Vous êtes sur la page 1sur 112

The Design and Construction of a Pilot-Scale Compost Reactor for the Study

of Gas Emissions from Compost under Different Physical Conditions

Edsel A. Phillip
Department of Bioresource Engineering
McGill University, Montreal

June 2010

A thesis submitted to McGill University in partial fulfillment of the requirements of the


degree of Master’s of Science

© Edsel A. Phillip, 2010

1
Abstract
Composting is generally accepted as an environmentally benign process for
organic waste disposal. However, when not properly managed, composting can result in
the emission of toxic and environmentally hazardous gases, including CH4, NH3, N2O
and CO. Due to the potential negative consequences of composting, there is a need to
gain a better understanding of the physical conditions that affect these volatile emissions
in order to better control them. The objective of this project was to construct a pilot-scale
compost reactor, as a platform to study the potential impact of temperature, O2
concentration, airflow rate, and moisture content on the gaseous emissions from compost.
The pilot-scale reactor was able to control the temperature and O2 concentration inside of
the compost using an automated control algorithm, and continually measure the
concentration of CO, CO2, CH4, NH3, and N2O under time-varying temperature and O2
concentration conditions, using FTIR spectroscopy.

2
Résumé
Composter pour disposer des déchets organiques est généralement perçu comme
un processus bénin et sans impact négatif. Cependant, si mal gérée la dégradation des
déchets par compostage peut émettre des gaz toxiques et dangereux pour l’environnement
tel que les composés : CH4, NH3, N2O et CO. Ces conséquences potentiellement
négatives ont soulevé le besoin de mieux comprendre les conditions physiques qui
affectent l’émission de ces composés volatiles afin de mieux les contrôler. L’objectif de
ce projet consista en la construction d’un composteur intelligent expérimental pour
étudier l’impact potentiel de la température, la concentration en oxygène, l’aération et
l’humidité sur les émissions gazeuses émanant du compost. Le composteur fut capable de
contrôler la température et la concentration en oxygène du compostage et de mesurer en
temps continu la concentration en CO, CO2, CH4, NH3, et N2O sous différentes
conditions de température et concentration d’O2. Le contrôle de la température et de la
concentration en oxygène fut possible grâce à un algorithme de contrôle automatique
intégré au composteur et la mesure des concentrations des composés toxiques par la
méthode de spectroscopie FTIR.

3
Acknowledgements
I want to first thank my fiancée and future wife, Jessica, for her infinite patience,
love and support. Secondly I would like to express my gratitude to my supervisor, Dr.
Grant Clark at McGill University, whose advice and guidance were essential to the
successful completion of this thesis. I would also like to thank my colleagues in the
Ecological Engineering Research Group at McGill for their friendship during my time at
McGill. A special thank you is extended to Scott Manktelow, Michael Schwalb, and
Yvan Gariepy for their help with my shop and laboratory work. Finally, I want to send
my heartfelt appreciation to my family: Mom, Dad, and Denley. Without their constant
encouragement and support I would not have been able to complete this journey.

“Accept finite disappointment, but never lose infinite hope.”


-Dr. Martin Luther King Jr.

4
Authorship and Manuscript
This thesis is written in a manuscript-based format (Chapter 3). The contributions
of authors are as follows: (1) E.A. Phillip - design and construction of pilot-scale
composting system, planning and executing experiments and chemical analysis of
compost, writing of manuscripts and performing statistical analysis on results; (2) O.G.
Clark - supervision of thesis work, reviewing thesis manuscript. The manuscript was
submitted to the Journal of Biological Engineering, American Society of Agricultural
and Biological Engineers (ASABE).

5
Table of Contents
Abstract ............................................................................................................................... 2
Résumé ................................................................................................................................ 3
Acknowledgements ............................................................................................................. 4
Authorship and Manuscript................................................................................................. 5
Table of Contents ................................................................................................................ 6
List of Tables ...................................................................................................................... 8
List of Figures ..................................................................................................................... 8
Abbreviations .................................................................................................................... 10
Chapter 1: General Introduction ....................................................................................... 11
1.1 Waste Generation .................................................................................................... 11
1.1.1 Sources of Waste in Canada ............................................................................ 13
1.2 Waste Management ................................................................................................. 14
1.2.1 Landfilling........................................................................................................ 15
1.2.2 Incineration ...................................................................................................... 17
1.2.3 Recycling ......................................................................................................... 18
1.2.4 Composting ...................................................................................................... 19
1.3 Project Description.................................................................................................. 29
1.4 Project Objectives ................................................................................................... 29
1.5 Scope of Project ...................................................................................................... 30
1.6 Thesis Format.......................................................................................................... 30
Chapter 2: Literature Review ............................................................................................ 32
2.1 Introduction ............................................................................................................. 32
2.2 Reactor Construction .............................................................................................. 33
2.3 Process Control ....................................................................................................... 34
2.3.1 Temperature Control ........................................................................................ 35
2.3.2 Oxygen Concentration Control ........................................................................ 36
2.3.3 Aeration Control .............................................................................................. 37
2.3.4 Moisture Content Control ................................................................................ 37
2.5 Emissions Monitoring ............................................................................................. 38
2.5.1 Dedicated Gas Analyzers ................................................................................. 38
2.5.2 Chemical Traps ................................................................................................ 39
2.5.3 Gas Chromatography ....................................................................................... 39
2.5.4 Multigas Analysis ............................................................................................ 39
Connecting Statement to Chapter 3 .................................................................................. 40
Chapter 3: A Pilot-Scale Compost Reactor for the Study of Gas Emissions from Compost
........................................................................................................................................... 41
3.1 Introduction ............................................................................................................. 41
3.2 Previous Research ................................................................................................... 42
3.2.1 Temperature Control ........................................................................................ 42
3.2.2 Oxygen Concentration Control ........................................................................ 43
3.2.3 Aeration Control .............................................................................................. 43
3.2.4 Moisture Content Control ................................................................................ 44
3.2.5 Emissions Monitoring ...................................................................................... 44
3.3 Methods and Materials ............................................................................................ 45

6
3.3.1 Reactor Construction ....................................................................................... 45
3.3.2 Process Control ................................................................................................ 47
3.3.3 Emissions Monitoring ...................................................................................... 52
3.3.4 Software Control .............................................................................................. 53
3.4 Results and Discussion ........................................................................................... 55
3.4.1 Thermal Resistance of Compost Reactor ......................................................... 55
3.4.2 Properties of Compost Mixture ........................................................................ 55
3.4.3 Temperature Regulation................................................................................... 60
3.4.4 Oxygen Regulation .......................................................................................... 65
3.4.5 Combined Control ............................................................................................ 68
3.4.6 Emissions Monitoring ...................................................................................... 71
3.4.7 General Discussion .......................................................................................... 75
3.4.8 Applications and Significance of Findings ...................................................... 78
3.5 Conclusion .............................................................................................................. 80
Chapter 4: General Conclusion ......................................................................................... 82
References ......................................................................................................................... 84
Connecting Statement to Appendix A .............................................................................. 90
APPENDIX A ................................................................................................................... 91
A.1 Introduction ............................................................................................................ 91
A.2 Brief Review of Literature ..................................................................................... 93
A.2.1 Compost Modelling......................................................................................... 93
A.2.2 Control Modelling ........................................................................................... 94
A.3 Methods and Materials ........................................................................................... 94
A.3.1 Compost Model ............................................................................................... 94
A.3.2 Heater model ................................................................................................... 99
A.3.3 Heat Exchanger Model.................................................................................. 101
A.3.4 Air Supply Model.......................................................................................... 103
A.3.5 Fan Model ..................................................................................................... 104
A.4. Results and Discussion........................................................................................ 104
A.4.1 Open-loop system, no control ....................................................................... 104
A.4.2 Closed-loop system, no control ..................................................................... 105
A.4.3 Temperature Control ..................................................................................... 106
A.4.4 Oxygen Control ............................................................................................. 107
A.4.5 Combined Control ......................................................................................... 110
A.5 Conclusion ........................................................................................................... 111

7
List of Tables
Table 1 - Municipal waste production in the U.S, Japan, and the U.K............................. 13
Table 2 - Potential gas emissions from compost .............................................................. 28
Table 3 - Reactor volume .................................................................................................. 33
Table 4 - Compost reactor insulating material .................................................................. 34
Table 5 - Number of temperature sampling points ........................................................... 36
Table 6 - Thermal resistance of composting reactor ......................................................... 55
Table 7 - Moisture content of dog food ............................................................................ 55
Table 8 - Carbon content of dog food ............................................................................... 55
Table 9 - Nitrogen content of dog food ............................................................................ 56
Table 10 - Moisture content of shredded paper ................................................................ 56
Table 11 - Carbon content of shredded paper ................................................................... 56
Table 12 - Nitrogen content of shredded paper ................................................................ 56
Table 13 - Moisture content of wood chips ...................................................................... 57
Table 14 - Carbon content of wood chips ......................................................................... 57
Table 15 - Nitrogen content of wood chips ...................................................................... 57
Table 16 - Compost material chemical composition ........................................................ 58
Table 17 - Compost recipe ................................................................................................ 60
Table 18 - Results of temperature regulation at 40°C....................................................... 61
Table 19 - Results of temperature regulation at 55°C....................................................... 62
Table 20 - Results of temperature regulation at 20°C....................................................... 63
Table 21 - Results of temperature regulation at 26°C....................................................... 64
Table 22 - Results of O2 concentration regulation at 9%.................................................. 66
Table 23 - Results of O2 concentration regulation at 14%................................................ 67
Table 24 - Performance of combined temperature and O2 control #1 .............................. 69
Table 25 - Performance of combined temperature and O2 control #2 .............................. 71

List of Figures
Figure 1 - Total amount of waste disposed and diverted in Canada ................................. 12
Figure 2 - Residential waste stream in Canada. ................................................................ 14
Figure 3 - Windrow composting systems ......................................................................... 25
Figure 4 - Metallurgie des Appalaches compost turner .................................................... 26
Figure 5 - Vertical flow, packed bed reactor (Haug, 1993) .............................................. 27
Figure 6 - 3D drawing of compost reactor ........................................................................ 46
Figure 7 - 3D drawing of heat exchanger ......................................................................... 49
Figure 8 - 3D drawing of centrifugal fan .......................................................................... 50
Figure 9 - 3D diagram for compost reactor with control apparatus. ................................. 51
Figure 10 - Schematic diagram of composting reactor and control apparatus ................. 52
Figure 11 - Composting trial with no control. .................................................................. 59
Figure 12 - Temperature regulation at 40 C° of empty reactor ........................................ 61
Figure 13 - Temperature regulation at 55°C of full reactor .............................................. 62
Figure 14 - Temperature regulation at 20°C of empty reactor. ........................................ 63
Figure 15 - Temperature regulation at 26°C of full reactor .............................................. 64
Figure 16 - Oxygen concentration regulation at 9% of empty reactor ............................. 66

8
Figure 17 - Oxygen concentration regulation at 14% of full reactor. ............................... 67
Figure 18 - Combined temperature and O2 control #1 ...................................................... 69
Figure 19 - Combined temperature and O2 control #2 ...................................................... 70
Figure 20 - Continuous monitoring of CH4 and CO2 from compost sample .................... 72
Figure 21 - Temperature and O2 concentration of continuous monitoring of compost
reactor ....................................................................................................................... 73
Figure 22 - Continuous monitoring of CH4 and CO2 from compost reactor .................... 74
Figure 23 - On/off control actions, O2 control .................................................................. 76
Figure 24 - Increase in O2 concentration overshoot/undershoot around set point ............ 77
Figure 25 - Compost reactor baffles to prevent preferential airflow ................................ 78
Figure 26 - Block diagram for generic chemical process ................................................ 92
Figure 27 - Closed-loop feedback control system ............................................................ 92
Figure 28 - Schematic of compost bed divided into layers ............................................... 95
Figure 29 - Schematic of pilot-scale composting system ................................................. 96
Figure 30 - Conceptual model of pilot-scale compost reactor .......................................... 96
Figure 31 - Conceptual model of heater ........................................................................... 99
Figure 32 - Schematic diagram of heat exchanger.......................................................... 101
Figure 33 - Conceptual model of air supply ................................................................... 103
Figure 34 - Simulation results, open-loop system .......................................................... 105
Figure 35 - Simulation results, closed-loop system ........................................................ 106
Figure 36 - Temperature control trial: 25°C set point .................................................... 107
Figure 37 - Oxygen control trial: 0.15 kg/m3 set point ................................................... 108
Figure 38 - Oxygen control trial: 0.115 kg/m3 set point, control overshoot .................. 109
Figure 39 - Oxygen control trial: 0.03 kg/m3 set point. .................................................. 110
Figure 40 - Combined O2 and temperature control: 0.03(kg/m3) ................................... 111

9
Abbreviations
BNQ Bureau De Normalisation du Quebec
C Carbon
C:N Carbon to nitrogen ratio
CCME Canadian Council of Ministers of the Environment
CFM Cubic feet per minute
CH4 Methane
CO Carbon monoxide
CO2 Carbon dioxide
COM Component Object Model
EDC Endocrine disrupting compounds
FTIR Fourier Transform infrared
GC Gas chromatography
GHG Greenhouse gas
GPIB General Purpose Interface Bus
H2O Water
H2SO4 Sulfuric acid
K Potassium
KOH Potassium hydroxide
LPM Liters per minute
N Nitrogen
NaOH Sodium hydroxide
NH3 Ammonia
NO3- Nitrate
NOx Nitrogen oxides
O2 Oxygen
P Phosphorous
PBC Polychlorinated biphenyls
PID Proportional-integral-derivative
RMSE Root mean square error
SA Surface area
SO2 Sulfur dioxide
SO42- Sulfate
UNSD United Nations Statistics Division
V Volume
w.b. Wet basis
wt. Weight

10
Chapter 1: General Introduction
Waste is a by-product of day-to-day human activity, and can be simply defined as
any material that is unwanted or considered undesirable by its producer. The Economist
special report on waste, Talking Rubbish, stated “wherever people have been - and some
places where they have not - they have left waste behind” (Anonymous, 2009). Waste can
be classified by its physical state (i.e. solid, liquid, gas), its place of origin (e.g.
residential, commercial, institutional, agricultural) or by its composition (e.g. organic,
hazardous, metal, glass). In 2006, 837 kg of per capita non-hazardous solid waste was
produced in Canada, an increase of 10% over 2002 (Statistics Canada, 2009). This figure
does not include waste that was produced and managed on-site (e.g. backyard burning).
Population growth, urbanization, technological development and increased
economic activity all result in the generation of increasingly large amounts of waste
(Rhyner & Schwartz, 1995). As developed and developing countries continue to expand
their industrial production, increase the proportion of their population that lives in cities,
and the standard of living of their citizens grow, the trend towards increased waste
production may be inevitable. In Canada, there has been a steady increase in the amount
of solid waste produced in the four years between 2002 and 2006 (Fig. 1), and similar
trends can be observed in other industrialized countries (Table 1). Consequently, there is
an increasing need for municipal and regional governments in Canada, and around the
world, to find effective and economical methods to divert and dispose of solid waste.
This section will provide an overview of waste generation in Canada, and its
primary sources. A list of solid waste management techniques that are most commonly
used in Canada will be explored - landfilling, incineration, recycling, and composting -
with an in-depth examination of composting provided, which is the focus of this research.

1.1 Waste Generation


Quantifying the global waste production each year is a difficult task because data
from different countries may be of varying quality, incomplete or non-existent. However,
using the most recent country data available (1998-2007) the United Nations Statistics
Division (UNSD) estimated that the total amount of municipal waste collected in 2007
worldwide was 1.25 billion tonnes (UN, 2009). This did not include industrial and

11
commercial waste that was produced around the world. In 2006, Canada produced
approximately 35 million tonnes of non-hazardous waste. Figure 1 shows the total
amount of waste that was diverted (i.e. recycled and composted) and disposed of (i.e.
landfilled and incinerated), on a per capita basis, from 2002 to 2006 (Statistics Canada,
2009). There was a steady increase in the total amount of waste produced in Canada, and
the percentage of waste that was diverted actually decreased in those four years.
Household sources of waste accounted for approximately 40% of the total waste
generated in Canada; industrial, commercial and institutional waste producers, including
construction, renovation and demolition projects accounted for 60% of the total waste
generated.

1200
Disposed
Diverted
1000

800

863.5
600 789.9
768.1

400

200

211.8 222.7 237.9

0
2002 2004 2006
Year

Figure 1 - Total amount of waste disposed (landfilled and incinerated) and diverted
(recycled and composted) in Canada (per capita)

The province of Alberta was the largest per capita producer of waste in Canada, and
Nova Scotia was the lowest producer of waste on a per capita basis in 2006. The
increasing trend in waste production in Canada is mirrored in other OECD countries;

12
Table 1 shows the increase in total municipal waste production in the United States,
Japan, and the United Kingdom from 1990 to 2005 (OECD, 2008).

Table 1 - Municipal waste production in the U.S, Japan, and the U.K from 1990 to
2005
Country 1990 2005
1,000 tonnes
United States 186,167 222,863
Japan 50,441 51,607
United Kingdom 27,100 35,077

1.1.1 Sources of Waste in Canada


There are three sources of non-hazardous solid waste in Canada (Statistics Canada 2005):
1. Residential: waste from primary and seasonal households. These wastes can
include durable and non-durable goods.
2. Construction, renovation and demolition: waste generated by the
construction industry that includes materials such as concrete, metal, cardboard,
and wiring.
3. Industrial, commercial and institutional: waste generated by commercial
entities such as shopping centers, restaurants and offices; institutional waste
generated by schools, hospitals and government facilities; and industrial waste
generated by resource extraction, processing, and manufacturing (e.g. oil & gas,
mining).
By mass, organic materials originating from kitchens and yards make up the largest
component of household waste. Newspapers and other paper fibers make up the second
highest portion. Figure 2 shows the breakdown of the residential waste stream in Canada
in 2002 (Statistics Canada 2005).

13
Other
Organics
18%

Metal
4% 40%
Glass 3%

9%
Plastics

26%

Paper

Figure 2 - Residential waste stream in Canada.

1.2 Waste Management


Waste management is an important economic sector in industrialized countries. In
2002, there were 1,785 waste management businesses operating in Canada, generating
revenues of $4.1 billion, which represented a 19% increase over 2000 (Statistics Canada
2005). There are various techniques for waste management, each suited for different
waste streams, applications, and geographical locations, and all of them carry
environmental and financial costs, and all require careful management (Anonymous,
2009, p. 3). Exploring the wide range of waste management techniques (anaerobic
digestion, enzymatic hydrolysis, autoclaving, gasification, pyrolysis, etc.) is beyond the
scope of this research. However, the four techniques of solid waste management that are
most common in Canada will be presented in the following four sections: (1) landfilling;
(2) incineration; (3) recycling; and (4) composting. One should note that that these four
approaches are not mutually exclusive and are often used in combination with one
another. For each waste management technique, a brief definition and an overview of the
technology will be provided. The advantages and disadvantages of the different
technologies will also be discussed.

14
1.2.1 Landfilling
Landfilling is the primary method for the disposal of both organic and inorganic
solid waste in Canada. A landfill can be defined as the site where refuse is disposed of by
burial, under layers of earth and the waste material is slowly degraded through anaerobic
decomposition. In 2000, landfills accepted 23 million tonnes of waste in Canada,
accounting for 95% of the total disposed waste (5% was incinerated) (Statistics Canada
2005). Landfill technology has improved over the years, employing methods to capture
hazardous gases that are emitted from landfills, and using liners and drainage pumps to
prevent leachate from contaminating ground water.

1.2.1.1 Advantages
The advantages of landfilling as a means of solid waste disposal include the low
waste processing costs, as the waste destined for landfills is directly disposed of without
any pre-processing or treatment (Bove & Lunghi, 2006). Furthermore, when equipped
with the proper technology, gases emitted from a landfill can be transported by pipeline
to be used for power generation. Methane (CH4) is the gas that is emitted in the highest
concentrations from landfills, and is highly combustible. Methane combustion can be
used to drive electricity generation, and limit the amount of CH4 that is released into the
atmosphere.

1.2.1.2 Disadvantages
Although modern landfills are designed to minimize their environmental impact,
there are three primary negative impacts resulting from landfills: (1) the emission of
environmentally hazardous gases; (2) the contamination of ground water through landfill
leachate; and (3) excessive land use. These three impacts are examined below.

Gas Emissions
The primary gas emissions from landfills, by order of concentration, are CH4,
CO2, N2, and O2 (Bagchi 1994). Methane is a potent green house gas (GHG) that has a
relative warming capacity 25 times that of CO2. In 2002, 3% of Canada’s GHG emissions
came from the CH4 production of landfills (Statistics Canada 2005). Methane is also
highly explosive in concentrations between 5% and 15%, which is a work safety concern

15
at landfill sites. Landfills continue to produce gases for many years after a site is closed
(Rhyner and Schwartz 1995) and consequently the monetary and environmental costs of
maintaining a landfill continue for many years after the landfill has been
decommissioned.

Leachate
Leachate is a liquid that is produced when water or other liquids filter through, or
come in contact with waste material. The toxic components of leachate are: (a) heavy
metals, such as arsenic, lead, mercury, and zinc; (b) organic compounds in concentrations
above recommended exposure limits for humans (Rhyner and Schwartz 1995); and (c)
endocrine disrupting compounds (EDC) (Bertanza and Pedrazzani 2007). EDCs interfere
with the endocrine system in many organisms, including humans, thus effecting growth,
metabolism, and tissue function. It is believed that EDCs can impair the functions of
hormones even at trace levels (Asakura, Matsuto, and Tanaka 2004). When landfill
leachate is allowed to contaminate groundwater sources and the surrounding soil, it can
have a deleterious effect on human health and the health of aquatic animals, as well as
plant growth. The collection and treatment of leachate is possible, using wastewater
management techniques, but the resulting sludge must ultimately be disposed of.

Land Use
The amount of land available for landfill construction is a major consideration for
countries that want to consider landfilling as a means of solid waste disposal. Once the
decision has been made to build a landfill, the selection of a site is important, and must be
chosen carefully so as to limit the environmental impact that the landfill can have on the
surrounding ecosystem. Unsuitable sites for landfill construction include: flood plains,
wetlands, land near airports, geological fault zones, and seismic impact areas (Rhyner
and Schwartz 1995). The recovery of land that has been used as a landfill can take several
decades, and may never be suitable for residential or commercial development because of
the potential dangers posed by the explosive nature of CH4 gas. Problems of establishing
trees and vegetation on and around landfill sites have also been reported (Flower et al.
1978).

16
1.2.2 Incineration
Incineration is the reduction of waste by combustion. After incineration the non-
combustible waste material, also called ash, is disposed of in a landfill. Thus, incineration
must be coupled with landfilling in order to constitute a complete method for waste
disposal (Cross, 1972). In 2000, 21 incinerators disposed of 1.1 million tonnes of solid
waste in Canada, which accounted for 5% of the total amount of disposed waste
(Statistics Canada, 2005).

Advantages
The heat energy released from the burning of waste material can be used to
produce hot water and steam for the production of electricity. In 2005, Denmark
produced 13.5% of total domestic heating from cogeneration of electricity from
incineration (Schwartz 2009). Research investigating the use of recycled ash from
incinerators, to be used as a construction material (cement additive or road base), has
been conducted to try to recycle the waste product of incineration (Park, Park, and Heo
2007; Pan et al. 2008; Huang et al. 2006; Birgisdóttir et al. 2007). Incineration of waste
has encountered increased interest as a method for waste disposal in countries where
space for waste disposal is limited (Asia and Europe), because the incineration of solid
waste can reduce the volume of waste by as much as 90% (Rand, Haukohl, and Marxen
2000). For example, in Japan, 77% of municipal solid waste is incinerated or gasified
(Inoue, Yasuda, and Kawamoto 2009). This is the primary advantage of incineration over
other means of solid waste disposal.

Disadvantages
The exhaust gases from waste incineration facilities can have a serious influence
on the environment, and on human health (Hu & Shy, 2001). The types of emissions
resulting from combustion depend on the waste material, the type of incinerator, and the
completeness of combustion. However, the major components of municipal solid wastes
combustion are CO2, H2O, SO2, nitrogen oxides (NOx), particulates, and toxic chemicals
like polychlorinated biphenyls (PBCs), dioxins, and heavy metals (Rhyner and Schwartz
1995). The results from studies attempting to find any correlations between waste
incineration and various health impacts, such as lung cancer, heart disease, and blood

17
levels of heavy metals, vary greatly. In a review of 22 epidemiological studies conducted
on the health effects of waste incineration, Hu & Shy (2001) found that the proximity to
incinerators influenced the levels of some organic chemicals and heavy metals in the
body, but had no effect on respiratory symptoms or pulmonary function. Air pollution
control systems can be installed on incinerators to reduce or eliminate the toxic and
environmentally hazardous gases that are released into the atmosphere from the
incineration of waste. However, the ash that is produced by incineration must ultimately
be disposed.

1.2.3 Recycling
Recycling is defined as the collection, separation and processing of waste material
for manufacture into raw materials or new products (Statistics Canada 2005). In 2002, 6.6
million tonnes of non-hazardous waste material was recycled in Canada, and mixed paper
was the largest component of the recycled materials. The growth in recycling services at
the local and municipal level was sparked by an increasing concern for the environmental
impacts of landfills in the 1980s and 1990s in Canada (Statistics Canada 2005).
Recycling has been identified as an effective method to deal with inorganic wastes,
including plastics, glass, metal, and paper. A major factor in considering the economic
viability of recycling operations is the market for recyclables, because the recycling of
waste products to make raw materials is an alternative to using primary raw materials for
the manufacturing of products (e.g. aluminum for soda cans, printer paper, etc).

Advantages
The advantages of recycling include: (a) the diversion of waste materials away
from landfills, and the associated negative impacts that result from landfilling; (b)
reduced energy and water resource use when aluminum, steel, paper, and glass are
manufactured using recycled material as opposed to primary raw materials; and (c) the
reduction of mining wastes (Rhyner and Schwartz 1995).

Disadvantages
The negative impacts of recycling vary with the type of material that is processed
(paper, glass, plastic, metal). These impacts can include: (a) sludge produced at paper

18
recycling mills that can be toxic due to de-inking processes that employ harsh chemicals;
(b) waste water from cleaning of plastics (Rhyner and Schwartz 1995); (c) the economic
costs of operating recycling facilities; and (d) the capital costs to repurpose waste
material. The transportation costs associated with the collection of recyclable material is
a drawback of recycling but it is not unique to recycling, and is present in all of the waste
management techniques discussed in the preceding and subsequent sections.

1.2.4 Composting
Composting is the controlled decomposition of organic waste, occurring under aerobic
conditions, that causes the development of thermophilic temperatures (greater than 50°C),
and results in an end product that can be beneficially applied to the land as a soil
amendment and fertilizer (Haug, 1993). The practice of composting can be traced back to
as early as 4000 B.C. when human beings transferred from a hunter-gatherer lifestyle into
a sedentary lifestyle, and began to dispose of organic urban waste in pits outside of their
dwellings. This waste would eventually be applied to agricultural lands (Martin &
Gershuny, 1992; Uhlig, 1976). Agricultural, human and animal residues were also used
as fertilizers by early civilizations in India, China, Japan, and South America (Diaz et al.,
2007). One of the first documented management practices for organic wastes was made
by Sir Albert Howard (Howard, 1935), who developed the composting procedure known
as the “Indore process”, which initially involved only animal manure, but evolved into
the decomposition of other biodegradable material (Diaz et al., 2007). Haug (1993, p. 22)
states that “the Indore method represented the first organized plan for composting in the
modern era”. The Indore process developed by Howard is described below (Howard,
1935):
Step 1. Place a layer of brush on the ground to provide a base for the heap
Step 2. Build the pile in layers, first using a 6-in. layer of “green matter”
like crop wastes or leaves. Next add a 2-in layer of manure, which in turn
is covered by a light layer of topsoil and limestone.
Step 3. Repeat the layering until the pile reaches a height of about 5 ft. Turn the
pile at about 6 week intervals for about 3 months.

19
The advances in composting practice and research have been substantial since Howard’s
publication of “The Manufacture of Humus by the Indore Process”, and other important
early work in composting, like J.I. Rodale’s “The Complete Book of Composting”
(Rodale, 1960). Current composting research includes diverse topics such as microbial
community identification (VanderGheynst & Lei, 2003), mathematical modeling of the
composting process (Petric & Selimbaöi, 2008), odour management and control (Fraser
& Lau, 2000), and the study of gas emission from compost (He et al., 2000; Hellebrand &
Kalk, 2001). In 2002, 1.2 million tonnes of organic wastes were composted in Canada, at
350 centralized composting facilities (Statistics Canada, 2005). This number did not
include the amount of organic waste that was handled by residential composting or on-
site industrial composting.

1.2.4.1 Overview
Composting became a popular method of dealing with organic waste in the early
1990’s. In The Practical Handbook of Compost Engineering, Haug (1993, p.1) gave the
following definition for composting, “the biological decomposition and stabilization of
organic substrates, under conditions that allow development of thermophilic temperatures
as a result of biologically produced heat, to produce a final product that is stable, free of
pathogens and plant seeds, and can be beneficially applied to the land”. Diaz et al. (2007,
p. 26) provides the following definition for composting, “[the] biodegradation process of
a mixture of substrates carried out by a microbial community composed of various
populations in aerobic conditions and in the solid state”. Golueke (1977, p. 2) defines
composting in relation to its waste management capacity, “composting is a method of
solid waste management whereby the organic component of the solid waste stream is
biologically decomposed under controlled conditions to a state in which it can be
handled, stored, and/or applied to the land without adversely affecting the environment”.
The Bureau de Normalisation du Québec (BNQ), like Haug, make reference to the
characteristic temperature increase in compost in their definition of composting, “A solid
mature product resulting from composting, which is a managed process of bio-oxidation
of a solid heterogeneous organic substrate including a thermophilic phase” (BNQ, 2005).
Although different authors and institutions have offered different definitions of what

20
constitutes composting, there are several important common aspects of the definitions
given in the literature:
1. Decomposition of organic substrates
2. It is a controlled process
3. The process operates under thermophilic conditions (greater than 50°C)
4. Aerobic decomposition
5. Results in a stable end product
The decomposition of organic substrates during composting is carried out by different
types of aerobic microorganisms, including bacteria, fungi, and actinomycetes. The type
and relative abundance of the different microorganisms present in compost depends on
the type of substrate and the stage of the composting process, as different microbial
communities will dominate at different times during the composting process. During
composting, microorganisms use O2 to convert organic matter into compost (the end
product) and produce carbon dioxide (CO2), water, nitrate (NO3-), sulfate (SO42-) and
heat. This relationship is represented by the following expression (Chiumenti et al.,
2005).
organic matter  O2 CO2  H 2O  NO3  SO4 2  heat

The composting process can be separated into four stages, each with its own
characteristic temperature range, duration, and dominant microbial community. The four
stages of composting are described below. One should note that these ranges are not
completely distinct and overlap one another.

Stage 1: Mesophilic Phase (25-40°C)


The mesophilic (moderate temperature) phase is the starting phase of the
composting process. The easily degradable compounds of the biomass (sugars and
proteins) are degraded by primary decomposers (bacteria, fungi and actinomycetes), and
their biological activity induces a temperature rise in the compost (Diaz et al., 2007).

Stage 2: Thermophilic Phase (35-65°C)


The temperature rise caused by the mesophilic organisms in Stage 1 inhibits their
growth, and causes them to become inactive. Thermophilic (high temperature) organisms

21
dominate during Stage 2. The rate of decomposition continues to increase until a
temperature of approximately 65°C is reached. Temperatures exceeding 55°C inhibits
fungal growth, thus bacteria and actinomycetes are the dominant classes of
microorganisms active in the compost above 55°C (Diaz et al., 2007).

Stage 3: Cooling Phases


The easily degradable substrate is exhausted and microbial activity is reduced,
causing a temperature decrease in the compost. Classes of both bacteria and fungi start to
degrade more recalcitrant compounds such as cellulose (Diaz et al., 2007).

Stage 4: Maturation Phase


The quality of the compost begins to decline as non-degradable compounds are
formed and become predominant (Diaz et al., 2007). The Guidelines for Compost
Quality, published by the Canadian Council of Ministers of the Environment (CCME),
specify that compost must meet one of the following three requirements at the time of
sale and distribution to be considered mature and stable (CCME, 2005): (1) The
respiration rate is less than, or equal to, 400 milligrams of O2 per kilogram of volatile
solids per hour; (2) the CO2 evolution rate is less than, or equal to, 4 milligrams of carbon
in the form of CO2 per gram of organic matter per day; or, (3) the temperature rise of the
compost above ambient temperatures is less than 8°C.

1.2.4.2 Factors Affecting the Composting Process


Several process parameters have been identified as important influencing factors
in the composting process. Descriptions of these process parameters are given below.

Substrate Composition
A compost substrate - the organic material that will undergo decomposition - can
be characterized by the availability of nutrients (Golueke, 1977). The nutrients available
to the microorganisms in compost, the concentration of those nutrients and the relative
abundance of the nutrients play a critical role in the performance of composting and the
quality of the end product. The macronutrients required by the microorganisms include
carbon (C), nitrogen (N), phosphorous (P) and potassium (K) (Diaz et al., 2007; Golueke,
1977). One of the most important nutrient ratios in the composting process is the ratio

22
between carbon and nitrogen content (C:N). The organisms involved in composting
require approximately 25 times more carbon than nitrogen, so when selecting a
composting mixture, a C:N ratio from 25-30 (on a mass basis) should be maintained to
ensure rapid composting (Cundiff & Mankin, 2003). Achieving an optimal C:N ratio in a
compost mixture is important because an excessive amount of carbon (high C:N ratio)
will slow the composting process, and excessive nitrogen (low C:N ratio) will elevate the
NH3 and N2O emission levels from the compost, inducing odour problems (Cundiff &
Mankin, 2003). Other desirable nutrient ratios exist (e.g. N:P), however they rarely need
to be considered in the composting of waste material because they are generally available
in adequate quantities, and their concentrations are not limiting (Golueke, 1977).

Temperature
The temperature of the compost is commonly used to identify the evolution and
state of the composting process because it is the easiest to monitor (Block, 1999; Keener
et al., 2000; Ressetti et al., 1999). The temperature rise in compost is the result of thermal
energy released by the microorganisms responsible for organic matter decomposition in
compost. The thermophilic bacteria in compost can cause the temperature in the compost
to rise to 60-70°C. These elevated temperatures, however, will limit bacterial growth and
eventually limit any further temperature rise. In special cases where extreme
thermophiles (hyperthermophilic bacteria) are part of the compost fauna, temperatures
can reach in excess of 80°C, as was demonstrated in the industrial composting developed
by Sanyu Company in Japan (Oshima et al., 2007). Oshima et. al. (2007) hypothesize that
the large size of the compost bed allows for very high temperatures to develop, and thus
favouring hyperthermophilic bacterial communities.
Diaz et. al. (2007) state that one of the three reasons to convert organic matter into
compost is to reduce the presence of agents that are pathogenic to humans, animals and
plants to levels that do not constitute a health risk, and it is the thermophilic phase that is
the most important phase in eliminating pathogenic organisms. The Canadian Council of
Ministers of the Environment specifies that compost should be maintained at 55°C or
greater for three days for an in-reactor and aerated static pile composting systems (see
section 1.2.4.3 for descriptions), and at least 15 days for windrow composting piles to

23
reduce any potential health concerns related to pathogenic organisms found in compost
substrates (Statistics Canada, 2005).

Moisture Content
Water provides the medium for chemical reactions, transport of nutrients, and
movement of microorganisms between particles (Cundiff & Mankin, 2003). The physical
properties of the compost substrate and the bulking agent used in the compost will affect
the optimal moisture content, however, a moisture content of 60% allows for optimal
decomposition in compost (Campbell et al., 1990). An excessive amount of water in
compost will reduce the free air space between the compost particles, inhibiting the
movement of air through the compost matrix and consequently the supply of O2 to
aerobic microorganisms. Biological activity is minimal at a moisture content below 12%,
however, in practice the moisture content should not be allowed to drop below 40-50%
(Golueke, 1977).

Oxygen Concentration
Compost is characterized by the decomposition of organic substrates under
aerobic conditions, thus O2 must be supplied to the compost to maintain those conditions.
According to Cundiff and Mankin (2003), the minimum O2 concentration required for
microbial growth is 5%, a fourth the concentration of O2 in ambient air. The advantages
of aerobic over anaerobic decomposition include: (1) fewer objectionable odors; (2) a
temperature rise that results in the destruction of pathogenic organisms; and (3) faster
decomposition (Golueke, 1977).

Aeration Rate
The purpose of aeration in composting is to maintain adequate O2 concentration
and moderate temperature. Aeration ensures that the air surrounding the substrate
particles is consistently replenished with O2 to reduce the number of pockets of anaerobic
respiration. As the composting process progresses, and the physical structure of the
compost changes (e.g. compaction), the aeration can be adjusted accordingly. Given all of
the parameters that have been discussed above, temperature, O2 concentration and

24
aeration have been identified as the most important factors influencing the composting of
organic wastes (Campbell et al., 1990).

1.2.4.3 Composting Systems


There are two general classifications for the wide variety of composting systems:
(1) Reactor systems, where the composting occurs within a closed reactor. These systems
are also referred to as “in-vessel” systems; and (2) non-reactor systems, where
composting takes place outside of a reactor, also referred to as “windrows”.

Non-reactor Systems
Non-reactor systems can be further divided into two groups, (1) “turned”, where
the compost pile is torn down and reconstructed in order to provide aeration to the
compost; and (2) “static”, where the compost pile is not disturbed and aeration is
provided by forcing air through the compost pile. Figure 3 shows an illustration of a basic
non-reactor composting system with the ventilation system positioned at the base of the
pile. Aeration in a static compost pile can be achieved through updraft aeration, where
ambient air is forced through the compost pile, or by creating negative pressure in the
compost pile through a downdraft aeration scheme.

Figure 3 - Windrow composting systems (Haug, 1993)

When forced aeration is not used the industrial compost tuner shown in Figure 4 is used
to agitate a compost windrow to provide aeration.

25
Figure 4 - Metallurgie des Appalaches compost turner
(Photo credit: Ménart Belgium)

Due to the inexpensive equipment required and limited materials handling, windrow
composting systems are inexpensive compared to reactor systems (Diaz et al., 2007).

In-reactor Systems
In-reactor composting systems are also referred to as “bioreactors” because the
active stage of composting inside of a reactor. The substrate is placed in a reactor for 7 to
15 days and is placed in a windrow for a curing phase. Figure 5 shows a vertical in-vessel
composting system where the compost is fed into the reactor at the top and finished
compost is removed from the bottom of the reactor. A fan at the base of the reactor
provides aeration.

26
Figure 5 - Vertical flow, packed bed reactor (Haug, 1993)

In-reactor systems typically employ forced aeration and a mechanized form of


turning, which increases the cost of construction and operation of in-reactor systems
compared to windrow composting (Diaz et al., 2007). The aim of compost technology
(in-vessel and windrow composting) is to maintain the processes parameters that affect
composting at an optimum level, to produce a high quality end product in the shortest
amount of time.

Advantages
The advantages of composting as a means of solid waste management include the
destruction of plant diseases, weed seeds, insects, and insect eggs. It can reduce the
toxicity of organic matter, and can also function as a soil amendment because it contains
valuable nutrients, including nitrogen and phosphorus (Haug, 1993). Bari et. al (2000)
provide the following advantages of composting: (a) potential recovery of organic waste
material in the form of compost for utilization in agriculture or other applications (e.g. as

27
organic fertilizer or soil conditioner); (b) effective inactivation of pathogenic bacteria
present in the organic waste; and (c) stabilization and volume reduction of the waste
materials prior to the environmentally sound final disposal in landfills.

Disadvantages
Although composting holds many advantages over landfilling as a means of
organic waste disposal, the composting process, when not properly controlled, can result
in the emissions of toxic and environmentally hazardous gases. Volatile emissions from
composting may include CH4, NH3, N2O and CO (Beck-Friis et al., 2001; Brown et al.,
2008; Hellebrand & Kalk, 2001).

Table 2 - Potential gas emissions from compost


Gas Impact
Carbon monoxide (CO) Toxic to humans and in significant quantities can be fatal
Methane (CH4) Greenhouse gas
Nitrous oxide (N2O) Greenhouse gas
Ammonia (NH3) Highly water soluble and toxic to aquatic animals

The composition of exhaust gases from compost depends on the control of important
compost process parameters. For example, limiting the amount of O2 available to the
microorganisms within the compost will allow anaerobic bacteria to dominate and results
in the production of CH4. Other disadvantages and challenges in composting include:
(a) the cost of specialized composting equipment that is used in the preparation of
compost feedstock, the aeration of compost using mechanized turning or forced aeration,
and refining the final compost product (Chiumenti et al., 2005); (b) the inability of
composting to deal with inorganic wastes; (c) the control of odours from compost; (d)
and the effort required to maintain compost quality from contaminated input streams,
because of the requirement that compost prepared for market not contain any identifiable
contaminants like rocks, glass particles, and large pieces of metal or plastic (Chiumenti et
al., 2005).
Due to the potential negative consequences posed by toxic and environmentally
hazardous gases, resulting from composting, there is a need to gain a better understanding
of the physical conditions that affect these emissions in order to better control them.

28
Pilot-scale composting systems are required to perform systematic composting
simulations under different operating conditions because they provide the requisite
degree of control and replication of process parameters in order to accurately study
composting dynamics (Mason & Milke, 2005). Full-scale composting systems, by
contrast, are not well suited to study the effect these process parameters have on the
composting process due to the lack of fine control over these parameters (Hogan et al.,
1989).

1.3 Project Description


This project involved the design and construction of a 200 L pilot-scale
composting reactor in which four composting process parameters were controlled:
temperature, O2 concentration, airflow rate through the composting matrix, and moisture
content of the initial compost mixture. Using MATLAB® data acquisition and hardware
control software tools, the compost reactor was able to automate the control of
temperature and O2 concentration inside of the compost reactor, and measure the volatile
emissions from the compost - CO, CO2, CH4, NH3, N2O - using a Fourier Transform
Infrared gas analyzer.

1.4 Project Objectives


There were two main objectives of this research:
1. Design and construct a 200 L compost reactor that allows for the control of four
composting process parameters:
i. Temperature of the compost matrix;
ii. Oxygen concentration in compost reactor headspace;
iii. Airflow rate: through compost matrix;
iv. Moisture content of compost.
2. Provide the ability to continuously measure the gas concentrations of CO, CO2,
CH4, NH3, and N2O from the composting reactor, using FTIR spectroscopy, under
different controlled operating conditions.
The automated compost reactor could also be used to study the relationship between the
important physical factors that affect different biological processes (e.g. anaerobic
digestion of manure), as well as their impact on the volatile emissions from these

29
processes. The system can also be used in the creation of new mathematical models of
composting and for validation of existing models of the composting process, and compost
control systems.

1.5 Scope of Project


This project was principally concerned with the design, construction, and testing
of a pilot-scale compost reactor that would allow for the study of volatile emissions from
compost (CO, CO2, CH4, NH3, N2O), under different operating conditions. The results of
this project consisted of a series of experiments to validate the design of the reactor, and
verify its ability to control the temperature and O2 concentration of the compost
accurately, and measure gas emissions from compost. Repeated experiments to
systematically study the effect of the different physical process parameters on volatile
emissions from compost were not conducted. The statistical analysis of the results
obtained was limited to evaluating the performance of the automated control system in
maintaining the two controlled process parameters - temperature and O2 concentration - at
a set point. This project should be viewed as an engineering design problem, intended to
achieve certain design goals, and bound by certain constraints (time, money, technical
resources, etc.). This project will serve as a proof of concept for future pilot-scale
compost reactors, and as a platform to perform systematic repeated studies to examine the
impact of compost process parameters on volatile emissions from compost.

1.6 Thesis Format


Chapter 1 is a discussion of waste generation and four different techniques for
waste management: landfilling, incineration, recycling and composting. The advantages
and disadvantages of each technique are discussed. The potential negative consequences
resulting from composting, namely the volatile emissions, forms the basis of the two
primary objectives of this research. Chapter 2 is comprised of a review of the different
lab and pilot-scale composting systems that are described in the literature, including
different techniques for controlling four process parameters: temperature, O2
concentration, aeration, and moisture content. Different methods for gas analysis are also
reviewed. Chapter 3 is comprised of a paper that describes the design, construction, and
testing of the pilot-scale compost reactor. Chapter 4 includes a general conclusion for the

30
research, a list of potential improvements for the pilot-scale compost reactor, and
potential future research directions. Appendix A contains a paper that describes the
development of a mathematical model of the pilot-scale composting system described in
this thesis, which can be used to study advanced control techniques for compost. This
paper does not appear as a separate chapter in the main body of the thesis, because an
adequate literature review of the topic was not conducted to justify its inclusion.

31
Chapter 2: Literature Review
2.1 Introduction
Composting is the aerobic decomposition of organic substrates by microbial
communities under thermophilic temperatures, and it results in nutrient rich and stable
end-product, that can be beneficially applied to the land (Haug, 1993). Composting can
be viewed as a complex three-phase (solid, liquid, gas) system, comprised of different
microbial communities interacting with their physical and chemical environment to
produce highly variable and non-linear dynamics. Due to the complexity inherent in
composting, precise apparatus have been devised to accurately control and measure
composting process parameters, on small and large scales, in order to gain a better
understanding of compost dynamics and various environmental impacts. The need to
develop tools to properly control and measure compost processes that are representative
of full-scale composting operations, has led to the development of lab-scale compost
reactors that have been used in composting research. Pilot-scale (medium size) and
bench-scale (small size) compost reactors provide the requisite control of compost
parameters to study the dynamics of composting systems in a systematic manner. In
contrast, on-site experiments at municipal or industrial composting facilities are
expensive, difficult to control and very labour intensive (Petiot & De Guardia, 2004).
Hogan et al. (1989, p. 1082) clearly identifies the problems of studying composting
dynamics as large scales: “Composting experimentation in the field is costly and difficult
to control, and it essentially excludes certain aspects of quantification. This indicates a
need for a laboratory apparatus in which field-like behavior can be simulated. Since field
practices vary widely, are rarely formalized or consistently applied, and tend to be highly
empirical, the precise nature of the system to be simulated is not necessarily obvious”.
This section is a review of the compost reactors that are described in the literature,
including different reactor shapes and insulating material, in addition to the different
techniques for controlling temperature, O2 concentration, airflow rate, and moisture
content within a compost reactor.

32
2.2 Reactor Construction

Size & shape:


Compost is characterized by the biological heat produced that is the result of
microbial decomposition of organic material that can sustain thermophilic temperatures
(greater than 50°C). To ensure that thermophilic temperatures can be attained and
sustained in the compost, reactor shape, reactor volume and the type of insulating
material must be considered in the design of composting reactors. The increase in
temperature of the compost can be described using a simplified energy balance equation:

In  Production  Out  Accumulation (Equation 1)

In represents the thermal energy brought into the system by the air entering the compost,
Production represents the biochemical heat generated by microorganisms in the compost
and Out represents conductive heat loss through the reactor walls (Mason & Milke,
2005). If the thermal losses through the reactor wall are too great, thermophilic
temperature will not be attained. The amount of biochemical heat produced is
proportional to the volume (V) of substrate, and the heat loss is proportional to the
surface area (SA) of the reactor. Therefore the SA:V ratio is an important design
consideration when building a compost reactor (Magalhaes et al., 1993). Cylindrical
compost reactors were the most common shape described in the literature (Petiot & De
Guardia, 2004). The volume of compost reactors reviewed in the literature range from 0.4
L to 1780 L. Table 3 lists several reactor volumes used by different authors.

Table 3 - Reactor volume


Author Reactor volume (L)
Schwab et al., 1994 1780
VanderGheynst et al., 1997 770
Smårs et. al, 2001 200
Barrington et. al, 2002 105
Stombaugh & Nokes, 1996 81
Magalhaes et al., 1993 0.4

The advantage of small volume reactors is the ease with which process parameters (e.g.
temperature or O2 concentration) can be controlled. However, in bench-scale or pilot-
scale reactors this increase in control can be accompanied by a deviation from the

33
dynamics that occur in full scale composting facilities (Mason & Milke, 2005). Large
volume reactors are able to attain higher temperatures compared to smaller reactors due
to a decrease in the SA:V ratio in larger reactors that reduces conductive heat loss
through the reactor walls. However, in large-scale composting facilities, interactions
between the factors that influence the composting process cannot be easily monitored
(Campbell et al., 1990), and maintaining homogeneous conditions throughout the whole
mass to be composted is difficult in full-scale systems (Deschamps et al., 1979).

Reactor Insulation
Compost is self-insulating, and heat is conserved within the system (Hogan et al.,
1989); however, to ensure that the compost is able to attain thermophilic temperatures
during composting many authors have insulated their composting reactors, which
decreases the value of the Out term in Equation 1. Table 4 lists some of the insulating
materials used by different authors.

Table 4 - Compost reactor insulating material


Author Insulating Material
Smårs et al., 2001 Foam rubber
Sikora et al., 1983 Styrofoam
Hong et al., 1983 Polystyrene
Magalhaes et al., 1993 Fiberglass
Whang & Meenaghan, 1980 Cement

In selecting an insulating material, one must consider the convenience of installing the
insulating material and its thermal resistance.

Compost Placement
Many authors have elevated the compost above the floor of the reactor in order to
provide uniform aeration (Bono et al., 1992; Campbell et al., 1990; Fraser & Lau, 2000;
Smårs et al., 2001), or separated the compost at different levels in the reactor to achieve
the same goal (Deschamps et al., 1979).

2.3 Process Control


The factors that affect the composting process include: temperature, O2
concentration, aeration rate, substrate composition, the physical properties of the

34
substrate (e.g. porosity, particle size, bulk density), moisture content, and time (Cundiff
& Mankin, 2003). Different lab-scale composting reactors are described in the literature,
each employing different control strategies for temperature, O2 concentration, aeration
and moisture content. The strategies for controlling these four parameters in in-reactor
composting systems are reviewed below. The goal of process control for in-reactor
systems is to optimize the environmental conditions for the microorganisms within the
compost to provide maximum degradation rates of substrates (Fraser & Lau, 2000).

2.3.1 Temperature Control


Temperature management in a compost reactor can be divided into two groups:
(1) self-heating - a reactor relying solely on microbial heat production to reach and
maintain process temperatures, and that has no temperature control besides some external
insulation (Campbell et al., 1990). Self-heating systems typically remain at thermophilic
temperatures for only one week (Ashbolt & Line, 1982); (2) fixed temperature - a reactor
in which a desired compost temperature is imposed and maintained by means of external
heating or cooling (Campbell et al., 1990). The advantage of a fixed temperature
composting system is that certain composting processes can be monitored at a desired
temperature and by turning off the temperature control apparatus, the system can be
converted into a self-heating system to study the natural time varying dynamics of
composting.
Different methods for controlling the temperature of the compost matrix include
heating and cooling the air stream (Smårs et al., 2001), submerging the compost reactor
in a temperature controlled water bath (Hong et al., 1983; Sikora et al., 1983; Suler &
Finstein, 1977) or water jacket (Körner et al., 2003; Mote & Griffis, 1979), placing the
reactor in an incubator where the temperature of the surrounding air is controlled (Hogan
et al., 1989) and using a heating tape that is attached to the exterior of an insulated reactor
(Magalhaes et al., 1993). The reactor built by Smars et. al. (2001) was the only design
found in the literature that cooled the compost air stream to decrease the temperature of
the compost substrate. The system described by Viel et al. (1987) used a water jacket heat
exchanger to decrease the temperature of the compost. A benefit of being able to both

35
heat and cool the compost is the ability to study composting phenomenon under constant
temperatures below the temperature naturally attained by the compost.
Monitoring the temperature of the compost is essential when constructing a
system to control the temperature. The lab-scale systems described in the literature have
often used thermocouples to measure temperature (Bari et al., 2000; Fraser & Lau, 2000;
Hogan et al., 1989; Tseng et al., 1995; VanderGheynst et al., 1997), or an unspecified
temperature transducer or probe (Ashbolt & Line, 1982; Smårs et al., 2001). The location
of sampling points vary - inside of the compost, inside of the air stream for forced
aeration systems, inside of the water bath (Suler & Finstein, 1977) or heating jacket - as
does the number of sampling points. Table 5 contains a short list of the different number
of sampling locations that have been employed in lab-scale systems described in the
literature.

Table 5 - Number of temperature sampling points


Author Number of Sampling locations
Smårs et al., 2001 16
VanderGheynst et al., 1997 8
Suler & Finstein, 1977 3
Magalhaes et al., 1993 1

2.3.2 Oxygen Concentration Control


The majority of closed-loop compost reactors reviewed in the literature used
compressed air as a source of O2 (Magalhaes et al., 1993; Sikora et al., 1983; Smårs et al.,
2001) to ensure the aerobic decomposition of organic material. In a closed composting
system, O2 must be added to the compost reactor in order to meet the O2 demands of the
aerobic microorganisms. The O2 concentration in ambient air (21%) is approximately
four times higher than the 5% needed for a reasonable composting rate (Cundiff &
Mankin, 2003). However, Fraser and Lau (2000, p. 276) stated that “minimum desirable
interstitial O2 concentration has been reported as 12 to 14%”.
The lab-scale compost reactors reviewed in the literature measured O2
concentration by sampling the exhaust gas in the headspace of the reactor, above the
compost (Sikora et al., 1983; Smårs et al., 2001; Suler & Finstein, 1977) or inside of the
composting matrix itself. Most of the systems described in the literature used dedicated

36
O2 gas analyzers (Fraser & Lau, 2000; Hogan et al., 1989; Noble et al., 1997; Smårs et
al., 2001; Tseng et al., 1995) to measure O2 concentration.

2.3.3 Aeration Control


Aeration in an open composting pile serves three purposes: Air must be supplied
to fulfill the O2 demand of the microorganism for aerobic decomposition. Secondly, air
must be supplied to remove water from wet compost. This purpose is not a concern for
this research because the initial moisture content of the compost was set, and no water
left the closed system. Third, air must be supplied to remove heat from the composting
pile (Haug, 1993). Again, this is not a concern in this research because heat will be
removed by cooling the recirculation air, and evaporative cooling is not present in a
closed system. The majority of the compost reactors described in the literature use forced
aeration, as opposed to passive aeration, and typically provided a constant airflow rate in
the upwards direction (Mason & Milke, 2005). However, downward aeration and
intermittent aeration has also been used, which did not have a significant impact on
degradation rates or temperature profiles in the compost (Bari et al., 2000). Using forced
aeration eliminates the need to turn the compost frequently in order to ensure aerobic
conditions are maintained in the compost. None of the reactors reviewed in the literature
specified how the aeration rate through the compost matrix was measured. This can be
done by using a hot wire anemometer to measure air velocity in the air ducts of the
compost reactor, and by measuring the pressure drop across the air circulation fan.
However, the size of lab-scale systems makes it difficult to accurately measure the
airflow rate through the compost matrix because of the turbulent airflow that is created by
the frequent bends in the air ducts.

2.3.4 Moisture Content Control


The moisture content of compost can be found using the following equation:

Wet wt.  Dry wt.


%Moisture   100% (Equation 2)
Wet wt.

“Moisture content can be controlled either by setting it initially by feed conditioning,


controlled manually by water addition, or maintained by open-loop control of water

37
addition” (Fraser & Lau, 2000, p.276). In order to prevent drying of the compost, many
compost reactor designs described in the literature include a mechanism to humidify the
compost circulation air (Cronje et al., 2003; Scholwin & Bidlingmaier, 2003;
VanderGheynst & Lei, 2003). Saturating the circulation air in the compost prevents
evaporative cooling in an open system (Campbell et al., 1990). However, some authors
have periodically added water to the compost mixture (Barrington et al., 2002; Suler &
Finstein, 1977) to maintain certain moisture content. It is important to maintain the
moisture content above a certain threshold, because below a certain moisture content
microbial activity will effectively cease. In their model of the composting process,
Stombaugh and Nokes (1996) assumed no microbial growth was possible below a
moisture content of 20%, and growth was not limited above 40% moisture.
The automatic control of moisture content of a compost reactor is not commonly
found in the literature, and this may be due to the difficulty of monitoring moisture
content in-situ (Fraser & Lau, 2000). Extracting samples from the compost reactor and
drying them to determine the amount of water present in the compost is a simple
procedure for periodically determining the moisture content of the compost.

2.5 Emissions Monitoring


The lab-scale compost reactors that are described in the literature use four main
methods for identifying and quantifying the gas emissions from compost. These four
techniques are described below.

2.5.1 Dedicated Gas Analyzers


Individual gas analyzers, which provide continuous monitoring of a specific
sample gas, have been used to quantify the concentration of different gas compounds
(CO2, O2, NH3, N2O) from compost (Hogan et al., 1989; Noble et al., 1997; Smårs et al.,
2001; Tseng et al., 1995). Dedicated gas analyzers provide continuous monitoring of the
sample gas, however a drawback of dedicated gas analyzers is that separate units are
required for each gas compound. This limits the flexibility of the gas monitoring system
if different substrates are used, and different exhausts gases are expected.

38
2.5.2 Chemical Traps
Some authors have used NaOH or KOH and H2SO4 to capture CO2 and NH3,
respectively, from the compost exhaust gases (Bono et al., 1992; Campbell et al., 1990;
Deschamps et al., 1979; Magalhaes et al., 1993; Mote & Griffis, 1979; Sikora et al.,
1983; Suler & Finstein, 1977). The concentrations of each gas compound were
determined through chemical analysis of the resulting solution. The primary drawback of
using a chemical trap to quantify compost exhaust gases is that it does not provide
continuous monitoring capability, only the cumulative emissions over time.

2.5.3 Gas Chromatography


Gas chromatography (GC) has been used to quantify the concentration of exhaust
gases from compost (Ashbolt & Line, 1982; Cappaert et al., 1976). All of the compost
reactors reviewed in the literature relied on periodic sampling of compost gases and off-
line analysis of gas samples, and thus the continuous monitoring of exhaust gases was not
possible.

2.5.4 Multigas Analysis


In order to quantify the relative concentrations of gases in compost exhaust gases,
some authors have used multigas analysis techniques such as photoacoustic gas analysis
(Blanes-Vidal et al., 2008; Hellebrand & Schade, 2008; Osada & Fukumoto, 2001) or
FTIR spectroscopy (Carballo et al., 2008; Fukumoto et al., 2003) to measure all of the
gas compounds simultaneously, and in real-time. FTIR spectroscopy relies on the
characteristic absorption spectrum of a gas sample in the infrared region of the
electromagnetic spectrum to determine the concentration of a sample gas (Gasmet, 2009).
In contrast, photoacoustic gas analysis relies on the detection of the acoustic signal that is
produced from the absorption of pulsed light energy from the sample gas, to quantify the
concentrations of the sample (Lumasense, 2010). A principal drawback of both FTIR and
photoacoustic gas analysis is that they are unable to directly identify gases, and can only
be used quantify the concentration of gases that are expected to be in the sample.

39
Connecting Statement to Chapter 3
Lab-scale compost reactors can be used to study the effect of process parameters
on gas emissions from compost. Chapter 3 describes the design, construction and testing
of a 200 L compost reactor that was built in order to be able to study the impact of
temperature, O2 concentration, aeration rate and compost moisture content on the
emission of CO, CO2, CH4, NH3, N2O from compost. This chapter is drawn from a
manuscript prepared for publication in the Journal of Agricultural Engineering Research,
and was co-authored by E. A. Phillip, M.Sc. candidate at McGill University; and Dr.
O.G. Clark, Supervisor and Assistant Professor at McGill University. The format of the
paper has been changed to be consistent with this thesis.

40
Chapter 3: A Pilot-Scale Compost Reactor for the Study of Gas
Emissions from Compost
3.1 Introduction
Composting can be defined as the “biological decomposition and stabilization of
organic substrates, under conditions that allow development of thermophilic temperatures
as a result of biologically produced heat, to produce a final product that is stable, free of
pathogens and plant seeds, and can be beneficially applied to the land” (Haug, 1993, p.1).
Composting is considered an environmentally benign method for organic waste disposal
(Magalhaes et al., 1993), and does hold important advantages over other means of solid
waste disposal (e.g. land filling). However, the composting process, when not properly
managed, can result in the emissions of toxic and environmentally hazardous gases. The
gas emissions from composting can include CH4, CO, NH3, and N2O (Beck-Friis et al.,
2001; Brown et al., 2008; Hellebrand & Kalk, 2001). Due to the potential negative
consequences of composting, there is a need to gain a better understanding of the
physical conditions that influence these gas emissions in order to better control them.
Pilot-scale composting systems are well suited to perform systematic composting
simulations under different operating conditions because they provide the requisite
degree of control and replication of process parameters in order to accurately study
composting dynamics (Mason & Milke, 2005). Full-scale composting systems by contrast
are very difficult to control (Hogan et al., 1989), and thus are not suited for systematic
composting studies.
The two goals of this research were to: (1) design and construct a 200 L compost
reactor where four process parameters were controlled: temperature, O2 concentration,
airflow rate, and moisture content; and (2) enable the quantification of CO, CO2, CH4,
NH3, and N2O gas from the compost reactor under different operating conditions, using
Fourier Transform Infrared (FTIR) spectrometry. The results from this research allow for
the examination of the impact of different physical process parameters on the emissions
of gases from compost, and it will also provide a platform to study different control
techniques for dynamic biological systems.

41
3.2 Previous Research
The factors that affect the composting process include temperature, O2
concentration, aeration rate, substrate composition, physical properties (e.g. porosity),
and moisture content (Cundiff & Mankin, 2003). Different bench and pilot-scale compost
reactors are described in the literature, each employing different control strategies for
temperature, O2 concentration, aeration rate, and moisture content. These strategies are
reviewed below.

3.2.1 Temperature Control


Temperature regulation in a compost reactor can be divided into two groups:
(1) self-heating - where the thermal energy generated by the microorganisms in the
compost is the only mechanism responsible for the temperature rise inside of the compost
reactor; and (2) fixed temperature composting reactor - where the temperature of the
compost is controlled using an external heat (or cooling) source. Different methods for
controlling the temperature of the compost include heating and cooling the air stream
(Smårs et al., 2001), submerging the compost reactor in a temperature controlled water
bath (Sikora et al., 1983), placing the reactor in an incubator where the temperature of the
surrounding air is controlled (Hogan et al., 1989) and heating tape attached to the exterior
of the composting reactor (Magalhaes et al., 1993). The reactor described by Smars et. al.
(2001) was the only design found in the literature that cooled the compost air stream,
allowing for cooling of the compost. The system described by Viel et al. (1987) used a
water jacket heat exchanger to decrease the temperature of the compost.
Monitoring the temperature of the compost is essential when constructing a
system to control the temperature of compost. The bench and pilot-scale systems
described in the literature have often used thermocouples to measure temperature (Bari et
al., 2000; Fraser & Lau, 2000; Hogan et al., 1989; Tseng et al., 1995; VanderGheynst et
al., 1997), or an unspecified temperature transducer or probe (Ashbolt & Line, 1982;
Smårs et al., 2001). The location of temperature sampling points vary - insides of the
compost, inside of the air stream for forced aeration systems, inside of the water bath
(Suler & Finstein, 1977) or heating jacket - as do the number of measuring points. In
order to ensure that the composting matrix is able to attain thermophilic temperatures

42
(greater than 50°C) during composting, many authors have insulated their composting
reactor. Insulating materials included foam rubber (Smårs et al., 2001), StyrofoamTM
(Sikora et al., 1983), polystyrene (Hong et al., 1983), fiberglass (Magalhaes et al., 1993)
and cement (Whang & Meenaghan, 1980).

3.2.2 Oxygen Concentration Control


The majority of closed-loop composting reactors described in the literature used
compressed air as a source of O2 (Magalhaes et al., 1993; Sikora et al., 1983; Smårs et al.,
2001) to ensure the decomposition of organic material occurred under aerobic conditions.
When an open-loop system was used, where the compost gases were exhausted into the
atmosphere, rather than being re-circulated back into the compost, the O2 present in
ambient air was sufficient to satisfy the O2 demand of the compost, provided that
adequate aeration was provided.
The bench and pilot-scale compost reactors described in the literature measured
O2 concentration by sampling the exhaust gas the headspace of the reactor, above the
compost (Sikora et al., 1983; Smårs et al., 2001; Suler & Finstein, 1977) or inside of the
composting matrix itself. Most of the systems described in the literature used dedicated
O2 gas analyzers (Fraser & Lau, 2000; Hogan et al., 1989; Noble et al., 1997; Smårs et
al., 2001; Tseng et al., 1995) to measure O2 concentration.

3.2.3 Aeration Control


The majority of compost reactors described in the literature use forced aeration, as
opposed to passively aerated systems, and typically provided a constant airflow rate in
the upwards direction (Mason & Milke, 2005). In order to ensure that gas flow was
evenly distributed throughout the compost matrix authors placed the compost on a
perforated floor (Barrington et al., 2002; Magalhaes et al., 1993; Smårs et al., 2001;
VanderGheynst et al., 1997). None of the reactors reviewed in the literature specified
how the aeration rate through the compost matrix was measured.

43
3.2.4 Moisture Content Control
In order to prevent drying of the compost, many compost reactor designs
described in the literature included a mechanism to humidify the compost circulation air
(Cronje et al., 2003; Scholwin & Bidlingmaier, 2003; VanderGheynst & Lei, 2003).
However, some authors have periodically added water to the compost mixture to
maintain a certain moisture content (Barrington et al., 2002). It is important to note that in
open-loop composting systems moisture is lost through the exhaust gases, and therefore
water must be added to the compost the maintain the desired moisture content. In closed-
loop composting systems, the periodic addition of water is not necessary because water
does not leave the system, provided that it is airtight.

3.2.5 Emissions Monitoring


The bench and pilot-scale composting systems that are described in the literature
use four main methods for identifying and/or quantifying the gas emissions from
compost. These four techniques are listed and described below.

3.2.5.1 Dedicated Gas Analyzers


Individual gas analyzers, which provide continuous monitoring of specific sample
gases, have been used to quantify the concentration of gas compounds (CO2, O2, NH3,
N2O) from compost (Hogan et al., 1989; Noble et al., 1997; Smårs et al., 2001; Tseng et
al., 1995). Dedicated gas analyzers provide continuous monitoring the sample gas,
however a drawback of this method is that separate units are required for each gas
compound. If the compost substrate is changed, and different exhaust gases are expected,
additional gas analyzer must be employed.

3.2.5.2 Chemical Traps


Some authors have used NaOH or KOH and H2SO4 to capture CO2 and NH3,
respectively, from the compost exhaust gases (Bono et al., 1992; Campbell et al., 1990;
Deschamps et al., 1979; Magalhaes et al., 1993; Mote & Griffis, 1979; Sikora et al.,
1983; Suler & Finstein, 1977). The concentrations of each gas compound were
determined through chemical analysis of the resulting solution.

44
3.2.5.3 Gas Chromatography
Gas chromatography (GC) has been used to quantify the concentration of exhaust
gases from compost (Ashbolt & Line, 1982; Cappaert et al., 1976). All of the compost
reactors reviewed in the literature relied on periodic sampling of compost gases and off-
line analysis of gas samples.

3.2.5.4 Multigas Analysis


In order to quantify the relative concentration of gases in compost exhaust gases,
some authors have used multigas analysis techniques such as photoacoustic gas analysis
(Blanes-Vidal et al., 2008; Hellebrand & Schade, 2008; Osada & Fukumoto, 2001) or
Fourier Transform Infrared (FTIR) spectroscopy (Carballo et al., 2008; Fukumoto et al.,
2003) to measure exhaust gases simultaneously, and in real time. A principal drawback of
both FTIR and photoacoustic gas analysis is that they are unable to identify gas
compounds, and can only quantify the concentration of gases that are expected to be in
the sample.

3.3 Methods and Materials

3.3.1 Reactor Construction


A 200 L cylindrical plastic waste disposal barrel with a sealable lid was used for
the compost reactor. Two 100 mm (4 in.) diameter plastic duct connectors were installed
at opposite sides at the bottom of the reactor, located 100 mm above the base of the
reactor, to accommodate the solid metal ducts that would re-circulate the heated and
cooled compost gases. Rigid metal ducts, as opposed to flexible ducts, were used to
circulate the compost air because of the ease with which they could be insulated with
bubble wrap. Two 100 mm (4 in.) diameter plastic duct connectors were installed into the
lid of the barrel to accommodate the metal ducts. Three holes were drilled into the side of
the reactor at 0.33 m, 0.55 m and 0.74 m from the bottom of the reactor to accommodate
three thermocouples (Type K). Two layers of insulating bubble wrap sheets were
wrapped around the reactor for thermal insulation. Two 40 mm thick Styrofoam disks,
covered in one layer of bubble wrap, were cut to cover the top and bottom of the reactor.

45
A hose barb fitting was installed at the base of the reactor to accommodate the O2 /N2
polyurethane supply tube. Figure 6 shows a 3D drawing of the compost reactor.

Figure 6 - 3D drawing of compost reactor

A metal grate was installed inside of the reactor, supported by v-shaped metal bars, in
order to suspend the compost substrate above the base of the compost reactor; this
allowed for gases to flow uniformly through the composting matrix from the bottom to
the top of the reactor. Mosquito netting was placed on top of the metal grate to prevent
small particles from falling to the bottom of the reactor. A drain was installed in the floor
of the compost reactor to collect compost leachate, and it was reintroduced at the top of
the reactor to maintain the moisture content.

3.3.1.1 Thermal Resistance of composting reactor


In order to perform an energy balance of the composting system, the thermal
resistance of the reactor is required. Three incandescent light bulbs of different power
ratings (60W, 100W, 200W) were used to heat the insulated reactor. The temperature
inside the reactor was measured using the three thermocouples (Type K), connected to a
data logger (Model 34970A, Agilent Technologies, Santa Clara, CA), at three different
heights (0.33 m, 0.55 m and 0.74 m) from the bottom of the reactor. The power supplied
by the light bulb was calculated using a digital multimeter (Model 117, Fluke Electronics,

46
Mississauga, ON). It was possible to find the thermal resistivity of the reactor, under
steady state conditions, using a modified Fourier equation:

Q AT
H  (Equation 3)
t R
H Heat flow, (J/s)
Q Difference in heat energy, (J)
t Time interval, (s)
A Area, (m2)
T Temperature gradient between inside (Ti) and outside (To) reactor, (K)
T  Ti  To , (K)
R Thermal resistance, (K·s·J-1)

The average between three the thermocouples was used for the value of Ti, and used in
the calculation for thermal resistivity.

3.3.1.2 Properties of composting material


The composting material used in the experiments was prepared by mixing dog
food (Nestlé Purina, Vevey, Switzerland) with wood chips, shredded paper, and water.
Dog food was chosen as a substrate because it serves as a good model for organic food
waste and its composition is consistent from batch to batch (VanderGheynst et al., 1997).
The initial moisture content was selected to be 60 % with a C:N ratio of 26:1 (on a mass
basis).
The nitrogen content was calculated using the Kjeldahl method (Bremner, 1965). In order
to determine the carbon content of the dog food, shredded paper, and wood chips were
dried in a 105°C oven for 24 hrs, and then put into a 550°C furnace for 6 hours. The
remaining ash was weighed and used to find the carbon content of the sample using the
following formula (Anonymous, 1951):

100 %ash
%carbon  (Equation 4)
1.8

3.3.2 Process Control


Four process parameters were controlled in the composting reactor: temperature,
O2 concentration, aeration rate, and moisture content. A description of the apparatus used
to control each process parameter, and the control program written in MATLAB®

47
(Version R2009a, The MathWorks, Natick, MA), is provided below. The aeration rate
and moisture content control were performed manually, and are also described below.

3.3.2.1 Heating Control


Temperature was measured using three thermocouples (Type K) located at 0.33
m, 0.55 m and 0.74 m from the bottom of the compost reactor. The measured temperature
was recorded as described in section 3.3.1.1. A 600-watt heater (Model AHF-06120,
Omega Engineering, Stamford, CT) was secured into the middle of a 0.15 m (6 in.) metal
duct connector and was used to heat the compost gases that would in turn heat the
compost. The heater was controlled using the data logger and a proportional voltage
controller (Model 10PCV2415, Crydom inc., San Diego, CA). The data logger produced
a 0-10V DC control signal that was amplified by the proportional voltage controller to
0-120V AC to drive the heater.

3.3.2.2 Cooling Control


A heat exchanger (Model M10-80, Lytron Inc.,Wobrum, MA) was used to
decrease the temperature of the compost gases that would in turn cool the compost. Cold
water (5°C) from the laboratory sink was used as the cooling fluid for the heat exchanger
and the circulation gases from the compost were sent through the heat exchanger. A
solenoid valve, controlled via the data logger, was used to control the flow of cold water
to the heat exchanger.

48
Figure 7 - 3D drawing of heat exchanger

A 0.11 m3/s (230 CFM) fan (Model W2S130-AA25-44, ebm-papst Inc., Farmington, CT)
was used to blow the compost air over the fins of the heat exchanger to cool the compost
gases.

3.3.2.3 Oxygen Concentration Control


A galvanic cell O2 sensor (Model SO-223, Apogee Instruments, Logan, UT) was
used to measure the O2 concentration in the compost exhaust gas. The sensor was
suspended in the headspace of the compost reactor, and the data logger recorded the O2
concentration measurements. The compressed air supply in the laboratory was connected
to a solenoid valve that controlled the supply of air to the compost reactor. A tank of
compressed N2 was connected to a second solenoid valve, and controlled the supply of N2
to the compost reactor. The air and N2 supply was heated to ambient air temperature
using an in-line heater (Model AHP-3471, Omega Engineering, Stamford, CT). Both
solenoid valves were controlled via the data logger. The compressed air was used to
replenish the O2 inside of the reactor. The N2 was used to actively depress the O2
concentration of the compost in order to be able to study compost dynamics as a
consequence of a forced depression of O2 concentration, in addition to a natural decrease
in O2 concentration, due to microbial respiration. Moreover, the compost reactor was not
perfectly airtight, and consequently the N2 was necessary to attain low levels O2
concentrations that could not have been attained through microbial respiration alone.

49
3.3.2.4 Aeration Control
An inline centrifugal fan (Model FR 150, Fantech, Sarasota, FL) was used to
aerate the compost reactor. A 100 mm (4 in.) diameter metal duct was used to connect the
fan in series with the heater. The flow rate of the fan was manually controlled using a
rotary dial. During the composting experiments, the fan operated at a constant power
setting that resulted in a pressure difference across the fan of 117 Pa (12 mm of H2O).
The pressure drop across the fan was measured using a U-shaped liquid column,
fabricated at McGill University.

Figure 8 - 3D drawing of centrifugal fan

The airflow rate through the compost was measured in the vertical exhaust pipe
connected to the circulation fan and heater above the compost at the beginning of the
compost experiment shown in Figure 21, using a hot-wire anemometer (Model VT100,
Kimo Instruments, Montpon, France). The air velocity was measured at three points in
the pipe, and the following formula was used to estimate the volumetric airflow rate:
n
1
Q  Aduct V (Equation 5)
n i1 i
Q Volumetric flow rate (m3/s)
Aduct Area of duct (m2)
n Number of sampling points
Vi Air velocity (m/s)

50
The volumetric flow rate was found to be 0.13 m3/s. Because of the configuration of the
air circulation ducts, which did not have any long, uninterrupted straight sections, the
calculated airflow rate can only be a rough estimate of the true volumetric flow rate. The
frequent bends in the metal ducts make it difficult to determine the true value for the flow
rate.

3.3.2.5 Moisture Control


The moisture content of compost mixture (dog food, shredded paper, and wood
chips) was set at the beginning of each trial, by adding water to the solid ingredients. Any
leachate that collected at the base of the compost reactor was re-introduced at the top of
the compost, maintaining a constant moisture content throughout each composting
experiment. Figure 9 shows a 3D drawing of the composting system described above.

(a)

(c) (d)
(b)

Figure 9 - 3D diagram of compost reactor with control apparatus. (a) fan; (b)
heater; (c) thermocouples; (d) heat exchanger.

3.3.2.6 Evaluating Control Performance

The performance of the two automated control systems (temperature or O2


concentration) was evaluated by calculating the mean of the values of the measured
variable after the set point was reached after the initiation of the control algorithm. The

51
root mean square error (RMSE) was used to calculate the precision of the control
algorithm in maintaining the measured variable at the set point:

1

 x x 
2
RMSE  (Equation 6)
n 1..n i
n Number of observations
xi ith measured value
x Set point

3.3.3 Emissions Monitoring


A Fourier Transform Infrared (FTIR) gas analyzer (Model CX-400, Gasmet,
Helsinki, Finland) was used to measure the exhaust gas from the headspace of the
compost reactor. An aquarium pump (Model AquaClear 20, Rolf C. Hagen, Inc.,
Mansfield, MA), suspended in the headspace of the composting reactor, delivered the gas
sample to the analyzer at a flow rate of 0.07 L/s (4 LPM). DrieriteTM (W.A Hammond
Drierite CO. LTD., Xenia, OH) was used to remove any water from the gas sample
before it entered the gas analyzer. The gas analyzer was connected to a desktop computer
via an RS-232 cable, and communicated with the MATLAB® control software using
Component Object Model (COM) communication interface (Microsoft, 2007). Figure 10
shows a schematic of the complete pilot-scale compost reactor.

Figure 10 - Schematic diagram of composting reactor and control apparatus

52
The FTIR measures the absorption spectrum of the gas sample to determine the
chemical composition of the gas sample. Using a library of spectra for different sample
gases, the FTIR computes the linear combination of reference spectra that best fits the
spectrum obtained from the gas sample, and determines the constituent gas
concentrations of the sample. Calcmet™ (Version 11.08, Gasmet, Helsinki, Finland), the
FTIR software, automatically converted the absorption spectrum into a volumetric
concentration (ppm or %) for the each constituent gas (Gasmet, 2009).

3.3.4 Software Control


The MATLAB® Instrument Control ToolboxTM (Version 2.8, The MathWorks,
Natick, MA) was used to communicate with the data logger, using the GPIB
communication protocol (IEEE-488), in order to control the solenoid valves for the O2
and N2 supply and heat exchanger cooling fluid, as well as control the voltage level to the
heater. The data logger was also used to record the temperature and the O2 concentration
inside of the reactor. The temperature and O2 concentration process parameter were
controlled using an automated on/off control algorithm, where a control action (e.g. turn
on heater) was initiated when the monitored variable (e.g. temperature) declined below a
certain threshold (i.e. set point). The pseudo code for the heating control algorithm is
given below.

function control_heater
%Measure temperature from three thermocouples
T=measure_temperature
%Calculate error signal using set point and measured temperature
error=set_point-T
%Determine if control action should be activated
if(error>threshold)
%Turn on heater
heater_on
else
%Turn off heater
control_on=0

53
Because the on/off algorithm employed to control the temperature of the compost
required one temperature value (T), the temperature measurements from the three
thermocouples were converted into one temperature value using the following equation:

T  0.3T1  0.4 T2  0.3T3 (Equation 7)


T Temperature used for control algorithm (°C)
T1 Thermocouple 1 temperature measurement (°C)
T2 Thermocouple 2 temperature measurement (°C)
T3 Thermocouple 3 temperature measurement (°C)

The coefficients in Equation 7 were chosen arbitrarily by the author to provide more
weight to the temperature measurement taken at the center of the compost matrix because
it is less affected by any edge affects a the bottom and the top of the reactor, and should
more closely model a large-scale system.
A master controller function was created to coordinate the control of the
temperature and O2 concentration, in addition to measuring the gas emissions from the
compost, and writing the output data to a text file. The pseudo code for the controller
function is given below.

function controller
%Loop until experiment is complete
while running==1{
%Activate heater control
control_heater
%Activate cooler control
control_cooler
%Activate air supply control
control_oxygen
%Activate nitrogen supply control
control_nitrogen
%Measure emissions with FTIR
measure_gas
%Plot temperature and oxygen concentration results to screen
display_results}
end

54
3.4 Results and Discussion

3.4.1 Thermal Resistance of Compost Reactor


The thermal resistance of the compost reactor was calculated, and the results are shown
in Table 6.

Table 6 - Thermal resistance of composting reactor


Heat flow Trial 1 Trial 2 Trial 3 Thermal Resistivity (K·s·J-1)
(J/s) ∆T (K) ∆T (K) ∆T (K)
60 17.2 16.5 16.5 0.29
100 26.4 26.2 26.1 0.26
200 48.9 46.4 46.6 0.24

The results indicate that the mean thermal resistance of the composting vessel was
0.26   0.02 K·s·J-1. Although this value is not used in this paper, it is required to make
any calorimetric calculations for the composting process, or when attempting to model
the pilot-scale compost reactor, as was done in the paper contained in Appendix A.

3.4.2 Properties of Compost Mixture


The moisture content, carbon content, and nitrogen content of each composting
ingredient - dog food, shredded paper, and wood chips - were determined, and the results
are given in Tables 7 through 15.

Table 7 - Moisture content of dog food


Sample no. Sample (g) Dried sample (g) Moisture Content (%)
1 8.2 7.4 6.4
2 9.2 8.4 7.4
3 8.6 7.8 6.8

Table 8 - Carbon content of dog food (See Equation 4)


Sample no. Dried Sample (g) Ash (g) Carbon content (%) [a]
1 7.4 0.6 51.1
2 8.4 0.7 50.9
3 7.8 0.6 51.3
[a]
See equation 4.

55
Table 9 - Nitrogen content of dog food
Sample no. Sample (g) Nitrogen Content (%)
1 0.231 2.55
2 0.204 3.3
3 0.212 2.6

Table 10 - Moisture content of shredded paper


Sample no. Sample (g) Dried sample (g) Moisture Content (%)
1 9.9 9.1 8.1
2 9.7 9 7.2
3 8.9 8.2 7.9

Table 11 - Carbon content of shredded paper


Sample no. Dried Sample (g) Ash (g) Carbon content (%)[a]
1 0.741 0.114 47.0
2 0.835 0.131 46.8
3 0.994 0.154 46.9
[a]
See equation 4.

Table 12 - Nitrogen content of shredded paper


Sample no. Sample (g) Nitrogen Content (%)
1 0.244 0.04
2 0.228 0.04
3 0.243 0.06
4 0.230 0.1
5 0.227 0.21
6 0.260 0.03

The results obtained from the nitrogen content analysis of shredded paper yielded widely
varying results (Table 12). This may have been due to the low nitrogen content in paper
and the inability of the employed protocol to accurately detect low levels of nitrogen.
Consequently, the nitrogen content for paper determined by Forge et. al (2003) of 0.24%
was used in the determination of the compost recipe (See Table 16).

56
Table 13 - Moisture content of wood chips
Sample no. Sample (g) Dried sample (g) Moisture Content (%)
1 80.9 69.7 13.8
2 61.8 52.1 15.7
3 70.0 57.5 17.9

Table 14 - Carbon content of wood chips


Sample no. Dried Sample (g) Ash (g) Carbon content (%)[a]
1 11.592 1.246 49.6
2 11.922 1.270 49.6
3 11.035 1.343 48.8
[a]
See equation 4.

Table 15 - Nitrogen content of wood chips


Sample no. Sample (g) Nitrogen Content (%)
1 0.238 0.23
2 0.256 0.23
3 0.247 0.17

The values for moisture content, carbon content and nitrogen content were used to
determine the proportions of each ingredient to be used in the compost mixture to attain a
moisture content of approximately 60% and a C:N ratio of approximately 26:1 on a dry
basis. Equation 8 was used to determine the C:N ratio of the compost mixture used for all
of the compost experiments presented in this paper.

mass of C in a  C in b
C:N  (Equation 8)
mass of N in a  N in b
C:N Mass ratio of carbon to nitrogen (kgC kgN-1)
C Carbon
N Nitrogen
a Dog food
b Shredded paper

The carbon and nitrogen content of the wood chips were not used in the calculation of the
overall C:N of the compost mixture because it was assumed that the nutrients present in

57
the wood were not readily available to microorganisms during the relatively short (5-7
days) duration of the composting experiments presented in this paper.

Equation 9 was used to determine the moisture content of the compost mixture:

mass of water in a  mass of water in b  mass of water in c


MC  (Equation 9)
total mass of all ingredients
MC Moisture content
a Dog food
b Shredded paper
c Wood chips

Table 16 provides an overview of the physical and chemical properties of the compost
recipe used for all of the composting experiments described in this paper.

Table 16 - Compost material chemical composition


Ingredient Mass Dry matter Carbon Nitrogen C:N
(kg) content (%) content (%) content (%) ratio
M (SD) M (SD) M (SD)
Dog food 18.8 93 (0.5) 51.1 (0.03) 2.8 (0.4) 18
Wood chips 24 84 (2.0) 49.3 (0.5) 0.3 (0.06) 142
Paper 1.4 92.5 (0.2) 46.9 (0.09) 0.08 (0.15)[a] 183[b]
M: Mean
SD: Standard Deviation
[a] Result obtained from experimentation.
[b] Literature value from Forge et. al. (2003) used for determination of C:N of compost recipe. 44% C,
0.24% N.

Composting Trial Without Control


To ensure that the chosen compost recipe (dog food, wood chips, and paper) was
suitable for controlled trials, and that it was able to attain thermophilic temperatures
(greater than 50°C), the chosen mixture was placed in the reactor without any control
employed. The results of this trial are shown in Figure 11.

58
60 25

24
50

Oxygen Concentration (%)


23
40

Temperature 22

30
21

20
20

10
19
Oxygen Concentration

0 18

0 50 100 150 200 250


Time (h)

Figure 11 - Composting trial without the use of control apparatus

The results illustrated in Figure 11 demonstrated that the compost mixture was able to
attain thermophilic temperatures inside of the compost reactor without the aid of external
heating. Compressed air was supplied to the compost continuously to ensure that the O2
concentration inside of the reactor did not limit aerobic microbial activity, which would
occur at a concentration below 5% (Cundiff & Mankin, 2003). These results also
illustrated that the compressed air supply was able to adequately replenish the O2
consumed by the microorganisms in the compost.

System Testing
Four sets of experiments were conducted to test the functionality and performance
of the pilot-scale compost reactor: (1) independent control of temperature (heating and
cooling); (2) independent control of O2 concentration; (3) combined temperature and O2
control experiments using two different set points for each experiment; and (4)
continuous monitoring of gas emissions. The temperature and O2 concentration control
experiments were conducted with both an empty compost reactor and the reactor filled

59
with compost. The compost recipe described in Table 17 was used for all of the
experiments. The results of the four sets of experiments are presented and discussed
below.

Table 17 - Compost recipe


Ingredient Mass (kg w.b.)
Wood chips 30.9
Shredded Paper 1.4
Dog food 18.8
Water 50

3.4.3 Temperature Regulation

3.4.3.1 Heating
In order to test the temperature control algorithm, and its ability to raise the
temperature of the compost reactor and maintain the temperature at a set point, an
experiment was conducted to elevate the temperature of the empty compost reactor to
40°C. The results from this experiment are shown in Figure 12. These results indicated
that the heating apparatus (200 W heater) and the on/off control algorithm were able to
control the temperature above the ambient air temperature. The on/off control algorithm
was implemented to activate a control action when the measured variable fell below a
lower threshold (0.5°C below set point) and activate the opposite control action when the
measured variable exceeded an upper threshold (0.5°C above set point). This on/off
control algorithm created a dead band around the set point, where no control actions
occur, and resulted in the sinusoidal behaviour of the measured variable observed for all
the control experiments presented in this paper.

60
45

40

35

30

25

20

15

10

0
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35
Time (h)

Figure 12 - Temperature regulation at 40 C° of empty reactor

The performance evaluation of the results shown in Figure 12 is given in Table 18.

Table 18 - Results of temperature regulation at 40°C


Set Point (°C) 40
Mean Temperature (°C) 38.6[a]
RMSE (°C) 1.5
[a]
After set point was attained

In order to test the ability to raise the temperature of the compost reactor filled
with compost, and maintain the temperature at a set point, an experiment was conducted
to elevate the temperature of the full compost reactor to 55°C. The results from this
experiment are shown in Figure 13.

61
60

50

40

30

20

10

0
0 5 10 15 20 25
Time (h)

Figure 13 - Temperature regulation at 55°C of full reactor

The performance evaluation of the results shown in Figure 13 is given in Table 19.

Table 19 - Results of temperature regulation at 55°C


Set Point (°C) 55
Mean Temperature (°C) 54.5[a]
RMSE (°C) 0.5
[a]
After set point was attained

The temperature control system was able to successfully raise the temperature of
the compost to the set point, and maintain the temperature of the compost at that set
point. The heat exchanger was not required to depress the temperature of the compost
because the temperature in this experiment never exceeded the upper threshold value
(0.5°C above the set point) that would have activated the heat exchanger. The response
time of the system to a temperature control action was slower in the full reactor compared
to the empty reactor because of the increased thermal capacitance of the full reactor
compared with the empty reactor.

62
3.4.3.2 Cooling
In order to test the temperature control algorithm, and its ability to depress the
temperature of the compost reactor and maintain the temperature at a set point, an
experiment was conducted to decrease the temperature of the empty compost reactor to
20°C. The results from this experiment are shown in Figure 14. These results indicated
that the heat exchanger and the on/off control algorithm were able to control the
temperature below the ambient air temperature.

23

22.5

22

21.5

21

20.5

0
20
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4
Time (h)

Figure 14 - Temperature regulation at 20°C of empty reactor.

The performance evaluation of the results shown in Figure 14 is given in Table 20.

Table 20 - Results of temperature regulation at 20°C


Set Point (°C) 20
Mean Temperature (°C) 20.5[a]
RMSE (°C) 0.5
[a]
After set point was attained

63
In order to test the ability to depress the temperature of the compost reactor filled
with compost, and maintain the temperature at a set point, an experiment was conducted
to depress the temperature of the full compost reactor to 26°C. The results from this
experiment are shown in Figure 15.

30

29.5

29

28.5

28

27.5

27

26.5

26

25.5

25

24.5
0
0 1 2 3 4 5 6 7
Time (h)

Figure 15 - Temperature regulation at 26°C of full reactor

The performance evaluation of the results shown in Figure 15 is given in Table 21.

Table 21 - Results of temperature regulation at 26°C


Set Point (°C) 26
Mean Temperature (°C) 25.6[a]
RMSE (°C) 0.4
[a]
After set point was attained

The temperature control system was able to successfully depress the temperature
of the active compost to the set point, and in conjunction with the heater, maintain the
temperature of the compost at the set point. The compost went through an initial period of
active cooling until the set point temperature was reached. Thereafter, only active heating

64
was applied to maintain compost at the set point. As was the case for heating, the
response time of the system to the temperature control was slower in the full reactor
compared to the empty reactor.
The compost reactors that are described in the literature were able to control the
temperature of compost, however, of the reactors reviewed in the literature, only the
reactor described by Smårs et. al. (2001) cooled the compost air stream to reduce the
temperature of the compost. This method of temperature control ensures a uniform
heating and cooling of the compost, in contrast to using a water bath as was done by
Sikora et. al. (1983). As was mentioned above, the ability to actively depress the
temperature of the compost allows for the examination of different compost processes
(e.g. gas emissions) at a temperature lower than the ‘natural’ state of the compost,
without having to modify another process parameter (i.e. aeration rate in an open system).
This is not possible for self-heating compost reactors, or the reactors that do not have to
ability to depress the temperature of the compost (Magalhaes et al., 1993).

3.4.4 Oxygen Regulation


In order to test the O2 concentration control algorithm, and its ability to regulate
the O2 concentration of the compost reactor, an experiment was conducted to depress the
O2 concentration of the empty compost reactor to 9%. The results from this experiment
are shown in Figure 16. These results indicate that the O2 regulation apparatus
(compressed air and N2) and the on/off control algorithm were able to control the O2
concentration within the compost reactor.

65
20

18

16

14

12

10

0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16
Time (h)

Figure 16 - Oxygen concentration regulation at 9% of empty reactor

The performance evaluation of the results shown in Figure 16 is given in Table 22.

Table 22 - Results of O2 concentration regulation at 9%


Set Point (%) 9
Mean O2 concentration (%) 10[a]
RMSE (°C) 0.48
[a]
After set point was attained

In order to test the ability to control the O2 concentration of the compost reactor
filled with compost, and maintain the O2 concentration at a set point, an experiment was
conducted to depress the O2 concentration of the full compost reactor to 14%. The results
from this experiment are shown in Figure 17.

66
25

(A)

20

15

10

No O2 control O2 control
0
0 10 20 30
26.2 40
26.3 50
26.4 60
26.5 70
26.6 80
26.7
Time (h)

Figure 17 - Oxygen concentration regulation at 14% of full reactor. Note: At point


(A), sampling frequency was changed from 1 600 Hz to 1 10 Hz.

The performance evaluation of the results shown in Figure 17 is given in Table 23.

Table 23 - Results of O2 concentration regulation at 14%


Set Point (%) 14
Mean O2 concentration (%) 13.8[a]
RMSE (°C) 0.5
[a]
After set point was attained

The composting trial shown in Figure 17 started with the O2 control system
deactivated, and the O2 concentration inside of the reactor decreased as a consequence of
the consumption of O2 by the microorganisms in the compost. The O2 control system was
activated when the O2 concentration declined to approximately 12%. Once the O2
concentration control system was activated it was able to successfully increase the O2
concentration inside of the reactor to the set point, and maintain the O2 concentration at
the set point. In contrast to the temperature regulation trials described above, the full

67
reactor responds as quickly to the O2 concentration control actions as the empty reactor.
This is due to the ease at which the air is able to move through the compost and replenish
the O2 that has been depleted by the microorganisms, or displace the O2 with the N2
supply.
None of the compost reactors reviewed in the literature had the capability to
actively depress the O2 concentration inside of the compost reactor. The reactors
described in the literature relied on the consumption of O2 by the microorganisms present
in compost to depress the O2 concentration. The ability to force an O2-depleted state in
compost allows for the examination of dynamics of anaerobic. The use of compressed air
is a simple method to provide O2 to the compost reactor, and is likely the reason that all
of the reactors reviewed in the literature used this method (Magalhaes et al., 1993; Sikora
et al., 1983; Smårs et al., 2001).

3.4.5 Combined Control


In order to test the ability to control temperature and O2 concentration of compost
within the reactor simultaneously, two experiments were conducted: (1) temperature and
O2 concentration control using active cooling and depression of O2 concentration; and (2)
temperature and O2 concentration control using passive cooling and depression of O2
concentration. The first experiment used the heat exchanger to cool the compost and the
compressed N2 to depress the O2 concentration of the compost. The second experiment
relied on the conductive and convective heat losses from the reactor, and the microbial
consumption of O2 to depress the temperature and O2 concentration in the reactor,
respectively. The results of these two experiments are given below.

3.5.5.1 Combined Control #1


Figure 18 illustrates the results obtained from the control of the compost at a
temperature of 40°C and an O2 concentration of 11%.

68
45 16

44 14
Oxygen
43 12

Oxygen Concentration (%)


42 10

41 8
Temperature
40 6

39 4

38 2

0 No O2 control O2 control
37 0
0 1 2 3 4 5 6 7
Time (h)

Figure 18 - Combined temperature and O2 control #1

The performance evaluation of the results shown in Figure 18 is given in Table 24.

Table 24 - Performance of combined temperature and O2 control #1


O2 Set Point (%) 10
Mean O2 concentration (%) 10.5[a]
RMSE (%) 0.6
Temperature set point (°C) 40
Mean temperature (°C) 39.3[a]
RMSE (°C) 0.8
[a]
After set point was attained

The composting trial shown in Figure 18 started with the O2 and temperature
control systems deactivated. The O2 concentration inside of the reactor decreased to
approximately 14%, and the temperature of the compost increased to approximately
43°C, both as a consequence of the biological activity of the microorganisms in the
compost. The O2 control system was activated when the O2 concentration declined to

69
14%, and the temperature reached the temperature set point of 40°C. Once the O2
concentration and temperature control systems were activated, they were able to
successfully maintain the O2 concentration and the temperature inside of the reactor,
simultaneously, at the set points for both process parameters. When the O2 concentration
control system was initially activated, there was a slight decrease in the temperature due
to the injection of compressed air into the reactor (between hour 5 and 6). The
temperature control system was able to overcome the effect of injecting the cold
compressed air into the reactor.

3.4.5.2 Combined Control #2


Figure 19 illustrates the results obtained from the control of the compost at a
temperature of 52°C and an O2 concentration of 9%.

60 20

Oxygen
18
50
16

14

Oxygen Concentration (%)


40
12

30 10

8
Temperature
20
6

4
10
2
Temperature and O2 control
0 0
0 5 10 15 20 25 30 35 40 45 50
Time (h)

Figure 19 - Combined temperature and O2 control #2

The performance evaluation of the results shown in Figure 19 is given in Table 25.

70
Table 25 - Performance of combined temperature and O2 control #2
O2 Set Point (%) 9
Mean O2 concentration (%) 9.0[a]
RMSE (%) 0.1
Temperature set point (°C) 52
Mean temperature (°C) 51.6[a]
RMSE (°C) 0.5
[a]
After set point was attained

The composting trial shown in Figure 19 started with the O2 concentration and
temperature control systems deactivated. The O2 concentration inside of the reactor
decreased to approximately 8%, and the temperature of the compost increased to
approximately 53°C, both as a consequence of the biological activity of the
microorganisms in the compost. The O2 concentration control system and the temperature
control system were activated simultaneously when the O2 concentration declined to 8%,
and the temperature increased to 53°C. Only the heater and the compressed air supply
were used to maintain the temperature and O2 concentration at the set points of 52°C and
9%, respectively. The decline in O2 concentration inside of the reactor was due to
microbial respiration and the decline in temperature was due to conductive and
convective heat losses through the reactor walls and metal ducts. Once the O2
concentration and temperature control systems were activated, they were able to
successfully maintain the O2 concentration and the temperature inside of the reactor,
simultaneously, at the set points for both process parameters.

3.4.6 Emissions Monitoring


Two experiments were conducted to test the emissions monitoring system of the
compost reactor: (1) 35 minute continuous emissions monitoring from a 2 kg sample of
compost; and (2) 7 hour continuous emissions monitoring from the compost reactor.
These two experiments are described below.

3.4.6.1 Bench-scale Experiment


A 2 kg compost sample was taken from the reactor and placed in a zip-lock
plastic bag, in order to test the continuous gas monitoring control software. An aquarium

71
pump was inserted into a plastic bag, and delivered a sample gas stream to the FTIR gas
analyzer. The gas analyzer continuously measured the concentration of CH4 and CO2
inside of the bag. The pump was then removed from the bag and the FTIR continued to
measure the gas concentrations of CH4 and CO2 in ambient air. The results of this test are
shown in Figure 20.

1.2 70
Ambient Compost Ambient
air air
60
1
carbon dioxide
methane 50

CH4 Concentration (ppm)


0.8

40
0.6
30

0.4
20

0.2
10

0 0
0 0.1 0.2 0.3 0.4 0.5 0.6
Time (h)

Figure 20 - Continuous monitoring of CH4 and CO2 from compost sample

This experiment was conducted to verify the proper functioning of the Calcmet™
software, the FTIR gas analyzer, and the MATLAB® monitoring software program that
was written by the author.

3.4.6.2 Pilot-scale Experiment


Continuous monitoring of the gas emissions from the composting reactor was
performed during the combined temperature and O2 concentration control experiment
shown in Figure 19, in order to test the gas analysis functionality of the compost reactor.
An aquarium pump was suspended in the headspace of the compost reactor, and delivered
a continuous gas sample to the FTIR gas analyzer. The temperature and O2 concentration

72
evolution in the compost reactor is shown in Figure 21. When the temperature rose in the
compost, the O2 concentration in the reactor declined, due to increased microbial
respiration. The compost was able to attain a temperature of approximately 53°C and the
O2 concentration declined to approximately 8%. The FTIR gas analyzer was used to
monitor the concentration of CO2, CH4, CO, N2O, and NH4 during the period of peak
microbial activity. These results are shown in Figure 22.

60 20

18
50
Oxygen 16

14
40

O2 Concentration (%)
12

30 10

8
Temperature
20
6
Emission
monitoring 4
10
2

0 0
0 5 10 15 20
Time (h)

Figure 21 - Temperature and O2 concentration of continuous monitoring of compost


reactor

73
60 18

Oxygen
16
50

Gas Concentration (O2(%), CO2(%), CH4


14

40 12

Temperature 10
30
8

20 6

Carbon dioxide 4
10
Methane 2

0 0
10.2 11.2 12.2 13.2 14.2 15.2 16.2
Time (h)

Figure 22 - Continuous monitoring of CH4 and CO2 from compost reactor

Figure 22 demonstrates the ability of the compost reactor to measure the gas
concentrations of several gases simultaneously, while also measuring the temperature and
O2 concentration of the compost reactor. A decrease in the O2 concentration coincides
with an increase in the CO2 and CH4 concentration; the increase in CO2 is likely due to
the increased respiration rate during the period of maximum microbial activity, while the
increase in the concentration of CH4 is likely due to anaerobic micro-environments that
appear when the O2 concentration declines, and anaerobic microorganisms begin to
produce CH4. Repeated controlled studies should be conducted to systematically examine
the influence of composting parameters of emission rates. The concentration of NH3, CO
and N2O gas were measured during the experiment shown in Figure 22, however, the gas
concentrations were not included in the graphical results.
FITR gas analysis allows for the on-line monitoring of emissions of multiple
compounds simultaneously, in contrast to dedicated gas analyzers used by Hogen et. al.
(1989) or GC gas analysis used by Capaert et. al (1976). Moreover, the system described
in this paper integrates the process control software with the gas measurement software,

74
which provides for maximum flexibility with regards to sampling times during periods of
interest during composting trials.
There are two main advantages to using a single software program to both
monitor exhaust gases and perform process parameter control: (1) different temperature
and O2 concentration set points can be dynamically set through a software command in
response to the exhaust gas profile. This type of sophisticated monitoring and control is
not possible when using non-programmable solid-state controllers; and (2) advanced
control algorithms are easily implemented and modifiable in software, which may not be
the case for solid-state controllers. Furthermore, some bioprocesses are not suited for
conventional PID control and advanced computer control must be used (Shimizu et al.,
1993).

3.4.7 General Discussion

3.4.7.1 Control algorithm


The control algorithm used for all of the controlled composting experiments
described in this paper was an on/off algorithm for temperature and O2 control. The
algorithm dictated that when the observed parameter (e.g. O2 concentration) exceeded a
certain threshold a control action would be activated (e.g. turn on N2 supply), and when
the observed parameter fell below the threshold, the opposite control action would be
activated when necessary (e.g. turn on compressed air supply). This resulted in the
sinusoidal behaviour observed for all of the controlled composting trials presented in this
paper. Figure 23 illustrates the relationship between the observed control variable (O2
concentration) and the control action (compressed air supply). The control signal for the
N2 supply is not show in Figure 23.

75
20 4

18
Oxygen concentration
16
3
14
Upper threshold
12

10 2

Lower threshold
6
Control Action On
1
4

0 Off
0

0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16


Time (h)

Figure 23 - On/off control actions, O2 control

An upper and lower threshold was used to identify a set point to eliminate the
possibility that both air supply and N2 supply (or heater and cooler) would be activated at
the same time. The control actions were activated only when the observed variable came
within the specified threshold of the set point, and eliminated a constant turning on and
off of the control apparatus. This created a dead band, delineated by the horizontal dotted
lines in Figure 23.

3.4.7.2 System State And Control


Two O2 set points were used during the O2 concentration regulation trials of the
full reactor described in section 3.4.4. The O2 concentration of the compost was allowed
to decrease to 12% before the control system was activated. The first set point was 14%
and is shown in Figure 17, and the second set point was 8%. Figure 24 illustrates that the
further away the O2 concentration set point is from the initial state of the compost, before

76
the control system was activated, the more pronounced the overshoot and undershoot are
as a result of the control algorithm.

25

Begin Control

20
A

15

Set point 1
`
10

Set Point 2

0
0 10 20 30 40 50 60 70 80 90 100
Time (h)

Figure 24 - Increase in O2 concentration overshoot/undershoot around set point

This phenomena can be tested for different O2 concentration and temperature set points at
different times during the composting process, and may provide some insights into the
different control algorithms that may be required for O2 concentration and temperature
depending on how far the set points are from the current state of the system.

3.4.7.3 Preventing Preferential Airflow


Compost trials extended beyond two weeks, the compost began to compact. As a
result of this compaction, the compost receded from the walls of the reactor, and
preferential airflow occurred along the reactor walls. In this state the heater and heat
exchanger were unable to heat or cool the compost matrix, because the heated or cooled
air did not flow through the compost pore space. For this reason, the compost was mixed
with an auger every five days of composting. A possible modification as suggested by

77
some authors is the use of plastic baffles to counter the tendency of the air to travel along
the reactor wall (Fig. 25) (Hogan et al., 1989).

Figure 25 - Compost reactor baffles to prevent preferential airflow

This would eliminate the need to manually mix the compost and allow for long,
uninterrupted composting trials.

3.4.8 Applications and Significance of Findings


The compost reactor described in this research is not useful on its own, but its
value is derived from the research that can be performed using its capability to control
temperature and O2 concentration, airflow rate, and moisture content, while measuring
gas emissions. A case in point is the pilot-scale compost reactor described by Smårs et al.
(2001) and the different authors who have used their system to examine different aspects
of composting such as NH3 and N2O emissions from compost (Jarvis et al., 2009), the
development of actinobacteria populations during composting (Steger et al., 2007), and
improved composting times for household waste using mesophilic temperature control
(Smars et al., 2002). Below is a description of three different research topics that can, and
have been, explored using the compost reactor described in this paper.

3.4.8.1 Gas emissions from municipal/industrial compost facilities.


In 2006, high levels of CO emissions were observed in the municipal composting
facility of the City of Edmonton, which was a potential health and safety risk to workers.
A research project was initiated by researchers at McGill University, The University of
Alberta, and the City of Edmonton, to identify any correlations between the CO emission
rates and the biological and physicochemical properties of the compost using bench-scale

78
experiments (EWMCE, 2010). Although the bench experiments that were conducted
using 250 ml containers container were able to assess the effect of temperature and O2
concentration on CO emissions, pilot-scale composting experiments using the reactor
described in this paper could better mimic the physical conditions of the full-scale
composting facility (temperature, aeration, moisture content, and oxygen concentration)
thus providing a more accurate model of the full-scale facility. The 250 ml containers
were sealed, and gas samples were drawn from the containers and analyzed using GC.
The bench-scale experiments were not able to model the down-draft aeration system that
is employed at the municipal composting facility in Edmonton, which the pilot-scale
reactor described in this paper is able to model. The municipal composting facility in
Edmonton is not well equipped to conduct repeated composting experiments.

3.4.8.2. Compost control modeling


Composting is a complex process, which involves the interaction of physical,
chemical and biological elements that ultimately give rise to adaptive and non-linear
system behaviour. Consequently, controlling compost, particularly at a large-scale, can
prove challenging. With a large-scale system, the simple on/off control algorithm that was
employed throughout this project may not have been sufficient to control the temperature
and oxygen concentration of the compost bed. Some authors have sought to create
advanced control algorithms for compost control in order to control compost quality
(Scholwin & Bidlingmaier, 2003). As part of the author’s M.Sc. degree requirements, a
project course was completed in which the main goal was to create a virtual platform to
test different control techniques for compost, and the result of this project course are
presented as a paper in Appendix A. The pilot-scale reactor described in this paper was
used as the physical model on which the computational model presented in Appendix A
was based. The reactor will be used to validate advanced control algorithms for
composting, without the need to conduct time intensive physical experiments.

3.4.8.3. Parameterization of mathematical models


Numerous mathematical models of composting have been described in the
literature (Higgins & Walker, 2001; Kaiser, 1996; Petric & Selimbaöi, 2008; Stombaugh
& Nokes, 1996). Modelling serves several purposes that include improving system

79
understanding and predicting system performance (Mason, 2006). The creation of an
accurate model of composting requires physical validation, and the compost reactor
described in this paper can serve as a physical model for validation of computational
models of composting. Additionally, because of its ability to measure gas emissions
under different operating conditions, the reactor can be used to construct mathematical
models that are able to predict the emissions of certain gases (e.g. CH4 and CO) under
different conditions, aiding in the design of operating procedures for industrial/municipal
composting facilities to limit the emission of harmful gases. Currently, the reactor
described in this paper is being used to validate a finite element model of an in-vessel
composting system being developed by Courvoisier & Clark (2009) in the Ecological
Engineering Research Group at McGill University, in the same lab as the author.

3.5 Conclusion
The pilot-scale compost reactor presented in this paper was able to successfully
control the temperature and O2 concentration, simultaneously, using a simple on/off
control algorithm implemented using the MATLAB® programming environment. The
aeration rate and moisture content of the compost were controlled manually using a
variable rate fan and a predefined amount of water at the outset of each compost
experiment, respectively. This system was able to demonstrate the ability to continuously
monitor the volatile emissions from compost, using FTIR spectroscopy, under time-
varying temperature and O2 concentration conditions. The advantage of FTIR gas
analysis over other technologies (e.g. GC, chemical traps) is that it allows for the real-
time monitoring of multiple gas compounds simultaneously, providing added flexibility
to the system to study different biological substrates that may have different emission
profiles. This functionality will form the basis of future work that will be concerned
correlating physical process parameters (temperature, O2 concentration, aeration rate)
with the volatile emissions from compost. The software interface that was developed to
monitor and control the temperature and O2 concentration of the compost provides a
platform to study the effectiveness of more sophisticated control algorithms (e.g. model
based control, adaptive control) for compost. This system can also be used to validate and

80
develop mathematical models of the composting system, and also study the possible
chemical and biological factors that affect emissions from compost.

81
Chapter 4: General Conclusion
There has been a steady increase in the amount of solid waste produced in
Canada, and as a result, governments at all levels are in need of economically viable, and
environmentally sustainable methods of solid waste disposal. Composting is considered
an environmentally benign technique for organic solid waste disposal, however, when not
properly managed, composting can result in the emission of environmentally hazardous
and toxic gases such as CO, CH4, N2O and NH3. Due to the potential negative
consequences of composting, there is a need to better understand the physical parameters
that affect the volatile emissions from compost, in order to better control them. Pilot-
scale composting systems are well suited for these types of systematic composting trials
because of their ability to accurately control composting process parameters during
repeated studies. The goal of this research was to: (1) design and construct a 200 L
compost reactor where four process parameters were controlled: temperature, O2
concentration, airflow rate, and moisture content; and (2) enable the quantification of CO,
CO2, CH4, NH3, N2O gas from the compost reactor under different operating conditions,
using Fourier Transform Infrared (FTIR) spectrometry. The pilot-scale composting
system described in this research was able to successfully control the temperature, O2
concentration using a software algorithm and the aeration rate through the compost, and
continuously measure the volatile emissions from compost under time varying
temperature and O2 concentration conditions.
A simple on/off control algorithm was used for this research to control the
temperature and O2 concentration of the compost. The use of a software system to control
these compost process parameters allows for more sophisticated control algorithms to be
used, without the need to purchase new control hardware or software.
This project will serve as a platform for further work to fully examine the impact of both
physical and biochemical process parameters on volatile emissions from compost.
Although the compost reactor designed in this project was able to meet the two
project objectives, there are several notable improvements that could be made that would
improve the performance of the reactor. These improvements are:

82
1. Use of a stainless-steel reactor, with airtight connection ports and air circulation
ducts. This would ensure that ambient air did not enter the reactor and moisture
did not exit the reactor.
2. Use of a computer-controlled, variable speed circulation fans for automated
control of aeration rate. This would allow for compensation in the fan speed to be
made when compost compaction reduces the bulk airflow rate through the
compost over time.
3. Increase the number of thermocouples to more fully characterize the temperature
profile of the compost.
4. Increase the number of O2 sensors to characterize the O2 concentration inside of
the pores of the compost matrix, in addition to the O2 concentration in the
compost headspace.
5. Set the temperature of the compressed air and N2 supply to the current
temperature of the compost so that O2 control has no effect on temperature
control. This could be accomplished by controlling the voltage to the in-line
heating.
6. Baffles installed around the wall of the reactor to prevent preferential airflow (see
section 3.4.7.3).

Future research involving the pilot-scale compost reactor described in this thesis provides
a valuable platform for a variety of experimental work in composting. Projects using this
system that will be implemented in the short term include: (1) the examination of the
physio-chemical and biological mechanisms responsible for CO emissions from compost;
(2) the validation of advanced control techniques for in-vessel composting systems; and
(3) the physical validation of a finite element model of the composting process.

83
References

Anonymous. 2009. Talking Rubbish, The Economist, Feb 28.


Anonymous. 1951. Second interim report of the interdepartmental committee on
utilization of organic wastes. N. Z. Eng., 6, 1-12.
Asakura, H., Matsuto, T. and Tanaka, N. 2004. Behavior of endocrine-disrupting
chemicals in leachate from MSW landfill sites in Japan. Waste Management,
24(6), 613-622.
Ashbolt, N. J. and Line, M. A. 1982. A Bench-Scale System to Study the Composting of
Organic Wastes. J Environ Qual, 11(3), 405-408.
Bagchi, A. 1994. Design, construction, and monitoring of Landfills. John Wiley & Sons,
Inc., New York.
Bari, Q. H., Koenig, A. and Guihe, T. 2000. Kinetic analysis of forced aeration
composting-I. Reaction rates and temperature. Waste Management and Research,
18(4), 303-312.
Barrington, S., Choiniere, D., Trigui, M. and Knight, W. 2002. Compost airflow
resistance. Biosystems engineering, 81(4), 433-441.
Beck-Friis, B., Smårs, S., Jönsson, H. and Kirchmann, H. 2001. Structures and
Environment: Gaseous Emissions of Carbon Dioxide, Ammonia and Nitrous
Oxide from Organic Household Waste in a Compost Reactor under Different
Temperature Regimes. Journal of Agricultural Engineering Research, 78(4), 423-
430.
Bequette, B. 2002. Process control: modeling, design, and simulation. Prentice Hall
Press Upper Saddle River, NJ, USA.
Bertanza, G. and Pedrazzani, R. 2007. Presence of EDCs (Endocrine Disrupting
Compounds) in landfill leachate and municipal wastewaters. Management of
Pollutant Emission from Landfills and Sludge, 75.
Birgisdóttir, H., Bhander, G., Hauschild, M. Z. and Christensen, T. H. 2007. Life cycle
assessment of disposal of residues from municipal solid waste incineration:
Recycling of bottom ash in road construction or landfilling in Denmark evaluated
in the ROAD-RES model. Waste Management, 27(8), 75-84.
Blanes-Vidal, V., Hansen, M. N., Pedersen, S. and Rom, H. B. 2008. Emissions of
ammonia, methane and nitrous oxide from pig houses and slurry: Effects of
rooting material, animal activity and ventilation flow. Agriculture, Ecosystems
and Environment, 124(3-4), 237-244.
Block, D. 1999. Composting for erosion control in Texas. Biocycle, 40(9), 40-41.
BNQ. 2005. National Standard of Canada: Organic Soil Conditioners - Composts.
Reference No. CAN/BNQ 0413-200/2005. Sainte-Foy, Qc. Bureau de
Normalisation du Quebec.
Bono, J. J., Chalaux, N. and Chabbert, B. 1992. Bench-scale composting of two
agricultural wastes. Bioresource technology, 40(2), 119-124.
Bove, R. and Lunghi, P. 2006. Electric power generation from landfill gas using
traditional and innovative technologies. Energy Conversion and Management,
47(11-12), 1391-1401.

84
Bremner, J. M. 1965. Total nitrogen. Methods of soil analysis. Part 2. Agronomy, 9,
1149-78.
Brown, S., Kruger, C. and Subler, S. 2008. Greenhouse gas balance for composting
operations. Journal of Environmental Quality, 37(4), 1396-1410.
Campbell, C. D., Darbyshire, J. F. and Anderson, J. G. 1990. The composting of tree bark
in small reactors: self-heating experiments. Biological wastes, 31(2), 145-161.
Cappaert, I., Verdonck, O. and DeBoodt, M. 1976. Composting of bark from pulp mills
and the use of bark compost as a substrate for plant breeding. Part II. The Effect
of Physical Parameters on the Composting Rate of Bark. Growth Experiments
with Bark Compost. Compost Sci, 17, 18-20.
Carballo, T., Gil, M. V., Gomez, X., Gonzalez-Andres, F. and Moran, A. 2008.
Characterization of different compost extracts using Fourier-transform infrared
spectroscopy (FTIR) and thermal analysis. Biodegradation, 19(6), 815-830.
CCME. 2005. Guidelines for Compost Quality. Reference No. PN 1340. Winnipeg, MB.
Canadian Council of Ministers of the Environment.
Chiumenti, A., Chiumenti, R., Diaz, L. F., Savage, G. M., Eggerth, L. L. and Goldstein,
N. 2005. Modern composting technologies. JG Press, Emmaus, PA.
Courvoisier, P., Clark, O.G. 2009. Modélisation numérique d’un compost. In Proc. CSBE
Journée d’information scientifique et technique en Génie Agroalimentaire. Saint-
Hyacinthe, Quebec, March 25, 2009.
Cronje, A., Turner, C., Williams, A., Barker, A. and Guy, S. (2003). Composting under
controlled conditions. Environmental Technology, 24(10), 1221-1234.
Cross Jr, F. L. 1972. Handbook of Incineration. Technomic Publishing Company,
Wesrport, CT.
Cundiff, J. S. and Mankin, K. R. 2003. Dynamics of biological systems. ASAE-American
Society of Agricultural Engineers, St. Joseph, MI.
Deschamps, A. M., Henno, P., Pernelle, C., Caignault, L. and Lebeault, J. M. 1979.
Bench-scale reactors for composting research. Biotechnology Letters, 1(6), 239-
244.
Diaz, L. F., De Bertoldi, M., and Bidlingmaier, W. 2007. Compost science and
technology. Elsevier, Amsterdam, Netherlands.
EWMCE. 2010. Emission of carbon monoxide during composting of municipal solid
waste - Report presented to the City of Edmonton. Edmonton, AB. Edmonton
Waste Management Center of Excellence.
Finger, S. M., Hatch, R. T. and Regan, T. M. 1976. Aerobic microbial-growth in
semisolid matrices - heat and mass-transfer limitation. Biotechnology and
Bioengineering, 18(9), 1193-1218.
Flower, F. B., Leone, I. A., Gilman, E. F. and Arthur, J. J. 1978. A study of vegetation
problems associated with refuse landfills. US Environmental Protection Agency
Publication, 600, 2-78.
Forge, T. A., E. Hogue, G. Neilsen, and D. Neilsen. 2003. Effects of organic mulches on
soil microfauna in the root zone of apple: implications for nutrient fluxes and
functional diversity of the soil food web. Applied Soil Ecology 22 (1):39-54.
Fraser, B. S. and Lau, A. K. 2000. The effects of process control strategies on composting
rate and odor emission. Compost Science and Utilization, 8(4), 274-292.

85
Fukumoto, Y., Osada, T., Hanajima, D. and Haga, K. 2003. Patterns and quantities of
NH3, N2O and CH4 emissions during swine manure composting without forced
aeration - effect of compost pile scale. Bioresource technology, 89(2), 109-114.
Gasmet. 2009. CX-Series FTIR Gas Analyzer Instruction and Operations Manual. Ver.
26.1. Helsinki, Finland. Gasmet Technologies, Inc.
Golueke, C. G. 1977. Biological reclamation of solid wastes. Rodale Press, Emmaus, PA.
Haug, R. T. 1993. The Practical Handbook of Compost Engineering. Lewis Publishers,
Boca, Raton, FL.
He, Y., Inamori, Y., Mizuochi, M., Kong, H., Iwami, N. and Sun, T. 2000. Measurements
of N2O and CH4 from the aerated composting of food waste. The Science of the
Total Environment, 254(1), 65-74.
Hellebrand, H. and Kalk, W.D. 2001. Emission of carbon monoxide during composting
of dung and green waste. Nutrient Cycling in Agroecosystems, 60(1), 79-82.
Hellebrand, H. J. and Schade, G. W. 2008. Carbon Monoxide from Composting due to
Thermal Oxidation of Biomass. Journal of Environmental Quality, 37(2), 592-
598.
Higgins, C. W. and Walker, L. P. 2001. Validation of a new model for aerobic organic
solids decomposition: simulations with substrate specific kinetics. Process
Biochemistry, 36(8), 875-884.
Hogan, J. A., Miller, F. C. and Finstein, M. S. 1989. Physical modeling of the composting
ecosystem. Applied and Environmental Microbiology, 55(5), 1082-1092.
Hong, J. H., Matsuda, J. and Ikeuchi, Y. 1983. High rapid composting of dairy cattle
manure with crop and forest residues. Trans. Am. Soc. Agric. Eng., 26(2), 533-
541.
Howard, A. 1935. Manufacture of humus by the Indore Process. Journal of the Royal
Society of Arts, 84, 26-59.
Hu, S. W. and Shy, C. M. 2001. Health effects of waste incineration: a review of
epidemiologic studies. Journal of the Air and Waste Management Association,
51(7), 1100-1109.
Huang, C. M., Yang, W. F., Ma, H. W. and Song, Y. R. 2006. The potential of recycling
and reusing municipal solid waste incinerator ash in Taiwan. Waste Management,
26(9), 979-987.
Inoue, K., Yasuda, K. and Kawamoto, K. 2009. Report: Atmospheric pollutants
discharged from municipal solid waste incineration and gasification-melting
facilities in Japan. Waste Management and Research, 27(6), 617.
Jarvis, A., Sundberg, C., Milenkovski, S., Pell, M., Smars, S., Lindgren, P. E. and Hallin,
S. 2009. Activity and composition of ammonia oxidizing bacterial communities
and emission dynamics of NH3 and N2O in a compost reactor treating organic
household waste. Journal of Applied Microbiology, 106(5), 1502-1511.
Kaiser, J. 1996. Modelling composting as a microbial ecosystem: a simulation approach.
Ecological Modelling, 91(1-3), 25-37.
Keener, H. M., Dick, W. A. and Hoitink, H. A. J. 2000. Composting and beneficial
utilization of composted by-product materials. Soil Science Society of America
Book Series, 315-342.

86
Körner, I., Braukmeier, J., Herrenklage, J., Leikam, K., Ritzkowski, M., Schlegelmilch,
M. and Stegmann, R. 2003. Investigation and optimization of composting
processes-test systems and practical examples. Waste Management, 23(1), 17-26.
Lumasense. 2010. Photoacoustic Detection. Frankfurt, Germany. Lumasense
Technologies. Available at: www.lumasenseinc.com. Accessed 29 May 2010.
Magalhaes, A. M. T., Shea, P. J., Jawson, M. D., Wicklund, E. A. and Nelson, D. W.
1993. Practical Simulation of Composting in the Laboratory. Waste Management
Research, 11(2), 143-154.
Martin, D. L. and Gershuny, G. 1992. The Rodale book of composting. Rodale Books.
Emmaus, PA.
Mason, I. G. 2006. Mathematical modelling of the composting process: A review. Waste
Management, 26(1), 3-21.
Mason, I. G. and Milke, M. W. 2005. Physical modelling of the composting environment:
A review. Part 1: Reactor systems. Waste Management, 25(5), 481-500.
Microsoft. 2007. COM: Component Object Model Technologies. Redmond, WA.
Microsoft Corporation. Available at: www.microsoft.com/com. Accessed 29 May
2010.
Mote, C. R. and Griffis, C. L. 1979. A system for studying the composting process.
Agricultural Wastes, 1(3), 191-203.
Noble, R., Fermor, T. R., Evered, C. E. and Atkey, P. T. 1997. Bench-scale preparation
of mushroom substrates in controlled environments. Compost science and
utilization, 5, 32-43.
OECD. 2008. OECD Environment Data, Compendium 2006-2008. Paris, France.
Organization for Economic Co-operation and Development.
Osada, T. and Fukumoto, Y. 2001. Development of a new dynamic chamber system for
measuring harmful gas emissions from composting livestock waste. Water science
and technology: a journal of the International Association on Water Pollution
Research, 44(9), 79.
Oshima, T., Moriya, T., Kanazawa, S. and Yamashita, M. 2007. Proposal of
Hyperthermophilic Aerobic Composting Bacteria and Their Enzymes in Space
Agriculture. 21(4), 121-123.
Park, J. S., Park, Y. J. and Heo, J. 2007. Solidification and recycling of incinerator
bottom ash through the addition of colloidal silica (SiO2) solution. Waste
Management, 27(9), 1207-1212.
Petiot, C. and De Guardia, A. 2004. Composting in a Laboratory Reactor: A Review.
Compost Science and Utilization, 12(1), 69-79.
Petric, I. and Selimbaöi, V. 2008. Development and validation of mathematical model for
aerobic composting process. Chemical Engineering Journal, 139(2), 304-317.
Phillip, E.A. and Clark, O.G. 2009. A pilot-scale compost reactor for the study of gaseous
emissions from compost. Paper presented at ASABE Annual International
Meeting. Reno, Nevada, June 21-24, 2009.
Ramaswamy, S., Cutright, T. J. and Qammar, H. K. (2005). Control of a continuous
bioreactor using model predictive control. Process Biochemistry, 40(8), 2763-
2770.
Rand, T., Haukohl, J. and Marxen, U. 2000. Municipal solid waste incineration:
requirements for a successful project. The World Bank. Washington, DC.

87
Ressetti, R. R., Soccol, V. T. and Kaskantzis Neto, G. 1999. Aplicação da
vermicompostagem no controle patogênico do composto de lodo de esgoto.
Sanare. Revista Técnica da SANEPAR, Curitiba, 12(12), 61-70.
Rhyner, C. R. and Schwartz, L. J. 1995. Waste management and resource recovery. CRC
Press.
Rodale, J. I., Rodale, R., Olds, J., Goldman, M. C., Franz, M. and Minnich, J. (1960). The
complete book of composting. Rodale Books. Emmaus, PA.
Scholwin, F. and Bidlingmaier, W. 2003. Fuzzifying the composting process: A new
model based control strategy as a device for achieving a high grade and consistent
product quality. In Proc. Fourth International Conference of ORBIT. Perth,
Australia, April 20-May 2, 2003, 739-751.
Schwab, B. S., Ritchie, C. J., Kain, D. J., Dobrin, G. C., King, L. W. and Palmisano, A.
C. 1994. Characterization of compost from a pilot plant-scale composter utilizing
simulated solid waste. Waste Management and Research, 12(4), 289.
Schwartz, H. 2009. Landfill vs Incineration: Is Canada Ready to Change. Waste
Management, 29-31.
Shimizu, K., Miura, K., Shioya, S. and Suga, K. 1993. An overview on the control system
design of bioreactors. Advances in Biochemical Engineering Biotechnology, 50,
65-65.
Sikora, L. J., Ramirez, M. A. and Troeschel, T. A. 1983. Laboratory Composter for
Simulation Studies. J Environ Qual, 12(2), 219-224.
Smårs, S., Beck-Friis, B., Jönsson, H. and Kirchmann, H. 2001. Structures and
Environment: An Advanced Experimental Composting Reactor for Systematic
Simulation Studies. Journal of Agricultural Engineering Research, 78(4), 415-
422.
Smårs, S., Gustafsson, L., Beck-Friis, B. and Jönsson, H. 2002. Improvement of the
composting time for household waste during an initial low pH phase by
mesophilic temperature control. Bioresource technology, 84(3), 237-241.
Statistics Canada. 2005. Solid Waste In Canada. Reference No. 16-201-XIE. Ottawa, ON.
Statistics Canada.
Statistics Canada. 2009. Human Activity and the Environment: Annual Statistics.
Reference No. 16-201-X. Otawa, ON. Statistics Canada.
Steger, K., Jarvis, A., Vasara, T., Romantschuk, M. and Sundh, I. 2007. Effects of
differing temperature management on development of Actinobacteria populations
during composting. Research in Microbiology, 158(7), 617-624.
Stombaugh, D. P. and Nokes, S. E. 1996. Development of a biologically based aerobic
composting simulation model. Transactions of the ASAE (USA).
Suler, D. J. and Finstein, M. S. 1977. Effect of temperature, aeration, and moisture on
CO2 formation in bench-scale, continuously thermophilic composting of solid
waste. Applied and Environmental Microbiology, 33(2), 345.
Suler, D. J. and Finstein, M. S. 1977. Effect of Temperature, Aeration, and Moisture on
CO2 Formation in Bench-Scale, Continuously Thermophilic Composting of Solid
Waste 1. Applied and Environmental Microbiology, 33(2), 345-350.
Tseng, D. Y., Chalmers, J. J., Tuovinen, O. H. and Hoitink, H. A. J. 1995.
Characterization of a bench-scale system for studying the biodegradation of
organic solid wastes. Biotechnology progress, 11(4).

88
Uhlig, H. 1976. Die Sumerer. Bertelsmann.
UN. 2009. Municipal Waste Collection. United Nations. New York, NY. Available at:
http://unstats.un.org. Accessed 22 May 2010.
VanderGheynst, J. S., Gossett, J. M. and Walker, L. P. 1997. High-solids aerobic
decomposition: pilot-scale reactor development and experimentation. Process
Biochemistry, 32(5), 361-375.
VanderGheynst, J. S. and Lei, F. 2003. Microbial community structure dynamics during
aerated and mixed composting. Transactions of the ASAE, 46(2), 577-584.
Viel, M., Sayag, D., Peyre, A. and Andre, L. 1987. Optimization of in-vessel co-
composting through heat recovery. Biol. Wastes, 20(3), 167-185.
Whang, D. S. and Meenaghan, G. F. 1980. Kinetic model of composting process.
Compost science/land utilization, 21(3), 44-46.

89
Connecting Statement to Appendix A

Appendix A contains a paper written by E.A. Phillip for the Special Problems in
Bioresource Engineering course (BREE 608), taken during the Winter 2010 semester in
the department of Bioresource Engineering, at McGill University. The paper describes
the development and testing of a computational model of the pilot-scale compost reactor
and control system that is described in this thesis, and it is currently being prepared for
publication in the journal, Bioresource Technology. The format of the paper has been
changed to be consistent with this thesis. This paper does not appear as a separate chapter
in the main body of the thesis, because an adequate literature review of the topic was not
conducted to justify its inclusion.

90
APPENDIX A

A Computational Model of a Pilot-Scale Compost Reactor and its Control Apparatus


1
E. A. Phillip, 1O.G. Clark
1
Department of Bioresource Engineering, McGill University
21 111 Lakeshore Rd., Ste. Anne de Bellevue, QC, H9X 3V9, Canada
E-mail: edsel.phillip@mail.mcgill.ca

A.1 Introduction
In a review of the literature on mathematical modelling of composting, Mason
listed some of the uses and benefits of modelling:

“Mathematical modelling has been widely utilised in science and engineering in


order to improve understanding of the behaviour of systems, explore new
theoretical concepts, predict system performance and, in an increasing number of
cases, aid in the solution of practical design problems. In the latter context,
mathematical models offer the potential to reduce, or even replace, the need for
physical experimentation when exploring new material and/or process options”
(Mason, 2006, p. 1).

Modelling complex biological systems, and the systems that are used to control them, can
enable scientists and engineers to investigate the relative performance of different control
methodologies, without the need to conduct laboratory or pilot-scale experiments.

Process Control
The goal of process control is the maintenance of a process (biological, chemical,
mechanical) within a specified boundary, and the reduction of variation within a process.
A conceptual block diagram for a generic process is shown in Figure 26.

91
Figure 26 - Block diagram for generic chemical process (Bequette, 2002)

To control the process represented in Figure 26, a closed-loop control system can be used
to obtain the desired output response; a block diagram of a closed-loop feedback control
system is shown in figure 27.

Figure 27 - Closed-loop feedback control system

The desired system response (set point) is fed into the feedback system and is compared
with the actual output of the system. The controller uses the difference between the set
point and the output, called the error signal, to control the process. The feedback control
system illustrated in Figure 27 is the foundation for many mechanical and electrical
control systems. However, for chemical and biological systems that are highly complex,
non-linear, and contain multiple interactions between manipulated and output variables,
the simple feedback control system depicted in Figure 27 may not suffice to perform
accurate control of process parameters. Consequently, more advanced feedback control
algorithms have been devised to handle such complex systems. These algorithms include
adaptive control, model predictive control or intelligent control that uses artificial
intelligence techniques (e.g. neural networks, expert systems).

92
Project objectives
In order to explore different control strategies for composting, from simple on/off
control to complex adaptive or model predictive control, a model of a composting system
is required that encompasses the compost, the compost vessel, and the actuators that are
used to control to system (i.e. heaters, coolers, fan). The two primary objectives of this
research were to:

1. Develop a computational model of the pilot-scale compost reactor described by


Phillip and Clark (2009), including the aeration fan, heater and heat exchanger for
temperature control, and compressed air supply for O2 concentration control.

2. Combine the computational model of the composting system with pre-existing


control software to demonstrate how simulated composting trials would be
performed, using different control strategies.

The results of this work will allow researchers and industrial composting operators to
evaluate the performance of different process control techniques, which would otherwise
take significant time, effort and money to perform.

A.2 Brief Review of Literature

A.2.1 Compost Modelling


Mathematical modelling of composting first appeared in the literature in 1976
(Finger et al., 1976). The solution to heat and mass balance equations in time and space
provided the physical basis for the majority of the compost process models. The models
that have appeared since 1976 have treated model parameters as lumped over the entire
reactor (Haug, 1993) or distributed into horizontal layers (Stombaugh and Nokes, 1996).
In addition to mathematical models of composting, physical modelling of composting
systems, using bench and pilot-scale composting vessels, have also been described in the
literature (Smårs et al., 2001; VanderGheynst et al., 1997), each employing different
techniques for temperature and O2 concentration control. No studies in the literature have

93
been found that attempt to model a compost reactor along with its associated control
systems.

A.2.2 Control Modelling


Various advanced control techniques have been studied in the field of
biotechnology (Ramaswamy et al., 2005) and fermentation (Shimizu et al., 1993).
However, such studies could not be found in the literature for control of composting
systems.

A.3 Methods and Materials


In order to develop a computational model of the pilot-scale compost reactor
described by Phillip and Clark (2009), the mathematical model of composting, developed
by Stombaugh and Nokes (1996), was implemented in MATLAB®, along with models of
the heater, heat exchanger, compressed air supply, and co-axial fan, that were used to
control the temperature, O2 concentration and aeration rate inside of the composting
vessel described by Phillip and Clark (2009). Each model is described below.

A.3.1 Compost Model

A.3.1.1 Conceptual Model


Open-loop system
The compost model developed by Stombaugh and Nokes (1996) was based on the
growth processes of the microorganisms involved in composting, and assumed that this
biological activity could be described using Monod kinetics. Conceptually, the compost
in the model developed by Stombaugh and Nokes (1996) was divided into n horizontal
layers, as shown in Figure 28. The properties within each layer were considered constant
(i.e each layer was considered to be a control volume). The results of their model
provided a description of how the microbial biomass changed with time, and how this
microbial growth impacted the temperature substrate concentration, moisture content, and
O2 concentration within the compost bed (Stombaugh and Nokes, 1996).

94
Figure 28 - Schematic of compost bed divided into layers

In the compost model described by Stombaugh and Nokes (1996), the properties of the
air entering the compost (layer 1) were representative of ambient air conditions (i.e. O2
concentration: 0.2992 kg/m3, temperature: 21°C etc.). However, when describing a
closed-loop composting system, the air entering the compost would originate from the air
exiting the compost, and thus would inherit the properties of the exiting air. The closed-
loop configuration of the compost system that was described by Phillip and Clark (2009)
is given below.

Close-loop system
A schematic of the closed-loop pilot-scale composting system described by
Phillip and Clark (2009) is given in Figure 29.

95
Figure 29 - Schematic of pilot-scale composting system

The conceptual model of this system is similar to that of the system described by
Stombaugh and Nokes (1996), with the compost bed divided into n layers, however for
the closed-loop composting system, the entering air conditions were taken from the
exiting air conditions. Figure 30 illustrates the conceptual model of the composting
system described by Phillip and Clark (2009), and was implemented for this project. The
‘Control’ block represents the heat, air supply and heat exchanger shown in Figure 29.

Figure 30 - Conceptual model of pilot-scale compost reactor

96
A.3.1.2 Mathematical model
The Stombaugh and Nokes (1996) compost model described how the
concentration of microbial biomass changed with time and the impact of microbial
growth on compost substrate concentration, O2 concentration and temperature within the
compost bed. The five differential equations that governed the compost dynamics are
given below:

1. Microorganism concentration
dCX
 (  k d )CX
dt

2. Substrate concentration
dCS 1 dCX dCX
  CX , 0
dt YX / S dt dt

dCS dCX
 CX , 0
dt dt

3. Oxygen concentration

dCO2 j dCS V
 YO2 / S  (CO2 j 1  CO2 j )
dt dt VL

4. Water concentration

dCmwj dCS V  a
dt
 YW / S
dt

VL
H xj  H ej 

5. Energy Balance (Temperature)



dT T dC S V a Q
qmj
dt
 hc
dt

VL
hxj  hej  lj
VL

The reader can refer to Stombaugh and Nokes (1996) for a detailed explanation of the
model development and a description of the model parameters.

97
A.3.1.3 Computational model
The five differential equations given in section A.3.1.2, along with the appropriate
psychrometric relationships, were implemented in MATLAB® as a set of difference
equations that defined the incremental change of a process variable (dCx) for an
incremental time step (∆t).

Cx (t  t)  Cx (t)  dCx (t)

The simulation model from Stombaugh and Nokes (1996) used a time step of one hour,
and all of the rates specified in the model were per hour (e.g. kg/h). In contrast, the
computational model developed for this project used a time step of one second. The
MATLAB® program used to implement the mathematical model of the composting
systems was structured into two nested loops:

for time  1: end _ of _ simulation


for layer  2 : number _ of _ layers  1
%Calculate change in layer properties : ( dP1..n )
%Update layer properties : P1..n (time  t )  P1..n ( time)  dP1..n
end
% Run control : PC1..n  control _ a lg orithm( P1..n ( time, number _ of _ layers  1))
%Re circulate air : P1..n ( time  t,1)  PC1..n
end

The inner loop traversed each layer of the compost for one time step, and the outer loop
defined the progression of time in the simulation. After the program updated the
parameters of each layer (inner loop), control actions were run on the exiting air from the
top layer of the compost, and the output of the controller models were used as input to the
compost model at layer 1. In addition to the Stombaugh and Nokes (1996) compost
model, individual models for each of the control actuators (heater, heat exchanger, air
supply, and fan) were created to fully represent the in-vessel composting system shown in
Figure 29. The models for each control actuator are described in the following section.

98
A.3.1.4 Model Assumptions:

 For a full list of assumptions for the Stombaugh and Nokes compost model see
Stombaugh and Nokes (1996).
 The length of re-circulation ducts was zero, meaning that there was no lag time
between the air exiting the compost and the air entering the compost.
Additionally, the processed air (heated, cooled) reached the first layer of the
compost instantaneously (no lag time).
 There are no uncertainty or variability measures included in the compost model.

A.3.2 Heater model

A.3.2.1 Conceptual Model


A heater was used to increase the temperature of the compost recirculation air,
which would in turn increase the temperature of the compost (Phillip and Clark, 2009). In
order to determine the temperature of the air exiting the heater, the following
conservation of energy equation was used:
Enthalpy out  Enthalpy in  Enthalpy added by heater

Figure 31 - Conceptual model of heater

The enthalpy of the air exiting the heater was used to solve for the temperature of the air
exiting the heater.

99
A.3.2.2 Mathematical Model
ha  C pa t a  hr(C pw t a  hwe )
ha Enthalpy of entering air (kJ/kg)
Cpa Specific heat capacity of dry air (kJ/kg°C)
ta Temperature of entering air (°C)
hr Humidity ratio (kg/kg)
Cpw Specific heat capacity of water vapour (kJ/kg°C)
hwe Evaporation heat of water (kJ/kg)

Qh  v w (hb  ha )
 w   a (1 hr)
Qh Heat flow rate of heater (kJ/s)
v Volumetric air flow (m3/s)
ρw Density of moist air (kg/m3)
ρa Density of dry air (kg/m3)
hb Enthalpy of exiting air (kJ/kg)

hb  ha  C pa (t b  t a )  hrC pw (t b  t a )
tb Temperature of heated air (°C)

Temperature of layer 1 after air has been heated:

Tout 
v  t b   VL  t a 
v  VL
VL Volume of layer (m3)

A.3.2.3 Model Assumptions

 When a voltage was applied to heater, the resulting heat energy was
instantaneous.
 Zero heat loss to air surrounding heater (i.e. all of heat energy was transferred to
re-circulation air).
 Perfect mixing of heated gases in layer 1.

100
A.3.3 Heat Exchanger Model

A.3.3.1 Conceptual Model


In order to remove heat from the compost recirculation air and cool the compost a
heat exchanger was employed (Phillip and Clark, 2009). In a heat exchanger, cold fluid is
passed through metal tubing which has metal fins connected to it. Air passes over the fins
and heat is exchanged between the warm air and the cold fluid. The efficiency of the heat
transfers is dependent on the specific heat capacities of the two fluids (air and water), the
area of contact, and the properties of the conducting material (copper). If the heat
exchanger effectiveness (ε) is known, the output temperature of the air can be found,
given the input temperature of the air and the water.

Figure 32 - Schematic diagram of heat exchanger

A.3.3.2 Mathematical Model


The following equations were used to determine the output temperature of air
exiting the heat exchanger.

q
T1,out  T1,in  
C1
T1,out Output air temperature (K)
T1,in Input air temperature (K)

q Heat transfer rate (kJ/s)

C1 Capacitance rate of air (kJ/s K)

101
 
q   C min (T1,in  T2,in )
T2,in Input water temperature (K)

C min Minimum capacitance rate between air and water (kJ/s K)
 Heat exchanger effectiveness

   

if C  C , C
 
1 2 2
C min 

 else, C1


C1  m1c1

 
C 2  m2 c2
c1 Specific heat of air (kJ/kg K)
c2 Specific heat of water (kJ/kg K)

C1 Capacitance rate of air (kJ/s K)

C2 Capacitance rate of water (kJ/s K)

 
m1  1 V 1
 
m 2  2 V 2

m1 Mass airflow rate (kg/s)

m2 Mass of water (kg/s)
1 Density of air (kg/m3)
2 Density of water (kg/m3)

V1 Volumetric flow rate of air (m3/s)

V2 Volumetric flow rate of water (m3/s)

A.3.3.3 Model Assumptions

 The specific heat capacity of air and water did not change with temperature.
 Heat exchanger efficiency was 0.9.

102
A.3.4 Air Supply Model

A.3.4.1 Conceptual Model


In order to calculate the increase in O2 concentration of the first layer of compost
due to the injection of air, the mass of O2 in the injected air was calculated and added to
the mass of O2 in the re-circulated air. This would result in a new O2 concentration for
layer 1.

Figure 33 - Conceptual model of air supply

A.3.4.2 Mathematical Model


Assuming an ideal gas, the molar concentration of air in the first layer of the
composting vessel was given by:

n P mol
 ( )
VL RT m3

n Number of moles (mol)


VL Volume of layer 1(m3)
P Pressure (kPa)
R Gas constant ( J K  mol )
T Temperature (K)

The density of O2, ρO2, is 16 g/mol, and the volume of air injected from the compressed
lab was given by:

V  v t

v Airflow rate (m3/sec)
t Time duration (sec)

103
Therefore, the mass of O2 added into the vessel was:
n
O  V
massO2 added 
2
V ( kg)
1000
ρO2: Density of oxygen (g/mol)

With this information, it was possible to calculate the change in O2 concentration of the
first layer of the compost (location of air injection) using the following equation:

initial _ massO2  massO2 added


[O2 ]out 
LayerVolume

A.3.4.3 Model Assumptions

 When air was added to the system, excess air was expelled from layer 6, so that
the total mass of air in the vessel remained constant.

A.3.5 Fan Model


The composting model defined by Stombaugh and Nokes included an airflow rate
parameter. Varying the value of the airflow rate in the model was considered the
equivalent of varying the air speed on the recirculation fan. The fan attached to the heat
exchanger in the system described by Phillip and Clark (2009) was ignored for this
model, even though it had an impact on the overall airflow rate through the compost
when the heat exchanger was on.

A.3.5.1 Model Assumptions

 There was uniform airflow through the compost bed (i.e. no preferential flow).

A.4. Results and Discussion

A.4.1 Open-loop system, no control


In order to validate the composting model that was implemented in MATLAB®,
the simulation results reported by Stombaugh and Nokes (1996) were compared with the

104
results from the MATLAB® simulation, using identical input values. The results of the
MATLAB® simulation is given in Figure 34.

(1) (2)

(3) (4)

Figure 34 - Simulation results, open-loop system. (1) Microorganism concentration,


(2) Substrate concentration, (3) Oxygen concentration, (4) Temperature of compost

The results from the simulation shown in Figure 34 confirmed that the computational
model developed in MATLAB® gave the same results as the model developed by
Stombaugh and Nokes (1996). For a detailed discussion of the simulation results, see
Stombaugh and Nokes (1996).

A.4.2 Closed-loop system, no control


Figure 35 shows the simulation results of the closed-loop composting system
represented in Figure 29, without employing any process control. These results showed
that as the biological activity inside the compost increased (microorganism
concentration), the temperature of the compost increased along with the O2 consumption
of the microorganisms that caused a decrease in the O2 concentration inside of the
composting vessel. When the O2 inside of the vessel was exhausted, the microorganism

105
concentration declined to zero (microorganisms became inactive) and the temperature
returned to ambient temperature (21°C).
(1) (2)

(3) (4)

Figure 35 - Simulation results, closed-loop system. (1) Microorganism concentration,


(2) Substrate concentration, (3) Oxygen concentration, (4) Temperature of compost

In contrast to the open-loop simulation (Fig. 34), there was little difference between the
five different compost layers for the four parameters that were being monitored during
the closed-loop simulation.

A.4.3 Temperature Control


Figure 36 shows the simulation results of a temperature control trial, where the
temperature set point was 25°C. The O2 inside of the vessel was exhausted by the
biological activity inside of the vessel; as a result, there was no heat being produced by
the compost after hour 25 of the simulation, and the heat was supplied exclusively by the
heater.

106
(1) (2)

(3) (4)

Figure 36 - Temperature control trial: 25°C set point. (1) Microorganism


concentration, (2) Substrate concentration, (3) Oxygen concentration, (4)
Temperature of compost

Because of the on-off control algorithm that was used for the temperature, a sinusoidal
behaviour was observed for the temperature in each later.

A.4.4 Oxygen Control


Figure 37 shows the results of an O2 control trial, where the O2 concentration set
point was 0.15 kg/m3. The O2 control system was able to maintain the O2 concentration at
the set point for approximately 10 hours (~hr 60 to hr 70), using an air supply flow rate of
0.0003 m3/s. However, the rate of O2 consumption by the microorganisms exceeded the
O2 supply from hour 70 to hr 80 of the simulation, and the O2 concentration within each
layer of the compost began to decrease. Once the rate of microbial activity decreased, the
O2 control system was able to restore the O2 concentration to the set point. Note: the O2
concentration was measured at layer 5.

107
(1) (2)

(3) (4)

Figure 37 - Oxygen control trial: 0.15 kg/m3 set point. (1) Microorganism
concentration, (2) Substrate concentration, (3) Oxygen concentration, (4)
Temperature of compost

This simulation demonstrates the need for a more sophisticated control algorithm for the
O2 concentration insides of the vessel. The on/off algorithm employed in this simulation
could not adjust the addition of air to meet the time-varying O2 demands of the
microorganisms inside of the compost. A proportional control algorithm, where the rate
of air supply is proportional to the error between the set point and the measured O2
concentration, could improve the control performance:

O2  k[O2 set  O2current ]

O2 Air supply rate (kg/sec)
k Constant
O2 set Oxygen concentration set point (%)
O2current Measured O2 concentration (%)

When the airflow rate was increased to meet the maximum O2 demands of the
microorganisms, there was significant overshoot in the O2 concentration above the set
point, outside of the period of maximum O2 consumption (Fig. 38).

108
(1) (2)

(3) (4)

Figure 38 - Oxygen control trial: 0.115 kg/m3 set point, control overshoot. (1)
Microorganism concentration, (2) Substrate concentration, (3) Oxygen
concentration, (4) Temperature of compost

When the O2 concentration set point was low (0.075 kg/m3, 10% concentration of O2 in
ambient air) the on/off control algorithm was able to maintain the O2 concentration at the
set point (Fig. 39).

109
(1) (2)

(3) (4)

Figure 39 - Oxygen control trial: 0.03 kg/m3 set point. (1) Microorganism
concentration, (2) Substrate concentration, (3) Oxygen concentration, (4)
Temperature of compost

Limiting the O2 availability had the effect of slowing the overall growth rate of the
microorganisms. The microorganism concentration reached 1.5 kg/m3 after
approximately 70 hours when the O2 set point was 0.15 kg/m3, compared to when the O2
set point was 0.03 kg/m3 where it took approximately 180 hours for the microorganism
concentration to reach 1.5 kg/m3.

A.4.5 Combined Control


Figure 40 shows the simulation results of a combined O2 and temperature control
trial, were the O2 set point was 0.03 kg/m3 and the temperature set point was 22°C.
At the beginning of the simulation, the temperature was below the set point and control
algorithm dictated that the heater should be turned on to increase the temperature of the
compost. When the temperature reached the set point, the heat exchanger was activated to
decrease the temperature of the compost, but the biologically produced heat gradually
increased the temperature of the compost, and the heat exchanger became ineffectual in
reducing the temperature of the compost. The temperature rise at the beginning the

110
simulation, compared to results in Figure 39, was more rapid due to active heating. The
heat exchanger had the effect of slowing the temperature rise of the compost, but could
not limit the compost temperature to the set point.
(1) (2)

(3) (4)

Figure 40 - Combined O2 and temperature control: 0.03(kg/m3), 22°C. (1)


Microorganism concentration, (2) Substrate concentration, (3) Oxygen
concentration, (4) Temperature of compost

The model heat exchanger, operating at maximum efficiency, given the cooling liquid
(water) was unable to remove a sufficient amount of heat from the compost recirculation
air in order to reduce the temperature of the compost to the set point

A.5 Conclusion
A computational model of a pilot-scale composting systems was created with
MATLAB®. The model was used to simulate controlled composting experiments using
an on/off control algorithm for temperature and O2 concentration. The control algorithm
was able to control the O2 concentration and temperature inside of the virtual composting
vessel independently. However, the system was unable to control both temperature and
O2 concentration simultaneously. Further work will be conducted to: 1) match the
computational model parameters to that of the physical system described by Phillip and

111
Clark (2009); and (2) implement a proportional control algorithm for O2 concentration
and temperature control within the computational model.

References
For references, see main reference listing on pg. 84.

112

Vous aimerez peut-être aussi