: Temperature should be maintained between 18° C and 46° C in
order to enhance reaction kinetics without drying out the media.2
Bed Height: A minimum 3 m (10 ft) bed height is recommended for optimum H2S
removal. A 6 m bed is suggested if mercaptans are present.3 A more
conservative estimate recommends a bed height of 1.2 to 3 meters.4
Superficial Gas Velocity: The optimum range for linear velocity is reported as 0.6-3
m/minute.3
Mass Loading: Surface contaminant loading should be maintained below 10 g
S/min/m2 bed.4
Moisture Content: In order to maintain activity, 40% moisture content, plus or
minus 15%, is necessary. Saturating the inlet gas helps to maintain this.2
pH: Addition of sodium carbonate can maintain pH between 8-10. Some sources
suggest addition of 16 kg sodium carbonate per m3 of sponge initially to
ensure an alkaline environment.2
Pressure: While not always practiced, 140 kPa is the minimum pressure
recommended for consistent operation.3
Sources: 1 Revell (2001), 2 Kohl and Neilsen (1997), 3 Anerousis and Whitman (1985),
4
Maddox and Burns (1968), 5 Taylor (1956)

Using the design constraints described in Table 2.7, a suitable iron sponge

system can be designed for a generic farm biogas application with characteristics

shown in Table 2.5. These results are presented in Table 2.8 below.
18

Table 2.8: System Characteristics of 15-lb Iron Sponge Design at AA Dairy


Number of Vessels 2 in series (Lead/Lag)
Vessel Dimensions 0.91 m diameter x 1.52 m high
Empty Bed Residence Time 120 seconds total
Gas Flow Rate 0.94 m3/min
Mass of Sponge 800 kg each
Air Recirculation Rate 2.4% - 3.7%
Performance Estimates
Low Loading High Loading
(1000 ppm H2S) (4000 ppm H2S)

Expected Bed Life 72-315 days 18-79 days


Annual Sponge Consumed 930-4070 kg 3,710 – 16,300 kg
Annual Sponge Costs $250 -$1,075 $985-$4,300

Biogas operations currently using iron sponge are located in Cooperstown,

NY, Little York, NY, and Chino, CA, among others. H2S levels at one farm digester

were consistently reduced from as high as 3600 ppm (average 1350 ppm) to below 1

ppm using a 1.5 m diameter x 2.4 m deep iron sponge reactor (Vetter et al. 1990).

Commercial sources for iron sponge include Connelly GPM, Inc., of Chicago,

IL, and Physichem Technologies, Inc., of Welder, TX. Both companies provide media

for around $6 per bushel (~50 lb), and note that shipping costs may be more

significant than actual media costs. Varec Vapor Controls, Inc., sells their Model-235
treatment units for around $50,000, including the cost of initial media. Such a unit

could last up to two years before change-out would be necessary (Wang 2000).

While the benefits of using iron sponge include simple and effective operation,

there are critical drawbacks to this technology that have lead to decreased usage in

recent years. The process is highly chemical intensive, the operating costs can be

high, and a continuous stream of spent waste material is accumulated. Additionally,

the change-out process is labor intensive and can be troublesome if heat is not

dissipated during regeneration. Perhaps most importantly, safe disposal of spent iron

sponge has become problematic, and in some instances, spent media may be
19

considered hazardous waste and require special disposal procedures. Landfilling on-

site is still practiced, but has become riskier due to fear of the need for future

remediation.

SulfaTreat®

SulfaTreat® is a proprietary sulfur scavenger, consisting mainly of Fe2O3 or

Fe3O4 compounds coated onto a proprietary granulated support and marketed by the

SulfaTreat® Company of St. Louis, MO. SulfaTreat® is used similarly to iron sponge
in a low-pressure vessel with down-flow of gas and is effective with partially or fully

hydrated gas streams.

Conversion efficiency in commercial systems is in the range of 0.55 - 0.72 kg

H2S/kg iron oxide, which is similar to, or slightly higher than, values reported for

batch operation of iron sponge (Kohl and Neilsen 1997). Particles range in size from

4 to 30 mesh with a bulk density of 1120 kg/m3 in place, and sell for roughly $0.88/kg

(Taphorn 2000).

Multiple benefits over iron sponge are claimed due to uniform structure and

free-flowing nature. SulfaTreat® is reportedly easier to handle than iron sponge, thus

reducing operating costs, labor for change-out, and pressure drops in the bed. Also,

SulfaTreat® claims to be non-pyrophoric when exposed to air and thus does not pose a

safety hazard during change-out. Buffering of pH and addition of moisture are not

necessary as long as the inlet gas is saturated.

Drawbacks associated with this product are similar to iron sponge; the process

is non-regenerable, chemically intensive, and spent product can be problematic or

expensive to dispose of properly. The manufacturer has suggested that spent product

may be used as a soil amendment or as a raw material in road or brick making, but
20

they state that every customer must devise a spent-product disposal plan in accordance

with local and state regulations.

For AA Dairy, a two-vessel arrangement (series) is proposed by the SulfaTreat

Company to ensure maximum removal while maintaining manageable bed sizes.

Proprietary rectangular vessels in a “Lead/Lag” arrangement, with the most fouled bed

contacting the gas first, are used (Taphorn 2000). Transportation, installation, and

disposal costs are not included in the system as described in Table 2.9 below.

Table 2.9: System Characteristics of SulfaTreat® Design at AA Dairy


Number of Vessels 2 in series (Lead/Lag)
Vessel Dimensions 1.22 m x 1.65 m x 1.83 m
Vessel Costs $8,000 for two
Gas Flow Rate 0.94 m3/min
®
Mass of SulfaTreat 3,636 kg each
Air Recirculation Rate 2.4%
Performance Estimates
Low Loading High Loading
(1000 ppm H2S) (4000 ppm H2S)

Expected Bed Life (one vessel) 345 days 86 days


Total Pressure Drop (kPa) 0.4 0.4
Annual SulfaTreat® Consumed 3,850 kg 15,450 kg
Annual SulfaTreat® Costs $3,400 $13,500

Sulfur-Rite®

Sulfur-Rite® is also a dry-based iron-oxide product offered by GTP-Merichem.

Sulfur-Rite® is unique in their claim that insoluble iron pyrite is the final end product.

Sulfur-Rite® systems come in prepackaged cylindrical units that are recommended for

installations with less than 180 kg sulfur/day in the gas and flow rates below 70

m3/min. Company literature claims spent product is non-pyrophoric and landfillable

and has 3-5 times the effectiveness of iron sponge. Sulfur-Rite® also has many of the
21

disadvantages of the iron-oxide scavengers previously mentioned. System design and

cost estimates for an installation similar to AA dairy are presented in Table 2.10.

Table 2.10: System Characteristics of Sulfur-Rite® Design at AA Dairy


Number of Vessels 1-Carbon Steel unit
Vessel Dimensions 2.29 m diameter x 3.43 m high
Vessel Costs $43,600 (vessel only)
Gas Flow Rate 0.94 m3/min
Mass of Sulfur-Rite® 9,100 kg
Performance Estimates
Low Loading High Loading
(1000 ppm H2S) (4000 ppm H2S)

Expected Bed Life 420 days 98 days


Annual Product Consumption 7,900 kg 33,900 kg
Annual Sulfur-Rite® Costs $5,560 $23,840

Media-G2®

Media-G2® is an iron-oxide-based adsorption technology originally developed

by ADI International, Inc., for removal of arsenic from drinking water. Recently ADI

has begun testing Media-G2® for the removal of H2S from gas streams with promising

results. Landfill gas and biogas installations will serve as the primary market for their

technology, which incorporates iron oxides onto a diatomaceous support.

Lab scale and pilot scale trials indicate that treatment of up to 30,000 ppm H2S
is possible, spent product is non-hazardous, and Media-G2® can remove up to 560 mg

H2S/g solid. This is achieved by being able to regenerate the matrix with air up to 15

times. Each adsorption cycle removes about 35-40 mg H2S/g media. A two-vessel

system design (parallel) is recommended for continuous operation, as 8-hour

regeneration cycles are estimated at full scale. Vessels are designed for approximately

60-second empty-bed residence times. Approximate product costs are estimated at

$1060/m3.
22

Only two full-scale plants have been installed to date; Brookhaven Landfill in

NY, and a farm based anaerobic digester installed by Enviro-Energy Corporation in

Tillamook, OR (McMullin 2002). Although no full scale operational results were

available, a system design summary is proposed in Table 2.11 below.

Table 2.11: System Characteristics of Media-G2® Design at AA Dairy


Number of Vessels 2 in parallel
Vessel Dimensions 0.91 m diameter x 1.52 m high
Gas Flowrate 0.94 m3/min
Empty Bed Residence Time 62 seconds (with one offline)
Mass of Media-G2® 760 kg each
Air Recirculation Rate 2.4%
Performance Estimates
Low Loading High Loading
(1000 ppm H2S) (4000 ppm H2S)

Expected Bed Life (one vessel) 190 days 47 days


Annual Media-G2® Consumption 1,460 kg 5,900 kg
Annual Media-G2® Costs $2,050 $8,290

2.5.1.2. Zinc Oxides

Zinc oxides are preferred for removal of trace amounts of hydrogen sulfide

from gases at elevated temperatures due to their increased selectivity over iron oxide

(Chiang and Chen 1987). Typically in the form of cylindrical extrudates 3-4 mm in

diameter and 8-10 mm in length, zinc oxides are used in dry-box or fluidized-bed

configurations. Hydrogen sulfide reacts with zinc oxide to form an insoluble zinc

sulfide via Equation 2.5 (Kohl and Neilsen 1997).

ZnO + H2S = ZnS + H2O (2.5)

The equilibrium constant for the reaction is given with Equation 2.6.
23

Kp = PH2O/PH2S (2.6)

Where: PH2O is the partial pressure of water vapor in the gas phase
PH2S is the partial pressure of hydrogen sulfide in the gas phase

As shown in Figure 2.3, the equilibrium constant decreases rapidly with

temperature. Therefore, at very high temperatures equilibrium is approached, but as

temperature decreases, reaction kinetics are drastically reduced to impractical levels.

Figure 2.3: Equilibrium Constant for the Reaction ZnO + H2S = ZnS + H2O.
Source: Kohl and Neilsen (1997) pg. 1307.

Zinc-oxide processes are available in several forms for operation at

temperatures from about 200° C to 400° C. Maximum sulfur loading is typically in

the range of 30-40 kg sulfur/100 kg sorbent for these processes. Puraspec®, marketed

by IC Industries of Great Britain, is a proprietary combination of zinc oxides that

boasts more effective performance in the temperature range of 40° C to 200° C.

Nevertheless, performance is preferable at 200° C to 40° C, so operation below 150° C

is rarely practiced. Spent product may also contain over 20% sulfur (by weight).

Formation of zinc sulfide is irreversible and zinc oxide is not very reactive with
24

organic sulfur compounds. If removal of mercaptans is also desired, catalytic

hydrodesulfurization to convert these compounds to the more reactive hydrogen

sulfide is needed first (Kohl and Neilsen 1997).

2.5.1.3. Alkaline Solids

Alkaline substances, such as hydrated lime, will react with acid gases like H2S,
SO2, CO2, carbonyl sulfides and mercaptans in neutralization reactions. Usually

liquid-based scrubbers are used, but fixed-beds of alkaline granular solid can also be

used in a standard dry box arrangement with up-flow of gas. Molecular Products Ltd.,

of Great Britain, markets a product called Sofnolime-RG®, which is claimed to be a

synergistic mixture of hydroxides that react with acid gases. Predominant reactions

are shown in Equations 2.7-2.8 (Kohl and Neilsen 1997)

2NaOH + H2S Æ Na2S +2H20 (2.7)

Ca(OH)2 + CO2 Æ CaCO3 + H2O (2.8)

To achieve significant removal of H2S, CO2 must also be concurrently reduced

at the cost of extremely high product utilization. Sofnolime® can remove about 180 L
of CO2/kg of media. At this efficiency, it would require over 3,020 kg/day of
Sofnolime® to remove all of the CO2 from 1350 m3 biogas/day, assuming 40% CO2

concentration by volume.

2.5.1.4. Adsorbents

Adsorbents rely on physical adsorption of a gas-phase particle onto a solid

surface, rather than chemical transformation as discussed with the previous dry

sorbents. High porosity and large surface areas are desirable characteristics, enabling
25

more physical area for adsorption to occur. Media eventually becomes saturated and

must be replaced or regenerated. If regeneration of the media is economical or

desirable, it can be achieved by using one of the processes described in Table 2.12

below. During regeneration, H2S rich gas is released and must be exhausted
appropriately or subjected to another process for sulfur recovery (Yang 1987).

Table 2.12: Processes for Adsorbent Regeneration


Regeneration
Description
Process
Regeneration takes place primarily through heating. The
Temperature differences between the equilibrium loadings at the two
Swing Adsorption temperatures represent net removal capacity. Considerable
(TSA) energy and time are required to heat and cool the bed. TSA is
often achieved by preheating a purge gas.
Regeneration is achieved by lowering the pressure of the bed
and allowing the adsorbate to desorb. Typically adsorption
Pressure Swing
takes place at elevated pressures to allow for regeneration at
Adsorption (PSA)
atmospheric pressure or under slight vacuum. PSA is
relatively fast compared to TSA
A non-adsorbing gas containing very little of the impurity is
Inert Purge passed through the bed, reducing the partial pressure of
adsorbate in the gas-phase so that desorption occurs.
A purge gas that is more strongly adsorbed than the impurity is
Displacement used to desorb the original contaminant. Steam regeneration,
Purge while mostly a thermal process, also regenerates through
displacing some of the original adsorbate.

Molecular Sieves (Zeolites)

Zeolites are naturally occurring or synthetic silicates with extremely uniform

pore sizes and dimensions and are especially useful for dehydration or purification of

gas streams. Polar compounds, such as water, H2S, SO2, NH3, carbonyl sulfide, and

mercaptans, are very strongly adsorbed and can be removed from such non-polar

systems as methane. About 40 different zeolite structures have been discovered and

properties of the four most common ones are described in the Table 2.13.
26

Table 2.13: Basic Types of Commercial Molecular Sieves


Source: Kohl and Neilsen, (1997), pg. 1043

Adsorption preference, from high to low, is: H2O, mercaptans, H2S, and CO2.

Not all mercaptans are adsorbable on type 4A or 5A molecular sieves because of pore

size limitations. Consequently, 13X is preferred for complete sulfur removal from

natural-gas streams. Because contaminants are essentially competing for the same

active adsorption spots, a graphical representation of multiple adsorption zones in a

molecular sieve bed might occur as in Figure 2.4.


27

Figure 2.4: Adsorption Zones in a Molecular Sieve Bed, Adsorbing Both Water
Vapor and Mercaptans from Natural Gas.
Source: Kohl and Nielsen, (1997), pg 1071

A design method for natural-gas purification by 5A molecular sieves,

developed by Chi and Lee (1973), can be used to estimate approximate bed-sizes and

media-life for a zeolites process at AA Dairy. Minimum pressures of 3500 kPa, and

maximum CO2 concentration of 5%, were verified for their model, but for the
following calculations a 40% CO2 concentration is used (Chi and Lee 1973). Table

2.14 shows characteristics for a sample 5A-molecular-sieve system for AA Dairy.


28

Table 2.14: Summary of 5A Molecular Sieve Design at AA Dairy


Low Loading High Loading
(1000 ppm H2S) (4000 ppm H2S) (units)
Gas Flow rate 1,400 1,400 m3/day
Operating Pressure 500 500 psig
Operating Temp. 25 25 °C
Bed Life 24 24 hours
Bed Height 1.4 2.0 m
Bed Diameter 0.6 0.6 m
Bed Volume 0.39 0.58 m3
Bed Wt. 262 391 kg

As calculated, roughly 250-400 kg of zeolite would be needed on a daily basis,


and therefore would not be economical without a regeneration process.

Activated Carbon

Granular activated carbon (GAC) is a preferred method for removal of volatile

organic compounds from industrial gas streams. Heating carbon-containing materials

to drive off volatile components forms GAC’s, which have a highly porous adsorptive

surface. Utilization of GAC’s for removal of H2S has been limited to removing small

amounts, and primarily from drinking water. If H2S is the selected contaminant to be

removed, GAC’s impregnated with alkaline or oxide coatings are utilized.

Impregnated Activated Carbons

Coating GAC’s with alkaline or oxide solids enhance the physical adsorptive

characteristics of the carbon with chemical reaction. Sodium hydroxide, sodium

carbonate, potassium hydroxide (KOH), potassium iodide, and metal oxides are the

most common coatings employed.

Distributors of impregnated activated carbon include Calgon Carbon


Corporation (Type FCA® carbon), Molecular Products, Ltd. (Sofnocarb KC®), US
29

Filter-Westates, and Bay Products, Inc. Typically, 20-25% loading by weight of H2S
can be achieved, which is up from 10% as seen with regular GAC.

An example of particular interest was the use of a non-regenerable KOH-

activated-carbon bed (Westates) for removal of H2S from anaerobic-digester and

landfill gas for use in a fuel cell. Oxygen (0.3-0.5% by volume) was added to facilitate

conversion of H2S to elemental sulfur. Two beds, 0.6 m in diameter by 1.5 m high,

were piped in series and run with space velocities of 5300/hr. Inlet H2S concentration

ranged from 0.7-50 ppm, averaging 24.1 ppm, and 98+% removal was demonstrated.

A loading capacity of 0.51 g S/g carbon was reported, which is substantially greater

than the normally reported range of 0.15 - 0.35 g S/g carbon for KOH-carbon. Media

costs were estimated at $5/kg for the adsorbent. Pretreatment system capital costs

(including sulfur removal, blowers and coalescing filters) were estimated to be

$500/kW (Spiegel, et al. 1997; Spiegel and Preston 2000).

Assuming loading capability of 25% and design with a 100 kW generator,

costs and performance might appear as represented in Table 2.15.

Table 2.15: System Characteristics for KOH-Impregnated


Activated Carbon at AA Dairy
Number of Vessels 2 in series (Lead/Lag)
Vessel Dimensions 0.6 m diameter x 1.5 m high
System Capital Cost $50,000
Gas Flow Rate 0.94 m3/min
Mass of Carbon 250 kg each
O2 Recirculation Rate 0.3%
Performance Estimates
Low Loading High Loading
(1000 ppm H2S) (4000 ppm H2S)

Expected Bed Life (one vessel) 340 days 85 days


Annual Carbon Consumption 270 kg 1075 kg
Annual Carbon Costs $1,250 $5,435
30

2.5.2. Liquid H2S Removal Processes

Liquid-based H2S removal processes have replaced many dry-based

technologies for natural-gas purification due to reduced ground-space requirements,

reduced labor costs, and increased potential for elemental-sulfur recovery. Gas-liquid

contactors, or absorbers, are used which increase surface area and optimize gas contact

time. If a reversible reaction is employed, regeneration columns are operated in

conjunction with the absorber to facilitate continuous processing. A generic

absorber/regenerator flow scheme is presented in Figure 2.5.

Stripping Stripping
Solution Gas Out
Clean Gas Out

Sour Gas In Stripping


Gas In

Figure 2.5: Generic Absorber/Stripper Schematic

As indicated, the stripper gas contains the displaced H2S if it has not been
converted to elemental sulfur in the process. When the sulfide level is high, the sour

stripping gas can be sent to a Claus plant for elemental-sulfur recovery. When the

reaction is irreversible, a simpler bubble column may be used in place of an absorber.

Liquid-based H2S removal processes can be grouped into liquid-phase oxidation

processes, alkaline-salt solutions, and amine solutions. Physical adsorption of H2S

into a liquid, such as water, is discussed in the next section.


31

2.5.2.1. Liquid-Phase Oxidation Processes

Iron- and Zinc-Oxide Slurries

Iron-oxide slurry processes historically mark the transition between dry-box

technologies and modern liquid-redox processes. The basic chemistry is similar to

that for the dry oxide reactions. H2S is reacted with an alkaline compound in solution

and then iron oxide to form iron sulfide, as shown in Equations 2.9-2.10.

Regeneration is achieved by aeration, converting the sulfide to elemental sulfur, as


shown in Equation 2.11 (Kohl and Neilsen 1997).

H2S + Na2CO3 = NaHS + NaHCO3 (2.9)

Fe2O3.3H2O + 3NaHS + 3 NaHCO3 = Fe2S2.3H2O + 3 Na2CO3 + 3H2O (2.10)

2Fe2S2.3H2O +3O2 = 2Fe2O3.3H2O + 6S (2.11)

Several side reactions can occur, forming thiosulfates and thiocyanates, which

continually deplete the active iron oxide supply. Commercial processes that were

available in the past include the Ferrox (1926), Gluud (1927), Burkheiser (1953),

Manchester (1953), and Slurrisweet (1982) processes (Kohl and Neilsen 1997).

A zinc-oxide liquid-based process, known as Chemsweet® (Natco, Inc.), has


achieved some success in more recent years. The proprietary powder, consisting of

zinc oxide, zinc acetate, and dispersant, is mixed with water and used in a simple

bubble column. The reaction mechanisms are presented in Equations 2.12-2.14 below

(Kohl and Neilsen 1997).

ZnAc2 + H2S = ZnS +2HAc (2.12)

ZnO + 2HAc = ZnAc2 + H2O (2.13)

ZnO + H2S = ZnS +H2O (2.14)


32

Low pH is maintained to avoid CO2 absorption and vessel corrosion while


encouraging RSH and COS removal. Pipeline-gas specifications are easily met, but

the high cost of non-regenerable reactant usually limits use of this process to removing

trace amounts of sulfur.

Quinone and Vanadium Metal Processes

The redox cycle shown in Figure 2.6 depicts how hydrogen sulfide is

converted to elemental sulfur using quinones.(Kohl and Neilsen 1997)

+ H2S +S
Reduction

+ O2 + H2O
Oxidation

Figure 2.6: Reduction-Oxidation Cycle of Quinones

Processes using quinones with vanadium salts, such as the Stretford process,

account for a large portion of the liquid-based natural-gas purification market today,

although chelated-iron processes are surpassing them. Because of high capital and

operating costs and significant thiosulfate byproduct formation, quinone-based H2S


technologies are generally not used for smaller gas streams.

Chelated-Iron Solutions

Chelated-iron solutions utilize iron ions bound to a chelating agent and are

gaining popularity for H2S removal. The LO-CAT® (US Filter/Merichem) and

SulFerox® (Shell) processes currently dominate the chelated-iron H2S removal market.
Basic redox reactions employed for adsorption and regeneration are as shown in

Equations 2.15-2.16.
33

2Fe3+ + H2S = 2Fe2+ + S + 2H+ (2.15)

2Fe2+ +(1/2)O2 + H2O = 2Fe3+ + 2OH- (2.16)

The LO-CAT® process is potentially attractive for biogas applications because


it is 99+% effective, the catalyst solution is nontoxic, and it operates at ambient

temperatures, requiring no heating or cooling of the media. Multiple configurations of

the LO-CAT® process are available and Figure 2.7 below depicts a standard system.

Figure 2.7: Conventional Flow Diagram for LO-CAT® Process


Source: Kohl and Nielsen (1997), pg 809.

LO-CAT® systems are currently only recommended and economical for

facilities with over 200 kg S/day. Landfills and wastewater treatment plant digesters

have implemented LO-CAT® H2S removal systems successfully, and LO-CAT® plants

producing less than 500 kg of S/day are designed to produce thickened slurry, so use

of a separate thickener vessel is not required. The thickened slurry may have some

value as a fertilizer amendment in certain agricultural applications. The two principal


34

operating costs are for power for pumps and blowers, and chemicals for catalyst

replacement due to losses via thiosulfate and bicarbonate production (Kohl and

Neilsen 1997).

Le Gaz Integral Enterprise of France markets the Sulfint® and SulFerox® iron-
chelate processes targeted for gas streams with 100-20,000 kg S/day and high

CO2/H2S ratios. CO2 will not be removed significantly and 50% -90% of mercaptans

can be removed with either low or high-pressure applications. Sulfur removal with

SulFerox® costs around $0.24-$0.3 per kg, and filtration using a plate-and-frame filter

is sufficient to recover elemental sulfur (Smit and Heyman 1999).

Other Liquid-Based Processes

Nitrite solutions are sometimes used when simple process configurations are

desired, requiring only a bubble-column contactor and mist eliminator. An overall

reaction is represented with Equation 2.17.

3H2S + NaNO2 = NH3 + 3S + NaOH + some NOx (2.17)

In the presence of CO2, the NaOH is neutralized to produce sodium carbonate


and bicarbonate. As seen, the reaction products are ammonia and NOx, which may be
just as problematic as H2S to deal with. Nevertheless, the spent slurry is non-

hazardous and non-corrosive, the equipment is simple and low cost, and change-out of

spent adsorbent is easy. Sulfa-Check® (NL Industries, Inc.) and Hondo HS-100®

(Hondo Chemicals, Inc.) are two commercially available nitrite-based media. Design

guidelines include: (Kohl and Neilsen 1997)

1.) Optimum efficiency in the temperature range of 24° C to 43° C.

2.) Maximum superficial velocity of gas should be below 0.05 m/sec.

3.) 6.3×10-6 liters of solution are required per m3 of gas per ppm of H2S.
35

4.) Liquid height in meters should be 0.76 times the logarithm of H2S
concentration in ppm.

Using these criteria and gas characteristics described in Table 2.5, a vessel 0.61

meters in diameter, and 2.3 – 2.7 meters in liquid height should be employed.

Permanganate and dichromate solutions can also be used to completely remove

traces of H2S. Spent media is also non-regenerable and the high costs of chemicals

limit the use of this process.

2.5.2.2. Alkaline Salt Solutions

As with alkaline solids, acid gases such as H2S and CO2 react readily with

alkaline salts in solution. Regenerative processes employ alkaline salts including

sodium and potassium carbonate, phosphate, borate, aresenite, and phenolate, as well

as salts of weak organic acids. Since H2S is adsorbed more rapidly than CO2 by

aqueous alkaline solutions, some partial selectivity can be attained when both gases

are present by ensuring fast contact times at low temperatures (Kohl and Neilsen

1997).

Caustic Scrubbing

Hydroxide solutions are very effective at removing CO2 and H2S, but are non-

regenerable. Mercaptans form less-strongly-bound mercaptides, which are

regenerable at high temperatures, and commercial caustic-plants have operated with

this specialty.

The Dow Chemical Company developed a low-residence-time absorber for the

selective removal of H2S. Tests indicated reduction of 1000 ppm H2S to less than 100

ppm (in the presence of 3.5% CO2 @ 1400 m3/day), with a gas-residence time of 0.02

sec, pressure drop of 14 kPa, and liquid-to-gas ratio of 0.004 l/m3. Disposal of the
36

liquid effluent was a major problem. Also, the presence of higher CO2 concentrations
would lead to higher chemical utilization.

Other Alkaline Salt Processes

The Seaboard process (ICF Kaiser) was the first commercially applied liquid

process for H2S removal and used a sodium-carbonate absorbing-solution with air

regeneration. The overall chemical reaction is shown in Equation 2.18:

Na2CO3 + H2S = NaHCO3 + NaHS (2.18)

Removal efficiencies of 85% – 95% were realized, but the occurrence of side

reactions and problems with disposal of the foul air, containing H2S, has restricted use
of this process. Variations on the Vacuum Carbonate process (ICF Kaiser), which also

employ carbonates, have replaced the Seaboard process by enabling vacuum capture

of the foul stripping-gas and reducing the steam requirement needed for regeneration.

Many other processes are available at ambient and elevated temperatures that

use alkaline-salt solutions for removal of CO2 and H2S from gas streams. However,

the complexity of these processes makes them unattractive for H2S removal from

small biogas streams.

2.5.2.3. Amine Solutions

Amine processes constitute the largest portion of liquid-based natural-gas

purification technologies for removal of acid gases. They are attractive because they

can be configured with high removal efficiencies, designed to be selective for H2S or

both CO2 and H2S, and are regenerable. Drawbacks of using an amine system, as with

most liquid-based systems, are more complicated flow schemes, foaming problems,

chemical losses, higher energy demands, and how to dispose of foul regeneration air.
37

Alkanolamines generally contain a hydroxl group on one end and an amino

group on the other. The hydroxyl group lowers the vapor pressure and increases water

solubility, while the amine group provides the alkalinity required for absorption of

acid gases. The dominant chemical reactions occurring are as shown in Equations

2.19–2.23 (Kohl and Neilsen 1997).

H2O = H+ + OH- (2.19)

H2S = H+ + HS- (2.20)


CO2 + H2O = HCO3- + H+ (2.21)

RNH2 + H+ = RNH3+ (2.22)

RNH2 + CO2 = RNHCOO- + H+ (2.23)

Typically used amines include monothanolamine (MEA), diethanolamine

(DEA), methyldiethanloamine (MDEA), and diisopropanolamine (DIPA). Adsorption

is typically conducted at high pressures with heat regeneration in the stripper. Glycol

solutions, mentioned in the next section, are also employed to enhance physical

absorption characteristics of the acid gases. The basic flow-scheme for an

alkanolamine acid-gas removal process is depicted in Figure 2.8.


38

Figure 2.8: Flow Scheme for Alkanolamine Acid-gas Removal Processes


Source: Kohl and Nielsen (1997), pg 58

Sulfa-Scrub® (Quaker Chemical Company) is a triazine-based sorbent


developed to selectively remove H2S from gas streams with minimal corrosion and

non-hazardous spent media. Sulfa-Scrub® has been used in scavenging applications

without regeneration, and media consumption was around 5.3×10-6 - 6.7×10-6 l/m3 per

ppm of H2S in the feed gas. This corresponds to generation of 10-40 liters per day of
spent non-regenerable slurry from an operation similar to AA Dairy’s. Further

information on the design and operation of alkanolamine plants can be found in Gas
Purification, Kohl and Nielsen (1997).

2.5.3. Physical Solvents

When acid gases make up a large proportion of the total gas stream, the cost of

removing them with heat-regenerable processes, such as amines, may be out of line

with the value of the treated gas. Physical solvents, where the acid gases are simply
39

dissolved in a liquid and flashed off elsewhere by reducing the pressure, have been

employed with limited success. Since these processes depend on partial-pressure

driving forces, some product will invariably be lost, especially at higher pressures.

2.5.3.1. Water Washing

Liquids with increased solubilities for CO2 and H2S are typically chosen over
water, but the principal advantages of water as an absorbent are its availability and low

cost. Absorption of acid gas produces mildly corrosive solutions that can be damaging

to equipment if not controlled. Table 2.16 indicates Henry’s law constants for biogas

components in water.

Table 2.16: Henry’s Law Constants at 25° C and 1-Atmosphere


CH4 1.5 x 10-4 M/atm
CO2 3.6 x 10-2 M/atm
H2S 8.7 x 10-2 M/atm

As seen, H2S has a slightly higher solubility than CO2, but costs associated

with selective removal of H2S using water scrubbing have not yet shown competitive

with other methods. Therefore, water scrubbing will probably only be considered for

the simultaneous removal of both H2S and CO2. Experimentally derived equilibrium
constants for mixtures of CH4, CO2, and H2S have been determined and can be used to

calculate water and gas flow rates, as well as vessel dimensions (Froning, et al. 1964).

2.5.3.2. Other Physical Solvents

Solvents such as methanol, propylene carbonate, and ethers of polyethylene

glycol, among others, are offered as improved physical solvents. Criteria for solvent

selection include high absorption capacity, low reactivity with equipment and gas
40

constituents, and low viscosity. Thermal regeneration techniques are still needed in

most cases to achieve pipeline-quality gas. Additionally, loss of product can be higher

with these solvents, as levels as high as 10% have been reported (Kohl and Neilsen

1997).

The Selexol process (Union Carbide) utilizes dimethylether of polyethylene

glycol (DMPEG) as a purely physical solvent. In 1992, Union Carbide reported 53

Selexol plants operating, of which 15 were designed for selective removal of H2S and
8 were in service for landfill-gas purification. Like water scrubbing, the cost for

selective H2S removal has not yet shown to be competitive and this process will most

likely only be considered for applications in which upgrading to relatively pure

methane is desired (Wellinger and Linberg 2000).

The Sulfinol Process (Shell Oil Company) is unique because it couples

improved physical solvents with chemical amine agents to boost removal efficiencies.

This method can easily produce pipeline-quality gas, but has yet to be demonstrated as

economical for small-scale biogas H2S removal.

2.5.4. Membrane Processes

Membranes operate based on differing rates of permeation through a thin

membrane, as dictated by partial pressure. Because of this, 100% removal efficiency

is not possible in one stage, and some product will inevitably be lost. Two types of

membrane systems exist: high pressure with gas phase on both sides, and low pressure

with a liquid adsorbent on one side. Membranes are generally not used for selective

removal of H2S from biogas but are becoming more attractive for upgrading of biogas
to natural-gas standards because of attributes such as reduced capital investment, ease

of operation, low environmental impact, gas dehydration capability, and high

reliability.
41

Kayhanian and Hills (1987) studied high-pressure membrane purification

specifically for the purification of anaerobic-digester gas. Cellulose acetate

membranes operating at 25°C, 550 kPa, and a stage cut (ratio of permeate flow rate to

non-permeate flow rate) of 0.45 performed the best for removal of CO2 and H2S, and
reduced 1000 ppm H2S to 430 ppm (Kayhanian and Hills 1988). Three-stage units

treating landfill gas have achieved product gases with over 96% CH4 but utilize

separate H2S removal systems to extend the membrane life, which is typically in the

range of three to five years (Wellinger and Linberg 2000).

Low-pressure gas-liquid membrane processes have recently been developed

specifically for upgrading of biogas and operate at around atmospheric pressure and

25°C –35°C. Initial trials indicate that 2% H2S concentrations can be reduced to less

than 250 ppm using NaOH or coral solutions for the liquid. Amine solutions can be

employed for preferential CO2 removal and traditional liquid regeneration techniques

employed for the solvent. This process is still in a developmental stage but may prove

to be desirable in the future (Eriksen, et al. 1999).

2.6. ALTERNATIVE H2S CONTROL METHODS

2.6.1. In-Situ (Digester) Sulfide Abatement

Iron chlorides, phosphates, and oxides can be added directly to the digester to

bind with H2S and form insoluble iron sulfides. McFarland and Jewell (1989) studied
the effects of digester pH and addition of insoluble iron phosphate directly to

digesters, pointing out that addition of FeCl3, although regularly practiced, is often

inconsistent and inconclusive for reducing H2S. Lab studies showed that increasing

pH from 6.7 to 8.2 through the addition of phosphate buffers reduced gaseous sulfide

emissions from 2900 to 100 ppm, while increasing soluble sulfide concentrations from
42

18 to 61 mg/l. Soluble sulfide levels around 120 mg/l begin to inhibit CH4 production.
Addition of insoluble iron (3+) phosphate up to FePO4-Fe:SO42--S ratios of 3.5,

reduced gaseous sulfide levels from 2400 to 100 ppm (McFarland and Jewell 1989).

Ferric phosphate (FePO4) and ferric oxide (Fe2O3) are able to lower HS-

concentrations in the digester via Equations 2.24 and 2.25 (Jewell, et al. 1993).

2 FePO4 H2O + 3 H2S Æ Fe2S3 + 2 H3PO4 + 2 H2O (2.24)

Fe2O3 H2O + 3 H2S Æ Fe2S3 + 4 H2O (2.25)

This method may be effective as a partial removal process for reducing high

H2S levels, but usually must be used in conjunction with another technology for
removal down to about 10 ppm H2S. Concern also exists that accumulation of

insoluble iron sulfides might cause premature buildup in a digester (Jewell, et al.

1993).

Richards, et al. (1994), studied a unique, in-situ, method for methane

enrichment whereby the leachate from a semi-continuously fed and mixed (SCFM)

reactor was purged of CO2 in an external, air-purged, stripper. This process took

advantage of differing solubilities for CO2 and methane, and it produced gas with over

98% CH4. No monitoring of H2S was conducted. This process has limited application
to SCFM or CSTR reactors, and further testing is needed to determine practical design

and operating requirements for larger-scale operation (Richards, et al. 1994).

2.6.2. Dietary Adjustment

Diet composition influences sulfur content in animal wastes, which directly

impact sulfides emitted from anaerobically digested manure. Sulfur is a required

nutrient for animal health and cannot be completely eliminated from a diet. Shurson,

et al. (1998), have reduced H2S levels from anaerobic swine-manure lagoons by 30%
43

through careful manipulation of a nutritional swine diet. Animal performance and

ammonia emissions were not studied in this experiment. Dietary adjustment is

generally not used for sulfide reduction because diets are typically optimized for

product yields and animal health, rather than sulfur levels in the excrement.

Furthermore, a complete reduction in H2S can never be effected, so additional H2S


abatement processes are needed (Shurson, et al. 1998). However, limiting sulfur

containing chemicals or high sulfate content waters from inadvertently entering the

digester could be a simple way to reduce H2S emissions somewhat.

2.6.3. Aeration

A simple technique for H2S reduction, now practiced in Europe, includes

air/oxygen dosing into the biogas. Air is carefully admitted to the digester or biogas

storage tank at levels corresponding to 2-6% air in biogas. It is believed effectiveness

is based on biological aerobic oxidation of H2S to elemental sulfur and sulfates.

Inoculation is not required, as Thiobacillus species are naturally occurring at aerobic

liquid-manure-wetted surfaces. Full scale digesters have claimed 80-99% H2S

reduction, down to 20-100 ppm, by adding <5% air with a simple air pump. Yellow

clusters of sulfur are deposited on surfaces, increasing chances of corrosion. Care

must also be taken to avoid explosive gas mixtures (Nijssen, et al. 1997).

2.7. BIOLOGICAL H2S REMOVAL METHODS

2.7.1. History and Development

To biologically address the problem of malodorous air, open-bed soil filters

began to be used in the 1920’s and industrial soil biofilters first appeared in the United
44

States during the 1950’s, but operation was not well understood (Carlson and Leiser

1966). Sulfur compounds are a major component of malodor in gases and are

produced during biochemical reduction of inorganic or organic sulfur compounds.

Many soils do exhibit a small chemical adsorption capacity for H2S that is heavily
dependent on the iron content of the soil (Bohn and Fu-Yong 1989). It has since been

determined that sustained effectiveness of soil or other biofiltration beds arises

primarily from microbial oxidation of organic compounds, leading to biomass

formation and nontoxic odorless products, or oxidation of inorganic compounds (such

as sulfides), which supply energy to cells and produce odorless compounds like

elemental sulfur and sulfate in the process (Ottengraf 1986).

Biologically active agents have since been used in a variety of process

arrangements, such as biofilters, fixed-film bioscrubbers, and suspended-growth

bioscrubbers (Dawson 1993). These processes may also be effective at removing

multiple contaminants from a gas stream, increasing their functionality. Fluidized-bed

bioreactors have recently been tested for simultaneous removal of H2S and NH3 with

promising results (Chung, et al. 2001). It is also possible to achieve co-treatment of

volatile organic compounds and H2S in the same biofilter (Devinny, et al. 1999).

Reviews on exhaust gas purification for odor control, microbiological

treatment of H2S containing gases, and biofiltration for air pollution control have been

published that summarize the current state of the art (Ottengraf 1986, Jensen and

Webb 1995, Devinny, et al. 1999). A general biofiltration process may include the

elements illustrated in Figure 2.9.


45

Figure 2.9: Biofiltration System Schematic


Source: Swanson and Loehr (1997)

There are many companies specializing in the design and operation of

biofilters for pollution control, including TRG biofilters, Bohn Biofilter Corporation,

Biorem Technologies, Biocube, Inc., and Envirogen, among others.

2.7.2. Biological Sulfur Cycles

Sulfur exists in many oxidation states from +6 (SO4) to –2 (H2S).

Transformations take place at significant rates both chemically and biologically. The

global balance of sulfur is depicted in the Figure 2.10.


46

Figure 2.10: The Global Sulfur Cycle.


(Artificial emissions are derived from human activities. An asterisk indicates a
process that is partially or solely due to microbial action.)
Source: Madigan, et al.(2000), pg 688.

Most sulfur in the Earth’s crust is contained in sulfate minerals, such as

gypsum, and sulfide minerals, such as iron pyrite. The oceans constitute the most

significant reservoir for sulfur, mostly in the form of inorganic sulfate. The biological

sulfur-cycle due to microbial involvement is depicted in Figure 2.11.


47

Figure 2.11: Biological Redox Cycle for Sulfur


Source: Madigan, et al. (2000), pg 687.

H2S is produced from bacterial sulfate-reduction in Desulfovibrio,

Desulfobacter, Desulfuromonas, and many other hyperthermophilic Archaea, via the

overall pathway shown in Equation 2.26.

SO42- + 8 H Æ H2S + 2 H2O + 2 OH- (2.26)

Once present, other microorganisms can use H2S oxidation to gain energy.
Various groups of organisms can oxidize reduced sulfur compounds under aerobic or

anaerobic conditions, including:

• Colorless sulfur bacteria (aerobic, i.e.: Thiobacillus, Beggiatoa, Thiothrix, etc.)

• Green sulfur bacteria (anaerobic, phototrophic, i.e.: Chlorobium, etc.)

• Purple sulfur bacteria (anaerobic, phototrophic, i.e.: Chromatium, Thiocapsa, etc.)


48

Colorless sulfur-oxidizing bacteria are most widely used to oxidize H2S using
oxygen as an electron acceptor. This process is preferred because growth rates are

significantly higher and there are no light intensity requirements. Thiobacillus species

are thought to account for a majority of sulfide oxidation, via the sulfite-oxidase

pathway as described in Figure 2.12 below.

Figure 2.12: Steps in the Oxidation of Sulfur Compounds by Thiobacillus Species.


Source: Yang (1992), pg 24.

The most energy is released when sulfide is oxidized completely to sulfate as

seen in Equation 2.27. Sulfide oxidation often occurs in steps with elemental sulfur as

an intermediate product, as seen with Equations 2.28-2.29, and in oxygen-limited

environments, oxidation may proceed only to elemental sulfur, producing less energy.

Cells can either deposit sulfur inside or outside of their cell membranes. Other

reduced sulfur compounds, such as thiosulfate, can also be oxidized for energy as seen

in Equation 2.30.
49

H2S + 2 O2 Æ SO42- + 2 H+ (∆G0’ = -798.2 kJ/rxn) (2.27)

HS- + ½ O2 + H+ Æ S0 + H2O (∆G0’ = -209.4 kJ/rxn) (2.28)

S0 + H2O + 1½ O2 Æ SO42- +2 H+ (∆G0’ = -587.1 kJ/rxn) (2.29)

½ S2O32- + ½ H2O + O2 Æ SO42- + H+ (∆G0’ = -409.1 kJ/rxn) (2.30)

Table 2.17 references several studies on biofiltration of H2S with specific


bacterial populations.

Table 2.17: Specific Microorganisms Studied for Biofiltration of H2S


MICROORGANISM REFERENCE
Degorce-Dumas, et al. (1997), Nishimura and Yoda
Thiobacillus species
(1997), Koe and Yang (2000), Oh, et al. (1998)
Thiobacillus thioxidans Cho, et al. (2000), Cadenhead and Sublette (1990)
Sublette, et al. (1994), Sublette and Sylvester
Thiobacillus denitrificans
(1987a, 1987b)
Thiobacillus thioparus Cho, et al. (1992), Cadenhead and Sublette (1990)
Thiobacillus ferrooxidans Jensen and Webb (1995)
Thiobacillus novellas Chung, et al. (1998)
Thiobacillus versutus Cadenhead and Sublette (1990)
Thiobacillus neopolitanus Cadenhead and Sublette (1990)
Pseudomonas putida Chung, et al. (1996, 2001)
Hyphomicrobium Zhang, et al. (1991)
Xanthamonas species DY44 Cho, et al. (1992)

Reaction products from many of these microorganisms include sulfates and H+,

which form sulfuric acid in the leachate and reduce the pH. Some Thiobacillus

species are acidophilic and therefore function adequately at low pH, but organic media

tends to degrade under these conditions causing plugging and increased pressure drop.

Investigations of different media have been done with respect to H2S removal in

biofilters, as summarized in Table 2.18.


50

Table 2.18: Media Tested for Biofiltration of Hydrogen Sulfide


ORGANIC MEDIA REFERENCE
Soil Carlson and Leiser (1966)
Furusawa, et al. (1984), Hartikainen, et al. (2001),
Peat Kim, et al. (1998), Zhang, et al. (1991), Degorce-
Dumas, et al. (1997), Cho, et al. (1992)
Rands, et al. (1981), Yang (1992), Wani, et al.
Compost
(1999), Sun, et al. (2000)
Sludge Yang (1992), Degorce-Dumas, et al. (1997)
Pig Manure/Sawdust Elias, et al. (2002)
Wood Bark and Waste Langenhove, et al. (1986), (Wani, et al. (1999)
Activated Carbon Oh, et al. (1998)
Rock Wool Kim, et al. (1998)
INORGANIC MEDIA REFERENCE
Lava Rock Cho, et al. (2000)
Poly-propylene Rings Koe and Yang (2000)
Calcium-alginate Beads Chung, et al. (1996)
Fuyolite and Ceramics Kim, et al. (1998)

Desirable attributes for biofilter support media include high surface area, low

pressure-drop characteristics, good moisture retention properties, durable in their

active environment, can provide a source of nutrients for an active biolayer, and

support a diverse community of microbes. The trade-off between organic and

inorganic media is traditionally that organic composts have vibrant microbial

populations and form extremely active biolayers, but degrade quickly at low pH and

have higher pressure-drops than some inorganic carriers.

2.7.3. Example Applications

A comprehensive study of operational parameters, design, and basic kinetic

modeling for removing H2S from air with composted sewage sludge and yard waste

was conducted by Yang and Allen (1994). Variables studied include temperature,

residence time, concentration, loading rate, compost sulfate level, acidity, and water
51

content. H2S removal efficiencies greater than 99.9% were noted using yard waste
composts and inlet concentrations ranging from 5 to 2650 ppm. Maximum

elimination capacities for the composts ranged from 11.5 to 130 g S/m3-solids/hr

(Yang and Allen 1994a, 1994b).

Elias, et al. (2002), operated an H2S biofilter using compressed composted pig

manure and sawdust for 2500 hours with over 90% removal efficiency. H2S loading

in air was in the range of 10–45 g H2S/m3-solids/hr with empty-bed residence times of

13-27 seconds. No chemical additions were needed for buffering or nutritive reasons

during operation. Elemental sulfur was the main sulfur product accumulated (87.5%

of sulfur) in the bed. Only a small pH drop was noticed, so leaching of heavy metals

was not significant (Elias, et al. 2002).

Manure composts have been used for biofiltration of other compounds as well.

Chou and Cheng (1997) tested composted pig- and cow-manure media, mixed with

woodchips and activated sludge, for removing methyl ethyl ketone (MEK), and

achieved removal rates of around 50 g/m3-solids/hr (Chou and Cheng 1997).

Cardenas-Gonzalez, et al. (1995), compared properties of immature and mature yard-

waste and horse-manure composts for biofiltration of VOC’s and found that the horse-

manure compost had higher microbial activity and shorter acclimation time, but was

not as stable for long term operation (Cardenas-Gonzalez, et al. 1999).

Degorce-Dumas, et al. (1997), tested biofilter columns with peat and dry

wastewater sludge on actual biogas (characterized as 50-60% CH4, 40-50% CO2, and
0.5-2.0% H2S), mixed 2:1 with air. The H2S concentration in the gas stream was

measured at 3260 mg/m3 (2375 ppm), and the column maintained 100% removal

efficiency for 10 days at a loading rate of 129 g H2S/m3-solids/hr. Autoclaved

compost, used as a control, showed only 60% H2S removal under similar conditions.

A Henry’s law calculation indicated that the abiotic removal efficiency cannot come
52

only from H2S absorption into water, but must also be from chemisorption (Degorce-
Dumas, et al. 1997).

Gadre (1989) also passed actual biogas from a lab scale anaerobic digester

(~55% CH4, ~42.5% CO2, and 2.04% H2S) through a 50-mL glass-bead-packed

biotrickling filter washed with innoculum isolated from distillery wastewater. The

collection vessel for the wastewater was open to the atmosphere and assumed to

contain aerobic Thiobacillus species due to a pronounced drop in pH to 3.0. 69.5%

H2S removal was achieved at a loading rate of 187 mg H2S/day (Gadre 1989).

Nishimura and Yoda (1997) performed a similar experiment with a more methane rich

biogas (~80% CH4, ~20% CO2, and 2000 ppm H2S), and were able to achieve 99.5%

reduction in H2S with a gas flow of 40 m3/hr in a 3 m3 bubble column reactor

(Nishimura and Yoda 1997).

A German patent issued to Neumann, et al. (1990), describes a method for

removal of H2S from biogas utilizing Thiobacillus ferooxidans in a packed bed of peat

or refuse-compost. Although flow rates and oxidation rates are not mentioned,

experimental results indicate a product gas with 59.8% CH4, 30.8% CO2, undetectable

H2S, 9.1% N2, and 0.5% O2, was produced from an inlet gas with 65% CH4, 34.0%

CO2, 1.0% H2S, 0.0% N2, and 0.0% O2 (Jensen and Webb 1995).

To investigate inorganic media supports for durability during low-pH H2S

biofiltration, Cho, et al. (2000), specifically immobilized Thiobacillus thioxidans on

porous lava rock. The rock showed favorable moisture retention and resisted

excessive pressure drops. Increased removal capacities up to 428 g S/m3-solids/hr


were reported with space velocities of 300-hr-1 (Cho, et al. 2000). In a previous study,

Cho, et. al. (1992), reported 89%+ removal of dimethyl-sulfide, methanethiol,

dimethyl-disulfide, and H2S with a Thiobacillus thioparus biofilter treating exhaust

gas from a night-soil (septic sludge) treatment plant (Cho, et al. 1992).
53

Koe and Yang (2000) also tested Thiobacillus thioxidans with plastic packing

and found that for gas-retention times greater than five seconds and a loading rate

below 90 g S/m3-solids/hr, 99% H2S removal was obtained (Koe and Yang 2000).
H2S levels up to 10,000 ppm were oxidized by Sublette, et. al. (1994), with

pure cultures of Thiobacillus denitrificans in less than 2 seconds under anoxic

conditions. Here, added nitrate, rather than oxygen, served as the terminal electron

acceptor. These reaction times indicated that limitations in H2S removal were due to

mass transfer rather than biological limitations. Up to 97% reduction in inlet H2S

levels were achieved (Sublette, et al. 1994).

Anaerobic bacteria have also been used for H2S oxidation from gas streams.

Cork, et al. (1983), provided the first demonstration of photoautotrophic growth of a

Chlorobium species with continuous inorganic gas feed (3.9% H2S, 9.2% CO2, 86.4%

N2, and 0.5% H2). The reaction was completed in a 1-L clear vessel with an external

light source. With a removal efficiency of 99.6% H2S, elemental sulfur and biomass

were the main reaction products (Cork, et al. 1983). Subsequent work with

Chlorobium species has been done (Kim, et al. 1997; Henshaw and Zhu 2001;

Kobayashi, et al. 1983) and an economic estimate of $2.82–$4.24 per thousand

standard cubic meters for gas desulfurization has been made (Basu, et al. 1996).

A few commercial biological processes exist specifically for energy gas

desulfurization. The Biopuric process (Biothane Corporation) has designed and tested

systems for H2S removal from gas streams similar to those found at agricultural
anaerobic-digestion facilities (1500–7000 m3 gas/day and 1000–27000 ppm H2S) with

consistent removal efficiencies over 97%. These systems generally cost $75,000–

$100,000 for capital investments alone (Lanting and Shah 1992).

Another biological process targeted at sour gas desulfurization is the Thiopaq

Process (UOP). SO2 and H2S are absorbed in a traditional chemical scrubber with
54

sodium bicarbonate solution (Ruitenberg, et al. 1999). Pressures in the scrubber are

often elevated to 6000 kPa to enhance absorption (Janssen, et al. 2000). The spent

liquid is then regenerated in a separate bioreactor, producing elemental sulfur. 99%

H2S removal is reported and 90-95% of the sulfur is recoverable. A $1.7 million
Thiopaq installation removed 2.76% H2S from 2000 m3 gas/hr to less than 10 ppm,

while recovering 96% of the sulfur. Operating costs were estimated at $65/day for

nutrients, 75 kWh/hr of electricity, and 180 kg/day of NaOH (UOP 2000).

2.8. RESEARCH STATEMENT

With integrated farm energy systems, the opportunity exists for improvements

in gas processing by utilizing on-farm compost and biological processes for H2S

removal. While there is a wealth of operational and research experience on

biofiltration for odor control, there is a relatively limited amount of information on

biofiltration for gas processing to purify biogas. No studies, to this author’s

knowledge, exist where anaerobically digested and composted dairy-manure has been

tested for its capability to selectively remove H2S, often at elevated levels, from

biogas. The following research directly addresses this need.


CHAPTER

3. MATERIALS AND METHODS

An experimental approach was used to investigate the suitability of digested

cow-manure compost for removal of H2S from biogas. Test reactors were constructed,

packed with compost, and exposed to actual biogas on-site at AA Dairy. The

experimental setup was located in the enclosed, but non-environmentally controlled,

engine room.

3.1. REACTORS

3.1.1. Small Reactors

Four reactor columns were built using 0.10 m (4 inch) ID, schedule-40 white

polyvinyl chloride (PVC) pipe. Each reactor is 0.5 m in length with female adapters

and male cleanout plugs on each end. Although clear pipe would have been desirable

for observing the compost, white pipe was used due to budget limitations. Type 316

stainless steel woven wire discs (0.032 inch wire with 8 x 8 wires per inch) were glued

into the columns 0.1 m from the end for packing support. Plastic 6.35 mm (¼ inch)

barbed fittings were placed in the center of each cleanout plug and 0.05 m from the

column ends for gas delivery and sampling with 6.35 mm flexible PVC tubing. The

small reactors are depicted in Figure 3.1. Two of the small reactors were equipped

with liquid leachate recycle capability using a Cole-Parmer peristaltic pump with dual

heads, as shown in Figure 3.2. Three-millimeter holes were drilled in all small

columns at 0.05, 0.25, and 0.45 m lengths for thermocouple insertion.

55
56

0.05 m
OUT T
1
0.5 m
G

3 T 0.3 m

T
IN 1 0.05 m

0.1 m

LEGEND:
T = Thermocouple 1 - PVC End Caps
G = Gas Sampling Port 2 - Top of Bed
IN = Gas Sample Inlet 3 – PVC Pipe
OUT = Gas Sample Outlet 4 – Wire Screen

Figure 3.1: Schematic of Small Columns


57

0.05 m
OUT T
1
4
0.5 m
G

2
0.3 m

3 T

T
IN 1
0.05 m

5
0.1 m

LEGEND:
T = Thermocouple 1 - PVC End Caps
G = Gas Sampling Port 2 - Top of Bed
IN = Gas Sample Inlet 3 - PVC Pipe
OUT = Gas Sample Outlet 4 - Wire Screen
5 – Peristaltic Pump

Figure 3.2: Schematic of Small Columns with Leachate Recycle


58

3.1.2. Large Reactors

Two identical 0.16 m (6.355 inch) ID by 1.5 m length clear PVC (ALSCO

Industrial Products, Inc.) columns were constructed with 3.175 mm wall thickness.

Each column was divided into three sections for easy inter-column gas sampling,

accessible compost loading, and to prevent compaction, as seen in Figure 3.3.

Specially constructed couplers lathed from 0.15 m (6 inch) ID schedule-40 white-PVC

pipe were used in conjunction with plastic draw latches and neoprene O-rings (size

362) to seal the sections. The same stainless steel screen used in the small reactors

were reinforced with 6.35 mm aluminum bars and glued into place as a bed support.

Silicone sealant was used to seal any leaks because of insufficient o-ring seating.

Barbed plastic fittings (6.35 mm) were installed before and after each bed

section for gas sampling with 6.35 mm clear-PVC flexible tubing. Inlet and outlet gas

ports were 6.35 mm plastic ball valve fittings. Thermocouples were inserted into the

center of each bed section and near the gas inlet and outlet ports for temperature

measurement.

PVC and 316 stainless steel materials were chosen because they are not

affected by exposure to methane, carbon dioxide, hydrogen sulfide, or dilute

concentrations (<75%) of sulfuric acid (Cole Parmer 2002). All reactors were

pressure tested in the laboratory to 34.5 kPa using compressed air, and no leaks were

detected.
59

T 0.05 m
OUT 1
G
1.5 m
3 T

0.3 m
T

0.4 m
T
2

G
T
IN
1 0.05 m

LEGEND:
T = Thermocouple 1 - PVC End Caps
G = Gas Sampling Port 2 - Wire Screen
IN = Gas Sample Inlet 3 - PVC Pipe
OUT = Gas Sample Outlet 4 – Plastic Latches
5 – PVC Couplers

Figure 3.3: Schematic of Large Columns


60

3.2. EXPERIMENTAL SETUP ON SITE

Piping was established so that a small portion of biogas could be diverted to

the test columns and returned to the original pipeline upstream of the engine, as shown

in Figure 3.4. Existing piping enabled digester gas at 0.75–1 .0 kPa to enter three 250-

W blowers, boosting the pressure to around 3 kPa. From here gas passed through a

solenoid valve that regulated the positive pressure of the engine intake to about 1.5

kPa. As shown in Figure 3.4, a 5 cm (2 inch) PVC tee was installed with a ball valve

(Banjo Corporation) between the blowers and the solenoid valve. 5 cm ID PVC

piping was run approximately 5 m to the column test area and terminated with an end-

cap with 6.35 mm barbed fitting installed. From here, pumps were used to boost gas

pressure and the biogas stream could be split, as needed, depending on the

experimental configuration.

An exhaust manifold consisting of 3 m of 3.173 mm (1.25 inch) ID, schedule-

40, white PVC pipe was installed. A specially constructed steel box, containing a

rectangular automobile air filter element, was placed downstream of the manifold to

protect the engine from any particulate blow-over from the experimental columns.

The tested biogas was then returned to the main biogas pipeline downstream of the

solenoid valve. Another Banjo™ ball valve was placed at the return intersection to

enable complete shut-off of the experiment stream when not in use.


Figure 3.4: Experimental Setup at AA Dairy
62

Two vacuum pumps were used to provide the additional head needed for the

experimental columns; a larger pump and a smaller pump for use with respective

columns. The larger pump was a 370-W Sargent-Welch 1402 Duoseal single-phase

pump. Biogas was fed to the vacuum inlet of the pump through a 3 m length of 6.35

mm ID brass tubing and exhausted to a multi-ported manifold on the positive pressure

side of the pump. A Speedaire airline oil-removal filter and pre-filter were installed

on the inlet side of the pump to remove oil, particulates and some moisture. The pump

is specified to deliver 9.5 m3/hr at standard temperature and pressure, but only
delivered 3.3 m3/hr due to resistance on the inlet side. The original pump configuration

produced a significant amount of oil mist in the outgoing gas stream. A mist

eliminator was designed, constructed, and installed on the outlet side of the pump by

John Poe Tyler (Tyler 2001). The smaller pump was a “Neptune Dyna-Pump” vacuum

diaphragm pump and was operated without inlet or outlet filters.

Compressed air was available at the test site and regulated by a single-stage

Speedaire regulator before being delivered to a block of multiple,variable-area flow-

meters, or “rotameters” (Dwyer Instruments, Inc.). Biogas and air were mixed

together at a tee connector on the inlet side of the humidification vessels.

To humidify the incoming gas, all inlet gas streams were bubbled through 1-L

plastic Nalgene bottles initially filled with 750 ml of distilled and deionized water. A

schematic of the simple humidification and mixing vessel is provided in Figure 3.5.
63

Air
Biogas Inlet
Inlet Mixed
Sample
Gas
Outlet

Figure 3.5: Humidifier and Air/Biogas Mixing Vessel

Two of the smaller columns were outfitted with a leachate recycle loop

consisting of a Cole-Parmer peristaltic pump with dual heads, as shown in Figure 3.2.

With a tubing size of 1/16”, flow rates up to 7 ml/min were achievable. The pump

forced liquid to the top of the column where a drop would form. Another stainless

steel screen was placed 0.05 m below the recycle inlet port to disperse the droplet as it

fell across the top of the media.

Gas flow rates were controlled with Gilmont Accucal™ variable-area flow-

meters (Barnant Company). Rates are measured by visually correlating the center of

the float with a graduated scale, calibrated for liters per minute of air at standard

temperature and pressure. Correction Equations 3.1-3.3 can be applied when

measuring gases other than air under non-standard conditions: (Gilmont 1993)

0 . 00120 (3.1)
q G0 = q 0A
ρ G0

ρ G0 = ∑1 xi ρ i0
i
(3.2)

P 530
qG' = qG0 ⋅ (3.3)
760 T
64

Where qG0 = Standard gas flow


q A0 = Standard air flow reading from meter
qG' = Gas flow at P (operating pressure) and T (operating temperature)
with volume corrected to measurement at standard conditions

ρ G0 = Density of gas in g/ml at standard conditions


xi = Mole fraction of ith gas component
ρ i0 = Density of ith gas component in g/ml at standard conditions

Assuming measurement at standard temperature and pressure, and

approximating biogas as 60% methane ( ρ CH


0
4
= 0.00072 gm/ml), and 40% carbon

dioxide ( ρ CO
0
2
= 0.00198 gm/ml), the corrected rotameter value calculated with

Equation 3.1, for biogas, becomes qG0 = 0.992 q A0 . Therefore, readings for air and

biogas are nearly the same.

3.3. GAS SAMPLING AND MEASUREMENT

Flexible PVC sample lines (6.35 mm ID) were run from the column gas

sampling ports to the inlet ports of a 16-channel multi-position valve (Valco

Instruments Co., Inc.). The desired sample line was selected with a digital controller,

and the outlet line from the switching valve was connected through a rotameter to the

H2S detector. Since the columns were operated under slight positive pressure, opening
the appropriate sample path allowed gas flow to the detector.

3.3.1. Electrochemical Sensor

An electrochemical, Toxi-Plus single gas detector (Biosystems, Inc.) was used

for H2S monitoring. The detector was equipped with a 0-100 ppm 4HS CiTiceL® H2S
65

sensor (City Instruments, Inc.). A plastic calibration adapter provided by Biosystems,

Inc., enabled direct sampling of a gas stream, provided a flow rate near 1 liter per

minute.

The electrochemical sensor has been carefully designed to minimize the effects

of common interfering gases, but some interfering gases may still have either a

positive or negative effect on the sensor readings. Table 3.1 indicates deviation of

measured H2S values with respect to a number of substances. The table is not meant
to be complete, as there may be other gases to which the sensor responds.

Table 3.1: Cross Sensitivity Data for Electrochemical H2S Sensor


(Number reported is sensor response to 100 ppm of selected test gas)
Compound Response
CH4 0
CO2 0
CO <2
Cl2 -20
(CH3)S 10
CH3SH 45
H2 < 0.5
NO2 -20
O3 -30
SO2 < 20
Source:City Technolgy Limited (2002)

Calibration of the sensor is recommended on a daily basis and was done with
certified-standard 24.7 ppm H2S in nitrogen (Empire AirGas, Elmira, NY) supplied in

a size-33A cylinder containing 850 liters of product. A dual-stage, stainless-steel

regulator (CGA #330: Matheson Tri-Gas) delivered calibration gas to the sensor.

Since H2S concentrations in the biogas were outside the range of the sensor, an

air dilution method was developed and utilized. Compressed air delivered through a

rotameter at 3.9 liters per minute was continuously combined with 0.1 liters per

minute of sample gas at a plastic tee, creating a 40-fold dilution of the sample stream.
66

The gas mixture then entered a 1-liter sealed plastic mixing vessel similar to the

humidification vessel depicted in Figure 3.5, but without water. The effluent from the

mixing vessel was sent directly to the electrochemical sensor for measurement.

The H2S measurement protocol with the electrochemical sensor was as


follows:

1) Select desired sample channel from multi-position valve

2) Adjust gas sample flow rate from outlet of multi-position valve to 0.1 lpm

3) Adjust air flow rate to 3.9 lpm

4) Let gas mix and allow meter reading to stabilize (~4 minutes)

5) Repeat steps 1-4 for additional sample streams

3.3.2. Gas Sampling Tubes

Gas detection tubes employing chemical reaction with lead acetate (Sensidyne)

were used for additional measurement of H2S in the sample streams. The reaction in

Equation 3.4 occurs in the detector tube, causing a brown stain to form that can be

directly read for H2S concentration.

H2S + Pb(CH3COO)2 Æ PbS + 2CH3COOH (3.4)

A model 8014-400A (Matheson-Kitagawa) hand aspirated pump was used to

draw known volumes of sample through the detector tubes. Hydrogen-sulfide

detector-tubes (Kitagawa type 120SA) with a measuring range from 100-2000 ppm

were utilized. The scale on the detector tube is calibrated for 20°C and atmospheric

pressure. The manufacturer provides a temperature correction table and pressure

correction equation for measurements at non-standard conditions. H2S gas detector


tubes also exhibit cross-sensitivity interferences similar to those described for the

electrochemical sensor.
67

3.3.3. Gas Chromatography

The Mass Spectrometry Facility, Department of Chemistry and Chemical

Biology, at Cornell University performed gas chromatography/mass spectrometry

analysis on one sample of raw digester gas. The sample was delivered in a 0.3 liter

Tedlar® gas sampling bag with a rubber septum valve (Cole-Parmer).

3.4. TEMPERATURE MEASUREMENT

Gas temperatures before and after each column, bed temperatures, and ambient

air temperatures were monitored with thermocouples and a computerized data

acquisition system. The hardware and software systems used are modified versions of

those described by Hall (1998).

Data acquisition hardware components from Computer Boards, Inc., were

utilized. Type-T (copper-constantan) thermocouples with welded and silicone coated

ends were inserted into the center of the reactor cylinders at locations previously

displayed Figures 3.1-3.3. Thermocouple wires were then connected to an EXP-32

thermocouple multiplexor board capable of handling 32 differential inputs and

equipped with onboard amplification and cold junction compensation. A CIO-DAS-

08, 12-bit analog-to-digital conversion board was installed in a Gateway PC with

Pentium I processor for multiplexor control and data acquisition.

A software program was written in Pascal 6.0, using the DOS 3.1 operating

system, to display and log desired temperatures. Temperatures are recorded into

temporary storage every 15 seconds for 15 minute periods. After each period, the

average, standard deviation, maximum, and minimum for each input channel are
stored into a computer data file and the cycle is repeated. Further documentation of the

software is provided in the program itself.


68

3.5. PRESSURE MEASUREMENT

Pressure measurements were made using Magnehelic® analog pressure sensors

(Dwyer Instruments, Inc.) with 0-1” H2O, 0-2” H20, and 0-5 psi ranges. Pressure drop

across a bed section was easily measured by disconnecting the appropriate sample

lines entering the multi-position valve and connecting them across the pressure sensor,

as shown in Figure 3.4.

3.6. COMPOST CHARACTERIZATION

The compost tested in this study was taken from the “finished compost” pile

on-site at AA Dairy. This medium consists only of anaerobically-digested separated-

solids that have been composted, using an outdoor windrow system, for at least 60

days. Samples from three spots around the pile were mixed together in a 15-liter

plastic pail to create representative samples. Tests performed on compost after

exposure to biogas may be from a specific part of the test column or from a well-

mixed sample of the entire contents, as stated in each method.

3.6.1. Moisture Content

Approximately 10 grams of sample were placed in aluminum weighing dishes,

and then into an 85°C oven for 24 hours. Weights of the samples before and after

were compared to determine percent moisture content. Samples from the tested

columns were taken from the external surface of the column cores at 0.05 m, 0.15 m,

and 0.25 m along its length.


69

3.6.2. Void Fraction:

A modified version of the “5-gallon pail” method, as described by Nicolai and

Janni (2001), was utilized to determine percent void fraction. The following

procedure was used to test the fresh compost:

1) A plastic pail was filled with 5 liters of water and a “full” line marked on

the inside of the pail. The water was then removed.

2) The pail was filled about 1/3 full with compost and dropped ten times from

a height of 15 cm onto the floor.

3) Compost was added to fill the pail 2/3 full. The pail was again dropped ten

times from 15 cm onto the floor.

4) Compost was added to the full line and dropped ten times again.

5) Compost was added to the full line once more.

6) Water was added to saturate the compost to the “full” line and the volume

of water was recorded.

The void fraction is calculated by dividing the volume of water added by the

original solids volume.

3.6.3. Bulk Density

Similarly, the density of 5-liters of fresh compost was determined by weighing

it in an unpacked form and packed form, as described above.

3.6.4. Particle Size Distribution

Distributions of particles were determined by passing them through a stack of


sieves arranged in decreasing mesh-size order and recording the weight retained on
70

each sieve. The mesh-sizes of each sieve were 9.52 mm, 2.36 mm, 1.168 mm, and 0.5

mm. The compost was placed in the largest sieve first and shaken for 2 minutes.

3.6.5. pH

The pH value for the compost was determined by mixing 20 ml of sample with

20 ml of distilled and deionized water in a 100 ml beaker and using a calibrated

Beckman Φ200 pH meter to determine the pH of the solution. As with moisture

analysis, samples from the biogas-exposed compost were taken from the external

surface of the column cores at 0.05 m, 0.15 m, and 0.25 m along the length.

3.6.6. Trace Element Analysis

The Cornell Nutrient Analysis Laboratory (CNAL) performed trace element

analyses by total microwave digestion on 1-liter media samples. For analysis of

biogas-exposed compost, representative mixtures of the entire column contents were

submitted.

3.6.7. Sulfate Content

Samples of fresh compost were tested for sulfate concentrations by ProDairy

(Syracuse, NY).

3.7. OPERATIONAL PROCEDURES

Columns were packed with compost according to steps 2-5 in the “void

fraction” measurement protocol. Each section of bed was filled to 30.5 cm (12
inches), creating bed volumes of 2.47 liters for the small columns and 6.24 liters for

each section of the larger columns, or 18.71 liters for the entire large columns.
71

In all of the trials, a 2:1 mixture of biogas-to-air was used as the test gas for the

columns. This formulation, similar to that tested in peat biofilters by Degorce-Dumas,

et al. (1997), was chosen for two major reasons: 1) To ensure an aerobic environment

to facilitate microbial oxidation with thiobacillus, pseudomonas, or other species that


may exist in composts, and 2) To maintain safe operation outside of the explosion

limits for methane (5-15% in air) and hydrogen sulfide (4-45% in air).

Six different trials were run in these experiments, as described in Table 3.2.

Table 3.2: Summary of Experimental Trial Conditions


Bed Total Biogas Air Empty Bed Linear
Leachate
Trial Volume Flow Flow Flow Residence Velocity
Recycle
(l) (lpm) (lpm) (lpm) Time (sec) (m/min)
1 18.7 11.2 7.5 3.7 98 0.55 No
2 18.7 11.2 7.5 3.7 98 0.55 No
3 2.45 1.5 1.0 0.5 98 0.19 No
4 2.45 1.5 1.0 0.5 98 0.19 No
5 2.45 1.5 1.0 0.5 98 0.19 Yes
6 2.45 1.5 1.0 0.5 98 0.19 Yes

These flow rates and empty-bed residence times were picked to maintain H2S

contaminant loading between 50 – 150 g/m3-solids/hr with expected H2S inlet

concentrations varying from 1000 – 3000 ppm.

In order to test the dilution method and column apparatus, a smaller column

without packing was run with 2:1 biogas-to-air mixture at 1.5 liters per minute for

three hours. Inlet and outlet H2S concentrations were compared using the
electrochemical sensor method previously described.
CHAPTER

4. RESULTS AND DISCUSSION

4.1. ORIGINAL COMPOST CHARACTERISTICS

The finished manure compost was dark brown in color and moist, but not

dripping, when squeezed by hand. The particles of fresh compost crumbled apart like

wet soil when touched. The odor characteristics changed from pungent and offensive

as fresh manure to earthy and soil-like as finished compost.

Figure 4.1: AA-Dairy “Field of Dreams” Cow-Manure Compost

Table 4.1 summarizes the results from characterization of the manure compost

as used in the columns.

72
73

Table 4.1: Cow-Manure Compost Characterization


Specification Value Std. Dev.
Water Content (Loss on drying, wt%) 72.9% 0.9%
Void Fraction 0.35 0.05
Density (g/L)
Unpacked 604 13
Packed 757 18
Particle Size Distribution (wt % retained on sieves)
Sieve Sieve Size (mm)
3/8 in. 9.52 16%
#7 2.8 62%
#8 2.36 9%
#14 1.168 10%
#35 0.5 3%
pH (in solution) 7.19 0.04
Trace Element Analysis (µg/g dry basis)
Sulfur 7,316
Aluminum 2,145
Arsenic < det.
Boron 67
Cadmium < det.
Calcium 49,910
Chromium 2
Cobalt 2
Copper 479
Iron 5,487
Lead < det.
Magnesium 13,615
Manganese 500
Molybdenum 11
Nickel 39
Phosphorous 11,869
Potassium 13,548
Sodium 3,277
Vanadium 7
Zinc 347
Note: <det = below analyte detection limit
Std. Deviations below 1% of reading for all values.
Sulfur Analysis (µg/g dry basis)
Sulfate 73
=
Sulfide as S2 579
74

4.2. OPERATIONAL SUMMARY

Trials 1 and 2, using the larger columns, were run for 16 hours during June 10-

12, 2002. Trials 3 and 4, utilizing the smaller columns with no leachate recycle, were

operated for 1,057 continuous hours from July 5 through August 19, 2002. Trials5

and 6, using the smaller columns with leachate recycle, were also started July 5, 2002,

and run for 50.5 and 216.5 hours, respectively.

Due to safety concerns during trials 1 and 2, the 370-W pump was not operated

while unattended. The high output of the pump could create an asphyxiating or

explosive air safety problem in the event of a leak. During operation, moisture would

accumulate in the pump’s inlet oil-filter bowl and have to be emptied every 4 hours.

Moisture also quickly appeared in all of the variable-area flow meters, causing

oscillation of flow. The flow meters were left at their original setting during this

oscillation. Despite attempts to seal large reactor joints with silicone sealant, gas and

liquid leaks were noticed at various times during operation. Because of this

observation, no data is reported for trials 1 and 2.

Trials 3-6 were started simultaneously using the 370-W vacuum pump, but, in

order to allow unattended operation, it was replaced with the smaller Neptune Dyna-

Pump after 10 hours of operation. Results from trials 3 and 4 are the major focus of

the following discussion because prolonged and consistent operation was logged.

Additionally, only compost from trials 3 and 4 were analyzed for chemical

composition after the trials. Trials 5 and 6, which included leachate recycle, were both

terminated prematurely due to increasing pressure drop in the bed and the inability to

provide sufficient flow to all columns with the smaller pump.

Moisture also condensed in the variable-area flow meters within the first hour
of trials 3-6. Flow rates were double-checked daily by temporarily inserting a clean,
75

dry, flow meter into the sample line. A small and unmeasured amount of liquid was

periodically lost during manual removal of sample lines to attach pressure sensors or

flow meters. The moisture was in the form of water vapor expelled due to the sudden

drop in column pressure to atmospheric, and estimated to be a less than a milliliter per

day per column.

The humidification flasks had to be replenished with 500 ml of water only

once after 700 hours of operation. Condensation was also evident in the clear PVC

lines to and from the columns.

4.3. PRESSURE MEASUREMENTS

For both columns 3 and 4, the pressure drop measured across the bed was

consistently 0.1 inches of H2O, the smallest graduation on the analog pressure scale.
Maximum pressure drops of 0.25 inches of H2O were recorded during the first few

hours only. Column 5, which included leachate recycle, displayed an increase in

pressure drop across the bed from 0.1 to 2.0 inches of H2O during operation.

Similarly, column 6 also displayed an increase in pressure loss from 0.1 to 3.0 inches

of H2O during the first 50 hours of operation. The pressure difference then receded to

around 1.0 inch of H2O for the rest of the trial. Pressure drops across beds are

depicted in Figure 4.2.


76

Pressure Drop Across Bed - Trials 3-6

3.5

3
Column 3 - No leachate Recycle
2.5 Column 4 - No leachate Recycle
Pressure (in. H2O)

2 Column 5 - Leachate Recycle


Column 6 - Leachate Recycle
1.5

0.5

0
0 200 400 600 800 1000
Run Time (total hrs)

Figure 4.2: Pressure Drop Across Bed for Trials 3-6

The decreases in observed pressure drops are most likely due to the

development of preferential flow patterns within the bed. The pressure range of 0.1-3

inches H20/ft for a 0.2 m/mim linear gas velocity corresponds with published compost
pressure drop data (Yang 1992); Nicolai and Janni 2001).

The following physical factors determine pressure drop across the filter bed:

1) Particle size distribution

2) Porosity

3) Particle shape and orientation

4) Gas viscosity and density

5) Superficial gas velocity

6) Height of the bed


77

Many empirical models have been developed to predict pressure drop through

a stationary bed of particles. The most well know models include the Ergun, Hukill,

and Shedd relations (McGuckin, et al. 1999). The Ergun equation, Equation 4.1

below, can be used to calculate expected pressure drop, including associated values.

∆P 150µ (1 − ε ) v0 1.75(1 − ε )ρv 02


2

= + (4.1)
L ε 3 (ΦD p )2 ε 3 (ΦD p )
∆P
Where = Pressure drop per unit length
L
µ = Gas viscosity ≈ 1240x10-7 Poise @ 20°C
ε = Porosity ≈ 0.35
v 0 = Superficial gas velocity ≈ 0.19 m/min
Dp = Characteristic dimension of particle ≈ 2.5 mm
Φ = Particle sphericity ≈ 0.7 for activated carbon
ρ = Gas density ≈ 1.16 kg/m3 @ 20°C

With these parameters, the calculated pressure drop for compost is 19.4 Pa/m,

or 0.078 inches of H20/m. This is lower than pressure drops measured experimentally.
Using a characteristic dimension (Dp) of 1.0 mm for particle size resulted in a

calculated pressure drop of 0.15 inches H20/ft, which is in closer agreement with the

measured data. The relatively low gas velocities used here resulted in minimal

pressure drops and did not require the use of bulking agents, as might be necessary in

larger reactors or those with decreased gas residence times.

4.4. TEMPERATURE MEASUREMENTS

The temperature measurement system logged data every 15 minutes when

operating properly. Unfortunately, interruptions to the power supply caused data loss

on multiple occasions and prevented any data collection after hour 431. The

temperature data collected for each column are in Figures 4.3-4.6.


78

Temperature - Column 3

45

40
Temp. (C) (15 minute average)

35

30

25

20
Inlet Gas Media
15
Outlet Gas Ambient
10
0 50 100 150 200 250 300 350 400 450
Run Time (hrs)

Figure 4.3: Temperatures (15-Minute Average) for Column 3

Temperature - Column 4

45

40
Temp (C) (15 minute average)

35

30

25

20
Inlet Gas Media

15
Outlet Gas Ambient

10
0 50 100 150 200 250 300 350 400 450
Run Time (hrs)

Figure 4.4: Temperatures (15-Minute Average) for Column 4


79

Temperature - Column 5

45

40
Temp (C) (15 minute average)

35

30

25

Inlet Media
20
Outlet Ambient
15
0 5 10 15 20 25 30 35 40 45 50
Run Time (hrs)

Figure 4.5: Temperatures (15-Minute Average) for Column 5

Temperature - Column 6

40

35
Temp (C) (15 minute average)

30

25

20
Inlet Media

Outlet Ambient
15
0 20 40 60 80 100 120 140 160 180 200
Run Time (hrs)

Figure 4.6: Temperatures (15-Minute Average) for Column 6


80

The cyclical variation is due to diurnal temperature fluctuation. The ambient

temperature was recorded inside of the engine room and is different than actual

outdoor temperatures. Outdoor temperature, as well as the digester temperature,

influences the inlet-gas temperature, which is why it is possible to observe an inlet-gas

temperature lower than ambient temperature. In all trials, it is observed that the bed

temperature is typically between 0-12°C higher than the inlet temperatures, as shown
in Figure 4.7. This is an indication of heat being generated in the bed, which may be

attributed to an exothermic biological, chemical, or physical adsorption reaction, and

could potentially be used to track bed activity or viability.

Difference Between Bed Temperature and Inlet-Gas Temperature For


Trials 3-6

14

12
Column 5
10 Column 6
Temp. Diff. (Bed-Inlet) [C] .

Column 3
8
Column 4
6

0
0 5000 10000 15000 20000 25000
-2

-4

Time (minutes)

Figure 4.7: Temperature Difference Between Bed and Inlet-gas for Columns 3-6
81

Table 4.2 below summarizes the maximum and minimum temperatures seen in
each trial.

Table 4.2: Summary of Temperature Extremes for Trials 3-6


Trial Ambient Temperature Inlet Temperature Bed Temperature
# Min(°C) Max(°C) Min(°C) Max(°C) Min(°C) Max(°C)
Column 3 16.7 36.5 13.9 37.8 20.6 42.0
Column 4 16.7 36.5 15.4 38.9 21.5 43.3
Column 5 22.5 33.6 21.4 35.0 24.7 41.9
Column 6 16.7 33.6 14.8 34.5 19.1 39.5

Literature indicates that the optimum range for biological H2S oxidation is in
the range of 30 – 40°C. In general, severe reduction in activity occurs below 10°C and
more moderate reduction of activity above 50°C (Yang 1992). Therefore, these
experiments were operated in the mid-to-high range of known optimum temperature
conditions.
A simplified heat balance can be written to estimate the heat generation in the
bed. Heat accumulation within a packed bed can be written as the sum of the
individual conductive, convective, evaporative and metabolic heat contributions.
Radiation losses are neglected and the media is assumed to maintain its uniform
structure. The overall relation is described by Equation 4.2.

Qaccumulation = Qconduction + Qconvection + Qevaporation - Qgeneration (4.2)

Considering only maximal gradients, unsteady-state heat accumulation can be


written as Equation 4.3 (Gutierrez-Rojas, et al. 1996).

 dTbed 
ρC p   = UA p (Tamb − Tbed ) + G ' ( ρ in hin − ρ out hout ) + (4.3)
 dt 
GAo Qw ( y in − y out ) − Q gen
82

Where: ρb = Bulk comp0st density, packed ≈ 750 g L-1


Cp = Medium heat capacity
U = Overall heat transfer coefficient
Ap = Specific heat transfer area parallel to gas flow
≈ 104 m2 m-3media
Tamb = Ambient Temperature of air ≈ 30°C
Tbed = Bed Temperature ≈ 33°C
G’ = Specific gas flow rate ≈ 0.01 m3gas m-3bed s-1
ρin = Inlet gas density ≈ 1.09 kg m-3 @ 30°C
ρout = Outlet gas density ≈ 1.08 kg m-3 @ 33°C
hin = Inlet gas enthalpy (@ Tin = 30°C) ≈ 744,080 J kg-1
hout = Outlet gas enthalpy (@ Tout = 33°C) ≈ 744,302 J kg-1
G = Gas mass velocity ≈ 0.0023 kg m-2 s-1
A0 = Specific heat transfer perpendicular to gas flow
≈ Ap ≈ 104 m2 m-3media
Qw = Latent heat of vaporization ≈ 2430 kJ kg-1 @ 30°C
yin = Water mass fraction of inlet gas ≈ 0.028 kgwater kg-1gas
yout = Water mass fraction of outlet gas ≈ 0.028 kgwater kg-1gas
Qgen = Heat generation from reaction in the bed
t = Time, seconds

Assuming heat accumulation is zero at steady state, and representative steady-


state temperatures (Tamb and Tbed) are grossly estimated from the temperature data for
column 6 (hour 50), many of the physical parameters can be determined. The overall
heat transfer coefficient (U) can be calculated with Equation 4.4 and stated values.

1
U = ≈ 2.29 J m - 2 s -1 K -1 (4.4)
1 1 δ
+ +
hi ho k
Where hi = Inside film coefficient
ho = Outside film coefficient
δ = Reactor wall thickness ≈ 0.00635 m
k = Reactor wall thermal conductivity ≈ 0.17 J m-1 s-1 K-1

The inside film coefficient (hi) for packed bed columns with laminar
airflow (Re<2000) was calculated with Equation 4.5-4.6 (Perry, et al. 1997)
83

0.365
λg  DpG 
hi = 3.6   ≈ 42.65 J m -2 s -1 K -1 (4.5)
D p  µ g ε 
Where λg = Gas thermal conductivity ≈ 0.027 J m-1 s-1 K-1 @ 30°C
µg = Gas viscosity ≈ 1275x10-7 Poise @ 30°C
ε = Porosity ≈ 0.35
G = Gas mass velocity ≈ 0.0023 kg m-2 s-1
Dp = Characteristic dimension of particle ≈ 2.5 mm
Dr G
Re =
µ g ≈ 1.8 (4.6)

Where Re = Reynolds number, dimensionless


Dr = Reactor diameter ≈ 0.1 m

The outside film coefficient (ho) for a vertical cylinder exposed to natural
convection is determined with Equations 4.7-4.9 (Perry, et al. 1997).

0.59kair 1

(Gr Pr ) ≈ 2.66 J m- 2 s-1 K -1


4
ho = (4.7)
L
Where kair = Thermal conductivity of air ≈ 0.26 J m-1 s-1 K-1
L = Length of heat transfer surface = 0.3 m

L3 ρ air
2
gβ∆T
Gr = ≈ 1.02x107 (4.8)
µ 2
air

Where Gr = Grashoff number, dimensionless


ρair = Ambient air density ≈ 1.16 kg m-3 @ 30°C
g = Acceleration due to gravity ≈ 9.81 m s-2
β = Volumetric coefficient of Thermal Expansion
= Tamb-1 ≈ 0.0033 K-1
∆T = Temperature difference ≈ 33°C – 30°C = 3°C
µair = Viscosity of ambient air ≈1860x10-7 Poise @ 30°C

C pair µ air
Pr = ≈ 0.72 (4.9)
k air
Where Pr = Pandtl number, dimensionless
Cpair = Heat capacity of air ≈ 1.007 J g-1 K-1
Plugging these calculated and assumed values into Equation 4.3, the heat
generated by reaction can be estimated as:
84

Qgen =7.68x104 J s-1 m3bed

Equation 2.27, in chapter 2, gives the theoretical maximum energy produced


from stoichiometric oxidation of H2S. Assuming that all of this energy is given off to
the bed (which is clearly an over-assumption, as 60% of energy is typically utilized by
the cell for growth and maintenance), it is estimated that a maximum of 4.85x102 J s-1
m3bed may be generated solely from oxidation of 1500 ppm H2S to sulfate. This is
much lower than the calculated heat generation value and indicates that there are likely
additional or different exothermic reactions occurring. This is expected since compost
and other organic compounds in the gas are known to degrade in aerobic
environments. Assuming a worst-case scenario where all of this heat is derived from
the oxidation of methane (∆Hrxn = -8.9x105 J/mol CH4), 47% of the methane would be
consumed. This calculation indicates that there is the potential for significant methane
reduction, but further testing is needed to determine the fate of CH4 as methane was
not monitored and no temperature controls were practiced in this experiment.

4.5. HYDROGEN SULFIDE MEASUREMENTS

4.5.1. Electrochemical Sensor

To quantify the error attributed to the electrochemical sensor and dilution


method, linear error propagation can be done. Summing the variances from the sensor
reading and the flow-meter readings for air and gas entering the dilution chamber, the
cumulative standard deviation, neglecting any cross sensitivity interference, given by
Equation 4.10.

σH 2S
= 1600 + (2.42 ⋅ 10−4 )(Sensor Reading) 2 ≅ ± 40 [ppm] (4.10)
85

Interference would likely raise this error, but the actual amount is difficult to
calculate without knowing all biogas components.
Results from testing the column without packing showed that the inlet and
outlet concentrations remained equal, indicating no apparent interference from the
column materials.
Columns 3 and 4 operated continuously for 1057 hours with an average inlet
H2S concentration around 1500 ppm. Removal efficiency is calculated by dividing the
difference between the inlet and outlet concentrations by the inlet concentration (in
ppm). Similarly, linear error analysis including only instrument variances indicates
that standard deviation for removal efficiency is less than 1.41%. The concentrations
and removal efficiencies for columns 3 and 4 (with associated error bars) are given
with Figures 4.8-4.11.

H2S Concentrations - Column 3

2500
H2S in
2000 H2S out
H2S (ppm)

1500

1000

500

0
0 200 400 600 800 1000
Run Time (total hrs)

Figure 4.8: H2S Concentrations for Trial 3


(lines are polynomial fits meant only to guide the eye)
86

Removal Efficiency - Column 3

100%
90%
80%
Removal Eff. (%)

70%
60%
50%
40%
30%
20%
10%
0%
0 200 400 600 800 1000
Run Time (total hrs)

Figure 4.9: H2S Removal Efficiency During Trial 3


(line is polynomial fit meant only to guide the eye)

H2S Concentrations - Column 4

2500
H2S in
2000 H2S out
H2S (ppm)

1500

1000

500

0
0 200 400 600 800 1000
Run Time (total hrs)

Figure 4.10: H2S Concentrations for Trial 4


(lines are polynomial fits meant only to guide the eye)
87

Removal Efficiency - Column 4

100%

80%
Removal Eff. (%)

60%

40%

20%

0%
0 200 400 600 800 1000
Run Time (total hrs)

Figure 4.11: H2S Removal Efficiency During Trial 4


(line is polynomial fit meant only to guide the eye)

Column 3 operated consistently above 80% removal efficiency until the last
150 hours when efficiencies declined to 50%. Column 4 started off with high removal
efficiency for the first 250 hours but then dipped to around 50% efficiency by 500
hours. Expecting to see a continued reduction, column 4 was allowed to continue
running but displayed a temporary rebound in removal efficiency to 80% by hour 900.
Similarly to column 3, a drop in efficiency occurred for the last 150 hours of operation
and both columns were shut down for media inspection when the removal efficiency
of both columns dropped below 50%.
The elimination capacity of columns 3 and 4 ranged from 24 – 112 and 16 –
118 g H2S/m3-solids/hr, respectively, as calculated assuming atmospheric pressure and
25°C. The total mass of H2S removed from the gas during these experiments are 135
and 127 g H2S, respectively for columns 3 and 4. An average flowrate and
concentration per sampling interval were used for calculation. This elimination
88

capacity approaches the maximum of 130 g H2S/m3-solids/hr reported for organic


media (Yang and Allen 1994; Degorce-Dumas, et al. 1997).
It is not clear why one column behaved differently than the other under similar
conditions. One possible explanation could assume inherent variability due to
biological systems. The fact that a rebound in removal efficiency occurred in column
4 may support the idea of biological system upset and recovery. A purely chemical
breakthrough would not be expected to behave in this manner.
The dip in efficiency in column 4 around hour 500 is also accompanied by a
lesser dip in efficiency in column 3. The exact reason for these reductions are not
known, but maximum bed-temperatures during this time exceeded 40°C, which may
have disrupted a biological or physical adsorption mechanism. Temporarily
insufficient moisture availability may also be a possible explanation.
It appears that the simultaneous decline in removal efficiencies of both
columns after hour 900 occurs in conjunction with a measured increase in inlet gas
H2S concentration. At hour 900, a minimum inlet H2S concentration of 480 ppm was
recorded, followed by an increase to over 2000 ppm in 6 days. This quick increase in
loading may also have had the effect of upsetting the system and leading to decreased
removal efficiency. It is also possible that the chemical or physical reaction
mechanism was becoming exhausted at this point.
The inlet and outlet H2S concentrations and removal efficiencies for columns 5
and 6 (with associated error bars) are represented in Figures 4.12-4.13. Removal
efficiencies were above 90% and 87% for columns 5 and 6, respectively, throughout
their operation.
89

H2S Concentrations and Removal Efficiency - Column 5

2500 100%

2000 H2S in 80%

Removal Eff. (%)


H2S out
H2S (ppm)

1500 60%
Removal Eff
1000 40%

500 20%

0 0%
0 10 20 30 40 50
Run Time (total hrs)

Figure 4.12:H2S Concentrations and Removal Efficiency for Column 5


(lines are polynomial fits meant only to guide the eye)

H2S Concentrations ad Removal Efficiency - Column 6

3000 100%
2500 80%
2000 H2S in Removal Eff. (%)
H2S (ppm)

H2S out 60%


1500 Removal Eff.
40%
1000
500 20%

0 0%
0 50 100 150 200
Run Time (total hrs)

Figure 4.13:H2S Concentrations and Removal Efficiency for Column 6


(lines are polynomial fits meant only to guide the eye)
90

4.5.2. Gas Detector Tubes

Gas samples from AA Diary were sporadically tested for H2S using lead-
acetate H2S detector tubes since November 2000. Table 4.3 shows H2S readings in the
gas prior to the experiments, and more frequently during the experiments (July-Aug.).

Table 4.3: H2S Gas Detector Tube Readings for AA Dairy Raw Digester Gas
November 13, 2000 3600
March 4, 2001 2200
July 1, 2001 3400
July 13, 2002 1400 (660)
July 15, 2002 1400 (1380)
July 20, 2002 1300 (1680)
July 27, 2002 1150 (1440)
August 5, 2002 1200 (1280)
August 19, 2002 1700 (1900)
August 22, 2002 1900
Note: Numbers in (parentheses) are H2S levels determined by the electrochemical
sensor on a 2:1 biogas:air sample. Provided for comparison.

Levels as high as 3600 ppm have been measured. H2S levels in the digester
gas during the experiment appear significantly lower than in the past. This difference
may be due to seasonal variation, variability in the feedstock, chemical differences in
the digester, or sampling error.
The readings from the gas detector tubes are consistently lower that expected
when compared with the electrochemical sensor method. One would expect the
electrochemical sensor method to produce a reading value 2/3 lower than the detector
tube, because the gas sample measured by the electrochemical sensor is diluted with
air, and the H2S detector tubes are always used to measure raw biogas. The low
readings might be attributed to normal error within the reported +/- 25% for the gas
detector tubes, or additionally from an incomplete seal from the sampling port to the
tube itself, allowing for gas to bypass the tube, resulting in a lower reading.
91

4.5.3. Gas Chromatography

A calibration standard gas was not tested with the biogas sample so
quantitative levels of each gas component are not accessible. Rather, a qualitative
assessment of the major gas components in biogas was completed, as follows.
By taking the relative contributions of the peak readings, rough estimates of
gas composition can be determined, as presented in Table 4.4.

Table 4.4: GC-MS Results for AA Dairy Digester Gas


Component Peak Height % Total % Dry Basis
8
Total 1.80*10 -- --
Methane 1.01*108 56.24% 62.49%
Carbon dioxide 6.00*107 33.33% 37.04%
5
Hydrogen sulfide 5.90*10 0.33% 0.36%
NH3 and CH4 isotopes 1.75*105 0.10% 0.11%
2
Dimethyl Sulfide 6.60*10 <0.01% <0.01%
Water 1.80*107 10.00% --

As seen here, the relative moisture content of the gas in this sample is about
10%. On a dry-basis, the methane and carbon dioxide concentrations are about what
was expected, ~60% and ~40%, respectively. The H2S content is also in the range of
a few tenths of a percent, which is in agreement with the other testing methods used
here. Other compounds were not identified due to the absence of specific calibration
gases for testing these compounds.
Figure 4.14 depicts the actual GC-MS results for the digester gas. Each graph
is for a specific mass element. Results for H2S, CO2, total gas, and water are
represented from top to bottom, respectively. Similar results for dimethyl sulfide and
ammonia are not shown here. Each graph is scaled to 100% of the signal intensity,
and the peak height value is given with the scales to the right.
H2 S

CO2

Total

H2 O

Figure 4.14: GC-MS Results for AA Dairy Digester Gas.


93

4.6. BIOGAS-EXPOSED-COMPOST ASSESSMENT

Columns 3 and 4 were opened 7 days after shutdown for compost inspection.
The solids slid out of the columns easily and appeared dry and whitish-yellow in color
as seen in Figure 4.15. No testing was done to determine if this coloration had any
correlation to sulfur content. The smell of rotten eggs was evident during opening.

Figure 4.15: Pictures of Columns after Exposure to Biogas for 1057 hours.

4.6.1. Moisture

The outer layers of the columns were crumbly to the touch and showed definite
signs of drying. Both compost volumes were sliced lengthwise for examination. In
94

both instances, the cores of the columns were darker brown in color, moist to the
touch, and appeared similar to the original compost.
Moisture measurements from 0.05, 0.15, and 0.25 m along the column lengths
are shown in the Table 4.5.

Table 4.5: Moisture Contents Along Bed Depth


Sample / Location Bottom Middle Top
Fresh Compost 73% 73% 73%
Column 3 44% 63% 76%
Column 4 41% 69% 60%

As seen, there is moisture loss near the inlet, down from an original content of
~70%. As expected with a temperature rise in the bed, humidification of the inlet gas
was not sufficient for maintaining moisture in the compost. Also, the evidence of
localized drying indicates preferential gas flow around the edges of the column.
Water content is an indirect measurement of water availability, whereas water
activity measurements are more useful, but difficult to obtain. Sufficient water
availability is critical for proper biological function and may have the greatest
detrimental effect on biofilter performance if inadequate (Mysliwiec, et al. 2001).
Studies by Yang and Allen (1994) indicate that removal efficiency drops linearly
below 30% wet weight water content in H2S biofilters. Although no water content this
low was measured, it is reasonable to assume that better performance could be
achieved with more uniform and increased wetting. Additionally, an improved water
management technique is desirable because organic media are known to become
hydrophobic when dried, making them hard to rewet and maintain acceptable water
potentials (Bohn and Bohn 1999).
95

4.6.2. pH

As with the moisture content, pH samples from 0.05, 0.15 and 0.25 m along
the column lengths were recorded and Table 4.6 shows these results.

Table 4.6: pH Levels Along Bed Depth


Sample / Location Bottom Middle Top
Fresh Compost 7.2 7.2 7.2
Column 3 6.9 4.9 4.6
Column 4 6.7 7.1 4.6

The most significant pH drops are noticed in the upper half of the beds. The
largest pH decreases are noticed where moisture contents remained higher. A drop in
pH is an indication of possible sulfate formation in the bed, which is a major
biological pathway for sulfide oxidation. One possible explanation for this might be
that biological activity, and thus sulfate formation, was curtailed by low moisture
content.
An extreme pH drop, as seen with some H2S biofilters (Devinny, et al. 1999),
was not noticed here. This may be due to the incomplete oxidation to elemental sulfur
or natural buffering capacity of the compost. Additional buffering to maintain a
neutral pH was not practiced, but could easily be achieved via crushed limetone
addition in the bed.

4.6.3. Trace Element Analysis

The results for the trace element analysis are shown in the Table 4.7.
96

Table 4.7: Elemental Analysis of Raw and Tested Compost


Element Compost* Sample 3* Sample 4*
Sulfur 7,316 103,752 121,490
Aluminum 2,145 774 849
Arsenic < det. < det. < det.
Boron 67 43 47
Cadmium < det. < det. < det.
Calcium 49,910 28,884 35,858
Chromium 2 < det. 0.1
Cobalt 2 < det. < det.
Copper 479 332 359
Iron 5,487 2,160 2,481
Lead < det. < det. < det.
Magnesium 13,615 8,892 11,934
Manganese 500 320 353
Molybdenum 11 15 21
Nickel 39 25 28
Phosphorous 11,869 7,261 8,068
Potassium 13,548 9,412 10,020
Sodium 3,277 2,356 2,528
Vanadium 7 3 3
Zinc 347 236 256
* all values in ug/g dry basis
< det. = below analyte detection limit

As seen, the original total sulfur content of the compost is about 7000 µg/g
(0.7%) on a dry basis. Columns 3 and 4 average around 110,000 µg/g for total sulfur
content. This indicates a net increase of about 100,000 µg/g of total sulfur in the
media itself during operation, and final sulfur content over 10% dry basis. These
results are encouraging because they confirm that sulfur is being removed from the gas
stream and being sequestered in the bed through biological, chemical or physical
mechanisms.
Besides an average 1440% increase in sulfur, molybdenum is the only other
element with a measured increase (66% average). All the remaining measured
97

elements decreased by less than 100% on a dry basis. Some metal leaching is
expected to occur in acidic biological environments, but was not measured here.
Yang and Allen (1994) have shown that sulfate levels above 30 mg/g inhibit
H2S biofilter performance. The original level in the compost (<100 µg/g) was well
below this threshold. Sulfate levels in the used composts were not measured.
An accurate sulfur balance is hard to accomplish because the final dry mass of
the columns and sulfur levels in leachate were not measured. The amount of sulfur
removed from the gas stream using the electrochemical measurement method is
calculated as 127 ± 15 g for column 3, and 120 ± 15 g for column 4. Assuming that
the dry mass of the columns stayed the same throughout operation, the trace element
analysis would predict an expected removal of 175 and 207 g S for columns 3 and 4,
respectively. The discrepancy may be due to a decrease in total dry matter during
operation, or significant loss of sulfur in the leachate. Also, the detected smell of H2S
during column opening might be an indication of sulfur reduction occurring and re-
generation of H2S while the columns were no longer exposed to an oxygen source.

4.7. DISCUSSION

These results indicate that a simple and minimally managed system, comprised
of passing a mixture of 2:1 biogas-to-air through cow-manure compost, can be
effective in removing sulfur from a biogas stream. Although using an electrochemical
sensor is affordable and good for tracking relative changes in concentration, it is
limited in measuring absolute concentrations because of cross-sensitivity interference
and dilution error. H2S gas detector tubes and a GC-MS analysis corroborate the
readings from the electrochemical sensor.
98

It is not clear from this study whether sulfur removal is due mostly to
biological, chemical, or physical phenomena, but sulfur is clearly being removed from
the gas stream and accumulated in the compost. The upset and recovery trend in
removal efficiency for column 4, noticeable pH drops, prolonged performance, and
relatively high elimination capacities are suggestive of a biological mechanism.
The fact that significant removal rates were achieved with minimal moisture
control and no temperature or pH controls indicate that further research on system
capacity and optimization are warranted. Initial progress is predicted simply by
maintaining temperature between 30-40°C and moisture contents above 50%. Other
researchers have achieved biological H2S removal with residence times as low as 1.6
seconds, indicating that higher elimination capacities may be possible with lower gas
residence times (Gabriel, et al. 2002). Major questions to be answered include: What
happens to methane during this process? What is the upper elimination capacity? What
sulfur products are formed? How does the compost change during the process? What
is the microbial makeup? And what are the optimum operating conditions?

4.8. SCALE-UP CONSIDERATIONS

These experiments were conducted on biogas side-streams containing roughly


1/1000 of the full-scale gas flow. 2.5-liter reaction columns with 100-second empty-
bed residence times for 2:1 biogas-to-air mixtures produced the aforementioned
results. Assuming linear scale-up for discussion, a 2500 L, or 2.5 m3, bed volume
might produce similar results. Although linear scaling would most likely not apply,
there are many reasons to believe that performance can be improved by optimizing
temperature, moisture, loading, and pH conditions.
99

Additionally, in these experiments, excess oxygen (in air) was added, but the
critical amount of air required may be substantially less. Reduced oxygenation levels
and empty-bed residence times, as determined with further testing, would decrease bed
size, as long as the maximum H2S elimination capacity was not exceeded.
Unfortunately, lower gas-residence times may require bulking of the compost to keep
pressure drops reasonable, and bed sizes would then need to be increased.
In scaling-up, there are two potential process-configurations that can be
envisioned with cow-manure compost. Ideally, a biological process with a stable
microbial environment would be established to promote continuous and prolonged
H2S removal. In this scenario, required nutrients are available in the compost or gas,
or added periodically, and sulfates and other metabolites are removed, when
necessary, with a water-rinse. Labor and media costs are lower because bed change-
outs would only be necessary less than once a year. A second configuration is
envisioned where the cow-manure compost is placed in a vessel with relatively few
process controls, operated until the bed is completely ineffective, and replaced with
fresh compost as a batch process. The loss in efficacy may be from exhaustion of
chemical or physical adsorption capacity, loss of biological activity from insufficient
nutrients, or because of increased pressure drop from compost degradation or plugging
of the void spaces with sulfur or biomass.
A comparison of these compost-based processes with an iron sponge system,
sized for AA Dairy, is included in Table 4.8. Media costs for the compost are based
on the average market value of $0.02/kg for AA Dairy cow-manure compost.
Transportation costs are zero for the compost, whereas the iron sponge has a
transportation cost for shipping from Chicago, IL. As seen, both manure processes
show potential for improvement over iron sponge. The Equivalent Uniform Annual
Cost (EUAC) is calculated assuming a $30,000 capital cost for all systems, 20-year
100

system life cycle, and 8% interest rate and including all other annual costs (Newnan
2000). The iron sponge system has an EUAC from $4,437 - $15,057 per year or $5.35
- $6.43 per kg H2S removed. The cow-manure systems have estimated EUAC’s of
$6,527 and $3,832 per year, respectively, or $6.30 and $3.70 per kg H2S removed.
There is a range of costs for the iron sponge system due to variations from low (1000
ppm H2S) to high (4000 ppm H2S) loading. Since only one concentration level (1500
ppm H2S average) was tested here, it is not clear how the cow-manure compost
systems would respond, and subsequently how costs would vary, based on different
loadings. Therefore, the symbol “±” is used to remind the reader of this variability in
Table 4.8. As indicated here and in the table, the continuous cow-manure system
appears to provide the most significant improvement and savings over the iron sponge
system.
101

Table 4.8: Estimated Comparison of Cow-Manure-Compost and Iron-Sponge


H2S-Removal Systems at AA Dairy
Cow-Manure Cow-Manure
Iron Sponge
(Batch) (Continuous)
CHARACTERISTICS
# Vessels 2 (Lead/Lag) 2 (Lead/Lag) 2 (Series)
0.91 m ID x 1.52 m 0.91 m ID x 1.77 m 0.91 m ID x1.77 m
Vessel Sizes
high (0.99 m3 each) high (1.15 m3 each) high (1.15 m3 ea.)
Biogas Flow Rate 0.94 m3/min 0.94 m3/min 0.94 m3/min
Air Addition Rate 2.4 – 3.7 % < 33 % < 33 %
3 3
Total Flow Rate 0.97 m /min 1.41 m /min 1.41 m3/min
Empty-Bed 60 sec (1-bed) 49 sec (1-bed) 49 sec (1-bed)
Residence Time 120 sec (both beds) 98 sec (both beds) 98 sec (both beds)
Mass of Bed 800 kg each 860 kg each 860 kg
Bed Life 18 - 315 days 40 days ± > 1 year
Annual Media
930 – 16,300 kg 15,480 kg ± < 1720 kg
Utilization
COSTS
Capital $10,000-$50,000 $10,000-$50,000 $10,000-$50,000
Media (per year) $250 - $4,300 $300 ± $35 ±
Transportation
$500 - $2,500 0 0
(per year)
$500 ±
$500 - $3,000 $1,000 ±
Operating and (1 change-out @ 8
Labor (per year)
(1-20 change-outs (9 ± change-outs @
hrs. + blower
@ 12 hrs. ea. + 8 hrs. ea. + blower
electric + H2O +
blower electric) electric + H2O)
nutrients)
Media Disposal
$130 - $2,200 ± $2,170 ± $240 ±
(per year)*
TOTALS
Total EUAC**
($/kg H2S $5.35-6.43 $6.30 ± $3.70 ±
removed)
* Kellog (1996) estimated disposal costs of $0.14/kg-spent media. It is assumed here that disposal
costs for all three processes would be the same. Costs may be higher or lower depending on whether
the material is disposed of on-site or if it is transported to a landfill and treated as hazardous waste.
(See Appendix A).
** EUAC = Equivalent Uniform Annual Cost assuming $30,000 capital cost, 8% interest rate, and 20
year life-cycle.
CHAPTER

5. SUMMARY AND CONCLUSIONS

The two-part objective of this study was to determine currently available H2S
removal technologies suitable for use with farm biogas systems, and to test the
feasibility of utilizing on-farm cow-manure compost as an H2S adsorption medium.
There are many chemical, physical, and biological methods currently available for
removal of H2S from an energy gas stream, as summarized below.

5.1. CURRENTLY AVAILABLE H2S REMOVAL METHODS

5.1.1. Dry-Based Processes

Traditionally, dry-based chemical methods have been used for sulfur removal
from gas streams with less than 200 kg S/day, and still appear to be competitive
because they are simple and effective. Estimated capital costs for farm systems are
around $10,000-$50,000 and media costs estimated between $250-$23,840 per year.
Relatively high labor costs for materials handling and disposal are incurred. Major
drawbacks include a continually produced waste stream of spent media, and growing
environmental concern over appropriate waste disposal methods. The most
competitive products appear to be Iron sponge, Media-G2®, and KOH-impregnated
activated carbon. Molecular sieves may be competitive with an appropriately
designed regeneration step. Table 5.1 summarizes dry-based, H2S removal processes
and Table 5.2 summarizes estimated costs for application at AA Dairy.

102
Table 5.1: Summary Table Comparing Dry-Based H2S Removal Processes for Farm Biogas
Operating Compounds Regen- Media Costs
Packing ($/kg H2S Notes
Conditions Removed erable? removed)
Changeout can be labor intensive.
Ambient Temp. (65- 2-3 times in
Iron Sponge H2S and Variability in process efficiency but
115 F), Wet gas, 60 batch mode 0.35 - 1.55
(Iron Oxide) mercaptans much practical experience with
sec residence time only
digester gas
Ambient Temp. (65- Non-pyrophoric and easier handling
Sulfa Treat® H2S and
115 F), Wet gas, 60 No 4.85 - 5.00 characteristics compared to iron
(Iron Oxide) mercaptans
sec residence time sponge
Ambient Temp. (65-
Sulfur Rite® H2S and Available in prepackaged modules,
115 F), Wet gas, 60 No 7.95 - 8.50
(Iron Oxide) mercaptans Forms iron pyrite
sec residence time
Ambient Temp. (65- 15 times in
Media-G2® H2S and Requires multiple regenerations to
115 F), Wet gas, 60 batch mode 2.90 - 3.00
(Iron Oxide) mercaptans obtain estimated removal efficiency
sec residence time only
Removes all water before sulfur
Ambient Temp., Water,mercap-
Molecular Sieve compounds. The proposed design
High Pressure (500 tans, H2S, Yes
(Adsorbent) would have a 1-day bed life operated
psig+) slight CO2
at 500 psi
Impregnated
Ambient Temp. and H2S and KOH carbon has been used
Activated Carbon No 1.75 - 2.00
Pressure mercaptans successfully for anaerobic digester gas
(Adsorbent)
Continuous or batch processes are
Estimated Cow-
Ambient Temp. and envisioned (Table 4.8). It is not clear
Manure Compost H2S No 0.03 – 0.29
Pressure whether other compounds are
Processes
oxidized in addition to H2S.
Table 5.2: Summary Table Comparing Dry-Based H2S Removal Processes for AA Dairy
AA DAIRY ESTIMATES (all costs do not include taxes, shipping, installation or operation)
Estimated
# of Vessels Annual Media Annual Media
Packing Bed Life (days) Capital Suppliers
and Sizes Consumption (kg) Costs ($)
Cost
(D=diameter, Low High Low High Low High (Vessels (Low loading=1000ppm
H=Height) loading Loading loading Loading loading Loading Only) ($) High loading=4000 ppm)
two - 0.91 m Connelly GPM,
Iron Sponge 930 - 3,710 - 250 - 985- 10,000 -
D x 1.52 m H 72 - 315 18 - 79 Physichem, Varec
(Iron Oxide) 4,070 16,300 1,075 4,300 50,000
each Vapor Control
two - 1.2 m
Sulfa Treat®
x1.6 m x 1.8 345 86 3,850 15,450 3,400 13,500 12,000 SulfaTreat
(Iron Oxide)
m each
®
Sulfur Rite one – 2.3 m
420 98 7,900 33,900 5,560 23,840 43,600 US Filter/ Merichem
(Iron Oxide) D x 3.4 m H
two - 0.91 m
Media-G2® 10,000 -
D x 1.52 m H 190 47 1,460 5,900 2,050 8,290 ADI International
(Iron Oxide) 50,000
each
Molecular
one – 0.6 m Many - See
Sieve 24 hours 24 hours 250/day 400/day n/a n/a n/a
Dx2mH Thomasregister.com
(Adsorbent)
Impregnated
two – 0.6 m
Activated Many - See
D x 1.5 m H 340 85 270 1075 1250 5435 <50,000
Carbon Thomasregister.com
each
(Adsorbent)
Estimated one or two – Note: (low and high
Cow-Manure 0.91 m D x 10,000 – loading values are for
40 360 1720 15480 35 300
Compost 1.77 m H 50,000 batch and continuous
Processes each process, respectively)
105

5.1.2. Liquid-Based Chemical and Physical Processes

These processes have not traditionally been used for low-flow and low-
pressure applications due to increased costs of operating at high pressure, high energy-
requirements for recirculation pumps and regeneration vessels, and higher media
costs. The LO-CAT® iron-chelate process shows potential for being viable with small
biogas systems, especially if the thickened sulfur-slurry produced has agricultural
value. Also, nitrate solutions, such as Sulfa-Scrub®, should be investigated for system
suitability. Water scrubbing and the Selexol Process are both used for upgrading of
biogas to natural gas but do not appear economically competitive for selective H2S
removal at this time.

5.1.3. Membrane Separation

Selective H2S removal from biogas is not yet practical with this technology,
but upgrading of biogas to natural gas is an emerging market for membrane processes.
Low-pressure gas-liquid membrane systems are in the developmental stage and show
promise for the future.

5.1.4. In-Situ Digester Sulfide Control

Research has shown that addition of iron compounds to the digester effectively
inhibits some H2S production, but this method may only be considered a partial
removal process, requiring another technology for removal down to about 10 ppm.
Also, there is concern about solids buildup in the digester using this method.
106

5.1.5. Biogas Aeration

Digestion facilities in Europe have reported significant H2S removal by adding


small amounts of air (2-6%) to the digester headspace or storage tank. This extremely
simple and affordable technique is believed to be biological in nature, and may reduce
initial H2S levels to around 100 ppm. This process requires further testing in the
United States as it shows great promise.

5.1.6. Biological Removal Techniques

Operational experience with biological oxidation of H2S from gas streams is


growing and there are companies, such as Biothane and UOP, who specialize in this
area. Currently available systems are more capital intensive than available dry-based
systems, but offer advantages of lower labor costs and improved environmental
impact. The research in the second part of this study examined the feasibility of using
cow-manure compost as an economic and effective H2S removal process for integrated
farm energy systems.

5.1.7. Comparison of Characteristics of H2S Removal Processes

Table 5.3 summarizes the characteristics of available H2S removal processes


from biogas.
107

Table 5.3: Summary of H2S Removal Process Characteristics


Applicability Operating/ H2S < Environ-
Removal Capital Ease of Regen-
for farm Media 250 mental
Technique Costs operation erable
biogas Costs ppm impact
(+=high) (+=low) (+=low) (+=easy) (+=yes) (+=yes) (+=low)
Iron Oxides + + +/- +/- +/- + -
Zinc Oxides - + - +/- - + -
Alkaline Solids - + - +/- - + -
Impregnated
Activated + + +/- +/- - + -
Carbons
Molecular Sieves +/- +/- - - + + -
Chelated Iron
+/- - - - +/- + +/-
Solutions
Nitrite Solutions +/- + - + - + -
Alkaline Salt
- - - - - + -
Solutions
Amine Solutions - - - - + + -
Water Scrubbing +/- +/- +/- +/- + +/- -
Physical Solvent
+/- +/- +/- +/- + +/- -
Scrubbing
Membrane: Low
+/- - - - +/- +/- +/-
Pressure
Membrane: High
- - - +/- n/a +/- +
Pressure
Digester pH
+/- + +/- +/- n/a - +/-
Control
Digester Iron
+ + + + n/a - +/-
Addition
Dietary
+/- + + +/- n/a - +
Adjustment
Air Dosing to
+ + + + n/a +/- +
Biogas
Commercial
Biological +/- - +/- +/- +/- + +
Processes
Estimate Cow-
Manure Compost + + + + n/a + +
Processes
n/a = not applicable desirable
+/- = neutral - = undesirable
108

5.2. TESTING OF COW-MANURE COMPOST

Initial testing of cow-manure compost indicates that it has potential as an


effective and economic medium for H2S removal. PVC test columns were constructed
and a 2:1 biogas-to-air mixture passed through the columns containing anaerobically
digested cow-manure compost. The tests of most significance were run for 1057
hours with an empty-bed gas-residence time near 100 seconds and inlet H2S
concentrations averaging 1500 ppm, as measured by electrochemical sensor with a
40:1 sample dilution.
Removal efficiencies over 80% were recorded for the majority of the trial.
Elimination capacities recorded were between 16–118 g H2S/m3-solids/hr. This is
significant considering only minimal moisture and no temperature or pH controls were
implemented. Temperature in the bed varied from 19-43°C and the moisture contents
in the spent column ranged from 41-70%, with pH values from 4.6 to 6.9. It is not
clear whether the major mechanism for sulfur removal from the gas stream is
biological, chemical or physical, but it is known that the sulfur content in the compost
increased by over 1400%, verifying sequestration of sulfur in the solid. These initial
results indicate that future work is warranted for examining the suitability of cow-
manure compost as a biofiltration medium for use with biogas.
Two potential process-configurations are envisioned for scale-up with cow-
manure compost. In one, continuous biological activity is promoted by optimizing
process conditions for long-term operation. In the other, a relatively simpler batch
system with fewer process controls is established, where the compost is used for its
chemical, physical, or neglected biological activity, and changed-out more frequently.
Both processes compare favorably to other dry-based process, as illustrated in Tables
4.8, 5.1, 5.2, and 5.3, assuming minimal methane oxidation.
CHAPTER

6. FUTURE WORK AND RECOMMENDATIONS

Although the present study partially fulfills its objectives, there are limitations
that should be addressed and new directions to be explored in future research. The
biogas summary completed in the background section of this study could be expanded
in the following areas:
• A Life Cycle Assessment (LCA) comparison of the economic, environmental
and social impacts for the most competitive H2S removal technologies should
be performed. Operation and maintenance costs, as well as appropriate
disposal costs, should be included.
• Alternative disposal and reuse techniques for spent adsorption media should be
studied. The agricultural nutrient-value of spent media and biological
regeneration techniques for sulfur clogged iron sponge should be investigated.
• An LCA for other biogas purification processes, such as CO2 reduction, water
removal, particulate filtration and removal of other gas contaminants, should
be conducted.
• The processing requirements for specific gas-utilization technologies should be
compiled. Specific uses should include biogas in boilers, engines,
microturbines, fuel cells, Stirling engines, upgrading biogas to natural-gas
quality or using biogas for hydrogen production, among others
This study was also effective as a proof-of-concept, indicating that cow-
manure compost can be used as economical and effective H2S adsorption media.

109
110

Further testing and verification of these results are necessary and the following
experimental modifications are recommended:
• Measurement of gaseous sulfur compounds should be done via gas
chromatography with a flame photometric detector for increased accuracy.
• Experiments should take place in a controlled temperature environment to
minimize variation due to temperature fluctuations.
• Moisture should be maintained around 50% wet basis. The implementation of
downward water flush could maintain bed moisture and serve to rinse out any
sulfur accumulating in the bed.
• Clear columns should be used to observe any drying or microbial plugging that
might be occurring.
Additional testing needs to be done to better determine and quantify the
reactions taking place. The following tests should be performed.
• The fate of methane during operation should be monitored and determined.
• The sulfur species in the medium and effluent gas, including sulfates and
elemental sulfur, should be measured to account for sulfur reactions.
• Jar tests should be performed on abiotic samples of cow-manure compost in
order to assess the purely physical and chemical adsorption capacity for H2S.
• The effect of aeration of the biogas should be quantified.
• A biological assessment of the major active microbial communities should be
performed.
• An assessment of the agricultural nutrient value, compost stability, and metal
leaching in the spent filter media could be performed.
Further long-term operation and bench-scale optimization are desired before
scale-up to pilot and full scales. A life cycle assessment should then be conducted for
determining overall economic and environmental benefits.
APPENDIX A: H2S Scavenger Media Disposal

A study by the Gas Research Institute estimated 1996 change-out and disposal
costs for various media as follows (Kellog 1996). These costs are merely given in
Table A.1. to show that disposal costs can easily exceed media costs, as are expected
for transportation, installation and operating costs.

Table A.1. Approximate Media Change-out and Disposal Costs (1996 est.)
Caustic $0.25/gal
Triazines $1.00-2.00/gal
Non-regenerable Amines $1.00-2.00/gal
SulfaTreat $0.10-0.30/lb
SulfaScrub(Nitrates) $0.10-0.20/lb
Iron Sponge $0.10-0.30/lb
Source: Kellog (1996)

The Gas Research Institute sponsored extensive research on economic analysis


of sulfur scavenging processes for gas streams with low sulfur production in the early
to mid 1990’s. The reader is directed to three papers for further information: 1)
Evaluation of H2S Scavenger Technologies, (Foral and Al-Ubaidi 1994). 2) GRI Field
Evaluation of Liquid-Based H2S Scavengers in Tower Applications at a Natural-Gas
Production Plant in South Texas, (Fisher and Dalrymple 1994). 3) Field Evaluation of
Solid-Based H2S Scavengers for Treating Sour Natural Gas, (Fisher and Shires 1995).
These works resulted in a design program, GRI-CalcBase, which can be used
for economic evaluation of different scavengers with various plant configurations.
Results of running the program with the system parameters from AA Dairy indicated
that dry scavengers such as Iron Sponge and SulfaTreat were the most promising to
investigate further.

111
REFERENCES

Anerousis, J. P. and S. K. Whitman (1985). "Iron Sponge: Still a Top Option for Sour
Gas Sweetening." Oil and Gas Journal February 18: 71-76.

Basu, R., E. C. Clausen, et al. (1996). "Biological Conversion of Hydrogen Sulfide


into Elemental Sulfur." Environmental Progress 15: 234-238.

Bohn, H. L. and K. H. Bohn (1999). "Moisture in Biofilters." Environmental Progress


18: 156-161.

Bohn, H. L. and H.-C. Fu-Yong (1989). "Hydrogen Sulfide Sorption by Soils." Soil
Sci. Soc. Am. J. 53: 1914-1917.

CADDET (2001). Upgrading of Biogas to Natural Gas - Technical Brochure No. 154.
Centre for Renewable Energy (CADDET), Oxfordshire, UK: 4 p. www.caddet-
re.org/assets/no154.pdf

Cadenhead, P. and K. L. Sublette (1990). "Oxidation of Hydrogen Sulfide by


Thiobacilli." Biotechnology and Bioengineering 35: 1150-1154.

Capstone Turbine Corporation (2002). C30 Biogas Spec-Sheet . Chatsworth, CA: 3 p.


www.microturbine.com/documents/specsheetlandfill.pdf.

Cardenas-Gonzalez, B., S. J. Ergas, et al. (1999). "Characterization of Compost


Biofiltration Media." Journal of the Air and Waste Management Association 49: 784-
793.

Carlson, D. A. and C. P. Leiser (1966). "Soil Beds for the Control of Sewage Odors."
Journal of Water Pollution Control Federation 38(5): 829-840.

Chi, C. W. and H. Lee (1973). "Natural Gas Purification by 5A Molecular Sieves and
its Design Method." AIChE Symposium Series 134(69): 95-101.

Chiang, A. S. T. and Y.-W. Chen (1987). "Selective Removal of Hydrogen Sulfide


Pollutant with Zinc oxide." Proc. of 4th Air Pollution Control Technical Conference.
National Taiwan University, Taipei: 199-206.

Cho, K. S., M. Hirai, et al. (1992). "Degradation of hydrogen sulfide by Xanthomonas


sp. strain DY44 isolated from peat." Applied and Environmental Microbiology 58(4):
1183-1189.

112
113

Cho, K.-S., M. Hirai, et al. (1992). "Enhanced Removal Efficiency of Malodorous


Gases in a Pilot-Scale Peat Biofilter Inoculated with Thibacillus thioparus DW44."
Journal of Fermentation and Bioengineering 73: 46-50.

Cho, K. S., H. W. Ryu, et al. (2000). "Biological Deodorization of Hydrogen Sulfide


Using Porous Lava as a Carrier of Thiobacillus thiooxidans." Journal of Bioscience
and Bioengineering 90: 25-31.

Chou, M. S. and W. H. Cheng (1997). "Screening of Biofiltering Material for VOC


Treatment." Journal of the Air and Waste Management Association 47: 674-681.

Chung, Y. C., C. Huang, et al. (2001). "Biotreatment of Hydrogen Sulfide- and


Ammonia- Containing Waste Gases by Fluidized Bed Bioreactor." Journal of the Air
and Waste Management Association 51: 163-172.

Chung, Y. C., C. Huang, et al. (1998). "Comparison of Autotrophic and Mixotrophic


Biofilters for H2S Removal." Journal of Environmental Engineering 124: 362-367.

Chung, Y. C., C. Huang, et al. (1996). "Biodegradation of Hydrogen Sulfide by a


Laboratory-Scale Immobilized Psedomonas putida CH11 Biofilter." Biotechnology
Progress 12(6): 773-778.

Chynoweth, D. P. and R. Issacson (1987). Anaerobic Digestion of Biomass. Elsevier


Science Publishing Co., Inc., New York: 279 p.

City Technolgy Limited (2002). Cross Sensitivity Table for 7H Series Electrochemical
H2S Sensors. Portsmouth, UK: 1 p.

Cole Parmer (2002). Chemical Resistance Chart from the Cole-Parmer Catalog
Vernon Hills, IL: R17-R26.

Cork, D. J., R. Garunas, et al. (1983). "Chlorobium limicola forma thiosulfatophilum:


Biocatlyst in the Production of Sulfur and Organic Carbon from a Gas Stream
Containing H2S and CO2." Applied and Environmental Microbiology 45(3): 913-918.

Crynes, B. L., Ed. (1978). Chemical Reactions as a Means of Separation: Sulfur


Removal. Chemical Processing and Engineering Series. Marcel Dekker, Inc., New
York: 345 p.

Dawson, D. S. (1993). "Biological Treatment of Gaseous Emissions." Water


Environment Research 65: 368-371.

Degorce-Dumas, H. R., S. Kowal, et al. (1997). "Microbiological Oxidation of


Hydrogen Sulphide in a Biofilter." Canadian Journal of Microbiology 43: 263-271.
114

Devinny, J. S., D. E. Chitwood, et al. (1999). "Co-Treatment of of VOC's in Low-pH


Sulfide Biofilters." Paper presented at Air and Waste Management Associations 92nd
Annual Meeting and Exhibition, St. Louis, Missouri: 9 p.

Devinny, J. S., M. A. Deshusses, et al. (1999). Biofiltration for Air Pollution Control.
Boca Raton, FL, Lewis Publishers: 299 p.

Elias, A., A. Barona, et al. (2002). "Evaluation of a packing material for the
biodegradation of H2S and product analysis." Process Biochemistry 37(8): 813-820.

Eriksen, K., T. Jensby, et al. (1999). "The Upgrading of Biogas to be Distributed


Through the Natural Gas Network - Environmental Benefits, Technology, and
Economy." The Academician Online 1(1). www.chartered.net/biogas.htm.

Fisher, K. S. and D. A. Dalrymple (1994). GRI Field Evaluation of Liquid-Based H2S


Scavengers in Tower Applications at a Natural Gas Production Plant in South Texas.
Chicago, Illinois, Gas Research Institute: GRI Report #94/0437.

Fisher, K. S. and T. M. Shires (1995). Field Evaluation of a Solid-Based H2S


Scavenger for Treating Sour Natural Gas. Chicago, Illinois, Gas Research Institute:
GRI Report #94/0161.

Foral, A. J. and B. H. Al-Ubaidi (1994). Evaluation of H2S Scavenger Technologies.


Chicago, Illinois, Gas Research Institute: GRI Report #94/0197.

Froning, H. R., R. H. Jacoby, et al. (1964). "New K-Data Show Value of Water
Wash." Hydrocarbon Processing and Petroleum Refiner 43(4): 125-130.

Furusawa, N., I. Togashi, et al. (1984). "Removal of Hydrogen Sulfide by a Biofilter


with Fibrous Peat." Journal of Fermentation Technology 62: 589-594.

Gabriel, D., J. Brown, et al. (2002). "Short Contact Time Biotrickling Filter for Odor
Treatment." Proceedings of the 2002 USC-TRG Conference on Biofiltration, Newport
Beach, CA: 267-276.

Gadre, R. V. (1989). "Removal of Hydrogen Sulfide from Biogas by


Chemoautotrophic Fixed-Film Bioreactor." Biotechnology and Bioengineering 34:
410-414.

Gilmont (1993). Accucal™ Flow-meter Instruction Sheet. Gas Instruments Division.


Barnant Company, Barrington, IL: #A-1299-445: 6 p.

Gutierrez-Rojas, M., S. A. A. Hosn, et al. (1996). "Heat Transfer in Citric Acid


Production by Solid State Fermentation." Process Biochemistry 31(4): 363-369.
115

Hall, S. G. (1998). Temperature Feedback Control via Aeration Regulation in


Biological Composting Systems. Ph. D. Thesis in the Department of Agricultural and
Biological Engineering. Cornell University, Ithaca, NY: 200 p.

Hartikainen, T., et al. (2001). "Carbon Disulfide and Hydrogen Sulfide Removal with
a Peat Biofilter." Journal of the Air and Waste Management Association 51: 387-392.

Henshaw, P. F. and W. Zhu (2001). "Biological Conversion of Hydrogen Sulphide to


Elemental Sulphur in a Fixed-Film Continuous Flow Photo-Reactor." Water Research
35: 3605-3610.

Janssen, A. J. H., H. Dijkman, et al. (2000). "Novel biological processes for the
removal of H2S and SO2 from gas streams." Chapter in Environmental Technologies
to Treat Sulfur Pollution. P. N. L. Lens and L. H. Pol (Eds.). IWA Publishing,
London, UK: 547 p.

Jensen, A. B. and C. Webb (1995). "Treatment of H2S-containing gases: A review of


microbiolocal alternatives." Enzyme Microb. Technol. 17(1): 2-10.

Jensen, J. K. and A. B. Jensen (2000). "Biogas and Natural Gas Fuel Mixture for the
Future." Written for the 1st World Conference and Exhibition on Biomass For Energy
and Industry, Sevilla: 8 p. http://uk.dgc.dk/pdf/Sevilla2000.pdf.

Jewell, W. J., H. R. Capener, et al. (1978). Anaerobic Fermentation of Agricultural


Residue: Potential for Improvement and Implementation. Cornell University, Ithaca,
NY: 427 p. US DOE Report #EY76S0229817.

Jewell, W. J., R. J. Cummings, et al. (1993). Energy and Biomass Recovery from
Wastewater. Cornell University, Ithaca, NY: 370 p. GRI Report #93/0192.

Kayhanian, M. and D. J. Hills (1988). "Membrane Purification of Anaerobic Digester


Gas." Biological Wastes 23: 1-15.

Kellog, M. W. (1996). GRI Scavenger CalcBase™ Software. Gas Research Institute,


Chicago, Illinois. GRI Project #96/0482.

Kim, B. W., K. W. Chang, et al. (1997). "Effect of light source on the microbiological
desulfurization in a Photobioreactor." Bioprocess Engineering 17: 342-348.

Kim, N. J., M. Hirai, et al. (1998). "Comparison of Organic and Inorganic Carriers in
Removal of Hydrogen Sulfide in Biofilters." Environmental Technology 19: 1233-
1241.
116

Kobayashi, H. A. and e. al. (1983). "Use of Photosynthetic Bacteria for Hydrogen


Sulfide Removal from Anaerobic Waste Treatment Effluent." Water Research 17(5):
579-587.

Koe, L. C. C. and F. Yang (2000). "A Bioscrubber for Hydrogen Sulphide Removal."
Water Science and Technology 41(6): 141-145.

Kohl, A. and R. Neilsen (1997). Gas Purification. Golf Publishing Company, Houston,
Texas: 1395 p.

Langenhove, H. V., E. Wuyts, et al. (1986). "Elimination of Hydrogen Sulphide from


Odorous Air by a Wood Bark Biofilter." Water Research 20(12): 1471-1476.

Lanting, J. and A. S. Shah (1992). "Biological Removal of Hydrogen Sulfide from


Biogas." 46th Purdue Industrial Waste Conference Proceedings 46: 709-714.

Lusk, P. (1996). Deploying Anaerobic Digesters: Current Status and Future


Possibilities, Prepared for the National Renewable Energy Laboratory under
Subcontract #CAE-6-13383-01 and sponsored by the Regional Biomass Energy
Program of the US DOE. www.biogasworks.com/Reports/weec96.htm.

Ma, J. (2002). Spatial Analysis of the Potential for Dairy Manure as a Renewable
Energy Resource in New York State. M.S. Thesis in Biological and Environmental
Engineering. Cornell University, Ithaca, NY: 108 p.

Maddox, R. N. and M. D. Burns (1968). "Iron-Oxide Process Design Calculations."


Oil and Gas Journal 54 (August).

Madigan, M. T., J. M. Martinko, et al. (2000). Biology of Microoraganisms (Brock).


Prentice Hall, Upper Saddle River, NJ: 991 p.

McFarland, M. J. and W. J. Jewell (1989). "In-Situ Control of Sulfide Emissions


During the Thermophilic (55C) Anaerobic Digestion Process." Water Research
23(12): 1571-1577.

McGuckin, R. L., M. A. Eiteman, et al. (1999). "Pressure Drop through Raw Food
Wast Compost Containing Synthetic Bulking Agents." Journal of Agricultural
Engineering Research 72: 375-384.

McMullin, M. (2002). Personal Communication with S. Zicari. ADI, International,


Inc., New Brunswick, Canada.

Minott, S. J. (2002). Feasibility of Fuel Cells for Energy Conversion on the Dairy
Farm. M.S. Thesis in Biological and Environmental Engineering. Cornell University,
Ithaca, NY: 130 p.
117

Mysliwiec, M. J., J. S. VanderGheynst, et al. (2001). "Dynamic Volume Averaged


Model of Heat and Mass Transport Within a Compost Biofilter: I. Model
Development." Biotechnology and Bioengineering 73: 282-294.

Newnan, D.G., J.P. Lavelle, et al (2000). Engineering Economic Analysis.


Engineering Press, Austin, TX: 694 p.

Nicolai, R. E. and K. A. Janni (2001). Determining Pressure Drop Through Compost-


Wood Chip Biofilter Media. Written for the ASAE Annual International Meeting,
Sacramento, CA: 7 p. Paper #014080.

Nijssen, J. M., S. J. F. Antuma, et al. (1997). Perspectiveven Mestvergisting op


Nederlandse Melkveebederijven (Dutch), Praktijkonderzoek Rundvee, Schapen en
Paarden (PR). 122: 39.

Nishimura, S. and M. Yoda (1997). "Removal of Hydrogen Sulfide From An


Anaerobic Biogas Using A Bio-Scrubber." Water Science and Technology 36: 349-
356.

Oh, J. K., D. Kim, et al. (1998). "Development of effective hydrogen sulphide


removing equipment using Thiobacillus sp. IW." Environmental Pollution 99: 87-92.

OSHA (2002). Occupational Safety and Health Administration Hazardous Pollutants


List. U.S. Department of Labor, Washington, D.C. www.OSHA.gov.
Ottengraf, S. P. P. (1986). Exhaust Gas Purification. Chapter (12) in Biotechnology 8.
H. J. Rehm and G. Reed (Eds.). VCH Verlagsgesellschaft, Weinheim, Germany: 425-
452.

Pellerin, R. A., L. P. Walker, et al. (1987). "Biogas Cogeneration Operating


Experience with a Dairy Farm Plug Flow Digester." Transactions of the ASAE 3(2):
303-313.

Pellerin, R. A., L. P. Walker, et al. (1988). "Operation and Performance of Biogas-


Fueled Cogeneration Systems." Energy in Agriculture 6: 295-310.

Perry, R. H., D. W. Green, et al., Eds. (1997). Perry's Chemical Engineers Handbook.
McGraw Hill, New York: 2640 p.

Rands, M. B., D. E. Cooper, et al. (1981). "Compost Filters for H2S Removal from
Anaerobic Digestion and Rendering Exhausts." Journal of Water Pollution Control
Federation 53: 185-189.

RDA (2000). BiogasWorks. Web page maintained by Resource Development


Associates, Washington, D.C. www.BiogasWorks.com.
118

Revell, C. (1997). Dear Mr. Iron Sponge Plant Operator. Connelly-GPM, Inc.,
Chicago, IL: 9 p.

Revell, C. (2001). Personal Communication with S. Zicari. Connelly-GPM, Inc.,


Chicago, IL.

Richards, B. K., F. G. Herndon, et al. (1994). "In-Situ Methane Enrichment in


Methanogenic Energy Crop Digesters." Biomass and Bioenergy 6(4): 275-282.

Roos, K. F. and M. A. Moser (Eds.) (2000). The AgSTAR Handbook, Prepared for the
US Environmental Protection Agency, Washington, D.C.:EPA#430B9715.

Ruitenberg, R., H. Dijkman, et al. (1999). "Biologically Removing Sulfur from Dilute
Gas Flows." Journal of Materials May: 45.

Schomaker, A. H. H. M., A. A. M. Boerboom, et al. (2000). Anaerobic Digestion of


Agro-Industrial Wastes: Technical Summary on Gas Treatment. Dutch Energy from
Waste Biomass Program (EWAB), Utrecht, Netherlands: 27 p.

Shurson, J., M. Whitney, et al. (1998). Nutritional Manipulation of Swine Diets to


Reduce Hydrogen Sulfide Emissions. University of Minnesota, Department of Animal
Science and Department of Biosystems and Agricultural Engineering.
http://manure.coafes.umn.edu/odor/sulfur.html.

Smit, C. J. and E. C. Heyman (1999). Present Status of the Sulferox Process.


Presented at the 9th GRI Sulfur Recovery Conference, San Antonio, Texas: 10 p.
www.westfieldengineering.com/images/ShellPaper.pdf

Spiegel, R. J. and J. L. Preston (2000). "Test Results for Fuel Cell Operation on
Anaerobic Digester Gas." Journal of Power Sources 86: 283-288.

Spiegel, R. J., J. C. Trocciola, et al. (1997). "Test Results for Fuel-Cell Operation on
Landfill Gas." Energy 22(8): 777-786.

STM Power (2002). Pure Energy on Demand, Brochure for STM Stirling Engines.
STM Power, Ann Arbor, MI: 6 p.
www.stmpower.com/Markets/STMPureEnergyBrochure.pdf

Sublette, K. L., R. P. Heskth, et al. (1994). "Microbial Oxidation of Hydrogen Sulfide


in a Pilot-Scale Bubble Column." Biotechnology Progress 10: 611-614.

Sublette, K. L. and N. D. Sylvester (1987a). "Oxidation of Hydrogen Sulfide by


Continuous Cultures of Thiobacillus denitrificans." Biotechnology and Bioengineering
29: 753-758.
119

Sublette, K. L. and N. D. Sylvester (1987b). "Oxidation of Hydrogen Sulfide by


Mixed Cultures of Thiobacillus denitrificans and Heterotrophs." Biotechnology and
Bioengineering 29: 759-761.

Sun, Y., C. J. Clanton, et al. (2000). "Sulfur and Nitrogen Balaces in BIofiltes for
Odorous Gas Emission Control." Transactions of the ASAE 43(6): 1861-1875.

Swanson, W. J. and R. C. Loehr (1997). "Biofiltration: Fundamentals, Design and


Operations Principals, and Applications." Journal of Environmental Engineering: 538-
546.

Taphorn, D. (2000). SulfaTreat-410HP Estimated Performance Sheet quoted for S.


Zicari. SulfaTreat, St. Louis, MO.

Taylor, D. K. (1956). "Natural Gas Desulfurization (Parts 1-4)." Oil and Gas Journal
(November 5, 19, and December 3, 10): 4 p.

Tyler, J. P. (2001). Personal Communication with S. Zicari. Cornell University, Ithaca,


NY.

UOP (2000). THIOPAQ Process Technical Data Sheet. Paques Biosystems, The
Netherlands: 3 p. www.uop.com/techsheets/thiopaq.pdf.

Vetter, R. L., D. J. Friederick, et al. (1990). Full Scale Anaerobic Digester and Waste
Management System for a 300 Cow Dairy. Proceedings of the Sixth International
Symposium on Agriculture and Food Processing Wastes, ASAE, Chicago, IL: 236-
249.

Walker, L. P., R. A. Pellerin, et al. (1985). "Anaerobic Digestion on a Dairy Farm:


Overview." Energy in Agriculture 4: 347-363.

Wang, E. (2000). Personal Communication with S. Zicari. Varec Vapor Control, Inc.,
Cypress, CA.

Wani, A. H., A. K. Lau, et al. (1999). "Biofiltration Control of pulping odors –


hydrogen sulfide: performance, macrokinetics and coexistence effects of organo-sulfur
species." Journal of Chemical Technology and Biotechnology 74: 9-16.

Wellinger, A. and A. Linberg (2000). Biogas Upgrading and Utilization - IEA


Bioenergy Task 24. International Energy Association, Paris, France: 20 p.

XENERGY, Inc. (2002). Toward a Renewable Power Supply: The Use of Bio-based
Fuels in Stationary Fuel Cells. Prepared for the Northeast Regional Biomass Program,
Washington, D.C.: 63 p.
120

Yang, R. T. (1987). Gas Separation by Adsorption Processes. Stoneham, MA,


Butterworth Publishers: 352 p.

Yang, Y. (1992). Biofiltration for control of Hydrogen Sulfide. Ph. D. Thesis in


Environmental Engineering. University of Florida, Gainesville, FL: 199 p.

Yang, Y. and E. R. Allen (1994a). "Biofiltration Control of Hydrogen Sulfide. 1.


Design and Operational Parameters." Journal of Air and Waste Management
Association 44: 863-868.

Yang, Y. and E. R. Allen (1994b). "Biofiltration Control of hydrogen Sulfide. 2.


Kinetics, Biofilter Performance and Maintenance." Journal of Air and Waste
Management Association 44: 1315-1321.

Zapffe, F. (1963). "Iron Sponge Removes Mercaptans." Oil and Gas Journal August
19.

Zhang, L., M. Hirai, et al. (1991). "Removal Characteristics of Dimethyl Sulfide,


Methanethiol and Hydrogen Sulfide by Hyphomicrobium sp. I55 Isolated from a Peat
Biofilter." Journal of Fermentation and Bioengineering 72: 392-396.
104

You might also like