Vous êtes sur la page 1sur 201

Effect of different modes of aeration on composting of

Title solid waste in a closed system

Author(s) Bari, Quazi Hamidul

Citation

Issue Date 1999

URL http://hdl.handle.net/10722/31235

The author retains all proprietary rights, (such as patent


Rights rights) and the right to use in future works.
CHAPTER 1

INTRODUCTION

1.1 General

Increasing living standards of the human being through technological development


and population growth lead to an increase in the generation rate of solid wastes.
Future living standards and the quality of environment essentially depend on the
management policies for solid wastes together with the management of other
pollution problems. The increasing rate of solid waste generation from different
sources, limited landfill space and more stringent environmental regulations for new
landfill sites and incinerators have increased the waste disposal fees in different
countries of the world, especially in the developed countries. Therefore,
municipalities and local governments are under heavy pressure to find sustainable
and cost-effective waste management policies. The present waste management
hierarchy emphasizes reduce, reuse, recycle, recover and stable residue. After
reducing and reusing there will be substantial amounts of waste left to be managed
and a large portion of this waste is biodegradable; for example, source separated
organic solid wastes including household refuse, market wastes and agricultural
wastes or sewage sludge. The biodegradable portion could be managed either by
recycling and recovering through biological treatment, or disposal to landfills.

Usually the organic wastes which are disposed directly or together with other wastes
to landfills create further long term problems by producing secondary pollutants
including methane, ammonia, hydrogen sulfide, volatile organic compounds and
leachate through anaerobic decomposition. These pollutants are entering either the
1
atmosphere as gases or the ground water as leachate. Hence new landfills require
separate units for long term collection, treatment and monitoring of these pollutants.
However, biological treatment offers a cost-effective sustainable solution for the
biodegradable organic wastes. Biological treatment is a very effective process for
recovering waste materials and for minimizing the above mentioned problems in
landfills by stabilizing the organic wastes in the shortest period of time. In practice,
the main biological process applied for solid wastes is composting (Haug 1993,
Tchobanoglous et al. 1993).

Composting is the biological degradation of highly concentrated biodegradable


organic wastes in the presence of oxygen (aerobic decomposition) to carbon dioxide
and water, whereby the biologically generated waste heat is sufficient to raise the
temperature of the composting mass to the thermophilic range (50 to 65 oC). The
final product of composting is a stable humus-like material known as compost. The
popularity of composting has increased in the past decade due to several
environmental benefits such as:

fast conversion of the organic solid waste to a biologically stable end product
(Christensen and Nielsen 1983),
recovery of waste material in the form of compost for utilization in agriculture,
horticulture or other applications, as a soil conditioner, potting soil, organic
fertilizer and landscaping material,
effective hygienization of pathogenic bacteria present in the organic waste
(Willson 1983, Obeng and Wright 1987),
stabilization and volume reduction of the waste materials prior to
environmentally sound final disposal in landfills,
cheap and effective solid waste treatment method (Stentiford et al. 1985a, de
Bertoldi et al. 1988, Magalhaes et al. 1993, Sesay et al. 1998).

In Europe, the entire organic portion of municipal solid waste is often composted.
Until recently, that approach had been less common in the USA (Tardy and Beck
1995). Currently many large cities in China and Southeast Asia are planning to erect

2
or improve existing municipal waste composting plants with capacities of up to 1500
t/d (Raninger 1996). Sewage sludge composting has also become very common
since the 1970s in the USA (Miller 1991).

In Hong Kong the organic wastes constitute about 25% of the waste disposed of in
landfills (Draft Waste Reduction Plan for Hong Kong 1997) and about 57% of
domestic waste (Environment Hong Kong 1995). To reduce the pollution problems
in the landfills, it may be necessary for Hong Kong to establish new composting
plants for separately collected organic waste. A composting plant for livestock
wastes with 75 t/d capacity is already successfully operated in Hong Kong.

Different composting technologies depending on the quality of the initial substrate,


processing time and process control have been practiced. The main technologies are
forced aeration, mechanical turnover in a reactor or in a windrow composting pile.
The reactors could be static or slowly rotating and the windrow could be formed in
an open field or inside a shelter. Furthermore the process could be batch or
continuous; however, the batch process is normally applied for large-scale
composting (Sikora et al. 1981, Epstein et al. 1983, Benedict et al. 1986).

Recent applications of large-scale composting have often been plagued by


technological problems, lack of understanding of the biological fundamentals, and
unsuitable waste materials resulting in unacceptable product quality as well as
environmental nuisances such as odor, leachates, and contaminated aerosols. Several
large municipal solid waste composting facilities in the USA including those in
Portland, Oregon; Dade County, Maryland; and Pembroke Pines, Florida; have been
shutdown because of odors (Epstein 1997). In Hong Kong, the former Chai Wan
composting facility also represents such a negative example. The forced aeration
composting process has been proposed to overcome the above problems (ORCA
1991). Variations of this process such as closed container composting or tunnel
composting have been widely applied in recent years in Europe (Wannholt 1995,
Czermak and Gruneklee 1995). However, few scientific results or evaluations have
been published (Baumann et al., 1995) and more work needs to be done to determine

3
the best control methods and specifications (Hyatt 1995). Manser and Keeling
(1996) stated that the composting process has been widely studied but is still not
fully understood.

A large number of bench and small-scale experiments were performed to evaluate


the effect of aeration, temperature, and moisture content on composting using static
(Magalhaes et. al. 1993, Garcia et al. 1996) and continuous or intermittently mixed
(Schulze 1960, Bech et al. 1984) reactors. These bench and small-scale reactors,
ranging in volume from 0.4 to 16 L are helpful to understand effects under ideal and
precisely controlled conditions. For understanding the critical field conditions with
regard to variations or gradients in temperature, moisture or other physico-chemical
parameters, it is necessary to conduct tests on a pilot-scale basis. Some small-scale
(Hogan et al. 1989, Garcia et al. 1996) and pilot-scale (Kuter et al. 1985) studies
showed only temperature gradients. Effects of different modes of aeration, specially
air recirculation and reuse of hot spent air of one reactor in another reactor, or
multilayer analysis of different physico-chemical parameters were not investigated
so far. Therefore, in this study, forced aeration composting in specifically designed
pilot-scale reactors was investigated together with bench-scale tests, in order to
obtain a better understanding of the above mentioned and other important factors and
to extend the findings for better design of future composting plants.

1.2 Objectives

In this study, forced aeration composting tests using two heat insulated closed pilot-
scale reactors were conducted to evaluate the following:

1. distribution of physico-chemical parameters including temperature and moisture


throughout the height of the composting mass;
2. rate and extent of organic waste degradation during composting; and
3. the order of reaction rates (degradation rates) and their dependence on
temperature.

The tests were designed to evaluate the effects on the above phenomena of the
following aeration modes: upflow, downflow, upflow/downflow, and internal
4
recirculation in a single-reactor system as well as reuse of spent air in a two-reactor
system in series. An attempt was also be made to develop a simple mathematical
model to simulate and predict the extent of organic waste degradation. For
comparison, several bench-scale composting tests were performed using similar
waste mixtures under forced aeration and self-heating conditions. Implications of the
results with regard to design and operation of large-scale composting are discussed.

1.3 Organization of Thesis

The importance of composting in the field of solid waste management and its
technological problems are discussed in Chapter 1. In Chapter 2, past and present
composting practices are reviewed. Chapter 2 also describes (i) the factors affecting
the composting process and their interrelated effects; and (ii) measurement
techniques of product quality. Theoretical considerations which form the basis for
prediction of organic waste degradation during composting are reviewed in Chapter
3. Detailed experimental programs and analytical methods for physico-chemical
parameters are presented in Chapter 4. Experimental programs of Chapter 4 include
pilot-scale composting tests, bench-scale composting tests and self-heating tests to
assess the product quality. Chapter 5 contains the results of the experiments
described in Chapter 4. Chapter 6 contains the discussion and interpretation of
experimental results. It mainly describes the effect of different modes of aeration
during pilot-scale tests on temperature, oxygen consumption rate, moisture content,
extent of degradation, product stability and on other physico-chemical parameters.
Kinetic analysis of the experimental results, based on the theoretical considerations
described in Chapter 3, and their application are presented in Chapter 7. In Chapter
8, engineering significance of pilot-scale composting and bench-scale composting
are discussed. Chapter 9 outlines the conclusions of this research and comments on
possible future research. Appendix I contains experimental results of self-heating
tests and a detailed estimation of the material balance for pilot-scale composting
tests.

5
CHAPTER 2

LITERATURE REVIEW

2.1 Historical Overview

Small-scale composting of nightsoil and other organic wastes has been used
successfully in Chinese agriculture since times immemorial. The late British
agronomist Sir Albert Howard conducted studies and experiments, between the years
1924 to 1931, in India that established the basic principles of composting (Spellman
1997). His process is known as Indore Process. The Indore Process represented the
first organized plan for composting in the modern era (Haug 1993). That process was
usually conducted in pits or piles of 9 x 4 x 0.9 m deep or high. Preparation of the
compost pile consisted of the alternate thick and thin layers of vegetable waste (152
mm) and animal manure (51 mm), respectively. The layering was repeated until a
height of 0.9 m was reached. The pile was turned after 16, 30 and 60 days with
intermittent re-moisturing and the whole process was completed after 90 days.

The first full-scale refuse composting facility in Europe was established in the
Netherlands in 1932 and was a modification of the Indore process (Epstein 1997). In
the 1940s, several mechanical composting systems were introduced in Europe (Kuter
1995). Composting activities in Australia, Japan, New Zealand and United States
during the 1950s are also reported (Snell 1957, Mantell 1975). However, during that
early stage of composting most large-scale plants were established in Europe. After
30 to 50 years, several early researches including Influence of temperature on the
microorganisms (Waksman et al. 1939); Evaluation of inoculums in composting
(Golueke et al. 1954); Rate of composting (Wiley and Pearce 1955, Snell 1957);
and Optimum moisture content and rate of oxygen consumption during composting
6
(Schulze 1960, 1962a) are still widely cited in the literature and books regarding
composting. Many advances have been made in the field of sludge composting since
the 1970s because of the demise of the open dump (Kuter 1995, Golueke and Diaz
1996) and its several advantages, as mentioned in the previous chapter, over other
disposal systems.

2.2 Composting Substrates

A wide variety of organic wastes generated from different sources or combinations


thereof can be used as compost substrate. The wastes should be free from
uncompostable materials such as plastics, glass, metal objects and hazardous
compounds in order to produce good quality compost. The major sources of organic
wastes are the organic fraction of municipal solid wastes, park and yard wastes,
industrial wastes, agricultural wastes and sewage sludge. The major examples of
waste generated from these sources are as follows:

Organic fraction of municipal solid wastes mainly from households, restaurants,


and supermarkets such as food residue, vegetable trimmings, unused portions of
fruit, paper, etc.
Park and yard wastes including grass clippings, leaves and paper.
Industrial wastes from food and wood processing industries, namely fishery by-
products (Brinton and Seekins 1994), fish and meat wastes, vegetables and
grains, paper and pulp sludge (Sesay et al. 1997), waxed corrugated cardboard
(Kunzler and Farrell 1996), corn fiber, wheat chaff, saw dust, wood shavings,
etc.
Agricultural wastes such as crop residue and livestock wastes, mainly animal
manure (Mondini et al. 1996, Sartaj et al. 1997).
Sewage sludge or biosolids from municipal wastewater treatment plants.

The fundamental constituents of the organic wastes are carbohydrates and sugar, fat,
protein, hemicellulose, cellulose, lignin and ash (Waksman 1939, Skitt 1972, Serra-
Wittling et al. 1996). The percentage of each constituent is different in different
7
organic wastes. The easily biodegradable constituents are carbohydrate, sugar, fat
and protein (de Bertoldi et al. 1983, Epstein 1997). These are present in higher
percentage in food wastes. Although hemicellulose, cellulose and lignin enriched
materials are less biodegradable (Snell 1957, Crawford 1985), they are often added
with other wastes for proper adjustment of C/N ratio, moisture content and porosity.
Examples of less biodegradable substrates are paper, straw and sawdust. The
appropriate process technology, quality and subsequent utilization of compost
considerably depend on the type and initial quality of the substrates.

2.3 Microorganisms

A large variety of microorganisms are responsible for composting. The


microorganisms are abundant in air and soil as well as in the wastes to be
composted. The odor emission from a garbage containing bag after one or two days
of storage represents the presence of these microorganisms and their decomposition
activities in the wastes. Microorganisms can be classified into three main groups
based on their temperature ranges for growth (Brock and Madigan 1991, Prescott et
al. 1996) as illustrated in Table 2.1. Various genera of bacteria, fungi, algae and
protozoa belong to each group.

Table 2.1 Classification of microorganisms based on their


temperature ranges for growth
Microorganisms Temperature range in oC
Psychrophilic 0 - 20
Mesophilic 20 - 45
Thermophilic > 45

The main microorganisms responsible for biological degradation in composting are


bacteria, actinomycetes and fungi (Golueke 1977, Composting 1985, Thambirajah et
al. 1995, Polprasert 1996) of mesophilic and thermophilic groups. Diaz et al. (1993)
noted that the bacteria and fungi are characterized by the successive appearance of
their mesophilic and thermophilic forms. Although a large number of mesophilic

8
bacteria can survive at 60 oC by forming colonies in the compost material, they
contribute almost nothing to the degradation of organic matter in composting
(Nakasaki et al. 1985b). The optimum temperature range for growth of fungi is 40 to
50 oC (Kane and Mullins 1973, Prescott et al. 1996) and the upper temperature limits
are 60 to 62 oC (Brock and Madigan 1991). It was shown that fungi were completely
absent in the peak temperature phase of 60 to 70 oC during composting of oil palm
empty-fruit-bunches and different livestock wastes (Thambirajah et al. 1995).
Nakasaki et al. (1985a) indicated that actinomycetes develop far more slowly than
most bacteria or fungi and are rather ineffective competitors when nutrient levels are
high and they became more active in a later period at 60 oC during the composting of
sewage sludge in a small reactor of 28 L volume. Furthermore, it was also found that
the actinomycetes are more tolerant to high temperature (Waksman 1939). Aerobic
thermophilic composting is a dynamic process which is brought about by the
combined activities of a rapid succession of mixed microbial populations, each
suited to an environment of relatively limited duration and each being active in the
decomposition of one particular type or group of organic materials (Skitt 1972). The
recovery of microorganisms and their specific activity from temperature and other
shocks during composting have not been studied so far.

It was found that no special inocula or seed is required for satisfactory composting of
garbage and other organic refuse (Golueke et al. 1954). Recently Nakasaki et al.
(1985c) observed that the rate of composting sewage sludge was mainly controlled
by the degradability of solid substrates and not by the kinds of microorganisms
inhabiting the compost. However, inocula are needed for special kinds of waste such
as wool waste (Tiwari et al. 1989) and hazardous waste.

2.4 Composting Processes

Composting is the controlled biological decomposition of organic waste into a


stable, pathogen free end product. Although this biological decomposition can take
place under aerobic or anaerobic conditions, composting is mainly considered as an
aerobic process (Schulze 1960, Willson 1983, Clark et al. 1984, Stentiford et al.

9
1985a, Obeng and Wright 1987, Darbyshire et al. 1989). Furthermore, most
practiced and controlled composting processes are aerobic (Rogalski and Charlton
1995, Manser and Keeling 1996, Martin 1998). Anaerobic composting is the
digestion or fermentation of organic matter under anaerobic conditions
(Tchobanoglous et al. 1993, Epstein 1997) and is generally applied for the
production of biogas, mainly methane (de Bertoldi et al 1988, Diaz et al. 1993, Six et
al. 1994). Anaerobic composting has higher odor potential because of the nature of
many intermediate metabolites (Haug 1993), whereas aerobic composting minimizes
the potential of nuisance odors (Metcalf & Eddy 1991). The decomposition rate in
anaerobic composting is also very slow. Therefore, in this study, composting will
only refer to the aerobic composting process.

The decomposition of organic wastes in composting can be described by the


following equation:

microoganisms
Organic Wastes + O 2 Compost + CO 2 + H 2 O + NH 3 + Heat (2.1)

As indicated in Equation 2.1, microorganisms decompose or oxidize the organic


compounds to simple, stabilized end products, with the production of heat. During
the process oxygen is consumed and carbon dioxide, water, and often ammonia are
released. The heat energy is partially used for cell synthesis of the microorganisms.
However, the heat production is sufficient to raise the temperature up to the
thermophilic range.

The composting process can be explained in many ways. According to several


researchers (Gray et al. 1971a, Obeng and Wright 1987, Polprasert 1996), the
composting process can be divided into four phases as related to temperature,
namely, (a) lag phase, (b) growth phase, (c) thermophilic phase, and (d) maturation
phase. These phases are illustrated in Figure 2.1.

10
70

60

Temperature in oC 50

40

Bacteria
30
Bacteria, fungi, Bacteria, fungi, and
20
Lag Phase

actinomycetes higher organisms


Growth
Phase

10
Thermophilic Phase Maturation Phase
0
Time

Figure 2.1 Different phases in the composting process as related to


temperature (Polprasert 1996, modified).

(a) Lag phase: this is the time phase necessary for the microorganisms to acclimatize
with the new environment in the composting mass. Haug (1993) suggested that
the lag phase can be avoided in actual practice by using product recycle to supply
an acclimated seed.

(b) Growth phase: in this phase the mesophilic microorganisms start to decompose
readily degradable carbohydrates, sugar and proteins and to generate heat. In this
initial phase simple organic acids are produced by acidogenic bacteria and these
cause a drop in pH. The rapid accumulation of heat raises the temperature up to
the mesophilic level (from ambient to 45 oC).

(c) Thermophilic phase: At temperatures above 40 to 45 oC the activity of


mesophiles is reduced and degradation is taken over by the thermophilic bacteria
and fungi. Finally the temperature rises to the highest level of about 50 to 65 oC
or more. At 60 oC the thermophilic fungi die off and the reaction is kept going by
spore forming bacteria and actinomycetes (Sterritt and Lester 1988). Generally,
the degradation rates are higher in this phase and mostly the protein, lipid, fat
and hemicellulose fractions are broken down (Composting 1985). In this phase,

11
the pH changes to the alkaline range. This phase is also significant for pathogen
destruction. The duration of the optimum thermophilic phase can be extended,
until the nutrients become limiting to the survival of the microorganisms, by
manipulating the aeration. However, this phase may continue for several weeks
or months depending on the composting system and initial waste mixture.
Different volatile compounds are released together with carbon dioxide, water
and ammonia nitrogen as indicated in Equation 2.1. However, the amounts of
emitted NH3-N and volatile organic compounds are very small in comparison to
CO2 and H2O (Epstein 1997).

(d) Maturation phase: As the rapidly degradable material is depleted, the reaction
rate slows down and the temperature again decreases from the thermophilic to
the mesophilic level. This mesophilic phase is quite longer in the maturation
phase than in the growth phase. Sterritt and Lester (1988) state that the bacterial
and fungal biomass generated in the earlier phases of composting can become
food for a succession of higher organisms which may be associated with the
process, including protozoa, rotifers and nematodes. However, in this phase
fungi and actinomycetes are also found quite active as shown by Strom (1985b).
As the process approaches completion, the nutrients become rate limiting and the
temperature eventually returns to ambient level. The pH drops again slightly but
usually remains on the alkaline side of neutrality. The maturation phase may last
for a period of months depending on the desired product quality and utilization.
However, the duration of maturation phase can be minimized using a controlled
second stage composting process.

It should be noted that different parts of the composting mass may, at the same time,
achieve either mesophilic or thermophilic temperatures and the duration of these
phases also may be different. The shape of the temperature curve (for example
Figure 2.1) depends on the initial waste materials being composted and the
composting methods (Epstein 1997). However, in a controlled reactor system the
variation in temperature distribution can be improved.

12
2.5 Classification of Composting Systems

Various classifications of composting technologies are mentioned in different


references including some additional terms such as mesophilic composting,
thermophilic composting, vermi-composting, passively aerated composting, etc.
However, the technologies are often classified into two broad groups, namely, open
(non-reactor system) and closed (in-vessel or reactor system) system (Haug 1993, de
Bertoldi et al. 1985, Wannholt 1995, Manser and Keeling 1996, Mathur 1998).
Sometimes, a combination of these systems is also practiced. Mostly, the closed
system is used for the initial high decomposition stage and after that maturation or
curing takes place in the open system. Based on the above mentioned references a
detailed classification of composting systems is presented in Figure 2.2 and each
system is discussed in the following sections.

Composting System

Open or Non-reactor Closed or Reactor


system system

windrow Aerated static Aerated turned Batch Continuous


pile pile process process

Horizontal or Vertical
inclined materials
materials flow
flow

Figure 2.2 Detailed classification of composting systems.

2.5.1 Open or Non-reactor System

Open or non-reactor systems entail the formation of long windrows or piles from a
composting substrate. The width and height of the piles ranges from 2 to 4.5 m and 1
to 2.5 m, respectively. The windrow or pile could be formed in an open field or
13
under a shade. These systems are further classified as (i) windrow, (ii) aerated static
pile (forced aeration) and (iii) aerated turned pile (forced aeration).

(i) Windrow: In the turned windrow system, aeration is provided by mechanical


turning and natural ventilation. Mobile equipment provides frequent turning during a
composting period of several weeks. Turning also promotes uniform decomposition
and sanitization of the end product. Due to the emission of offensive odors, long
processing time, large area requirements, and other limitations, this system is not
usually appropriate in urban areas except for yard wastes. Turned windrow systems
need a long time to produce a stable compost. Garcia et al. (1992) reported that
different mixtures of wastes were composted over three months in the open air with
periodical turning and were left to mature afterwards for a further four months.
Therefore, in this case the total processing time was about seven months. Sharma et
al. (1997) noted that, depending on the properties of the input substance and turning
frequency rate, the active phase of the composting lasts for three to nine months.
Long processing times from four to eight months are common for open systems
(Canet and Pomares 1995, Gies 1995, Goldstein 1997).

(ii) Aerated static pile (forced aeration): In the aerated static pile system the
composting piles are formed on a porous aeration base or air distribution system.
The base includes a layer of wood chips or straw which are placed over perforated
pipes. Usually air blowers or fans are used for forced aeration. In upflow (positive
pressure) aeration, air is blown through the perforated pipes and then distributed
through the aeration base. In downflow (negative pressure) aeration, air is sucked
from the pile via the aeration base. A layer of mature compost is used for cover of
the aerated composting pile to improve the insulation of the pile. Instead of forming
windrows, some systems have been using rectangular cross-sectioned long open
channels with concrete side walls and aeration base at the floor (Gies 1993).

(iii) Aerated turned pile (forced aeration): In this system, rails or tracks are so
installed on the top of the aeration base at the floor of rectangular cross-sectioned
long open channels with concrete side walls, that a turning machine can move. It

14
may overcome the problems associated with uneven settling of materials and
excessive drying because water can be added during the turning operation (Manser
and Keeling 1996).

2.5.2 Closed or Reactor System

Closed or reactor systems are sophisticated methods in which composting is


conducted within a fully closed system. High quality compost, maximum
composting efficiency and minimization of environmental nuisances are achievable,
if a proper controlling system is adopted. In practice, two to three weeks of
composting time have been assigned for the initial high decomposition stage.
However, most closed composting systems have a separate curing and storage step
after the initial stage. Sometimes curing is also carried out in the reactor, though
normally curing is performed in an open system. The closed system can be further
classified into two processes, namely, (i) batch process and (ii) continuous process.

(i) Batch process: The reactors used in the batch process are cubical or cylindrical in
shape. Sometimes open top reactors are also used. Usually, the height of these
reactors is 2 to 2.5 m and/or diameter 2.5 to 3 m. Aeration is provided from the
bottom of the reactor. The waste materials are re-mixed and re-moistured
periodically. These are conducted either inside or outside using a mechanical agitator
or a front end loader, respectively. Some of the batch reactors may have the
provision to rotate slowly or intermittently. Examples of this process are container,
box and drum composter.

(ii) Continuous process: Different types of continuous processes have been


developed. These can be classified as (a) horizontal and inclined material flow
reactor and (b) vertical material flow reactor. In both types, the waste materials may
remain under mixing or non-mixing condition, though they are continuously moving
from one end to another or from top to bottom, respectively. Aeration is also
provided from the bottom floor. One common example of horizontal and inclined
material flow reactors is the tunnel reactor. In the tunnel reactor, the waste is fed at

15
one end and the compost is discharged at the other end. An example of vertical
material flow reactors is the silo reactor. In this process, waste is fed at the top and
the compost is removed at the bottom.

However it should be noted that, sometimes, the tunnel composter is used as a batch
process. The three tunnel composters at Ipswich (UK) are loaded with a waste
mixture for fourteen days for initial high degradation and after that the waste is
allowed a three months curing period in the open (Barnes 1998).

2.6 Factors Affecting the Composting Process

The rate of microbial activity or degradation in the composting mass depends on


certain important physical and chemical factors which should be considered in the
design and operation of a composting process. These factors are particle size, C/N
ratio, water content, temperature, pH and aeration. During composting, the first five
factors change with time whereas aeration supplies oxygen, removes excess moisture
and heat and thereby plays as a key role in process control. All factors are discussed
separately in the following sections.

2.6.1 Particle Size and Porosity

The particle size of the substrate can vary depending on the physical nature of the
waste materials. The particle size of composting materials should be as small as
possible so as to allow for efficient aeration and to be easily decomposed by
bacteria, fungi and actinomycetes (Jeris and Regan 1973b, Polprasert 1996), because
smaller particles have a greater surface to volume ratio. Thus more surface area is
available for microbial decomposition and subsequently the composting efficiency is
increased. Gray (1971b) reported that size reduction reduces the depth for oxygen
diffusion and microbial advance within the particle, aids the homogenizing of an
initially heterogeneous material and improves insulation. Diaz et al. (1993) reported
that in practice the minimum permissible particle size is that one at which the
porosity required for proper aeration into the composting mass can still be attained

16
and maintained. This maintenance depends upon the structural strength of the
particles. Schulze (1962b) suggested that the bulk weight of the prepared feed
mixture should be between 400 to 500 kg/m3 and moisture content between 50 to
60% so that a minimum of 30% free air space is provided.

The distribution of various sized particles in a waste or compost sample can be


determined by sieve analysis. This is a mechanical method used for the
determination of particle size distribution of coarse-grained soil. In sieve analysis the
sample is passed through a series of test sieves from larger to smaller mesh and the
weight of the sample retained in each sieve is determined. The particle size
distribution of a sample is presented as a curve, plotting the cumulative percentage
by weight passing each sieve versus the sieve size on a semi-logarithmic plot. Smith
(1994) noted that for soil samples a uniform sample has an almost vertical curve; a
well graded sample has a curve spread evenly across the chart; and a poorly graded
sample will stretch across the chart, but may be deficient in intermediate sizes.

Different types of screening or shredding machines are used, respectively, for sorting
or cutting the substrate to get the required particle size. Usually a sorting or cutting
process is applied for municipal solid wastes and agricultural residue. Shredding is
also applied for oversize bulky materials to make them suitable as structural support
in the composting of sludge and livestock wastes.

2.6.2 C/N Ratio

The carbon to nitrogen (C/N) ratio is the most important indicator of the availability
of nutrients for microbial use in composting. Nitrogen is the major nutrient required
by microorganisms in the assimilation of carbon substrate from organic wastes.
Phosphorus is next in importance while potassium, magnesium, sulfur, calcium and
trace quantities of several other elements all play a part in cell metabolism (Skitt
1972). Diaz et al. (1993) state that according to experience almost without exception
all other nutrients than nitrogen are present in the typical organic waste in adequate
amounts and ratios. Carbon is an energy source essential for cell growth and nitrogen

17
is the major source for proteins needed by the microorganisms (Miller and Jones
1995).
Nakasaki et al. (1992) examined the effects of the C/N ratio on the thermophilic
composting of ground garbage using a 5.1 kg mixture in each of the runs. Among
three runs of different C/N ratios; 13.9, 22.4 and 30.9, the rise in the rate of CO2
evolution occurred earliest at a C/N ratio of 22.4 and the final conversion of carbon
was the highest in the same run. The optimum range of the C/N ratio for most types
of composting substrate is 20 to 30 (Golueke 1977, Obeng and Wright 1987, Vallini
and Pera 1989, Golueke and Diaz 1990, Fernandes and Sartaj 1997). Deviation from
this range leads to a slower composting rate. High C/N ratios inhibit the growth of
microorganisms and thereby reduce the decomposition rate. At low C/N ratios
nitrogen could be lost as ammonia, especially at conditions of high temperature and
forced aeration. Extremely high amounts of nitrogen in a composting mass can form
enough ammonia to be toxic to microbial population, further inhibiting the process
(USEPA 1992). High or low C/N ratios can be adjusted by adding high nitrogen or
carbon rich wastes, respectively. Sawdust, wheat straw, grass clippings, dry leaves,
etc. are examples of carbon-rich materials, whereas poultry manure, slaughterhouse
waste, sewage sludge, etc. are nitrogen-rich. The C/N ratio decreases during the
composting process as carbon is lost in the form of carbon dioxide.

2.6.3 Moisture Content

Moisture content (MC) is often termed as water content (WC). Microorganisms need
a certain amount of water for their metabolism and reproduction. Water is the key
ingredient that transports substances within the composting mass and makes the
nutrients physically and chemically accessible to the microorganisms (USEPA
1992). Biological activity is greatly reduced at substrate moisture contents below
40% (Epstein 1997, Composting 1985) although some claim that composting can be
achieved at moisture contents as low as 30% (Skitt 1972).

The upper limit of moisture content may vary from 55 to 85% depending on the
structural strength of the materials at wet condition and availability of the pore space

18
for proper aeration. Fibrous or bulky material such as straw or wood chips can
absorb relatively large quantities of water and still maintain their structural integrity
and porosity (Haug 1993). According to Golueke (1977) the maximum permissible
moisture content for straw and rice hulls is 75 to 85% and for municipal refuse and
manure 55 to 65%.

Wiley and Pearce (1955) conducted composting experiments of mixed garbage using
six continuously mixed (0.03 to 0.60 rpm) batch reactors of 57 L volume each. They
performed replicate tests for low moisture (40 to 53%), medium moisture (55 to
69%) and high moisture (72 to 77%). The peak temperatures for the three tests were
54, 64 and 43 oC, respectively; and the overall highest temperature prevailed in the
composting mass with medium moisture content. They concluded that the optimum
moisture content of mixed garbage was 55 to 69%, and moisture content below 50%
or above 72% did not permit optimum composting under their test conditions.

Gray et al. (1973) suggested that the range of moisture should be 50 to 70%.
However, the range of optimum moisture content cited in many other references is
50 to 60% (Schulze 1962, Hay and Kuchenrither 1990). A significant amount of
moisture could be removed at thermophilic temperature during forced aeration
composting. Sometimes it is necessary to adjust moisture content through irrigation
or other means such as internal air recirculation or reuse of hot air from one reactor
to another. The suitability of the latter methods is investigated in this research.

Excessive moisture can have several adverse effects on the subsequent processes
after the final decomposition phase. Finstein et al. (1986a) state that the final
compost should be dry, since it decreases weight and bulk and improves materials
handling, processing operation, storage and transport.

2.6.4 pH

Considerable changes in pH value occur during the composting process. In the


beginning, the formation of organic acids and carbon dioxide lower the pH value to

19
approximately 5 or less, whereas with process progress the pH value reaches up to 8
to 8.5 (USEPA 1992, Tchobanoglous et al. 1993, Sharma et al. 1997). McKinley and
Vestal (1985) proposed that the increase in pH may be an indirect indicator of high
levels of microbial activity in composting. Jeris and Regan (1973c) obtained
maximum thermophilic composting rates at a pH range of 7.5 to 8.5, by measuring
O2 consumption rate or CO2 evolution rate.

The pH level desired for microbial activity is between 5 and 7 (Miller and Jones
1995). Savage et al. (1988) state that in practice an initial pH level of 5.5 or 8 would
not have a serious effect on composting time and overall composting progress. If the
pH is too low or high then the lag phase could be prolonged at the beginning. Three
days of lag period were observed during composting of sewage sludge containing
lime with pH value of 11 (Nakasaki et al. 1985a). Galler and Davey (1971, cited by
Finstein and Morris 1975) reported that a slower initial temperature rise was
associated with the more acidic samples, and the temperature remained below 45 oC
until pH 7.0 was exceeded. However, for composting an optimum pH of 5.5 to 8.5,
with little variation, is recommended in most of the literature (Inbar et al. 1990,
Miller 1991, Kapetanios et al. 1993).

2.6.5 Temperature

Temperature affects the composting process in several ways. The species and
number of microorganisms change with increments in temperature. The
decomposition rate and the heat production is affected by these microorganisms. Due
to accumulation of heat the temperature rises first to the mesophilic phase (25 to 45
o
C) followed by the thermophilic phase (over 45 oC). After the thermophilic phase
the temperature again decreases to the mesophilic phase and finally to ambient level.
A typical temperature pattern is similar to that presented in Figure 2.1. Usually the
thermophilic temperature is attained after 1 to 4 days. Schwab et al. (1994) reported
that the compost mass in a pilot-scale plant self-heated from an ambient temperature
of 27 oC to about 55 oC in the first 24 hours. Lau et al. (1992) found that a
temperature of 55 oC was reached within 2 days and remained above this level for 4

20
to 7 days in the experimental reactor of 55 L. Vallini and Pera (1989) found that the
attainment of thermophilic temperature took a longer time; the temperature remained
at 27 to 30 oC for the first 8 days and then gradually rose to 60 oC after 15 days.

Hay and Kuchenrither (1990) reported that temperatures above 60 oC are common in
large windrows and in certain instances the temperature stayed above 71 oC for
several days. Maximum reported thermophilic temperatures were 74 oC for a pilot-
scale reactor (VanderGheynst et al. 1997), 75 oC in a forced aerated composting pile
(Willson 1983), 78 oC at the center of an aerated pile (Epstein et al. 1976), and 80 oC
at the top layers of a pilot-scale pile (Bhamidimarri and Pandey 1996). Although
these upper temperatures are beneficial for pathogen destruction, they are harmful to
the major groups of the microorganisms responsible for biological decomposition. It
was found that the decomposition rate was greatly reduced above 60 oC (Sular and
Finstein 1977, MacGregor et al. 1981) and 55 oC (McKinley and Vestal 1984, 1985).
Kuter et al. (1985) indicated that the microbial activity was greatest when mean
compost temperatures were 40 to 60 oC in a full-scale vessel system. Waksman et al.
(1939) state that organic matter degradation rate was maximum at 65 oC in the
composting of horse manure and wheat straw.

Bach et al. (1984) conducted an experimental investigation on the rate of


composting, as monitored by the measurement of CO2 evolution rate, of dewatered
sludge mixed with rice husks and crushed rice husks using a continuously mixed
isothermal reactor of 16 L. The CO2 evolution rate was maximum at around 60 oC.
In another experiment (Bach et al. 1985) using an autothermal packed bed reactor of
28 L, the optimum temperature was also found to be close to 60 oC. Strom (1985a)
observed that the bacterial species diversity decreased markedly in a laboratory
composter (of 4.5 L size) at 60 oC, but was similar for other runs incubated at 49, 50
and 55 oC. According to this result the maximum desirable composting temperature
based on bacterial species diversity is 60 oC. The results obtained by Bach et al.
(1984, 1985) and Strom (1985a) are consistent with the findings of other workers
(Nakasaki et al. 1985d).

21
Although there is some variation in the optimum temperature range due to variations
in waste materials and operational practices, in most of the cases it is reported as 55
to 60 oC. On the other hand, an acceptable level of pathogen destruction results when
temperatures are maintained at 55 oC or above for 3 consecutive days (Burge et al.
1981, Golueke 1983). Thus from the above discussion the range of optimum
temperature appears to be 55 to 60 oC.

2.6.6 Oxygen and Aeration

Aeration supplies oxygen to the microorganisms for aerobic biological degradation


of organic wastes in the composting mass and is a key process control parameter of
the forced aeration composting system. Insufficient or mal-distributed aeration leads
to the onset of anaerobic conditions, with a consequent decrease in rate of
decomposition and evolution of offensive odor (Skitt 1972). This occurs when the
oxygen concentration in the air within the mass falls below 5 percent (Schulze 1960,
Fernandes and Sartaj 1997) or 10% (Sular and Finstein 1977). Aeration requirements
for biological degradation can be determined from the stoichiometric reaction of
organic waste oxidation.

As a process control parameter the important application of aeration is to supply


oxygen and to remove excess heat generated by the microbial activity for
maintaining the optimum temperature. The air supply needs for temperature and
moisture control typically are ten or more times greater than those for biological
decomposition, so that when these needs are met, biological oxygen demands also
will be safely satisfied (Kuter 1995). Leton and Stentiford (1990) noted that oxygen
concentrations in excess of 18% are common during the highly active
biodegradation stage in order to keep the temperature in the composting mass within
favorable levels (40 to 60 oC) without any serious side effects. Several systems have
been applied to provide aeration for composting. The main systems are mechanical
turning and forced aeration via air blower or fan. Forced aeration composting is
usually applied to static windrow, static pile and to most of the reactor systems.

22
Usually at the early stage of composting excess heat should be removed, to maintain
a temperature below 60 oC, via high rate of aeration. On the other hand, at the later
stage (maturation stage) a low aeration rate is needed to maintain the aerobic
biological degradation process and at the same time keeping the composting mass
warm enough for thermophilic microorganisms and effective pathogen destruction.
Wiley and Pearce (1955) studied the effect of different aeration rates in continuously
mixed reactors containing approximately 16 kg of garbage mixture with a moisture
content of 52 to 58%. The effect of aeration was evaluated from the temperature
curve during composting. Low aeration rates of 0.17 to 0.28 L/min/kg VS resulted in
a late peak temperature. Medium aeration rates of 0.39 to 1.26 L/min/kg VS
provided a peak temperature after four days and after that the temperature declined
slowly. High aeration rates of 1.41 to 3.35 L/min/kg VS also provided a peak
temperature after four days but after that the temperature declined rapidly. In another
study (Composting 1985) on sludge composting it is suggested that an aeration rate
of 0.30 to 0.83 L/min/kg dry solids, provided on an intermittent basis, results in
oxygen levels from 5 to 15% throughout the pile.

To control the aeration rate a number of strategies have been designed and practiced
for the composting process. These range from simple manual control systems to
more sophisticated computer control system using temperature, oxygen or carbon
dioxide feedback as the controlling variables and air supply as the manipulated
variable (Stentiford and Neto 1985, Haug 1993). However, the control system should
be able to satisfy peak, average and minimum aeration requirements.

2.6.7 Interrelated Effects of The Factors

In the composting process the effects of the key factors, namely temperature,
moisture content, and aeration are interrelated. Schulze (1960) conducted
composting experiments with synthetic garbage in a small drum. An average of 40%
of the volatile matter was oxidized during 14 days of active decomposition phase.
The initial and final percent ash was low in all experiments, due to the fact that the
garbage mixture used consisted of 90 to 96% volatile solids. The C/N ratio decreased

23
from an average of 28.6 to 19.9. The oxygen consumption rates varied directly with
temperature between 27 and 63 oC, with values ranging from 1 to 5 mg O2 per gram
initial volatile matter per hour. The temperature coefficient Q10 of the oxidation
reaction was found to be 1.9. It was suggested that to maintain aerobic conditions
during peak oxygen demands, an air supply of 0.38 to 0.43 L/min.kg VS (560 to 620
m3 per ton of initial volatile material per day) is necessary. Respiratory quotients for
the overall process varied between 0.8 and 0.9.

In another experiment, Schulze (1962a) studied the relationship between temperature


and O2 consumption rates during periodic feeding composting of mixed organic
waste materials in a 208 L (55 gal) intermittent rotating drum. The temperature of
organic waste material was 53 to 70 oC. The loss of volatile matter through the
process was 37 to 45%. The average detention time was 7 to 18 days, however the
material removed from the drum consisted of a mixture of particles with residence
times ranging from a minimum of one day to a maximum of weeks or months. He
plotted O2 consumption rates on a log scale versus temperature and found a linear
relation according to the following equation:

y = a10kT
or y = 0.1*100.028T
where, y = O2 consumption rate, mg O2/gm volatile matter.hr
a = constant
T = Temperature, oC

The reactors used by Schulze were intermittently mixed and operated as a batch or a
periodic feeding continuous process. However it remains necessary to find the
relationship between temperature and O2 consumption rates for a temperature feed-
back control aeration system with different aeration modes. Furthermore the question
of reaction order should be addressed.

Jeris and Regan (1973a) studied different environmental parameters for optimum
composting of mixed refuse using three types of experimental setup. Those were

24
bench-scale composter (42 L volume, charged with 1 to 3 kg waste mixture), shake
flasks (500 mL each) and Warburg respirometer (125 mL each). They found
optimum oxygen consumption rates of 0.06 to 0.08 L/min.kg-VS (3.7 to 4.5 mmole
O2/day-gVS) within the temperature range of 50 to 64 oC. They also found that
newsprint and stabilized municipal refuse have very similar and low degradation
rates of about 10% of the rate obtained for mixed refuse.

Campbell and Darbyshire (1990a) indicate that, during composting of tree bark,
aeration and moisture content have a direct effect on the composting process and that
a 60% moisture content was optimal for decomposition. Forced aeration was found
essential only in the initial rapid self-heating phase in their small reactors of 26 L
which had no temperature control. In another study (Campbell and Darbyshire
1990b), they found that high temperature could be maintained for long periods in an
adiabatic reactor (of 16 L) whereby the temperature could be controlled by aeration.
Although the reactors used by Schulze and Campbell and Darbyshire do not
represent a process with full height of composting mass, they are very useful for
determining the optimal physico-chemical parameters.

A forced aeration composting process was investigated by Epstein et al. (1976).


Their pile dimension was 12x6x2.5 m high. Suction mode aeration was applied for
the first 16-20 days and after that air was blown into the pile for another 8-10 days
by centrifugal blower. The exhaust gas in the suction mode aeration was deodorized
by passing it through a biofilter. They concluded that either digested or raw sludge
can be composted in an aerated pile without releasing objectionable odors.
Destruction of total coliforms, fecal coliforms, and salmonellae was much greater in
forced aeration composting than in windrow composting.

Willson (1983) conducted similar pilot tests using downflow mode aeration. Six
different fixed aeration rates were applied in six small compost piles. He found that
aeration rates can vary over a wide range with little effect on developing temperature
essential for destruction of pathogenic organisms and that the higher rates remove
more moisture. During most of the composting period, the average temperature

25
among the piles ranged from 60 to 70 oC. Although Epstein et al. (1976) and Willson
(1993) used forced aeration, they did not control the process temperature.

Darbyshire et al. (1989) conducted forced aeration (upflow) composting for the
decomposition of coniferous bark in a static pile of 14 m3 volume and 1.2 m height.
They indicate that the optimum temperature was below 60 oC when comparing the
piles with controlled and non-controlled temperature via forced aeration. However,
controlling the temperature did not prolong the length of the thermogenic phase
above 40 oC in their pile. A positive temperature gradient along the direction of air
flow was observed. Temperature gradients along the air flow direction for both
upflow and downflow aeration in pilot-scale reactors and in static piles were also
observed by Koenig and Bari (1998) and Stentiford et al. (1985b), respectively.

Fernandes and Sartaj (1997) and Sartaj et al. (1997) compared natural, forced
(upflow) and passive aeration methods in different static piles. The compost piles
had an initial volume of 5 m3, were 1.2 m in height and were monitored over a
period of three months. They found that the average temperature inside the passively
and naturally aerated piles stayed above 55 oC for a longer period than in the forced
aerated pile. Conversely, forced aeration methods were more effective in providing
oxygen. Shorter times to reach stability were achieved by using forced and passive
aeration as compared to natural aeration. They also observed non-homogeneous
temperature distribution inside the piles of all types with a higher temperature zone
of 60 to 65 oC at the middle and a lower temperature zone of 20 to 25 oC near the
surface.

In another study (Garcia et al. 1996), temperature and moisture gradients were
observed in a small reactor of 15 L volume. Aeration was controlled at and below
55 oC by temperature feed-back and timer, respectively. Two types of waste
mixtures with initial moisture content of 60% were composted for 120 and 220
hours, respectively. The observed temperature gradient was 6.5 to 8 oC/10 cm
(calculated from Figures 1 and 2 of their paper). The moisture content in the upper
layers were 75 to 71% and in the lower layers 56 to 46%, respectively. Hogan et al.

26
(1989) reported temperature gradients of 6.4 oC and 8.4 oC per 10 cm in insulated
closed reactors of 45 cm height with and without heat conduction control,
respectively.

Hofer and Kugler (1995) conducted aeration experiments on composting in closed


reactors of 20 m3 capacity. During the 5 to 8 weeks of operation, the reactors were
emptied every week for re-mixing and re-moisturing the waste mixture. Upflow
mode aeration with air recirculation was applied for the first 4 weeks followed by
downflow mode without recirculation. However, no special effect of air recirculation
on composting was presented. Furthermore, aeration was controlled only for the
upflow aeration period, but the temperature in the top layers was mostly 70 oC.
During the downflow aeration period the temperature in the bottom layers was
mostly above 80 oC. During both aeration modes they observed temperature
gradients along the air flow directions.

Pilot-scale experiments were performed by VanderGheynst et al. (1997) using four


different fixed aeration rates and two initial moisture contents. The reactors had a
volume of 770 L with a height of 2.65 m. They used only upflow aeration mode
with no temperature control. In all experiments the maximum temperatures of 60 to
74 oC and the positions where they occurred varied with initial moisture content and
aeration rate. Their work with replicate tests showed that these pilot-scale reactors
were capable of providing reproducible results.

Since the degradation rates of a given organic substrate in the composting process
depend on temperature, moisture and on other physico-chemical parameters, forced
aeration and control of air flow rate is applied for maintaining high degradation rates
at the optimum conditions. As discussed above the main problem of forced aeration
composting is the non-homogeneous distribution of temperature, moisture and other
physico-chemical parameters over the height of the composting mass. For example,
if upflow aeration mode is applied, the temperature of the bottom layers remains
close to ambient temperature while high temperature is observed in the top layers,
and vice-versa in the downflow aeration mode. Furthermore, the measured

27
temperatures or other parameters in large-scale composting plants are usually not
representative for the whole composting mass and vary according to aeration mode
as well as location of measurement points in the composting mass.

2.7 Product Stability

The main objective of composting is to produce a biologically stable or mature and


pathogen free humus-like end product i. e. compost, which can be beneficially used
as a soil conditioner or for other purposes. The biological stability or maturity of
compost affects its successful utilization in agriculture. A compost with a very low
decomposition rate, after a thermophilic phase where easily decomposable organic
compounds are completely oxidized, can be described as stable. It should be noted
that a strict distinction between the terms stability and maturity is not possible and
they are often used interchangeable. If the compost is not sufficiently stabilized, it
will cause subsequent problems in storage, transportation, utilization and in the final
disposal site. Instable compost emits offensive odor, decreases plant growth due to
the presence of phytotoxic substances (basically ammonia, ethylene oxide and
organic acids, etc.). Jimenez and Garcia (1989, 1992a) reported that the rapid
decomposition of an immature compost may cause a decrease of the O2
concentration in the soil and as a result the creation of an anaerobic and strongly-
reducing environment at the level of the root system. This causes an increase of the
solubility of heavy metals in the soil and inhibition of plant seed germination by the
production of phytotoxic substances. Zucconi et al. (1981) noted that the
introduction of decomposing organic material in soil may damage plant roots.
Manios et al. (1989) indicated the inhibitory effect of phytotoxicity on the
germination of lettuce seeds.

A number of physical, biological and chemical methods have been suggested and
evaluated to measure the degree of stability of compost. Among those, the self-
heating test, temperature decline, color and odor, microbial respiration, seed
germination and plant growth, C/N ratio, pH and other chemical tests are
significant. Although none of these tests has been widely accepted (Willson and

28
Dalmat 1986), some of them are very useful. It has been suggested that more than
one parameter is required to determine compost maturity (Inbar et al. 1990, Hsu and
Lo 1998). Generally, three approaches, namely metabolic activity, respiration and re-
heating potential as measured by carbon dioxide evolution, oxygen uptake and the
self-heating test, respectively, are suggested to measure compost stability (Brinton et
al. 1995, Feldman 1995, Switzenbaum et al. 1997a, Switzenbaum et al. 1997b).
Willson and Dalmat (1986) noted that, since the respiration rate of aerobic
microorganisms is proportional to their heat production, carbon dioxide production
or their oxygen consumption, measurement of any of these parameters would be an
indication of stability. The tests which are extensively applied in this research are
discussed separately in the following sections.

2.7.1 Self-heating Test

The self-heating test and its procedure were developed in Germany (LAGA 1985,
BGK 1994). It is a simple test and an excellent indicator of biological stability, if
properly conducted (Koenig 1997), and does not require sophisticated equipment
(Zimmerman 1991). Brinton et al. (1995) reported that the self-heating test integrates
a number of factors present in normal composts such as temperature, aeration, etc.
and therefore may provide data that correlate well with field observations about
compost behavior. In the self-heating test, suitably prepared waste samples of
optimally adjusted moisture content are loosely filled into Dewar bottles (volume =
1.5 L, inner diameter = 100 mm) open to the atmosphere. A temperature sensor is
inserted for monitoring. The bottles are then kept at room temperature of
approximately 23 oC. If the waste is not yet biologically stable, it will further
degrade aerobically, generating heat which will cause a temperature rise. Usually,
the maximum temperature Tmax is reached after two to five days. The test ends after
Tmax is culminated and rapidly declining temperatures are observed, at the latest after
ten days. Tmax attained is used as an indicator of biological stability to define the
stability index SI. The SI of waste is classified in degrees from I to V, in ascending
order of biological stability, and ranges from raw, unstabilized waste (SI = I) to
completely stabilized waste (SI = V). The temperature curves of typical self-heating

29
tests for fresh compost (SI = I) and mature compost (SI = V) are shown in Figure
2.3, with Tmax, A72 (area under the temperature curve after 72 hours) and Imax
(maximum temperature increasing rate) indicated in the graph.

80
Imax= y/x
70 Fresh compost
Tmax Mature compost
x
60 Ambient temperature (Ta)
Temperature in oC

y
50

40
T-Ta
30

20

10 A72

0
0 4 8 12 16 20
Time in days
Figure 2.3 The temperature curve of a typical self-heating test for fresh
compost, matured compost, Tmax, A72 and Imax are shown.

Tmax also corresponds approximately with the area under the temperature curve A72
(after 72 hours). Based on the result of the self-heating test, the biological stability of
waste and compost can be classified as shown in Table 2.2 (LAGA 1985).

Table 2.2. Classification of waste according to degree of biological stability


Tmax, in oC 20-30 30-40 40-50 50-60 60-70 >70
Imax, in oC/h <0.3 0.3-0.45 0.45-0.8 0.8-1.4 1.4-2.0 >2.0
A72, in oC.h <1700 1700- 2000- 2500- 3000- >3500
2000 2500 3000 3500
RAa in mg O2/g VS-h 0-0.5 0.5-1.0 1.0-1.5 1.5-2.0 >2.0
Degree of V IV III I II I
biological stability stable, Un-
(Stability index SI) mature stable
a
Imax = maximum temperature increase; RA = respirometric acitivity; Based on
Iannotti et al. (1994) as cited by Epstein (1997)

30
2.7.2 Temperature Decline

The final drop in temperature after the thermophilic phase to or near ambient, may
be considered as a measure of end product stability. Golueke (1977) stated that once
the temperature has dropped to 40-45 oC, the material can be stored indefinitely
without causing problems. Kapetanios et al. (1993) considered temperature drop
from the thermophilic phase to 25-30 oC after 90 days as an indicator of maturity of
compost in a turned windrow. Jimenez and Garcia (1989) reported that a compost is
mature enough when its temperature remains more or less constant and does not vary
with the turning-over of the material. However, it should be assured that a
temperature drop due to the influence of other factors such as moisture and oxygen
deficiency, inactivation of the microorganisms at high temperature, etc. is excluded.
Sometimes the re-heating potential of the end product in a container or a non-aerated
pile is considered as a measure of stability by researchers (Schulze 1960, Fernandes
and Sartaj 1997).

2.7.3 Color and Odor

The color of the final product after effective decomposition should be dark brown or
almost black. Gray et al. (1973) reported that the product from farm and garden
wastes is gray-black or brown-black in color. Brown color is one of the
specifications of final compost produced from wastewater sludge of some projects in
the USA (USEPA 1989). Generally the offensive odor of the waste material
decreases during the first stages of decomposition and almost disappears by the end
of the composting process. Jimenez and Garcia (1989) noted that a characteristic
odor similar to that of damp forest ground should be detected. However, evaluation
of stability based on color and odor mostly depends on the experience of the
operator.

2.7.4 C/N Ratio

The C/N ratio is often used as an index of compost maturity. Jimenez and Garcia
(1992b) noted that the C/N ratio is one of the most important parameters in the
31
control of the composting process and in the determination of the degree of maturity
of the newly formed organic materials. A high C/N ratio in immature compost causes
nitrogen immobilization in the soil and affects plant growth. According to Inbar et
al. (1990), a C/N ratio below 20 is indicative of an acceptable maturity in the final
product, a ratio of 15 or even less being preferable. Sometimes this value may be
above 20 in relatively mature compost since part of the organic carbon may be in the
form of compounds more resistant to bio-degradation and not immediately available
to microorganisms (Jeris and Regan 1973a, Jimenez and Garcia 1989). Viel et al.
(1987) performed a composting test of a mixture of agricultural industrial waste and
saw dust in a 100 L reactor. They showed that the C/N ratio of 34 to 35 in the initial
mixture decreased to 20 to 22 after 120 days of composting. However, the final C/N
depends on the initial composition of the waste to be composted.

Another indicative criterion of compost stability as proposed by Morel et al. (1985)


is the ratio:
final( C / N )
initial(C / N )

Jimenez and Garcia (1989) suggest that a reasonable estimation of this ratio could be
less than 0.75 in compost of more than 120 days, preferably of the order of 0.60
(compost of about 180 days).

2.7.5 pH

The pH is one of the indicators of the progress of the composting process. As


mentioned previously, at the beginning of composting, the formation of organic
acids and carbon dioxide lower the pH value to approximately 5 or less, whereas
with further progress the pH value reaches up to 8 to 8.5. The pH of the final product
considerably depends on the type of initial waste mixture. The pH ranges reported
for finished compost produced from municipal solid waste were 7.0 to 7.5 (Canet
and Pomares 1995), for sewage sludge 5 to 9 (Kuter 1995), mixed garden waste
compost 7.5 to 7.8 (Keeling et al. 1996), and for manures 5 to 6 (Lau et al. 1992,

32
Gies 1995). However, for plant growth, pH levels between 5 and 8 are acceptable
(Kuter 1995, Manser and Keeling 1996).

2.8 Odor Control Using Biofilter

Emission of odor is an important concern in composting facilities. As mentioned in


Chapter 1, several solid waste composting facilities have been shut down because of
odors. Many factors including type of substrate and operational procedure are
responsible for emission of odor in a composting facility. The major reported odor
groups in composting include: fatty acids, ammonia and other nitrogen containing
compounds, ketones, aromatics, and inorganic and organic sulfur compounds
(Williams and Miller 1992a, Epstein 1997).

Biofilters have been used for the treatment of odors and volatile organic compounds
in exhaust gases collected from wastewater treatment facilities and composting
operations (Metcalf and Eddy 1991, Haug 1993, Hodge and Devinny 1995).
However, this technology is still under development (Toffey 1997). Biofiltration
means that a odorous airflow is passing through a layer of filter material (compost,
filamentous peat, etc.), followed by biodegradation of the captured odor components
(Van Der Hoek and Qosthoek 1985).

This process offers a cost-effective solution for the treatment of large volumetric
airstreams containing low levels of highly odorous compounds (Williams and Miller
1992a, Deshusses et al. 1995). Yang and Allen (1994) indicated that the biofiltration
process was effective for low level H2S emissions with 99.9% removal efficiency.
Torres et al. (1997) reported that biofilters removed more than 99% H2S and 70%
volatile organic compounds. More than 99% removal of various odorous compounds
through biofilter are presented by Fischer et al. (1990). Therefore, in most cases the
odor removal efficiency was found to be very high. The recommended design and
operating parameters for biofilter are presented in Table 2.3.

33
Table 2.3. Recommended design and operating parameters for compost
biofilter. Modified from Haug (1993)
Filter media biologically stable compost, peat, granular activated carbon,
soil a,b,c
Media depth 1 - 1.5 m b,c
Air distribution uniform manifold with a base of rock
Moisture content 50 to 70%
pH 6.5 to 7.5 d
Temperature near ambient, or 15 to 45 oC
Air loading rate < 100 m3/h-m2, unless pilot testing supports higher loading
Air residence time 30 to 60 sec, unless pilot testing supports a shorter time
a
Ottengraf 1986; Hodge and Devinny 1995; cTorres et al. 1997; dToffey 1997
b

2.9 Utilization of Compost

Beneficial uses of compost are increasing mainly in the following fields: in


landscaping as a soil conditioner, in agriculture as a organic fertilizer and in
horticulture as a growing media (USEPA 1992, Nieveen 1997, Wong et al. 1999). A
number of benefits can be achieved by applying compost to the land. The organic
matter in the compost improves the physical properties of soil, including enhanced
aggregation, increased soil aeration, lower bulk density, less surface crusting,
increased moisture holding capacity and improved infiltration (Kuter 1995), which
are necessary for crop production. Haug (1993) stated that compost contains
valuable nutrients including nitrogen, phosphorus, and a variety of essential trace
elements. It can improve the growth and vigor of crops in commercial agriculture
and home related uses. Manser and Keeling (1996) reported that compost, by
improving soil structure, increases the ability of soil to resist erosion from water and
wind. Other uses of compost include: (i) daily cover for landfills (Frangipane and
Vismara 1996); (ii) top dressing on golf courses and lawns, land reclamation, high-
way shouldering (USEPA 1989); and (iii) media for biofilter (Williams and Miller
1992b). Compost is also used as a filtering media for treating stormwater runoff
which carries pollutants such as sediments, solubilized metals, oil and grease, and
nutrients (Conrad 1995).

34
CHAPTER 3

THEORETICAL CONSIDERATIONS

3.1 Reaction Rates


In order to describe the changes that occur in the waste material during composting,
and the extent of these changes, it is important to know the reaction rates or
degradation rates. The reaction rates are significant for designing or optimizing a
composting process because they directly affect the processing time, size of the
reactor or pile area and the product quality. A high degradation rate indicates lower
capital and operational cost for a composting plant. Finstein and Miller (1985) noted
that, for any given processing duration, the higher the rate the more stable and easily
handled the residue and this facilitates storage, transport, and final disposal with a
minimal cost. The rate of water vaporization is directly related to the rate of waste
decomposition and water removal is an objective in sludge treatment because dry
residue is more easily managed (Finstein et al. 1986b). The elevated temperature
from 55 to 60 oC for a longer period is vital in sanitization and this can be
maintained by the higher degradation rates. Rapid decomposition of the putrescible
material may help to control odor at the composting site. In addition, by using the
reaction rates a simple model to predict the extent of degradation could be worked
out.

The reactions which occur in a composting process may be considered as a change in


the quantity of biodegradable volatile solids (BVS). The daily BVS change can be
determined by monitoring daily oxygen consumption, since this is proportional to
the BVS degradation rate. The proportionality constant or ratio of kg oxygen

35
consumed per kg BVS degraded would be different for different types of wastes.
Haug (1993) reported that the stoichiometric demand for oxygen varies from a low
of about 1 kg O2/kg organic matter for highly oxygenated substrates such as starch
and cellulose to a high of about 4 kg O2/kg organic matter for saturated
hydrocarbons. Koenig and Tao (1996) indicated that an average of 1.2 kg O2 were
consumed per kg of VS degraded for a mixture of canteen wastes and waste paper.
However, the O2 consumption rate correlates with the BVS degradation rate.

The general equation to define the rate of three different types of reaction, namely,
zero, first, and second order is,

dBVS
= kBVS n (3.1)
dt

where BVS = biodegradable volatile solids, in kg


t = time, in days
dBVS
= rate of change in mass of BVS (the negative sign indicates
dt
that the quantity of BVS decreases with time)
n = reaction order, 0, 1, or 2
k = rate constant, units depend on reaction order

3.1.1 Zero Order Reaction


Zero order reactions proceed at a rate which is independent of the quantity of BVS
(biodegradable volatile solids). From Equation 3.1 for zero order reaction,
substituting n = 0,
dBVS
= kBVS 0 = k (3.2)
dt
where k = zero order rate constant, in kg.day-1
Integrating equation 3.2 and substituting for BVS = BVS0 at t = 0, it follows that
BVSt = BVS0 - kt (3.3a)
where BVS0 = initial BVS at t = 0
BVSt = remaining BVS after time t
If k is not constant over time, Equation 3.2 can be solved numerically as

36
BVSt = BVS0 - (k1 + k2 + .......... kn)t (3.3b)
where k1, k2, .......... kn = the reaction rate constant for time intervals t1, t2,
.......... tn, respectively and t1 = t2 = .......... = tn = t
Therefore, the percentage degradation of BVS after any time t is
BVS 0 BVS t
* 100
BVS 0

kt
= * 100 or (3.4a)
BVS 0

=
( k 1 + k 2 +LL k n )t * 100 (3.4b)
BVS 0

3.1.2 First Order Reaction


First order reactions proceed at a rate which is proportional to the quantity of
remaining BVS. From Equation 3.1 for first order reaction, substituting n = 1,
dBVS
= kBVS1 = kBVS (3.5)
dt
where k = first order rate constant, in day-1
Integrating Equation 3.5 and substituting for BVS = BVS0 at t = 0, it follows that

BVS t = BVS 0 e kt (3.6a)


If k is not constant over time, the numerical solution is

BVS t = BVS 0 . e k 1 t . e k 2 t LLe k n t or (3.6b)

BVS t = BVS 0 . e ( 1 2
k + k +LL k n ) t

It can be shown that for small exponents e-k1t is approximately (1- k1t)
or e-kt = (1 - k1t)(1 - k2t) .......... (1- knt)
Therefore, the percentage degradation of BVS after time t is
BVS 0 BVS t
* 100
BVS 0

(
= 100 1 e kt ) or (3.7a)

= 100[1 - (1 - k1t)(1 - k2t) .......... (1-knt)] (3.7b)

37
3.1.3 Second Order Reaction
Second order reactions proceed at a rate which is proportional to the second power
of the quantity of remaining BVS. From Equation 3.1 for second order reaction,
substituting n = 2,
dBVS
= kBVS 2 (3.8)
dt
where k = second order rate constant, in kg-1.day-1
Integrating Equation 3.8 and substituting for BVS = BVS0 at t = 0, it follows that
BVS 0
BVS t = (3.9a)
1 + ktBVS 0

If k is not constant over time, from Equation 3.9a, for a small interval of time a
numerical solution can be derived as follows:
BVS 0
BVS1 =
1 + k 1BVS 0 t
BVS1
BVS 2 =
1 + k 2 BVS1t

BVS 0
or BVS 2 =
1 + ( k 1 + k 2 ) BVS 0 t

BVS 0
similarly BVS t = (3.9b)
1 + ( k 1 + k 2 + k 3 +LL k n ) BVS 0 t

Therefore, the percentage degradation of BVS after time t is


BVS 0 BVS t
* 100
BVS 0
ktBVS 0
= * 100 (3.10a)
1 + ktBVS 0

=
( k 1 + k 2 + k 3 +LL k n )BVS 0 t * 100 (3.10b)
1 + ( k 1 + k 2 + k 3 +LL k n ) BVS 0 t

3.2 Effect of Temperature on Reaction Rates


The rate of biological reactions increases with temperature up to a limited range,
suitable for microorganisms. Above this range of temperature the activity of
enzymes, responsible for mediating the biological reaction, decreases due to enzyme

38
denaturation. A frequently quoted rule of thumb known as the vant Hoff rule states
that the reaction rate doubles for a 10 oC temperature rise. Equation 3.11 proposed
by the Swedish chemist Arrhenius, has been used to estimate the effect of
temperature over a limited range for biological reactions. The Arrhenius equation is:
d(ln k ) Ea
= (3.11)
dT RT 2
where k = reaction rate constant
Ea = activation energy, kJ/mol
R = ideal gas constant, 8.314*10-3 kJ/mol
T = absolute temperature, oK
Equation 3.11 can be integrated to give the expression
Ea 1
ln k = ln C (3.12)
R T

where C = frequency factor, a constant


From Equation 3.12, it can be shown that the plot of ln k versus 1/T will be a straight
line. Equation 3.12 can be formulated for practical application as
Ea
ln C
RT
k=e (3.13)
Ea

RT
or k = C. e (3.14)
The reaction rate constant k at any temperature can be obtained by Equation 3.14.

39
CHAPTER 4

METHODOLOGY

4.1 Scope of Study

As discussed in previous chapters, recent applications of large-scale composting


have often been plagued by (i) technological problems, such as vertical temperature
and moisture gradients and biological product stability; (ii) lack of understanding of
the biological fundamentals and energy relationships; and (iii) environmental
emissions such as odor and leachates. In this study therefore, the experimental
techniques and methods were so designed, based on the suggested optimum
conditions of different physical and chemical factors as discussed in Chapter 2, as to
improve the distribution of temperature, moisture and other physico-chemical
parameters throughout the height of the composting mass in a closed system and
thereby increase the rate and extent of organic waste degradation and minimise the
environmental emissions. To this purpose, the effects of different operating
conditions on the composting process in a closed system were investigated, with
special emphasis on the effect of various modes of forced aeration on:
1. vertical distribution of physico-chemical parameters including temperature and
moisture in the composting mass.
2. rate and extent of degradation of organic wastes during composting.
3. biological product stability.

The effect of the following selected forced aeration modes were studied using two
heat insulated closed pilot-scale reactors:
1. Upflow aeration (blowing or positive pressure)
40
2. Downflow aeration (suction or negative pressure)
3. Alternate upflow/downflow aeration
4. Internal air recirulation in a single-reactor system
5. Reuse of spent air in a two reactor-system

To accomplish the above objectives, extensive pilot-scale and bench-scale


composting tests were conducted with similar mixtures of organic solid wastes. A
number of self-heating tests were also conducted to assess the biological stability of
compost produced in each pilot-scale test. An attempt was also made, based on the
theories described in Chapter 3, to develop a simple mathematical model to simulate
and predict the extent of organic waste degradation.

All together 18 pilot-scale tests were performed, either as a first, second or third
stage composting. Nine tests were performed as a first stage composting of mostly
20 to 30 days duration each, using a single reactor system with or without internal air
recirculation. Eight of the first stage tests were follow by their corresponding second
stage composting tests of also 20 to 30 days duration each, using either a single
reactor or a two-reactor system in series. The remaining test was conducted as a third
stage composting test of 11 days duration following its corresponding second stage
composting test. The first stage composting tests were conducted using fresh organic
solid waste mixture, whereas the second or third stage composting tests were
conducted using the fresh compost produced in the corresponding first or second
stage composting tests. After first or second stage composting the product was re-
mixed and adjusted with water, if necessary, for the second or third stage
composting.

All pilot-scale tests were conducted using two specially designed heat insulated
closed reactors. The pilot-scale reactors were so designed that they represent the
central core of an aerated static pile or a closed composting system. It was assumed
that variations of all physico-chemical parameters, including temperature and
moisture along the radial horizontal directions were minimum and along the vertical
direction were maximum. Therefore, the pilot-scale reactors with a full height of
composting mass of 1.7 m were capable of simulating the variations of most
41
parameters as they varied in the field-scale composting process along the vertical
direction. The rates of aeration were also controlled, for an optimum temperature of
55 to 60 oC, via a temperature feed-back control system.

A series of self-heating tests were conducted to determine the product stability and
the amount of remaining BVS at the end of each stage of the pilot-scale tests. A total
of 16 sets of self-heating tests were conducted, with each set using a series of 6
vacuum flasks of 1 L volume. These 6 vacuum flasks were used for the individual
samples collected separately from six different layers of the composting mass after
the pilot-scale tests.

Four runs of bench-scale tests were conducted under different conditions. The
objective of run 1 was to compare the effect of two aeration conditions such as
natural and forced aeration in the bench-scale test. The same waste mixture was used
for four bench-scale tests of run 1. The objective of runs 2 (two tests) and 3 (two
tests) was to compare the results with the corresponding first and second stage pilot-
scale tests. The objective of run 4 was to investigate the effect of initial moisture
content on VS degradation and overall temperature. Twelve bench-scale tests were
conducted under run 4, using the same waste mixtures but of various moisture
content. Vacuum flasks of 1 L volume were used as reactors for the bench-scale
tests.

4.2 Experimental Set-up

4.2.1 Pilot-scale Reactor

Two batch reactors were made using 10 mm thick Perspex acrylic sheet. Each
reactor had a total volume of about 200 L, with a height of 2.2 m and a cross-
sectional area of 0.09 m2 (0.3 x 0.3 m). The reactors were insulated with a 0.1 m
thick layer of polyurethane on all sides. A perforated false floor of stainless steel was
situated inside the reactor at 0.5 m from the bottom to facilitate the air distribution
and on which the waste material was placed. The waste mixture was put into the

42
reactor from the top, which could be easily opened. After the end of the test the
compost was removed through an opening of 0.3 x 0.3 m located above the false
floor and placed at one side of the reactor, which remained closed during the test.
Thermocouples for measurement of temperature during the composting tests were
placed inside the reactor at different heights of the composting mass: at 10 cm, 30
cm, 50 cm, 70 cm, 90 cm, and 110 cm, respectively, starting from the bottom.

The schematic diagrams of the experimental pilot scale reactor as a single-reactor


and two-reactor composting system are shown in Figures 4.1 and 4.2, respectively.
The pipe and valve arrangements for upflow, downflow, and internal air
recirculation in the single reactor system are illustrated in Figure 4.1. For simplicity,
only the pipe arrangement in the upflow aeration mode is shown for the two-reactor
system in Figure 4.2. The reactors were fixed on separate moveable steel frames
which facilitated work within a small place in the laboratory, but are not shown in
the figures.

Leachate was stored in diluted form at the bottom of the reactors in an intermediate
storage tank and intermittently recycled to the top of the column. The leachate was
collected at the end of each test from the bottom of the reactor through a valve.
Actually tap water was stored in the leachate storage tank for making up moisture
losses of the composting mass and this water diluted the leachate. The diluted
leachate was recycled by a recycling pump (Sanso Circulating Pump, model no.
PMD-411F, Sanso Electric Co. Ltd., Japan) and was uniformly distributed by a
sprayer.

Forced aeration was provided by a central compressor (maximum pressure 90 psi,


Faculty of Engineering, The University of Hong Kong), with an option for
continuous or intermittent aeration. The pressure was reduced by using an air valve
and an air regulator (SMC, model no. AF M3000-02, Japan). Continuous or
intermittent aeration was performed using a control valve

43
Temperature

Relative Humidity Sensor

Downflow Spent Air

O2 Measuring
Point

Compost-
Monitoring ing
Mass
Data
Clean
Storage Insulation
Air
Process
Control
Internal
Air Air Air
Pressure Velocity Recycling
Pump
Fresh Air

Air Blower Upflow


Leachate
Fresh water makeup

Leachate overflow Pump

Reactor Biofilter

Figure 4.1 Schematic diagram of an experimental pilot-scale composting reactor and


biofilter. Upflow, downflow and internal air recirculation pipe lines are shown.

44
Temperature

Relative Humidity Sensor O2 Measuring O2 Measuring


Point Point
Temperature
Used
Hot Air

Spent Air

Insulation

Monitoring

Data Clean
Storage Compost- Air
ing
Process Mass
Control

Air Air
Velocity Pressure

Fresh Air
Leachate
Recycling
Air Blower
Pump
Leachate

1st Reactor 2nd Reactor Biofilter

Figure 4.2 Two-reactor composting system with reuse of spent air in the upflow
aeration mode and biofilter.

45
(AB31-02-3, CKD Corporation Japan), which was controlled by the datalogger
software program depending on the monitored temperature. After leaving the
reactor, the hot spent air cools down naturally in a condenser and the condensates
were collected regularly. Finally, prior to discharge to the atmosphere, the spent air
was filtered through a biofilter as shown in Figures 4.1 and 4.2. The total height of
the biofilter was 1.3 m with a cross-sectional area 0.09 m2. The filter media was
made of a lower layer (0.2 m) of gravel and an upper layer (0.9 m) of mature
compost of optimum moisture content. The excess condensate was collected from
the bottom valve. Water could be added from the opening located at the top of the
biofilter to prevent dryness of the media.

The pipes for air recirculation or reuse were insulated with a 5 cm thick layer of
Fanyalon to prevent heat loss and minimize condensation. Internal air recirculation
was performed by an air recirculation pump (model no. 7530-65, Cole-Parmer
Instrument CO. USA). Internal air recirculation could be varied according to the air
recirculation ratio r which is defined as the ratio of recirculated air over fresh air
input, in percent. The flow of reused air was equal to the flow of spent air from first
stage composting, i.e. approximately equal to fresh air input.

4.2.2 Bench-scale Reactor

Bench-scale tests were conducted using vacuum flasks of 1 L volume (Thermos


Limited, England) as a bench-scale reactor. About 400 g of waste mixture are
necessary to fill each of the bench-scale reactors. The waste inside the reactor was
compacted loosely to provide proper porosity for natural and forced aeration. After
putting the waste mixture into the reactor, the inlet was covered by packing foam to
prevent excessive heat loss. The temperature of the waste mixture was continuously
monitored through thermocouple by a datalogger. Aeration was provided by a small
air pump on a timer schedule basis of 15 minutes aeration followed by 15 minutes
pause. For the self-heating test only natural aeration was allowed. After final
declination in temperature close to ambient, the tests were discontinued. Figure 4.3
shows the experimental setup.

46
Thermocouple

Aeration pipe

Compost

Computer Datalogger Aerator Vacuum flask

Figure 4.3 Experimental setup of the bench-scale test. A similar setup was used for
self-heating tests for determination of biological stability of compost.

4.3 Experimental Program and Aeration Modes

4.3.1 Pilot-scale Tests

All together eighteen pilot-scale composting tests were performed as a first, second
or third stage, using either a single reactor or a two-reactor system in series, under
different modes of aeration. The modes of aeration studied were upflow, downflow,
alternate upflow/downflow, and internal recirculation in a single-reactor system as
well as reuse of spent air in a two-reactor system in series. The composting stages,
reactor systems, aeration modes and duration of the different tests conducted are
summarized in Table 4.1.

To achieve the desired temperature and oxygen levels during active composting,
variation of aeration rates was considered as one of the key process control methods.
In this research, aeration was controlled by temperature feed-back, because of ease
of measurement and reliability as an indicator of the composting process (Stentiford
et al. 1985a, 1985b).

47
Table 4.1. Experimental program and setup of pilot-scale composting tests
Test Stage Reactor Aeration mode Duration
name system days
1a 1 SR Upflow, internal recirculation (r = 20-50%) 31
1b 2 SR Upflow/downflow 21 (52)*
2a 1 SR Downflow 33 (33)
3a 1 SR Upflow, internal recirculation (r =100- 29
200%)
3b 2 TR Upflow/downflow for 17 days, reuse of air 39 (68)
from stage 1 (4a)
4a 1 SR Upflow followed by downflow after 28 days 49
4b 2 SR Upflow/downflow alternatively every 3-4 21 (70)
days
5a 1 SR Upflow, internal recirculation (r = 20-80%) 21
5b 2 TR Upflow/down flow for 9 days, reuse of air 20
from stage 1 (6a)
5c 3 SR Upflow/downflow alternatively every 3-4 11 (52)
days
6a 1 SR Upflow/downflow, aeration mode changed 30
irregularly 4 times over 30 days
6b 2 SR Downflow 21 (51)
7a 1 SR Upflow/downflow alternatively every 3-4 23
days
7b 2 TR Upflow flow for 1 day, reuse of air from 32 (55)
stage 1 (8a)
8a 1 SR Upflow 28
8b 2 TR Upflow/downflow for 10 days, reuse of air 28 (56)
from stage 1 (9a)
9a 1 SR Downflow 24
9b 2 SR Upflow 32 (56)
SR = Single-reactor system, TR = Two-reactor system, r = recirculation ratio
a = first stage, b = second stage, c = third stage.
*
Total duration including first, second and third stages

Based on previous composting tests by Koenig and Tao (1996), using the same pilot-
scale reactor and similar organic waste, the range of air flow rate was set at 17 to 20
L/min (equivalent to about 0.67 L/min per kg initial VS) during the first two days.
This was followed by an air flow rate of 8 to 10 L/min (equivalent to about 0.33
L/min per kg initial VS) according to the following rule: (i) if the temperature at any
measuring point in the composting mass exceeded 60 oC, continuous aeration; (ii)
between 55 oC and 60 oC, intermittent aeration according to a pre-set cycle of 5
minutes aeration and 5 minutes pause; and, (iii) for temperatures in the compost

48
mass below 55 oC, intermittent aeration according to a pre-set cycle of 5 minutes
aeration followed by 10 minutes pause.

4.3.2 Bench-Scale Tests

Four runs of bench-scale tests were conducted under different operating conditions.
The operating conditions of different bench-scale tests are summarized in Table 4.2.
Among the four bench-scale tests of run 1, two tests were performed under natural
aeration or self-heating condition and another two tests were performed under two
different forced aeration conditions. The objective of this run was to compare the
effect of two aeration conditions such as natural and forced aeration in the bench-
scale test. The same waste mixture was used for all the bench-scale tests of run 1.

In run 2, two bench-scale tests under natural aeration were conducted in parallel with
first stage pilot-scale tests 9a using the same waste mixture as used for the pilot-scale
test. Two bench-scale tests were also performed in run 3, in parallel with the second
stage pilot-scale test 9b using the same waste mixture. The objective of runs 2 and 3
was to compare the results with the corresponding pilot-scale tests.

Table 4.2 Operating conditions and duration of different bench-scale tests


Run Test Aeration condition Parallel pilot- Duration
scale test in days
1 1NA1 Natural aeration - 52
1NA2 Natural aeration - 52
1FA1 Forced aeration, 0.34 L/min.kg VS - 52
1FA2 Forced aeration, 0.21 L/min.kg VS - 52

2 2NA1 Natural aeration 9a 52


2NA2 Natural aeration 9a 22

3 3NA1 Natural aeration 9b 32


3NA2 Natural aeration 9b 32

4 4NA1 to Natural aeration - 21


4NA12
The objective of run 4 was to investigate the effect of initial moisture content on VS
degradation and overall temperature. Twelve bench-scale tests using the same waste

49
mixture were conducted under run 4. The waste mixture (compost of 8a) was air
dried to a moisture content below 30% and after that the moisture content was
adjusted to between 30 and 80% for the different tests.

All bench-scale and pilot-scale composting tests and required physico-chemical


analyses were performed in the Environmental Engineering Laboratory of the
Department of Civil Engineering, The University of Hong Kong.

4.4 Waste Materials Used

4.4.1 Type and Source of Waste

For the composting tests, organic solid wastes from the University of Hong Kong
were used. These consisted of two types of waste, namely food waste (from the
student canteen) and mixed office waste paper. In addition, sawdust from the
University Industrial Center was used as a bulking agent. The few non-compostable
and oversized waste components like plastic trays, aluminium cans, etc. were
manually removed from the food waste prior to use and analysis. The food wastes
mainly consisted of rice, noodles, bread, vegetables, meat and fish residue. The
mixed waste paper was shredded to a suitable particle size (1 to 1.5 cm) by a cutting
machine (Sam Chin Y-type high speed granulator SC 100, Sam Chin Cold Water
Machines Ltd.). The average initial moisture content of the raw, presorted food
waste, wastepaper and saw dust was 72.08, 8.47 and 8.31% and volatile solids (VS)
content was 94.32, 95.25 and 98.86%, respectively. Photographs of food waste and
shredded waste paper, together with the final stable compost are presented in Figure
A.1 of Appendix I.

4.4.2 Preparation of Initial Waste Mixtures

4.4.2.1 Bench-scale Test

Separately weighed food waste, shredded waste paper and sawdust were mixed by
hand in a plastic box to an homogeneous waste mixture. The amount of waste
50
mixture needed for bench-scale tests was smaller than that for pilot-scale tests.
However, for the parallel bench-scale tests, the same waste mixture as prepared for
pilot-scale tests was used.

4.4.2.2 First Stage Pilot-scale Test

The presorted food waste, shredded waste paper and sawdust were weighed
separately before mixing and then mixed uniformly for about 15 minutes in a
concrete mixing machine of 220 L volume. The moisture content of the initial waste
mixtures was adjusted to about 50-60% for optimum consistency and porosity by
adding sawdust.

4.4.2.3 Second and Third Stage Pilot-scale Test

The fresh compost produced by the first or second stage composting tests were used
as a initial feed material for subsequent second or third stage composting tests. After
first or second stage composting the product was re-mixed manually using a shovel
and adjusted with water, if necessary, for the second or third stage composting.

4.5 Process Monitoring and Data Acquisition

During the composting process, a number of changes occur in its physico-chemical


and biological properties. For determining the rate and extent of degradation of
biodegradable organic wastes for the above mentioned aeration modes, different
physico-chemical parameters were monitored over time which are discussed in the
following section.

4.5.1 Temperature

The temperature at the different depths of the waste mixture in the reactors: at 10
cm, 30 cm, 50 cm, 70 cm, 90 cm, and 110 cm, respectively, starting from the bottom,
and the ambient temperature were continuously monitored by datalogger (21X
Micrologger, Campbell Scientific Inc. USA). The datalogger was programmed in

51
such a way that, using datalogger software program (PC208, 1989, Campbell
Scientific Inc. USA), it measured temperature every 5 seconds for each point and
stored average hourly data (average of 720 data/hour) in different locations allocated
for different measuring points. If there was any change beyond the set point
temperatures it actuated its control panel to control the aeration within these 5
seconds. The stored data in the datalogger were downloaded to a personal computer
at routine intervals for subsequent analysis. Temperature of bench-scale tests were
also monitored by similar datalogger.

4.5.2 Air Velocity

Air velocity of the incoming air was monitored by the datalogger via hot wire
sensing probe and air velocity transmitter (model no. 640-0, Dwyer Instruments Inc.,
USA) and average hourly (average of 720 data/hour) data was stored. The hot wire
sensing probe was inserted in a long straight transparent and smooth duct (1.2 m
long and 31 mm in diameter) to measure the air velocity. A digital flow-meter
(model no. 730, Humonics Inc. USA) was used for the conversion of air velocity
(m/s) data to air flow rate (L/min) data by a calibration curve of air flow rate versus
air velocity. A Rotameter was connected just after the air flow regulator for the
instantaneous check of the flow rate and to check the above mentioned conversion of
air velocity to air flow rate.

4.5.3 Air Pressure

Air pressure (static) in mm of water of the incoming air was monitored by the
datalogger through pressure transducer (Hynewell Co.) and average hourly (average
of 720 data/hour) data was stored. The air pressure was monitored to ensure positive
and regular air flow into the reactors and for this purpose it was measured in the
same duct used for the measurement of air velocity.

4.5.4 Oxygen Content in the Spent Air

52
Percentage oxygen content in the spent air was measured by means of an oxygen
meter (MSI 150-Basic, MSI Elektronik GmbH). It was measured at a frequency of
about 8 to 10 times daily during the initial decomposition phase and after that at least
three times per day when the variation in the measurement was less.

4.5.5 Relative Humidity

Relative humidity was monitored by the datalogger via a relative humidity sensor
(HMP 130Y humidity and temperature transmitters, Vaisala Inc., USA). The
humidity sensor was placed near the experimental setup.

4.5.6 Height of the Composting Mass

The decrease in height of the composting mass was monitored in one or two day
intervals for the first one or two weeks of the test duration and after that it was done
weekly. However, it was not possible to monitor height during the upflow aeration
mode due to poor visibility created by the accumulated condensate inside the
transparent upper cover.

4.5.7 Leachate and Condensate

During the composting tests the production of leachate was almost zero and the
liquid inside the leachate tank was a mixture of previously stored tap water, leachate
and condensate. However, the water level in the leachate storage tank was monitored
at one or two day intervals. The water level monitoring was important during the
downflow aeration mode, internal air recirculation, and series connection since it
changed markedly by the partial accumulation of condensate. The water level was
also recorded before and after the leachate recirculation or water addition from the
leachate tank. For all types of tests, a large amount of condensate was produced
daily from the air outlet condenser and accumulated in a vessel connected after the
condenser. The condensate was regularly collected, measured and then preserved for
further analysis.

53
4.6 Sampling Program and Preparation

4.6.1 Waste and Compost

4.6.1.1 Sample Collection

At the beginning of each composting stage about 1.5 kg sample was collected from
the 65 to 75 kg of waste mixture or compost for subsequent analyses. This sample
was produced by thoroughly mixing 12 to 15 grab samples of 80 to 100 gm each.
The final composting mass of each test was divided into 6 layers along the height of
the reactor. The layers were denoted as L1, L2, L3, L4, L5 and L6, starting from
bottom. Each layer was approximately 20 cm thick, except the top layer, which had a
different thickness for different tests depending on the reduction of overall height of
the composting mass. About 500 gm sample per layer (mixture of 6 to 8 grab
samples) was collected separately. The fresh samples were used for immediate
analyses of moisture content (MC), total solids (TS), volatile solids (VS), pH,
ammonia nitrogen (NH3-N) and nitrate nitrogen (NO3-N), whereas for other analyses
the samples were dried in an air recirculating oven at 44 oC.

4.6.1.2 Sample Preparation

As mentioned above, the fresh samples were used for immediate analysis of MC,
VS, pH (of fresh sample), NH3-N and NO3-N, whereas for other analyses the
samples were dried in an air recirculating oven at 44 oC. The samples were mixed
several times during drying to obtain a uniformly air dried sample with moisture
content within 4 to 6% (Austrian Standard S 2023 1986). Usually the drying took 10
to 15 days. Part of the well mixed air dried samples was then reduced by a laboratory
blender (model no. 32BL79, Waring Commercial, USA) to particle size 0.63 mm
(98% mass fraction 2 mm and 80% mass fraction 0.63 mm) for further analysis.

4.6.2 Leachate and Condensate

Leachate samples were collected, in diluted form with stored tap water in the
leachate storage tank, at the end of each experiment, but also before the
54
establishment of series connection of two reactors. Condensate samples were
collected in one or two day intervals during the initial stage of decomposition, when
the production was high. After that the sampling frequency was reduced depending
on the production rate. After collection of both leachate and condensate samples the
pH was measured immediately, then the samples were preserved according to the
technique recommended by Standard Methods for the Examination of Water and
Wastewater (1995).

4.7 Physico-chemical Analyses

At the beginning and end of each stage of the composting tests, the initial waste
mixture and/or compost were analyzed with regard to the most relevant physico-
chemical parameters which are discussed in the following section.

4.7.1 Waste and Compost

For waste and compost at least three replicate samples were analyzed for all physico-
chemical parameters.

4.7.1.1 Moisture Content (MC)

A fresh sample of approximately 100 g was weighed accurately to 0.1 g and then
dried to constant mass at 105 3 oC which usually took 24 hours. It was then cooled
in a desiccator and weighed again. The percentage loss in weight, based on the initial
weight of fresh sample, was reported as moisture content (Austrian Standard S 2023
1986).

4.7.1.2 Volatile Solids Content (VS)

The dried sample from the moisture determination was placed in a muffle furnace at
550 oC for 5 hours, cooled in a desiccator, and weighed. The percentage loss in
weight, based on oven dry weight, was reported as volatile solids.

55
4.7.1.3 pH Fresh Sample

Approximately 10 g of fresh sample was suspended in 100 mL of distilled water


overnight. Then the pH of the supernatant was measured electrometrically (pH
meter, model 420A, Orion Research Inc. USA) (Gray et al. 1973).

4.7.1.4 pH Air Dried Sample

Approximately 10 g of sample of air dried sample prepared as described in section


4.6.1.2 was suspended with 100 mL 0.01 mol/L calcium chloride dihydrate solution
overnight. Then the pH of the supernatant was measured electrometrically (pH
meter, model 420A, Orion Research Inc. USA) (Austrian Standard S 2023 1986).

4.7.1.5 Conductivity

Ten g of air dried sample prepared as described in section 4.6.1.2 was suspended in
water at 1:10 mass ratio and then filtered after three hours overhead shaking.
Measurement of conductivity was performed with a conductivity measuring cell.
(Conductivity meter, model no 33, Yellow Spring Instrument Co. Inc., USA)
(Austrian Standard S 2023 1986).

4.7.1.6 Organic Carbon

Organic carbon of air dried sample prepared as described in section 4.6.1.2 was
determined by photometric measurement (at 578 nm, Lambda 2 UV/VIS
Spectrophotometer, Perkin-Elmer Corporation, GmbH) after dichromate reflux of
suitable amount of about 0.1 g of sample accurately weighed to 0.001 g according to
the method described in Austrian Standard S 2023.

4.7.1.7 Total Kjeldahl Nitrogen (TKN)

Total Kjeldahl Nitrogen of air dried sample prepared as described in section 4.6.1.2
was determined according to Austrian Standard S 2023, using a micro digester and a
rapid distillation apparatus (model no 60300-01 AT and 65000-00 H, respectively,

56
Labconco Corporation, USA) with a sample size of 0.5 to 1.0 g accurately weighed
to 0.001 g.

4.7.1.8 Ammonia Nitrogen (NH3-N)

Ammonia nitrogen in the extraction solution (1 part fresh sample to 4 parts 0.0125
mol/L CaCl2.2H2O/L solution) of sample was determined by distillation with caustic
soda and titrimetric method as described in Austrian Standard S 2023. Only the fresh
final samples after every stage were analysed for NH3-N.

4.7.1.9 Nitrate Nitrogen (NO3-N)

Nitrate nitrogen in the extraction solution (as produced in section 4.7.1.8) of fresh
sample was determined by the method described in Austrian Standard S 2023 and
photometric measurement (at 210 nm, Lambda 12 UV/VIS Spectrophotometer,
Perkin-Elmer Corporation, GmbH). The specific procedure used was as follows:

Pipette 5 mL extraction solution into each of two 25 mL graduated flasks.


Add 1 mL sulfuric acid (10 %) to each flask and make up to 25 mL with calcium
chloride working solution (0.0125 mol/L).
Add into one of the solutions 3 to 4 coppered zinc granules to reduce the nitrate.
leave standing overnight (18 to 20 hours maximal).
Determine the concentration of nitrate by measuring the absorbance difference
between the reduced and non-reduced solution at 210.0 nm. The difference is
measured by inserting the non-reduced solution into the sample beam and the
reduced solution into the reference beam. The result is obtained in mg NO3-N/5
mL sample. Further calculation was needed to convert the result to percentage
NO3-N of dry solids.
Only the fresh final samples after every composting stage were analysed for NO3-N.

4.7.1.10 Sieve Analysis

This method was used to determine the distribution of various sized particles in the
oven dried sample of initial waste material and produced compost. The samples were
57
passed through a standard series of sieves: 11.2, 8.0, 5.6, 4.0, 2.8, 2.0, 1.4, 0.63 mm
mesh size. The weight of the sample retained in each sieve was determined and
percentage finer or the cumulative percentage by weight passing each sieve was
calculated. The particle size distribution was presented as a curve, percentage finer
versus sieve size on a semi-logarithmic plot. The effective size and uniformity
coefficient as defined below were used to characterize the particle size distribution
curve (Smith 1994).
Effective size: is the maximum particle size of the smallest 10 percent and is
denoted as D10.
Uniformity coefficient: is the ratio of the maximum particle size of the smallest
60 percent to the effective size and is denoted as U, i. e. U=D60/D10. A uniform
sample will has a coefficient approaching 1, whereas a well graded sample has a
high uniformity coefficient.

4.7.1.11 Porosity

Porosity of the composting mass inside the reactor was calculated using the
following equation:
V Vwater VVS VFS
Porosity = *100% (4.1)
V
where V = Total volume of the undisturbed composting mass
Vwater = Volume of water in the composting mass
VVS = Volume of volatile solids
VFS = Volume of fixed solids
The volumes of water, volatile solids, and fixed solids are calculated from their
known mass as determined before and after each test divided by their approximate
density of 1.0, 1.1 and 2.4, respectively. The units used for the calculation of mass
and volume are kg and L, respectively.

4.7.2 Condensate and Leachate

The samples of condensate and leachate were regularly analysed for pH, chemical
oxygen demand (COD) and ammonia nitrogen (NH3-N).
58
4.7.2.1 pH

The pH of leachate and condensate were measured electrometrically (pH meter,


model 420A, Orion Research Inc. USA).

4.7.2.2 Chemical Oxygen Demand (COD)

Chemical Oxygen Demand was determined by the closed reflux, titrimetric method
as described on page 5-14 of Standard Methods for the Examination of Water and
Wastewater (1995).

4.7.2.3 Ammonia Nitrogen (NH3-N)

Ammonia Nitrogen was determined by the preliminary distillation, titrimetric


method as described on page 4-77 of Standard Methods for the Examination of
Water and Wastewater (1995).

4.8 Biological Product Stability Tests

4.8.1 Self-heating Test

The self-heating test was conducted according to the procedure recommended by


BGK (1994). Simple 1.0 L vacuum flasks were used instead of the recommended
expensive 1.5 L Dewar bottles. In addition, the mouth of the vacuum flasks was
covered with packing foam to prevent excessive heat loss. The temperature changes
were monitored by thermocouples instead of the conventional thermometer reading.
The experimental setup of a self-heating test is similar to the bench-scale composting
test as shown in Figure 4.3, only the forced aeration was not applied. At the
beginning and end of each test, total mass, total solids, and volatile solids of the
samples were determined.

59
A total of 16 sets of self-heating tests were conducted, with each set using a series of
6 vacuum flasks of 1 L volume. These 6 vacuum flasks were used for the individual
samples collected from the 6 different layers L1, L2, L3, L4, L5 and L6 of the
composting mass at the end of the pilot-scale tests. Each self-heating test was
denoted by the respective name of the pilot-scale test and layer number. For example
8bL4 means the self-heating test of the sample collected from the fourth layer (L4)
after completion the second stage pilot-scale test 8b.

60
CHAPTER 5

EXPERIMENTAL RESULTS

5.1 Pilot-scale Tests

Detailed presentation of the results obtained from eighteen pilot-scale tests is made
in the following sections. The results cover initial composition of waste mixture,
temperature variations, air supply, oxygen consumption, degradation of organic
matter, moisture content, and product stability.

5.5.1 Initial Composition of Waste Mixtures

The mix ratio and amount (based on the physico-chemical analysis of individual
component of waste mixture and/or waste mixture) of initial composition of the
waste materials used for the different pilot-scale tests are presented in Table 5.1. The
initial components of the waste mixtures of all first stage tests were food waste,
waste paper and sawdust except in test 4a, where produced compost was used
instead of saw dust. As mentioned in the previous chapter, the re-mixed and re-
moistened (if necessary) compost of the first stage test was used as a feed material
for the corresponding second stage test. However, in most of the cases water was
added to obtain an optimum moisture content. The initial moisture content of the
waste mixtures varied between 55 to 62% for first stage tests and 54 to 65% for
second stage tests, thus staying within the recommended range as discussed in
section 2.6.3. The range of initial weight of waste mixture put inside the reactor for
first and second stage tests were 55 to 77 kg and 36 to 59 kg, respectively.

61
In all waste mixtures, volatile solids amounted to more than 90% of the total solids;
however, the volatile solids originating from paper and sawdust were considered less
readily biodegradable due to their high cellulosic content (discussed in section 2.2).
Readily biodegradable food waste constituted between 43% and 55% of the total
solids and of the volatile solids, or 16% to 22% of the fresh initial waste mixture.

Since it was intended to use real waste instead of synthetic waste in the experiments
to simulate real plant conditions, the composition of the initial waste mixtures for the
first stage composting tests differed somewhat due to the slightly varying
characteristics of the food waste from the canteen. However these slight differences
are thought to have a negligible effect on the results which mainly depend on
aeration modes.

5.1.2 Temperature

In all pilot-scale tests, temperatures measured varied over time and height of
composting mass according to the aeration mode applied. The temperature patterns
at different layers of different pilot-scale tests are presented in Figures 5.1 to 5.17.
The average temperatures are shown together with oxygen consumption rates in
Figures 5.22 to 5.38. During the first two days of most first stage tests, evaporative
cooling reduced the temperature in the layers near air inlet by 5 to 9 oC from the
ambient temperature (23-25 oC). Evaporative cooling also happened successively in
top or bottom layers of test 7a, where alternate upflow/downflow aeration was
applied. In all first stage tests, 1 to 1.5 days of lag period were observed in the
temperature curves of all layers and after that the temperature started to rise
gradually, whereas the temperatures in most second stage tests rose quickly within
the first 6 to 12 hours to thermophilic range (above 50 oC). The second stage tests
did not show any lag period, because the microorganisms present in the compost
(produced from first stage test) were already acclimated with the environment.
However, effects of evaporative cooling on the temperature curves were observed
during some second stage tests.

Table 5.1 Initial composition and mix ratio of waste material used for different first,
62
second, and third stage composting tests
Test Waste % Wet Total H2O in % H 2O Total Volatile % VS
material/ ratio mass, total of total solids solids of TS
components kg mass, kg mass (TS), kg (VS), kg
1a food waste 55.55 42.89 29.60 13.28 12.82
paper 11.82 9.13 0.77 8.35 7.95
saw dust 12.52 9.67 0.80 8.87 8.78
water 20.09 15.51 15.51 ----- -----
waste mixture 100.00 77.20 46.68 60.50 30.50 29.54 96.90
1b compost of 1a 100.00 55.56 35.55 64.00 20.01 19.07 95.30
2a food waste 54.44 40.49 28.59 11.90 11.32
paper 11.67 8.68 0.74 7.94 7.56
saw dust 11.67 8.68 0.72 7.96 7.87
water 22.22 16.53 16.53 ----- -----
waste mixture 100.00 74.38 46.58 62.60 27.8 26.76 96.30
3a food waste 80.00 50.99 36.49 14.50 13.68
paper 10.00 6.37 0.54 5.83 5.55
saw dust 10.00 6.37 0.53 5.84 5.78
waste mixture 100.00 63.73 37.56 58.90 26.17 25.01 95.60
3b compost of 3a 100.00 52.55 34.30 65.30 18.24 17.03 93.40
4a food waste 80.00 59.46 42.62 16.84 15.59
paper 10.00 7.43 0.63 6.80 6.48
compost 10.00 7.43 3.69 3.74 3.17
waste mixture 100.00 74.32 46.94 58.30 27.38 25.54 93.30
4b compost of 4a 100.00 59.75 38.84 53.80 20.91 18.98 90.80
5a food waste 66.83 44.91 32.74 12.17 11.33
paper 10.49 7.05 0.60 6.45 6.14
saw dust 10.49 7.05 0.59 6.46 6.39
water 12.18 8.19 8.19 ----- -----
waste mixture 100.00 67.20 42.12 59.60 25.09 23.86 95.10
5b compost of 5a 100 41.84 23.20 53.40 18.63 17.33 93.00
5c compost of 5b 100 31.05 17.21 55.40 13.84 12.72 91.90
6a food waste 74.05 47.98 36.30 11.68 11.28
paper 11.62 7.53 0.63 6.89 6.56
saw dust 11.62 7.53 0.62 6.90 6.83
water 2.70 1.75 1.75 ----- -----
waste mixture 100.00 64.79 39.3 60.66 25.47 24.67 96.85
6b compost of 6a 100.00 36.85 21.88 59.38 14.97 14.02 93.65
7a food waste 75.01 41.56 29.52 12.03 11.41
paper 12.63 7.00 0.59 6.41 6.10
saw dust 12.35 6.84 0.54 6.27 6.20
waste mixture 100.00 55.40 30.65 55.32 24.71 23.71 95.95
7b compost of 7a 100.00 41.50 23.80 57.35 17.70 16.35 92.37
8a food wastea 71.33 45.41 32.15 13.26 12.10
paper 11.89 7.57 0.64 6.93 6.60
saw dust 11.89 7.57 0.63 6.94 6.87
water 4.88 3.11 3.11 ----- -----
waste mixture 100.00 63.66 36.53 57.40 27.13 25.57 94.20
8b compost of 8a 100.00 41.80 24.13 57.72 17.67 16.03 90.72
9a food wasteb 75.00 45.25 32.41 12.84 12.27
paper 12.5 7.54 0.64 6.90 6.57
saw dust 12.5 7.54 0.63 6.91 6.83
waste mixture 100.00 60.33 33.68 55.82 26.65 25.67 96.33
9b compost of 9a 100.00 41.28 24.67 59.76 16.61 15.62 94.04
a
contained excess amount of milk and sweet food waste, b did not contain chicken food waste

63
During most of the test period, the temperatures of the composting mass were
maintained below the maximum set point 60 oC. Due to quick degradation of well
mixed compost by acclimated microorganisms, rapid temperature increases up to 70
o
C were observed at the beginning of all second stage tests. Peak temperatures above
the maximum set point were also observed successively after the change of aeration
mode (airflow direction) in the case of alternate upflow/downflow aeration mode.
Generally four to eight hours were needed after the change of aeration mode to
reverse the temperature from maximum to minimum or minimum to maximum at top
or bottom layers of the composting mass, which is clearly indicated in Figure 5.12
for test 7a and in other Figures where alternate upflow/downflow aeration mode was
applied.

In tests 1a, 3a and 5a, upflow aeration was employed with internal recirculation
ratios of 20-50%, 100-200% and 20-80%, respectively. In test 8a only upflow
aeration was applied with no internal air recirculation. The average temperature in
tests 1a, 3a and 5a remained above 45 oC for more than 17, 9 and 16 days,
respectively, while in test 8a the temperature was above 45 oC for 6 days only.
During most of the test period, the average temperature was above 40 oC in tests 1a
and 5a and below 40 oC in test 3a, where excessive internal air recirculation may
have cooled down the composting mass. In contrast, the average temperature was
below 40 oC during most of the test period of test 8a.

The narrow band width between minimum and maximum temperature curves in tests
3a and 5a demonstrates clearly that the internal recirculation of air leads to a more
homogeneous temperature distribution throughout the composting mass as compared
to no recirculation. Although the band in test 1a is regular and narrower than in other
tests with no recirculation, it is wider than in tests 3a and 5a, owing to low internal
recirculation ratio.

In tests 3b, 5b, 7b and 8b after 17, 9, 1 and 10 days reuse of spent air from first stage
tests 4a, 6a, 8a and 9a, respectively, was applied in a two reactor system. The effect
of this series connection on the temperature curves is very clearly shown in

64
80
110 cm

70 90 cm
70 cm

60 50 cm
Temperature in oC

30 cm

50 10 cm

40

30

20 Ambient Temperature

10

0
0 5 10 15 20 25 30
Time in days

Figure 5.1 Temperature variation at different heights of composting mass during test
1a, upflow aeration with internal air recirculation, r = 20 to 50 % (hourly data are
plotted and all the symbols are only used to identify the curves).

80
110 cm
70 90 cm
70 cm
60 50 cm
Temperature in oC

30 cm
50 10 cm

40

30

20 Ambient Temperature

U = upflow D = downflow
10

0
0 5 10 15 20 25 30
Time in days

Figure 5.2 Temperature variation at different heights of composting mass during test
1b, upflow and downflow aeration.

65
80
Aeration was stopped for 3 hours

70

60
Temperature in oC

50

40
110 cm
30 90 cm
70 cm
20 50 cm
Low aeration rates for 24 hours
30 cm
10 Average Ambient
10 cm
Temperature 24.68 oC
0
0 5 10 15 20 25 30
Time in days

Figure 5.3 Temperature variation at different heights of composting mass during test
2a, downflow aeration.

66
80
110 cm
70 90 cm
70 cm
60 50 cm
Temperature in oC

30 cm
50 10 cm

40

30

20 Ambient Temperature

10

0
0 5 10 15 20 25 30
Time in days

Figure 5.4 Temperature variation at different heights of composting mass during test
3a upflow aeration with internal air recirculation, r = 100 to 200 %.

80
110 cm
70 90 cm
70 cm
60 50 cm
Temperature in oC

30 cm
50 10 cm

40 Series connection

30
Ambient Temperature
20

U D U U and Reuse of spent air


10

0
0 5 10 15 20 25 30 35 40 45
Time in days

Figure 5.5 Temperature variation at different heights of composting mass during test
3b, upflow/downflow aeration for 17 days, reuse of air from 4a as a series
connection.

67
80

70 U D

60
Temperature in oC

50

40

30 110 cm
90 cm

20 70 cm
Average Ambient 50 cm
10 Temperature 20.28 oC 30 cm
10 cm
0
0 5 10 15 20 25 30 35 40 45
Time in days

Figure 5.6 Temperature variation at different heights of composting mass during test
4a, upflow followed by downflow aeration after 28 days.

80
U D U 110 cm
70 90 cm
70 cm
60 50 cm
Temperature in oC

30 cm
50 10 cm

40

30

20
Ambient Temperature
(line without symbol)
10

0
0 5 10 15 20 25 30
Time in days

Figure 5.7 Temperature variation at different heights of composting mass during test
4b, upflow/downflow aeration.

68
80
110 cm
70 90 cm
70 cm
60 50 cm
Temperature in oC

30 cm
50 10 cm

40

30

20
Ambient Temperature

10

0
0 5 10 15 20 25 30
Time in days

Figure 5.8 Temperature variation at different heights of composting mass during test
5a, upflow aeration with internal air recirculation, r = 20 to 80 %.

80
D U U and Reuse of spent air U
70

60
Temperature in oC

50 Series connection

40
110 cm
90 cm
30
70 cm
50 cm
20 Ambient Temperature
30 cm
10 cm
10

0
0 5 10 15 20 25 30
Time in days

Figure 5.9 Temperature variation at different heights of composting mass during test
5b, upflow/downflow aeration for 9 days and then reuse of spent air from 6a.

69
80
110 cm
90 cm
70 Downflow for 3 hours only
70 cm
50 cm
60
30 cm
Temperature in oC

10 cm
50

40

30

20 Ambient Temperature

U D UD
10

0
0 5 10 15 20 25 30
Time in days

Figure 5.10 Temperature variation at different heights of composting mass during


test 6a, upflow/downflow aeration.

80
110 cm
70 90 cm
70 cm
60 50 cm
Temperature in oC

30 cm
50 10 cm

40

30

20
Ambient Temperature
(curve without symbol)
10

0
0 5 10 15 20 25 30
Time in days

Figure 5.11 Temperature variation at different heights of composting mass during


test 6b, upflow aeration.

70
80
D U D U D U D U 110 cm
70 90 cm
70 cm
60 50 cm
Temperature in oC

30 cm
50 10 cm

40

30

20
Ambient Temperature
(curve without symbol)
10

0
0 5 10 15 20 25 30
Time in days

Figure 5.12 Temperature variation at different heights of composting mass during


test 7a, upflow/downflow aeration alternatively for every 3 to 4 days.

80
110 cm
70 90 cm
70 cm
60 50 cm
Temperature in oC

30 cm
50 10 cm

40

30

20
Ambient Temperature
(curve without symbol)
10

0
0 5 10 15 20 25 30
Time in days

Figure 5.13 Temperature variation at different heights of composting mass during


test 7b, upflow aeration for 1 day and then reuse of spent air from 8a.

71
80
110 cm
70 90 cm
70 cm
60 50 cm
Temperature in oC

30 cm
50 10 cm

40

30

20
Ambient Temperature
(curve without symbol)
10

0
0 5 10 15 20 25 30
Time in days

Figure 5.14 Temperature variation at different heights of composting mass during


test 8a, upflow aeration.

80
110 cm
70 90 cm
70 cm
60 50 cm
Temperature in oC

30 cm
50 Series connection 10 cm

40

30

20
Ambient Temperature

10 U D U and Reuse of spent air

0
0 5 10 15 20 25 30
Time in days

Figure 5.15 Temperature variation at different heights of composting mass during


test 8b, upflow and downflow aeration for 10 days, after that reuse of spent air from
9a.

72
80
110 cm
70 90 cm
70 cm
60 50 cm
Temperature in oC

30 cm
50 10 cm

40

30

20
Ambient Temperature
(curve without symbol)
10

0
0 5 10 15 20 25 30
Time in days

Figure 5.16 Temperature variation at different heights of composting mass during


test 9a, downflow aeration.

80
110 cm
70 90 cm
70 cm
60 50 cm
Temperature in oC

30 cm
50 10 cm

40

30

20 Ambient Temperature
(curve without symbol)
10

0
0 5 10 15 20 25 30
Time in days

Figure 5.17 Temperature variation at different heights of composting mass during


test 9b, upflow aeration.

73
Figures 5.5, 5.9, 5.13 and 5.15. In all these cases, a narrower band width between
minimum and maximum temperature curves was observed after series connection
thus creating a most homogeneous temperature distribution throughout the
composting mass. In test 5b, the series connection was disconnected one day before
termination of the test which again caused a wider band width between minimum
and maximum temperature. In the two reactor system as a series connection, the
range of temperature difference between source first stage reactor and receiving
second stage reactor was 10 to 15 oC.

5.1.3 Air Supply and Oxygen Consumption

Air supply in L/min (or m3/d) was calculated from hourly monitored air velocity (in
m/sec) and from a calibration curve (see section 4.5.2). Hourly monitored relative
humidity and ambient temperature were used to determine dry air supply at ambient
condition. Daily and cumulative air supply for composting tests 6a, 6b, 9a and 9b are
shown in Figures 5.18 to 5.21 as examples. At the initial high decomposition phase
of all first and second stage tests the air supply was higher due to the higher heat
removal demand. Usually most second or successive peak demands were observed
after a change in air flow direction or in the case of series connection from a first
stage test to a second stage test. In test 6a, a second peak supply was needed as
shown in Figure 5.18, due to change of air flow direction from upflow to downflow
aeration, which may again have accelerated the degradation at the bottom layers of
the composting mass; after that the air demand decreased gradually. The total air
demand was always less in the second stage composting tests as compared to the
first stage tests as indicated in Figures 5.19 and 5.21.

The O2 concentration in the waste air fluctuated between 10-20%, with an average of
about 18%. The minimum O2 concentration in the waste air was observed at the
beginning of each test, when the degradation of VS and the resulting O2 demand
were highest, whereas the maximum O2 concentrations occurred towards the end of
each test. After observing an O2 concentration in the waste air between 18.5 to
19.5%, most of the first stage tests were terminated.

74
20 350

18
Daily air supply 300
16 Cumulative air supply

Cumulative air supply in m3


14 250
Daily air supply in m3/d

12
200
10
150
8

6 100

4
50
2

0 0
0 2 4 6 8 10 12 14 16 18 20 22 24 26 28 30
Time in days

Figure 5.18 Daily and cumulative air supply during first stage composting test 6a.

20 160

18
Daily air supply 140
16 Cumulative air supply
120
14 Cumulative air supply in m3
Daily air supply in m3/d

100
12

10 80

8
60
6
40
4
20
2

0 0
0 2 4 6 8 10 12 14 16 18 20 22 24 26 28 30
Time in days

Figure 5.19 Daily and cumulative air supply during second stage composting test
6b.

75
20 350

18
300
16 Daily air supply
Cumulative air supply

Cumulative air supply in m3


14 250
Daily air supply in m3/d

12
200
10
150
8

6 100

4
50
2

0 0
0 2 4 6 8 10 12 14 16 18 20 22 24 26 28 30
Time in days

Figure 5.20 Daily and cumulative air supply during first stage composting test 9a.

20 160

18
Daily air supply 140
16 Cumulative air supply
120 Cumulative air supply in m3
14
Daily air supply in m3/d

100
12

10 80

8
60
6
40
4
20
2

0 0
0 2 4 6 8 10 12 14 16 18 20 22 24 26 28 30
Time in days

Figure 5.21 Daily and cumulative air supply during second stage composting test
9b.
76
1.4 14 70

1.2 12

Cumulative O2 consumption, kg
60
O2 consumption rate, kg/d

O2 consumption rate
1.0 10

Temperature in oC
Cumulative O2 consumption
Average temperature
50
0.8 8

0.6 6
40

0.4 4
30
0.2 2

0.0 0 20
0 5 10 15 20 25 30
Time in days

Figure 5.22 Oxygen consumption rate, cumulative oxygen consumption and average
temperature during test 1a.

1.0 10 70
O2 consumption rate
Cumulative O2 consumption
Cumulative O2 consumption, kg

0.8 8 60
O2 consumption rate, kg/d

Average temperature
Temperature in oC

0.6 6 50

0.4 4 40

0.2 2 30

0.0 0 20
0 5 10 15 20 25 30
Time in days

Figure 5.23 Oxygen consumption rate, cumulative oxygen consumption and average
temperature during test 1b.

77
1.0 10 70

Cumulative O2 consumption, kg
0.8 8 60
O2 consumption rate, kg/d

Temperature in oC
0.6 6 50

0.4 4 40

O2 consumption rate
0.2 2 30
Cumulative O2 consumption
Average temperature

0.0 0 20
0 5 10 15 20 25 30
Time in days
Figure 5.24 Oxygen consumption rate, cumulative oxygen consumption and average
temperature during test 2a.

78
1.0 10 70
O2 consumption rate
Cumulative O2 consumption

Cumulative O2 consumption, kg
0.8 8 60
O2 consumption rate, kg/d
Average temperature

Temperature in oC
0.6 6 50

0.4 4 40

0.2 2 30

0.0 0 20
0 5 10 15 20 25 30
Time in days

Figure 5.25 Oxygen consumption rate, cumulative oxygen consumption and average
temperature during test 3a.

1.0 10 70
O2 consumption rate
Cumulative O2 consumption Cumulative O2 consumption, kg
0.8 8 60
O2 consumption rate, kg/d

Average temperature

Temperature in oC
0.6 6 50

0.4 4 40

0.2 2 30

0.0 0 20
0 5 10 15 20 25 30 35 40 45
Time in days

Figure 5.26 Oxygen consumption rate, cumulative oxygen consumption and average
temperature during test 3b.

79
1.0 18 70
O2 consumption rate
16
Cumulative O2 consumption

Cumulative O2 consumption, kg
0.8 60
O2 consumption rate, kg/d
Average temperature 14

Temperature in oC
12
0.6 50
10

8
0.4 40
6

0.2 4 30
2

0.0 0 20
0 5 10 15 20 25 30 35 40 45
Time in days

Figure 5.27 Oxygen consumption rate, cumulative oxygen consumption and average
temperature during test 4a.

1.0 10 70
O2 consumption rate
Cumulative O2 consumption Cumulative O2 consumption, kg
0.8 8 60
O2 consumption rate, kg/d

Average temperature

Temperature in oC
0.6 6 50

0.4 4 40

0.2 2 30

0.0 0 20
0 5 10 15 20 25 30
Time in days

Figure 5.28 Oxygen consumption rate, cumulative oxygen consumption and average
temperature during test 4b.

80
1.0 10 70
O2 consumption rate
Cumulative O2 consumption

Cumulative O2 consumption, kg
0.8 8 60
O2 consumption rate, kg/d
Average temperature

Temperature in oC
0.6 6 50

0.4 4 40

0.2 2 30

0.0 0 20
0 5 10 15 20 25 30
Time in days

Figure 5.29 Oxygen consumption rate, cumulative oxygen consumption and average
temperature during test 5a.

1.0 10 70
O2 consumption rate
Cumulative O2 consumption
0.8 8 Cumulative O2 consumption, kg 60
O2 consumption rate, kg/d

Average temperature

0.6 6 Temperature in oC
50

0.4 4 40

0.2 2 30

0.0 0 20
0 5 10 15 20 25 30
Time in days

Figure 5.30 Oxygen consumption rate, cumulative oxygen consumption and average
temperature during test 5b.

81
1.0 12 70
O2 consumption rate
Cumulative O2 consumption

Cumulative O2 consumption, kg
10
0.8 Average temperature 60
O2 consumption rate, kg/d

Temperature in oC
8
0.6 50

0.4 40
4

0.2 30
2

0.0 0 20
0 5 10 15 20 25 30
Time in days

Figure 5.31 Oxygen consumption rate, cumulative oxygen consumption and average
temperature during test 6a.

1.0 10 70
O2 consumption rate
Cumulative O2 consumption
0.8 Average temperature 8 Cumulative O2 consumption, kg 60
O2 consumption rate, kg/d

0.6 6 Temperature in oC
50

0.4 4 40

0.2 2 30

0.0 0 20
0 5 10 15 20 25 30
Time in days

Figure 5.32 Oxygen consumption rate, cumulative oxygen consumption and average
temperature during test 6b.

82
1.2 10 70
O2 consumption rate
Cumulative O2 consumption

Cumulative O2 consumption, kg
1.0 Average temperature 8 60
O2 consumption rate, kg/d

Temperature in oC
0.8
6 50

0.6

4 40
0.4

2 30
0.2

0.0 0 20
0 5 10 15 20 25 30
Time in days

Figure 5.33 Oxygen consumption rate, cumulative oxygen consumption and average
temperature during test 7a.

1.0 10 70
O2 consumption rate
Cumulative O2 consumption
Cumulative O2 consumption, kg
0.8 Average temperature 8 60
O2 consumption rate, kg/d

Temperature in oC
0.6 6 50

0.4 4 40

0.2 2 30

0.0 0 20
0 5 10 15 20 25 30
Time in days

Figure 5.34 Oxygen consumption rate, cumulative oxygen consumption and average
temperature during test 7b.

83
1.4 12 70
O2 consumption rate
Cumulative O2 consumption
1.2

Cumulative O2 consumption, kg
Average temperature
10
60
O2 consumption rate, kg/d

1.0

Temperature in oC
8
50
0.8
6
0.6
40
4
0.4
30
0.2 2

0.0 0 20
0 5 10 15 20 25 30
Time in days

Figure 5.35 Oxygen consumption rate, cumulative oxygen consumption and average
temperature during test 8a.

1.2 10 70
O2 consumption rate
Cumulative O2 consumption
Cumulative O2 consumption, kg
1.0
Average temperature 8 60
O2 consumption rate, kg/d

Temperature in oC
0.8
6 50

0.6

4 40
0.4

2 30
0.2

0.0 0 20
0 5 10 15 20 25 30
Time in days

Figure 5.36 Oxygen consumption rate, cumulative oxygen consumption and average
temperature during test 8b.

84
1.0 10 70
O2 consumption rate
Cumulative O2 consumption

Cumulative O2 consumption, kg
0.8 Average temperature 8 60
O2 consumption rate, kg/d

Temperature in oC
0.6 6 50

0.4 4 40

0.2 2 30

0.0 0 20
0 5 10 15 20 25 30
Time in days

Figure 5.37 Oxygen consumption rate, cumulative oxygen consumption and average
temperature during test 9a.

1.2 10 70
O2 consumption rate
Cumulative O2 consumption Cumulative O2 consumption, kg
1.0
Average temperature 8 60
O2 consumption rate, kg/d

0.8 Temperature in oC
6 50

0.6

4 40
0.4

2 30
0.2

0.0 0 20
0 5 10 15 20 25 30
Time in days

Figure 5.38 Oxygen consumption rate, cumulative oxygen consumption and average
temperature during test 9b.

85
Conversely, all second stage tests were terminated with a O2 concentration in the
waste air greater than 20%. A few first stage tests were terminated at a lower O2
concentration value than the range mentioned above to simulate other effects such as
shorter duration in each stage etc. Oxygen consumption rate was estimated from
calculated daily dry airflow (at inlet) and the % O2 concentration measured at the
spent air. O2 consumption rate in kg/d, cumulative O2 consumption and average
temperature of each test are shown in Figures 5.22 to 5.38. Higher O2 consumption
rates, especially at the beginning of each test, result in higher average temperature in
the composting mass. The range of total O2 consumption in different first stage tests
was 7.5 to 17.0 kg for different aeration modes, amounts of waste mixture and
duration of each test. The range of total O2 consumption in the different second stage
tests was only 2.5 to 5.0 kg. Therefore the O2 consumption in second stage tests was
approximately one third to one fourth of that in the first stage tests.

5.1.4 Initial and Final Physico-chemical Characteristics of Waste


and Compost

The initial and final physico-chemical characteristics of the waste mixture and
compost for the different first, second and third stage pilot-scale composting tests are
presented in the following subsections and are summarized in Tables 5.2 to 5.7.

5.1.4.1 Total Waste Mass and Moisture Content

Total duration and changes in the amount of waste mass in the pilot-scale tests are
summarized in Table 5.2. The total mass of the produced compost in first and second
stage tests ranged between 30 to 50 kg depending on the initial weight and the final
moisture content. At the end of the first stage tests the range of total mass reduction
was 30 to 45% (16 to 35 kg) with an average of 42%. The total mass reduction range
for the second stage tests was 14 to 26% (6 to 11 kg) with an average of 19%.

The moisture characteristics of the waste and compost in the pilot-scale tests are
summarized in Table 5.3. The average moisture content of the end products of all
61
first stage tests remained above 42% except for tests 3a and 5a, where high and
moderate rate internal air recirculation was applied with no further moisture
addition. In the case of the second stage tests, the average moisture content
remained above 49%. At the end of the first stage tests the range of total moisture
removal was 30 to 66% (9 to 26 kg) with an average of 51%. The total moisture
removal range for the second stage tests was 12 to 34% (4 to 8 kg) with an average
of 23%. However, the total average mass reduction and moisture removal at the end
of first stage tests was 2.2 times higher than that of second stage tests. The average
duration of first stage and second stage tests were 27.4 and 25.0 days, respectively.

5.1.4.2 Total Solids and Volatile Solids

Changes in total solids and volatile solids in the pilot scale tests are shown in Tables
5.4 and 5.5. The range of total solids reduction in the first stage tests was 22 to 37%
(6 to 12 kg) with an average of 30%. The total solids reduction for the second stage
tests was 10 to 21% (2 to 4 kg) with an average of 15%.

The range of volatile solids reduction in first stage tests was 24 to 37% (6 to 11 kg)
with an average of 31%. In test 4a a reduction of 48.4% VS was achieved after 49
days. The volatile solids reduction for the second stage tests was 11 to 22% (2 to 4
kg) with an average of 16%. It can be worked out that first stage tests of
approximately 27.4 days duration constituted about 65% to 70% of the overall
volatile solids removal, while approximately 25 days duration of second stage
composting removed the other 30% to 35%. The range of overall volatile solids
reduction of first, second and third (if any) stage tests combined was 36 to 54% with
an average of 43%. Assuming no biodegradation of cellulosic matter in paper and
sawdust, the overall reductions in volatile solids amounted to between 75% and 89%
of the initial volatile solids content in the food waste.

5.1.4.3 pH

The pH values of the air dried samples of the initial waste mixture varied between
5.0 to 5.5 with an average of 5.25 according to the values presented in Table 5.6. All

62
the final values indicated in Table 5.6 are the average of six values determined for
six different layers. The final pH values of the air dried samples after the second
stage test varied between 5.7 to 6.5 with an average of 6.1. However, the pH values
of the final fresh samples were about one unit higher than those obtained after air
drying of the same samples, probably due to the presence of ammonia. The range of
pH values of the fresh sample was 6.1 to 8.0 with an average value 7.13. Therefore
the average pH of fresh samples was close to neutrality.

5.1.4.4 Conductivity

The conductivity of the initial waste samples varied between 1.7 to 4.2 mS/cm with
an average of 2.9 mS/cm, as indicated in Table 5.6. The range of conductivity of the
final samples after completion of all the stages was 4.5 to 7 mS/cm with an average
of 5.7 mS/cm. The conductivity value of the final compost after completion of every
stage of each run always increased from its starting value probably due to
accumulation of mineral salt in the remaining compost through VS reduction. The
average increment was 2.1 mS/cm.

5.1.4.5 Ammonia Nitrogen (NH3-N)

The NH3-N content of the final fresh samples is presented in Table 5.6 in g NH3-
N/kg dry solids. The range of NH3-N content after first stage tests was 1.6 to 4.9 g
NH3-N/kg dry solids and after second stage tests was 3 to 8 g NH3-N/kg dry solids.
The average values after first and second stage tests were 3.7 and 4.3 g NH3-N/kg
dry solids, respectively.

5.1.4.6 Nitrate Nitrogen (NO3-N)

The NO3-N content of the final fresh samples, in mg NO3-N /kg dry solids, is also
presented in Table 5.6. In most of the cases the NO3-N content was below the
detectable range. NO3-N was present only in the samples obtained after tests 1a, 1b,
3b, 6b and 9b. Most of these tests were second stage tests.

Table 5.2 Mass changes in waste mixtures during pilot-scale tests


63
Duration in Waste mixture
Test days Initial, kg Final, kg Removed, kg Removed, %
1a 31 77.20 42.26 34.94 45.25
1b 21 (52)* 55.56 46.26 9.30 16.74
2a 33 (33) 74.38 41.98 32.40 43.56
3a 29 63.73 32.28 31.45 49.35
3b 39 (68) 52.55 45.02 7.53 14.33
4a 49 74.32 39.45 34.87 46.91
4b 21 (70) 59.75 50.52 9.23 15.44
5a 20 67.20 34.90 32.30 48.06
5b 21 41.84 33.90 7.94 18.98
5c 11 (52) 31.05 30.89 0.16 0.51
6a 30 64.79 43.90 20.89 32.24
6b 21 (51) 36.85 27.62 9.23 25.04
7a 23 55.40 39.09 16.31 29.44
7b 32 (55) 41.50 32.32 9.18 22.12
8a 28 63.66 37.17 26.49 41.61
8b 28 (56) 41.80 36.15 5.65 13.51
9a 24 60.33 33.12 27.21 45.10
9b 32 (56) 41.28 30.30 10.98 26.60
* Total duration in days including first, second and third stages

Table 5.3 Moisture changes in waste mixtures during pilot-scale tests


Test Moisture in kg Moisture in %
Initial Additiona Final Removed Initial Final Removedb
1a 46.40 4.15 22.56 23.84 60.10 43.61 51.37
1b 35.55 4.23 30.55 5.00 63.98 63.35 14.06
2a 46.38 - 21.35 25.02 62.36 42.67 53.94
3a 37.56 - 12.57 24.99 58.93 31.01 66.53
3b 34.31 4.5 30.38 3.93 65.29 66.05 11.45
4a 46.94 - 24.49 22.45 63.16 61.66 47.82
4b 38.84 - 31.83 7.01 65.00 62.93 18.05
5a 42.13 - 15.36 26.77 62.68 37.29 63.45
5b 23.20 13.24 17.62 5.59 55.46 50.70 24.09
5c 17.21 - 17.42 -0.21 55.43 55.99 -1.22
6a 39.31 8.73 26.91 12.40 60.68 61.34 31.54
6b 21.88 0.81 14.39 7.49 59.38 49.44 34.23
7a 29.77 15.21 20.73 9.04 53.41 52.49 30.36
7b 23.80 8.88c 16.94 6.86 57.36 52.41 28.82
8a 36.53 - 17.49 19.05 57.39 47.05 52.15
8b 24.13 1.20 20.00 4.12 57.72 54.68 17.07
9a 33.68 - 12.90 20.78 55.83 38.95 61.69
9b 24.67 0.86 16.48 8.19 59.77 51.56 33.20
a
Sum of optionally added moisture during test, b % of initial moisture of each stage

Table 5.4 Changes in total solids during pilot-scale tests


Test Total solids (TS) in kg Total solids in %
64
Initial Final Removed Initial Final Removal* ()
1a 30.80 19.70 11.10 39.90 56.39 37.02
1b 20.01 15.71 4.30 36.02 36.65 21.48 (49.78)
2a 27.99 20.60 7.40 37.64 57.33 26.43 (26.43)
3a 26.18 19.71 6.47 41.07 68.99 24.71
3b 18.24 14.64 3.60 34.71 33.96 19.73 (39.57)
4a 27.38 14.96 12.42 36.84 38.34 45.36
4b 20.91 18.68 2.23 35.00 37.07 10.66 (51.20)
5a 25.09 19.94 5.55 37.32 55.99 22.12
5b 18.63 16.28 2.35 44.54 49.30 12.61
5c 13.84 13.47 0.37 44.57 44.01 02.67 (33.76)
6a 25.48 16.99 8.49 39.32 38.66 33.32
6b 14.97 13.23 1.74 40.62 50.56 11.62 (41.07)
7a 25.64 18.36 7.28 46.59 47.51 28.39
7b 17.70 15.38 2.32 42.64 47.59 13.11 (37.78)
8a 27.13 19.68 7.44 42.61 52.95 27.42
8b 17.67 15.87 1.81 42.28 45.32 10.24 (34.85)
9a 26.65 20.22 6.43 44.17 61.05 24.13
9b 16.61 13.82 2.79 40.23 48.44 16.80 (36.87)
* % of initial TS of each stage
() total % removal, determined on the basis of initial TS of the first stage

Table 5.5 Changes in volatile solids during pilot-scale tests


Test Volatile solids (VS) in kg Volatile solids in %
Initial Final Removed Initial Final Removal* ()
1a 29.79 18.76 11.03 96.70 95.28 37.02
1b 19.07 14.73 4.34 95.28 93.46 22.75 (51.35)
2a 26.48 19.16 7.33 95.13 92.98 27.68 (27.68)
3a 25.01 18.67 6.34 95.17 94.75 25.35
3b 17.03 13.40 3.63 93.37 91.54 21.31 (41.26)
4a 25.54 13.28 12.36 93.28 88.23 48.39
4b 18.98 16.79 2.19 90.77 89.91 11.53 (54.34)
5a 23.86 18.18 5.68 95.01 93.03 23.80
5b 17.33 14.93 2.41 92.99 91.77 13.91
5c 12.72 12.38 0.34 91.77 91.90 02.67 (36.16)
6a 24.37 15.89 8.49 95.68 93.69 34.84
6b 14.02 12.28 1.74 93.69 92.88 12.41 (42.93)
7a 24.16 17.01 7.16 94.25 92.37 29.63
7b 16.35 13.97 2.38 92.37 90.84 14.55 (39.88)
8a 25.57 18.07 7.49 94.26 91.81 29.29
8b 16.03 14.24 1.79 90.69 89.77 11.17 (37.18)
9a 25.61 19.27 6.33 96.09 95.30 24.72
9b 15.62 12.74 2.88 94.04 92.15 18.43 (38.59)
* % of initial VS of each stage
() total % removal, determined on the basis of initial VS of the first stage

Table 5.6 Final NH3-N and NO3-N and changes in pH and conductivity during
pilot-scale tests
Test NH3-N, NO3-N, pH Conductivity, mS/cm
65
g/kg-TS mg/kg-
TS
Final* Final* Initial Final Final* Initial Final
1a 4.89 66.68 4.86 5.34 - 1.70 1.66
1b 4.05 86.52 5.34 5.69 - 1.66 1.75
2a 3.49 0.00 5.27 5.82 - 2.60 3.22
3a 3.79 0.00 5.04 5.43 - 2.70 4.29
3b 8.05 10.40 5.43 5.97 - 4.29 5.88
4a 4.40 0.00 5.54 6.14 - 4.20 6.48
4b 3.82 0.00 5.40 5.78 - 6.50 7.00
5a 2.75 0.00 5.13 5.42 6.69 2.40 3.70
5b 3.97 0.00 5.42 6.19 7.72 3.68 4.94
5c 3.97 0.00 6.19 6.35 - 4.94 5.22
6a 4.01 0.00 5.54 6.17 7.59 2.95 4.90
6b 3.09 57.83 6.17 5.94 - 4.90 5.87
7a 4.22 0.00 5.55 6.14 - 3.68 4.54
7b 4.61 0.00 6.14 6.40 6.95 4.54 5.30
8a 3.97 0.00 5.17 5.93 6.35 3.20 4.61
8b 3.49 0.00 5.93 6.55 8.00 4.61 5.94
9a 1.65 0.00 5.21 5.58 6.07 2.60 3.38
9b 3.12 12.21 5.58 5.72 7.66 3.38 4.63
* Fresh solid samples were used of the analysis

Table 5.7 Changes in organic carbon (C), total Kjeldahl nitrogen (TKN) and
C/N ratio during pilot-scale tests
Test Organic Carbon %* TKN %* C/N ratio
Initial Final Initial Final Initial Final
1a 44.90 43.75 1.44 1.99 31.21 22.21
1b 43.75 43.90 1.99 2.09 22.21 21.29
2a 43.62 41.28 1.54 1.86 28.32 22.26
3a 45.15 43.75 2.02 2.29 22.35 19.24
3b 43.75 42.51 2.04 2.29 21.44 18.61
4a 44.33 40.69 2.13 2.38 20.81 17.15
4b 38.60 38.42 2.32 2.33 16.64 16.55
5a 44.41 42.81 1.61 2.43 27.58 17.59
5b 42.81 41.46 2.45 1.82 17.47 23.56
5c 41.46 42.01 1.82 2.09 22.78 20.10
6a 43.77 42.97 1.46 1.97 29.98 21.94
6b 42.97 43.01 1.97 2.08 21.81 20.68
7a 43.77 43.03 2.03 2.42 21.56 17.85
7b 43.03 42.34 2.42 2.44 17.78 17.39
8a 43.87 42.96 2.05 2.34 21.40 18.49
8b 42.96 40.10 2.34 2.23 18.49 18.00
9a 43.57 42.84 1.47 1.94 29.64 22.25
9b 42.84 41.91 1.94 2.06 22.25 20.39
* % of air dried solid samples
5.1.4.7 Organic Carbon

66
The average initial and final percentage organic carbon is presented in Table 5.7.
The organic carbon of initial and final samples varied between 44.9 to 38.4% of dry
solids. The percentage of organic carbon in the final samples was always found 2 to
3% less than that in the initial samples. The average organic carbon of all initial
waste mixtures before first stage tests and for final compost after second stage tests
were 44.15 and 41.71%, respectively.

5.1.4.8 Total Kjeldahl Nitrogen (TKN)

The average initial and final percentage of TKN is presented in Table 5.7. The TKN
of initial and final samples varied between 1.4 to 2.4% of dry solids. Due to the loss
of volatile solids, the percentage of TKN in the final samples was always found 0.4
to 0.6% higher than that in the initial samples. The average TKN of all initial waste
mixtures before first stage tests and for final compost after second stage tests were
1.75 and 2.20% of dry solids, respectively.

From Table 5.7 it can also be seen that the range of initial C/N ratios was 21 to 31
with an average of 25.9. The range of C/N ratios of final compost after second stage
tests was 16.5 to 21.3 with an average of 19.1. Therefore, C/N ratios decreased as
composting proceeded.

5.1.4.9 Sieve Analysis

Particle size distribution curves for tests 1a and 1b and tests 5a, 5b and 5c are shown
in Figures 5.39 and 5.40. The overall particles size varied between 0.6 to 10.0 mm.
The D10 of samples of tests 1a and 1b were 0.76 and 0.23 mm. The uniformity
coefficient of samples of these tests were 6 and 13, respectively. The D10 of samples
of tests 5a, 5b and 5c were 0.42, 0.30 and 0.17 mm. The uniformity coefficient of the
samples from these tests were 10, 11 and 16, respectively. Most D10 values were
calculated from extrapolation of curves presented in Figures 5.39 and 5.40. The
particle size was reduced after every composting stage and the D10 value gradually

67
100

1a
80 1b
Percentage finer

60

40

20

0
0.10 1.00 10.00 100.00
Particle size in mm

Figure 5.39 Particle size distribution curve of test 1a and 1b. The D10 of test 1a and
1b are 0.76 and 0.23 mm. The uniformity coefficient of the tests are 6 and 13
respectively. D10 value for test 1b was calculated from extrapolation of curve.

100

5a
80 5b
5c
Percentage finer

60

40

20

0
0.10 1.00 10.00 100.00
Particle size in mm
Figure 5.40 Particle size distribution curve of test 5a, 5b and 5c. The D10 of test 5a,
5b and 5c are 0.42, 0.30 and 0.17 mm. The uniformity coefficient of the tests are 10,
11 and 16 respectively. All D10 values were calculated from extrapolation.

68
became smaller. After every stage the compost also gradually became more well
graded as indicated by a higher uniformity coefficient than the initial one. More
stages may produce more well graded compost as comparing to the results of tests 1b
and 5c.

5.1.5 Physico-chemical Characteristics of Condensate and Leachate

Variations in production, pH, COD, and NH3-N of condensate over the whole
duration of some first and second stage pilot-scale tests are presented together in
Figures 5.41 to 5.46. A summary is given in Table 5.8.

5.1.5.1 Amount of Production

The amount of condensate collected on the sampling day is shown by the bar chart in
Figures 5.41 to 5.46. During the peak degradation period, the maximum condensate
production rate was 3 to 4 L/d. After that condensate was not collected frequently, as
the daily production rate greatly decreased to a lower quantity. The ranges of total
condensate produced in first and second stage tests were 7 to 26 L and 5 to 15 L,
respectively, as summarized in Table 5.8. About 1.5 to 3.5 times more condensate
was produced in first stage tests than in second stage tests.

The leachate production during the pilot-scale tests was very low. The range of total
leachate production over the whole duration of the tests was 0.0 to 0.9 L, either in
first or second stage tests (Table 5.8). The leachate production increased sometimes
in the intermediate leachate storage tank mainly due to the accumulation of
condensate during internal air recirculation or series connection of two reactors.

5.1.5.2 pH

The variations in pH of condensate produced during composting tests 5a, 7a, 3a, 6a,
5b and 7b are shown in Figures 5.41 to 5.46. During tests 5a and 6a, the pH dropped
from its initial value to about 3.75.

69
10 16 5

4
8 12

NH3-N, g
COD, g 3

COD, g
pH

NH3-N, g
6 8
pH
2
Condensate, L
4 4
1

2 0 0
4

Volume, L
2

0
0 5 10 15 20 25 30
Time in days

Figure 5.41 pH and total amount of COD and NH3-N in the produced condensate
during pilot-scale test 5a. Total amount of COD was calculated as concentration of
COD total volume of condensate collected on that day (as shown in figure). Same
procedure was spplied for NH3-N.

10 16 5

4
8 12

NH3-N, g
3
COD, g
pH

6 COD, g 8
NH3-N, g 2
pH
4 4
Condensate, L 1

2 0 0
4
Volume, L

0
0 5 10 15 20 25 30
Time in days

Figure 5.42 pH and total amount of COD and NH3-N in the produced condensate
during pilot-scale test 7a. Values except pH, of 11th and 23rd days were equally
distributed over the respective previous days.

70
10 16 5

4
8 COD, g 12
NH3-N, g

NH3-N, g
3

COD, g
pH
pH

6 8
Condensate, L
2

4 4
1

2 0 0
4

Volume, L
2

0
0 5 10 15 20 25 30
Time in days

Figure 5.43 pH and total amount of COD and NH3-N in the produced condensate
during pilot-scale test 3a.

10 16 5
COD, g
NH3-N, g
4
8 pH 12
Condensate, L
3 NH3-N, g
COD, g
pH

6 8
2

4 4
1

2 0 0
4
Volume, L

0
0 5 10 15 20 25 30
Time in days

Figure 5.44 pH and total amount of COD and NH3-N in the produced condensate
during pilot-scale test 6a.

71
10 4 20

16
8 3

NH3-N, g
12

COD, g
COD, g
pH

6 NH3-N, g
2
pH
8

4 Condensate, L 1
4

2 0 0
4

Volume, L
2

0
0 5 10 15 20 25 30
Time in days

Figure 5.45 pH and total amount of COD and NH3-N in the produced condensate
during pilot-scale test 5b.

10 4 20

16
8 3

NH3-N, g
12
COD, g
pH

6 COD, g 2
NH3-N, g 8
pH
4 1
Condensate, L 4

2 0 0
4
Volume, L

0
0 5 10 15 20 25 30
Time in days

Figure 5.46 pH and total amount of COD and NH3-N in the produced condensate
during pilot-scale test 7b.

72
After 6 and 12 days for tests 5a and 6a, respectively, the pH again sharply increased
to 8 as in the other first stage tests 7a and 3a. No decrease in pH of the condensate
from tests 7a and 3a were observed. During all second stage tests the pH values
always remained above 8 to 9 and at the end came closer to a value 8 as shown in
Figures 5.45 and 5.46.

The pH of leachate samples, diluted with the tap water stored in the intermediate
storage tank of the reactors, varied between 6 to 8.3 and 7.6 to 8.6, respectively, for
different first and second stage tests. However, leachate which was mixed with
condensate during different modified aeration modes such as downflow aeration,
upflow/downflow aeration, etc., showed higher pH values of about 8.

5.1.5.3 Chemical Oxygen Demand (COD)

During first stage tests, the initial total COD of the condensate varied between 6 to
15 g and after 5 to 10 days it sharply decreased to a value near zero, while at the
same time the pH increased to its maximum value as shown in Figures 5.41 to 5.44.
The initial total COD of the produced condensate during second stage tests varied
between 0.5 to 1.0 g and after 3 to 5 days it gradually decreased to a value near zero
as indicated in Figures 5.45 and 5.46. Throughout the second stage tests, the COD in
the condensate was much lower than that in first stage tests.

5.1.5.4 Ammonia Nitrogen (NH3-N)

At the beginning of all the first stage tests, the NH3-N content in the condensate was
very low, but after that the NH3-N content gradually increased as pH increased
towards maximum value. After reaching a maximum, the NH3-N production again
gradually decreased towards the end of the tests as shown in Figures 5.41 to 5.44. At
the beginning of the second stage tests, the NH3-N production was approximately
five times higher than in the first stage tests. However, at the end of the second stage
tests, the NH3-N production also gradually decreased as shown in Figures 5.45 and
5.46.

73
Table 5.8 Total volume, COD and NH3-N of produced condensate and leachate
Test Condensate Leachate
Production Total Total Production, L Total Total
L COD, g NH3-N, g Actual Diluted COD, g NH3-N, g
1a 25.76 50.81 11.16 0.45 5.58 12.95 -
1b 6.65 2.08 27.16 0.45 19.35 5.28 -
2a 12.65 9.98 4.64 *9.09 29.34 44.48 8.55
3a 16.16 13.53 11.57 **6.75 22.58 3.48 12.76
3b 8.05 3.83 50.87 *1.18 12.15 1.70 6.26
4a 16.69 50.13 53.87 *2.16 22.86 5.81 29.44
4b 5.32 0.79 18.87 0.90 20.70 - 7.41
5a 20.52 42.91 8.56 **8.78 23.67 9.28 2.72
5b 13.49 2.94 44.00 *1.88 8.93 0.80 2.55
5c 0.60 0.13 2.25 0.00 12.93 0.61 0.39
6a 15.74 49.58 10.61 0.27 7.74 0.42 0.74
6b 4.90 0.88 9.00 *2.70 24.57 0.86 10.39
7a 20.55 19.33 35.16 *2.74 10.43 3.33 1.53
7b 15.00 3.18 53.65 0.00 16.97 1.93 4.35
8a 10.95 12.32 4.50 0.00 18.09 7.42 0.32
8b 4.45 1.23 30.00 0.47 11.28 1.10 2.41
9a 7.01 14.90 2.06 *8.64 27.09 18.22 3.12
9b 5.60 1.08 18.33 0.00 16.00 0.61 0.59
* Condensate from downflow or upflow/downflow aeration increased the volume
** Condensate from internal air recirculation increased the volume

Table 5.9 Initial and final height and porosity of the composting mass
Test Height of composting mass in cm Porosity in %
Initial Final Change %Change Initial Final
1a 154 132 22 14.28 46.68 66.72
1b 143 110 33 23.07 62.74 55.35
2a 158 122 36 22.78 50.01 64.14
3a 160 125 35 21.88 51.54 69.41
3b 147 113 34 23.12 56.36 51.37
4a 136 81 55 40.44 42.05 48.95
4b 158 124 34 21.51 59.93 57.09
5a 160 125 35 21.87 48.73 66.88
5b 157 125 32 20.38 67.91 67.61
5c 135 118 17 12.59 72.37 68.51
6a 165 122 43 26.06 58.00 62.00
6b 143 113 30 20.98 73.00 74.00
7a 160 120 40 25.00 58.00 61.00
7b 153 117 36 23.52 67.00 66.00
8a 151 121 30 19.87 48.96 65.01
8b 160 135 25 15.62 68.60 68.22
9a 164 128 36 21.95 61.12 73.25
9b 160 135 25 15.62 68.68 73.06

5.1.6 Height and Porosity


74
A quick settlement from an initial height of the composting mass of 170 cm to about
140 cm occurred within the initial 2 to 4 days and then gradually settled about
another 5 cm during the rest of the test period as shown in Figures 5.47 and 5.48 for
first and second stage tests, respectively. The range of overall height reduction or
settlement in all the tests was 14 to 26 % (22 to 43 cm) with an average of 22% as
calculated from Table 5.9. Only in test 4a the height decreased by 40%, where food
content was higher than in other tests (61% of total dry solids). The initial and final
average porosities of waste and compost are shown in Table 5.9. The range of
porosity of the waste mixture was 47 to 74%. Sometimes the final porosity was
increased due to moisture removal or was decreased due to addition of moisture from
the top of the reactor.

5.1.7 Physical Appearance of the Compost

The color of the fresh compost from the first stage was yellowish brown to light
brown, whereas the mature (final) compost from the second or third stage
composting was brown to dark brown color with friable texture. The overall particle
size of the screened compost was less than 10 mm as mentioned in section 5.1.4.9.

75
180
2a
3a
Height of composting mass in cm
5a
160 7a
9a

140

120

100
0 5 10 15 20 25 30
Time in days

Figure 5.47 Reduction of height of composting mass during different first stage tests

180

3b
Height of composting mass in cm

4b
160 5b
7b

140

120

100
0 5 10 15 20 25 30
Time in days

Figure 5.48 Reduction of height of composting mass during different second stage
tests

5.1.8 Product Stability


76
The product stability of the composts as determined by different methods such as:
self-heating test, temperature decline, color and odor, C/N ratio, pH, and fungus
growth test are presented in this section.

5.1.8.1 Self-heating Test

A total of 96 self-heating tests were performed in 16 sets after completion of the


pilot-scale tests. The range of duration of these sets of self-heating tests were 15 to
30 days. The tests were terminated, once the temperature of the sample inside the
vacuum flask declined to 5 oC of ambient temperature. The results of the tests of
each set are presented in detail in Tables A.1 to A.16 in Appendix I.

5.1.8.1.1 Temperature

The temperature patterns of the self-heating tests for separate samples collected from
6 different layers of first stage test 8a and second stage test 8b are shown in Figures
5.49 and 5.50, respectively. All the samples from the first stage test 8a rose to peak
temperatures between 52 to 68 oC and the temperature remained above 50 oC for
long periods of time, whereas the samples from second stage test 8b peaked between
31 to 48 oC.

The biological stability index for the final product of each pilot-scale test was
determined on the basis of the average temperature curve of all layers. The average
temperature curves from the self-heating tests for the three different stage tests 5a,
5b and 5c are shown in Figure 5.51. This figure represents the typical temperature
curves for fresh, semi-mature and mature compost. Figure 5.52 illustrates the
temperature variation during self-heating tests of stored mixed samples collected at
different days (1, 7, 30 and 60 days) after completion of the second stage pilot-scale
test 9b. The two replicate self-heating tests (A and B) of each sample showed a very
similar temperature pattern. It is also clearly seen that the older samples produced
lower peak temperatures and thereby indicated less microbial activity than the less
aged samples.

77
Table 5.10 shows the maximum average temperature, average temperature
parameters and stability index (SI) calculated from six self-heating tests after each
pilot-scale test. In the self-heating test of samples from test 5a, a maximum
temperature of 64.25 oC was reached with A72 of 3594 oC.h, which indicates a
biological stability index of I according to Table 2.2. The self-heating tests of
samples from tests 5b and 5c attained maximum temperatures of 46.17 oC and 36.84
o
C (with A72 of 2506 and 1652 oC.h), respectively, resulting in stability indices of III
and V, respectively. It is of interest to note that the only other completely mature end
product with stability index V was obtained from two stage tests 3a and 3b after 68
days, where also reuse of spent air was applied. Composting in three stages with
reuse of spent air produced a completely stable end product in even shorter time
(about 50 days) as demonstrated by tests 5a, 5b and 5c.

As indicated in Table 5.10, all end products after first stage pilot-scale tests of 20 to
30 days duration were found to have stability index I and therefore had to be
considered biologically unstable. It appears, therefore, that even with the best
possible aeration practices of internal air recirculation and reuse of spent air a
minimum of 7 weeks (approximately 50 days) composting time is necessary to
achieve a completely stable end product.

5.1.8.1.2 Physico-chemical parameters

Duration and average values (of six self-heating tests) of physical parameters of the
self-heating tests after each pilot-scale test are presented in Table 5.11. The average
initial mass for the self-heating tests was 315 to 420 g (average of 84 tests is 370g) ,
except for tests 2a and 3a, where the mass was approximately 500 g due to higher
moisture content. Initial moisture content varied between 55 to 65% with an average
of 58.6%. Initial VS of the samples varied between 88.2 to 94.7 %. About 11 to 30%
VS with an average of 18.8% was degraded in those self-heating tests,

78
80
L1

70 L2
L3

60 L4
Temperature in oC

L5

50 L6

40

30

20
Ambient Temperature
10

0
0 5 10 15 20 25
Time in days

Figure 5.49 Temperature variation during self-heating tests of individual samples


collected separately from six layers after first stage pilot-scale test 8a (L1 = bottom
layer of 20 cm).

80
L1

70 L2
L3

60 L4
Temperature in oC

L5

50 L6

40

30

20
Ambient Temperature
10

0
0 5 10 15 20 25
Time in days

Figure 5.50 Temperature variation during self-heating tests of individual samples


collected separately from six layers after second stage pilot-scale test 8b.

79
80
5a (26.6)
70 5b (24.9)
5c (21.9)
60
() Average ambient
Temperature in oC

temperature, oC
50

40

30

20

10

0
0 5 10 15 20 25
Time in days

Figure 5.51 Average temperature variation during self-heating tests after first,
second and third stage pilot-scale tests: 5a, 5b and 5c respectively. Each curve is the
average of six self-heating test data.

80
01 days, A (23.8)

70 01 days, B (23.8)
07 days, A (24.6)

60 07 days, B (24.3)
Temperature in oC

30 days, A (26.1)

50 30 days, B (26.1)
60 days, A (23.6)
40 60 days, B (23.6)
() Average ambient
temperature, oC
30

20

10

0
0 5 10 15 20 25
Time in days

Figure 5.52 Temperature variation during self-heating tests of mixed samples


collected at different days after second stage pilot-scale test 9b. Two self-heating
tests (A and B) of each sample show a very similar pattern of temperature.

80
Table 5.10 Maximum temperature, calculated temperature parameters and
stability index (SI) . All the values are the average of six self-heating tests after
each pilot-scale test
Tests Ta Tmax Tmax-Ta Imax A72 Atotal tmax Duration of SI
o o o o o o
of C C C C/h C.h C.h day pilot-scale test
in days
2a 21.74 65.82 44.08 0.91 3545 18880 5.30 33 I
3a 19.95 60.29 40.34 0.94 3063 17733 4.92 29 I
3b 19.07 21.58 2.51 0.05 1602 7445 6.95 39 (68) V
4a 20.62 68.45 47.83 1.01 3123 15836 3.14 49 I
4b 21.91 33.73 11.82 0.16 1826 18736 6.89 21 (70) IV
5a 26.62 64.25 37.63 1.06 3594 21620 4.57 20 I
5b 24.95 46.17 21.22 0.39 2506 12855 6.19 21 III
5c 21.81 36.84 15.03 0.19 1652 15161 11.55 11 (52) V
6a 24.09 58.22 34.13 0.97 3471 13068 3.18 30 II
6b 22.77 36.85 14.08 0.29 2024 15936 8.55 21 (51) IV
7a 20.61 68.21 47.60 0.96 3414 21735 3.71 23 I
7b 20.52 54.52 34.00 0.78 3006 14713 3.41 32 (55) II
8a 21.24 60.28 39.04 0.90 3302 21056 3.74 28 I
8b 22.20 39.26 17.06 0.26 1670 20647 10.21 28 (56) IV
9a 22.19 54.09 31.90 0.76 2670 30635 10.34 24 III
9b 23.62 44.16 20.54 0.49 2719 18737 2.84 32 (56) III
Ta = ambient temperature; Tmax = maximum temperature; Imax = maximum temperature
increase; A72 = area under temperature curve after 72 hours; Atotal = area under temperature
curve after completion; tmax = time needed to reach Tmax, in days; in (), total duration of first,
second and third stages.

Table 5.11 Duration and different physical parameters of self-heating tests. All
the values are the average of six self-heating tests after each pilot-scale test
Tests Days of self- Massi MCi VSi TSi VSi VS g VS Porosityi
of heating test g % % g g % %
2a 16 494.28 64.50 92.98 174.79 162.51 37.36 22.60 55.01
3a 15 506.13 67.50 94.75 162.71 154.16 29.79 19.06 51.29
3b 15 414.70 65.33 91.54 136.83 125.28 -1.34 -1.07 62.23
4a 15 415.58 59.65 88.23 165.05 145.61 28.03 18.87 62.72
4b 28 396.60 62.93 89.91 146.46 131.66 5.75 4.39 64.19
5a 21 393.97 58.04 92.99 160.95 149.70 23.81 14.75 64.40
5b 20 335.43 57.28 91.77 142.75 130.97 7.47 5.78 69.84
5c 22 351.32 58.47 91.89 145.86 134.01 8.73 6.54 68.36
6a 12 335.00 57.91 93.69 140.98 132.01 14.77 11.22 69.74
6b 24 301.30 56.69 93.69 129.87 121.73 4.30 3.43 72.81
7a 20 417.22 59.38 92.37 169.55 156.77 29.55 18.57 62.33
7b 20 385.75 57.48 90.84 163.95 148.92 10.32 6.92 65.39
8a 21 401.98 55.78 91.70 177.79 163.14 25.81 15.44 63.94
8b 28 337.80 57.68 89.77 142.93 128.30 9.36 7.24 69.75
9a 30 386.60 57.86 95.33 162.80 155.18 46.77 30.10 64.95
9b 24 315.48 56.04 92.15 138.58 127.71 10.88 8.55 71.66
i = initial; MC = moisture content; VS = change in volatile solids.

81
where samples from first stage tests were used. In the self-heating tests after second
stage tests, only 3.4 to 8.5% VS with an average of 6.1% was degraded. The average
initial porosity was 62 to 73%, except in tests 2a and 3a where it amounted to 55 and
51%, respectively.

5.1.8.2 Temperature Decline

Final temperature drop coupled with higher O2 concentration in the spent air
represents completion of each pilot-scale test. All the first stage tests were
terminated, once the average temperature ranged between 30 to 50 oC. The reasons
of higher average end temperature about 45 oC in some of the first stage tests were
short test duration less than three weeks, alternate upflow/downflow aeration or
internal air recirculation. In the second stage tests were terminated at approximately
25 oC except those where reuse of hot spent air were applied. Although most of the
second stage tests terminated at 25 oC, some of them did not produce completely
stable end products as indicated in Table 5.10.

5.1.8.3 Color and Odor

The color of the end product (fresh compost) from the first stage composting of 20 to
30 days was yellowish brown to light brown, whereas the end product (mature
compost) of the second or third stage composting of another 30 days exhibited a
brown to dark brown color. The fresh compost emitted a moderately offensive odor
while mature compost after second stage had an earthy to slightly ammonia tinted
odor. After about one week of storage, the ammonia odor could not be perceived any
more in the mature compost.

5.1.8.4 C/N ratio

As mentioned above, according to Table 5.7 the C/N ratio of the initial waste
mixtures varied from 21 to 31 with an average of 25.9, while the C/N ratio in the
final compost after second or third stage pilot-scale tests ranged from 16.5 to 21.3
with an average of 19.1. This range of final C/N ratio is very close to 20, a value
suggested for maturity (Jimenez and Garcia 1989, Senesi 1989, Inbar et al. 1990).
82
Zucconi and Bertoldi (1987) recommended a final C/N ratio of 22 (cited by Villar et
al. 1993). On average the C/N ratio in different tests decreased by a value of 6.8. The
range of the ratio (final C/N)/( initial C/N) for the pilot-scale tests of maximum 70
days duration was 0.68 to 0.84 with an average of 0.76. These values are also close
to the suggested value of 0.75 after more than 120 days of composting period
(Jimenez and Garcia 1989).

5.1.8.5 pH

The final pH values of air dried samples after second stage tests varied between 5.7
and 6.5, which is the preferable range for most plants (Kuter 1995). The pH of the
final fresh samples varied between 7 to 8 which is comparable with the suggested
range 7.0 to 7.5 for mature compost produced from municipal solid waste (Canet and
Pomares 1995).

5.1.8.6 Fungus Growth

Fungus growth was observed after one day in the samples which were used to
measure the pH of air dried compost of different layers of one pilot-scale test. After
that observation, samples of different layers of tests 8a, 8b, 9a and 9b were prepared
(according to section 4.7.1.4 for pH determination of air dried sample) and then
incubated in an oven at 30 oC for 36 to 48 hours. After incubation, the different
intensities of fungus growth in the moist samples of test 8a (upflow aeration) are
clearly shown in Figure 5.53, by color difference from gray to brown. The compost
samples in the upper layers, which remained at higher temperature (average 50 oC)
and higher moisture content (70%) during the pilot-scale test, are of brown color
with SI = II (as calculated from self-heating test) and showed less fungus growth.
More fungus growth was obtained for the samples from the lower layers (gray color)
with SI = I which remained at lower temperature (average 26 oC) and moisture
content (22%) during the duration of the pilot-scale test. For test 9a (downflow
aeration), the fungus growth is shown in Figure 5.54. The compost samples of the
upper layers (gray color) with SI = II which remained at lower temperature

83
(average 35 oC) and lower moisture content (9%) during the pilot-scale test, showed
more fungus growth. In contrast, less growth was obtained for the samples from the
84
lower layers (brown color) with SI = III which remained at higher temperature
(average 53 oC) and moisture content (55%) during the pilot scale test. The raw,
uncomposted sample showed much fungus growth with yellowish color. No fungus
growth in the samples of second stage tests 8b and 9b was detected after 48 or even
72 hours and the color (dark brown) remained unchanged. Therefore, if the compost
is not biologically stabilized, there will be more growth and vice versa for the
stabilized compost. Although the fungus growth of the different layered samples
may have been affected by the temperature and moisture content at that layer during
the pilot-scale test, still by implication the overall treatment condition may be
qualitatively evaluated in a shorter period by applying incubation of the samples and
observation of fungus growth.

Bench-scale Tests

5.2.1 Temperature

Figures 5.55 and 5.56 show the comparison of temperature patterns of bench-scale
tests with natural aeration and forced aeration, respectively, under bench-scale run 1.
The first peaks obtained following a similar temperature pattern in all tests after 3.5
days and before raising to their second peaks, all the temperatures declined up to 30
to 40 oC until 5 to 7 days. After the second peaks, the temperatures followed a very
similar and regular pattern in both natural aerated tests 1NA1 and 1NA2 until the
end of the tests. The temperatures of forced aerated tests 1FA1 and 1FA2 followed a
different and more irregular pattern with several peaks after their second peaks. The
temperatures in tests 1FA1 and 1FA2 finally declined after 25 and 35 days,
respectively. As indicated in Table 5.12, the mean temperature of bench-scale tests
1NA1, 1NA2, 1FA1 and 1FA2 were 43.92, 46.72, 36.43 and 41.59 oC, respectively.
The areas under the temperature curves were 54300, 57750, 45040, and 51420 oC.h,
respectively. In test 1FA1 with a higher aeration rate, the mean temperature and the
area under the temperature curve was lower than in the other tests.
Temperature variations in bench-scale tests of run 2 with natural aeration and the
average temperature of corresponding pilot-scale test 9a are shown in Figure 5.57.

85
The test 2NA2 was terminated together with pilot-scale test 9a. Probably due to
temperature gradients along the height of the composting mass, the average
temperature curve of 9a stayed partly below the temperature curve of the bench-scale
tests. The overall mean temperatures of tests 2NA1, 2NA2 and 9a were 44.65, 48.11
and 44.73 oC, respectively, as illustrated in Tables 5.13 and 6.1. The areas under the
average temperature curves of test 9a and 2NA2 are comparable at 25720 and 26180
o
C.h, respectively.

Temperature variations in bench-scale tests of run 3 with natural aeration and


average temperature of pilot-scale test 9b are shown in Figure 5.58. The average
temperature curve of 9b also stayed below the temperature curves of the bench-scale
tests of run 3. The overall mean temperatures of tests 3NA1, 3NA2 and 9b were
38.24, 37.32 and 33.20 oC, respectively, as illustrated in Tables 5.13 and 6.1. The
areas under the average temperature curve of tests 3NA1, 3NA2 and 9b were 29330,
28620 and 25590 oC.h, respectively.

5.2.2 Physico-chemical Characteristics

Initial and final physico-chemical properties of waste and compost, masses and air
supplied in different bench-scale tests of run 1 are shown in Table 5.12. The duration
of all the tests was 52 days. Total mass reduction in tests 1FA1 and 1FA2 were
higher (43.53 and 47.59%) than in tests 1NA1 and 1NA2 (24.92 and 25.58%).
However, the %VS reduction range was 29.3 to 40 % in all tests with natural and
forced aeration except in test 1FA1 with 21.1 %. Higher aeration rate reduced the
final moisture content of the composting mass to 42.2% in test 1FA1, whereas the
final moisture content in tests 1NA1, 1NA2 and 1FA2 were 66.5, 60.5 and 52.5%,
respectively. Therefore, higher aeration (0.34 L/min.kg VS) in test 1FA1 removed
more moisture which dried the mass and thereby reduced the total VS degradation
and lead to lower temperature (Figure 5.56). Except for the irregular pattern of
temperature during forced aerated tests, the natural aerated tests 1NA1 and 1NA2
and lower forced aeration rate (0.21 L/min.kg VS) of test 1FA2 showed similar VS
degradation, mean temperature, and area under temperature curves. Therefore, it was

86
thought that natural aeration tests were suitable because of their simplicity for further
bench scale study.

From Table 5.13, the VS degradation in tests 2NA1 and 2NA2 after 52 and 22 days
were 39.93 (final MC was 65.4%) and 34.40% (final MC was 62.2%), respectively.
The VS degradation in corresponding pilot-scale test 9a was 24.72% (final was MC
39.0%) (Tables 5.3 and 5.5). Thus VS degradation in pilot-scale test 9a was 28%
lower than in bench-scale test 2NA2 after approximately same duration of 24 and 22
days, respectively. Therefore, the VS degradation in the first stage pilot-scale test 9a
is poorly related to the result of corresponding bench scale run 2 generally due to
higher moisture removal, probable higher moisture gradient and lower final moisture
content in the first stage tests (Table 5.3).

From Table 5.13, the VS degradation in tests 3NA1 and 3NA2 after 32 days were
16.66 (final MC was 61.2%) and 15.6% (final MC was 61.7%), respectively. The VS
degradation in the corresponding pilot-scale test 9b was 18.43% (final MC was
51.6%) (Table 5.5). Therefore, VS degradation in the second stage pilot-scale test 9b
was close to the result of the corresponding bench scale run 3. Due to lower moisture
removal, probably lower moisture gradient and higher final moisture content in the
second stage tests, it exhibited a slightly higher VS degradation (Table 5.3).

Masses and physico-chemical parameters of bench-scale tests of run 4, which was


conducted to investigate the effect of initial moisture content on VS degradation and
overall temperature, are summarized in Table 5.14. The tests 4NA1, 4NA2, 4NA3,
4NA4, 4NA5, 4NA6, 4NA7, 4NA8, 4NA9, 4NA10, 4NA11, and 4NA12 were
conducted with initial moisture content of 28.80, 35.10, 39.60, 43.80, 50.50, 52.70,
56.80, 59.30, 64.70, 69.92, 75.00, and 80.21%, respectively. The effects of initial
moisture content on VS degradation and overall temperature are discussed in
Chapter 6.

87
80

Natural aeration 1NA1


70
Natural aeration 1NA2

60
Temperature in oC

50

40

30

20 Ambient Temperature

10

0
0 5 10 15 20 25 30 35 40 45 50
Time in days

Figure 5.55 Temperature variations in bench-scale tests of run1 with natural


aeration.

80
Forced aeration 1FA1
70
Forced aeration 1FA2

60
Temperature in oC

50

40

30

20 Ambient Temperature

10

0
0 5 10 15 20 25 30 35 40 45 50
Time in days

Figure 5.56 Temperature variations in bench-scale tests of run1 with forced


aeration. Aeration rates of 0.34 and 0.21 L/min.kg VS were provided in tests FA1
and FA2, respectively.

88
80
Natural aeration 2NA1
70
Natural aeration 2NA2
Pilot-scale test 9a
60
Temperature in oC

50

40

30

20
Ambient Temperature
10

0
0 5 10 15 20 25 30 35 40 45 50
Time in days

Figure 5.57 Temperature variations in bench-scale tests of run2 with natural


aeration. A waste mixture similar as in pilot-scale test 9a was used.

80
Natural Aeration 3NA1
70
Natural Aeration 3NA2
Pilot-scale test 9b
60
Temperature in oC

50

40

30

20
Ambient Temperature

10

0
0 5 10 15 20 25 30 35 40 45 50
Time in days

Figure 5.58 Temperature variations in bench-scale tests of run3 with natural


aeration. A waste mixture similar as in pilot-scale test 9b was used.

Table 5.12 Initial and final physico-chemical parameters of waste and compost,
and amount of air supplied in different bench-scale tests of run 1. Duration of
89
all the tests was 52 days. Wet percentage of food waste, waste paper and
sawdust in initial mixture was 70, 15 and 15% respectively
Parameter Initial/Final 1NA1 1NA2 1FA1 1FA2
MC % Initial 58.74 58.74 58.74 58.74
Final 66.49 60.55 42.37 52.45
VS % Initial 95.00 95.00 95.00 95.00
Final 93.42 94.39 95.06 92.50
pH Initial 5.78 5.78 5.78 5.78
Final 6.10 5.65 5.33 5.44
OC % Initial 40.54 40.54 40.54 40.54
Final 37.06 41.84 42.87 40.69
TKN Initial 1.24 1.24 1.24 1.24
Final 1.43 1.32 1.67 1.72
C/N Initial 32.69 32.69 32.69 32.69
Final 25.92 31.70 25.67 23.66
Porosity Initial 68.45 68.46 71.59 64.33
Total mass, g Initial 384.40 384.30 331.70 437.50
Final 288.60 286.00 187.30 229.30
Change 95.80 98.30 144.40 208.20
% Change 24.92 25.58 43.53 47.59
H2O in g Initial 225.80 225.74 194.84 266.61
Final 191.89 173.17 79.36 120.27
Change 33.91 52.56 115.48 146.34
% Change 15.02 23.29 59.27 54.89
TS in g Initial 158.60 158.56 136.86 170.89
Final 96.71 112.83 107.94 109.03
Change 61.89 45.74 28.92 61.86
% Change 39.02 28.84 21.13 36.20
VS in g Initial 150.67 150.63 130.02 163.37
Final 90.35 106.50 102.61 100.85
Change 60.33 44.14 27.41 62.51
% Change 40.04 29.30 21.08 38.27
Air flow L/min.kg VS - - 0.34 0.21
Air supplied L - - 3421 2538
O2 supplied mg O2/gVS/hr - - 5.79 3.41
Temperature Mean 43.92 46.72 36.43 41.59
Maximum 56.88 61.13 59.74 67.63
o
C*hour 54300 57750 45040 51420

Table 5.13 Initial and final physico-chemical parameters of waste and compost
in bench-scale tests of run 2 and run 3.

90
Initial/ Run 2 Run 3
Parameter Final 2NA1 2NA2 3NA1 3NA2
Duration, d 52 22 32 32
MC % Initial 55.38 55.38 59.77 59.77
Final 65.44 62.19 61.18 61.73
VS% Initial 96.09 96.09 94.04 94.04
Final 94.24 93.18 90.46 93.21
Total mass, g Initial 432.60 391.70 386.70 347.20
Final 338.60 309.50 372.70 333.60
Change 94.00 82.20 39.50 39.10
% Change 21.72 20.98 10.21 10.49
H2O in g Initial 241.52 218.69 231.13 222.76
Final 221.57 192.46 212.42 205.92
Change 19.95 26.22 18.71 16.85
% Change 8.26 11.99 8.10 7.56
TS in g Initial 191.00 173.00 156.00 150.00
Final 117.00 117.00 135.00 128.00
Change 74.05 55.98 20.79 22.25
% Change 38.75 32.35 13.36 14.84
VS in g Initial 184.00 166.00 146.00 122.00
Final 110.00 109.00 141.00 119.00
Change 73.32 57.19 24.37 21.99
% Change 39.93 34.40 16.66 15.60
Temperature Mean 44.65 48.11 38.24 37.32
Maximum 61.01 63.17 55.85 60.30
o
C*hour 55720 26180 29330 28620

Table 5.14. Initial and final physico-chemical parameters of waste and compost
in bench-scale tests of run 4.
Tests Massi Massf Mass MCi MCf VSf VS VS Tmax
o
g g g % % % g % C
4NA1 183 173 10.00 28.80 27.93 90.86 5.28 4.46 59.23
4NA2 233 219 14.00 35.10 34.37 90.69 7.23 5.26 57.30
4NA3 256 241 15.10 39.60 40.15 89.97 10.98 7.80 53.73
4NA4 260 244 16.40 43.80 48.74 90.12 20.42 15.36 53.57
4NA5 337 308 28.80 50.50 51.40 90.80 15.78 10.39 55.59
4NA6 322 302 19.80 52.70 56.24 89.83 19.80 14.28 48.50
4NA7 369 347 22.20 56.80 59.01 89.87 17.29 11.93 49.71
4NA8 385 362 22.80 59.30 62.72 89.50 21.71 15.24 49.09
4NA9 448 427 21.00 64.70 68.27 89.90 22.10 15.35 45.64
4NA10 554 524 30.10 69.92 71.91 88.89 20.79 13.71 54.40
4NA11 722 694 28.00 75.00 77.46 90.06 23.37 14.24 43.69
4NA12 904 853 50.50 80.21 80.92 90.74 15.05 9.25 29.31
Massi = initial mass; Massf = final mass

91
CHAPTER 6

DISCUSSION AND INTERPRETATION OF


RESULTS

6.1 Pilot-scale Tests

6.1.1 Ambient conditions

The average values of the recorded hourly ambient temperature and relative
humidity are presented in Table 6.1. The range of average ambient temperature
during the pilot-scale tests was 19.9 to 26.7 oC with an average of 23.4 oC. The
range of average relative humidity during the pilot-scale tests was 55.2 to 70.9%
with an average of 62.9%. These ambient conditions reflected the Environmental
Engineering Laboratory condition. It should be noted that the bench scale tests were
also performed under these ambient conditions.

6.1.2 Temperature

Maximum and average temperature of the pilot-scale tests are presented in Table 6.1.
The range of the overall average temperature (average of hourly monitored
temperature of six layers of the composting mass over the whole duration) was 37.2
to 48.6 oC for first stage tests (mean of overall average 44.3 oC) and 31.0 to 40.7 oC
for second stage tests (mean of overall average 34.4 oC). The range of the maximum
temperature was 60.0 to 71.0 oC for first stage tests (mean of overall average 65.4
o
C) and 64.7 to 74.1 oC for second stage tests (mean of overall average 70.8 oC).
Therefore, the average temperatures and maximum temperatures of the composting

117
mass in second stage tests were, respectively, 10 oC lower and 5 oC higher than that
of first stage tests.

6.1.3 Air flow and Air Pressure

As mentioned in section 4.3.1, the air flow rate was maintained between 17 to 20
L/min (equivalent to about 0.67 L/min per kg initial VS) during the first two days.
This was followed by a continuous or intermittent air flow rate of 8 to 10 L/min
depending on the maximum temperature inside the pilot-scale reactor. The average
pressure of inflow air in the flow measuring pipe for different tests is shown in Table
6.2. The range of average pressure of inflow air during the pilot-scale tests was 0.0
to 4.3 cm of water with an average of 2.46 cm of water. The lower pressure was
obtained during the latter part of the tests, when minimum air flow was needed to
keep conditions aerobic inside the reactors.

6.1.4 Estimation of Initial Biodegradable Volatile Solids

Since the volatile solids originating from paper and sawdust were considered less
readily biodegradable due to their high cellulosic content (Koenig and Bari 1998),
the determination of the biodegradable fraction of VS or biodegradable volatile
solids (BVS) would give a better understanding of the waste degradability and final
product stability (Fernandes and Sartaj 1997). The term biodegradable volatile solids
(BVS) is generally used when referring to the degradation rates or oxygen
consumption rates (Haug 1993) The total initial biodegradable volatile solids (BVS)
for different pilot-scale tests were calculated from the sum of VS degraded in the
first stage, second stage and final self-heating test after second stage. It was thought
that the remaining BVS in the compost, after 70 to 90 days of effective degradation
in the first stage, second stage and in self-heating test, would not be more than 10%.
Therefore, the sum of VS degraded is increased by 10% to have a suitable estimation
of initial BVS. The calculation procedure and the estimated initial amounts of BVS
in the different pilot-scale tests are presented in Table 6.3.

118
Table 6.1 Maximum and average temperature in the composting mass, ambient
temperature and relative humidity
a
Test Days Overall Tmax of Tmax in the Tave.time, Tambient Relative
Tave, oC Tave, oC whole mass, oC o
C.hour , oC humidity, %
1a 31 44.26 60.89 67.00 32890 26.20 70.90
1b 21 39.20 58.30 71.90 20590 26.00 68.50
2a 33 48.63 66.96 71.00 38230 24.60 65.00
3a 29 39.00 56.05 64.90 25940 24.60 65.00
3b 39 31.92 62.41 73.10 29790 21.00 60.00
4a 49 48.00 53.98 60.80 56310 20.30 60.50
4b 21 31.55 52.65 69.34 15870 21.40 62.20
5a 20 48.56 57.34 62.50 23110 25.10 63.10
5b 21 40.77 65.10 74.10 20350 26.70 63.30
5c 11 30.56 51.30 60.00 10300 24.80 62.20
6a 30 43.55 55.66 62.95 31310 26.30 63.60
6b 21 32.72 52.00 69.40 16490 23.10 55.20
7a 23 44.40 65.40 70.20 24600 22.50 58.60
7b 32 32.90 70.14 73.70 25200 20.90 57.90
8a 28 37.20 55.28 68.80 24900 21.00 58.30
8b 28 33.00 55.30 70.50 22420 21.60 67.00
9a 24 44.73 54.50 60.00 25720 19.90 67.00
9b 32 33.20 55.40 64.70 25590 24.50 63.00
a
Maximum of hourly average temperature calculated from 6 layers

Table 6.2 Air supply and oxygen consumption in pilot-scale composting tests
b
Test Total air supplied Air Percentage O2 Total O2 Total kg O2
pressure in the spent air supplied O2 used/kg
fresh spent m3/kg in cm of min. ave. in kg used VS
m3 m3 VSa water in kg degraded
1a 358 - 12.01 1.98 10.8 17.63 95.83 14.12 1.28
1b 185 - 9.70 1.64 13.3 18.41 49.44 5.47 1.26
2a 359 - 13.55 4.22 16.2 18.56 96.43 9.90 1.35
3a 195 - 7.80 2.25 9.3 17.45 52.44 8.31 1.31
3b 132 175 7.75 2.95 16.2 19.91 73.69 3.82 1.05
4a 325 - 12.72 2.18 14.0 16.60 87.83 16.92 1.37
4b 121 - 6.37 1.17 12.7 19.15 32.70 2.57 1.78
5a 179 - 7.50 2.83 13.8 16.69 48.18 7.23 1.28
5b 70 109 4.03 4.31 13.4 18.82 44.41 3.61 1.48
5c 25 - 2.02 1.48 17.1 19.70 6.85 0.47 1.38
6a 307 - 12.59 2.87 10.0 17.90 82.52 11.59 1.33
6b 96 - 6.84 1.41 10.1 19.29 26.07 3.04 1.68
7a 198 - 8.19 3.97 15.3 17.46 53.76 8.85 1.22
7b 1 99 6.50 2.29 8.9 15.86 23.30 2.99 1.24
8a 190 - 7.43 2.33 13.9 17.30 51.65 10.28 1.35
8b 57 128 3.55 2.34 11.7 17.87 44.78 3.76 2.06
9a 180 - 7.02 3.13 14.7 17.23 48.91 8.51 1.33
9b 129 - 8.25 0.97 13.8 19.32 35.06 4.11 1.41
a
Initial VS at the beginning of each stage
b
Average of hourly values over the whole duration of the test

119
Table 6.3 Estimated values of initial biodegradable volatile solids (BVS) in the
pilot-scale runs. The BVS were estimated as 1.1 times the sum of VS degraded
in the first stage pilot-scale test (ST1), second stage pilot-scale test (ST2) and in
final self-heating test (SH)
Run Initial Initial Food Duration in days VS degraded in kg Estimated
TS, VS, VSi, ST1 ST2 SH Total ST1 ST2 SH Total BVS, 1.1 times
kg kg kg of total VS
1 30.80 29.79 12.82 31 21 - 52 11.0 3.91 - 15.34 16.87
3
2 27.99 26.48 11.32 33 - 16 49 7.33 - 3.57 10.90 12.39
3 26.18 25.01 13.68 29 39 15 83 6.34 3.98 0.00 10.32 11.35
4 27.38 25.54 15.59 49 21 28 98 12.3 2.80 0.46 15.62 17.18
6
5 25.09 23.86 11.33 20 32 22 74 5.68 2.95 0.99 9.62 10.58
6 25.48 24.37 11.28 30 21 24 75 8.49 1.97 0.48 10.94 12.03
7 25.64 24.16 11.41 23 32 20 75 7.16 2.44 1.00 10.60 11.66
8 27.13 25.57 12.10 28 28 28 84 7.49 2.02 1.16 10.67 11.74
9 26.65 25.61 12.27 24 30 24 78 6.33 3.55 1.12 11.00 12.10

6.1.5 Effect of Aeration modes

6.1.5.1 Temperature

Temperature distribution over time and height of selected pilot-scale composting


tests under different aeration modes are shown in Figures 6.1 to 6.8. These figures
illustrate daily temperature variations along the height of the composting mass.
Temperatures within 10 oC intervals from 15 to 75 oC are indicated by different
colors. Figures 6.1 to 6.4 show the temperature distribution over time and height for
upflow, upflow with internal air recirculation, downflow and upflow/downflow
aeration, respectively, in the first stage tests. Comparing Figures 6.1 and 6.2 for tests
8a and 5a, most of the time the mass in test 5a was above 35 oC and in test 8a was
below 35 oC. The temperature in the upper layers of the composting mass of test 5a
seldom exceeded 60 oC, whereas in test 8a about 40% of the mass remained in the
range of 65 to 75 oC for 2 days which is considered harmful to the microorganisms
responsible for biological degradation as discussed in detail in section 2.6.5.
Therefore upflow aeration with internal air recirculation can more effectively
maintain the optimum temperature inside the composting mass as shown in test 5a.
In test 9a downflow aeration exhibited a completely opposite profile to test 8a as
shown in Figure 6.3, with a comparatively better temperature distribution.
120
In test 7a with alternate upflow/downflow, the temperature distribution throughout
the height of the composting mass was more uniform than in the other tests. Higher
temperature regions (over 45 oC) were created periodically for every 3 to 4 days at
the top and bottom layers following alternate upflow/downflow aeration as shown in
Figure 6.4. The higher temperature region covered 40% mass from the top during
upflow aeration and 60% mass from the bottom during downflow aeration.
Therefore, alternate upflow/downflow aeration showed overall better control over
degradation and sanitization. The downflow aeration could remove less heat than
upflow aeration as demonstrated in the cases of tests 7a (during down flow) and 9a.

Figures 6.5 to 6.8 show the temperature distribution over time and height for second
stage tests 9b (upflow), 6b (downflow), 5b (upflow/downflow and reuse of spent air
from test 6a) and 7b (upflow and reuse of spent air from test 8a), respectively. All
second stage tests remained in thermophilic temperature for a shorter period of time
as compared to first stage tests which implied less microbial activity or degradation.
Normally all second stage tests cooled down to ambient temperature after 10 to 12
days. Tests 9b and 6b showed completely opposite patterns as expected for upflow
and downflow aeration as shown in Figures 6.5 and 6.6, respectively. In test 5b,
reuse of spent air from test 6a (as a two reactor system in series connection) was
applied after 10 days, which again boosted up the temperature from 32 oC to 50 oC
with uniform temperature distribution throughout the height as shown in Figures 6.7
and 5.9.

The vertical mean temperature profiles of different first stage tests are shown in
Figures 6.9 and 6.10. The vertical mean temperatures are determined as the mean of
the hourly temperature values at that point throughout the duration of the test. The
mean temperature profiles of tests 2a and 9a (downflow aeration only) and test 8a
(upflow aeration only) exhibited the most extreme variations with height, but
completely opposite trends. The highest mean temperatures observed were 63 oC and
55 oC in the bottom layer of tests 2a and 9a, respectively, and 47 oC in the top layer
of test 8a. More uniform temperature profiles were achieved by internal recirculation

121
110

Height of composting mass in cm


90

70 65-75
55-65
45-55
50 35-45
25-35
15-25
30

10
0 2 4 6 8 10 12 14 16 18 20 22 24
Time in days

Figure 6.1. Temperature distribution over time and height of first stage test 8a with
upflow aeration mode. Test duration was 28 days.

110

Height of composting mass in cm


90

70
65-75
55-65
45-55
50 35-45
25-35
15-25
30

10
0 2 4 6 8 10 12 14 16 18 20 22 24
Time in days

Figure 6.2. Temperature distribution over time and height of first stage test 5a,
upflow aeration mode with internal air recirculation (r = 20 to 80%). Test duration
was 19.8 days.

122
110

Height of composting mass in cm


90

70
65-75
55-65
45-55
50 35-45
25-35
15-25
30

10
0 2 4 6 8 10 12 14 16 18 20 22 24
Time in days

Figure 6.3. Temperature distribution over time and height of first stage test 9a with
downflow aeration mode. Test duration was 24 days.

110

Height of composting mass, cm


90

65-75
70 55-65
45-55
35-45
50 25-35
15-25

30

10
0 2 4 6 8 10 12 14 16 18 20 22 24
Time in days

Figure 6.4. Temperature distribution over time and height of first stage test 7a with
alternate upflow and downflow aeration mode. Test duration was 23 days.

123
110

Height of composting mass in cm


90

70 65-75
55-65
45-55
50 35-45
25-35
15-25
30

10
0 2 4 6 8 10 12 14 16 18 20 22 24
Time in days

Figure 6.5. Temperature distribution over time and height of second stage test 9b
with upflow aeration mode. Test duration was 32 days.

110

Height of composting mass in cm


90

70
65-75
55-65
45-55
50 35-45
25-35
15-25
30

10
0 2 4 6 8 10 12 14 16 18 20 22 24
Time in days

Figure 6.6. Temperature distribution over time and height of second stage test 6b
with downflow aeration mode. Test duration was 21 days.

124
110

Height of composting mass in cm


90

70
65-75
55-65
45-55
50 35-45
25-35
15-25
30

10
0 2 4 6 8 10 12 14 16 18 20 22 24
Time in days

Figure 6.7. Temperature distribution over time and height of second stage test 5b,
upflow and downflow aeration for 10 days and after that reuse of spent air from 6a.
Test duration was 21 days.

110

Height of composting mass in cm


90

70
65-75
55-65
45-55
50 35-45
25-35
15-25
30

10
0 2 4 6 8 10 12 14 16 18 20 22 24
Time in days

Figure 6.8. Temperature distribution over time and height of second stage test 7b,
upflow aeration for 1 day and after that reuse of spent air from 8a. Test duration was
32 days.

125
of air as shown in Figure 6.9 for tests 1a, 3a and 5a. The highest mean temperatures
in the top layers of 1a, 3a and 5a were 56 oC, 47 oC and 53 oC, respectively. For
tests 4a, 6a and 7a, where alternatively upflow and downflow aeration was applied,
the mean temperature was highest in the middle layers with values ranging from 47
o
C to 54 oC. Therefore, intermittent reversal of upflow and downflow aeration

125
increased the overall mean temperature as compared to continuous unidirectional
flow aeration, with the highest temperature found in the middle layers.

The mean temperature profiles of the different second stage tests are shown in
Figures 6.11 and 6.12. The effect of different aeration modes is less pronounced in
the second stage tests than in the first stage tests. However, the temperature profiles
of test 6b (downflow aeration only) and test 9b (upflow aeration only) also exhibit
the most extreme variations with opposite trends similar to first stage tests 2a and 8a.
The mean temperature profiles of other second stage composting tests with modified
aeration show more uniform temperature distribution.

The most uniform mean temperature profile was obtained by reuse of spent air from
first stage tests as demonstrated by the profiles of tests 5b and 8b. The profiles for
these tests after series connection, as shown in Figure 6.12, are almost vertical mean
uniform temperature distributions. Although series connection was also applied in
test 7b, the temperature distribution was not uniform. In tests 5b and 8b the series
connection was applied after 9 and 10 days, respectively, once the thermophilic
temperature of the composting mass cooled near to ambient. On the other hand, the
series connection in test 7b was applied one day after the start of the second stage.
This connection quickly elevated the temperatures of all the layers of test 7b above
60 oC, but it also cooled down sharply after 5 to 6 days as shown in Figures 5.13 and
6.8. Thereafter it did not produce any second peak in temperature like the other
similar tests 5b and 8b. This rapid temperature rise in all layers of test 7b may have
inactivated most of the microorganisms responsible for degradation.

126
120

Height of composting mass in cm 100


1a, upflow (r=20-50%)
2a, downflow
80 3a, upflow (r=100-200%)
4a, upflow and downflow

60 5a, upflow (r=20-80%)

40

20

0
20 30 40 50 60 70
Temperature in oC

Figure 6.9 Mean temperature profile in the composting mass over the whole
duration of different first stage composting tests

120
Height of composting mass in cm

100

6a, upflow/downflow
80
7a, upflow/downflow
8a, upflow
60 9a, downflow

40

20

0
20 30 40 50 60 70
Temperature in oC

Figure 6.10 Mean temperature profile in the composting mass over the whole
duration of different first stage composting tests

127
120

Height of composting mass in cm 100 1b, upflow and downflow


3b, series connection
4b, upflow/downflow
80
6b, downflow
7b, series connection
60 9b, upflow

40

20

0
20 30 40 50 60 70
Temperature in oC

Figure 6.11 Mean temperature profile in the composting mass over the whole
duration of different second stage composting tests

120
Height of composting mass in cm

100 5b, over all


5b, before series connection
5b, after series connection
80 8b, over all
8b, before series connection

60 8b, after series connection

40

20

0
20 30 40 50 60 70
Temperature in oC

Figure 6.12 Mean temperature profile in the composting mass over the whole
duration before and after the series connection of different second stage composting
tests

128
The vertical temperature gradient in first stage composting was highest for
unidirectional airflow either in the upflow or downflow mode, with observed values
of 6.5 oC to 7.5 oC per 25 cm height. For comparison, Hogan et al. (1989) reported
results of 16 oC and 21 oC per 25 cm in insulated closed reactors with and without
heat conduction control, respectively. With internal air recirculation, the temperature
gradient was reduced to between 3.8 oC and 4.0 oC per 25 cm, thus demonstrating
the beneficial effect of air recirculation on a more uniform temperature distribution.
The second stage composting tests exhibited a lower temperature gradient than the
first stage composting tests and ranged between 1.75 oC and 2.25 oC per 25 cm
which could be attributed to the relatively lower O2 consumption rates during the
second stage. However, reuse of spent air in the second stage appeared to achieve a
more uniform temperature distribution and lead to an increase in the mean
temperature, thereby accelerating degradation of the remaining organic matter.

Internal air recirculation was applied only with upflow aeration in the first stage tests
1a, 3a and 5a. More uniform distribution of temperature was achieved by internal
recirculation of air in tests 1a, 3a and 5a than in tests 2a, 8a and 9a which were
conventionally aerated. Recirculation ratios also affect the overall mean temperature.
The overall mean temperature in test 5a (r = 20 to 80%) was always 8 to 10 oC
higher than in test 3a (r = 100 to 200%), whereas a less uniform distribution of
temperature was observed in test 1a as compared to tests 3a and 5a, due to a lower
recirculation ratio of r = 20 to 50%. Therefore, the recirculation ratio r = 20 to 80%
exhibited a better temperature distribution than the higher r of 100 to 200% and the
lower r of 25 to 50%.

6.1.5.2 Oxygen Consumption Rate

The specific oxygen consumption rates, in mg O2/g VSinitial-h, of the first stage tests
are shown in Figures 6.13 to 6.14. The maximum specific O2 consumption rates in
the first stage tests varied between 0.8 to 2.2 mg O2/g VSinitial-h. The specific O2
consumption rates were increasing and decreasing fastest, covering a large area
under the curves in tests 3a and 5a which applied internal air recirculation, as

129
compared to tests 2a (downflow) and KT (upflow, conducted by Koenig and Tao
1996 using similar waste material).

In test 7a, the O2 consumption rate was also increasing very fast, but produced
several peaks due to alternate air flow reversal. Caution is needed, however, in
interpreting these values, since the VS content decreased over time, which would
have resulted in higher specific O2 consumption rates towards the end of each test.
Although test 8a showed the highest initial O2 consumption rate (due to more milk
contained in the food waste), it did not produce a more stable end product, but
produced another high peak in O2 consumption rates in the second stage test. After
15 to 20 days, the specific O2 consumption rate had decreased to a very low,
relatively stable value in all tests, indicating the completion of the first stage of
composting. If the VS of the food waste is considered as the sole biodegradable
fraction of the waste mixture and taken as the basis for determination of the specific
O2 consumption rate, then the rates were almost doubled. On this basis the maximum
specific O2 consumption rates varied from 2 to 3.75 mg O2/g VSfood-initial-h as shown
in Figure 6.16.

As indicated in Table 5.10, all end products after first stage tests were found to have
stability index I according to the self-heating test and therefore had to be considered
biologically unstable. However, the degree of the instability of these unstable end
products, under stability index I, can be indicated by the maximum specific O2
consumption rates during their respective second stage tests. Therefore, a higher
value of maximum specific O2 consumption rate in the second stage for a similar
type end product of the first stage would indicate a higher degree of instability.
Similarly, a lower value of the maximum specific O2 consumption rate would
indicate a lower degree of instability. The specific O2 consumption rates of some
typical second stage tests are shown in Figure 6.15. The maximum specific O2
consumption rates in the second stage tests varied between 1.75 to 3.25 mg O2/g
VSinitial-h. The specific O2 consumption rates were increasing and decreasing faster
in all second stage tests and most of the time they followed a very similar pattern.
After a comparatively shorter period of 8 to 10 days, the specific O2 consumption

130
rate had decreased to a very low and stable value in all these tests, indicating thus the
completion of the second stage of composting.

The maximum specific O2 consumption rates of 2.75 and 3.75 mg O2/g VSinitial-h
were observed in tests 8b and 9b, which are almost double than those of the other
second stage tests. This indicates that the initial composts for the second stage tests
8b and 9b were highly unstable as compared to those of the other second stage tests.
In fact the initial composts of these second stage tests were the end products of first
stage tests 8a and 9a, respectively. Furthermore, in tests 8a and 9a only conventional
upflow and downflow aeration were applied, respectively. Therefore it is clear that
the end products of these conventionally aerated tests 8a and 9a were most unstable
as they showed the highest maximum specific O2 consumption rate in Figure 6.15,
although test 8a showed a higher initial O2 consumption rate.

The O2 consumption rates, in kg/day, and the cumulative O2 consumption, in kg,


over the duration of the different tests are presented in Figures 5.22 to 5.38. The O2
consumption rate, in kg/day, is proportional to the specific O2 consumption rate by
the factor 0.001x24xkg VSinitial. It was found that the O2 consumption per unit of VS
degraded varied between 1.05 to 2.06 kg O2/kg VSdegraded, with a mean value of
1.35. This value agrees with common experience as previously discussed in section
3.1.

Table 6.2 presents a summary of the results on air and oxygen consumption. About
12.6% of the O2 in the supplied air was used for degradation of VS in the
composting tests, with a range between 7% and 19%. Obviously, a higher percentage
of O2 was consumed at the beginning of each test as well as when internal air
recirculation and reuse of spent air where employed. It should be noted, that the
amount of air supplied depends not only on the theoretical O2 requirement, but also
on its capacity to remove excess heat and water; therefore, total air supply exceeded
the theoretical air requirement for O2 by a factor of about 8. In tests 3b, 5b, 7b and
8b, about 57 to 99% of the supplied air was reused air as calculated from Table 6.2.
Therefore, it appears that internal air recirculation and reuse of spent air were able to
lower the fresh air required under the set rule for control of air flow rate.
131
2.5

2a
3a
2.0
5a
KT
mg O2/g.VSinitial-h

1.5

1.0

0.5

0.0

0 5 10 15 20 25 30
Time in days

Figure 6.13 Specific oxygen consumption rate during first stage composting tests, in
mg O2/g.VSinitial-h. KT is the upflow pilot scale test (Koenig and Tao 1996).

2.5

7a
8a
2.0
9a
mg O2/g.VSinitial-h

1.5

1.0

0.5

0.0

0 5 10 15 20 25 30
Time in days

Figure 6.14 Specific oxygen consumption rate during first stage composting tests, in
mg O2/g.VSinitial-h.

132
3.5

3b
3.0 7b
8b
2.5 9b
mg O2/g.VSinitial-h

2.0

1.5

1.0

0.5

0.0

0 5 10 15 20 25 30
Time in days

Figure 6.15 Specific oxygen consumption rate during second stage composting tests,
in mg O2/g.VSinitial-h.

4.0

2a
3.5
3a
7a
3.0
mg O2/g.VSfood initial-h

KT

2.5

2.0

1.5

1.0

0.5

0.0

0 5 10 15 20 25 30
Time in days

Figure 6.16 Specific oxygen consumption rate during first stage composting tests, in
mg O2/g.VSfood initial-h.
6.1.5.3 Total Air Consumption
133
Airflow was controlled by temperature feedback in order to maintain the temperature
in the composting mass within the range of 55 oC to 60 oC, with or without any
modifications in the mode of aeration. As calculated from Tables 5.5 and 6.2 on the
basis of initial VS, the total amount of fresh air consumed for tests 3a+3b, 5a+5b+5c,
and 8a+8b which were operated as two-reactor systems for the second stage, was
13.08, 11.47 and 11.43 m3/kg VSinitial producing a final product with stability index
V, V and IV, respectively. For the single reactor tests 4a+4b, and tests 6a+6b
however, total air consumptions were found to be much higher with 17.45 and 16.52
m3/kg VSinitial, while the final product of both these tests attained only stability index
IV. The single reactor system needs about 1.5 times more air than the two-reactor
system; in other words, to process waste containing one ton of VS in the single
reactor composting system requires 5400 m3 more fresh air. Since the two-reactor
system needs less air (average 12.0 m3/kg VSinitial) to produce a completely stable
compost, it also uses less energy for aeration and is therefore more economical. The
total air consumption in all tests appears quite economical because in some cases the
air consumption in the composting process has been reported as 37 to 62 m3/kg TS
(Haug 1993).

6.1.5.4 Moisture Content

Although the average moisture content of the final product appeared to be just
sufficient for microbial activity, especially in the case of unidirectional aeration
composting, the situation is completely different if the distribution of moisture
content is analysed on a layer basis. Final distribution of moisture content
throughout the composting mass (along the height of the pilot-scale reactors) at the
end of different first and second stage tests are shown in Figures 6.17 to 6.20. Final
moisture distributions were calculated from the moisture content of each layer at the
end of the experiments.

The distribution of moisture content of tests 2a and 9a (downflow aeration only) and
test 8a (upflow aeration only) exhibits the most extreme variations with height, but at
completely opposite trends. The lowest moisture contents observed were 11.75% and
6.41% in the top layer of tests 2a and 9a, respectively, and 18.51% in the bottom
134
layer of test 8a. The highest moisture contents were observed in the extreme opposite
layers of these respective tests in the range between 62 to 74%. At the end of the
tests, the composts from the upper layers for downflow aeration test and from the
lower layers for upflow aeration tests were not only extremely dry and inhibitive to
microbial activity as discussed in section 2.6.3, but also biologically unstable as
discussed in the next section 6.1.4.7. Therefore, the dry and mixed compost
produced in an unidirectional aerated static pile process does not provide any
guarantee for biological stability or maturity.

More uniform moisture distributions were achieved in first stage tests 6a and 7a by
air flow reversal, that is alternate upflow/downflow aeration, and addition of
moisture as shown in Figure 6.18. A total of 8.73 and 15.21 L of water were added
during the whole duration of tests 6a and 7a. The moisture content was highest in the
top layers with a value of 70% for both these tests. The moisture content in all the
layers stayed above 57% for test 6a and 42% for test 7a. Therefore, intermittent
reversal of aeration mode and addition of water during the tests increased the
uniformity in moisture distribution throughout the composting mass of the first stage
tests, as compared to upflow aeration with internal air recirculation and continuous
unidirectional flow aeration.

The air recirculation helped to a attain more uniform temperature distribution


throughout the composting mass as discussed in section 6.1.4.1. Similarly, the effect
of air recirculation on moisture distribution is more pronounced at the bottom layers.
In tests 3a and 5a, the moisture content in the first layers (20 cm thick from bottom)
were higher than in the second layers as illustrated in Figure 6.17. The subsequent
layers of these tests were drier due to higher air recirculation ratios than in test 1a. In
test 1a, the lower recirculation rate (r = 20 to 50%) did not improve the moisture
content in the lower layers as indicated by the final moisture distribution curve.
However, a more favorable overall distribution in the moisture content may remain
during the first 2 to 3 weeks of composting under internal air recirculation as

135
120

Height of composting mass in cm 100

80

60
1a
2a
40 3a
4a

20 5a

0
0 10 20 30 40 50 60 70 80
Moisture content (%)

Figure 6.17 Final moisture distribution throughout the composting mass at the end
of different first stage composting tests.

120
Height of composting mass in cm

100

80

60

40 6a
7a
8a
20
9a

0
0 10 20 30 40 50 60 70 80
Moisture content (%)

Figure 6.18 Final moisture distribution throughout the composting mass at the end
of different first stage composting tests.

136
120

Height of composting mass in cm 100

80

60

40 1b
3b

20 4b
5b

0
0 10 20 30 40 50 60 70 80
Moisture content (%)

Figure 6.19 Final moisture distribution throughout the composting mass at the end
of different second stage composting tests.

120
Height of composting mass in cm

100

80

60

40 6b
7b
8b
20
9b

0
0 10 20 30 40 50 60 70 80
Moisture content (%)

Figure 6.20 Final moisture distribution throughout the composting mass at the end
of different second stage composting tests.

137
indicated by test 5a (r = 20 to 80%). This test of 20 days duration showed the best
temperature distribution over time and comparatively more uniform final moisture
distribution than tests 1a and 3a. Therefore, a maximum period of 20 days for the
first stage composting with internal air recirculation is adequate. If it continues more
than 20 days, it will produce very dry compost with very little subsequent
degradation due to lack of moisture as indicated in the respective O2 consumption
curve, unless irrigation is applied.

The effect of different aeration modes is more pronounced on the moisture


distributions than on the temperature profiles for the second stage tests. However,
the moisture distributions of test 6b (downflow aeration only) and test 9b (upflow
aeration only) also exhibit the most extreme variations with opposite trends, similar
to the first stage tests 2a and 8a. Similarly as in the first stage tests, more uniform
moisture distributions were also achieved in those second stage tests where alternate
upflow/downflow aeration was applied as shown in Figures 6.19 and 6.20 for tests
4b and 8b. The ranges of moisture contents throughout the composting mass were 60
to 65% (average 63%) for test 4b and 50 to 60% (average 55%) for test 8b.
Therefore, intermittent reversal of upflow/downflow aeration increased the
uniformity in moisture distribution also in the second stage tests.

Reuse of spent air showed a more positive effect on the moisture distribution than
internal air recirculation, although reuse of spent air can only be applied in second
stage tests. In tests 3b, 5b and 8b alternate upflow/downflow aeration was applied
for 17, 9 and 10 days, respectively, before series connection for reuse of spent air
was used. More uniform distributions in moisture content were observed in tests 3b
and 8b due to prolonged reuse of spent air for 21 and 17 days, respectively, than in
test 5b with 10 days only. However, the initial moisture content (55%) in test 5b was
less than in the other tests and the initial material was composted for 20 days only in
the first stage test (5a), which may have required more air before series connection.
In test 7b, the air flow was always upflow and reuse of spent air started one day after
its start. The distribution was less uniform in this test than in tests 3b, 5b and 8b.
Furthermore, the series connection just after the start of the second stage test as in
test 7b affected adversely not only the moisture distribution, but also the temperature
138
pattern (discussed in section 6.1.4.1) and subsequent final product stability
(discussed in section 6.1.4.7). Therefore, reuse of spent air should be applied to the
second stage only 8 to 10 days after its start or after declination of the thermophilic
temperature and then it should be continued for at least another 10 days.

It was observed that the patterns of both mean temperature gradients and final
moisture distributions in the final composting mass for a particular test were quite
similar. Usually the lower mean temperature occurs in those layers which had lower
moisture content at the end. Lower mean temperatures lead to subsequent lower
biological decomposition, which is discussed in section 6.1.5.7.

6.1.5.5 pH and Conductivity

The distribution of pH of final samples throughout the height of the composting


mass of first and second stage tests is shown in Figures 6.21 to 6.24. Dry samples
were used to determine the pH of the compost of all the tests and in addition fresh
samples were also used to determine the pH for few tests (section 4.7.1.3 and
4.7.1.4). The distribution of pH determined from the dried sample of different tests
were mostly vertically uniform with little variation. The ranges of pH for first and
second stage tests were 5.25 to 6.25 and 5.50 to 6.50, respectively. However, the
highest variations were observed in the distribution of pH determined from fresh
samples. Comparing the pH distributions of dry and fresh samples of test 8a
(upflow) in Figure 6.22, the pH in the top and bottom layers for dry samples was 6.1
and 5.5, respectively, and for fresh samples 7.75 and 5.00, respectively, which
showed a big variation among the layers. Similar but opposite distributions were also
observed between dry and fresh samples in test 9a (downflow). In test 5a (upflow
with internal air recirculation), the distributions of pH of dry and fresh samples were
mostly uniform, although the pH values determined from the fresh samples were
always 1 unit higher than those of the dry samples. The distribution of pH for dried
samples in second stage tests were more uniform than in the first stage tests, but
fresh samples also showed big variations among the layers with respect to the air
flow direction.

139
120

Height of composting mass in cm


100

80

60
1a
2a
40 3a
4a

20 5a
5a

0
5 6 7 8 9
pH

Figure 6.21 Final pH distribution throughout the composting mass at the end of
different first stage composting tests. Broken line indicates the pH of fresh sample.

120
Height of composting mass in cm

100

80

6a
60
6a
7a
40 8a
8a

20 9a
9a

0
5 6 7 8 9
pH

Figure 6.22 Final pH distribution throughout the composting mass at the end of
different first stage composting tests. Broken lines indicate the pH of fresh sample.

140
120

Height of composting mass in cm


100

80

60

1b
40 3b
4b
5b
20
5b

0
5 6 7 8 9
pH

Figure 6.23 Final pH distribution throughout the composting mass at the end of
different second stage composting tests. Broken line indicates the pH of fresh
sample.

120
Height of composting mass in cm

100

80

60
6b
7b
40 8b
8b
9b
20
9b

0
5 6 7 8 9
pH

Figure 6.24 Final pH distribution throughout the composting mass at the end of
different second stage composting tests. Broken lines indicate the pH of fresh
sample.

141
In tests 5b, 8b and 9b (upflow with or without reuse of air), the pH profile of fresh
samples varied sharply with pH values of 7 to 7.25 in the bottom layers and 8 to 8.5
in the top layers, whereas the values for dried samples were 5.5 and 6.5 for bottom
and top layers. The probable cause of variations of pH determined from the fresh
samples was the variation in NH3-N concentration throughout the height of the
composting mass. On the contrary, the dried samples showed mostly uniform pH due
to NH3-N removal through oven drying at 44 oC. Therefore, when reporting the pH
of a particular compost it should be mentioned whether a fresh or air dried sample
was used for pH analysis.

Tests 5a and 6a required longer startup times than other tests to raise the average
temperature beyond 50 oC. For the first 5 and 12 days, respectively, the pH of the
composting mass as determined from the daily collected condensate remained quite
low in tests 5a and 6a, and varied between 3.4 to 4.8 at which composting efficiency
decreases considerably (Epstein 1997). Test 5a (upflow with internal air
recirculation) slowly reached 60 oC after 6 days as shown in Figure 5.8, but test 6a
did not reach 60 oC even after 12 days (until this time only upflow aeration was
applied for this test). After that the upflow aeration mode of test 6a was changed to
downflow for only three hours, which accelerated degradation and subsequently the
temperature quickly reached 60 oC within these three hours. The pH also changed
from 3.64 to about 7 within the next 2 days. Therefore, reversal of airflow direction
changed the 12 days of inefficient composting within three hours to high efficiency.

Variations in final conductivity throughout the height of the composting mass at the
end of first and second stage tests are illustrated in Table 6.4. Low conductivity
values were observed at the top layers of the tests with alternate upflow/downflow
aeration and addition of moisture from the top. However, it was discussed in section
5.1.4.4 that the average conductivity value of the final compost after completion of
every stage always increased from its initial value. The average increment of
conductivity in final compost after all stages (including first, second and third
stages) was 2.1 mS/cm or 73%.

142
Table 6.4 Final conductivity in the different layers of the composting mass at
the end of first and second stage tests in mS/cm
Tests Raw L1 L2 L3 L4 L5 L6
1a 1.70 2.02 2.11 2.06 1.67 1.62 0.49
1b 1.66 2.30 1.80 2.40 2.61 1.09 0.31
2a 2.60 3.80 3.00 2.95 2.85 3.70 3.00
3a 2.70 4.60 4.10 4.40 4.35 2.80 5.50
3b 4.29 5.50 5.90 6.50 7.30 4.20 -
4a 4.20 7.00 7.00 6.00 5.90 - -
4b 6.50 6.90 7.10 7.10 6.60 7.20 7.10
5a 2.40 3.32 3.50 4.52 3.90 3.12 3.70
5b 3.68 4.72 4.75 5.60 4.75 5.50 4.32
5c 4.94 5.10 5.20 5.30 5.40 5.40 4.90
6a 2.95 5.50 5.60 6.00 5.90 4.20 2.20
6b 4.90 5.30 6.00 6.00 6.00 6.00 5.90
7a 3.68 4.08 4.00 4.22 5.70 6.10 3.14
7b 4.54 4.95 5.08 5.40 5.70 6.50 4.18
8a 3.20 4.40 4.28 4.49 4.60 5.70 4.16
8b 4.61 5.70 6.00 5.60 6.15 6.15 6.05
9a 2.60 3.40 3.15 3.40 3.58 3.50 3.27
9b 3.38 4.30 4.60 4.70 4.75 5.30 4.15

6.1.5.6 NH3-N

Final NH3-N distributions throughout the composting mass at the end of different
first and second stage tests are illustrated in Figures 6.25 to 6.28. Although the
aeration modes affected the NH3-N distributions, generally they were more uniform
in the second stage tests than in the first stage tests. The effect of internal air
recirculation on NH3-N distribution is very clear as shown in Figure 6.25. In the
bottom layers of tests 1a, 5a and 3a the concentrations were 3.25, 3.8 and 5.75 NH3-
N g/kg-TS for r = 20 to 50%, 20 to 80% and 100 to 200%, respectively. In the
bottom and top layers of test 8a (upflow), the concentrations were 1.9 and 7.5 NH3-N
g/kg-TS. Similar but opposite patterns were obtained for tests with downflow
aeration. Therefore, the variations of NH3-N in the composting mass mainly
depended on the aeration modes (final aeration mode in the case of alternate
upflow/downflow aeration) and sometimes on the duration of tests.

143
120

Height of composting mass in cm


100

80

60

1a
40 2a
3a
4a
20
5a

0
0 1 2 3 4 5 6 7 8 9 10 11 12
NH3-N, g/kg

Figure 6.25 Final NH3-N distribution throughout the composting mass at the end of
different first stage composting tests.

120
Height of composting mass in cm

100

80

60

40 6a
7a
8a
20
9a

0
0 1 2 3 4 5 6 7 8 9 10 11 12
NH3-N, g/kg

Figure 6.26 Final NH3-N distribution throughout the composting mass at the end of
different first stage composting tests.

144
120

Height of composting mass in cm


100

80

60
1b

40 3b
4b
5b
20

0
0 1 2 3 4 5 6 7 8 9 10 11 12
NH3-N, g/kg

Figure 6.27 Final NH3-N distribution throughout the composting mass at the end of
different second stage composting tests.

120
Height of composting mass in cm

100

80

60

40 6b
7b
8b
20
9b

0
0 1 2 3 4 5 6 7 8 9 10 11 12
NH3-N, g/kg

Figure 6.28 Final NH3-N distribution throughout the composting mass at the end of
different second stage composting tests.

145
In summary, in most of the tests the distribution of pH and NH3-N, both determined
from fresh samples, showed similar patterns and was affected by the air flow
direction as discussed above. The respective tests where similar patterns in
distribution of pH and NH3-N were obtained were 8a, 9a, 5b, 8b and 9b. Therefore, it
appears that the variation in pH of fresh samples is mainly due to the variation of
NH3-N concentrations in different layers, i. e. for a particular layer the pH is usually
higher if the NH3-N concentration is higher in that layer.

6.1.5.7 Extent of Degradation and Product Stability

The amount of biodegradable volatile solids (BVS), which degraded in each first stage
test and second stage test and their overall sum are presented in Table 6.5. The daily
degradation of BVS of a particular test can also be illustrated by the respective daily
O2 consumption and cumulative O2 consumption curves as shown in Figures 5.22 to
5.38. The range of BVS reduction in different first stage tests was 52 to 72% of the
initial estimated BVS of each test, after different durations. If the degradation after
20 days (calculation based on the cumulative O2 consumption curves) for all the first
stage tests are separately considered, the range of degradation was 48 to 59% of
initial estimated BVS. In tests 1a, 3a and 5a, where upflow with internal air
recirculation was applied, the degradations were 58.91, 50.49 and 53.66%,
respectively. In these tests the degradation rates were higher in the first 15 to 20 days
and after that decreased considerably due to low moisture content as discussed in
section 6.1.4.4 and illustrated in the respective Figures 5.22, 5.25 and 5.29.
Therefore it can be concluded that a maximum period of 20 days is adequate for first
stage composting where upflow aeration with internal air recirculation is applied and
the maximum recirculation ratio should be 20 to 80%.

In tests 6a and 7a, where alternate upflow/downflow aeration and optional addition
of water were applied, the BVS degradations were 53.94 and 57.66%, respectively.
The percentage degradation in the tests using upflow with internal air recirculation
and alternate upflow/downflow (i.e. 1a, 3a, 5a, 6a and 7a) was higher than that of the
tests 2a, 4a and 9a where unidirectional air flow or only one change in air flow from
upflow to downflow were applied. The degradation in tests 2a, 4a and 9a were 47.88,
146
45.45 and 49.21% after 20 days, respectively. In test 8a with upflow aeration the
degradation was 57.43%, which might be due to the presence of comparatively more
milk products in the food waste, which was used as the feed material in this test.
Therefore, it can be concluded that upflow aeration with internal air recirculation
and alternate upflow/downflow aeration with optional addition of water accelerated
the rate of composting and thereby increased the extent of degradation among these
first stage tests.

Table 6.5 Biodegradable Volatile Solids (BVS) degraded () in first stage tests
(ST1), second stage tests (ST2) and overall sum
Test Days BVSi BVS BVS BVS Test Days BVS BVS BVS Days BVS
ST1 kg kg % %, in ST2 kg % % of Total Total
20 days initial
1a 31 16.87 11.03 65.38 58.91 1b 21 4.34 74.32 25.73 52 91.11
2a 33 12.38 7.33 59.21 47.88
3a 29 11.35 6.34 55.86 50.49 3b 39 3.63 72.46 31.98 68 87.84
4a 49 17.17 12.36 71.99 45.45 4b 21 2.19 45.53 12.75 70 84.74
5a 20 10.58 5.68 53.69 53.66 5b 21 2.41 49.18 22.78 52 81.57
6a 30 12.02 8.49 70.63 53.94 6b 21 1.74 29.79 14.48 51 85.11
7a 23 11.66 7.16 61.41 57.66 7b 32 2.38 52.89 20.41 55 81.82
8a 28 11.74 7.49 63.80 57.43 8b 28 1.79 42.12 15.25 56 79.05
9a 24 12.10 6.33 52.31 49.21 9b 32 2.88 50.00 23.80 56 76.12

In the second stage tests, the range of BVS degradation was 13 to 32% of the initial
estimated BVS after a range of 21 to 39 days. However, degradation in all second
stage tests was mostly completed within 10 to 15 days as shown by the cumulative
and daily O2 consumption curves in the respective Figures 5.23 to 5.38. Test 3b was
continued for 39 days to investigate the effect of long time reuse of air in a series
connection on degradation, and it was found that most of the degradation in this test
was also completed within 25 days as illustrated in Figure 5.26. From this figure it
can be seen that the O2 consumption rates after 25 days were very low and constant.
Therefore, it would not be unreasonable to assume that test 3b was almost completed
after 28 days instead of 39 days. Therefore, the total assumed duration of tests 3a+3b
is 57 days.

The total BVS degradations in tests 1a+1b, 3a+3b and 5a+5b were 91.11, 87.84 and
81.57% after 52, 57, and 52 days, respectively. The end product of both tests 3a+3b
147
and 5a+5b achieved complete stability with stability index V. A stability test for
tests 1a+1b was not carried out. In tests 6a+6b and 7a+7b the extent of degradation
was 85.11 and 81.82% after 51 and 55 days, respectively. The stability index of the
end product of tests 6a+6b was IV. However, in the end product of tests 7a+7b a
stability index of only III was achieved. The reactor in second stage test 7b was
connected with the reactor of test 8a as a two reactor system, just 1 day after its start,
which quickly raised temperatures in all the layers and after 8 days all the layers
simultaneously cooled down indicating no further degradation as illustrated in Figure
5.13. The temperature curves (in Figure 5.13) of test 7b were completely different
from other second stage tests. This different effect or unfavorable environment might
inactivate all those microorganisms which are responsible for degradation in
composting and finally the self-heating test showed the end product to have only
stability index III.

A comparatively lower degradation of 79.05 and 76.12% occurred in tests 8a+8b and
9a+9b, respectively, after the same duration of 56 days. The stability index of the
end products of these tests was IV and III, respectively. Upflow aeration was applied
in test 8a and reuse of spent air in a two reactor system was applied in test 8b,
whereas only downflow aeration was applied in test 9a and upflow in test 9b as a
single reactor system. Therefore, comparing tests 8a+8b and 9a+9b, the two reactor
system produced a more stable end product than the single reactor system in the
same total time. Although 84.74% BVS degraded in test 4a+4b in a single reactor
system, it required the longest processing time of 70 days to reach end product
stability index IV.

In summary, composts after second stage composting in the two-reactor system were
more stable or completely stable, whereas the end products of second stage
composting from the single reactor system did not achieve complete stability despite
the same or increased total composting time. Therefore, the two-reactor system
improves composting efficiency while lowering air (according to section 6.1.5.3) and
energy requirements.

148
6.1.6 Effect of Aeration Shutdown

Aeration was shut down in test 2a for 3 hours on the 11th day due to sudden power
failure, which significantly affected the temperature pattern as well as the O2
consumption rate. Again between the 16th and 17th day, the aeration rates became
very low for a 24 hour period due to low pressure in the air supply pipe.

6.1.6.1 Temperature

The effect of aeration shutdown on day 11 on the temperature of test 2a is illustrated


in Figure 5.3. After shutdown the temperatures of all the layers sharply increased to
68 oC which had an inhibitory effect on the microbial activity. The temperatures
went down sharply and created a steep temperature gradient (66 oC at bottom and 30
o
C at top of the composting mass) once aeration resumed. However, the aeration
shutdown subsequently decreased the temperature of all the layers by 12 to 14 oC.
Low aeration rates from the 16th to 17th day again increased the temperatures of all
the layers.

6.1.6.2 Oxygen Consumption Rate

The effect of aeration shutdown on the O2 consumption rate and on the average
temperature of test 2a is illustrated in Figure 5.24. The O2 consumption rate went
down considerably after aeration shutdown. The O2 consumption rate curve and
average temperature curve followed a similar pattern. Therefore, sudden aeration
shutdown reduced average temperature as well as O2 consumption rate and thereby
reduced overall microbial activity during composting.

6.1.7 Self-heating test

Few attempts have been made so far to establish some relationships among the
maximum temperature increase rate, maximum temperature, area under the
temperature curves, and VS reduction using data of more than 80 self-heating tests
which were applied for similar type of compost.

149
6.1.7.1 Relationship between g VS Reduction and Temperature

Correlation between maximum temperature increase rate (Imax) in oC/h and g volatile
solids reduction in self-heating tests is shown in Figure 6.29. The increment of Imax
followed the log rate, which was fast up to 15 g VS degradation and after that
became slower. The correlation coefficient was R2=0.63. Higher Imax indicates higher
degradation and thereby a less stable end product. Therefore it was confirmed that
unstable compost has a higher Imax as discussed in section 2.7.1.

Another correlation between maximum temperature in oC and g volatile solids


reduction in self-heating tests is shown in Figure 6.30. This correlation also followed
the log rate curve with correlation coefficient R2=0.69. There was only little
degradation when the maximum temperature remained below 30 oC, as expected.
However, higher degradation (30 to 40 g VS) generated high peak temperatures
above 60 oC indicating an unstable end product. Therefore it is also confirmed as
discussed in section 2.7.1 that higher peak temperatures above 50 to 60 oC are an
indication of unstable end product. However, up to 15 g VS degradation in a
specified vacuum flask (as specified in section 4.8.1) sharp variations in the peak
temperature were observed and the extent of degradation within this range could be
predicted by temperature from Figure 6.30, whereas over 15 g VS degradation, the
variation in the peak temperature was less. Therefore, the degradability over 15 g VS
in the self-heating test will be poorly predicted by the peak temperature.

6.1.7.2 Relationship between the Area under the Temperature Curves

Correlation between A72 (area under temperature curve after 72 hours) in oC.h and g
volatile solids reduction in the self-heating tests is shown in Figure 6.31. It followed
a linear relation with R2=0.49. It can be seen that more VS degradation produced a
higher A72 which indicated instability of the end product and confirmed the theory in
section 2.7.1.

The correlation between Atotal-ambient (area under net temperature increase curve after
completion of the self-heating test) in oC.h and g volatile solids reduction is shown

150
in Figure 6.32, which followed a linear relation. This correlation showed a higher
correlation coefficient, (R2=0.85) than the correlation between A72 and g VS
reduction. It is clearly shown that higher Atotal-ambient indicate higher VS degradation.
Therefore, the extent of degradation in a specified self-heating test can be
determined more precisely from the correlation between Atotal-ambient in oC.h and g VS
reduction as indicated in Figure 6.32.

The relationship between the areas under the (T-Ta) curve in oC.h for different layers
of pilot-scale tests (total of 6 layers) and their respective self-heating tests is shown
in Figures 6.33 to 6.35 for tests 3a, 8a and 9a, respectively, where T is the test
temperature either of the layers in the pilot-scale test or of the self-heating test of
that respective layer. Ta is the ambient temperature. For upflow aeration in the pilot-
scale tests 3a and 8a, the upper layer showed a higher value of the area under T-Ta
and the value gradually decreased for the subsequent bottom layers. In contrast, the
self-heating tests of the respective layers showed similar gradients but completely
opposite ones. The self-heating tests for the upper layers showed a lower value of the
area under T-Ta and the value gradually increased in the self-heating tests for the
subsequent bottom layers. Completely opposite trend was found for downflow
aeration in pilot-scale test 9a and its respective self-heating tests as shown in Figure
6.35. However the individual sums of the areas under (T-Ta) curves for a layer of the
pilot-scale test and its respective self-heating test showed approximately similar
values with an average of 21400 oC.h for both tests 3a and 8a. Therefore, the sums of
the areas under (T-Ta) curves of any layer of the pilot-scale test and its respective
self-heating tests are approximately equal, indicating the energy content of the
original material. The proportion of the two areas is a measure of the extent of
degradation achieved.

151
1.40

1.20

1.00
Imax in oC/h

0.80

0.60
y = 0.3313Ln(x) - 0.1911
R2 = 0.6301
0.40

0.20

0.00
0 10 20 30 40 50
g Volatile solids reduction

Figure 6.29 Correlation between Imax (maximum temperature increase) in oC/h and g
volatile solids reduction in self-heating tests.

80

70
Maximum temperature (Tmax) in oC

60

50

y = 12.954Ln(x) + 18.959
R2 = 0.6874
40

30

20
0 10 20 30 40 50
g Volatile solids reduction

Figure 6.30 Correlation between maximum temperature oC and g volatile solids


reduction in self-heating tests.

152
4500

4000

3500

3000
A72 in oC.h

2500

y = 40.718x + 1989.3
2000
R2 = 0.4882

1500

1000

500

0
0 5 10 15 20 25 30 35 40 45
g Volatile solids reduction

Figure 6.31 Correlation between A72 (area under temperature curve after 72 hours)
in oC.h and g volatile solids reduction in self-heating tests.

25000

y = 375.56x
R2 = 0.8493
20000
Atotal-ambient in oC.h

15000

10000

5000

0
0 5 10 15 20 25 30 35 40 45 50
g Volatile solids reduction

Figure 6.32 Correlation between Atotal-ambient (area under temperature curve after
completion) in oC.h and g volatile solids reduction in self-heating tests.

153
SH
L6
PS

L5 Average oC.h

Layer of compost L4

L3

L2

L1

0 5000 10000 15000 20000 25000 30000


o
Area under (T-Ta) curve in C.h

Figure 6.33 Relationship between the areas under (T-Ta) curve of different layers of
pilot-scale test 3a (upflow aeration, r = 100 to 200%) and respective self-heating
tests. Where T and Ta are test (in different 6 layers) temperature and ambient
temperature, respectively. SH = self-heating test, PS = pilot-scale test.

L6 SH
PS

L5
Layer of compost

L4
Average oC.h

L3

L2

L1

0 5000 10000 15000 20000 25000 30000


Area under (T-Ta) curve in oC.h

Figure 6.34 Relationship between the areas under (T-Ta) curve of different layers of
pilot-scale test 8a (upflow aeration) and respective self-heating tests.

154
6 SH
PS

5
Average oC.h
Layer of compost

0 5000 10000 15000 20000 25000 30000 35000 40000


o
Area under (T-Ta) curve in C.h

Figure 6.35 Relationship between the areas under (T-Ta) curve of different layers of
pilot-scale test 9a (downflow aeration) and respective self-heating tests.

6.1.8 Material Balance

At the beginning and the end of each test, either pilot-scale or bench-scale, the waste
input and output were weighed for determining the material balances. Two examples
of an estimated mass and energy balance for tests 1a and 1b are shown in Figure
6.36. Water balances for the same tests are shown in Figure 6.37. The detailed
calculation of the mass and energy balances and the water balances are presented in
Tables A2.1 and A2.2 in Appendix I. Mass and energy balances were estimated for
the degradation of 1000 g volatile solids, whereas the water balances were estimated
on a total mass basis. According to the mass and energy balance of test 1a, 1276 g O2
was used to degrade 1000 g VS which produced 1754 g CO2, 6g NH3-N and 516 g
H2O releasing 15000 kJ heat energy. About 50 to 58% of the released heat during
degradation was used for evaporation of the water in the composting mass in tests 1a
and 1b. According to the water balance of test 1a, 46.40 kg water was present in the
initial mix and a total 4.15 kg water was added intermittently during the test.
Metabolism of the VS produced 5.67 kg water. During the test, 7.90 kg water
escaped through the outlet air, 25.30 kg was collected as condensate and only 0.45
kg was leachate. The final compost of test 1a held 22.60 kg water. About 60% of

155
total initial water was removed by the air in test 1a, whereas in the second stage test
1b the percentage of water removed by the air was less than half (27.46%) of the first
stage test. Therefore, water removal in the first stage test was almost double that of
the second stage test.

1754 g 1733 g
CO2 CO2

1276 g 1000 g VS 6.0 g 1260 g 1000 g VS 19.5 g


O2 NH3 O2 NH3

516 g 508 g
H2O H2O

15000 kJ 15056 kJ

7450 kJ 7550 kJ 6312 kJ 8744 kJ


Used for H2O Heat Used for H2O Heat
evaporation loss evaporation loss

Test 1a Test 1b

Figure 6.36 Estimated mass and energy balances for the degradation of 1000 g
volatile solids (VS) for tests 1a and 1b.

7.90 3.49
escaped escaped
46.40 in 35.55 in
initial mix, 25.30 initial mix, 7.50
4.15 with 30.8 TS condensate 4.23 with 20.0 TS condensate
added 5.67 0.45 added 2.21 0.45
metabolism leachate metabolism leachate
22.60 30.22
in final in final
compost, compost,
with with
19.70 TS 15.70 TS

Test 1a Test 1b

Figure 6.37 Estimated total water balance for tests 1a and 1b. All the numbers are in
kg.
6.2 Bench-scale Tests
156
6.2.1 Natural Aeration and Forced Aeration

The temperature curves of most natural aerated bench-scale tests followed a regular
and almost similar pattern, whereas the pattern of the temperature curves of the
forced aerated tests were irregular, such as having several peaks in the curves. The
average temperature and the area under the temperature curve of the forced aerated
tests were lower than the natural aerated tests as discussed in section 5.2.1. However,
the natural aerated tests 1NA1 and 1NA2 and forced aerated test (0.21 L/min.kg VS)
1FA2 showed comparable mean temperature (43.92, 46.72 and 41.59 oC), area under
the temperature curves (54300, 57750 and 51420 oC.h) and VS degradation (40.04,
29.30 and 38.27%). In test 1NA2, less VS degraded probably due to dryness.
Therefore, it appeared that the extent of degradation in natural aerated and forced
aerated bench-scale tests was almost similar and that temperature curves of the
natural aerated bench-scale tests followed a reproducible pattern. It should be
mentioned that the four sets of duplicate self-heating tests conducted under the same
principle as a natural aerated bench-scale test using a mixed sample of test 9b, also
showed a very similar pattern in their temperature curves as illustrated in Figure 5.52
and discussed in section 5.1.8.1.1.

6.2.2 Relationship between Moisture Content and VS Degradation

A total of 12 bench-scale tests with varying moisture content from 28.80 to 80.21%
were conducted under run 4 to investigate the effect of initial moisture content on
VS degradation and on temperature parameters as discussed in section 5.2.1. The
relationship between % VS reduction and initial moisture content in bench-scale
tests is illustrated in Figure 6.38. The %VS reduction was highest in the moisture
range between 55 and 65%. VS reduction sharply decreased at a moisture content
below 30 to 40%.

6.2.3 Relationship between Moisture Content and Temperature

The relationship between the area under the temperature curve and the initial
moisture content in bench-scale tests under run 4 is illustrated in Figure 6.39. The
area under the temperature curve, in oC.h, was highest in the moisture range between
157
55 and 65%. It should be noted that a higher value of the area under the temperature
curve indicates higher heat production and thereby higher VS degradation. Therefore
it appears that the optimum initial moisture content with respect to %VS reduction
and the area under temperature curve for this particular type of waste was 55 to 65%
which confirms similar findings by others as discussed in section 2.6.3.

6.3 Comparison between Bench-scale and Pilot-scale Tests

The results of pilot-scale tests 9a and 9b were compared with the corresponding
bench-scale tests of runs 2 and 3 in section 5.2. The pattern of daily temperature
variations of the pilot-scale test (average temperature) and the bench-scale test were
not similar. However, it was found that the overall average temperature and the area
under the temperature curve of the pilot-scale tests were a little lower (as expected)
than those of the corresponding bench-scale tests as presented in section 5.2.1.
However, these temperature parameters have a closer similarity in both pilot-scale
tests and bench-scale tests than the average VS degradations. The VS degradation in
pilot-scale test 9a was 10% lower than in the corresponding bench-scale test 2NA2
within approximately same duration of 24 and 22 days, respectively. On the other
hand, the VS degradation in pilot-scale test 9b was 2.4% higher than the
corresponding bench-scale tests 3NA1 and 3NA2.

Therefore, the average temperature pattern and VS degradation in the pilot-scale


tests with full height are poorly related to the results of the bench scale tests.
Furthermore, the gradients in temperature, moisture and other physico-chemical
parameters throughout the height of the composting mass could not be properly
explained by the bench-scale tests. However, bench-scale tests were found most
suitable for the determination of optimum values of some important parameters such
as: moisture content, temperature, etc. for a particular type of waste mixture.

158
20

15
VS reduction in %

10

0
0 10 20 30 40 50 60 70 80 90 100
Initial moisture content in %

Figure 6.38 Relationship between % VS reduction and initial % moisture content in


bench-scale test run 4.

25000
Area under temperature curve in oC.h

20000

15000

10000

5000
0 10 20 30 40 50 60 70 80 90 100
Initial moisture content in %

Figure 6.39 Relationship between the area under temperature curve and initial %
moisture content in bench-scale test run 4.

159
CHAPTER 7

KINETIC ANALYSIS AND APPLICATION

7.1 Pilot-scale Test: Reaction Rates and Temperatures

7.1.1 Reaction Rates

The waste degradation rates (reaction rates) and their orders were studied for
describing the kinetics of the biological composting process under different aeration
modes. In this study, three different reaction orders namely zero, first, and second
order were considered as described in Chapter 3. The reaction rate constants k0 (zero
order), k1 (first order) and k2 (second order) were basically estimated from the daily
O2 consumption rates in kg/d, which are proportional to the BVS degradation by a
factor of 1.3, using Equations 3.2, 3.5 and 3.8, respectively. The relationships
between different reaction rate constants k0, k1 and k2 and the absolute temperature T
according to the Arrhenius equation (Equation 3.12) are shown in Figures 7.1 and
7.2 as ln(k) vs 1/T for tests 1a and 3a, respectively. Although this relation was
established for all tests, for the sake of simplicity, only tests 1a and 3a are presented.
The best correlation was obtained for the first order reaction with R2 = 0.7783 for
test 1a and R2 = 0.8445 for test 3a.

7.1.2 Prediction of BVS Degradation from Temperature

7.1.2.1 Average Temperature

From the calculated Arrhenius equation of each test, as in Figures 7.1 or 7.2 for tests
1a and 3a, and the recorded average temperature of the pilot-scale tests the
160
percentage BVS degradation could be calculated according to the three different
reaction orders. Equations 3.4b, 3.7b and 3.10b described in Chapter 3 are the
mathematical expression of predicted percentage BVS degradation for zero, first and
second order reaction, respectively. Figures 7.3 and 7.4 show the predicted % BVS
degradation for tests 1a and 3a under three different reaction orders. The predicted
degradation curves from the three different reaction orders are compared with the
actual BVS degradation curve and it can be shown that the first order prediction
produces the best fit with the actual data. Therefore, it appeared that the BVS
degradation in forced aeration composting follows a first order reaction.

Table 7.1 presents the first order reaction rate constants at 25 oC and 50 oC,
temperature coefficient Q10 and the activation energy Ea calculated from Arrhenius
equation for different tests. The range of reaction rate constants at 25 oC is 0.004 to
0.013 per day and at 50 oC 0.027 to 0.14. The range of Q10 in the first stage tests is
1.7-2.3 which compares well with Schulze (1960), and in the second stage tests 2.2
to 2.9 except for 6b and 8b.

o o
Table 7.1 First order reaction rate constants at 25 C and 50 C,
temperature coefficient Q10 and activation energy Ea
Test Ea Q10 k1 k1 Test Ea Q10 k1 k1
kJ/mol d , 25 C d , 50oC
-1 o -1
kJ/mol d , 25 C d , 50oC
-1 o -1

1a 56.91 1.97 0.007 0.044 1b 82.55 2.67 0.011 0.141


2a 30.20 1.43 0.011 0.027
3a 55.60 1.94 0.009 0.051 3b 85.03 2.75 0.010 0.147
4a 51.12 1.83 0.006 0.028 4b 89.53 2.90 0.008 0.123
5a 46.92 1.75 0.009 0.040 5b 66.98 2.21 0.007 0.059
6a 73.30 2.39 0.007 0.065 6b 124.11 4.37 0.003 0.155
7a 45.07 1.71 0.013 0.051 7b 69.40 2.28 0.006 0.052
8a 65.98 2.19 0.010 0.082 8b 110.00 3.70 0.003 0.096
9a 43.53 1.67 0.009 0.037 9b 88.88 2.87 0.004 0.068

The activation energy Ea varied from 44 to 90 kJ/mole in different tests. Similar


values of 92.05 kJ/mole at 50 to 70 oC for thermophilic bacteria (Nakasaki 1985d)
and 58 kJ/mole for Aerobacter aerogenes in continuous culture at 35 oC (McKinley
and Vestal 1984) were reported.

161
Corresponding temperature in oC
60 50 40 30
0

ln(k0) = 26.73 - 8911.5/T


R2 = 0.7338
-2

ln(k1) = 18.06 - 6845.3/T


ln(k)

-4 R2= 0.7783

-6 ln(k2) = 8.61 - 4535.1/T


R2= 0.4744

-8
0.00300 0.00310 0.00320 0.00330
1/T (T in oK)

Figure 7.1 The graphical presentation of Arrhenius equation for pilot-scale test 1a.
where k is reaction rate constant at temperature T in oK.

Corresponding temperature in oC
60 50 40 30
0

ln(k0) = 24.15 - 8098.2/T


R2 = 0.7365
-2

ln(k1) = 17.73 - 6688.5/T


ln(k)

-4 R2= 0.8445

-6 ln(k2) = 11.32 - 5278.7/T


R2= 0.7997

-8
0.00300 0.00310 0.00320 0.00330
1/T (T in oK)

Figure 7.2 The graphical presentation of Arrhenius equation for pilot-scale test 3a.
where k is reaction rate constant at temperature T in oK.

162
70

60
Percentage BVS degradation

50

40

30

Actual
20
Zero order
First order
10
Second order

0
0 5 10 15 20 25 30 35
Time in days

Figure 7.3 Actual and predicted %BVS degradation with time for different reaction
orders in test 1a.

70

60
Percentage BVS degradation

50

40

30

Actual
20
Zero order
First order
10
Second order

0
0 5 10 15 20 25 30 35
Time in days

Figure 7.4 Actual and predicted %BVS degradation with time for different reaction
orders in test 3a.

163
In wastewater treatment processes, Ea ranges from 8.4 to 84 kJ/mol (Metcalf & Eddy
1979) and for inactivation processes the range of inactivation energy Ed is between
209 and 418 kJ/mol (Bailey and Ollis 1986).

7.1.2.2 Using First Order Reaction at Different Constant Temperatures

Using the first order reaction rate constants k1 for test 1a, the predicted BVS
degradation with time for constant temperatures of 20, 30, 40, 50 and 60 oC could be
calculated from the following equations:
30 oC or 303oK, k1=exp(18.06-6845.3/303)=0.01076, d-1

BVSo BVS
% = 100[1 exp( kt )] = 100[1 exp( 0.01076t )] (7.1)
BVSo

40 oC or 313oK, k1=0.02215, d-1


BVSo BVS
% = 100[1 exp( 0.02215t )] (7.2)
BVSo

50 oC or 323oK, k1=0.04359, d-1

BVSo BVS
% = 100[1 exp( 0.04359 t )] (7.3)
BVSo

60 oC or 333oK, k1=0.08237, d-1

BVSo BVS
% = 100[1 exp( 0.08237 t )] (7.4)
BVSo
Figure 7.5 shows the predicted BVS degradation with time at different constant
temperatures for test 1a. For test 3a, using similar equations, the predicted BVS
degradation with time at different constant temperatures is shown in Figure 7.6. It is
clearly seen in Figure 7.5, that for the first 15 days the actual degradation curve of
test 1a almost followed the predicted BVS degradation of 50 oC. In fact the average
temperature of this test 1a was almost 50 oC for that period as shown in Figure 5.22.
After 15 days the reaction rates became slower as temperature fell towards 35 oC.
However, for test 3a, the actual degradation curve followed the predicted BVS
degradation of 50 oC for 10 days, because the average temperature of this test was
almost 50 oC for that 10 day period as shown in Figure 5.25.

7.1.2.3 Outlet Temperature


164
Since it is difficult to measure the temperature at different points in the large-scale
composting mass and thereby determine the average temperature, an attempt was
made to predict BVS degradation using outlet air temperature instead of average
temperature because it is easy to measure the outlet temperature of a closed
composting process. The correlation between average temperature and outlet
temperature for test 3a with upflow aeration plus internal air recirculation is made
and shown in Figure 7.7. Application of this correlation for another, similar test 5a
shows excellent prediction as shown in Figure 7.8. Another correlation between
average temperature in the composting mass and the outlet air temperature of tests 8a
and 9a is shown in Figure 7.9. Upflow and downflow aeration mode were applied in
tests 8a and 9a, respectively. The relationship obtained was average temperature is
0.8506 times outlet temperature, with a correlation coefficient R2=0.8506. Using this
relationship a prediction of %BVS degradation of another test 7a is made. In test 7a
alternate upflow/downflow aeration mode was applied. The predicted %BVS
degradation from outlet and average temperature of test 7a compared well with the
actual degradation in Figure 7.10. Therefore the %BVS degradation could be
predicted by easily measurable outlet air temperature instead of using the average
temperature which has to be calculated from the different measured temperatures
throughout the composting mass.

165
100

90 Actual

30 oC
80
Percentage BVS degradation
40 oC
70 50 oC
60 oC
60

50

40

30

20

10

0
0 5 10 15 20 25 30 35
Time in days

Figure 7.5 Actual and predicted %BVS degradation at different constant


temperature using first order reaction model of test 1a.

100

90 Actual

30 oC
Percentage BVS degradation

80
40 oC
70 50 oC
60 oC
60

50

40

30

20

10

0
0 5 10 15 20 25 30 35
Time in days

Figure 7.6 Actual and predicted %BVS degradation at different constant


temperature using first order reaction model of test 3a.

166
60

y = 0.90x
50 R2= 0.7571
Average temperature in oC

40

30

20

10

0
0 10 20 30 40 50 60 70
Outlet Temperature in oC

Figure 7.7 Correlation between average temperature and the outlet temperature of
composting tests 3a where upflow aeration mode with internal air recirculation was
applied

60

50
Percentage BVS degradation

40

30

20

Actual degradation

10 Prediction from average temperature


Prediction from outlet temperature

0
0 5 10 15 20 25
Time in days

Figure 7.8 Predicted percentage BVS degradation from outlet (using relationship
from Figure 7.7) and average temperatures and actual degradation of test 5a where
upflow aeration with internal air recirculation was applied

167
60

y = 0.87x
50 R2= 0.8506
Average temperature in oC

40

30

20

10

0
0 10 20 30 40 50 60 70
Outlet Temperature in oC

Figure 7.9 Correlation between average temperature and outlet temperature of


composting tests for upflow and downflow aeration mode of tests 8a and 9a.

70

60
Percentage BVS degradation

50

40

30

20
Actual degradation

10 Prediction from average temperature


Prediction from outlet temperature

0
0 5 10 15 20 25
Time in days

Figure 7.10 Predicted percentage BVS degradation from outlet (using relationship
from Figure 7.9)and average temperatures and actual degradation of test 7a where
alternate upflow and downflow aeration mode was applied
7.1.3 Multilayer Analysis
168
The degradation rate of a given organic substrate in the composting process depends
on the temperature, hence temperature gradients usually create a different extent of
degradation of BVS in different compost layers. Therefore, it is important to analyse
the rate and extent of BVS degradation in different layers of the composting mass.
For this detailed analysis, the final composting mass in each test is divided into 6
layers along the height of the reactor as described in section 4.6.1.1. Each layer,
starting from bottom was considered 20 cm thick except for the top one, which had a
different thickness for different tests. Initial BVS in each layer is proportional to the
amount of fixed solids present in that layer. The predicted %BVS degradation in
different layers was calculated using the first order reaction model as presented in
section 7.1.2.1 and the individual daily average temperature of that layer.

Examples of a relatively non-uniform and uniform predicted %BVS degradation in


different layers compared to predicted average degradation are shown in Figures
7.11 and 7.12 for tests 3a and 7a, respectively. In test 3a (upflow aeration and r =
100-200%), higher temperatures resulted in a higher percentage degradation in the
top layers. The degradation in the top and bottom layers was 74 and 36%,
respectively, whereas the average degradation was 56%. The variation in top and
bottom layers was 19% with respect to the average predicted degradation.
Comparatively less variation with respect to average degradation was obtained in
test 5a (upflow aeration and r = 20-80%, results are not shown here). In test 7a
(alternate upflow/downflow aeration and addition of moisture), the degradation in
different layers was more uniform than in other tests. The highest and lowest
percentage degradation in layer L3 and L6 was 73 and 56%, respectively, with an
average of 61%. The variation of upper and lower percentage degradation was
8.5% with respect to the average predicted degradation. Due to non uniform or
partial degradation, by implication less biological stability could be achieved in the
mixed product of the first stage composting process which could increase the O2
consumption rate at the beginning of the second stage composting (as discussed in
section 6.1.5.2).
Uniform degradation throughout the layers can be achieved in unidirectional aeration
composting process like test 3a, if frequent mixing of waste material is applied.

169
Figure 7.13 shows the positive effect of assumed mixing (applied after every seven
days) on the predicted %BVS degradation in different layers of test 3a, using the
first-order reaction model.

Using the relationships in Figures 7.11 and 7.12 for tests 3a and 7a, respectively, the
initial and final BVS, in kg, for the six different layers are shown in Figure 7.14.
Higher degradation occurred in the top layers of test 3a due to upward movement of
temperature and other physico-chemical parameters by the upflow aeration with
internal air recirculation. The degradation in the subsequent layers decreased
gradually. Similar results were also obtained in the other tests with unidirectional
upflow aeration or downflow aeration. Since the oC.h (area under the temperature
curve) of different layers (as presented in Figure 6.34) and the predicted %BVS
show a certain correlation, oC.h could be indicative of degradation.

100
average
90 L6
L5
Percentage BVS degradation

80
L4
70 L3
L2
60
L1
50

40

30

20

10

0
0 5 10 15 20 25 30
Time in days

Figure 7.11 Predicted %BVS degradation in different layers of test 3a using first
order reaction model.

170
100
average
90
L6
Percentage BVS degradation
80 L5
L4
70
L3

60 L2
L1
50

40

30

20

10

0
0 5 10 15 20 25 30
Time in days

Figure 7.12 Predicted %BVS degradation in different layers of test 7a using first
order reaction model.

100

90
Mixing at Mixing at Mixing at
7 th day 14 th day 21 st day
Percentage BVS degradation

80

70

60

50
average

40 L6
L5
30 L4
L3
20
L2
10 L1

0
0 5 10 15 20 25 30 35
Time in days

Figure 7.13 Predicted %BVS degradation in different layers of test 3a using first
order reaction model. Assumed mixing is applied after every 7 days.

171
In test 7a the amount of final predicted BVS in different layers was approximately
equal to that of test 3a. Therefore, significant variation in BVS degradation was
observed in the composting mass along the vertical directions when unidirectional
aeration were applied and this situation could be improved by applying alternate
upflow/downflow aeration and addition of moisture.

12

1.12 1.63

10
1.84
2.06

8
2.15
2.14
BVS in kg

6
2.05
0.29
2.14
0.62

4 0.83
0.71

2.35
0.95
1.97 0.77

0.67
2 1.25
0.58

1.84 1.71 0.59


1.18
0.56
0
3a, initial 3a, final 7a, initial 7a, final

Figure 7.14 Initial and final BVS in kg, in six different layers of test 3a and 7a.

172
7.2 Self-heating test: Heat Generation Rate and
Respirometric Activity

7.2.1 Heat Energy Balance

The temperature curve obtained in the self-heating test is the result of heat energy
changes in the waste sample over time. A simplified heat energy balance can be
formulated and described according to Koenig (1997) by the following Equation:

dT dVS dH 2 O
M. c = H U. A( T Ta ) L in J/h (7.5)
dt dt l dt e
change of heat biological heat loss of sensible loss of latent
energy in the generated by heat to heat due to
sample degradation of surroundings evaporation
VS in waste

where

M = mass of wet waste sample, with M = FS + VS + H2O, g


FS = fixed solids (inert mineral matter in waste), g
VS = Volatile solids (degradable organic matter in waste), g
H2O = Water content in the wet waste sample, g
c = specific heat capacity of wet waste sample, with
c = (FS.cFS + VS.cVS + H2O.c H2O)/M, J/g.oC

T = temperature of wet waste sample in vacuum flask, oC


Ta = ambient temperature, oC
t = time since start of test, h
Hl = heat energy generated by the degradation of VS, J/g
Le latent heat of evaporation, J/g
U = overall coefficient of heat transfer through top, side and
bottom of the filled vacuum flask, J/h.m2.oC
A = total surface of top, side and bottom of filled vacuum flask, m2

The heat energy balance alone is not sufficient for prediction of temperature and
would have to be coupled with a mass and a water balance. The waste in the vacuum
flask, used for the self-heating test, can exchange gases with the surrounding
173
atmosphere and hence constitutes a thermodynamically open system, leading to
changes in solids and water content over time. Since these changes cannot be
monitored separately in a simple manner, a simplified approach of analysis of the
energy balance is suggested, based on the temperature curve alone and easily
obtainable data. Dividing Equation 7.5 by M.c gives the derivative of the obtained
temperature curve versus time, in oC/h:

dT dVS H l U. A dH 2O Le
=
dt M. c
(T Ta )
dt M. c
in oC/h (7.6)
dt M. c
rate of rate of rate of rate of
temperature temperature temperature temperature
change in increase due decrease due to decrease due to
the sample to biological loss of sensible latent heat of
heat heat to evaporation
generation surroundings

The term (U.A)/(M.c) is the test-specific heat transfer coefficient kc in h-1. If the
temperature at the completion of the self-heating test drops again to Ta, then the total
area under the curve dT/dt is equal to zero and it can be shown that the integral over
time for Equation 7.6 reduces to:

VS. H l ( H 2 O + VS. f vs ) L e
= kc .F + in oC (7.7)
M ave . c M ave . c
total biological loss of sensible heat loss of latent heat due to
heat generation to surroundings evaporation

where
VS = change in VS over duration of test, g
H2O = change in H2O over duration of test, g
fvs = a factor used to estimate the amount of water produced by unit
VS degradation. An average value of 0.5 is used based on the
material balance of this study (see Figure 6.35).
Mave = average mass of sample during test, approximately equal to
(Minitial + Mfinal)/2, g
F = area under the temperature curve, oC.h

174
Mave, VS and H2O can be determined from initial and final data of the test and F
is found graphically. The approximate value of kc can also be found graphically as
the slope of the plot of the logarithm of (T-Ta) versus time, after the biological heat
generation rate has become negligible, i.e. when the temperature is rapidly and
exponentially declining towards the end of the self-heating test. Using available c
values (Marshall and Holmes 1979), the total biological heat generation can now be
directly estimated from Equation 7.7. From this, Hl can also be estimated and
compared with known values for a check.

Selected self-heating tests 7bL6 and 8aL6 were used to demonstrate the application
of the above equations and methods. The graphical determination of the test-specific
heat transfer coefficient kc of these respective tests is presented in Figures 7.15 and
7.17. The coefficient kc for tests 7bL6 and 8aL6 was 0.0199 and 0.0197 h-1,
respectively. The biological heat generation rate (sensible) is the sum of rate of
change of temperature (dT/dt) and the heat loss rate kc(T-Ta). The graphical
determination of the biological heat generation rate (sensible) is presented in Figures
7.16 and 7.18 for tests 7bL6 and 8aL6, respectively.

7.2.2 Estimation of Total Biological Heat Generation Rate

If it is assumed that the rate of heat loss from evaporation is proportional to the rate
of sensible biological heat generation, then the rate of total biological heat
generation can be determined by multiplying the rate of sensible biological heat
generation by the adjustment factor ra (ratio of total biological heat generation to loss
of sensible heat to surroundings):
k . F + ( H 2 O + VS. f vs ) L e / M ave . c
ra = c (7.8)
kc.F

It should be pointed out, however, that the adjustment factor r is test-specific and
hence not a constant.

175
4

Y = -0.01985 * X + 3.8101
3 R2 = 0.9853
ln(T-Ta)

0
0 25 50 75 100 125 150
Time in hours

Figure 7.15 Graphical determination of heat loss coefficient kc (the slope of the
curve = 0.0199 h-1) for the self-heating test 7bL6 between 108 and 252 hours.

1.4 80

1.2 Rate of temperature change, TC=dT/dt


Temperature/Heat change, oC/h

Heat loss rate, L=kc(T-Ta )


1.0
Heat generation rate (sensible), B=TC+L 60 Temperature in oC
0.8 Temperature curve

0.6

0.4 40

0.2

0.0
20
-0.2

-0.4

-0.6 0
0 100 200 300 400 500
Time in hours

Figure 7.16 Graphical determination of biological heat generation rate B (as sensible
heat only) for the self-heating test 7bL6.

176
4

Y = -0.01965 * X + 3.4211
3
R2 = 0.9896
ln(T-Ta)

0
0 25 50 75 100 125 150
Time in hours

Figure 7.17 Graphical determination of heat loss coefficient kc (the slope of the
curve = 0.0197 h-1) for the self-heating test 8aL6 between 168 and 312 hours.

1.4 80

1.2 Rate of temperature change, TC=dT/dt


Temperature/Heat change, oC/h

Heat loss rate, L=kc(T-Ta )


1.0
Heat generation rate (sensible), B=TC+L 60 Temperature in oC
0.8 Temperature curve

0.6

0.4 40

0.2

0.0
20
-0.2

-0.4

-0.6 0
0 100 200 300 400 500
Time in hours

Figure 7.18 Graphical determination of biological heat generation rate B (as sensible
heat only) for the self-heating test 8aL6.

177
7.2.3 Estimation of Respirometric Activity (RA)

It can be shown that, by multiplying the obtained biological heat generation rate, in
o
C/h, and the specific heat capacity c results in the specific biological heat
production rate, in J/h.g wet waste sample. From bioenergetic relationships it has
been well established that the biological consumption of 1 g O2 generates about
14,000 J of heat (Haug 1993, Koenig and Tao 1996, material balance in section
6.1.8), or conversely, that the generation of 14,000 J of biological heat consumes 1 g
of oxygen. Therefore, dividing the maximum specific biological heat production by
14,000 J/g O2 results in the maximum respirometric activity, in g O2/h.g wet waste
sample which can easily be converted to units of mg O2/g VS-h of the sample.

7.2.4 Respirometric Activity and Stability Index

The calculated and estimated values of kc, Hl, ra, RA, SI and other parameters of all
the self-heating tests (total 12 tests) after pilot-scale tests 8a and 8b are presented in
Table 7.2. The specific heat transfer coefficient kc ranged from 0.015 to 0.020 h-1.
The range of estimated heat energy generated by the degradation of VS in self-
heating tests was 12000 to 15000 kJ/kg and is close to the heat energy calculated
during pilot-scale tests as indicated in the mass and energy balances (see Figure
6.36). The range of ra is 1.12 to 1.86, usually with a higher value if a high amount of
moisture was evaporated as presented in Table 7.2. The estimated respirometric
activity (RA) of all these self-heating tests compares well with the stability index as
shown in Table 7.2. The RA values are higher (0.75 to 1.45 mg O2/g VS.h) in the
samples of first stage pilot-scale test 8a as the SI fluctuated between I to II (instable
compost). The RA values are lower (0.23 to 0.73 mg O2/g VS.h) in the samples of
second stage pilot-scale test 8b as the SI varied between IV and V (stable compost).
The RA values of the second stage samples are comparable to the values presented
in Table 2.1 cited by Epstein (1996) based on Iannotti et al. (1994), which reported
that values less than 0.5 mg O2/g VS.h are an indication of stability. The obtained
RA values are also comparable with values of 0.41 mg O2/g VS-h for stabilised
sludge samples after thermophilic aerobic digestion (Reza and Mike 1998) or with
EPA criteria for stabilized sludge of 2.3 mg O2/g VS-h (converted from value of
178
1.5 mg O2/g TS-h for 65% VS) (Switzenbaum et al. 1997). Therefore, the estimated
RA from the self-heating test shows a high potential to reliably assess the stability of
compost and other solid wastes.

Table 7.2 Relation among calculated respirometric activity (RA), stability index
and other self-heating test parameters of samples of pilot scale tests 8a and 8b
Test Tmax Imax A72 eH2O kc Hl ra RA SI
o o o
C C/h C.h g h-1 kJ/kg mgO2/g VS-h
8aL1 64.15 1.07 3605 88.88 0.015 12441 1.86 1.45 I
8aL2 63.68 0.97 3440 48.08 0.021 14862 1.28 0.92 I
8aL3 56.65 0.94 3383 24.68 0.020 12806 1.21 0.88 II
8aL4 56.27 0.63 2839 13.09 0.021 12155 1.15 0.75 II
8aL5 68.58 0.96 3504 36.44 0.020 15436 1.40 1.26 I
8aL6 52.35 0.81 3040 8.39 0.020 13870 1.12 0.93 II

8bL1 31.30 0.13 1651 8.66 0.018 12129 1.35 0.23 V


8bL2 36.23 0.11 1627 9.09 0.018 13215 1.27 0.27 V
8bL3 39.39 0.26 1589 0.015 V
8bL4 46.82 0.48 1678 13.51 0.017 13456 1.34 0.73 IV
8bL5 36.65 0.13 1666 10.21 0.014 12800 1.37 0.28 V
8bL6 45.16 0.42 1810 16.13 0.018 12024 1.28 0.65 IV
eH2O = Total H2O evaporation; kc = test specific heat transfer coefficient; Hl = heat energy generated
by the degradation of VS; ra = adjustment factor (ratio of total biological heat generation to loss of
sensible heat to surroundings).

179
CHAPTER 8

ENGINEERING SIGNIFICANCE

8.1 Pilot-scale Tests


The results obtained from the pilot-scale tests can be used for a better understanding
of the accelerated forced aeration composting process in existing composting plants
and may help in the design and operation of future plants. While the findings
confirm some design and operation practices already applied in modern composting
plants, they may also assist in identifying some actual and potential limitations.

In Europe, especially in Germany, Austria and the Netherlands, many large scale
plants (10,000 to 100,000 t/year capacity) employ the static-dynamic aerated pile
composting system with automatic turning machines in enclosed buildings, which is
closely approximated by the two-reactor system of this study, except that turnover
and/or re-mixing of the composting material may take place in shorter intervals of
once to twice every two weeks. Forced aeration is conducted either in the upflow or
downflow mode. In a few cases, reuse of spent air is applied in such a way that the
spent air from the beginning composting stages (initial 2 to 4 weeks of downflow) is
used for the upflow aeration of the final stages (last 7 to 10 weeks), producing a
completely mature compost of stability index V. An almost perfect example of the
two-reactor composting system is the huge composting plant of Bangkok
(approximately 400,000 t/year capacity) which is applying a two stage composting
process, whereby the spent air of the first stage (21 days of downflow) is used for
aeration of the second stage (21 days of upflow), achieving a very stable end
product.

180
Similar to the single reactor system with internal air recirculation (tests 1a, 3a and
5a) of this study, the increasingly popular small scale tunnel composters or container
composters (approximately 1,000 t/year capacity) for accelerated composting apply a
single stage, static composting process with internal recirculation of air. Alternate
upflow/downflow aeration with addition of moisture can be applied to small scale or
large scale enclosed composting facilities, for example box composters and aerated
static pile systems in enclosed halls.

The findings of this study also confirm some additional benefits of enclosed systems.
Controlled air flow in an enclosed system and application of a biofilter resulted in
complete elimination of odor problems. Furthermore, in the enclosed system
employed, condensate and leachate could be collected and recycled effectively as
demonstrated in tests 6a and 7a.

As mentioned above, the results of this study also highlight some of the limitations
associated with the forced aeration system of modern composting plants which
impede further acceleration of the composting process and attainment of
homogeneous product quality. These are: (i) in the large scale plants, internal
recirculation of air is technically not feasible, preventing full maximization of
degradation efficiency in the first stages due to steep vertical temperature gradients;
(ii) in the containerized, small scale plants, lack of turnover of the composting
material leads to loss of porosity and hence prevents efficient composting for
complete degradation after two to three weeks of processing time; and, (iii) general
inability to maintain continuously uniform temperature and moisture profiles over
the whole duration of the process and thus to subject all waste particles to the same
treatment conditions. While these limitations cannot be completely eliminated in any
aerated static pile system, modern design has to find the optimum combination of
frequency of material turnover as well as intensity, duration, and mode of aeration.

8.2 Bench-scale Tests


181
The bench-scale tests were found most suitable for determination of optimum values
of some important parameters such as: moisture content, temperature, etc. It was also
found that bench-scale tests provide easy control over experimental conditions.
Therefore, composting plants may apply bench-scale tests according to their need to
determine the optimum values of the above mentioned parameters for a particular or
a new type of waste mixture.

8.3 Self-heating Tests


The estimation of the respirometric activity according to the method described in
section 7.2 could be applied in many composting plants, where the self-heating test
is already used as an indicator of biological stability. For other or future plants, the
self-heating test could be regularly used to determine the stability index of the
compost as well as to estimate the respirometric activity, since the self-heating test is
a simple method and the required data are easily obtainable. A simple computer
program based on the theory described in section 7.2 could be used for the quick and
routine estimation of the respirometric activity of compost.

182
CHAPTER 9

CONCLUSIONS AND RECOMMENDATIONS

9.1 Conclusions
Based on the results of this study the following conclusions are drawn:

1. Significant vertical distribution of different physico-chemical parameters,


including temperature, moisture, pH and NH3-N was observed in the composting
mass along the airflow directions when unidirectional upflow or downflow
aeration were applied.

2. The upflow aeration mode produces high temperature, moisture content, pH and
NH3-N concentration in the top layers of the composting mass due to upward
forced movement of these parameters, and vice-versa for the downflow aeration
mode.

3. Alternate upflow/downflow aeration with addition of water achieved more equal


distribution of temperature, moisture and other physico-chemical parameters
throughout the height of the composting mass.

4. Application of alternate upflow/downflow aeration and internal air recirculation


accelerates the composting process.

5. A maximum period of 20 days is adequate for first stage composting applying


upflow aeration with internal air recirculation and a recirculation ratio of 20 to
80%.

6. First stage composting of 20 to 30 days duration did not produce a stable end
product as indicated by the self-heating test.

183
7. Composts after second stage composting in the two-reactor system were
completely stable, whereas the end products of second stage composting from
the single reactor system did not achieve complete stability despite the same or
increased total composting time.

8. The two-reactor system improves composting efficiency while lowering air and
energy requirements.

9. Reuse of spent air helps to attain a more uniform vertical temperature


distribution than periodic reversal of airflow direction in the second stage
composting.

10. About 50 to 58% of the released heat during composting in pilot-scale tests was
used for evaporation of moisture.

11. Water removal due to aeration in the first stage tests was almost double than in
the second stage tests.

12. Temperature dependence of the reaction rates of different aeration mode


composting tests followed the Arrhenius equation.

13. First order reaction rates provided the best fit with the actual reaction rates.

14. Percentage degradation of volatile solids depending on temperature could be


predicted well using a first order reaction model.

15. The extent of overall degradation in a closed composting system could be


predicted well based on outlet air temperature instead of average internal
temperature.

16. The extent of degradation in different layers of the composting mass could be
estimated by multilayer analysis using first order reaction rates.

17. Applying a simplified heat energy balance of the self-heating test, the
respirometric activity of compost can be estimated and compared well with
values from the literature.

18. Composting plants using the aerated static pile method need to find the optimum
combination of frequency of material turnover as well as intensity, duration, and
mode of aeration for achieving a stable end product in the shortest possible time.

184
9.2 Recommendations
Recommendations for future research are as follows:
1. Determination of optimum recirculation ratio via extensive composting tests with
variable air recirculation.

2. Effect of air recirculation during downflow aeration on vertical distribution of


physico-chemical parameters and extent of degradation.

3. Vertical distribution of microorganisms under different modes of aeration during


first and second stage composting.

4. Effects of different modes of aeration on composting of various types of wastes.

5. Predictive temperature model based on air flow rates and reaction rates.

6. Establishment of a relationship between estimated respirometric activity based


on the heat energy balance of the self-heating test and the practical respirometric
activity determined from respirometric test.

7. Relation between results of self-heating test and plant compatibility.

185
APPENDIX I

List of Tables and Figures of Appendix I

Table Page

A1.1 Temperature, stability index and other parameters of six self-heating tests
of samples collected from different six layers of pilot-scale test 2a ............ 196
A1.2 Temperature, stability index and other parameters of six self-heating tests
of samples collected from different six layers of pilot-scale test 3a ............ 196
A1.3 Temperature, stability index and other parameters of six self-heating tests
of samples collected from different six layers of pilot-scale test 3b............ 197
A1.4 Temperature, stability index and other parameters of six self-heating tests
of samples collected from different six layers of pilot-scale test 4a ............ 197
A1.5 Temperature, stability index and other parameters of six self-heating tests
of samples collected from different six layers of pilot-scale test 4b............ 198
A1.6 Temperature, stability index and other parameters of six self-heating tests
of samples collected from different six layers of pilot-scale test 5a ............ 198
A1.7 Temperature, stability index and other parameters of six self-heating tests
of samples collected from different six layers of pilot-scale test 5b............ 199
A1.8 Temperature, stability index and other parameters of six self-heating tests
of samples collected from different six layers of pilot-scale test 5c ............ 199
A1.9 Temperature, stability index and other parameters of six self-heating tests
of samples collected from different six layers of pilot-scale test 6a ............ 200
A1.10 Temperature, stability index and other parameters of six self-heating tests
of samples collected from different six layers of pilot-scale test 6b............ 200
A1.11 Temperature, stability index and other parameters of six self-heating tests
of samples collected from different six layers of pilot-scale test 7a ............ 201
A1.12 Temperature, stability index and other parameters of six self-heating tests
of samples collected from different six layers of pilot-scale test 7b............ 201
A1.13 Temperature, stability index and other parameters of six self-heating tests
of samples collected from different six layers of pilot-scale test 8a ........... 202
A1.14 Temperature, stability index and other parameters of six self-heating tests
of samples collected from different six layers of pilot-scale test 8b ........... 202
A1.15 Temperature, stability index and other parameters of six self-heating tests
of samples collected from different six layers of pilot-scale test 9a ........... 203
A1.16 Temperature, stability index and other parameters of six self-heating tests
of samples collected from different six layers of pilot-scale test 9b............ 203
A2.1 Material Balance of test 1a........................................................................... 204
A2.2 Material Balance of test 1b .......................................................................... 205

Figure Page

A.1 Waste food, waste paper and final compost ................................................. 206

195
Table A1.1 Temperature, stability index (SI) and other parameters of six self-
heating tests of samples collected from different six layers of pilot-scale test 2a
Test 2aL1 2aL2 2aL3 2aL4 2aL5 2aL6 Average
Massi, g 527.20 472.20 521.00 455.40 485.20 504.70 494.28
TSi, g 150.30 172.35 192.77 184.44 172.25 176.65 174.79
VSi, % 92.93 93.30 93.01 91.51 92.66 94.48 92.98
VSi, g 139.68 160.81 179.30 168.79 159.60 166.89 162.51
VS, g 14.11 29.21 43.06 34.91 29.58 35.89 37.36
VS, % 10.10 18.17 24.02 20.69 18.54 21.51 22.60
MCi, % 71.49 63.50 63.00 59.50 64.50 65.00 64.50
Porosityi 51.59 57.06 52.68 58.96 55.87 53.92 55.01
Ta, oC 21.74 21.74 21.74 21.74 21.74 21.74 21.74
Tmax, oC 64.98 64.57 68.66 63.29 69.95 63.46 65.82
Tmax-Ta, oC 43.24 42.83 46.92 41.55 48.21 41.72 44.08
o
Imax, C/h 0.73 0.88 1.10 1.01 0.99 0.76 0.91
A72, oC.h 3725 3724 3976 3550 3184 3113 3545
A72-a, oC.h 2156 2155 2407 1981 1615 1544 1976
Atotal, oC.h 16107 19559 21364 18740 18086 19428 18880
Atotal-a, oC.h 7361 10813 12618 9994 9341 10682 10135
tmax, d, 3.17 8.00 4.96 5.29 5.75 4.63 5.30
SI I I I I I I I
a = ambient; L = layers of pilot-scale composting mass(L1 = bottom layer).

Table A1.2 Temperature, stability index (SI) and other parameters of six self-
heating tests of samples collected from different six layers of pilot-scale test 3a
Test 3aL 1 3aL2 3aL3 3aL4 3aL5 3aL6 Average
Massi, g 526.20 477.70 508.70 445.40 469.00 609.80 506.13
TSi, g 183.41 153.84 180.07 158.21 157.09 143.61 162.71
VSi, % 95.35 93.80 94.50 94.04 95.26 95.53 94.75
VSi, g 174.89 144.30 170.17 148.78 149.63 137.19 154.16
VS, g 37.03 31.52 42.96 25.68 24.30 17.25 29.79
VS, % 21.17 21.84 25.25 17.26 16.24 12.57 19.06
MCi, % 65.14 67.80 64.60 64.48 66.51 76.45 67.50
Porosityi 49.47 54.10 51.25 57.36 54.89 40.64 51.29
Ta, oC 19.95 19.95 19.95 19.95 19.95 19.95 19.95
Tmax, oC 64.64 64.94 64.83 60.37 40.63 66.34 60.29
Tmax-Ta, oC 44.69 44.99 44.88 40.42 20.68 46.39 40.34
Imax, oC/h 0.89 1.30 1.18 0.72 0.53 0.99 0.94
A72, oC.h 3047 2977 3277 2951 2586 3542 3063
A72-a, oC.h 1531 1461 1761 1435 1070 2026 1547
Atotal, oC.h 21575 19900 19598 19009 13271 13044 17733
Atotal-a, oC.h 13914 12239 11937 11348 5610 5383 10072
tmax, d, 5.16 4.75 5.71 6.25 4.25 3.38 4.92
SI I I I I III I I

196
Table A1.3 Temperature, stability index (SI) and other parameters of six self-
heating tests of samples collected from different six layers of pilot-scale test 3b
Test 3bL1 3bL2 3bL3 3bL4 3bL5 Average
Massi, g 421.30 331.60 354.70 451.60 514.30 414.70
TSi, g 144.80 138.89 137.67 134.93 127.87 136.83
VSi, % 92.35 91.84 91.17 90.97 91.37 91.54
VSi, g 133.73 127.56 125.52 122.75 116.84 125.28
VS, g -4.44 -0.16 -1.34 3.73 -4.50 -1.34
VS, % -3.32 -0.12 -1.07 3.04 -3.85 -1.07
MCi, % 63.90 57.84 60.31 70.47 74.14 65.33
Porosityi 61.65 70.15 67.98 58.73 52.64 62.23
Ta, oC 19.07 19.07 19.07 19.07 19.07 19.07
Tmax, oC 21.38 21.14 21.26 21.78 22.33 21.58
Tmax-Ta, oC 2.31 2.07 2.19 2.71 3.26 2.51
Imax, oC/h 0.06 0.05 0.05 0.05 0.05 0.05
A72, oC.h 1580 1593 1604 1612 1621 1602
A72-a, oC.h 12 25 36 44 53 34
Atotal, oC.h 7369 7263 7335 7520 7737 7445
Atotal-a, oC.h 273 167 239 424 641 349
tmax, d, 7.50 6.12 6.16 7.45 7.50 7
SI V V V V V V

Table A1.4 Temperature, stability index (SI) and other parameters of six self-
heating tests of samples collected from different six layers of pilot-scale test 4a
Test 4aL1 4aL2 4aL3 4aL4 Average
Massi, g 526.70 340.90 447.70 347.00 415.58
TSi, g 194.88 160.56 162.38 142.37 165.05
VSi, % 88.84 87.50 87.01 89.58 88.23
VSi, g 173.14 140.50 141.28 127.54 145.61
VS, g 44.18 18.00 26.48 23.45 28.03
VS, % 25.52 12.81 18.74 18.39 18.87
MCi, % 63.00 52.90 63.73 58.97 59.65
Porosityi 33.99 54.57 44.29 55.32 47.04
Ta, oC 20.62 20.62 20.62 20.62 20.62
Tmax, oC 67.96 65.68 69.64 70.50 68.45
Tmax-Ta, oC 47.34 45.06 49.02 49.88 47.83
Imax, oC/h 1.10 0.88 0.86 1.20 1.01
A72, oC.h 4065 2820 3048 2557 3123
A72-a, oC.h 2532 1287 1515 1024 1590
Atotal, oC.h 19824 14416 14895 14209 15836
Atotal-a, oC.h 12154 6746 7225 6539 8166
tmax, d, 2.33 3.37 3.46 3.38 3
SI I I I I I

197
Table A1.5 Temperature, stability index (SI) and other parameters of six self-
heating tests of samples collected from different six layers of pilot-scale test 4b
Test 4bL1 4bL2 4bL3 4bL4 4bL5 4bL6 Average
Massi, g 431.90 388.30 392.50 360.30 383.00 423.60 396.60
TSi, g 152.55 149.50 154.35 144.37 142.85 135.15 146.46
VSi, % 89.30 90.09 89.66 90.05 89.72 90.63 89.91
VSi, g 136.22 134.68 138.40 130.01 128.17 122.49 131.66
VS, g 9.83 2.55 4.19 4.37 6.22 7.36 5.75
VS, % 7.21 1.90 3.03 3.36 4.85 6.01 4.39
MCi, % 64.68 61.50 60.67 59.93 62.70 68.10 62.93
Porosityi 60.95 65.01 64.70 67.61 65.45 61.42 64.19
Ta, oC 21.91 21.91 21.91 21.91 21.91 21.91 21.91
Tmax, oC 39.17 28.49 29.24 27.60 38.71 39.19 33.73
Tmax-Ta, oC 17.26 6.58 7.33 5.69 16.80 17.28 11.82
Imax, oC/h 0.21 0.09 0.09 0.08 0.21 0.27 0.16
A72, oC.h 1809 1775 1800 1773 1851 1949 1826
A72-a, oC.h 170 136 161 134 212 310 187
Atotal, oC.h 19740 18333 18402 17564 19438 18939 18736
Atotal-a, oC.h 4494 3087 3156 2318 4192 3693 3490
tmax, d, 6.45 9.88 8.04 5.83 6.58 4.58 6.89
SI IV V V V IV IV IV

Table A1.6 Temperature, stability index (SI) and other parameters of six self-
heating tests of samples collected from different six layers of pilot-scale test 5a
Test 5aL1 5aL2 5aL3 5aL4 5aL5 5aL6 Average
Massi, g 351.10 370.40 365.90 370.90 352.80 552.70 393.97
TSi, g 154.76 164.75 161.58 170.24 158.76 155.59 160.95
VSi, % 94.02 93.80 92.73 94.28 91.91 91.20 92.99
VSi, g 145.50 154.54 149.83 160.51 145.91 141.89 149.70
VS, g 21.66 31.08 4.91 31.24 32.88 21.07 23.81
VS, % 13.99 18.86 3.04 18.35 20.71 13.54 14.75
MCi, % 55.92 55.52 55.84 54.10 55.00 71.85 58.04
Porosityi 68.34 66.63 67.10 66.61 68.38 49.35 64.40
Ta, oC 26.62 26.62 26.62 26.62 26.62 26.62 26.62
Tmax, oC 67.02 58.37 67.54 68.07 66.63 57.89 64.25
Tmax-Ta, oC 40.40 31.75 40.92 41.45 40.01 31.27 37.63
Imax, oC/h 1.14 0.91 1.17 1.38 1.19 0.58 1.06
A72, oC.h 3743 3136 3550 3959 3957 3218 3594
A72-a, oC.h 1829 1222 1636 2045 2043 1304 1680
Atotal, oC.h 21481 23558 20053 21554 21985 21089 21620
Atotal-a, oC.h 8186 10263 6758 8259 8690 7794 8325
tmax, d, 4.50 5.17 3.66 3.66 4.42 6.00 4.57
SI I I I I I II I

198
Table A1.7 Temperature, stability index (SI) and other parameters of six self-
heating tests of samples collected from different six layers of pilot-scale test 5b
Test 5bL1 5bL2 5bL3 5bL4 5bL5 5bL6 Average
Massi, g 344.00 333.60 325.10 296.20 344.50 369.20 335.43
TSi, g 150.58 146.61 144.96 134.88 144.26 135.18 142.75
VSi, % 91.80 90.86 91.42 92.01 91.75 92.78 91.77
VSi, g 138.23 133.21 132.53 124.10 132.36 125.41 130.97
VS, g 8.30 4.64 5.06 9.16 1.63 16.03 7.47
VS, % 6.01 3.48 3.81 7.38 1.23 12.78 5.78
MCi, % 56.23 56.05 55.41 54.46 58.12 63.39 57.28
Porosityi 69.12 70.13 70.88 73.46 69.00 66.47 69.84
Ta, oC 24.95 24.95 24.95 24.95 24.95 24.95 24.95
Tmax, oC 54.75 56.05 46.93 36.58 37.49 45.19 46.17
Tmax-Ta, oC 29.80 31.10 21.98 11.63 12.54 20.24 21.22
Imax, oC/h 0.68 0.64 0.39 0.21 0.22 0.23 0.39
A72, oC.h 3075 2977 2556 2089 2193 2146 2506
A72-a, oC.h 1229 1131 710 243 347 300 660
Atotal, oC.h 8952 8919 14063 14416 14674 16104 12855
Atotal-a, oC.h 2798 2765 2813 3166 3424 4854 3303
tmax, d, 2.58 2.88 3.08 13.25 4.66 10.67 6.19
SI II II III IV IV III III

Table A1.8 Temperature, stability index (SI) and other parameters of six self-
heating tests of samples collected from different six layers of pilot-scale test 5c
Test 5cL1 5cL2 5cL3 5cL4 5cL5 5cL6 Average
Massi, g 352.30 346.00 355.80 336.60 365.00 352.20 351.32
TSi, g 142.61 148.54 152.82 145.07 151.91 134.22 145.86
VSi, % 91.80 91.59 91.42 92.01 91.75 92.78 91.89
VSi, g 130.91 136.05 139.71 133.49 139.38 124.53 134.01
VS, g 8.22 7.64 12.08 8.06 5.49 10.90 8.73
VS, % 6.28 5.62 8.65 6.04 3.94 8.75 6.54
MCi, % 59.52 57.07 57.05 56.90 58.38 61.89 58.47
Porosityi 68.23 68.92 68.05 69.74 67.14 68.07 68.36
Ta, oC 21.81 21.81 21.81 21.81 21.81 21.81 21.81
Tmax, oC 45.22 36.25 33.05 34.76 34.94 36.79 36.84
Tmax-Ta, oC 23.41 14.44 11.24 12.95 13.13 14.98 15.03
Imax, oC/h 0.43 0.11 0.11 0.21 0.19 0.10 0.19
A72, oC.h 1706 1640 1627 1641 1647 1649 1652
A72-a, oC.h 100 34 21 35 41 43 46
Atotal, oC.h 15560 15573 14664 14841 14885 15444 15161
Atotal-a, oC.h 4045 4058 3149 3326 3370 3929 3646
tmax, d, 6.25 14.83 12.75 9.83 8.83 16.83 11.55
SI III V V V V IV IV

199
Table A1.9 Temperature, stability index (SI) and other parameters of six self-
heating tests of samples collected from different six layers of pilot-scale test 6a
Test 6aL1 6aL2 6aL3 6aL4 6aL5 6aL6 Average
Massi, g 319.60 337.00 335.60 346.90 321.60 349.30 335.00
TSi, g 140.54 149.44 148.02 149.32 124.80 133.75 140.98
VSi, % 93.61 93.23 92.65 92.64 94.20 95.80 93.69
VSi, g 131.56 139.32 137.14 138.33 117.57 128.13 132.01
VS, g 11.23 17.63 15.92 16.04 17.16 10.65 14.77
VS, % 8.54 12.65 11.61 11.59 14.60 8.32 11.22
MCi, % 56.03 55.66 55.89 56.96 61.19 61.71 57.91
Porosityi 71.20 69.67 69.83 68.77 70.79 68.15 69.74
Ta, oC 24.09 24.09 24.09 24.09 24.09 24.09 24.09
Tmax, oC 57.13 56.85 54.36 70.3 63.3 47.36 58.22
Tmax-Ta, oC 33.04 32.76 30.27 46.21 39.21 23.27 34.13
Imax, oC/h 0.99 1.03 1.01 1.32 0.81 0.64 0.97
A72, oC.h 3764 3513 3323 4299 3088 2837 3471
A72-a, oC.h 1984 1733 1543 2519 1308 1057 1691
Atotal, oC.h 11105 13798 13585 14282 12295 13346 13068
Atotal-a, oC.h 3623 6316 6103 6800 4813 5864 5586
tmax, d, 1.88 3.04 6.00 1.58 4.50 2.08 3.18
SI II II II I I III II

Table A1.10 Temperature, stability index (SI) and other parameters of six self-
heating tests of samples collected from different six layers of pilot-scale test 6b
Test 6bL1 6bL2 6bL3 6bL4 6bL5 6bL6 Average
Massi, g 335.30 261.10 272.10 302.30 307.60 329.40 301.30
TSi, g 123.99 117.78 125.96 134.95 135.31 141.25 129.87
VSi, % 93.61 93.23 92.65 92.64 94.20 95.80 93.69
VSi, g 116.07 109.81 116.70 125.02 127.46 135.31 121.73
VS, g 4.30 0.84 2.73 4.66 5.20 8.09 4.30
VS, % 3.70 0.77 2.34 3.73 4.08 5.98 3.43
MCi, % 63.02 54.89 53.71 55.36 56.01 57.12 56.69
Porosityi 69.51 76.53 75.61 72.84 72.24 70.13 72.81
Ta, oC 22.77 22.77 22.77 22.77 22.77 22.77 22.77
Tmax, oC 35.18 29.9 27.4 38.36 37.03 53.23 36.85
Tmax-Ta, oC 12.41 7.13 4.63 15.59 14.26 30.46 14.08
Imax, oC/h 0.16 0.15 0.15 0.19 0.19 0.91 0.29
A72, oC.h 1768 1719 1719 1852 1882 3207 2024
A72-a, oC.h 90 41 41 174 204 1529 346
Atotal, oC.h 16395 15238 14544 17091 16002 16348 15936
Atotal-a, oC.h 3238 2081 1387 3934 2845 3191 2779
tmax, d, 8.16 8.16 17.63 9.00 6.04 2.29 8.55
SI IV V V IV IV II IV

200
Table A1.11 Temperature, stability index (SI) and other parameters of six self-
heating tests of samples collected from different six layers of pilot-scale test 7a
Test 7aL1 7aL2 7aL3 7aL4 7aL5 7aL6 Average
Massi, g 417.80 424.80 427.00 408.00 410.90 414.80 417.22
TSi, g 178.45 178.96 182.81 170.10 161.64 145.32 169.55
VSi, % 93.82 93.89 93.11 92.26 90.20 90.97 92.37
VSi, g 167.42 168.02 170.22 156.94 145.80 132.20 156.77
VS, g 37.00 39.38 37.90 27.02 12.09 23.92 29.55
VS, % 22.10 23.44 22.27 17.21 8.29 18.09 18.57
MCi, % 57.29 57.87 57.19 58.31 60.66 64.97 59.38
Porosityi 62.27 61.61 61.51 63.23 63.01 62.37 62.33
Ta, oC 20.61 20.61 20.61 20.61 20.61 20.61 20.61
Tmax, oC 70.4 70.1 69.49 70.7 66.29 62.25 68.21
Tmax-Ta, oC 49.79 49.49 48.88 50.09 45.68 41.64 47.60
Imax, oC/h 1.02 0.93 1.28 0.94 0.77 0.79 0.96
A72, oC.h 3770 3388 3773 3396 3074 3083 3414
A72-a, oC.h 2094 1712 2097 1720 1398 1407 1738
Atotal, oC.h 25449 26276 22962 20937 16312 18473 21735
Atotal-a, oC.h 15469 16296 12982 10957 6332 8493 11755
tmax, d, 3.29 4.88 2.75 3.71 3.67 3.96 3.71
SI I I I I I I I

Table A1.12 Temperature, stability index (SI) and other parameters of six self-
heating tests of samples collected from different six layers of pilot-scale test 7b
Test 7bL1 7bL2 7bL3 7bL4 7bL5 7bL6 Average
Massi, g 385.30 392.90 366.50 378.70 406.20 384.90 385.75
TSi, g 166.14 164.86 160.53 161.82 169.30 161.03 163.95
VSi, % 91.61 90.51 90.48 90.56 90.17 91.71 90.84
VSi, g 152.20 149.21 145.24 146.55 152.65 147.67 148.92
VS, g 16.98 5.63 6.71 10.72 7.40 14.49 10.32
VS, % 11.16 3.78 4.62 7.31 4.85 9.82 6.92
MCi, % 56.88 58.04 56.20 57.27 58.32 58.16 57.48
Porosityi 65.40 64.74 67.20 66.05 63.56 65.36 65.39
Ta, oC 20.52 20.52 20.52 20.52 20.52 20.52 20.52
Tmax, oC 51.55 52.57 47.01 60.62 55.79 59.55 54.52
Tmax-Ta, oC 31.03 32.05 26.49 40.10 35.27 39.03 34.00
Imax, oC/h 0.73 0.79 0.71 0.91 0.62 0.91 0.78
A72, oC.h 3030 3035 2834 3268 2748 3121 3006
A72-a, oC.h 1454 1459 1258 1692 1172 1545 1430
Atotal, oC.h 14049 14454 13474 16483 14712 15107 14713
Atotal-a, oC.h 4270 4675 3695 6704 4933 5328 4934
tmax, d, 2.88 3.00 2.92 4.04 3.92 3.71 3.41
SI II II III II II II II

201
Table A1.13 Temperature, stability index (SI) and other parameters of six self-
heating tests of samples collected from different six layers of pilot-scale test 8a
Test 8aL1 8aL2 8aL3 8aL4 8aL5 8aL6 Average
Massi, g 436.60 413.60 381.80 386.20 403.00 390.70 401.98
TSi, g 193.59 202.17 186.47 174.76 162.37 147.41 177.79
VSi, % 92.62 91.86 92.08 91.58 91.42 90.66 91.70
VSi, g 179.30 185.71 171.70 160.04 148.43 133.65 163.14
VS, g 38.00 35.79 27.26 19.91 20.30 13.58 25.81
VS, % 21.19 19.27 15.87 12.44 13.68 10.16 15.44
MCi, % 55.66 51.12 51.16 54.75 59.71 62.27 55.78
Porosityi 60.77 63.13 65.95 65.42 63.68 64.71 63.94
Ta, oC 21.24 21.24 21.24 21.24 21.24 21.24 21.24
Tmax, oC 64.15 63.68 56.65 56.27 68.58 52.35 60.28
Tmax-Ta, oC 42.91 42.44 35.41 35.03 47.34 31.11 39.04
Imax, oC/h 1.07 0.97 0.94 0.63 0.96 0.81 0.90
A72, oC.h 3605 3440 3383 2839 3504 3040 3302
A72-a, oC.h 2027 1862 1805 1261 1926 1462 1724
Atotal, oC.h 23939 25936 22242 18292 19158 16768 21056
Atotal-a, oC.h 13516 15513 11819 7869 8735 6345 10633
tmax, d, 2.96 4.46 3.79 3.58 3.67 4.00 3.74
SI I I II II I II I

Table A1.14 Temperature, stability index (SI) and other parameters of six self-
heating tests of samples collected from different six layers of pilot-scale test 8b
Test 8bL1 8bL2 8bL3 8bL4 8bL5 8bL6 Average
Massi, g 340.80 332.40 333.60 320.50 333.80 365.70 337.80
TSi, g 145.73 143.46 140.71 135.70 139.76 152.24 142.93
VSi, % 89.64 89.71 90.17 89.65 89.85 89.59 89.77
VSi, g 130.63 128.70 126.89 121.65 125.57 136.39 128.30
VS, g 6.69 8.02 - 9.79 7.24 15.06 9.36
VS, % 5.12 6.23 - 8.04 5.77 11.04 7.24
MCi, % 57.24 56.84 57.82 57.66 58.13 58.37 57.68
Porosityi 69.51 70.28 70.10 71.31 70.09 67.23 69.75
Ta, oC 22.20 22.20 22.20 22.20 22.20 22.20 22.20
Tmax, oC 31.30 36.23 39.39 46.82 36.65 45.16 39.26
Tmax-Ta, oC 9.10 14.03 17.19 24.62 14.45 22.96 17.06
Imax, oC/h 0.13 0.11 0.26 0.48 0.13 0.42 0.26
A72, oC.h 1651 1627 1589 1678 1666 1810 1670
A72-a, oC.h 205 181 143 232 220 364 224
Atotal, oC.h 19081 20469 19448 21599 20666 22618 20647
Atotal-a, oC.h 3075 4463 3442 5593 4660 6612 4641
tmax, d, 10.00 12.79 8.58 9.83 11.79 8.25 10.21
SI V V - IV V IV IV

202
Table A1.15 Temperature, stability index (SI) and other parameters of six self-
heating tests of samples collected from different six layers of pilot-scale test 9a
Test 9aL1 9aL2 9aL3 9aL4 9aL5 9aL6 Average
Massi, g 414.50 392.80 381.60 367.40 344.50 418.80 386.60
TSi, g 168.62 173.22 154.85 158.72 146.86 174.51 162.80
VSi, % 94.92 95.29 95.39 95.10 95.83 95.44 95.33
VSi, g 160.05 165.06 147.72 150.94 140.74 166.56 155.18
VS, g 38.87 41.64 33.66 36.02 37.20 46.46 46.77
VS, % 24.29 25.23 22.79 23.87 26.43 27.89 30
MCi, % 59.32 55.90 59.42 56.80 57.37 58.33 57.86
Porosityi 62.39 64.47 65.33 66.75 68.75 62.00 64.95
Ta, oC 22.19 22.19 22.19 22.19 22.19 22.19 22.19
Tmax, oC 37.85 59.62 51.3 60.55 56.3 58.91 54.09
Tmax-Ta, oC 15.66 37.43 29.11 38.36 34.11 36.72 31.90
Imax, oC/h 0.37 0.82 0.68 0.89 0.96 0.84 0.76
A72, oC.h 2156 3099 2625 2729 2511 2901 2670
A72-a, oC.h 710 1653 1179 1283 1065 1455 1224
Atotal, oC.h 22387 34092 30767 31851 29436 35276 30635
Atotal-a, oC.h 7420 19125 15800 16884 14469 20309 15668
tmax, d, 13.30 10.80 10.80 9.33 7.83 9.96 10.34
SI IV II III II II II III

Table A1.16 Temperature, stability index (SI) and other parameters of six self-
heating tests of samples collected from different six layers of pilot-scale test 9b
Test 9bL1 9bL2 9bL3 9bL4 9bL5 9bL6 Average
Massi, g 299.60 311.90 316.30 328.70 304.60 331.80 315.48
TSi, g 131.34 139.58 145.88 146.11 135.36 133.22 138.58
VSi, % 91.42 91.26 92.49 92.48 92.54 92.70 92.15
VSi, g 120.07 127.37 134.93 135.12 125.27 123.50 127.71
VS, g 7.33 6.18 9.99 7.35 12.76 10.81 10.88
VS, % 6.11 4.85 7.40 5.44 10.18 8.75 8.55
MCi, % 56.16 55.25 53.88 55.55 55.56 59.85 56.04
Porosityi 73.13 72.08 71.65 70.48 72.64 70.01 71.66
Ta, oC 23.62 23.62 23.62 23.62 23.62 23.62 23.62
o
Tmax, C 37.31 41.5 42.05 50.05 45.36 48.68 44.16
Tmax-Ta, oC 13.69 17.88 18.43 26.43 21.74 25.06 20.54
Imax, oC/h 0.23 0.37 0.49 0.70 0.50 0.67 0.49
A72, oC.h 2192 2596 2666 3117 2728 3014 2719
A72-a, oC.h 486 890 960 1411 1022 1308 1013
Atotal, oC.h 18488 17758 18006 19084 19207 19880 18737
Atotal-a, oC.h 4313 3583 3831 4909 5032 5705 4562
tmax, d, 4.67 2.88 2.79 2.00 2.79 1.88 2.84
SI IV IV III III III III III
Table A2.1 Material Balance of test 1a
203
O2 in kg 14.07
VS in kg 11.03
O2/VS 1.28
O2 gm 1276
CO2 gm 1754 C 39.9
TS initial in kg 30.8 H 58.4
TS final in kg 19.7 O 28.6 39.86 39.9 28.6 0.35
Ni % as TKN 1.45 N 0.4 O2 CO2 H2O NH3
Nf % as TKN 1.99 M.W. 1000 1276 1754 516 6.01
N gm 54.57
N gm/kg VS 4.95 Check total 2276 2276
mole N/kg VS 0.353
xH2O in RHS 515.64
H in LHS 58.35 H in VS C 6.0 0.151
O in VS 28.65 H 8.8
O 4.3 6 6 4.3 0.1
Heat Balance per kg VS N 0.05 O2 CO2 H2O NH3
C H O Total
g MW 478.8 58.4 457.6 994
% of 1000 gm 47.8 5.8 45.8 99.4
Adjust to 99.8% 48.0 5.8 46.0 100 C 6.0
Assume 0.2% S H2O 4.3 6 6 4.3 0.1
Using DuLong formula (low heating value) NH3 0.1 O2 CO2 H2O NH3
QL = 145C+610(H-O/8)+57.7S-10.6(9H)
QL = 6449 Btu/lb
QL = 15000 kJ/kg
in gm unit
Water Balance Total Basis (kg) 1754 CO2
H2O i mix 46.4 O2
H2O Metabolism 5.67 1275.6 1000 VS 6.0 NH3
H2O in 4.15
Total initial 56.22 516 H2O
H2O f mix 22.56
Leachate 0.45 15000 kJ Evaporation
Condensate 25.31 7450 7550 per kg DVS
H2O escape 7.9 H2O Heat 3.01 kg or
Total final 56.22 evp. Loss 2474 kJ/kg H2O
Moisture in inlet air 2.43 Mass and Energy
Balance

Heat required in mass change


VS 11
Ave temperature Enthalpy (kcal/kg) in kg (Total)
Tin and Tout in oC Water Steam 7.9 H2O escape
26.16 26.17 609 H 2 O in 46.4 25.3 Condensate
46.54 46.49 617 4.15 in initial mix 0.45 Leachate
Heat=Only evp. at Tin + all steam at Tout 5.67 H2O 22.6 H2O in
H2O evp. at Tin 19346 kcal metabolism final compost
Steam at Tout 312.6 kcal
Heat in kcal 19659 Water balance
Heat in kJ 82173 in out
Heat kJ/kg VS 7450 Check 56.22 56.2

Table A2.2 Material Balance of test 1b


204
O2 in kg 5.47
VS in kg 4.34
O2/VS 1.26
O2 gm 1260
CO2 gm 1733 C 39.39
TS initial in kg 20.01 H 59.87
TS final in kg 15.71 O 28.21 39.39 39.4 28.2 1.15
Ni % as TKN 1.99 N 1.15 O2 CO2 H2O NH3
Nf % as TKN 2.09 M.W. 1000 1260 1733 508 19.55
N gm 69.86 Check
N gm/kg VS 16.10 Total 2260 2260
mole N/kg VS 1.15
xH2O in RHS 507.82
H in LHS 59.87 H in VS C 6.0 0.152
O in VS 28.21 H 9.1
O 4.3 6 6 4.3 0.2
Heat Balance per kg VS N 0.18 O2 CO2 H2O NH3
C H O Total
g MW 472.7 59.9 451.4 984
% of 1000 gm 47.3 6.0 45.1 98.4
Adjust to 99.8% 48.0 6.1 45.7 100 C 6.0
Assume 0.2% S H2O 4.3 6 6 4.3 0.2
Using DuLong formula (low heating value) NH3 0.2 O2 CO2 H2O NH3
QL = 145C+610(H-O/8)+57.7S-10.6(9H)
QL = 6473 Btu/lb
QL = 15056 kJ/kg
in gm unit
Water Balance Total Basis (kg) 1733 CO2
H2O i mix 35.55 O2
H2O Metabolism 2.21 1260.4 1000 VS 19.5 NH3
H2O in 4.23
Total initial 41.99 508 H2O
H2O f mix 30.55
Leachate 0.45 15056 kJ Evaporation
Condensate 7.5 6312 8744 per kg DVS
H2O escape 3.485 H2O Heat 2.53 kg or
Total final 41.99 evp. Loss 2494 kJ/kg H2O
Moisture in inlet air 1.7 Mass and Energy
Balance

Heat required in mass change


VS 4.3
Ave temperature Enthalpy (kcal/kg) in kg (Total)
Tin and Tout in oC Water Steam H2O in 3.49 H2O escape
18.41 18.44 605 4.23 35.55 7.50 Condensate
37.81 37.79 614 in initial mix 0.45 Leachate
Heat=Only evp. at Tin + all steam at Tout 2.21 Meta 30.55 H2O in
H2O evp. at Tin 6447 kcal final compost
Steam at Tout 106.7 kcal Water balance
Heat in kcal 6553
Heat in kJ 27393 Check in out
Heat kJ/kg VS 6312 41.99 42

205

Vous aimerez peut-être aussi