Composting Review

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 91

This article was downloaded by: [McGill University Library]

On: 14 November 2012, At: 04:56


Publisher: Taylor & Francis
Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered
office: Mortimer House, 37-41 Mortimer Street, London W1T 3JH, UK

Critical Reviews in Environmental


Science and Technology
Publication details, including instructions for authors and
subscription information:
http://www.tandfonline.com/loi/best20

Solid Waste Management by Composting:


State of the Art
a a
S. Gajalakshmi & S. A. Abbasi
a
Centre for Pollution Control & Energy Technology, Pondicherry
University, Kalapet, Pondicherry, India
Version of record first published: 15 Jul 2008.

To cite this article: S. Gajalakshmi & S. A. Abbasi (2008): Solid Waste Management by Composting:
State of the Art, Critical Reviews in Environmental Science and Technology, 38:5, 311-400

To link to this article: http://dx.doi.org/10.1080/10643380701413633

PLEASE SCROLL DOWN FOR ARTICLE

Full terms and conditions of use: http://www.tandfonline.com/page/terms-and-conditions

This article may be used for research, teaching, and private study purposes. Any
substantial or systematic reproduction, redistribution, reselling, loan, sub-licensing,
systematic supply, or distribution in any form to anyone is expressly forbidden.

The publisher does not give any warranty express or implied or make any representation
that the contents will be complete or accurate or up to date. The accuracy of any
instructions, formulae, and drug doses should be independently verified with primary
sources. The publisher shall not be liable for any loss, actions, claims, proceedings,
demand, or costs or damages whatsoever or howsoever caused arising directly or
indirectly in connection with or arising out of the use of this material.
Critical Reviews in Environmental Science and Technology, 38:311400, 2008
Copyright Taylor & Francis Group, LLC
ISSN: 1064-3389 print / 1547-6537 online
DOI: 10.1080/10643380701413633

Solid Waste Management by Composting:


State of the Art

S. GAJALAKSHMI AND S. A. ABBASI


Centre for Pollution Control & Energy Technology, Pondicherry University,
Kalapet, Pondicherry, India
Downloaded by [McGill University Library] at 04:56 14 November 2012

One of the most versatile and remunerative techniques for han-


dling biodegradable solid wastes is composting. A large variety of
such wastes of plant, animal, and synthetic origins can be gainfully
composted, at scales varying from a household bin to a large indus-
try. Compost also has an inexhaustible market as a soil conditioner
and fertilizer. Apart from being a source of nitrogen, phosphorus,
potassium, and other nutrients for plants, compost is also believed
to suppress soil-borne diseases in plants. These virtues make com-
posting an ideal option for processing the enormous quantities of
biodegradable solid wastes that are generated in the world. In this
article, a review of the state of the art of compost science is pre-
sented, encompassing the history, the mechanism, the production,
the characteristics, and the effects of composting. It is hoped that
this effort would stimulate further interest in composting and in the
appreciation of its benefits.

1. INTRODUCTION

Unlike industrial or hospital solid waste, which is produced from clearly iden-
tifiable point sources and is regulated by governmental agencies, solid waste
associated with day-to-day living is generated by everyone. Due to increas-
ing population pressure on the land, and the ever-increasing loads of waste
generated every minute of the day, it has become difficult for governmental
agencies to cope with the challenge of handling the enormous quantities of
the waste. The situation is especially poignant in the third world; we see mu-
nicipal solid waste (MSW) strewn almost everywhere in the cities and towns

Address correspondence to Prof S. A. Abbasi, Centre for Pollution Control & Energy Tech-
nology, Puducherry University, Kalapet, Pondicherry 605 014, India. E-mail: Prof.S.A.Abbasi@
gmail.com

311
312 S. Gajalakshmi and S. A. Abbasi

of most developing countries. In the developed world, incessant generation


of massive quantities of MSW has brought the sanitary landfills and other
MSW handling facilities to the limits of their capacities.
Apart from being an unseemly sight, MSW also poses great risk to soil,
water, and human health. A major chunk, often more than half by weight, of
this MSW is made up of compostable materials (CPCB, 2006). Massive quan-
tities of biodegradable solids are also generated in the form of aquatic and
terrestrial weeds, leaf litter, and crop wastes. If left unharvested, the weeds
seriously pollute and deplete the land and the water resources. In developing
countries, leaf litter and crop waste is often burned in the open air to generate
fertilizer in the form of ash, but this not only destroys a great deal of carbon
and other nutrients but is also a source of air pollution and global warming.
Downloaded by [McGill University Library] at 04:56 14 November 2012

An important way in which MSW can be managed is to find techniques


and methodologies for utilizing such waste at scales varying from very small
(a single household) to very large (several neighborhoods or an entire town).
Composting appears to be among the most potent of the options. It is one of
the few technologies of which the benefit-cost ratio remains higher than 1 at
all scales of operation. It is also, like rearing of livestock, a technology that
can be managed by persons of a skill level ranging from plebeian to highly
sophisticated.
Even though the practice of leaving piles of putrecible organic waste to
biodegrade, into what we later came to call compost is believed to have
originated as soon as humankind learned its first lessons in agriculture, com-
posting has not been exploited for biowaste utilization to even a fraction of
its potential. For example, in the United States, which has been among the
largest producers of solid waste in the world, composting of MSW as a per-
centage of total waste generated was insignificant during the 1960s through
to the 1980s (U.S. EPA, 2000). It is only from the beginning of 1990 that the
composting activity has picked up. It was a mere 2.1% of the total amount
of MSW in 1990 but more than trebled, to 7.1%, by 2000. Composting is
also finding increasing acceptance in other developed countries for example
Australia (Larney and Hao, 2007). This trend is expected to continue (Fitz-
patrick et al., 2005), and may acquire a more steep upwards slope in the 21st
century, if the benefits of composing are brought more and more strongly to
global attention. This review is one of the attempts in this direction.

2. THE COMPOSTING PROCESS: AN OVERVIEW

Composting involves the conversion of organic residues of plant and animal


origin into manure. It is largely a microbiological process based upon the
activities of several bacteria, actinomycetes, and fungi (Bharadwaj, 1995).
The main product is rich in humus and plant nutrients; the by-products are
carbon dioxide, water, and heat (Abbasi and Ramasamy, 1999).
Solid Waste Management by Composting 313

In the composting process, aerobic microorganisms use organic matter as


a substrate. The microorganisms decompose the substrate, breaking it down
from complex to intermediate and then to simpler compounds (Epstein, 1997;
lpek et al., 2002). During composting, compounds containing carbon and ni-
trogen are transformed through successive activities of different microbes
to more stable organic matter, which chemically and biologically resembles
humic substances (Pare et al., 1998). The rate and extent of these transfor-
mations depend on available substrates and the process variables used to
control composting (Marche et al., 2003).

microbial
Fresh organic + O2 stabilized organic + CO2 + H2 O + heat
metabolism
Downloaded by [McGill University Library] at 04:56 14 November 2012

waste material
(compost)

In nature, composting takes place when leaves pile up and begin to


decay. Eventually the decayed leaves are returned to the soil, where living
roots reclaim the nutrients from the remains of the leaves. Ancient peo-
ple dumped food wastes in piles near their camps, and found that the
wastes rotted and formed habitat for the seeds of many food plants that
sprouted there. Perhaps this led to the realization that dump heaps were
good places for food crops to grow, and humans began to put seeds there
intentionally.
By all accounts, recycling of organic residues through composting
appears to be an ancient practice. It has acquired ever greater rele-
vance, and in the present times the use of composting to turn organic
wastes into resource should be practiced with a sense of urgency as land-
fill space becomes increasingly more scarce and expensive (He et al.,
1995).

2.1. Temperature-Dependent Microbial Stages in a Typical


Composting Process
The first microorganisms to colonize a heap of biodegradable solid waste
are mesophilic bacteria, actinomycetes, fungi, and protozoa. They grow be-
tween 10 and 45 C (Cooperband, 2000) and break down easily degradable
components such as sugars and amino acids (Hellmann et al., 1997). The
degradation of fresh matter starts as soon as it is piled into heap. Due to the
oxidative action of microorganisms, the temperature increases. Eventhough
there is a drop in pH at the very beginning of composting, caused by the
formation of volatile fatty acids, the subsequent degradation of acids brings
about an increase in pH.
When the temperature of a waste heap reaches 4550 C, thermophilic
microorganisms replace mesophilic ones (Hellman et al., 1997). The second
314 S. Gajalakshmi and S. A. Abbasi

phase is called the thermophilic phase and can last several weeks. It is the
active phase of composting: Most of the organic matter is degraded and
consequently most oxygen is consumed in this phase. According to Tuomela
et al. (2000), lignin degradation also starts during this phase. Indeed, the
optimum temperature for thermophilic micro-fungi and actinomycetes which
mainly degrade lignin is 4050 C. Above 60 C, these microorganisms cannot
grow and lignin degradation is slowed down (Hellman et al., 1997).
After the thermophilic phase, which corresponds to the peak of degrada-
tion of fresh organic matter, the microbial activity decreases, as does the tem-
perature. This is termed the cooling phase. The compost maturation phase
then begins when the compost temperature falls to that of the ambient air.
During this phase, mesophilic microorganisms colonize the compost heap
Downloaded by [McGill University Library] at 04:56 14 November 2012

and slowly degrade complex organic compounds such as lignin. This last
phase is important because humus-like substances are produced in this phase
to form mature compost (Cooperband, 2000).
During composting, mineralization and humification occur simultane-
ously and are the main processes causing the degradation of the fresh organic
matter. During mineralization, transformations of nitrogenous compounds
occur involving several biochemical reactions. Degradation of protein, urea,
or uric acid produces ammonium ion (NH+ 4 ) (Hansen et al., 1990). Dur-
ing this process, high pH, high temperature, and moisture determine the
NH3 /NH+ 4 balance and the NH3 emission. The solubility of NH3 is reduced
by about 30% when temperature increases from 40 to 50 C, and when pH also
increases.
Another step of degradation is the nitrification, which transforms NH+ 4
into NO+ 3 (nitrate) by oxidation, under aerobic conditions. One of the by-
products of nitrification is N2 O. Although composting is essentially an aerobic
transformation of organic matter, anaerobic conditions can occur in pockets
of the waste heap where free oxygen is exhausted. It may lead to formation
of volatile fatty acids, which lower the pH of the anaerobic zone. Under these
conditions NO3 is reduced to N2 O and then to N2 . In addition, N2 O, NO, and
NO2 may be produced in a compost heap that is not completely aerobic.
Due to these reasons, steps must be taken to avoid anaerobic zones from
developing in a compost heap.
During composting, carbon is transformed into CO2 and is integrated
into humus-like substances as a result of humification. If anaerobic zones
form in a compost heap, methane can be released from such zones (Lopez-
Real and Baptista, 1996). According to Peigne and Girardin (2004), low redox
potential and high temperature provide suitable conditions for the develop-
ment of thermophilic methanogenic bacteria. Moreover, during the thermo-
hilic phase, oxygen is liberally consumed by aerobic microorganisms; the
subsequent reduction of oxygen concentration in the heap favors anaerobic
conditions for methane production (Ott, 1990).
Solid Waste Management by Composting 315

3. EVOLUTION OF COMPOST SCIENCE

The practice of converting animal manure and other biodegradable wastes


to compost is believed to have originated as early as agriculture (Fitzpatrick
et al., 2005). The earliest known written reference to composting is found
in clay tablets dated to the Akkadian empire, about 4300 years ago (Rodale
et al., 1960), but it is believed that the fertilizer value of aerobically degraded
organic matter, which we now call compost, was recognized much earlier.
There is evidence that Romans, Greeks, and the Bani Israel knew about
compost. The Bible and Talmud both contain numerous references to the
use of rotted manure straw, and mention of compost occurs in 10th- and
12th-century Arab writings, in medieval Church texts, and in Renaissance
Downloaded by [McGill University Library] at 04:56 14 November 2012

literature (Smith et al., 2007).


Indeed, references to the methods of generating compost and the ben-
eficial role of compost in agriculture abound in the Indian and the Chinese
literature from ancient times. Several other references indicate the worldwide
use of compost throughout recorded history. One of the high-profile exam-
ples is the composting system built in 1787 at the Mount Vernon estate of
George Washington, the first president of the United States (Higgins, 2001;
Pogue and Arner, 1997). Under Washingtons directions his farm manager had
constructed a facility to compost stable waste in a pit approximately 4.6
9.1 1.2 m. It had brick walls and a cobblestone floor, and was enclosed
by a wooden roof. Washington referred to this facility as a dung repository
or stercorary. It was situated close to the stables, and a wagon was used to
deliver the manure and the bedding into the pit. Later, it was used to remove
the finished compost for use in adjacent fields. Other luminaries who pro-
duced compost and promoted its use include Thomas Jefferson and James
Madison (Smith et al., 2007).
Even as the use of compost has been widely prevalent, scientific studies
detailing and quantifying the benefits of compost, and the systematic devel-
opment of composting technology, began only in the early 20th century. The
first attempt to give a scientific basis to the composting process and accord-
ingly design composting systems was made during 19241926 by Howard
and Wad (1931). The studies were conducted at the Institute of Plant Indus-
try situated at Indore, India, and the process developed there has come to
be called the Indore method.
Howards work was, in essence, a systematization of the traditional com-
posting practices being followed in China and India for centuries (Bhardwaj,
1995). It employed layered mixtures of high carbon to nitrogen ratio (C:N)
feedstocks, like leafy plant material, with low C:N materials like animal ma-
nure in an approximately 3:1 ratio. It typically involved: (1) the placement of
a layer of hard-pressed woody material on the ground to provide a base for
the compost pile; (2) putting over it successive layers of 20-cm thickness of
316 S. Gajalakshmi and S. A. Abbasi

carbonaceous material such as crop wastes and leaves, followed by 10 cm


thickness of nitrogenous material such as animal manure or sewage sludge;
(3) continuing the placing of layers as already described until the pile reached
a height of 1.53 m. The material was sprinkled with enough water to make
it damp but not soggy. The pile was covered with soild or hay to retain the
process heat, and was turned at 6-week intervals. The composting took 46
months.
There was also a variant of this method suitable for accomplishing the
composting in pits. Shredding of the material was recommended to expedite
its composting.
Howard wrote and spoke profusely of his work, which inspired other
researchers to study compost science. Numerous projects were established as
Downloaded by [McGill University Library] at 04:56 14 November 2012

a consequence during the 1920s and 1930s, including the ones by Moubray
(1943), and others (Gotaas, 1956).
The drawbacks of the Indore method are that it requires massive inputs
of labor in heap construction or filling of pits, turning, and other maintenance
operations. The method also entails significant losses of nitrogen and ammo-
nia, besides moisture. To overcome some of these disadvantages a variant of
this method was developed at Bangalore, India, by Acharya in 1939 (Misra
and Roy, 2007).
In the Bangalore method, trenches of about 1 m depth, and 1.52.5 m
width, are made at an appropriate place, generally on the outskirts of the
city. The compostable refuse is dumped into the trench or pit and spread
out with rakes or forked shovels to make a layer of about 15 cm thickness.
Night soil or dung is then placed over the refuse in a layer of about 5 cm.
The process is repeated until the trench or pit is filled up to about 30 cm
above the ground level and a final layer of compostable material is placed
on the top. At each layering, water is sprinkled over the material to make it
optimally moist. The above-ground material is made into a dome shape and
covered with about 2.5 cm mud-plaster. If all operations are properly carried
out, the compost is ready in 5 to 6 months.
A major advantage of the Bangalore method is its ability to process even
night soil.
Modifications of the Indore and the Bangalore methods have been at-
tempted across the world in order to speed up the process. The Chinese
high-temperature composting method (FAO, 1980) is one of the examples.
In it, after the substrate heap is erected, hollow bamboo pipes are inserted
into it vertically and horizontally. The resultant aeration causes the temper-
ature of the heap to rise to 6070 C within 23 days. The pipes are then
removed and the heap is plastered with mud. The plaster is broken after 15
days and the heap is turned thoroughly. If need be, moisture is adjusted to
appropriate levels. The heap is replastered and left for natural decomposition.
The compost is ready in less than half the time taken by the conventional
aerobic methods.
Solid Waste Management by Composting 317

The advancements in composting technology, which were triggered by


the Indore method, saw increasing use of machines and larger scales of op-
eration throughout the 20th century. The VAM composting system started at
the Netherlands in 1932 was a mechanized version of the Indore method
(Gotaas, 1956). Inspite of its early popularity, and continuous upgrades, its
technology became obsolete by 1989 (Haug, 1993). Other adaptations of the
Indore/Bangalore methods also served their purpose for the most part of
the 20th century and continue to do so even now, albeit to an increasingly
lesser extant than more technologically advanced systems. Even as the tradi-
tional open air or non-reactor systems were increasingly mechanized and
instrumented, several types of other systems were also introduced in which
composting was carried out inside vessels or reactors. The first such reactor
Downloaded by [McGill University Library] at 04:56 14 November 2012

was the rotary drum composting system (RDCS) developed in the US in the
1940s. In it the substrate is anaerobically fermented in a large rotary drum for
36 days, followed by windrow composting. After the initial spell of success
RDCS experienced a temporary slump in popularity, but has since staged a
comeback and is in use throughout the world at present (Fitzpatrick et al.,
2005).
By now, with the aid of instrumentation, mechanization, and increasingly
rigorous process control, it is possible to obtain high-quality compost in a
span of 34 weeks.

4. COMPOSTING SYSTEMS

In the broadest of terms, composting systems fall into two categories: the
fully or partially open to air systems and the in-vessel systems. In the
first category are systems ranging from the ones used from prehistoric times
to the windrow, static pile, and household systems used in the present
day. In the second category fall the tunnel systems, the RDCS (mentioned
briefly in the preceding section), and other in-vessel or reactor systems
of various designs. Depending on the location, the substrate, the scale of
operation, and the skills and the machinery available, one or the other type
of system is used.
Another way of categorizing composting system is on the basis of means
of aeration: the agitated or the static (Tchobanoglous et al., 1993). In the
former, the material to be composted is agitated mechanically to introduce
oxygen as well as to (and accordingly) control the temperature, and effect
mixing of the material. In the latter, the substrate remains static and air is
blown through it. The windrow and the static pile processes are examples
of the agitated and the static bed aeration systems, respectively.
It must be emphasized that terms like windrows, static piles, reac-
tor systems, and nonreactor systems used in compost science jargon are
neither precise, nor are they used in the same sense by different authors.
318 S. Gajalakshmi and S. A. Abbasi

For example, Tchobanoglous et al. (1993) describe windrows and static


piles as two distinct composting processes, whereas Manser and Keeling
(1996) group these and numerous other designs based on handling piles
of waste under a single umbrella: windrow composting systems. Likewise,
the terminology reactor and nonreactor systems (Fitzpatrick et al., 2005)
is imprecise because very similar biochemical and microbiological reactions
occur in all substrates undergoing composting, irrespective of the type of
pile or container in which the composting may be taking place. All com-
post heaps or piles are reactors in that sense. For example, when a mass
of waste is set for composting in a discarded packing box, one is in effect
starting a batch reactor. Be it the back-garden waste heap or a sophisticated
reactor of postmodern design, one or other kind of reactor is veritably in
Downloaded by [McGill University Library] at 04:56 14 November 2012

operation.
An overview of the major composting systems is presented next. The
defining characteristics of these systems are compared in Table 1.

4.1. Windrow Composting System


In windrow composting, the biowaste is laid out in parallel rows, 23 m high
and 34 m wide across the base. Windrows naturally acquire a trapezoidal
shape, with angles of repose depending on the nature of material.
The actual dimensions of the windrows depend on the type of equip-
ment that is used to turn the material to be composted. Before forming the
windrows, the material is processed by shredding and screening it to ap-
proximately 3 to 9 cm. The moisture content is adjusted to 5060%. If such
a windrow is left without any mechanical agitation or turning, limited aer-
ation would occur naturally via diffusion and convection currents, but the
composting would proceed very slowly, requiring more than a year for com-
pletion. The process is greatly speeded up by periodic turning, which not
only aerates the waste but also homogenizes it, leading to a more uniform
breakdown. Depending on the size of the operation, turning may be done by
anything from front-end loaders on very small sites, to self-propelled special-
ized turners, which straddle the windrows at larger facilities. The intervals
between turnings can be tailored to the stage of the process, being more
frequent early on, when oxygen demand is high, and becoming longer as
composting proceeds. Normally the windrows are turned up to twice per
week while the temperature rises to, and remains at, about 55 C. Complete
composting is accomplished in 3 to 4 weeks. Thereafter the compost is al-
lowed to cure for an additional 3 to 4 weeks without turning. During this
period, residual decomposable organic materials are further reduced by fungi
and actinomycetes.
As windrows are aerated primarily by natural or passive air movements,
brought about by convection and gaseous diffusion, the rate of air exchange
depends on the porosity of the windrow. Therefore, the size of a windrow
Downloaded by [McGill University Library] at 04:56 14 November 2012

TABLE 1. An Overview of Different Commonly Used Composting Systems


Windrow Static pile and Rotary drum Tunnel
composting contained pile Bin composting composting composting In-vessel
Criteria system systems system system system system
Size and form of The waste is laid out in Waste is laid out in Bins of different Rotary drum with Long perforated Consist of vessels
the heap parallel rows; 23 m parallel rows; sizes and 3 m or larger heavy-duty (reactors) of different
/container high and 34 m wide considerably taller materials are diameter is used conveyor shapes and sizes; most
across the base; and wider rows can used for pretreating enclosed inside a approximate the
acquires trapezoidal be had compared to the waste; the sealed casing of characteristics of
shape windrows; especially waste is then approximately plug-flow (tubular)
in contained pile windrowed square cross reactors or of
systems section moves the continuously stirred
waste though a tank reactors which are
tunnel; the system common in process
approximates a industry
plug flow reactor
Preprocessing Material is shredded Material is mixed using Material is The drum itself is Material is shredded Material is sorted to
and screened standard agricultural hand-sorted to a remove
equipment prevent non- precomposting uncompostables
compostables unit; it
from getting homogenizes
into the the waste and
composting bin sets its
decomposition
process going
Turning/aeration Frequency of turning is No turning is done; to The composting As in the Air is blown through A variety of mechanical
high during the early speed up the mass is either windrows the conveyor or and forced aeration
stages; progressively composting process, left to natural pan, and is systems are used
lesser with time a grid of aeration or aeration or exhausted from
Turning is exhaust piping is turning is done the casing top
accomplished using used, over which at periodic
machines according substrate piles are intervals with
to the scale of formed. simple garden
operation equipment
Composting 34 weeks 34 weeks 68 weeks 34 weeks 23 weeks 23 weeks
period
(Continued on next page)

319
Downloaded by [McGill University Library] at 04:56 14 November 2012

TABLE 1. An Overview of Different Commonly Used Composting Systems (Continued)

320
Windrow Static pile and Rotary drum Tunnel
composting contained pile Bin composting composting composting In-vessel
Criteria system systems system system system system
Curing period 34 weeks without turning 4 weeks or longer 34 weeks 34 weeks 34 weeks 34 weeks
Operation site Carried out in the outskirts Contained pile systems can Ideal for As in the windrows Suitable only Can be installed everywhere
of towns and cities to be used anywhere household where adequate at widely varying scales of
avoid disturbance to publi composting land area is operation
Operated as an addition to available
existing landfill
operations
Major features Thorough mixing of Since turning is not done, it Being small scale, Ability to Highly efficient at Enable large masses of waste
material is possible is less dependent on labor very effective co-compost a low-to-medium to be composted within
Low odor emission hand sorting of mixture of scales of much shorter land spaces
More flexible operation and material is sewage sludge operation Better public acceptance due
more precise control of possible; hence and municipal Overcomes to less forbidding
oxygen and temperature can ensure good waste, which is traditional appearance of the
conditions in the pile quality product otherwise compaction composting site Less
than would be obtained difficult to problems manpower requirements
in a windrow system achieve by other Minimized effect of external
methods factors such as rains and
other extreme weather
conditions
Consistent compost quality
Better odor control
Drawbacks Require large land area; can Decomposition progresses The potential of As in windrows Pose serious High capital and operational
cause odor problems, at slower rate, causing the bin composting feasibility cost.
particularly when material to remain on site can be realized problems at Episodes of odor release can
windrows are turned for a longer period only by ensuring larger scales; occur due to equipment
during periods of calm air Decreased ability to adjust public occupy more failure or system design
and temperature moisture in composting participation, floor surface, limitations
inversion mass after initial mix which until now since they are
Likely to release fungal Potential for drying in the has been long rather than
spores and other immediate vicinity of the difficult to high
bioaerosols aeration systems. achieve
Labor-intensive: some or Composted material can
other activity has to be be heterogeneous.
performed on the site
almost daily
Solid Waste Management by Composting 321

that can be effectively aerated is determined by its porosity. A light fluffy


windrow of leaves can be much larger than a wet dense windrow containing
manure. If the windrow is too large, anaerobic zones may occur near its cen-
ter, interfering with the composting. On the other hand, small windrows lose
heat quickly and may not attain temperatures high enough to kill pathogens
and weed seeds.
Windrows require large areas of land and can cause odor, problems
especially during the turning operations. They are also likely to release fungal
spores and other bioaerosols. Despite these drawbacks, windrows account
for the vast majority of centralized composting systems, possibly because
windrow composting is often carried out either as an addition to existing
landfill operations or on the outskirts of towns and cities, thereby reducing
Downloaded by [McGill University Library] at 04:56 14 November 2012

nuisance to people living nearby.


Machines of diverse designs are used to turn up windrows; these include
conventional earthmoving machines, side-cutting windrow turners, straddle
turners, and preaeration drums. If the substrate is prone to rapid compaction,
such as sewage sludge, bulking agents like wood chips are added to prevent
anaerobic zones from developing in the windrows.
Windrows can also be operated in passive aeration mode. In this op-
tion air is supplied to the composting materials through perforated pipes
embedded in each windrow. Air flows into the open ends of the pipes and
through the windrow because of the chimney effect created as the hot gases
rise upward out of the windrow. When the composting period is completed,
the pipes are simply pulled out, and the base material is mixed with the
compost. This method of composting has been studied and used in Canada
for composting seafood wastes with peat moss, manure slurries with peat
moss, and solid manure with straw or wood shavings. Manure from dairy,
beef, swine, and sheep operations has been used for the purpose.
Shredding of material prior to its placement in windrows speeds up the
composting process significantly (Tognetti et al., 2007).

4.2. Static Pile and Contained Pile Systems


A static pile system, like windrows, also consists of waste laid out in parallel
rows. But the static pile, as its name suggests, is not turned and thus does
not have to conform to the dimensions of a mechanical turner, allowing the
rows to be considerably taller and wider. What mixing is needed can be
achieved using standard agricultural equipment, so these systems tend to
be significantly cheaper with respect to equipment, manpower, and running
cost. However, they enhance the land requirement, since decomposition pro-
gresses at a slower rate, causing the material to remain on site for a longer
period. In the State of Pondicherry, where the authors are located, and in the
nearby State of Tamil Nadu, India, it is a common practice among farmers to
322 S. Gajalakshmi and S. A. Abbasi

make large, semi-elliptical piles of biodegradable wastes, top each pile with
clay, and leave it to compost for about a year (Abbasi et al., 2000).
To speed up the composting, aerated static pile systems are used. These
consist of a grid of aeration or exhaust piping over which substrate piles are
formed. The initial height of the piles should be 58 feet, depending on the
porosity of the material, weather conditions, and the reach of the equipment
used to build the pile. Extra height is advantageous in the wintertime to
retain heat. It may be necessary to top off the pile with 6 inches of finished
compost and clay/garden soil. The topping of finished compost protects the
surface of the pile from drying, insulates it from heat loss, discourages flies,
and filters ammonia and potential odors generated within the pile.
Aeration can occur passively via the piping system, but the piles are
Downloaded by [McGill University Library] at 04:56 14 November 2012

usually provided with an individual blower for more effective aeration con-
trol. Disposable corrugated plastic drainage pipe is used commonly for air
supply. Blower operation is controlled by a timer, or in some systems by a
microcomputer, to ensure a preset temperature profile. Since the piles are
not subjected to additional turnings, the selection and initial mixing of raw
materials are critical. Otherwise, poor air distribution and uneven compost-
ing can occur. The pile must have good structure, as well as be laid in a
manner that porosity is maintained through the entire composting period.
This generally requires a fairly stiff bulking agent such as straw or wood
chips. Because of their large size, wood chips pass through the process only
partially composted. They are usually screened from the finished compost
and reused as bulking agents for an additional two or three cycles. The mat-
eral is composted for 3 to 4 weeks, and is then cured for another 4 weeks
or longer. Shredding and screening of the cured compost is done to im-
prove the quality of the final product. To improve process and odor control,
all or significant portions of the system in newer facilities are covered or
enclosed.
A variant of the static pile is the contained pile. In contained pile
systems the material to be composted is contained between walls over a
perforated floor through which air is blown. The advantage of contained piles
over static piles is that since the material is stacked up between the walls,
there is no longer any problem of the pile width reducing with height due to
sliding of material along the angle of repose. Aeration is also more efficient
because the presence of walls ensures that all the pumped air is forced to
go upwards preventing any sideways diffusion. Leachate is also easier to
manage. Most significantly, a contained pile system can be cut off from the
atmostphere by putting a roof over its walls, thereby achieving significant
odor containment. Consistent with the prevailing lack of standardization in
the terminology of compost science, some authors refer to enclosed static
piles as bins. But it appears logical to restrict the use of bin composting
to the systems in which commonly used waste bins or discarded cartons are
employed.
Solid Waste Management by Composting 323

4.3. Bin Composting


At household level, or at small livestock-raising units, composting is usu-
ally done in bins of different sizes and materials ranging form cement bins
to emptied cartons. As in unturned windrows and static piles, composting
can occur simply by natural aeration, but does so very slowly. Turning the
waste at periodic intervals (which is usually done with shovels) and periodic
sprinkling of water to maintain moisture at adequate levels can reduce the
composting time to 68 weeks.
Bin composting represents a simple and low technology that can be
accomplished with very ordinary and common contraptions. Bin compost-
ing also represents one the most primitive of composting methods that is
Downloaded by [McGill University Library] at 04:56 14 November 2012

used in the present day, greater in sophistication only in comparison to the


practice of simply leaving heaps of putrescible organic material on land to
compost.
A major advantage of bin composting is that with it being a small-
scale operation, very effective hand sorting of the material is possible to
prevent undesirables from going into the composter bin. This can en-
sure a fairly good quality product even when handled by barely skilled
persons.
Bin composting has the potential of lessening the enormous quantities
of biodegradable waste that is presently discarded by the households and
that either lies strewn on roadsides and other common property areas as in
the developing countries or stresses to breaking point the waste processing
facilities as in the developed world. If most of the households can individually
compost their biodegradable waste, or do so within consortia of households,
a great proportion of solid waste can be treated at the source. Attempts have
been made everywhere to realize this potential of household composting
but have met with very limited success thus far.
For example, Gajalakshmi (2003) had launched a major campaign to
segregate and compost the waste generated at the residential quarters of the
Pondicherry University faculty. Bins of two different colors were provided to
each household with compostables and non-compostables written clearly
on each bin in English and Tamil. Members of each household were met with
by volunteers who explained the virtues of the initiative and emphasized that
by cooperating with it everyone would be contributing to a better environ-
ment. There was an initial burst of enthusiasm, which soon cooled off, and
eventually all the bins ended up collecting mixed waste. Worse, by and by,
the earlier habit of tossing off the waste onto the nearby land returned in full.
Other trial schemes in the past have had similar fate: great initial enthusiasm
tapering off, leaving most bins unused or misused (Wright, 1998). One hopes
that with the solid waste problem getting out of hand, and public awareness
for environmental protection increasing by the day, household composting
would soon begin to make a significant impact.
324 S. Gajalakshmi and S. A. Abbasi

4.4. Rotary Drum Composters


The rotary drum, for which processes of this type are named, is a pretreat-
ment device, rather than the core of the technology. A large-diameter (3 m
or larger), slowly rotating drum is fed at one end with crude organic waste,
water, and, in some facilities, sewage sludge. As the waste mixture is tumbled
in its passage down the length of the drum, the material is gradually broken
down and is mixed intimately with oxygen and water. The consequent in-
crease in granularity causes increasingly intense biological activity leading to
a well-established decomposition process within a few hours in the drum.
The waste is then windrowed until composting is complete.
Drum-based systems are among the oldest of the mechanical processing
Downloaded by [McGill University Library] at 04:56 14 November 2012

technologies (Fitzpatrick et al., 2005), and after many years of being under-
used they are back in fashion. Their major attraction lies in their ability to
co-compost a mixture of sewage sludge and municipal waste, with an ease
almost impossible to achieve with other systems.

4.5. Tunnel Composters


A common problem associated with the sizing of composting systems is
that the substrate gets increasingly compacted as the composting proceeds
further. This entails idling of an increasing fraction of the system space in
each run.
In an attempt to overcome this problem, tunnel composters were devel-
oped. These consist of a long perforated heavy-duty conveyor enclosed in a
sealed casing of approximately square cross section. Air is blown through the
conveyor pan, and is exhausted from the casing top, while waste is loaded
onto one end of the conveyor and compost is discharged from the other. In
common with cylindrical static digesters, a high-temperature zone through
which the material must pass exists, but in this technology its position de-
pends upon the speed of the floor conveyor.
The tunnel composter is a fully continuous process; the waste is moved
through the machine at a consistent rate. A static digester, even with continu-
ous recirculation through a series of units, can never quite match thisthere
is always some point at which the materials are moving more slowly than
elsewhere. However, a tunnel composter is limited by the maximum prac-
tical width to which it is possible to build a conveyor, and by the power
requirements to drive it. In addition, power must be provided to overcome
friction between the material and the side casing, while in a static digester
that power is supplied by gravity (Manser and Keeling, 1996).
While tunnel composters prove highly efficient at low to medium scales
of operation, they begin to pose serious feasibility problems at larger scales.
The running speed of the floor conveyor is determined by the in-feed rate
of raw material, not by the volume loss during the process as it is in a static
Solid Waste Management by Composting 325

digester. It is not possible to slow the material down progressively as its


volume reduces. Therefore, the total enclosed volume must be equivalent to
the volume of waste input multiplied by the number of days of treatment.
If a process is to receive 917 m3 of waste per day and to treat it for up to
10 days, the tunnel of the composter would have to be 1.3 miles (2140 m)
long!
A number of devices operating in series could overcome the problem of
excessive length, and running each successive machine at a slower conveyor
speed could take advantage of material volume losses. But the mechanical
complexity would develop beyond reasonable operating reliability. There
will simply be too many moving parts. It would therefore appear that tunnel
composters, while they can overcome the traditional compaction problems
Downloaded by [McGill University Library] at 04:56 14 November 2012

experienced with static digesters, do so at the expense of capacity and me-


chanical simplicity. They also occupy much more floor surface, since they
are long rather than high, and may therefore bear an additional cost burden
(Manser and Keeling, 1996).

4.6. In-Vessel Systems


In-vessel composting systems are amenable to more rigorous designing and
engineering than conventional systems. This helps in achieving better pro-
cess efficiency, process control, and optimization. Most such systems either
approximate the characteristics of plug flow (tubular) reactors or of contin-
uously stirred tank reactors which are common in process industry. Perhaps
for this reason in-vessel systems are also referred as reactor systems.
The first composting process to use a vessel was the rotary drum com-
posting system (RDCS) discussed earlier. As mentioned, the drum was es-
sentially a pretreatment device rather than a composting unit. Indeed, the
composting bins described earlier are more true reactor systems, as they
are very close to ideal batch reactors in their shape and operation. In later
years a number of in-vessel composting systems based on the principles of
reactor engineering have been developed with vessels as diverse in shapes
and sizes as vertical towers, horizontal-rectangular tanks, horizontal-circular
tanks, and circular rotating tanks. Some of the methods combine the virtues
of the windrow and the aerated pile systems in an attempt to overcome the
deficiencies and exploit the attributes of each system.
Various forced aeration and mechanical turning devices are used to op-
timize aeration in the in-vessel systems.
A common feature of all these systems is that the waste is made to
undergo decomposition within an enclosed space, which makes it possi-
ble for the process conditions to be rigorously controlled. The systems also
enable large masses of waste to be composted within much shorter land
spaces than conventional composting methods. But the use of machinery
326 S. Gajalakshmi and S. A. Abbasi

and power places significant cost burden on in-vessel systems, making


them more expensive than the conventional systems on a ton for ton
basis.

4.7. Vermicomposting
Vermicomposting is the term given to the process of conversion of biodegrad-
able matter by earthworms into vermicast. In the process, a major fraction of
the nutrients contained in the organic matter is converted to more bioavail-
able forms.
The first step in vermicomposting occurs when earthworms break the
substrate down to small fragments as a prelude to ingesting the substrate.
Downloaded by [McGill University Library] at 04:56 14 November 2012

The earthworms possess a grinding gizzard that enables the mincing of the
substrate. This increases the surface area of the substrate, facilitating micro-
bial action (Chan and Griffiths, 1988). The substrate is then ingested and
goes through a process of digestion brought about by numerous species
of bacteria and enzymes present in the worm gut. During this process, im-
portant plant nutrients such as nitrogen, potassium, phosphorus, and cal-
cium present in the feed material are converted into forms that are much
more water-soluble and bioavailable to the plants than those in the parent
substrate.
The earthworms derive their nourishment from the microorganisms that
grow upon the substrate particles. At the same time, they promote further
microbial activity in the residuals so that the fecal material, or casts, that
they produce is much more fragmented and microbially active than what
was ingested. Worms can digest several times their own weight each day,
and since the retention time of the waste in the earthworm is short, large
quantities are passed through an average population of earthworms. As dis-
cussed earlier, in the normal composting process the substrate has to be
turned regularly or aerated in some way in order to maintain aerobic condi-
tions. In vermicomposting the earthworms take over the roles of both turning
and maintaining the organics in an aerobic condition, thereby eliminating the
need for mechanical or forced aeration.
Vermicomposting is not an exothermic process and, unlike composting,
does not lead to any perceptible rise in the vermireactor temperature. To
ensure that the earthworms remain maximally active, the vermireactor should
be kept at conditions of temperature and soil moisture as close to the given
earthworm species naturally preferred habitat as possible.
Almost any agricultural, urban, or industrial organic material can be ver-
micomposted, but many may need some form of preprocessing to make them
acceptable to earthworms. Such preliminaries can involve washing, precom-
posting, and macerating or mixing. Precomposting is particularly beneficial
in facilitating vermicomposting (Gajalakshmi et al., 2002a, 2005a; Nair et al.,
2006; Tognetti et al., 2007).
Solid Waste Management by Composting 327

4.7.1. SPECIES SUITABLE FOR VERMICOMPOSTING


It is generally known that epigeic species have a greater potential as decom-
posers of animal manure and similar soft organic wastes than anecics and
endogeics. This is due to the predominantly humus-consuming and surface-
dwelling nature of the epigeics. The commonly used epigeic species are
Eisenia foetida, Eudrilus eugeniae, and Perionyx excavatus (Abbasi and Ra-
masamy, 1999; Gajalakshmi and Abbasi, 2004a; Hartenstein et al., 1979; Graff,
1974; Haimi and Huhta, 1986; Kale et al., 1982; Reinecke and Venter, 1987;
Kumar, 1994). These species are prolific feeders and can feed upon a wide
variety of degradable organic wastes. They exhibit high growth rate. Eisen-
sia foetida has a wider tolerance for temperature than E. eugeniae and P.
Downloaded by [McGill University Library] at 04:56 14 November 2012

excavatus, which allows the species to be cultivated in areas ranging in soil


temperature from 540 C.
Epigeics are also able to vermicompost harder waste such as waste
paper but anecics steal a march over them in feeding upon such wastes. This
is because anecics, being geophytophagous, have greater ability to break
harder substrate than epigeics (Gajalakshmi et al., 2001a, 2001b, 2002b).
4.7.2. CONVENTIONAL STEPS INVOLVED IN VERMICOMPOSTING
Vermicomposting can be done in pits, concrete tanks, well rings, or in
wooden or plastic crates appropriate to a given situation (Ismail, 1997). If
done in pits, it is preferable to select a composting site under shade, in the
upland or an elevated level, to prevent water stagnation during rains.
A setup for producing vermicompost should have the following at-
tributes:

1. It should have adequate provision for earthworms to live, feed, and breed;
such provision should confirm to the habits of the earthworm species used
in the setup.
2. It should be kept optimally moist and close to neutral pH.
3. It should safeguard against insects and predators so as to prevent harm
to the earthworms.
4. It should have adequate provision for periodic harvesting of vermicast
and renewal of feed.

According to Ismail (1997), a typical vermicomposting unit may be set


up by first placing a basal layer of vermibed comprising of broken bricks
or pebbles (34 cm) followed by a layer of coarse sand to a total thickness
of 67 cms to ensure proper drainage. This may be followed by a 15-cm
moist layer of loamy soil. Into this soil may be inoculated about 100 locally
collected earthworms (about 50 surface and 50 subsurface varieties).
Small lumps of cattle dung (fresh or dry) may then be scattered over the
soil and covered with a 10-cm layer of hay. Water may be sprayed till the
328 S. Gajalakshmi and S. A. Abbasi

entire setup is moist but not wet. Less water kills the worms and too much
chases them away. The unit may be kept covered with broad leaves like
those of coconut or palmyrah. Old jute bags can also be used for covering.
Watering the unit should be continued and the unit monitored for 30 days.
The appearance of juvenile earthworms by this time may be taken as a
healthy sign. Organic refuse may be added from the 31st day as a spread
on the bed after removing the fronds. The spread should not exceed 5 cm
in thickness at each application, though addition of this amount of matter
can be done everyday. According to the author, it is advisable for a beginner
to spread the feed only twice a week, watering to requirement. After a few
applications, the refuse may be turned once without disturbing the bed. The
day enough refuse has been added into the unit, watering may be done and
Downloaded by [McGill University Library] at 04:56 14 November 2012

45 days later the compost will be ready for harvest.


As the organic refuse changes into a dark brown compost, addition of
water should be stopped (42nd day). This would move the worms into the
vermibed. The compost may be harvested and the harvested compost placed
in the form of a cone on ground in bright sunlight. This will facilitate worms
present in the compost to move to the lower layers. The compost pile may
be spread for about 24 to 36 hours, and the worms may be removed from
the lower layers of the compost.
4.7.3. HIGH-RATE VERMICOMPOSTING
The concept of high-rate vermicomposting has been developed by
Gajalakshmi and Abbasi (Gajalakshmi and Abbasi, 2003; Gajalakshmi et al.,
2001c, 2001d, 2002c, 2005a). It is based on the premise that vermireactors
are essentially tanks in which earthworms are made to feed upon animal
manure and/or other biodegradable solid wastes, and the system does not
require continuous inputs of other forms of energy for their operation. As
such, the cost of the tanks constitutes the major component of the system
cost. Hence in order to maximize benefit from such reactors, it is essential to
minimize the reactor volume for a given vermicast output. The authors have
achieved these objectives to a significant extent by the following means:

1) Attainment of maximum sustainable earthworm density. In an attempt to


improve the efficieny of the vermireactors in terms of vermicast produc-
tion per unit time and per unit digester volume, the feed was first com-
posted and then subjected to vermireactors operated with much higher
earthworm densities: 50150 adults per liter of digester volume, more than
used in conventional vermireactors (7 animlas per litrer of digester vol-
ume). Such high-rate vermireactors were operated with different sub-
strates, different earthworm species, and different earthworm densities
over long spells (of 6 months and more). An earthworm population den-
sity was deemed sustainable if the animals gained weight, produced off-
springs, and displayed negligible mortality. The studies revealed that in
Solid Waste Management by Composting 329

most cases maximum sustainable earthworm densities were as high as


150 animals per liter of reactor volume. At such densities over 90% of the
feed was vermiconverted within 1015 days (Gajalakshmi and Abbasi,
2003; Gajalakshmi et al., 2002c, 2005a). Clarke et al. (2007) report that
in the vermicomposting of sewage sludge by Eisenia foetida, high-rate
vermicomposting is sustainable at feeding rates upto 1 kg wet sludge per
kilogram earthworm per day.
2) Modification in the design of epigeics-based vermireactors. Epigeics, such
as Eudrilus eugeniae, Eisenia fetida, and Perionyx excavatus, being es-
sentially humus-feeding earthworms, dwell in the humus and leaf litter
lying over the topsoil and have very shallow burrows. In order to make
the most of this characteristic of the epigeics, the authors have simplified
Downloaded by [McGill University Library] at 04:56 14 November 2012

the design of the vermireactors by doing away with layers of gravel, sand,
and soil, normally used for preparing vermibeds. Instead,, a moistened
thick cotton cloth saturated with water was laid at the bottom of the ver-
mireactors and the feed was placed over it. This simple modification in
design saved 75% of the reactor space that was earlier being occupied
by gravel, sand and soil in the conventional systems without causing any
adverse impact on either the vermicast production or the health, repro-
duction, or survivability of the earthworms. In other words, the utilizable
fraction of the reactor volume was increased threefold, thereby enabling a
much higher feed throughput per liter of the digester (Gajalakshmi et al.,
2005b).
3) Finding the most effective reactor geometry. Studies on reactors of identical
volume but different surface area-to-height ratios have revealed that the
higher the surface area, the more efficient is the vermicast production per
unit reactor volume, subject to the condition that the height of the feed
column should not be less than 3 cm. At lesser heights, sufficient space
is not available to the earthworms for lying low when they tend to rest
during the day (Elizabeth, 2006).

4.7.4. BENEFITS OF VERMICOMPOST


In India, as also many other parts of the world, vermicasts are believed to
have several components that improve the soil to which they are applied
(Kumar, 1994; Ismail, 1997; Joshi and Kelkar, 1951; Abbot and Parker, 1981).
The perceived, and sometimes demonstrated, benefits include improvement
in the water retention capability of the soil, and better plant availability of the
nutrients in the vermicasts compared to the parent (pre-vermicomposted)
material (Ismail, 1998; Curry and Byrne, 1992). The magnitude of the transfor-
mation of phosphorus forms is considerably higher in the case of earthworm-
inoculated organic wastes, showing that vermicomposting may prove to be an
efficient technology for providing better phosphorus nutrition from different
330 S. Gajalakshmi and S. A. Abbasi

organic wastes (Reinecke et al., 1992; Ghosh et al., 1999). The castings of
earthworms may contain two to three times more available potassium than
the surrounding soil (Basker et al., 1993). Earthworm castings have a higher
ammonium concentration and water-holding capacity than bulk soil samples,
and they constitute sites of high denitrification potential (Elliot et al., 1990).
According to Parkin and Beery (1994), earthworm castings are enriched in
mineral N, and compared with the surrounding soil, vermicompost has lower
C/N ratio and pH than normal compost irrespective of the source of organic
waste. Microbial population is also considerably higher in vermicompost than
in compost (Chowdappa et al., 1999).
In Bangalore, earthworms successfully decomposed sugar factory resid-
uals and turned them into a soil nutrient that allowed farmers using the
Downloaded by [McGill University Library] at 04:56 14 November 2012

material to reduce chemical fertilizers by 50% (Logsdon, 1994).


Vermicasts are believed to contain enzymes and hormones that stimulate
plant growth and discourage pathogens (Ismail, 1997; Abbasi and Ramasamy,
1999; Szczeck, 1999). Vermicompost added to various container media sig-
nificantly inhibited the infection of tomato plants by Fusarium oxysporium f.
sp. Lycopercisi, and the protective effect increased in proportion to the rate
of application of vermicompost (Szczeck, 1999).
Although vermicast generated from animal dung is universally believed
to be beneficial to soil and plants, there are no reports giving evidence that
the same may be true of vermicasts generated from other sources. To explore
this area, the authors have conducted three studies on the impact of the
application of water hyacinth and neem compost/vermicompost on plants
(Gajalakshmi and Abbasi, 2002c, 2004b; Gajalakshmi, 2003).
The first experiment was qualitative, done at the kitchen gardens of
five farmers (Gajalakshmi and Abbasi, 2002). In each location 4-m2 plots
were marked out and the following common vegetables were planted: ladys
finger (Abelmoschus esculentus),brinjal (Solanum melongena), cluster bean
(Cyamopsis tetragonoloba), chili (Capsicum annum), and tomato (Lycoper-
sicon esculentum). Three of the plots were treated with water hyacinth ver-
micompost and two of the plots with an equal quantity of water hyacinth
compost. As these were qualitative studies basically to see whether the wa-
ter hyacinth compost/vermicompost discourages plant growth, no controls
of unfertilized plots were studied.
Observations on the five kitchen gardens revealed total absence of any
harmful effect of compost/vermicompost. Rather, the farmers view was that
the vegetables grew better than normal on the treated plots.
In the second study, saplings of ornamental plant Crossandra undu-
lafolia were grown with and without the presence of water hyacinth com-
post/vermicompost (Gajalakshmi and Abbasi, 2002). It was found that the
pots containing soil amended with water hyacinth compost had Crossandra
plants achieving significantly better height, larger number of leaves, more
favorable shoot:root ratio, greater biomass per unit time, and larger length
Solid Waste Management by Composting 331

of inflorescence. In terms of root length, quicker onset of flowering, and


harvest index, too, the treated plants on an average performed better than
the controls but the enhancement was not statistically significant.
The positive impact was more pronounced in plants treated with vermi-
compost; indeed in respect of all the nine parameters there was statistically
significant (at 95% confidence level) enhancement in performance. Of par-
ticular interest is the enhancement in the flower yield and harvest index
by vermicompost, as these attributes directly enhance the benefits from the
cultivation of Crossandra.
The third experiment was conducted to study the impact of the appli-
cation of vermicompost obtained from neem (Azadirachta indica) on the
growth and yield of the brinjal plant Solanum melongena Linn (Gajalakshmi
Downloaded by [McGill University Library] at 04:56 14 November 2012

and Abbasi, 2004b). Morphological and yield attributes were studied in brin-
jal saplings treated with neem vermicompost as compared to the untreated
saplings.
The plot supplemented with neem vermicompost had plants achieving
significantly better height, root length, greater biomass per unit time, quicker
onset of flowering, and enhancement in fruit yield. In terms of fertility coeffi-
cient and harvest index, too, in treated plots, there was statistically significant
enhancement in performance. With the supplementation of neem vermicom-
post after 2 months in control plots, there was increase in plant height, root
length, total biomass, and number of flowers and fruits produced.
Deolalikar and Mitra (1997a) have used vermicompost prepared from
paper mill solid waste for fertilizing aquacultural tanks and found an increase
in net primary productivity from 32.08 to 220.83 mg C/m/h. Vermicompost
application also showed better growth of Rohu fish (Labeo rohita) when
compared with other commercially availalable organic manures (Deolalikar
and Mitra, 1997b).

5. FACTORS THAT CONTROL THE COMPOSTING PROCESS

In order to control and optimize the composting process toward achieving


a product of desired quality, it is necessary to understand the factors that in-
fluence the process in one way or the other. A compost heap is a miniature
ecosystem where interactions between biotic and abiotic factors bring about
the desired changes. By providing a favorable environment for the growth
and activities of the desired biota in the system, good quality compost can
be produced. The criteria used in the evaluation of the composting process,
compost stability (maturity), and quality are based on the physical and chem-
ical characteristics of the organic material (Bertoldi et al., 1983; Forster et al.,
1993; Grebus et al., 1994; Tiquia and Tam, 2000a). These parameters include
a drop in temperature, degree of self-heating capacity, oxygen consumption,
332 S. Gajalakshmi and S. A. Abbasi

cation-exchange capacity, organic matter, nutrient contents, and C:N ratio


(Tiquia et al., 2002a).
The abiotic and biotic factors playing key role in the composting process
are described next.

5.1. Abiotic Factors


5.1.1. NATURE OF THE SUBSTRATE
All kinds of organic residues amenable to the enzymatic activities of the
microorganisms can be converted into compost if suitable conditions for
biodegradation are provided. As the substrate becomes the only source of
food to the microorganisms in a compost heap, the nature of the substrate
Downloaded by [McGill University Library] at 04:56 14 November 2012

is the most basic controlling factor in any composting process. Most of the
substrates are largely made up of polymers, which are insoluble in water. The
extracellular enzymes released by the microbes hydrolyze these polymers
into monomers, which then dissolve into water and enter the microbial cell
where further decomposition takes place (Ginkel, 1996).
The maturity of the compost also depends upon the nature of the
substrate (Zucconi and Bertoldi, 1987). Use of compostagronomic or
horticulturalis based on the composts chemical composition (Barker,
1997). If the substrate is of plant origin, then the main constituents are the car-
bonaceous compounds such as cellulose, hemicellulose, and lignin. Nitroge-
nous constituents (proteins) occur to a lesser extent. Protein constituents, cel-
lulose, and hemicellulose decompose easily. Although cellulosic substrates
form good raw material for composting, lignin, being a complex aromatic
polymer, is resistant to microbial attack to a considerable extent. However,
it is not entirely recalcitrant to microbial decomposition; it undergoes slow
degradation. The elevated temperature found during the thermophilic phase
is essential for rapid degradation of lignocellulose (Tuomela et al., 2000). A
number of fungi, particularly those belonging to the Basidiomycetes group,
are well known for their ability to decompose lignin (Muthukumar and Ma-
hadevan, 1983). Some bacteria and actinomycetes also posses lignolytic char-
acteristics (Bharadwaj, 1995).
The organic compounds in the biowaste could be divided into three
main fractions: (1) carbohydrates (polymers and simple sugars), (2) lignin,
and (3) nitrogen compounds. In the beginning of the composting process,
simple carbohydrates are converted to carbon dioxide and water, and degra-
dation of nitrogenous compounds results mainly in the production of ammo-
nia. In the later stages of composting, cellulose and hemicellulose are utilized
by the compost microflora and finally lignin is also subjected to degradation.
Besides mineralization, organic matter is converted to humic substances.
The porosity of the substrate plays a major role in the composting pro-
cess. Porosity facilitates gas exchange with the atmosphere, enabling the
aerobic metabolism to become dominant, liberating heat profusely.
Solid Waste Management by Composting 333

Materials that should not be included while setting up a composting pile


include soil, ashes from a stove or fireplace, and manure from carnivorous
(meat-eating) animals. Manure from herbivorous animals such as rabbits,
goats, cattle, horses, elephants, or fowl can be used, as it is much leaner in
proteins than the manure from carnivores. Once a pile is started, no further
substrate should be added; the reason is that it takes a certain length of
time for the substrate to break down and anything added has to start at the
beginning, thus lengthening the decomposition time for the whole pile.
5.1.2. CARBON/NITROGEN RATIO
The relative proportion of carbon and nitrogen is a major controlling factor
in the composting process. Carbon serves primarily as an energy source for
Downloaded by [McGill University Library] at 04:56 14 November 2012

the microorganisms, while a small fraction of the carbon is incorporated to


the microbial cells. Nitrogen is critical for microbial population growth, as it
is a constituent of protein that forms over 50% of dry bacterial cell mass. If
nitrogen is limiting, microbial populations will remain small and it will take
longer to decompose the available carbon. Excess nitrogen, beyond the mi-
crobial requirements, is often lost from the system as ammonia gas. Nitrogen
mineralization generally occurs in two phases, a rapid exponential immobi-
lization or mineralization phase, followed by a slow linear mineralization
phase. The C/N ratio of the substrate determines whether immobilization or
mineralization will dominate in the early stages of composting. The rate of
inorganic N release to the soil from composted manure depends on the rate
of decomposition of the organic matter and on subsequent turnover of the
decomposed C and N in soil. Release of plant available N from manure in
the soil is controlled by the balance of N immobilization and mineraliza-
tion, which in turn is controlled, to a large extent, by the C/N ratio of the
decomposing organic material (Cambardella et al., 2003). In the compost-
ing process, the substrate should achieve a C/N ratio of 30:1 for stimulating
degradation and immobilization of nitrogen. According to Golueke (1992),
rapid and entire humification of a substrate essentially depends on it ini-
tially having a C/N ratio between 25 and 35. The range of initial and final
ratios associated with the composting of different substrates is illustrated in
Table 2.

5.1.3. MOISTURE
Moisture is one of the composting variables that affects microbial activities, as
it provides a medium for the transport of dissolved nutrients required for the
metabolic and physiological activities of microorganisms. It is essential for the
decomposition process, as most of the decomposition occurs in the thin liquid
films on the surfaces of particles. Moisture content of 6070% is generally
considered ideal to start with. At later stages of decomposition, the ideal
moisture content may be 5060%. Moisture management requires a balance
between microbial activity and oxygen supply. Very low (<30%) or high
Downloaded by [McGill University Library] at 04:56 14 November 2012

TABLE 2. An Overview of the Types of Substrates Used for Composting in Recent Years (20002005), and Some Key Physicochemical Parameters

334
Associated With the Respective Processes

Number Authors Substrate Temperature profile Moisture pH C/N

1. Alburqueque et al., Olive mill by-product initial8 weeks 9 22.7


2006 (alperujo) + cotton 40 C 812
waste weeks
thermophilic
until 26th week
mesophilic
2. Castaldi et al., 2005 MSW + vegetal waste 4050%
(1:1)
3. Dresboll and Wheat straw + Reached 68 C; 7.68.9
Thorup-Kristensen, clover-grass >40 C for more
2005 than 3 weeks
4. Goyal et al., 2005 Sugarcane trash + cattle Maintained at 60% Initial 51.1
dung (4:1 dry weight) Final 28.3
5. Goyal et al., 2005 Sugarcane trash + cattle Maintained at 60% Initial 32.0
dung (1:1 dry weight) Final 20.5
6. Goyal et al., 2005 Press mud Maintained at 60% Initial 14.5
Final 11.7
7. Goyal et al., 2005 Poultry waste Maintained at 60% Initial 13.9
Final 20.8
8. Goyal et al., 2005 Water hyacinth Initial 2830 C; Maintained at 60% Initial 18.1
46 C in 14 days Final 16.1
9. Parades et al., 2005 Sewage sludge + cotton >55 C for 15
gin waste + olive mill days;
wastewater thermophilic up
to 56 days
10. Parades et al., 2005 Sewage sludge + cotton >55 C for 15
gin waste days,
thermophilic up
to 56 days
11. Perez-Murcia et al., 30% sewage sludge + 70 C 6.83 5.1
2005 70% chopped straw
Downloaded by [McGill University Library] at 04:56 14 November 2012

12. Raviv et al., 2005 Grape marc + cow 55 C for 12 days 5060% initially, Initial 7.9 Initial 23
manure then 4050% Final 7.1 Final 15.6
13. Raviv et al., 2005 Orange peels + cow 55 C for 12 days 5060% initially, Initial 6.7 Initial 30
manure then 4050% Final 7.1 Final 11.9
14. Raviv et al., 2005 Wheat straw + cow 55 C for 12 days 5060% initially, Initial 7.8 Initial 43
manure then 4050% Final 6.8 Final 13.1
15. Ros et al., 2005 Pine bark + urea 5.9 34.9
16. Ros et al., 2005 Pruning waste + coffee 8.4 12.7
wastes (3/1 dry wt)
17. Ros et al., 2005 Pruning waste + coffee 8.1 11.9
wastes (4/1 dry wt)
18. Ros et al., 2005 Pruning waste 8.1 27.1
19. Rotenberg et al., 2005 Fresh paper mill residuals 7.0 19.2
20. Rotenberg et al., 2005 Fresh paper mill residuals 7.0 36.3
+ bark bulking agent
21. Wei and Liu, 2005 Sewage sludge + wood Not above Not lower than 40% 10.1, 9.7, 9.4
waste (1:2 wet volume) 6065 C
22. Abdelhanmid et al., Oil rape seed, poultry 60 C 5060% Initial 7.1
2003 manure Final 8.0
23. Garcia et al., 2004 Sewage sludge + chip 6.9
wood (bulking agent)
24. Koivula et al., 2004 Catering waste 70 C at day 14; Initial 4.5
peaks on days Final 5.7
2,5,8,11,14
25. Kulcu and Yaldiz, 2004 Grass 54%, pepper 20%, Reached 50 C No addition of Final 8.9 0.01
tomato 10.6%, eggplant after day 6 water
15.4%
26. Kwon and Lee, 2004 Food waste Initial 44 C; 58.51% throughout Initial 7.94
58 C at day 2
27. Manios, 2004 Cucumber plant biomass Final 8.2 Initial 16.4
Final 12.37
28. Manios, 2004 Sewage sludge >45 C for 68 Final 7.5 Final 12.87
weeks
(Continued on next page)

335
Downloaded by [McGill University Library] at 04:56 14 November 2012

TABLE 2. An Overview of the Types of Substrates Used for Composting in Recent Years (20002005), and Some Key Physicochemical Parameters

336
Associated With the Respective Processes (Continued)

Number Authors Substrate Temperature profile Moisture pH C/N

29. Mondini et al., Cotton wastes 4065%


2004
30. Tang et al., 2004 Fresh cattle manure Maintained at 60 C Initial 70% Initial 7.58.0 Initial 28 Final 18
in thermophilic Maintained at Final 9.3
waste; after day 1, 6079%
65 C
31. Wang et al., 2004 Manure solids from dairy 52.5 1.14 Initial 8.54 0.09 Initial 25.1 0.85
farm + wheat straw Final 8.34 0.34 Final 8.5 1.01
32. Wang et al., 2004 Manure solids from dairy 55.8 0.75 Initial 8.76 0.06 Initial 33.0 1.22
farm + hardwood Final 8.57 0.03 Final 12.7 0.34
sawdust + wood
shavings
33. Wang et al., 2004 partially composted pig 47.8 2.71 Initial 8.88 0.07 Initial 18.7 0.66
manure + shredded Final 8.72 0.11 Final 15.3 1.68
wood (mostly oak)
34. Barrington et al., Liquid swine manure; pin 67 C after 2 days Initial 63% Initial 18
2003 shavings as bulking
agent
35. Barrington et al., Liquid swine manure; pin 67 C after 2 days Initial 63% Initial 18
2003 shavings as bulking
agent
36. Barrington et al., Liquid swine manure; 67 C after 2 days 65% during the
2003 wood shavings as process
bulking agent
37. Benito et al., 2003 Pruning waste, leaves & Initial 6.9 0.1 Initial 46.8 Final
grass clippings Final 8.1 0.2 32.9
38. Cambardella et Swine manurestraw 53.3 2.10% Initial 8.06 0.5 Initial 13.5 0.1
al., 2003 mixture blended with during process; final 8.2 1.05 Final 9.6 1.2
20% top soil compost final 65.2 2.44
39. Diaz et al., 2003 Liquid vinasse, solid 55 C 45% Initial 6.25 0.1 Initial 15.71 2.20
residue cotton waste final 8.38 1.72 Final 12.16 2.20
Downloaded by [McGill University Library] at 04:56 14 November 2012

40. Ellorieta et al., 2003 Pepper, bean, and cucumber Max 65 C at day 2; 4
wastes peaks on days 2, 5, 8,
10; 30 C on day 14;
65 C on day 18
41. Fukomoto et al., 2003 Fresh swine manure Max 70 C on day 25 Initial 6.5% Initial 8.1
Peaks on days 5, 8, Final 53.9% Final 6.9
17, 25, 30
42. Gaind and Gaur, 2003 Wheat straw Initially maintained Initial 7.48 Initial 35.4
at 100% (w/v) Final 6.92 Final 16.4
43. Gaind and Gaur, 2003 Wheat straw + Mussoorie Initial 7.30 Final 22.80
rock phosphate (2% P2 O5 ) Final 6.50
44. Gaind and Gaur, 2003 Wheat straw + Mussoorie Initial 7.44 Final 17.7
rock phosphate (2% Final 6.92
P2 O5 )+ 10% fly ash
(inoculated with Bacillus
polymyxa)
45. Gaind and Gaur, 2003 Wheat straw + Mussoorie Initial 7.48 Final 16.4
rock phosphate (2% Final 6.88
P2 O5 )+ 20% fly ash
(inoculated with Bacillus
polymyxa)
46. Gaind and Gaur, 2003 Wheat straw + Mussoorie Initial 7.52 Final 16.7
rock phosphate (2% Final 6.79
P2 O5 )+ 40% fly ash
(inoculated with Bacillus
polymyxa)
47. Gaind and Gaur, 2003 Wheat straw + Mussoorie Initial 7.57 Final 16.8
rock phosphate (2% Final 6.87
P2 O5 )+ 60% fly ash
(inoculated with Bacillus
polymyxa)
(Continued on next page)

337
Downloaded by [McGill University Library] at 04:56 14 November 2012

338
TABLE 2. An Overview of the Types of Substrates Used for Composting in Recent Years (20002005), and Some Key Physicochemical Parameters
Associated With the Respective Processes (Continued)

Number Authors Substrate Temperature profile Moisture pH C/N

48. Gestel et al., 2003 Biowaste (vegetable, fruit, Max 75 C on day 5560% during the Final 17
garden waste and 12; 3 peaks on process
paper) days 4, 12, 16
49. Gomez and Bernal, 65% solid olive mill 55 C maintained Initial 5.89 Initial 40.1
2003 waste, 35% olive leaves Final 8.46 Final 14.1
50. Hartileb et al., 2003 Municipal biowaste Max 70 C; 29 55% during process; Initial 6.8 Initial 25
days- final 40%
thermophilic
phase and
3078 days for
cooling phase
51. Hernandez et al., 2003 Coffee pulp 60 C 70%; reduced to
5861% by the
end
52. Horiuchi et al., 2003 Wood chips, wheat lirean Initial 20 C Max Initially 60% ter
62 C after day 1
53. Levy and Taylor, 2003 Bedding materials + 5.2 15.1
pelted carcasses from
mink farm
54. Levy and Taylor, 2003 Horse manure + bedding 5.8 12.2
straw from a stable
55. Levy and Taylor, 2003 Sewage sludge + MSW 7.4 6.7
56. Luo and Netravali, Chicken manure >55 C during the Initially 50%
2003 process
57. Ma et al., 2003 Anthracene in benzene Maintained
sprinkled on garden 56.559.5 C
soil; kitchen waste,
cooked rice and
vegetable waste
Downloaded by [McGill University Library] at 04:56 14 November 2012

58. Manios and Stentiford, Green waste Day 1 55 C; max 68 C Final pH 6.71
2003 at day 8; 6 peaks on
days 8,10,13,15,18,23;
>55 C for 30 days
59. Marche et al., 2003 Paper mill sludge 60 C 55% during the Remained
process constant
60. Mari et al., 2003 Olive press cake 65 C on day 6; did not Water added as Initial 6.0 Initial 28.7 Final
exceed 45 C after first temperature Final 7.5 21.7
6 weeks of >30 C
composting
61. Mari et al., 2003 Olive press cake + olive 65 C Initial 5.3 Initial 29.0
mill waste water Final 8.5 Final 25.5
62. Soumare et al., 2003a Malian farm compost Final 8.44 0.09 Final 16.8
63. Soumare et al., 2003b Belgian compost from an Final 7.20 0.01 Final 16.2
industrial composter
64. Steger et al., 2003 Organic fraction of 55 C till day 9; then
household waste stepwise lowering to
35 C
65. Stelmachowski et al., Sewage sludge 7.777.93 1416.3
2003
66. Zorpas et al., 2003 Dewatered anaerobically Thermophilic phase 15 Initial 4050% Final 7.25 Final 11.73
stabilized primary days Final 31.5%
sewage sludge (DASPS)
67. Akanbie and Togun, Maize stover and poultry Initial 7.5
2002 manure Final 5.6
68. Baheri and Meysami, Flare pit soil silt 63%, clay 50%
2002 37%
(Continued on next page)

339
Downloaded by [McGill University Library] at 04:56 14 November 2012

TABLE 2. An Overview of the Types of Substrates Used for Composting in Recent Years (20002005), and Some Key Physicochemical Parameters
Associated With the Respective Processes (Continued)

340
Number Authors Substrate Temperature profile Moisture pH C/N

69. Beauchamp et al., 2002 Poultry manure 10%, Within 3 days Initial 7.0 Initial 50.0
chicken broiler floor 40 C, reached Final 6.0 Final 24
litter 10% , raw DPS 60 C within 2
80% weeks and
remained above
50 C from week
416
70. Beauchamp et al., 2002 De-inking paper sludge 5560 C even Initial 7.8 1.6 Initial 29.4 1.0
90%(v/v);pig slurry and when ambient Final 7.4 1.7 Final 8.1 1.3
liquid poultry temperature
manure10% was below 0 C
71. Boulter et al., 2002 Chicken manure + bone Final 40.60% Final 24.12
meal ash + bark mix +
soyabean meal +
milorganite
72. Boulter et al., 2002 Chicken manure + Final 40.60% Final 29.22
paunch manure
(remains in rumens of
slaughtered cattle) +
bark mix
73. Coventry et al., 2002 Onion wastes, urea Initial 80%
Final 8084%
74. Diaz et al., 2002a Depotassified beet vinasse Max 47 C on day 55% in thermophilic Initial 6.20 Initial 12.0
20; peaks on phase; no water Final 8.40 Final 8.0
days 20, 40 added in any
other time of
process
75. Diaz et al., 2002a Depotassified beet vinasse 55 C on day 6; 55% in thermophilic Initial 6.52 Initial 19.0
peaks on days phase; no water Final 8.0 Final 12.0
6, 18, 50, 73 added in any
other time of
process
Downloaded by [McGill University Library] at 04:56 14 November 2012

76. Diaz et al., 2002b Vinasse 80%, grape marc Incubated at 55 5 during the Initial 5.2 Initial 23.2
20% thermophilic process Final 8.11 Final 22.7
chambers at
55 C
77. Eghball, 2002 Feedlot manure compost 7.28.3 8.6311.35
78. EL-Masry et al., 2002 Waste materials from leafy Incubated at 55 C 5060%
fruit orchards, garden
fallen leaves, crop
plants
79. Guardia et al., 2002 Green waste (branches, 78, except 5 on
leaves, weeds, grass day 30
cuttings)
80. Hoyas et al., 2002 Waste sludge from gelatin Thermophilic 50% during the Initial 6.70 Initial 30.1
grenatine manufacturers phase day 514; process Final 9.20 Final 13.1
max 65 C on
day 8
81. Jang et al., 2002 Food garbage Thermophilic 6067% Initial 4.8 Initial 8.85
phase day 10; Final 8.8 Final 13.3
58 C from day 8
to 17
82. Korboulewsky et al., Sewage sludge + green 6.3 14.59
2002a wastes + pine bark
83. Korboulewsky et al., Sewage sludge + green 70 C 35.6% 8.0 13.4
2002b wastes + pine bark
84. Lodha et al., 2002 Pearl millet 4851 C at 30 cm Final 18.2
residuesoildungurea 6062 C at 60
cm
85. Lodha et al., 2002 Cluster Final 9.4
beansoildungurea
86. Lodha et al., 2002 Weeds-soildungurea Final 9.4
87. Lodha et al., 2002 Cauliflowersoildung Final 11.6
urea
(Continued on next page)

341
Downloaded by [McGill University Library] at 04:56 14 November 2012

TABLE 2. An Overview of the Types of Substrates Used for Composting in Recent Years (20002005), and Some Key Physicochemical Parameters
Associated With the Respective Processes (Continued)

342
Number Authors Substrate Temperature profile Moisture pH C/N

88. Lopez et al., 2002 Sun-dried plants of green Incubated at 30 C Initial 6.58 Initial 29.1
bean, pepper, and for 3 months Final 8.3, 8.24,
cucumber (1:1:1) 8.83
89. Mbuligwe et al., 2002 Organic household waste 3445 C day 1 to 5070% Initial 6.5 Initial 37.1
4; 4664 C day Final 7.0 Final 20.1
5 to 13; 2533 C
day 14 to 26
90. Mondini et al., 2002 Cotton carding waste and 65% Initial 25
yard waste Final 11.38
91. Parades et al., 2002 Olive mill waste sludge + 65 C after first Not less than 40% Initial 6.6 Initial 23.4
cotton gin (bulking week Final 8.5 Final 11.6
agent)
92. Parades et al., 2002 Olive mill waste sludge + 54 C after first Not less than 40% Initial 6.4 Initial 33.3
maize stover (bulking week Final 8.7 Final 13.0
agent)
93. Reuveni et al., 2002 Cow manure 65% >50 C for 50 days 4555% Final 14.3
Chicken manure 15%
Wheat straw 20%
94. Smars et al., 2002 Organic household waste Above 40 C Initial 65%
during day 56
95. Soumare et al., 2002 Malian farm compost Final 8.44 0.09 Final 16.8
96. Soumare et al., 2002 Belgian compost from an Final 7.20 0.01 Final 16.2
industrial composter
97. Tiquia et al., 2002a Leaves, grass clippings, 70 C after first 60% maintained Initial 9.43 0.20 Initial 30.1
shredded bark day; maintained during Final 8.35 0.25 0.001
till day 14 experiment Final 20.1 0.99
98. Tiquia et al., 2002c Poultry litter (mix of Average Initial 65%; no 7.10 during Initial 15.1
poultry manure, wood temperature further process Final 24.1
shavings, waste feed attained during adjustment in
and feathers) process; middle moisture
63 C, bottom
58 C; top 54 C;
surface of pipes
48 C
Downloaded by [McGill University Library] at 04:56 14 November 2012

99. Veeken and Hamelers, Biowaste of MSW Constant at 55 C 40% at 67 weeks


2002
100. Zorpas et al., 2002 Clinoptilolite, dewatered Maintained 60 C
anaerobically stabilized
primary sewage sludge
101. Arthur et al., 2001 South African tree leaves 2135 C 4060%
throughout the
composting;
after a month
decline in
temperature
102. Beck-Friis et al., 2001 Organic household waste 60 C at day 6 Initial 22
103. Ferrer et al., 2001 Fresh grape waste (90%), 54.1 C on day 11 Initial 73.56% Initial 3.56 Initial 26.94
hen dopppings 10% by Final 7.0 Final 13.57
weight
104. Ghorpade et al., 2001 Extruded poly lactic acid External heat 2056% Initial 6.0
sheets (1.5 mm thick) applied at Final 4.0
constant
temperature of
52 C
105. Hanajima et al., 2001 Fresh cattle manure Max 80 C at pile 78%
centre; 12 days
thermophilic
phase
106. Hartileb et al., 2001 Municipal biowaste 70 C day 0 - 29 Initial 55% Initial 6.8 Initial 25
107. Hassen et al., 2001 Domestic waste Turned when Adjusted to 50% Intial 24.40 3.67
Windrow 1 inner constantly Final 11.03 0.47
Windrow 2 temperature of Final 11.03 0.55
Windrow 3 the pile reached Final 10.73 0.28
or exceeded
70 C
(Continued on next page)

343
Downloaded by [McGill University Library] at 04:56 14 November 2012

TABLE 2. An Overview of the Types of Substrates Used for Composting in Recent Years (20002005), and Some Key Physicochemical Parameters

344
Associated With the Respective Processes (Continued)

Number Authors Substrate Temperature profile Moisture pH C/N

108. Jeong and Kim, 2001 Food waste Initial 17 C; max 55% maintained Initial 18.5
61 C at day 9; during the Final 8.5
totally 6 peaks process
on days 1, 3, 5,
7, 9, 11
109. Khalil et al., 2001 MSW sieved to eliminate 4858 C 23.540.0 C 6.767.10 20.025.0
fraction larger than 5 cm
110. Li et al., 2001 Sewage sludge + sawdust 6070% Initial 7.16 Initial 31.59
Final 4.79 Final 18.67
111. Li et al., 2001 Sewage sludge + pig 6070% Initial 8.18 Initial 29.41
manure +sawdust Final 6.85 Final 14.39
112. Madejon et al., 2001a Depotassified beet Final 8.4 Final 9.6
vinasse and other solid
agricultural waste
113. Madejon et al., 2001b Grape marc 76%; vinasse 58 C on day 4; 2 4550% Final 7.6 Final 12
20% weeks Final 25%
thermophilic
phase
114. Smars et al., 2001 Half frozen source Max 63 C at day 6 Initial 65% Average Initial 22
separated household 57.570%
waste
115. Smith et al., 2001 Garden refuse, fruit and 70% during the Final 6.79 Final 12.5
vegetable waste process
116. Smith et al., 2001 Garden refuse without Maintained at 70% 5.80 14.9
turning
117. Smith et al., 2001 Garden refuse with Maintained at 70% 5.97 14.5
turning
118. Smith et al., 2001 Market + garden refuse Maintained at 70% 6.36 13.1
without turning
119. Smith et al., 2001 Market + garden refuse Maintained at 70% 6.48 12.5
with turning
Downloaded by [McGill University Library] at 04:56 14 November 2012

120. Solano et al., 2001 Sheep manure Thermophilic Maintained above


waste 13 weeks 40%
121. Sommer, 2001 Cattle deep litter 6070% in first 50% Final 810
week
122. Tiquia et al., 2001 Poultry litter 70 C on day 4 65% Initial 6.9
Final 7.2
123. Tiquia et al., 2001 Poultry litter + yard 55 C in first 2
trimmings (1:2) days; peak by
day 4; bottom
of pile 60 C;
middle of pile
70 C; top of pile
71 C
124. Weppen, 2001 Organic waste 68 C in second
week
125. Wong et al.,2001 Soya bean leaves and 73 C after 2 week 6070% Initial 6.1 Initial 30
residue (1:1) Final 8.1 Final 18
126. Almendros et al., 2000 Plant materials (leaves 60%
and stem)
127. Eklind and Kirchman, Artificial organic Initial 28
2000a household waste, Final 6
potatoes, carrot, meat
and bone meal (65%,
15%, 13%, 7%)
128. Eklind and Kirchman, Organic household waste 58 C on day 7 50% Initial 4.5 Initial 28
2000b Final 4.9 Final 20
129. Garcia-Gil et al., 2000 MSW 7.9 8.7
130. Ghosh et al., 2000 Organic fraction of solid 59.5 C 45% Initial 26
hospital waste Final 14
131. Ghosh et al., 2000 Hospital solid waste Initial 24.0 C Maintained at 45% Initial 5.5 Initial 55.0
(HSW) under ambient Final 25.7 C Final 5.7 Final 54.4
conditions
(Continued on next page)

345
Downloaded by [McGill University Library] at 04:56 14 November 2012

TABLE 2. An Overview of the Types of Substrates Used for Composting in Recent Years (20002005), and Some Key Physicochemical Parameters

346
Associated With the Respective Processes (Continued )

Number Authors Substrate Temperature profile Moisture pH C/N

132. Ghosh et al., 2000 HSW + one year lod Initial 24.0 C Maintained at 45% Initial 5.6 Initial 52.7
composted waste under Final 26.3 C Final 5.9 Final 52.3
ambient conditions
133. Ghosh et al., 2000 HSW + cow manure Initial 40.0 C Maintained at 45% Initial 6.9 Initial 27.0
under lab conditions Final 59.5 C Final 8.6 Final 10.2
134. Ghosh et al., 2000 HSW + autoclaved cow Initial 40.0 C Maintained at 45% Initial 6.9 Initial 27.0
manure under lab Final 58.2 C Final 8.5 Final 12.2
conditions
135. Ghosh et al., 2000 HSW + horse manure Initial 40.0 C Maintained at 45% Initial 6.2 Initial 28.5
under lab conditions Final 58.2 C Final 8.6 Final 12.7
136. Ghosh et al., 2000 HSW + cow manure Initial 22.0 C Maintained at 45% Initial 6.2 Initial 28.5
under ambient Final 53.1 C Final 8.7 Final 13.2
conditions
137. Ghosh et al., 2000 HSW + food waste under Initial 40.0 C Maintained at 45% Initial 4.0 Initial 28.5
lab conditions Final 67.8 C Final 8.2 Final 12.4
138. Ghosh et al., 2000 HSW + food waste under Initial 24.0 C Maintained at 45% Initial 4.0 Initial 28.5
ambient conditions Final 62.3 C Final 10.6 Final 13.5
139. Haider and Karlson, Dehydrated manure 48% Vaied from 3 to 75% Initial 6.5
2000 water 35% 34 C Final 6.3
140. He et al., 2000 Food waste 4547 C between
days 5 and 8
and stabilized
after day 10
141. Larney et al., 2000 Cattle manure >45 Cin 132 days Initial 71%,
in winter, 65 C final 68%, after
in 98 days in curing 40%
summer on day
6
142. Lazzari et al., 2000 Filter pressed sewage 65 C on day 14 Initial 7.1
sludge; ligneous and Final 8.5
green waste
Downloaded by [McGill University Library] at 04:56 14 November 2012

143. Nakasaki et al., 2000 Sludge from fish 50 C for 8 days Initial 55% Initial 7.3
processing WTP Final 7.0
144. Prince et al., 2000 Coir pith 6.70 26.10
145. Schloss and Walker, Big red puppy meal (dog 65 C 5165% Initial 18
2000 meal)
146. Tiquia and Tam, 2000a Chicken litter Max 65 C on first Initial 65% and no Initial 8.3
week further water Final 7.4
added during the
process
147. Tiquia and Tam, 2000b Pig sludge and spent pig 65 C at pile centre Initial 8.6 0.32 Initial 17.58 1.15
litter at day 7 Final 6.9 Final 14.0
148. Wong and Fang, 2000 Dewatered anaerobically >50 C after 23 6.9
digested sewage sludge days
+ saaw dust (bulking
agent) + lime at 0, 0.63,
1.0, 1.63%
149. Wong and Fang, 2000 Dewatered anaerobically 55 C, >50 C after 6070% Initial 7.3
stabilized primary 2 3 days Final 6.9
sewage sludge
150. Zorpas et al., 2000a Dewatered anaerobically Thermophilic 4050% Final 7.25 0.05
stabilized primary phase 15 days;
sewage sludge 6065 C in
reactor centre
151. Zorpas et al., 2000b Dewatered anaerobically Thermophilic 4050% Initial 7.0 7.25 Initial 13.0
stabilized primary phase 15 days Final 11.73
sewage sludge and
organic fraction of MSW

347
348 S. Gajalakshmi and S. A. Abbasi

moisture content (>75%) inhibits microbial activities due to early dehydration


or anaerobiosis (Bertoldi et al., 1983; Tiquia et al., 1996, 2002a). Excess mois-
ture will fill many of the pores between particles with water, thereby limiting
oxygen transport. This in turn would create anaerobic conditions and brings
about putrefaction, resulting in disagreeable odor and undesirable products.
On the other hand, if the composting substrate is supplied with insufficient
water, the growth and proliferation of microorganisms as well as the rate
of decomposition of the organic material would be slowed down or even
stopped. It is important, therefore, to ensure adequate moisture in each layer
of the compost heap. The moisture content of several composting systems
studied in recent years (20002005) is given in Table 2.
Downloaded by [McGill University Library] at 04:56 14 November 2012

5.1.4. OXYGEN AND TEMPERATURE


The decomposition process enhances the interplay between two of the key
environmental parameters, oxygen and temperature. The temperature within
a composting mass determines the rate at which many of the biological pro-
cesses take place and plays a selective role in the development and the
succession of the microbiological communities (Mustin, 1987). Temperature
and oxygen fluctuate in response to microbial activity, which consumes oxy-
gen and generates heat. Both are linked by a common mechanism of control:
aeration. Aeration is one of the components of the controlling process, as it
ensures the growth of adequate aerobic microbe populations and the devel-
opment of stabilizing temperature (Barrington et al., 2003). Aeration supplies
the depleted oxygen to the composting mixture and carries away excess heat
from the system. Inadequate oxygen may lead to the growth of anaerobic
microorganisms, which can produce odorous compounds.
Usually, in an aerobic system, the temperature rises to 5060 C in just
a few days and can even go up to 70 C in some cases. If done correctly,
a compost pile will heat to high temperatures within 24 to 48 hours. If it
doesnt, the pile is too wet or too dry or there is not enough green mate-
rial (or nitrogen) present. The high temperature rise in the compost heap
destroys weed seeds, pathogenic microorganisms, maggots, and worms, and
prevents fly breeding. This happening and the generation of antibiotics dur-
ing composting drastically reduce pathogens in the final compost. A temper-
ature in the range of 55 to 65 C ensures destruction of pathogenic organisms
(Finstein et al., 1987) (Table 6). A temperature of 65 C for at least 30 min-
utes is considered a critical threshold for plant pathogens (Bollen, 1969;
Lopez-Real and Foster, 1985; Bollen et al., 1989). Human pathogens are also
inactivated at high temperatures (Burge, 1983). The temperature and the
time interval required to destroy most common types of pathogenic microor-
ganisms and parasites are given in Table 3. The heat resistance of human
pathogens increases markedly under dry conditions (Cooper and Golueke,
1977). Therefore, wet conditions must prevail in the compost pile. Standards
Solid Waste Management by Composting 349

TABLE 3. Temperature and the Time Interval Required to Destroy Most Common Types of
Pathogenic Microorganisms and Parasites

Pathogen Temperature and time

Salmonella typhosa Further growth is stopped above 46 C; dies within 2030 minutes at
temperature of 5560 C
Salmonella sp. Dies within 60 and 20 minutes at a temperature of 55 and 60 C,
respectively
Shigella sp. Dies within 60 and 20 minutes at a temperature of 65 C
Note. From Burge (1983).

have been proposed by the U.S. Environmental Protection Agency (EPA) to


Downloaded by [McGill University Library] at 04:56 14 November 2012

judge pathogen destruction. The U.S. EPA recommends a five days period
at 55 C (U.S. EPA, 1999), whereas Bertoldi et al. (1988) suggest that a 3-day
period at 65 C and moist conditions is required (Table 6). According to Vin-
neras et al. (2003), to achieve inactivation of pathogens, the reactor has to be
sufficiently insulated so that the materials at the walls also attain high tem-
perature. When low-temperature areas are present, turning of the material
will increase the inactivation of pathogens. Therefore, it is important to have
adequate insulation of the contents, even in areas with high temperature
climates. With sufficient insulation, it is possible to reach temperatures over
60 C.
It is assumed that heat is primarily lost via natural convection when
the sides of the pile are insulated. Cold and oxygen-rich air enters the pile
from the bottom of the reactor where the compost pile is heated by aerobic
degradation. The air is warmed up and takes up water from the compost bed
(Kubler, 1982; Beukema et al., 1983). At the start of the process, oxygen is
supplied by air, which is in the pores of the compost bed. This amount of
oxygen is very small but high enough to ensure a small temperature raise
which triggers natural convection.
The maximum temperature of the composting process reaches 60
70 C, the temperature level where many microorganisms become less ac-
tive (Epstein, 1997). At the top of the pile, the temperature is slightly lower
due to conductive heat loss from the top to the surroundings. Over time,
the temperature gradually drops off as the degradation rate of organic mat-
ter becomes less. This course in composting will result in adequate stabi-
lization of organic matter, drying of the compost, and killing of pathogens
and weeds. According to Rynk et al. (1992) and Fernandes et al. (1994),
low temperature typically indicates low aerobic activity in the composting
pile.
Temperature alone is not a foolproof indicator of aerobic activity, as it
is a result of heat production and heat removal. Lack of aerobic activity can
only be confirmed by measuring the oxygen content within the compost bed
(Veeken et al., 2004). To attain temperatures high enough for heat activation
350 S. Gajalakshmi and S. A. Abbasi

throughout in the compost, the vessel has to be insulated to retain the heat
produced.
High temperature combined with high exchange rates of the air will
increase the ammonia losses (Vinneras et al., 2003). In a composting pile,
however, the rate of degradation is a result of metabolic activity of a mixed
microbial population that may originally include microorganisms with differ-
ent temperature optima. These microorganisms adapt to the environmental
temperature during composting and have a collective temperature optimum
(Topt ) at which respiration from the microbial community is highest. The
temperature profile of the composting processes involving different types of
substrates is illustrated in Table 2.
Not only is microbial metabolism highly temperature dependent, but
Downloaded by [McGill University Library] at 04:56 14 November 2012

it also dramatically influences the population dynamics (e.g., composition


and density) of microbes are dramatically influenced by temperature.
Temperature increase within composting materials is a function of ini-
tial temperature, metabolic heat evolution, and heat conservation (Miller,
1992).
Indeed, temperatures of composting material below 20 C have been
demonstrated to significantly slow or even stop the composting process.
Temperature in excess of 60 C has also been shown to reduce the activity
of the microbial community, and above this temperature, microbial activity
declines as the thermophilic optimum of microorganisms is surpassed (Miller,
1992). If the temperatures reach 82 C, the microbial community is severely
impeded (Finstein et al., 1986; Fermor et al., 1989). MacGregor et al. (1981)
found that optimum composting temperatures, based on maximizing decom-
position, were in the range of 5260 C.
5.1.5. AERATION
With composting being an aerobic biodegradation process, oxygen is its
lifeline. Adequate supply of air to the compost pile is, therefore, of utmost
importance. As briefly mentioned in section 5.1, aeration may occur naturally
by diffusion of ambient air into the porous matrix of a substrate pile but it
happens very frugally, making the composting time exceedingly long (a year
or so). To enable better aeration at appropriate stages of composting, either
the mechanical means of aeration, turning, is employed or air is supplied
through pipes with or without the aid of pumps.
Robertsson (2001) studied the effects of turning off the aeration in small
laboratory-scale compost reactors at three different stages: day 3 at the end
of the mesophilic phase, day 4 at the maximum temperature, and day 5 as
the temperature had begun to decline. He found that there is a tendency for
a greater amount of acetic acid and a smaller amount of other fatty acids to
be generated when aeration is turned off at the two later stages.
In natural aeration, there is diffusion of warm air upward from the sur-
face, which draws ambient air into the base of the compost pile. The path
Solid Waste Management by Composting 351

of ambient air entering the sides of the composting pile and leaving through
the top is a result of high internal temperatures, and is referred to as a chim-
ney effect (Hellman et al., 1997). The disadvantage of natural aeration is
the inability to readily change the physical conditions within the pile mass
(Stentiford, 1996).
In forced aeration, air is actively pushed into the pile in a manner and
to an extent that can be programmed to precision. It is designed to max-
imize the rate of microbial decomposition and optimize the utilization of
process heat. Care should be taken to ensure that forced aeration does not
lead to desiccation, and the resultant impediment in subsequent satges of
composting (Boulter et al., 2000).
Downloaded by [McGill University Library] at 04:56 14 November 2012

5.1.6. PH

The pH is another parameter that greatly affects the composting process.


The range of pH values suitable for bacterial development is 6.07.5, while
fungi prefers an environment in the range of pH 5.58.0 (Kapetanois et al.,
1993; Zorpas et al., 2003). Nakasaki et al. (1993) tested the pH dependency
of organisms active in the composting process and found the pH range of
78 to be optimum, whereas according to Bharadwaj (1995) the optimum
pH for most microorganisms is between 6.5 and 7.5. An initial phase charac-
terized by a low pH is often observed during composting of organic wastes
and perhaps especially of easily degradable energy-rich materials like house-
hold waste (Poincelot, 1974; Haug, 1993; Nakaskai et al., 1993; Kirchmann
and Widen, 1994). This is due to the formation of carbon dioxide and volatile
fatty acids. With the subsequent evolution of CO2 and utilization of VFAs, the
pH begins to rise and may reach even values exceeding 8.0 (Sharma et al.,
1997). Organic acids are produced during decomposition of the organic mat-
ter, but their existence is only transitory. Problems may arise if the material
obtained undergoes putrefaction, as appreciable amounts of troublesome or-
ganic acids are produced during anaerobic decomposition and may produce
malodour. However, a rise in pH beyond 7.5 could make the environment
alkaline, which may cause loss of nitrogen as ammonia. The growth of active
microorganisms is inhibited by temperature above about 40 C if short-chain
fatty acids and low pH are present (Smars et al., 2002).
Microbial tolerance to thermophilic temperature is reduced by the com-
bination of low pH and increasing concentrations of fatty acids. In studies
conducted by Beck-Friis et al. (2001), it was observed that there was an ini-
tial decreasing pH and increasing concentrations of short-chain fatty acids,
particularly lactic acid, during the start of the process. These acids may be
produced as a result of degradation of the more easily degraded nutrients
such as sugars, starch, and fats. In addition, it has been proven that the
degradation process can be enhanced by pH control or microbial inoculants
(Nakasaki et al., 1993, 1996; Smars et al., 2002).
352 S. Gajalakshmi and S. A. Abbasi

Although variation of pH is common across a compost pile, there is a


general trend for pH to become acidic at the onset of decomposition and
then rise, reflecting the loss of organic acids through volatilization and min-
eral decomposition, and the release of ammonia through mineralization of
organic nitrogen (Finstein and Morris, 1975). The acidity of the initial material
is a result of organic acids that are formed during fast degradation, which
often occurs before the material is even in a composting pile. Composting
increases pH from organic acid depletion, release of alkali ions, and the accu-
mulation of ammonia (Hellmann et al., 1997). Acceptable pH ranges should
be within levels tolerable to microorganisms. For fungi, the upper limit of
pH is a function of the precipitation of essential nutrients from the medium,
rather than pH itself. Hoitink and Kuter (1986) indicated the optimum pH
Downloaded by [McGill University Library] at 04:56 14 November 2012

range for decomposition is between 6.5 and 8.5. The pH affects the po-
tential for beneficial bacteria to colonize composts; below pH 5.0, bacterial
biocontrol agents are inhibited.
To curtail excessive ammonia loss, Hoitink and Kuter (1986) suggest
that pH should be below 7.4 in aerated composting systems. The pH is an
indicator of aeration levels within a composting pile. Well-aerated compost
piles generally have a high pH, whereas piles with anaerobic conditions have
decreased pH values (Boulter et al., 2000).
According to studies conducted by Nakasaki et al. (1993, 1996), the
degradation rate of organic matter in the pH-controlled experiment was faster
than in the experiment without pH control.
The pH profiles of several composting systems studied in recent years
is given in Table 2. A list of the important abiotic parameters associated
with the success of composting process, and the range in which they should
preferably remain, is presented in Table 4. The relationship between the

TABLE 4. Key Parameters That Influence the Composting Process and Their Optimum Values

Parameter Optimum value for composting

C/N ratio of the feed 25 to 35


Particle size 10 mm for agitated systems and forced aeration, 50 mm for long
heaps and natural aeration
Moisture content 50 to 60% (higher values when bulking agents are used)
Air flow 0.6 to 1.8 m3 air/day/kg volatile solids during thermophilic stage, or
maintain oxygen level at 10% or higher
Temperature 55 to 60 C held for 3 days
Agitation No agitation to periodic turning in simple systems and short bursts
of vigorous agitation in mechanized systems
pH control Normally not necessary
Heap size Any length, 1.5 m and 2.5 m wide for heaps using natural aeration.
With forced aeration, heap size depends on need to avoid
overheating
Activators Use of efficient cellulolytic fungi and biofertilizers
Note. Culled from Gaur (2000).
Solid Waste Management by Composting 353

TABLE 5. The Relation Between the Degree of Maturity, Temperature, and O2 Consumption
in a Compost System

Maximum O2 consumption O2 consumption


Degree of temperature (mg/g OS) (mg/g OS)
maturity ( C) (according to Jourdan) (according to Becker) Material status

I >60 >40 >80 Raw material


II 6050.1 408.1 8050 Fresh compost
III 5040.1 2816.1 5030 Fresh compost
IV 4030.1 166.1 3020 Matured compost
V 30 6 20 Matured compost
Note. Culled from Korner et al. (2003).
Downloaded by [McGill University Library] at 04:56 14 November 2012

degree of maturity of the compost, the temperature, and the O2 consumption


is presented Table 5.
The decrease in pH during the initial period of composting is expected
because of the acids formed during the metabolism of readily available car-
bohydrates. After the initial stage, the pH is expected to rise, with evolution
of free ammonia, and to stabilize or drop slightly again to near neutral as a
result of humus formation with its pH buffering capacity at the termination
of composting activity (Poincelot, 1974; Fogarty and Tuovinen, 1991).
In composting, carbohydrates are also broken down into humic and
fulvic acids (Spaggiari and Spigoni, 1986). However, the fulvic acid is subse-
quently degraded. This, together with ammonification of inorganic nitrogen,
accounts for neutral pH, which is generally attained at the end of the process
(MacGregor et al., 1981).

5.1.7. ELECTRICAL CONDUCTIVITY (EC)


Generally, it is found that EC increases during composting as volatile solids
(VS) are degraded and the amount of water-soluble salts increases on a to-
tal solids (TS) basis. At lower pH values, negatively charged surface sites
of organic matter are occupied by protons, which thus lowers CEC (Bolt
and Bruggenwert, 1988). A decrease in CEC results in a lower adsorption of
cations to organic matter and thus an increase in EC.

5.2. Biotic Factors


Composting involves a myriad of microorganisms. The composition and
magnitude of these microorganisms are important components of the com-
posting process. The microbes decompose the organic matter, and trans-
form the nitrogen component through oxidation, nitrification, and denitrifi-
cation (Golueke, 1992; Tiquia and Tam, 2000a). During the process, there
may be depletion of nutrients if the microbes incorporate minerals from the
waste into biomass (Golueke, 1992). However, composting may also involve
354 S. Gajalakshmi and S. A. Abbasi

sequential growth and degradation of subpopulations. Hence there may be


no significant change in overall levels of microorganisms or inorganic nutrient
requirement (Tiquia et al., 2002a).
Microbes cannot directly metabolize the insoluble particles of organic
matter. All biochemical reactions during composting are catalyzed by en-
zymes (Ayuso et al., 1996; Garcia et al., 1992a; Godden et al., 1983; Vuori-
nen, 1999, 2000). The microbes produce hydrolytic extracellular enzymes to
depolymerize the larger compounds (i.e., plant polymers, cellulose, hemicel-
lulose, and lignin) to smaller fragments that are water-soluble (Hankin et al.,
1976).
Downloaded by [McGill University Library] at 04:56 14 November 2012

5.2.1. BACTERIA
Bacteria play by far the most dominant role during the most active stages
of composting process because of their ability to grow rapidly on soluble
proteins and other readily available substrates (Strom, 1985a; Golueke, 1992;
Epstein, 1997). They may also attack more complex materials, or may exploit
substances released from the less degradable materials due to extracellular
enzyme activities of other organisms (Epstein, 1997).
Among bacteria that occur commonly in aerobically decomposing sub-
strate are species of Bacillus, Cellulomonas, Pseudomonas, Klebsiella, and
Azomonas (Nakasaki et al., 1985; Strom, 1985a, 1985b). Clostridium oc-
cur substantially in anaerobic conditions. Typical bacteria of the ther-
mophilic phase are species of Bacillus, e.g., B. subtilis, B. licheniformis,
and B.circulans. Strom (1985b) reports that as much as 87% of the randomly
selected colonies during the thermophilic phase of composting belong to the
genus Bacillus. Many thermophilic species of Thermus have been isolated
from compost at temperatures as high as 65 C and even 82 C (Beffa et al.,
1996a; Tuomela et al., 2000). Nitrosomonas spp. and Nitrobacter spp. are the
ammonium-oxidizing and nitrite-oxidizing bacteria respectively, present in
the compost heap (Focht and Verstraete, 1977).
Establishment of a large population of denitrifying bacteria suggests that
some anaerobic microhabitat exists within the compost piles. These micro-
habitats could have been developed within the piles partially due to the initial
high water content (65%) of the piles and partially because of the rich con-
tents of organic matter and nitrogen present in the substrate, which promote
microbial activity to the extent of causing depletion in O2 content in isolated
pockets within the piles. Moreover, some species of denitrifying bacteria may
be facultative and grow aerobically (Firestone, 1982). Some microbial genera
capable of denitrification are Bacillus, Flavobacterium, and Pseudomonas
(Tiquia et al., 2002a).
Nakasaki et al. (1985) found Bacillus spp. and Azotobacter spp. to be the
common mesophilic bacteria responsible for CO2 evolution during early stage
of composting when the temperature is <40 C. Mesophilic microorganisms
Solid Waste Management by Composting 355

are partially killed or poorly active during the thermogenic stage (4060 C)
(Beffa et al., 1996b). The diversity decreased as temperature increased, with a
shift from Pseudomonas, Achromobacter, Flavobacterium, Micrococcus, and
Bacillus to one dominated by Bacillus. Bacteria related to B. schlegelii, Hy-
drogenobacter spp., and particularly to the genus Thermus (T. thermophilus,
T. aquaticus) appear to be the main active microbes in hot compost (6580 C)
(Beffa et al., 1996b). Bacterial survival in high-temperature composting ma-
terial is possible through formation of microcolonies. Mesophiles are likely
to contribute little to compost degradation at these temperatures (Nakasaki
et al., 1985).
Microbial fermentation of carbohydrates generally results in an increase
in acidity (Garg and Neelakantan, 1982). Clostridium species commonly fer-
Downloaded by [McGill University Library] at 04:56 14 November 2012

ment glucose to yield butyl and ethyl alcohols and certain acids. Lactobacillus
lactis yields almost entirely lactic acid, while Lactobacillus bevis yields lactic
and acetic acids, ethyl alcohol, and carbon dioxide (Frobisher et al., 1974).

5.2.2. FUNGI
The role of fungi starts when simple, easily degradable substances such as
sugar, starch, and protein are acted upon by bacteria and the substrate is
predominated by cellulose and lignin, which normally occurs toward the
later stages of composting (curing process) (Bertoldi et al., 1983; Golueke,
1992; Tiquia et al., 2002b)
Most fungi are eliminated by high temperatures (Epstein, 1997), but they
commonly recover when temperatures are moderate (Tiquia et al., 2001), and
the remaining substrates are predominantly cellulose or lignin (Bertoldi et al.,
1983).
Being efficient consumers of carbon, fungi build up much higher
biomass than other microorganisms. The most commonly observed species
of celluloytic fungi in composting materials are Aspergillus, Penicillium, Rhi-
zopus, Fusarium, Chaetomonium, Trichoderma, Alternaria, and Cladiospo-
rium. Some of the species of Paecilomyces and Sporotrichum have also been
named as efficient degraders of lignocellulosic wastes (Kapoor et al., 1978;
Mandhulika et al., 1993).
White-rot fungi are known as the most efficient lignolytic microorgan-
isms. Phanerochaete chrysosporium is probably the best suited microorgan-
ism with this activity and it is often used as a reference. Among other well-
known white-rot fungi, Coriolus versicolor show even higher efficiency and a
wider range of lignolytic activities together with an important celluloytic activ-
ity. Phanerochaete flavidoalba causes preferential loss of lignin rather than of
cellulose and it is more efficient than P. chrysosporium on paper mill effluents.
The plant constituent that offers maximum resistance to biodegradation
is lignin. Yet, in spite of its substantial microbial recalcitrance, lignin does get
degraded by some fungi and a few bacteria. The most important among these
356 S. Gajalakshmi and S. A. Abbasi

are white-rot fungi belonging to Basidiomycetes. Species of Polyporus, Pleu-


rotus, Collybia, Poria, Fomes, Trametes, Sporotrichum, Cyathus, and Corio-
lus have also been found to degrade lignin (Muthukumar and Mahadevan,
1983).
Temperature is one of the most important factors affecting fungal growth.
Other important factors are sources of C and N, and the pH. During com-
posting, temperatures above 55 C discourage fungal growth. Fungi are ex-
cluded during the earlier high-temperature stage of the composting process
(Nakasaki et al., 1985; Miller, 1992). A moderately high level of nitrogen is
needed for fungal growth, although some fungi, mainly wood-rotting fungi,
grow at low nitrogen levels. Indeed, a low nutrient nitrogen level is often
a prerequisite for lignin degradation (Eriksson et al., 1990; Dix and Web-
Downloaded by [McGill University Library] at 04:56 14 November 2012

ster, 1995). However, low nutrient nitrogen is a rate-limiting factor for the
degradation of cellulose (Dix and Webster, 1995). Most fungi prefer an acidic
environment but tolerate a wide range of pH, with the exception of the
Basidiomycotina, which do not grow well above pH 7.5.
The majority of the fungi are mesophiles, which grow between 5 C and
37 C, with an optimum temperature of 2535 C (Dix and Webster, 1995).

However, in the compost environment the elevated temperature means


that the small group of thermophilic fungi is an important biodegradation
agent.
Cooney and Emerson (1964) define thermophilic fungi as fungi with a
maximum growth temperature of 50 C or higher and a minimum growth
temperature of 20 C. Thermotolerant species have a maximum growth tem-
perature of about 50 C and a minimum well below 20 C (Cooney and Emer-
son, 1964). But Crisan (1973) defines thermophilic fungi as thermotolerant
when it can survive in a temperature optimum of 40 C or higher. These
fungi are known to have celluloytic or lignolytic activity. Thermophilic fungi
that have been found growing in lignocellulose substrate or compost are
Taloromyces emersonii, T. thermophilus, Thermoascus auranticus, and Ther-
momyces lanuginosus.
The most effective lignin degraders are Basidiamycotina, but according
to Cooney and Emerson (1964) all Basidiamycotina are mesophilic. However,
a few Basidiamycotina grow well at elevated temperatures. Phanerochaete
chyrsoporium (Sporotrichum pulverulentum) is a white-rot fungus with an
optimum temperature of 3640 C and maximum temperature of 4649 C.
Ganoderma colosum is another white-rot fungus that is still capable of grow-
ing at 45 C and has an optimum temperature of 40 C (Adaskaveg et al., 1995).
In the genus Coprinus there are some species that have an optimum tem-
perature of above 40 C (Crisan, 1973). A thermophilic Ascomycotina, Ther-
moascus aurantiacus, has a high lignolytic capacity and has been isolated
from compost (Klopotek, 1962).
Waksman et al. (1939a, 1939b) studied composting on a large scale
at temperatures of 28, 50, 65, and 75 C. At 28 C, the population was
Solid Waste Management by Composting 357

heterogeneous, with bacteria being dominant throughout the whole period,


and fungi appearing later. Fungi, together with bacteria and actinomycetes,
formed the microbial population in the compost at 50 C.
The highest degradation of lignin occurred at 50 C, while degradation
was to some extent lower at 28 C and 65 C. At 75 C biodegradation of
lignin occurred, but 12% of the lignin was solubilized as a result of the
high temperature and alkaline reaction of the compost (Tuomela et al.,
2000).

5.2.3. ACTINOMYCETES
Like fungi, actinomycetes also utilize complex organic material. They tend to
Downloaded by [McGill University Library] at 04:56 14 November 2012

grow in numbers in the later stages of composting, and have been shown to
attack polymers such as hemicellulose, lignin, and cellulose (Bertoldi et al.,
1983; Epstein, 1997).
The actinomycetes that occur most frequently are Micromonospora,
Streptomyces, Nocardia, and Thermoactinomyces. Actinomycetes generally
show their activity at later stages of decomposition.
Actinomycetes are higher forms of bacteria, which form multicellular
filaments; thus they resemble fungi. They are primarily strict aerobic sapro-
phytes, and are common in many environments. Their ubiquity is a result
of their ability to utilize a wide range of carbon sources and to sporulate
prolifically. Actinomycetes colonize more slowly than bacteria and fungi.
Colonization is minimal in areas that are poorly aerated. They appear during
the thermophilic phase as well as the cooling and maturation phase of com-
posting, and can occasionally become so numerous that they are visible as
a white film on the surface of the compost. The genera of the thermophilic
actinomycetes isolated from compost include Nocardia, Streptomyces, Ther-
moactinomyces, and Micromonospora (Waksman et al., 1939b; Strom, 1985a).
Actinomycetes are able to degrade some cellulose and solubilize lignin, and
they tolerate higher temperatures and pH than fungi. Thus, actinomycetes
are important agents of lignocellulose degradation during peak heating, al-
though their ability to degrade cellulose and lignin is not as high as that of
fungi (Crawford, 1983; Godden et al., 1992; Tuomela et al., 2000). The acti-
nomycetes are thus well placed to exploit the compost environment as the
piles cool in the immediate post peak heat phase. Epstein (1997) reported
that during the cooling stage of composting, actinomycetes actively degrade
hemicellulose in the compost (Tiquia et al., 2002b).
With an optimum growth between 2530 C and pH of 59, these mi-
croorganisms are the most significant group of microbes in the degradation
of relatively complex, recalcitrant polymers. As actinomycetes develop more
slowly than most bacteria or fungi, they are ineffective competitors when
nutrient levels are high, but become more competitive as nutrient levels de-
crease (Nakasaki et al., 1985). Actinomycetes thermophilus, Streptomyces, and
358 S. Gajalakshmi and S. A. Abbasi

Micromonospora spp. are common in compost. Although optimum growth


temperatures fall in the mesophilic range, obligate thermophiles such as Ther-
moactinomycetes and Saccharomonospora spp. have been isolated. Certain
species of actinomycetes are more tolerant of high temperatures, becoming
increasingly active as temperatures approach and surpass 60 C (Nakasaki
et al., 1985).
Cellulose is not an obligate carbon source for fungi and actinomycetes,
which are the microorganisms mainly responsible for cellulose degradation,
and the addition of readily metabolizable substances has been shown to
accelerate the decomposition of cellulose. It is thought that, by initially uti-
lizing the more available C sources, the population of cellulose degraders can
develop to a large size. Once the more available C source becomes limiting,
Downloaded by [McGill University Library] at 04:56 14 November 2012

the microorganisms adapt to the cellulose, with the overall effect being an in-
crease in cellulose hydrolysis (Alexander, 1961). Since cellulose degradation
is largely attributed to fungi and actinomycetes, which are characterized by
the formation of hyphae, it is possible that frequent turning of the compost
resulted in the breaking of the hyphae and, subsequently, in reduced activity
of the cellulose degraders.

6. ENZYMES INVOLVED IN THE COMPOSTING PROCESS

It is well known that all biochemical reactions during composting are cat-
alyzed by enzymes (Ayuso et al., 1996; Garcia et al., 1992a; Godden et al.,
1983; Vuorinen, 1999, 2000).
Enzymes in the composts can be classified as intracellular (enzymes
inside viable cells) or extracellular (enzymes outside the cells). Intracellular
enzymes are enzymes that catalyze biochemical reactions occurring within
the cells. Conversely, extracellular enzymes are enzymes purposefully re-
leased exterior to cells, generally to catalyze the degradation of polymeric
substances (i.e., plant polymers, cellulose, hemicellulose, and lignin) too
large to cross the cellular membrane. The intracellular and extracellular en-
zymes cannot be distinguished in compost suspension. However, after a brief
incubation period, extracellular groups of enzymes can be identified to which
a large portion of enzymes in soils and composts belong (Vuorinen, 1999,
2000).
In composting, the soluble organic matter in the starting material is ini-
tially assimilated by the microorganisms (Rynk et al., 1992). Microbes in the
compost pile cannot directly metabolize the insoluble particles of organic
matter. Microorganisms produce hydrolytic enzymes to depolymerize the
larger compounds (i.e., lignin, cellulose, and hemicellulose) to smaller frag-
ments that are water-soluble (Hankin et al., 1976; Priest, 1984; Tate, 1995;
Tiquia et al., 2002c). The water-soluble components dissolve in the water
Solid Waste Management by Composting 359

transport across the cytoplasmic membrane and are finally assimilated by the
microorganisms (Tabatabai, 1994).
The mineralization of organic N during composting, which involves the
release of N from nonpeptide CN bonds in amino acids, and urea is medi-
ated by enzymes such as amidohydrolases and dehydrogenases (Tabatabai,
1994). Alkaline and acid phosphatases are important enzymes in organic P
minieralization and plant nutrition. Phosphatases are the enzymes that cat-
alyze the hydrolysis of organic P esters to orthophosphate; arylsulfatase re-
moves the sulfate group from organic compounds; leucine-amino peptidase
is an enzyme that catalyzes the hydrolysis of proteins into individual amino
acids; and -glucosidase is involved in the hydrolysis of cellobiose. These
are key reactions in soils and composts (Garcia et al., 1992a; Tabatabai, 1994;
Downloaded by [McGill University Library] at 04:56 14 November 2012

Vuorinen, 1999).
Godden et al. (1983) found that cellulase, invertase, and alkaline phos-
phatase activities increased during early stages of composting of cattle ma-
nure, and remained constant during the thermophilic and curing period. On
the contrary, Ayuso et al. (1996), Garcia et al. (1992a), and Diaz-Burgos et al.
(1993) observed that the activities of phosphatase, urease, and protease de-
crease during sewage sludge composting.
Various hydrolytic enzymes are believed to control the rate at which
various substrates are degraded. Enzymes are the main mediators of various
degradative processes (McKinley et al., 1985; Tiquia et al., 1996). Three im-
portant enzymes, cellulase, xylanase, and protease, are responsible for hy-
drolysis of cellulose, hemicellulose, and proteins, respectively.
Dehydrogenase and catalase are intracellular enzymes that are involved
in microbial oxidoreductase metabolism. Catalase is an oxidoreductase as-
sociated with aerobic microbial activity (Rodriguez-Kabana and Truelove,
1982). Urease and protease hydrolyse nitrogen compounds to ammonium,
using urea and low-molecular-weight protein substrates, respectively.

7. PARAMETERS USED TO EVALUATE COMPOST MATURITY

Several indicator variables have been proposed for monitoring the com-
posting process and evaluating the stability of the compost. The terms sta-
bility and maturity are both commonly used to define the degree of de-
composition of organic matter during the composting process even if they
are conceptually different. Compost stability refers to the level of activ-
ity of the microbial biomass and can be determined by O2 uptake rate,
CO2 production rate, or the heat released as a result of microbial activ-
ity (Iannotti et al., 1994; Conti et al., 1997). Compost maturity refers to
the degree of decomposition of phytotoxic organic substances produced
during the active composting stage (Wu et al., 2000). However, stability
and maturity usually go hand in hand, since phytotoxic compounds are
360 S. Gajalakshmi and S. A. Abbasi

produced by the microorganisms in unstable composts (Zucconi et al., 1985).


The application of immature compost can result in inhibited seed germina-
tion, root destruction, and suppressed plant growth caused by, inter alia,
an excess of NH+ 4 , the presence of phenolic substances or organic acids
such as acetic acid, propionic acid, and n-butyric acid, N immobilization
due to high C/N ratio, and a decrease in oxygen concentration and re-
dox potential due to rapid decomposition of the compost with a possible
increase in the mobility of some trace metals (Smith and Hughes, 2004).
The indicator variables include physical, chemical, and biological parame-
ters of the organic material (Bertoldi et al., 1983; Forster et al., 1993; Grebus
et al., 1994; Tiquia and Tam, 2000a). These parameters include drop in
temperature, degree of self-heating capacity, oxygen consumption, biochem-
Downloaded by [McGill University Library] at 04:56 14 November 2012

ical parameters of microbial activities, analysis of biodegradable constituents,


phytotoxicity assays, cation-exchange capacity (CEC), organic matter nutrient
content, C/N ratio, and humus content and quality (Harada and Inoko, 1980;
Michel et al., 1996; Rynk et al., 1992; Tiquia et al., 2002a; Tiquia and Tam,
2002; Forster et al., 1993; Mari et al., 2003). Of these parameters, O2 con-
sumption or respiratory activity and heat production, which are indicative of
the amount of degradable organic matter still present, are inversely related
to stabilization (Zucconi and Bertoldi, 1987).
Compost stability refers to the situation when the rate of energy release
due to microbial degradation of the organic matter equals the rate of energy
loss to the environment. Under these conditions, the temperature of the com-
post remains constant and equals that of the ambient. Compost stability is
strongly related to the rate of microbial activity in the compost. Compost ma-
turity is used to describe product quality. It is based on utilization-oriented
definitions as follows: (i) greenhouse utilization: organic matter composted
to the degree of decomposition that has no adverse effects on container
grown plants; (ii) field application: organic matter composted to the degree
of decomposition that has no adverse effects on growth of various crops.
The widely used parameter for assessing compost stability is respira-
tion, as measured either by O2 uptake or CO2 evolution. Respiration must
be measured under well-controlled conditions of moisture and temperature,
and is affected by temporary anaerobic conditions (Iannotti et al., 1994). In
addition, Wu et al. (2000) showed that low CO2 evolution is not always an
indicator of a nonphytotoxic compost.

7.1. Respirometric Studies


Oxygen is consumed and carbon dioxide is evolved as a consequence of
microbial metabolism. Both the processes decline at late composting stages.
Respirometric studies, which determine the O2 consumption or CO2 produc-
tion caused by mineralization of the composts organic matter, have been
carried out in pure composts and in compost mixed with soil in a proportion
Solid Waste Management by Composting 361

compatible with agricultural use (Morel et al., 1979; Iannotti et al., 1993).
An insufficiently mature compost has a strong demand for O2 and high CO2
production rates due to intense development of microorganisms as a con-
sequence of the abundance of easily biodegradable compounds in the raw
material. For this reason, O2 consumption or CO2 production are indicative
of compost stability and maturity (Hue and Liu, 1995).
In almost all experiments that are respirometric units for studying com-
posting processes, the measurement of respiration activity is carried out at
standard temperature. In a composting pile, however, the rate of degradation
is a result of metabolic activity from a mixed microbial population that may
originally include microorganisms with different temperature optima. These
microorganisms adapt to the environmental temperatures during composting
Downloaded by [McGill University Library] at 04:56 14 November 2012

and have a collective temperature optimum (Topt ) at which respiration from


the microbial community is highest.
The potential of the microbial community in a compost pile to metab-
olize available carbon source at different environmental temperatures was
studied by Mari et al. (2003), who reported that the respiration measure-
ments at 35 C could be used as a general indicator of the potential metabolic
activity during composting. However, measurements at higher temperature
(48.5 C) were better indicators of the maximum respiration potential (respi-
ration at Topt ).
Because respiration rate is independent of the initial feedstock char-
acteristics, it should provide an indication of microbial activity and reflect
the degree of stabilization in open or closed composting systems. Microbial
respiration is affected by moisture, temperature, and oxygen and nitrogen
availability. Measurements of microbial respiration can be problematic as
they are subjected to environmental constraints aside from changes in car-
bon substrates characteristic of the decomposition process. CO2 evolution
can also be used for studying microbial activity because photosynthesis is
virtually negligible in the composting process.

7.2. Microbial Ratio Indices


Microbial ratio indices may be used as criteria for evaluating the level of
compost activity and maturity. Because degradation of complex molecules
by microbial consortia is more efficient than by microbes in isolation, changes
in microbial activities during composting have been measured. Although enu-
meration of microbes is a commonly suggested approach to estimate com-
munity characteristics and dynamics, most methods are not satisfactory. Serial
dilutions and spread plating on media are known to underestimate the total
number of microbes, partially because of unculturable microorganisms. In
addition, single colonies on plates may arise from several bacterial aggre-
gates. The isolation and description of various compost microorganisms may
362 S. Gajalakshmi and S. A. Abbasi

not be useful, as it is unlikely that any particular organism could be the chief
agent of decomposition throughout the variable conditions of composting
(Boulter et al., 2000).

7.3. ATP Content


ATP content has been considered as a microbial biomass indicator in soil
(Jenkinson and Ladd, 1981). It has also been used as an indicator of microbial
activity in a composting process and of the degree of composting.
ATP is considered to be suitable for the estimation of microbial activ-
ity because of its linear relationship with microbial cell mass (Oades and
Downloaded by [McGill University Library] at 04:56 14 November 2012

Jenkinson, 1979). Since the intracellular ATP content in microorganism varies


with the species and their specific growth rates in addition to cell numbers
in a compost, the ATP content in compost is regarded as the qualitative in-
dicator to reflect total microbial activity, including microbial cell mass and
their growth rates in a compost (Horiuchi et al., 2003).
A decrease in the ATP content in the composting process is due to less
microbial activity, since ATP is related to intact microbial cells (Jenkinson,
1988). ATP stabilization indicates maturity of the substrate when most bac-
teria have been replaced by specialized groups of microorganisms such as
actinomycetes and fungi, which are characterized by slower metabolism and
hence slower ATP production (Garcia et al., 1992a).

7.4. Arginine Ammonification


Ammonification of arginine, an essential amino acid, is a common pro-
cess in microorganisms (Alef and Kleiner, 1986). According to Forster et al.
(1993), arginine ammonification provides valuable information on the N sta-
tus in compost. Negative values indicate that considerable amounts of easily
degradable organic compounds with wide C/N ratios are still present, which
could lead to microbial immobilization of soil N after application of the com-
post. Arginine mineralized to ammonium during decomposition can readily
be extracted from soil or composts and measured. Lin and Brookes (1999)
reported that most heterotrophic bacteria are able to ammonify arginine.
However, according to Tiquia et al. (2002a), arginine ammonification may
not be used as a reliable indicator of microbial activities in yard trimmings
composting.

7.5. Dehydrogenase Activity


Dehydrogenase activity refers to a group of mostly endocellular enzymes that
catalyze the oxidation of soil organic matter (Forster et al., 1993). Dehydroge-
nase activity decreases with composting time of different organic feedstocks
Solid Waste Management by Composting 363

and remains stable after 2 or 3 months of the process (Benito et al., 2003;
Pelaez et al., 2004).

7.6. Water-Soluble Carbon/Dissolved Organic Carbon


The dissolved organic carbon (DOC) or water-soluble carbon (WSC) has
been proposed by a number of researchers as a parameter that consistently
decreases during the composting process. It is one of the most readily active
biological parameters in the compost, since organic carbon in the water ex-
tracts of immature MSW compost consists of sugars, hemicellulose, phenolic
substances, amino acids, peptides, and other easily biodegradable substances
(Hsu and Lo, 1999).
Downloaded by [McGill University Library] at 04:56 14 November 2012

The water to compost extraction ratio varies, as does the definition of


dissolved carbon. DOC is shown to be highly correlated to respiration (Chica
et al., 2003).
Garcia et al. (1992b) suggested that in a mature municipal compost the
amount of CO2 -C evolved should be less than 500 mg CO2 -C per 100 g total
organic C in the compost. More CO2 evolution indicates that compost has
not yet stabilized and needs further decomposition.

7.7. Nitrification
The ratio of NH4 -N to NO3 -N in the water extract has been suggested as an
index of maturity. However, the final value of NO3 depends on the source
material, and no particular level of NO3 or its ratio to NH4 can be relied
upon as an indicator of compost biomaturity (Mathur et al., 1993; Bernal
et al., 1998a). It should also be noted that the increase in NO3 N is gradual
over a lengthy period of time, and thus the determination of the point at
which the increase begins is difficult.
Compost maturity can also be defined in terms of nitrification. When the
NH+ 4 concentration decreases and NO3 -N concentration increases, it suggests
that intensive biological decomposition has been slowed down and that the
compost is mature enough (Finstein and Miller, 1985; Benito et al., 2003).
The maximum ratio of NH4 -/NO3 -N for a mature compost as suggested by
Bernal et al. (1998a) is 0.16.
Another indicator of activity is N2 O, a by-product of nitrification and
denitrification. N2 O is produced when temperatures is low, and after the
thermophilic stage during declining temperature (Hellmann et al., 1997).

7.8. Soluble Organic Carbon/Organic Nitrogen


The water-soluble organic C to organic nitrogen ratio (sol org C/org N) has
been suggested as a maturity index by several research groups (Garcia et al.,
364 S. Gajalakshmi and S. A. Abbasi

1991; Hue and Liu, 1995; Bernal et al., 1998b). A threshold value of 56 for
mature composts has been reported by Chanyasak et al. (1983) and Jimenez
and Garcia (1989). Using this ratio as a maturity index is only possible when
all organic N species can be accounted for, that is, when there are no N
losses during the analysis.

7.9. Carbon/Nitrogen Ratio


The C/N ratio represents very good index of maturity level for the organic
substance, as it significantly affects the microbiological growth. The activities
of the heterotrophic microorganisms involved in the process are dependent
Downloaded by [McGill University Library] at 04:56 14 November 2012

upon the nitrogen and carbon content. These microorganisms use carbon as
the energy source, whereas the nitrogen is used for synthesis of proteins.
During the oxidation reactions that involve the release of carbon dioxide,
the major portion of the carbon (approximately two-thirds) is used by the
microorganisms as the energy source, while the remaining portion serves to
form protoplasm of the cells, along with nitrogen, phosphorus, potassium,
and other microelements. Nitrogen, in the form of ammonium ions, is gener-
ally required as a major nutrient. Ammonia is produced from deamination of
proteins, hydrolysis of urea, and probably also from purines and pyrimidines.
Excessive carbon presence slows the microbiological activities, whereas
excessive nitrogen, allowing rapid decomposition, causes big nitrogen loss
through volatilization. The C/N ratio narrows as the composting progresses
because of the conversion of organic carbon to carbon dioxide. Higher C/N
ratios require more time for completion of the maturation phase. It can be
said that the C/N ratio of a compost should not be too high, as an application
of such composts can result in immobilization of available nitrogen, causing
an N deficiency in plants (Kostov et al., 1991; Bannick and Joergensen, 1993).
Conversely, C/N values of composts must not be too low, as N mobilization
and subsequent N toxification and efflux to groundwater may occur (Jimenez
and Alvarez, 1993; Kapetanois et al., 1993; Brink, 1995).
What should be the C/N ratio of ideal compost? Opinions vary; different
workers have deemed different C/N ratios as ideal, as presented in Table 6.
Extended composting periods will reduce long-term N availability, pri-
marily since as composting proceeds a higher proportion of N will be con-
verted to available, organic forms, either within the microbial biomass or
incorporated into developing humic acid substances (Keeling et al., 1994).
In the former case, release of N is dependent upon microbial death or preda-
tion by grazing protozoa and nematodes. In the latter, humic substances are
relatively recalcitrant, having a high half-life within the soil, thus requiring
longer time to release nitrogen (Keeling et al., 1995).
According to Jimenez and Garcia (1992) a C/N ratio lower than 12 for
municipal waste compost indicates a good degree of maturity. Furthermore,
Solid Waste Management by Composting 365

TABLE 6. Values of Different Parameters Considered Suitable for Composting Systems


Parameter and values Reference
CEC
60 cmol/kg (minimum value) Harada and Inoko, 1980
67 cmol/kg (minimum value) Jimenez and Garcia, 1992
CEC/Corg ratio
1.7 mmol g1 (lowest limit) Roig et al., 1988
1.9 mmol g1 (lowest limit) Jimenez and Garcia, 1992
C HA /C FA
1.9 Jimenez and Garcia, 1992
C/N ratio
Below 20 Poincelot, 1974; Golueke, 1981
Less than 18 De Baere et al., 1986
Lower than 12 Jimenez and Garcia, 1992
1525 Mathur et al., 1993; Canet and Pomares,
Downloaded by [McGill University Library] at 04:56 14 November 2012

1995
CW (water-soluble organic carbon)
<1.7% Bernal et al., 1998a
CW /Norg ratio
0.55 Bernal et al., 1998a
<0.7 Hue and Liu, 1995
NH+4 -N, mg kg
1

0.04% (400 mgkg1 ) Zucconi and De Bertoldi, 1987


NH+4 / NO3 -N ratio
0.16% Bernal et al., 1998a
Temperature
>55 C: maximize sanitation Stentiford, 1996
>4555 C: maximize the biodegradation rate
>3540 C: maximize microbial diversity
Below 20 C: significantly slows down or even stops Mosher and Anderson, 1977
the composting process
In excess of 60 C: reduces the activity of microbial Miller, 1992
community and at still higher temperatures, the
microbial activity declines further
If temperature reaches 82 C: microbial community Nell and Wiechers, 1978; Finstein et al.,
is severely impeded 1986; Fermor et al., 1989
5260 C: optimum temperature based on MacGregor et al., 1981; Bach et al.,
maximizing decomposition 1984; McKinley et al., 1985
70 C: inhibitory to many microbes Golueke, 1972
Moisture
50% minimal requirement for rapid increase in Liang et al., 2003
microbial activity, while a range of 6070%
provides maximum activities
6070% Hoitink and Kuter, 1986
4550% limiting; less than 35% (on a weight basis) Golueke, 1972; Frost et al., 1992
reduces microbial activity
1215%: minimum level at which bacterial activity Golueke, 1972
occurs
To induce plant disease suppression, moisture content Hoitink et al., 1997
must be high enough (at least 4050% w/w) to
allow bacterial and fungal colonization of the
substrate after peak heating
pH
67.5: bacteria generally require this Golueke, 1992
5.58.0: fungi can generally tolerate this Golueke, 1992
5.09.0: actinomycetes Goodfellow and Williams, 1983
6.58.5: optimum for decomposition Hoitink and Kuter, 1986
<5.0: bacterial biocontrol agents are inhibited Hoitink et al., 1997
<7.4 in aerated composting systems to curtail Hoitink and Kuter, 1986
excessive ammonia loss
Germination index (GI)
>50% Zucconi et al., 1981a, 1981b
366 S. Gajalakshmi and S. A. Abbasi

in well-humified soils the C/N ratio is close to 10 and the addition to soils
of materials with a C/N ratio below 15 may not alter the microbiological
equilibrium of the soil. Therefore, a C/N ratio of lower than 12 may indicate
the suitability for addition of soil. According to Sharma et al. (1997), a good
quality compost has a C/N ratio of the order of 1520.
Rapid and entire humification of a substrate essentially depends on initial
C/N ratio, which should be between 25 and 35, with a pH of 67.5 (Golueke,
1992). A C/N ratio below 20 is indicative of proper compost maturity, with a
ratio of 15 or less being preferred (Poincelot, 1974). A C/N ratio of 7, as well
as quotient (final ratio/initial ratio) being low, indicates that it should not
induce immobilization of mineral nitrogen in soil (Casale et al., 1995). It has
been stated that the C/N ratio of mature compost should ideally be about 10,
Downloaded by [McGill University Library] at 04:56 14 November 2012

but this is hardly ever achievable, due to the presence of recalcitrant organic
compounds, or materials that resist decomposition due to their physical or
chemical properties (Mathur, 1991). Some authors reported that a C/N ratio
below 20 is indicative of an acceptable maturity (Poincelot, 1974; Golueke,
1981), with a ratio of 15 or even less being preferable (Jimenez and Garcia,
1989).
The composting of materials with low C/N ratio result in more N losses
than in high C/N ratio wastes (Reddy et al., 1979; Sanchez-Monedero et al.,
2001).
Most of the nitrogen found in a composting mixture is organic, princi-
pally as part of the structure of proteins and simple peptides. A small part
of this organic nitrogen is mineralized to ammonia by ammonification re-
actions resulting from the microbial activity developed. The ammonia thus
formed undergoes different processes depending on the condition of the
mixture being composted. For example, it may be dissolved (as ammonium)
and then immobilized by the microorganisms of the mixture, which use it as
nitrogen source and transform it again into organic nitrogen. Alternatively,
it may be volatilized and be given off, as happens when the mixture is at a
high temperature with a pH of above 7.5 (Witter and Lopez-Real, 1987).
A lack of oxygen leads the organisms to use the nitrate as an oxygen
source, which results in denitrification and stops nitrification (Tisdale et al.,
1985). During the nitrification process, the nitrifying bacteria lower the pH
of the medium due to the liberation of hydrogen ions, a process that can be
summarized in the following equations:

Nitrosomonas bacteria: 2NH+ +


4 + 3O2 2NO2 + 4H + 2H2 O

Nitrobacter bacteria: 2NO


2 + O2 2NO3

Of these nitrogen transformations, those undergone by the inorganic forms,


ammonium and nitrate, are the most interesting from an agricultural point of
view since in these forms they can be assimilated directly by the root system
of plants (Sanchez-Monedero et al., 2001).
Solid Waste Management by Composting 367

Morisaki et al. (1989), Parades et al. (1996) and Witter and Lopez-Real
(1988) emphasized the importance of adding lignocellulosic materials as
bulking agents to reduce nitrogen losses.
The ratio between inorganic forms of nitrogen has been used as a crite-
rion for assessing compost maturity. At the end of the process the concentra-
tion of nitrates should be higher than that of ammonium, indicating that the
process has been prepared under adequate conditions of aeration (Finstein
and Miller, 1985). A high concentration of NH4 -N in compost indicates insta-
bility, and according to Zucconi and Bertoldi (1987), it should not exceed
0.04% in mature compost (Table 6). Bernal et al. (1998a) established a limit
of 0.16% for the relation between ammonium nitrogen and nitrate nitrogen
as an index of maturity in composts of very different origin.
Downloaded by [McGill University Library] at 04:56 14 November 2012

The compost quality for agricultural use depends on its inorganic nitro-
gen content. Compost is primarily an organic N source and releases N slowly,
i.e., on the order 0.3% to 6% of the compost N per year for municipal solid
wastes (MSW) compost (McConnell et al., 1993; Sikora et al., 1980; Tester
et al., 1977). Because of the slow mineralization rate of the compost and the
need for immediate availability of fertilizer N to plants, several authors have
suggested the addition of N and P to compost.
Compost having a high C/N ratio may immobilize soil N in microbial
biomass. Indeed, compost with a C/N ratio above 30 results in N immobi-
lization (Tester et al., 1977; Sikora et al., 1980; McConnell et al., 1993). A C/N
ratio between 12 and 14 corresponds to that observed for organic amendment
with high N liberation (Bauduin et al., 1986).
Nyns (1986) observed losses in ammonia in compost heaps that had
low carbon/nitrogen ratios. However, nitrogen is generally stabilized in the
process by being transformed into microbial protein.

7.10. Cation Exchange Capacity (CEC)


Cation exchange capacity (CEC) has been heavily studied as a potential in-
dicator of compost maturity. The CEC in an organic material increases as a
function of humification due to the formation of carboxyl and phenolic func-
tional groups (Lax et al., 1986). CEC values of 60 and 67 cmol kg1 described
by Harada and Inoko (1980) and Jimenez and Garcia (1992), respectively, are
the minimum values for a city refuse compost to be considered sufficiently
mature.

7.11. Humification
Since stabilization or maturation also implies the formation of some
humic-like substances, the degree of organic matter humification is gener-
ally accepted as a criterion of maturity. Studies in this respect refer to the
humification ratio, humification index, percent of humic acid, humic acid
368 S. Gajalakshmi and S. A. Abbasi

to fulvic acid ratio, and the chemical, physicochemical, and spectroscopic


characterization of humus like substances.
A CHA /CFA ratio (HAhumic acid; FAfulvic acid) of 1.9 has been pro-
posed as a maturity index of city refuse and sewage sludge compost (Jimenez
and Garcia, 1992) (Table 6).
The humic acid concentration increases during the process; however,
this increase represents a very small percentage in comparison with the pre-
existing humified fraction. Therefore, this index alone cannot be used in
order to estimate the degree of organic matter evolution during the compost-
ing process.
The phenolic fraction, unlike the carbohydrate fraction, is composed
of more complex molecules. This phenolic fraction is generated during the
Downloaded by [McGill University Library] at 04:56 14 November 2012

partial degradation of lignin and is more resistant to degradation due to its


aromaticity. However, the water-soluble phenol fraction, which has a more
simple structure and a smaller molecular size, is sensitive to the transforma-
tions occurring during composting and has been used as indicator of the com-
posting process (Saviozzi et al., 1987). Haninen and Lilja (1994) and Saviozzi
et al. (1987), using a mixture of wastewater sludge from a paper-processing
factory and chopped wheatstraw and a mixture of slaughter wastes, respec-
tively, observed that this fraction behaved otherwise, decreasing rapidly dur-
ing the first months of composting and then increasing to levels higher than
the initial values. Their observations suggested that phenol polymerization
and its degradation predominated over the lignin degradation since the phe-
nol fraction did not decrease during the composting process. Both the sol-
uble carbohydrate and phenol fractions of the soluble organic matter play
a very important role in organic matter degradation. Their concentrations
are in continuous equilibrium as a result of their formation from more com-
plex polymers and degradation as a consequence of microbial activity. No
mathematical correlations were observed between the soluble carbohydrate
fraction and the principal humification indices, probably because such car-
bohydrates are the principal carbon sources of the microbial flora that are
mainly responsible for the degradation that takes place. The participation of
soluble carbohydrates in the humification process is therefore masked.

7.12. Germination Index (GI)


The degree of maturity can also be revealed by biological methods involving
seed germination and by measuring the root length of the ensuing plants
(Zucconi et al., 1981a). Immature composts may contain phytotoxic sub-
stances such as phenolic acids and volatile fatty acids (Kirchmann and Widen,
1994), thereby inhibiting the germination and growth of angiosperms. It had
been suggested that a GI (a factor of relative seed germination and rela-
tive root elongation) of 80% indicated the disappearance of phytotoxins in
Solid Waste Management by Composting 369

composts (Zucconi et al., 1981b; Tiquia et al., 1996). GI values greater than
50%, according to Zucconi et al. (1981b), indicate a phytotoxin-free compost.

8. CORRELATIONS BETWEEN THE MATURITY AND STABILITY


OF COMPOST

Lignin concentration highly correlates with GI and inversely with Cw /Norg


ratio, which can be considered as a maturity index.
The Cw /Norg (water-soluble carbon to organic N ratio) is the parameter
that is correlated with most factors and at the highest level of probability, es-
Downloaded by [McGill University Library] at 04:56 14 November 2012

pecially lignin as already mentioned; C F A (fulvic acid), GI, and CEC correlate
to a lesser degree of probability with the CHA /CFA (humic acid to fulvic acid
ratio) and C/N ratios. This suggests that this factor is the most suitable for
describing composts degree of maturity.
According to Haider (1992), the lignin/Norg ratio controls C mineraliza-
tion in plant residues and a low ratio implies a rapid mineralization rate. A
high C/N ratio implies a high proportion of mineralizable C, while the nature
of the organic matter (CW, CFA , lignin, etc.) would influence the rate of the
process (Bernal et al., 1998b).
The application of immature compost can result in inhibited seed germi-
nation, root destruction, and suppressed plant growth (Zucconi et al., 1981a,
1981b; Garcia et al., 1992a; Marambe and Ando, 1990). This is caused by,
inter alia, an excess of NH+4 , the presence of phenolic substances or organic
acids such as acetic acid, propionic acid, and n-butyric acid, N immobilization
due to a high C/N ratio, and a decrease in oxygen concentration and redox
potential due to rapid decomposition of the compost with a possible increase
in the mobility of some trace metals (Chanyasak et al., 1983; Garcia et al.,
1992b; He et al., 1992; Marambe and Ando, 1990; Keeling et al., 1994). Since
cellulose is the most abundant structural polysaccharide of plant cell walls
(Alef and Nannipieri, 1995) and probably the most abundant organic com-
pound in nature (Alexander, 1961), its microbial degradation is a key process
in the decomposition of plant debris and plays, therefore, an important role
in composting. The degree of degradation of cellulose (Smith and Hughes,
2001) can thus be chosen as another parameter to monitor the composting
process. Because the formation of carboxyl groups during the composting
process contributes greatly to the cation exchange capacity (CEC) of compost
(Lax et al., 1986), CEC was included as a further maturity parameter. Other
parameters monitored were: (1) fluorescein diacetate hydrolysis, which is a
simple, sensitive and rapid technique to estimate microbial activity (Schnurer
and Rosswall, 1982); (2) arginine ammonification rate, which is suggested
as a simple method to determine microbial activity potentials in soil (Alef
and Kleiner, 1986) and to characterise the N status in compost (Forster et al.,
370 S. Gajalakshmi and S. A. Abbasi

1993); (3) absorbance of compost extracts, as an indicator of the amount


of soluble organic matter; and (4) cress germination and growth on com-
post, a commonly used simple method to test whether a compost contains
phytotoxic compounds.
According to Forster et al. (1993), arginine ammonification provides valu-
able information on the N status in compost. Negative values indicate that
considerable amounts of easily degradable organic compounds with wide
C/N ratios are still present, which could lead to microbial immobilization
of soil N after application of the compost. The results of the arginine am-
monification test carried out on samples taken during the composting process
showed that the N concentration in the composts had not been limiting, since
the NH+ 4 production rates were positive at all times. With time, however, a
Downloaded by [McGill University Library] at 04:56 14 November 2012

small but continuous decline in arginine ammonification was observed in


all composts. This decline could have been caused by decreasing micro-
bial activity in the compost. Alef and Kleiner (1986, 1987) demonstrated that
arginine ammonification is an accurate indicator of microbial activity in soil.
However, Alef and Kleiner (1987) cautioned that the application of the test
might be limited to samples in which the energy source is abundant and no
ammonification might be observed.

9. BENEFITS OF COMPOSTING
9.1. Compost for Soil Application
Compost intended to serve as soil amendment is applied in order to improve
soil fertility. In many cases prolonged intensive agronomical or horticultural
cultivation causes gradual depletion of soil organic matter. Reduced soil or-
ganic matter is frequently associated with lower soil biological activity and
with deteriorating soils physical properties. The result is overall reduced soil
fertility. Repeated application of various types of organic matter, and espe-
cially of compost, can reverse this negative process. Quantitative parameters
that are affected by organic amendments include soil biomass, soil respira-
tion, various enzymatic activities, nitrification rate, large aggregate stability,
water infiltration rate and hydraulic conductivity, and water-holding capacity.
If application of organic amendments is done properly and for relatively long
period (510 years, depending upon climatic conditions, soil management
and soil type), original soil fertility can be restored (Raviv, 2005).

9.2. Application as Soil Conditioner and Fertilizer


The application of compost has been shown to positively affect, inter alia, the
structure, porosity, water-holding capacity, compression strength, nutrient
content, and organic matter content of the soil (Mays et al., 1973; Smith,
Solid Waste Management by Composting 371

1996). One of the earliest reports on container production of ornamentals


with compost, describing research between 1948 and 1954, was published
in Belgium in 1956 (DeGroot, 1956).
Compost as organic fertilizer has no recognized negative impacts on
plant growth and the environment, and can possibly promote the former
(Vaughan and Malcolm, 1985). The addition of compost to soil can affect soil
fertility by modifying the physical, chemical, and biological properties of the
soil (Dick and McCoy, 1993). Of these, the physical changes include modi-
fications of soils bulk density, structure, strength, and water retention. The
chemical changes include the storage of organic plant nutrients, the soils ion
exchange capacity, and its buffering ability (Tiquia and Tam, 2002). Hence
the application of compost enhances the organic matter and nutrient content
Downloaded by [McGill University Library] at 04:56 14 November 2012

of the soil and improves soil structure (Diez and Krauss, 1997; Saviozzi and
Riffaldi, 1998; Cox et al., 2001).
Through its close to neutral pH value, compost helps arable land to have
quick decomposition of organic substances. In addition, through adsorbent
action, it minimizes the migration of contaminants into the environment.
Biologically, compost improves the plants nitrogen-absorbing capacity
by promoting the mineralization level of the soil, as well as increasing the
absorption capacity of natural components from the soil. This is mainly be-
cause of the presence of diverse species of numerous bacteria in the compost
(Sharma et al., 1997).
Fitzpatrick (2001) has discussed that the challenges to successful com-
post use in container media are nutrient content, soluble salt levels, com-
paction, and phytotoxicity.
Since compost products are made from organic materials, it is not surpris-
ing that they contain substantial levels of certain essential nutrients. However,
the concentration of these nutrients, particularly nitrogen (N) and potassium
(K), may sometimes not be sufficient high to provide complete nutritional
support for the crops. Hence the crops nutritional requirement must be met
with supplemental fertilization. Excessively high soluble salt levels in com-
post materials can be managed by leaching, where leaching would not pose a
threat to surface or groundwater resources, or by blending the compost with
substrates that have lower soluble salt levels. Indeed, many compost-based
potting media recommendations contain only 20 to 30% compost, as a means
of reducing damage that can be caused by high salt levels or other phytotoxic
substances that may be present in certain compost products (Raymond et al.,
1998).
Porosity is one of the more important physical parameters in container
media (Poole et al., 1981), because of the need for effective gas exchange in
the root zone. Some compost products have been reported to have satisfac-
tory pore space at the beginning of the plant production period but undergo
compaction during the production period (Fitzpatrick and Verkade, 1991).
Compost materials used as the complete, or stand-alone, rooting substrate
372 S. Gajalakshmi and S. A. Abbasi

are more likely to settle or compact during the production period, thereby
reducing the porosity of the medium. This phenomenon is more likely to
occur in fresh and immature compost products. This problem can be treated
either by allowing the compost product to age further, or by blending the
compost with materials that are not likely to undergo compaction during
production.
Compost products may contain phytotoxic materials that can come from
a variety of sources. The organic material from which the compost is made
may contain residues of substances that can be toxic to crops grown in the
end product compost. A more serious cause of phytotoxicity in compost
products can come from the composting process itself.
The technically correct minimum amount of time that an organic sub-
Downloaded by [McGill University Library] at 04:56 14 November 2012

stance must undergo composting in order to have a stable end product com-
post is variable. It depends on several factors, including the size of the com-
post pile, aeration and moisture status, the C/N ratio of the material, heat
levels and range during composting, etc. (Fitzpatrick, 2001), as discussed
earlier in section 8.
Given next is the gist of the studies conducted by the authors and others
on the impact of compost on plants.

9.2.1. FRUIT AND VEGETABLE PLANTS


In a study conducted by Kostov et al. (1996), application of compost (derived
from vine branch, rice husks, and flax) significantly increased the yield of
tomatoes and the quality of fruits compared to the soil treated with mineral
fertilizers and manure. Shiralipour et al. (1996) observed that application of
compost increased the height and dry weight of broccoli shoots and dry
weight of lettuce shoots in several different soil textures tested by them.
Experiments were conducted by Smith et al. (2001) to monitor the effects
of compost application and inorganic fertilizer addition on the growth of two
vegetable cropsSwiss chard (Beta vulgaris L. var. flavescens) and common
bean (P. vulgaris L. var. nanus). The four composts used in this study differed
in raw material (garden refuse with or without market waste) and the fre-
quency of turning during the composting processes (no turning, turned once
in 6 months). The composts were applied at three rates (0, 25, 50% m/m)
to a sandy soil with and without low applications of inorganic fertilizer. In
case of both plants, the soilcompost mixtures out-performed the soil alone
irrespective of the amount of fertilizer added. Swiss chard produced the high-
est total leaf fresh mass on composts made from market and garden refuse.
The yield was further significantly improved when the composts had been
turned or when the proportion of compost was increased from 25 to 50%,
but the addition of fertilizer had no significant effect on the total yield. The
positive effects of the composts that had been turned during the composting
process might have been the result of a slower, but continuous, release of
Solid Waste Management by Composting 373

nutrients throughout the experiment due to the higher cation-exchange ca-


pacities detected in the turned composts. In contrast to the results for Swiss
chard, the application of compost produced from garden and market refuse
that had undergone turning had a negative effect on the yield of common
bean, especially when the proportion of compost was increased to 50% and
when fertilizer was added. Common bean performed the best when grown
on soil mixed with compost made exclusively from garden refuse.
With the exception of the treatment that contained compost made
from turned market and garden refuse, adding fertilizer and increasing the
proportion of compost led to higher numbers of pods and seeds and in-
creased total seed dry mass. The contrasting responses of the two crops to
the applied composts are believed to be caused by differences in the crops
Downloaded by [McGill University Library] at 04:56 14 November 2012

nutrient requirements and salinity tolerances.


Another study was carried out with Swiss chard (Beta vulgaris L. var.
cicla) by Parades et al. (2005) on the effects of two composts(i) from cot-
ton gin waste and sewage sludge with addition of olive mill wastewater and
(ii) from cotton gin waste and sewage sludge without the addition of olive
mill wastewater. Five treatments were applied: mineral fertilizer and two
doses (30 and 60 tonnes ha1 ) of both composts. The olive mill wastewater
addition produced compost with lower organic matter and nitrate concen-
trations, higher electrical conductivity, and a stabilized and humified organic
matter similar to that of the compost produced without olive mill wastew-
ater. The olive mill wastewater compost application to soil did not injure
plants, producing a similar plant yield to both compost without olive mill
wastewater and inorganic fertilizer. Also, the accumulation of potentially toxic
heavy metals in plants cultivated with organic or mineral fertilisers did not
reveal significant differences. The olive mill wastewater compost application
to soil also improved the chemical and physicochemical properties of the
soil.
The authors conducted studies on some vegetables to dispel the appre-
hension of the farmers that compost/vermicompost obtained from a perni-
cious weed like water hyacinth may have deleterious effect on other plants
(Gajalakshmi and Abbasi, 2002). Qualitative studies were done at the kitchen
gardens with ladys finger (Hibiscus esculentus), brinjal (Solanum melon-
gena), cluster bean (Cyamopsis tetragonoloba), chili (Capsicum annum),
and tomato (Lycopersicon esculentum). The observations revealed total ab-
sence of any harmful effect of compost/vermicompost. Rather, the farm-
ers view was that the vegetables grew better than normal on the treated
plants.
Wei and Liu (2005) used a randomized complete block with a total num-
ber of 21 plots, with each plot measuring 2 2 m2 . Seven treatments were
done: with the incorporation of sewage sludge compost at the rates of 25,
50,100, 150, 200, and400 ton ha1 ; one plot without treatment served as con-
trol. These treatments were repeatedly conducted each year from 1998 to
374 S. Gajalakshmi and S. A. Abbasi

2000 to evaluate the effect of different rates of sewage sludge compost appli-
cation on rapeseed germination and development. The results of this study
show that the sewage sludge compost application (1) had no obvious effect
on the rapeseed germination and was beneficial for the plumelet develop-
ment at lower application rates (150 ton ha1 ); (2) generated positive yield
responses for barley and Chinese cabbage; and (3) increased the concentra-
tions of Cu and Zn in 020 cm soil and produced little effect on those of Cu
and Zn in barley grains and cabbage leaves. Therefore, the sewage sludge
compost should be applied to cropland soil at limited application rates (<150
ton ha1 ).
In a study conducted by Lee et al. (2004), the effect of food waste
(FW) composted with soil microorganisms was compared with commercial
Downloaded by [McGill University Library] at 04:56 14 November 2012

compost (CC) and mineral fertilizer (MF) on bacterial and fungal populations,
soil enzyme activities, and growth of lettuce in a greenhouse. Populations of
fungi and bacteria, soil biomass, and soil enzyme activities in the rhizosphere
of FW treatments significantly increased compared to control (CON), CC,
and MF treatments at 2, 4, and 6 weeks. The fresh weight of lettuce in FW
treatments was about 23 times higher than that in CC treatment at 4 and 6
weeks. The pH, EC, total nitrogen, organic matter, and sodium concentration
in FW treatments were generally higher than those in CON, CC, and MF
treatments.
It is known that organic matter introduced to soil stimulates soil microbial
populations and soil biological activity (Brady and Weil, 1999). Alvarez et al.
(1995) reported that addition of compost to soil increased the incidence of
bacteria in the tomato rhizosphere.
The capacity of composts made from three different combinations of
organic wastes (horse manure and bedding, mink farm wastes, municipal
solid waste [MSW], and sewage sludge) along with clarifier solids from a
chemo-thermomechanical pulp mill to enhance the growth of tomato (Ly-
copersicon esculentum L.) seedlings grown in nutrient-poor organic potting
soil was studied by Levy and Taylor (2003). Germination and seedling emer-
gence of tomatoes, cress (Lepidium sativum L.), or radish (Raphanus sativus
L.) were tested to assess phytotoxicity of the four amendments. Mink farm
compost and horse manure compost stimulated root and shoot growth of
tomato seedlings, but MSW compost and pulp mill solids were strongly in-
hibitory. MSW compost and unamended potting soil also inhibited seedling
emergence, and pulp mill solids produced stunting and deformities in radish
and cress seedlings. Both toxic constituents and nutrient imbalances may be
responsible for the growth-inhibiting effects of these amendments. Applica-
tion of pulp mill solids to agricultural soil without composting may lead to
deleterious effects on vegetable crops.
The effects of compost derived from municipal solid waste (MSW)
on plant stands, growth, and yields of snap beans (Phaseolus vulgaris L.)
was studied by Ozores-Hampton and Bryan (1993). MSW compost was
Solid Waste Management by Composting 375

incorporated into calcareous limestone soil at 0, 90, and 134 t ha1 . Plant
stands and bean yields were not different among treatments. However, fresh
shoot weight increased quadratically as rates of compost increased. In a sec-
ond experiment, fertilizer applications at planting were compared with a
control (no fertilizer). No additional compost was applied in experiment 2.
Plant stands were the highest when no fertilizer and 90 t ha1 of MSW were
used. Total snap bean yields were higher with fertilizer at planting than with
no fertilizer. The yield increased quadratically as rates of compost increased.
Experiments were conducted by Korboulewsky et al. (2002a) to study
the effects of sewage sludge compost on white wall rocket (Diplotaxis
erucoides L.) in comparison with mineral fertilization and control (without
any fertilizer) in a greenhouse experiment. The plants grown on the compost-
Downloaded by [McGill University Library] at 04:56 14 November 2012

amended soil showed a different growth dynamics: a significant delay on


flowering and a bigger root system. Both the compost and the fertilization
treatments increased biomass and seed yield. The experiment showed that
sewage sludge compost provides significant benefits to white wall rocket
growth without causing any disturbances of mineral nutrition.

9.2.2. AGRICULTURAL CROPS


The application of urban waste compost caused no visible phytotoxic effects
on corn plants; rather, it led to a significant increase in grain yield (Businelli
et al., 1996). In a study conducted by Prince et al. (2000), application of
coir pith compost at the rate of 5 g/kg soil produced a significant increase
over the control in the number of roots in the mulberry plant, Morus alba
L. When 10 g compost per 3 kg of soil was applied, the length of the roots
and nitrogen content of leaves were increased, whereas in the case of plants
with 25 g compost per 3 kg of soil, the leaf chlorophyll and the number of
branches and leaves also increased.
The influence of maize-stover compost and nitrogen fertilizer on growth,
shoot yield, and nutrient uptake of amaranth (Amaranthus cruentus L) was
studied by Akanbi and Togun (2002) over a period of 2 years. Twelve treat-
ments derived from a factorial combination of four levels of maize-stover
compost (0, 1.5, 3.0, and 4.5 t/ha) and three levels of N-fertilizer (from urea)
(0, 30, and 60 kg/ha) were tested. The plant growth parameters increased
significantly with increasing levels of applied compost from 0 to 4.5 t/ha soil.
The interactive effects between the various levels of N and compost were
significant for all the growth parameters considered. The highest values for
all the parameters (except dry matter yield) were recorded when 30 kg N
was combined with 3.0 t/ha compost, producing the highest dry matter yield,
which was not significantly different from combined 30 kg N/ha and 3.0 t/ha
compost.
Better crop production and development was observed under higher
N and compost rates. This suggested that increasing crop fertilization to an
376 S. Gajalakshmi and S. A. Abbasi

optimum rate enables the plants to produce their potential number of leaves
and leaf area. These invariably enhance plant photosynthetic activities and
hence more dry matter is produced (Bittenbender et al., 1998).

9.2.3. ORNAMENTAL PLANTS


The authors have conducted studies on the application of water hyacinth
compost on the growth and yield of a flowering plant, Crossandra undu-
laefoila (Gajalakshmi and Abbasi, 2002). Crossandra undulaefolia is an an-
giosperm and a free-branching ornamental perennial herb (Pandey, 1982).
The results indicated that the pots containing soil amended with water
hyacinth compost had plants achieving significantly better height, larger num-
Downloaded by [McGill University Library] at 04:56 14 November 2012

ber of leaves, more favorable shoot: root ratio, greater biomass per unit time,
and larger length of inflorescence. In terms of root length, quicker onset of
flowering, and harvest index, too, the treated plants on an average performed
better than the control but not in a statistically significant manner.
A greenhouse pot study was conducted by Ribeiro et al. (2000) to eval-
uate the use of a municipal solid waste compost (MSWC) as a fertilizer for
potted geranium. MSWC was mixed with a peat-based growing media at rates
of 0%, 10%, 20%, 30%, 40%, and 50% by volume. Plants grew in those forms
for 90 days, with no additional fertilization. Shoot dry weight, number of
leaves per plant, number of flower stems per plant, and number of flowers
per flower stem were significantly affected by the percentage of MSWC, with
greatest growth occurring at 10% and 20% MSWC. However, these plants
showed a low concentration of N and P in the leaves, showing that MSWC
provided only a part of the required N and P. At an MSWC percentage higher
than 20%, yield decreased, a likely result of salt stress. Therefore the MSWC
rates must be adjusted according to the conductivity of the applied compost
and to the salt tolerance characteristic of the plant species to avoid salt stress
and detrimental effects on plant growth.
The concentrations of K, Ca, Mg, Fe, Zn, Mn, Cu, and B in the leaves
of the 20% MSWC plants were within the adequate range proposed by Jones
et al. (1991) and Ribeiro et al. (2000).

9.3. Patterns of Plant Responses to Compost


There are many opinions about the use of compost in the growing substrate
for ornamental plants. Compost can be used as a stand-alone substrate, as
a component of the growing substrate, as an organic top dressing, or as an
incorporated soil amendment (Fitzpatrick, 1998).
Research has been conducted on annual, perennial, and woody orna-
mental plant responses to increasing concentrations of different compost
products in the growing substrate. Five basic patterns become evident when
reviewing the literature about compost use in container production: (1) no
Solid Waste Management by Composting 377

response, (2) plateau, (3) linear increase, (4) bell curve, or (5) decrease.
Some plants show no response to increasing percentages of compost in the
growth substrate. Plants grown in containers containing compost are equal
in size and quality to plants grown in substrates containing no compost. For
example, Klock-Moore (1999) noted that there was no difference in bego-
nia (Begonia semperflorens-cultorum) or impatiens (Impatiens wallerana)
shoot dry weight between the control substrate (60% Canadian sphagnum
peat, 25% vermiculite, 15% perlite by volume) and substrates containing 30%,
60%, or 100% compost made from recycled greenhouse media and yard trim-
mings.
The second response group is characterized as a plateau, in which plant
growth increases as the percentage of compost in the growing substrate
Downloaded by [McGill University Library] at 04:56 14 November 2012

increases up to a point, with no further increases at higher percentages.


Plants grown in substrates containing compost are larger or better quality
than plants grown in control substrates. For example, Bugbee and Frink
(1989) reported that shoot dry weight of Lemondrop marigold (Tagetes
erecta) plants grown in substrates containing 10%, 20%, 30%, 40%, and 50%
composted sewage sludge was greater than for control plants but did not
differ from one another.
Other crops show a linear increase in which plant growth increases
as the percentage of compost in the substrate increases. For example, for
impatiens and begonia shoot dry weight increased as the percentage of
biosolid and yard trimming compost increased from 0% to 100% in the
growing substrate.
Some crops show an increase in growth as the percentage of compost
in the substrate increases up to a point and then growth decreases as the
percentage of compost increases beyond that point (bell curve). As the per-
centage of some compost products in the substrate increases above 50%,
plant growth decreases as a result of high soluble salt concentrations, poor
aeration, heavy metal toxicity, and/or phytotoxicity (Shiralipour et al., 1992).
For example, salvia, dianthus (Dianthus chinensis), and petunia (Petunia x
hybrida) shoot dry weight increased from 0% to 60% and then decreased
(Klock-Moore, 2000; Klock-Moore and Fitzpatrick, 1997).
Finally, plant growth may decrease as the percentage of compost in
the growing substrate increases. For example, Klock-Moore and Fitzpatrick
(1997) observed that impatiens shoot dry weight decreased as the percentage
of composted municipal solid waste increased from 0% to 100%.
It is obvious from various studies that plant response varies with the
amount of compost in the growing substrate as well as the type of compost
being used (Klock-Moore, 2005).
Some compost products have been demonstrated to accelerate growth in
some crop species, thereby decreasing the production period for these crops.
Some compost products also have been demonstrated to have a suppressive
effect on some plant pathogens (Fitzpatrick, 2001). A number of workers
378 S. Gajalakshmi and S. A. Abbasi

have reported the drastic reduction or elimination of phytopathogens that


accompanies composting of wastes (Theodore and Toribio, 1995; Hointink
and Boehm, 1999).
Compost supports a rich diversity of microorganisms and hence extends
a potentially antagonistic community to phytopathogens (Coventry et al.,
2002). Addition of compost to container media can suppress soil-borne and
foliar plant pathogens due to the natural pesticides (pathogen inhibitors)
formed and antagonists developing during the composting process (Hointink
and Fahy, 1986; Saviozzi and Riffaldi, 1998; Hadar and Mandelbaum, 1992;
Zhang et al., 1998). Composting, if done properly, accompanies rise in
temperature of the substrate to 55 C and higher. This may denature weed
seeds and several pathogens (Crawford, 1983; De Ceuster and Hoitink, 1999).
Downloaded by [McGill University Library] at 04:56 14 November 2012

Even though reduction in phytopathogens during composting was mainly


due to high temperatures reached during the process (Ylimaki et al., 1983;
Yuen and Raabe, 1984), additional effects of other microorganisms may occur
during the curing phase (Hointink and Fahy, 1986).
In composting, thermophilic temperature, characteristic of the process,
acts as a strong and comparatively fast pasteurizing agent. Most pathogenic
forms are rapidly eliminated by the direct effect of the heat on vital con-
stituents such as DNA and proteins. Additionally, the antagonistic effects
resulting from the active growth of the compost microbial population further
hasten the elimination of pathogens (Finstein et al., 1980).
Compost must be stable and of consistent quality to be used successfully
for biological control of diseases in horticultural crops (Inbar et al., 1993).
Biological control rarely eliminates a pathogen, but rather reduces its num-
bers or its ability to produce diseases. Such control may be achieved with
little or no reduction in the pathogen population (Baker and Cook, 1974).
Consistent and sustained biological control of diseases can be achieved
in different compost-amended growing media as long as variables such as
consistency of the parent organic material, moisture content, salinity, car-
bon to nitrogen ratio, and process parameters are controlled in the compost
(Hoitink et al., 1991; Zinati, 2005).

ACKNOWLEDGMENT

S.A.A. thanks the Department of Biotechnology, Government of India, New


Delhi, for financial support, and S.G. thanks the Council for Scientific &
Industrial Research, New Delhi, for a research associateship.

REFERENCES

Abbasi, S.A., and Ramasamy, E.V. 1999. Biotechnological methods of pollution control.
Orient Longman (Univesities Press India Ltd.), Hyderabad, India.
Solid Waste Management by Composting 379

Abbasi, S.A., Ramasamy, E.V., Gajalakshmi, S., Khan, F.I., and Abbasi, N. 2000. A
waste management project involving engineers and scientists of a university, a
voluntary (non-governmental) organization, and lay peopleA case study. In
Proceedings of International Conference on Transdisciplinarity, Swiss Federal
Institute of Technology, Zurich, February 13.
Abbot, I., and Parker, C.A. 1981. Interactions between earthworms and their soil
environment. Soil Biology & Biochemistry, 13, 191197.
Abdelhamid, M.T., Horiuchi, T., and Oba, S. 2004. Composting of rice straw with
oilseed rape cake and poultry manure and its effects on faba bean (Vicia faba
L.) growth and soil properties. Bioresource Technology, 93, 183189.
Adaskaveg, J.E., Gilbertson, R.L., and Dunlap, M.R. 1995. Effects of incubation
time and temperature on in vitro selective delignification of silver leaf oak
by Ganoderma colossum. Applied and Environmental Microbiology, 61, 138
Downloaded by [McGill University Library] at 04:56 14 November 2012

144.
Akanbi, W.B., and Togun, A.O., 2002. The influence of maize-stover compost and
nitrogen fertilizer on growth, yield and nutrient uptake of amaranth. Scientia
Horiculturae, 93, 18.
Alburquerque, J.A., Gonzalvez, J., Garcia, D., and Cegarra, J. 2006. Composting of a
solid olive mill by-product (alperulo) and the potential of the resulting com-
post for cultivating pepper under commercial conditions. Waste Management,
26, 620626.
Alef, K., and Kleiner, D. 1986. Arginine ammonification as a simple method to esti-
mate soil microbial activity potentials in soil. Soil Biology and Biochemistry, 18,
233235.
Alef, K., and Kleiner, D. 1987. Applicability of arginine ammonification as indica-
tor of microbial activity in different soils. Biology and Fertility of Soils, 5, 148
151.
Alef, K., and Nannipieri, P. 1995. Cellulase activity. In: Alef, K., and Nannipieri, P.
(eds.), Methods in applied soil microbiology and biochemistry. Academic Press,
London, pp. 345373.
Alexander, M. 1961. Introduction to soil microbiology. Wiley, New York, pp. 163
182.
Almendros, G., Doradoa, J., GonzaAlez-Vila, F.J., Blanco, M.J., and lankes, U. 2000.
13
C NMR assessment of decomposition patterns during composting of forest and
shrub biomass. Soil Biology & Biochemistry, 32, 793804.
Alvarez, M.B., Gagne, S., and Antoun, H. 1995. Effect of compost on rhizosphere
microflora of the tomato and on the incidence of plant growth-promoting rhi-
zobacteria. Applied Environmental Microbiology, 61, 194199.
Arthur, G.D., Jager, A.K., and Van Staden, J. 2001. The release of cytokinins from
leaves of Gingko biloba during composting. Environmental and Experimental
Botany, 45, 5561.
Ayuso, M., Hernandez, T., Garcia, C., and Pascual, J.A. 1996. Biochemical and chem-
ical structural characterization of different organic materials used as manures.
Bioresource Technology, 57, 201207.
Bach, P. D., Shoda, M., and Kubota, H. 1984. Rate of composting of dewatered
sewage sludge in continuously mixed isothermal reactor. Journal of Fermenta-
tion Technology, 62, 285292.
380 S. Gajalakshmi and S. A. Abbasi

Baheri, H., and Meysami, P. 2002. Feasibility of fungi bioaugmentation in composting


a flare pit soil. Journal of Hazardous Materials, B 89, 279286.
Baker, K.F., and Cook, R.J. 1974. Biological control of plant pathogens. W.H. Freeman,
San Francisco.
Bannick, C.G., and Joergensen, R.G. 1993. Changes in N fractions during composting
of wheat straw. Biology and Fertility of Soils, 16, 269 - 274.
Barker, A.V. 1997. Composition and uses of compost. In: Rechcigl, J.E., and McKin-
non, H.C. (eds.), Agricultural uses of by-products and wastes. ACS Symposium
Series 668, Washington, DC, pp. 140162.
Barrington, S., Choiniere, D., Trigui, M., and Knight, W. 2003. Compost convective
airflow under passive aeration. Bioresource Technology, 86, 259266.
Basker, A., MacGregor, A.N., and Kirman, J.H. 1993. Exchangeable potassium and
other cations in non-ingested soil and cast of two species of pasture earthworms.
Downloaded by [McGill University Library] at 04:56 14 November 2012

Soil Biology and Biochemistry, 25, 16731677.


Bauduin, M., Delcarte, E., and Impens, R. 1986. Pour une valorisation agricole des
composts urbains. Bulletin de Recherches Agronomiques, Gembloux, 21, 349
357.
Beauchamp, C.J., Charest, M.H., and Gosselin, A. 2002. Examination of environmental
quality of raw and composting de-inking paper sludge. Chemosphere, 46, 887
895.
Beck-Friis, B., Smars, S., Jonsson, H., and Kirchmann, H. 2001. Emissions of carbon
dioxide, ammonia and nitrous oxide from organic household waste in a compost
reactor under different temperature regimes. Journal of Agricultural Engineering
Research, 78(4), 423430.
Beffa, T., Blanc, M., Lyon, P.F., Vogt, G., Marchiani, M., Fischer, J.L., and Aragano,
M. 1996a. Isolation of Thermus strains from hot composts 6080 C. Applied
Environmental Microbiology, 62, 17231727.
Beffa, T., Blanc, M., Marilley, l., Fischer, J.L., Lyon, P.F., and Aragno, M. 1996b. Tax-
onomic and metabolic microbial diversity during composting. In: The science
of composting, Part I, M. Bertoldi, P. Sequi, B. Lemmes, and T. Papi (Eds.),
Chapman and Hall, London, pp. 149161.
Benito, M., Masaguer, A., Moliner, A., Arrigo, N., and Palma, R.M. 2003. Chemical and
microbiological parameters for the characterization of the stability and maturity
of pruning waste compost, Biology and Fertility of Soils, 37, 184189.
Bernal, M.P., Parades, C., Monedero, S.A., and Cegarra, J. 1998a. Maturity and sta-
bility parameters of composts prepared with a wide range of organic wastes,
Bioresource Technology, 63, 9199.
Bernal, M.P., Navarro, A.F., Sanchez-Monedero, M.A., Roig, A., and Cegarra, J. 1998b.
Influence of sewage sludge compost stability and maturity on carbon and nitro-
gen mineralization in soil. Soil Biology and Biochemistry, 30, 305313.
Bertoldi, M. de, and Vallini, P.A. 1983. The biology of composting: A review. Waste
Management Research, 1, 157176.
Bertoldi, M. de, Zucconi, F., and Civilini, M. 1988. Temperature, pathogen control
and product quality. Biocycle, 29(2), 4350.
Beukema, K.J., Bruin, S., and Schenk, J. 1983. Three-dimensional natural convec-
tion in a confined porous medium with internal heat generation. International
Journal of Heat and Mass Transfer, 26, 451458.
Solid Waste Management by Composting 381

Bharadwaj, K.K.R. 1995. Improvements in microbial compost technology: A special


reference to microbiology of composting. In: Wealth from waste, Khanna, S.,
and Mohan, K. (Eds.), Tata Energy Research Institute, New Delhi, pp. 115135.
Bittenbender, H.C., Hue, N.V., and Hilary, B.K. 1998. Sustainability of organic fertil-
ization of macadamia with macadamia husk-manure compost. Communications
in Soil Science and Plant Analyses, 29(34), 409419.
Bollen, G.J. 1969. The selective effect of heat treatment of the microflora of a green
house soil, Netherlands Journal of Plant Pathology, 75, 157163.
Bollen, G.J., Volker, D., and Wijnen, A.P. 1989. Inactivation of soil-borne plant
pathogens during small scale composting of crop residues. Netherlands Jour-
nal of Plant Pathology, 95(suppl. 1), 1930.
Bolt, G.H., and Bruggenwert, M.G.M. 1988. Soil chemistry A. Basic elements, 4th ed.
Elsevier, Amsterdam.
Downloaded by [McGill University Library] at 04:56 14 November 2012

Boulter, J.I., Boland, G.J., and Trevors, J.T. 2000. Compost: A study of the develop-
ment process and end-product potential for suppression of turfgrass disease.
World Journal of Microbiology & Biochemistry, 16, 115134.
Boulter, J.I., Boland, G.J., and Trevors, J.T. 2002.Assessment of compost for sup-
pression of Fusarium patch (Microdochium nivale) and Typhula blight (Ty-
phula ishikariensis) snow molds of turfgrass. Biological Control, 25, 162
172.
Brady, N.C., and Weil, R.R. 1999. The nature and properties of soils, 12th ed., Prentice
Hall, Englewood Cliffs, NJ.
Brink, N. 1995. Composting of food waste with straw and other carbon sources for
nitrogen catching.Acta Agriculturae Scandinavica Section B Soil Plant Science,
45, 118123.
Bugbee, G.J., and Frink, C.R. 1989. Composted wasted as apeat substitute in peat-lite
media. HortScience, 24, 625627.
Burge, W.D., 1983. Monitoring pathogen destruction. Biocycle, 24(2), 4850.
Businelli M., Gigliotti G., and Giusquiani, P.L. 1996. Trace element fate in soil profile
and corn plant after massive applications of urban waste compost: A six-year
study. Agrochimica, 4, 145152.
Cambardella, C.A., Richard, T.L., and Russell, A. 2003. Compost mineralization in
soil as a function of composting process conditions. European Journal of Soil
Biology, 39, 117127.
Canet, R., and Pomares, F. 1995. Changes in physical, chemical and physicochemical
parameters during the composting of municipal solid wastes in two plants in
Valencia. Bioresource Technology, 51, 259264.
Casale, W.L., Minassian, V., Menge, J.A., Lovatt, C.J., Pond, E., Johnson, E., and
Guillermet, F. 1995. Urban and agricultural wastes for use as mulches on avocado
and citrus and delivery of microbial agents. Journal of Horticultural Science, 70,
315322.
Castaldi, P., Alberti, G., Merella, R., and Melis, P. 2005. Study of the organic matter
evolution during municipal waste solid composting aimed at identifying suitable
parameters for the evaluation of compst maturity. Waste Management, 25, 209
213.
Chan, P.L.S., and Griffiths, D.A. 1988. The vermicomposting of pre-treated pig ma-
nure, Biological Wastes, 24, 5769.
382 S. Gajalakshmi and S. A. Abbasi

Chanyasak, V., Katayama, A., Hirai, M.F., Mori, S., and Kubota, H. 1983. Effects
of compost maturity on growth of Komatsuna (Brassica rapa var. pervidis) in
Neubauers pot. II. Growth inhibitory factors and assessment of degree of ma-
turity by org.-C/org.-N ratio of water extract. Soil Science and Plant Nutrition,
29, 251259.
Chica, A., Mohedo, J.J., Martin, M.A., and Martin, A., 2003. Determination of the
stability of MSW compost using a respirometric technique. Compost Science
and Utilization,11, 169175.
Chowdappa, P., Biddappa, C.C., and Sujatha, S., 1999. Efficient recycling of organic
wastes in arecanut (Areca catechu) and cocoa (Theobroma cacao) plantation
through vermicomposting. Indian Journal of Agricultural Science, 69, 563
566.
Clarke, W.P., Taylor, M., and Cossins, R. 2007. Evaluation by respirometry of the
Downloaded by [McGill University Library] at 04:56 14 November 2012

loading capacity of a high rate vermicompost bed for treating sewage sludge
Bioreource Technology, 98, 26112618.
Conti, M., Arrigo, N., and Marelli, H. 1997. Relationship of soil carbon light frac-
tion, microbial activity, humic acid production and nitrogen fertilization in the
decaying process of corn stubble.Biology and Fertility of Soils, 25, 7578.
Cooney, D.G., and Emerson, R. 1964. Themophilic fungi: An account of their biology,
activities and classification, W.H. Freeman, San Francisco.
Cooper, R.C., and Golueke, C.G. 1977. Public health aspects of on-site waste treat-
ment. Compost Science, 18(3), 811.
Cooperband, L.R. 2000. Composting: art and science of organic waste conversion to
a valuable soil resource. Laboratory Medicine, 31, 283289.
Coventry, E., Noble, R., Mead, A., and Whipps, J.M. 2002. Control of Allium white
rot (Sclerotium cepivorum) with composted onion waste. Soil Biology and Bio-
chemistry, 34, 10371045.
Cox, D., Bezdicek, D., and Fauci, M. 2001. Effects of compost, coal ash and straw
amendments on testing the quality of eroded Palouse soil. Biology and Fertility
of Soils, 33,365372.
CPCB, 2006. Mangement of municipal solid waste, CPCB Division Activties, Planning
Division (PCP), Central Pollution Control Board, New Delhi, India.
Crawford, J.H. 1983. Composting of agricultural wastes: A review. Process Biochem-
istry, 18, 1418.
Crisan, E.V. 1973. Current concepts of thermophilism and the thermophilic fungi.
Mycologia, 65, 11711198.
Curry, J.P., and Byrne, D. 1992. The Role of earthworms in straw decomposition and
nitrogen turnover in arable land in Greenland. Soil Biology and Biochemistry,
24, 14091412.
De Baere, L., Verdonck, O., and Verstraete, W. 1986. High rate dry anaerobic com-
posting process for the organic fraction of solid wastes. Biotechnology and Bio-
engineering Symposium, 15, 321330.
De Ceuster, T.J.J., and Hoitink, H.A.J. 1999. Prospects for composts and biocontrol
agents as substitutes for methyl bromide in biological control of plant diseases.
Compost Science and Utilization, 7, 615.
DeGroot, R. 1956. Lutilisation du compost urbain dans la culture des plantes orna-
mentales de la region Gantoise. Revue de lagriculture (Bruxelles). 14, 517523.
Solid Waste Management by Composting 383

Deolalikar, A. V., and Mitra, A. 1997a. Vermicompost of paper mill solid waste: A
potential low cost source of organic manure for primary production in aquacul-
ture. In: Proceedings of the International Bioethics Workshop: Biomanagement
of Biogeoresources, J. Azariah et al. (eds.), University of Madras, 1619 January
1997.
Deolalikar, A. V., and Mitra, A. 1997b. Application of paper mill solid waste vermi-
compost as organic manure in rohu (Labeo rohita, Hamilton) cultureA com-
parative study with other commercial organic manure. In: Proceedings of the In-
ternational Biocthics Workshop: Biomanagement of Biogeoresources, J. Azariah
et al. (eds.), University of Madras, 1619 Janurary 1997.
Diaz-Burgos, M.A., Ceccanti, B., and Polo, A. 1993. Monitoring biochemical activity
during sewage sludge composting. Biology and Fertility of Soils, 16, 145150.
Diaz, M.J., Madejon, E., Lopez, F., Lopez, R., and Cabrera, F. 2002a. Composting of
Downloaded by [McGill University Library] at 04:56 14 November 2012

vinasse and cotton gin waste by using two different systems. Resources, Con-
servation and Recycling, 34, 235248.
Diaz, M.J., Madejon, E., Lopez, F., Lopez, R., and Cabrera, F., 2002b. Optimization of
the rate vinasse/grape marc for co-composting process. Process Biochemistry,
37, 11431150.
Diaz, M.J., Eugenio, M.E., Jimenez, L., Madejon, E., and Cabrera, F. 2003. Model-
ing vinasse/cotton waste ratio incubation for optimum composting. Chemical
Engineering Journal, 93, 233240.
Dick, W.A., and McCoy, E.I. 1993. Enhancing soil fertility by addition of compost.
In: Hoitink, H.A.J., and Keener, H.M (Eds.), Science and engineering of com-
posting: Design, environmental, microbiological and utilization aspects, Renais-
sance Publications, Worthington, OH, pp. 623644.
Diez, T., and Krauss, M. 1997. Effect of long-term compost application on yield and
soil fertility. Agrobiological ResearchZeitschrift Fur Agrar biologie Agrikultur-
chemie Oikologie, 50(1), 7884.
Dix, N.J., and Webster, J. 1995. Fungal ecology. Chapman and Hall, Cambridge, Great
Britain.
Dresboll, D.B., and Thorup-Kristensen, K. 2005. Delayed nutrient application affects
mineralization rate during composting of plant residues. Bioresource Technol-
ogy, 96, 10931101.
Eghball, B., 2002. Soil properties as influenced by phosphorous- and nitrogen-based
manure and compost applications. Agronomy Journal, 94, 128135.
Eklind, Y., and Kichmann, H. 2000a. Composting and storage of organic household
waste with different litter amendments I: Carbon turnover. Bioresource Technol-
ogy, 74, 115124.
Eklind, Y., and Kichmann, H. 2000b. Composting and storage of organic household
waste with different litter amendments II: Nitrogen turnover and losses. Biore-
source Technology, 74, 125133.
Elizabeth, S.R. 2006. Studies on vermireactors efficiency as a function of re-
actor geometry. MPhil dissertation, Pondicherry University, Pondicherry,
India.
Elliot, P.W., Knight, D., and Anderson, J.M. 1990. Denitrification in earthworm casts
and soil from pastures under different fertilizer and drainage regimes. Soil Biol-
ogy and Biochemistry, 22, 601605.
384 S. Gajalakshmi and S. A. Abbasi

El-Masry, M.H., Khalil, A.I., Hassouna, M.S., and Ibrahim, H.A.H. 2002. In vitro and
in vivo suppressive effect of agricultural composts and their water extracts on
some phytppathogenic fungi. World Journal of Microbiology and Biotechnology,
18, 551558.
Elorrieta, M.A., Estrella, F.S., Lopez, M.J., Garcia, V., and Moreno, J. 2003. Survival
of phytopathogenic bacteria during waste composting. Agriculture Ecosystems
and Environment, 96, 141146.
Epstein, E. 1997. The science of composting. Technomic Publishing, Lancaster, PA.
Eriksson, K.E.L., Blanchettes, R.A., and Ander, P. 1990. Microbial and enzymatic
degradation of wood and wood components. Springer, Berlin, Germany.
FAO. 1980. A manual of rural composting. FAO/UNDP Regional Project RAS/75/004
Field Document 15.
Fermor, T.R., Wood, D.A., and Lynch, J.M. 1989. Microbiological processes in com-
Downloaded by [McGill University Library] at 04:56 14 November 2012

post. In: International Symposium on Compost Production and Use, San Michele
AllAdige, Italy, pp. 282300.
Fernandes, L., Zhan, W., Patni, N.K., and Jui, P.Y. 1994. Temperature distribution and
variation in passively aerated static compost piles. Bioresource Technology, 48,
257263.
Ferrer, J., Paez, G., Marmol, Z., and Ferrer, A. 2001. Agronomic use of biotechnolog-
ically processed grape wastes, Bioresource Technology, 76, 3944.
Finstein, M.S., and Morris, M.L., 1975. Microbiology of municipal solid waste com-
posting. Advances in Applied Microbiology, 19, 113151.
Finstein, M.S., Cirello, J., Suler, D.J., Morris, M.L., and Strom, P.F. 1980. Microbial
ecosystems responsible for anaerobic digestion and composting, Journal of the
Water Pollution Control Federation, 52, 26752685.
Finstein, M.S., and Miller, F.C. 1985. Principles of composting leading to maximiza-
tion of decomposition rate, odor control and cost effectiveness. In: Composting
of agricultural and other wastes, J.K.R. Gasser (Ed.), Elsevier Applied Science
Publication, Barking, Essex, UK, pp. 1326.
Finstein, M.S., Miller, F.C., and Strom, P.F. 1986. Waste treatment composting as a
controlled system. Biotechnology, 8, 363398.
Finstein, M.S., Miller, F.C., Hogan, J.A., and Strom, P.F. 1987. Analysis of EPA guid-
ance on composting sludge, Part IBiological heat generation and temperature.
Biocycle, reprinted from January, February, March, and April 1987.
Firestone, M.K. 1982. Biological denitrification. In: F.J. Stevenson (Ed.), Nitro-
gen in agricultural soils, ASA, CSSA and SSSA, Inc. Madison, WI, pp. 289
326.
Fitzpatrick, G.E., and Verkade, S.D., 1991. Substrate influence on compost efficacy
as a nursery growing medium. Proceedings of the Florida State Horticultural
Society, 104, 308310.
Fitzpatrick, G.E. 1998. Compost and quality assurance, In: D.Tonnessen (Ed.), Com-
post use in Florida, Florida Department of Environmental Protection, Tallahassee.
Fitzpatrick, G.E. 2001. Compost utilization in ornamental and nursery crop produc-
tion systems. In: Stoffella, P.J., and Kahn, B.A. (Eds.), Horticultural cropping
systems, Lewis Publishers, Boca Raton, FL, pp. 135150.
Fitzpatrick, G.E., Worden, E.C., and Vendrame, W.A. 2005. Historical development
of composting technology during the 20th century. HortTechnology, 15(1), 48
51.
Solid Waste Management by Composting 385

Focht, D.D., and Verstraete, W. 1977. Biochemical ecology of nitrification and deni-
trification. Advances in Microbiology and Ecology, 1, 135214.
Fogarty, A.M., and Tuovinen, O.H. 1991. Microbiological degradation of pesticides
in yard waste composting. Microbiological Reviews, 55, 225233.
Forster, J.C., Zech, W., and Wurdinger, E. 1993. Comparison of chemical and micro-
biological methods for the characterization of the maturity of composts from
contrasting sources. Biology and Fertility of Soils, 16, 9399.
Frobisher, M., Hindsill, R.D., Crabtree, K.T., and Goodheart, C.R. 1974. Fundamentals
of microbiology. W.B. Saunders, Philadelphia, p. 108.
Frost, D.I., Toth, B.L., and Hoitink, H.A.J. 1992. Compost stability. Biocycle, 33, 6266.
Fukumoto, Y., Osada, T., Hanajima, D., and Haga, K. 2003. Patterns and quantities of
NH3 , N2 O and CH4 emissions during swine manure composting without forced
aerationEffect of compost pile scale. Bioresource Technology, 89, 109114.
Downloaded by [McGill University Library] at 04:56 14 November 2012

Gaind, S., and Gaur, A.C. 2003. Quality assessment of compost prepared from fly
ash and crop residue. Bioresource Technology, 87, 125127.
Gajalakshmi, S., 2003. Development of methods for the treatment and reuse of munic-
ipal and agricultural solid wastes appropriate for rural/suburban households.
PhD thesis, Pondicherry University, Pondicherry, India.
Gajalakshmi, S., and Abbasi, S.A. 2002. Effect of the application of water hyacinth
compost/vermicompost on the growth and flowering of Crossandra undulae-
folia, and on several vegetables. Bioresource Technology, 85, 197199.
Gajalakshmi, S., and Abbasi, S.A. 2003. High-rate vermicomposting systems for re-
cycling paper waste.Indian Journal of Biotechnology, 2, 613615.
Gajalakshmi, S., and Abbasi, S.A. 2004a. Earthworms and vermicomposting. Indian
Journal of Biotechnology, 3, 486494.
Gajalakshmi, S., and Abbasi. S.A. 2004b. Neem leaves as a source of vermicompost-
cum-pesticide. Bioresource Technology, 92, 291296.
Gajalakshmi, S., Ramasamy, E.V., and Abbasi, E.V. 2001a. Potential of two epigeic
and two anecic earthworm species in vermicomposting of water hyacinth. Biore-
source Technology 76, 177181.
Gajalakshmi, S., Ramasamy, E.V., and Abbasi, E.V. 2001b. Screening of four species
of detritivorous (humus-former) earthworms for sustainable vermicomposting
of paper waste. Environmental Technology, 22, 679685.
Gajalakshmi, S., Ramasamy, E.V., and Abbasi, E.V. 2001c. Towards maximizing output
from vermireactors fed with cowdung spiked paper waste. Bioresource Tech-
nology, 79, 6772.
Gajalakshmi, S., Ramasamy, E.V., and Abbasi, E.V. 2001d. Assessment of sustainable
vermiconversion of water hyacinth at different reactor efficiencies employing
Eudrilus eugeniae Kinberg. Bioresource Technology, 80, 131135.
Gajalakshmi, S., Ramasamy, E.V., and Abbasi, E.V. 2002a. Vermicompositng of dif-
ferent forms of water hyacinth by the earthworm Eudrilus eugeniae, Kinberg.
Bioresource Technology, 82, 165169.
Gajalakshmi, S., Ramasamy, E.V., and Abbasi, E.V. 2002b. Vermicomposting of paper
waste with the anecic earthworm, Lampito mauritii Kinberg, Indian Journal of
Chemical Technology, 22, 679682.
Gajalakshmi, S., Ramasamy, E.V., and Abbasi, E.V. 2002c. High-rate composting
Vermicomposting of water hyacinth (Eichhornia crassipes, Mart. Solms). Biore-
source Technology, 83, 235239.
386 S. Gajalakshmi and S. A. Abbasi

Gajalakshmi, S., Ramasamy, E.V., and Abbasi, S.A. 2005a. Composting-


vermicomposting of leaf litter ensuing from the trees of mango (Mangifera
indica) Kinberg. Bioresource Technology, 96, 10571061.
Gajalakshmi, S., Ramasamy, E.V., and Abbasi, S.A. 2005b. A highly cost-effective
simplification in the design of fast-pacedc vermireactors based on epigeic earth-
worms. Biochemical Engineering Journal, 22, 111116.
Garcia, C., Hernandez, T., and Costa, F. 1991. Changes in carbon fractions during
composting and maturation of organic wastes. Environmental Management, 15,
433439.
Garcia, C., Hernandez, T., Costa, F., and Ayusho, M. 1992a. Evaluation of maturity of
municipal waste compost using simple chemical parameters. Communications
in Soil Science and Plant Analyses, 23, 15011512.
Garcia, C., Hernandez, T., Costa, F., Ceccati, B., and Ciardi, C. 1992b. Changes in
Downloaded by [McGill University Library] at 04:56 14 November 2012

ATP content, enzyme activity and inorganic nitrogen species during composting
of organic wastes. Canadian Journal of Soil Science, 72, 243253.
Garcia, C., Pascual, J.A., Mena, E., and Hernanadez, T. 2004. Influence of the sta-
bilization of organic materials on their biopesticide effect in soils. Bioresource
Technology, 95, 215221.
Garcia-Gomez, A., Roig, A., and Bernal, M.P. 2003. Composting of the solid frac-
tion of olive mill wastewater with olive leaves: Organic matter degradation and
biological activity. Bioresource Technology, 86, 5964.
Garg, S.K., and Neelakantan, S., 1982. Bioconversion of sugarcane bagasse for cel-
lulase enzyme and microbial protein production. Journal of Food Technology,
17, 271279.
Gaur, A.C. 2000. Bulky organic manures and crop residues. In: H.L.S. Tandon (Ed.),
Fertilizers, organic manures, recyclable wastes and biofertilizers, Fertilizer De-
velopment and Consultation Organization, New Delhi, India.
Gestel, K.V., Mergaert, J., Swings, J., Coosemans, J., and Ryckeboer, J. 2003. Biore-
mediation of diesel oil-contaminated soil by composting with biowaste. Envi-
ronmental Pollution, 125, 361368.
Ghorpade, V.M., Gennadios, A., and Hanna, M.A. 2001. Laboratory composting of
extruded poly (lactic acid) sheets. Bioresource Technology, 76, 5761.
Ghosh, M., Chattopadhyay, G.N., and Baral, K. 1999. Transformation of phosphorous
during vermicomposting. Bioresource Technology, 69, 149154.
Ghosh, S., Kapadnisb, B.P., and Singh, N.B. 2000. Composting of cellulosic hospital
solid waste: A potentially novel approach. International Biodeterioration and
Biodegradation, 45, 8992.
Ginkel, V.J.T. 1996. Physical and biochemical processes in composting material. PhD
Thesis, Wageningen Agricultural University, Wageningen, The Netherlands.
Godden, B., Penninckx, M., Perard, A., and lannoye, R. 1983. Evolution of enzyme
activities during composting of cattle manure. European Journal of Applied Mi-
crobiology and Biotechnology, 134, 24412448.
Godden, B., Ball, A.S., Helvenstein, P., McCarthy, A.J., and Penninckx, M.J. 1992. To-
wards elucidation of the lignin degradation pathway in actinomycetes. Journal
of General Microbiology, 138, 24412448.
Golueke, C.G. 1972. Composting: A study of the process and its principles, Rodale
Press, Ephrata, PA.
Solid Waste Management by Composting 387

Golueke, C.G. 1981. Principles of biological resource recovery. Biocycle, 22, 36


40.
Golueke, C.G. 1992. Bacteriology of composting. Biocycle,33, 5557.
Goodfellow, M., and Williams, S.T. 1983. Ecology of Actinomycetes. Annual Review
of Microbiology, 37, 189216.
Gotaas, H.B. 1956. Composting, sanitary disposal and reclamation of organic wastes.
World Health Organization, Geneva, Switzerland.
Goyal, S., Dhull, S.K., and Kapoor, K.K. 2005. Chemical and biological changes during
composting of different organic wastes and assessment of compost maturity.
Bioresource Technology, 96,15841591.
Graff, O. 1974. Gewinnung von Biomasse aus Abfallstoffen durch Kultur des Kom-
postregenwurms Eisenia foetida (Savigny 1826). Landbauforschung Volkenrode,
2, 137142.
Downloaded by [McGill University Library] at 04:56 14 November 2012

Grebus, M.E., Watson, M.E., and Hoitink, H.A.J. 1994. Biological, chemical and
physical properties of composted yard trimmings as indicators of matu-
rity and plant disease suppression. Compost Science and Utilization, 2, 57
71.
Guardia, A.D., Brunet, S., Rogeau, D., and Matejka, G., 2002. Fractionation and char-
acterization of dissolved organic matter from composting green wastes. Biore-
source Technology, 83, 181187.
Hadar, Y., and Mandelbaum, R. 1992. Suppressive compost for biocontrol of soil
borne plant pathogens. Phytoparasitica, 20, 113116.
Haider, K. 1992. Problems related to the humification process in soil of temperate
climates. Soil Biochemistry, 7, 5594.
Haider, N., and Karlsson, S. 2000. Loss of Chimassorb 944 from LPDE and identifica-
tion of additive degradation products after exposure to water, air and compost.
Polymer Degradation and Stability, 74, 103112.
Haimi, J., and Huhta, V. 1986. Capacity of various organic residues to support ade-
quate earthworm biomass for vermicomposting. Biology and Fertility of Soils, 2,
2327.
Hanajima, D., Kuroda, K., and Haga, K., 2001. Enhancement of the thermophilic stage
in cattle waste composting by addition of tofu residue. Bioresource Technology,
78, 213216.
Haninen, K., and Lilja, R. 1994. Humification during the composting of slaughter
wastes. In: Senesi, N., and Miano, T. (Eds.), Humic substances in the global
environment and implications on human health, Elsevier Science, Berlin, pp.
12651272.
Hankin, L.H., Poincelot, R.P., and Anagnostakis, S.L. 1976. Microorganisms from com-
posting leaves: ability to produce extracellular enzymes, Microbial Ecology, 2,
296308.
Hansen, R.C., Keener, H.M., Dick, W.A.., Marugg, C., and Hoitink, H.A.J. 1990. Poul-
try manure composting: Ammonia capture and aeration control. Proceedings
of American Society of Agricultural Engineers meeting, Columbus, OH, paper
894075, pp. 2427.
Harada, Y., and Inoko, A. 1980. Relationship between cation-exchange capacity and
degree of maturity of city refuse composts. Soil Science and Plant Nutrition, 26,
353362.
388 S. Gajalakshmi and S. A. Abbasi

Hartenstein, R., Neuhauser, E.F., and Kaplan, D.L. 1979. Reproductive potential of
the earthworm Eisenia foetida. Oecologia (Berlin), 43, 329340.
Hartlieb, N., Ertunc, T., Schaeffer, A., and Klein, W. 2003. Mineralization, metabolism
and formation of non-extractable residues of [14 C]-labelled organic contaminants
during pilot-scale composting of municipal biowaste. Environmental Pollution,
126, 8391.
Haug, R.T. 1993. The practical handbook of compost engineering, Lewis Publishers,
London.
He, X.T., Traina, S.J., and Logan, T.J. 1992. Chemical properties of municipal solid
waste composts. Journal of Environmental Quality, 21,318329.
He, X.T., Logan, T.J., and Traina, S.J. 1995. Physical and chemical characteristics of
selected US municipal solid waste composts. Journal of Environmental Qual-
ity,24(3), 543552.
Downloaded by [McGill University Library] at 04:56 14 November 2012

He, Y.W., Inamori, Y., Mizuochi, M., Kong, H.N., Iwami, N., and Sun, T.H., 2000.
Measurements of N2 Oand CH4 from the aerated composting of food waste.
Science of the Total Environment, 254, 6574.
Hellmann, B., Zelles, L., Palojarvi, A., and Bai, Q. 1997. Emission of climate-relevant
trace gases and succession of microbial communities during open-windrow
composting. Applied and Environmental Microbiology, 63, 10111018.
Hernandez, D., Sanchez, J.E., and Yamasaki, K. 2003. A simple procedure for prepar-
ing substrate for Pleurotus ostreatus cultivation. Bioresource Technology, 90,
145150.
Higgins, A. 2001. Foreverdung: At Mount Vernon, a restored barn and pit turn
stable waste into black gold for the garden. <http://gwpapers.virginia.edu/
project/news/Washington/dung.html>
Hoitink, H.A.J., and Fahy, P.C. 1986. Basis for the control of soil borne pathogens
with composts. Annual Review of Phytopathology, 24, 93114.
Hoitink, H..A.J., and Kuter, S.A. 1986. Effects of composts in growth media on
soilborne pathogens. In: Y. Chen and Y. Avnimelech (Eds.), The role of or-
ganic matter in modern agriculture, Martinus Nijhoff Publishers, Dordrecht,
The Netherlands, pp. 289306.
Hoitink, H..A.J., and Boehm, M.J. 1999. Biocontrol within the context of soil mi-
crobial communities: a substrate-dependent phenomenon. Annual Review of
Phytopathology, 37, 427446.
Hoitink, H.A.J., Inbar, Y., and Boehm, M.J. 1991. Status of compost-amended potting
mixes naturally suppressive to soilborne diseases in floricultural crops. Plant
Disease, 75, 869873.
Hoitink, H.A.J., Stone, A.G., and Han, D.Y. 1997. Suppression of plant disease by
composts. Horticulture Science, 32, 184187.
Horiuchi, J.I., Ebie, K., Tada, K., Kobayshi, M., and Kanno, T. 2003. Simplified method
for estimation of microbial activity in compost by ATP analysis, Bioresource
Technology, 86, 9598.
Howard, A., and Wad, Y.D. 1931. The waste products of agriculture. Their utilization
as humus, Oxford University Press, London.
Hoyos, S.E.G., Juarez, J.V., Ramonet, A.C., Lopez, J.G., Rios, A.A., and Uribe, E.G.,
2002. Aerobic thermophilic composting of waste sludge from gelatin-grenetine
industry, Resources, Conservation and Recycling, 34, 161173.
Solid Waste Management by Composting 389

Hsu, J.H., and Lo, S.L. 1999. Recycling of separated pig manure: Characterization
of maturity and chemical fractionation of elements during composting. Water
Science and Technology, 40, 121127.
Hue, N., and Liu, J. 1995. Predicting compost stability. Compost Science and Utiliza-
tion, 3, 815.
Iannotti, D.A., Pang, T., Toth, B.L., Elwell, D.L.,Keener, H.M., and Hoitink, H.A.J.
1993. A quantitative respirometric method for monitoring compost stability.
Compost Science and Utilization, 1, 5265.
Iannotti, D.A., Grebus, M.E., Toth, B.L., Madden, L.V., and Hoitink, H.A.J. 1994.
Oxygen respirometry to assess stability and maturity of composted municipal
soild waste. Journal of Environmental Quality, 23, 11771183.
Inbar, Y., Hadar, Y., and Chen, Y. 1993. Recycling of cattle manure: The composting
process and characterization of maturity. Journal of Environmental Quality, 22,
Downloaded by [McGill University Library] at 04:56 14 November 2012

857863.
Ipek, U.,Obek, E., Akca, L., Arslan, E.I., Hasar, H., Dogru, M., and Baykara, O.
2002. Determination of degradation of radioactivity and its kinetics in aerobic
composting, Bioresource Technology, 84, 283286.
Ismail, S.A. 1997. Vermicology: The biology of earthworms. Orient Longman,
Hyderabad, India.
Ismail, S.A. 1998. The contribution of soil fauna especially the earthworms to soil fer-
tility. In: Proceedings of a workshop on organic farming, Ismail, S.A. (ed.), Insti-
tute of Research in soil Biology and Biotechnology, The New College, Chennai,
p. 9.
Jang, J.C., Shin, P.K., Yoon, J.S., Leed, I.M., Lee, H.S., and Kim, M.N. 2002. Glucose
effect on the biodegradation of plastics by compost from food garbage. Polymer
Degradation and Stability, 76, 155159.
Jenkinson, D.S. 1988. Determination of microbial biomass carbon and nitrogen in
soil. In: Wilkinson, J.K (Ed.), Advances in Nitrogen Cycling in Agriculture Ecosys-
tems, CAB International,Wallingford, UK, pp. 368386.
Jenkinson, D.S., and Ladd, J.N., 1981. Microbial biomass in soil: Measurement and
turnover. In: Paul, E.A., and Ladd, J.N (Eds.), Biochemistry, Marcel Dekker, New
York, pp. 414471.
Jeong, Y.K., and Kim, J.S. 2001. A new method for conservation of nitrogen in aerobic
composting processes. Bioresource Technology, 79, 129133.
Jimenez, I., and Garcia, P. 1992. Determination of maturity indices for city refuse
composts Agricultural Ecosystems and the Environment, 38, 331343.
Jimenez, E.I., and Garcia, V.P. 1989. Evaluation of city refuse compost maturity: A
review. Biological Wastes, 27, 115145.
Jimenez, E.I., and Alvaraez, C.E. 1993. Apparent availability of nitrogen in composted
municipal refuse. Bioogy and Fertility of Soils, 16, 313318.
Jones, J., Wolf, B., Jr., and Mills, H. 1991. Plant analysis handbook: A practical sam-
pling preparation, analysis and interpretation guide, Macro-Micro Publishing,
Athens, GA, pp. 2337.
Joshi, N.V., and Kelkar, B. 1951. The role of earthworms in soil fertility. Indian
Joournal of Agricultural Science, 22, 189196.
Kale, R.D., Bano, K., and Krishnamoorthy, R.V. 1982. Potential of Perionyx excavatus
for utilizing organic wastes. Pedobiologia, 23, 419425.
390 S. Gajalakshmi and S. A. Abbasi

Kapetanois, E.G., Loizidu, M., and Vakanas, G. 1993. Compost production from Greek
domestic refuse, Bioresource Technology, 43, 1316.
Kapoor, K.K., Jain, M.M., Mishra, M.M., and Singh, C.P. 1978. Cellulase activity, degra-
dation of cellulose and lignin and humus formation by cellulolytic fungi. Annals
of Microbiology, 129B, 613620.
Keeling, A.A., Mullett, J.A.J., and Paton, I.K., 1994. GC-mass spectrometry of refuse-
derived composts. Soil Biology and Biochemistry, 26, 773776.
Keeling, A.A., Griffiths, B.S., Ritz, K., and Myers, M. 1995. Effects of compost stability
on plant growth, microbiological parameters and nitrogen availability in me-
dia containing mixed garden-waste compost. Bioresource Technology, 54, 279
284.
Kirchmann, H., and Widen, P. 1994. Separately collected organic household wastes
chemical composition and composting characterisitcs. Swedish Journal of Agri-
Downloaded by [McGill University Library] at 04:56 14 November 2012

cultural Research, 24, 312.


Klock-Moore, K.A. 1999. Bedding plant growth in greenhouse waste and biosolid
compost. HortTechnology, 9, 210213.
Klock-Moore, K.A. 2000. Comparison of salvia growth in seaweed compost and
biosoilds compost. Compost Sceince and Utilization, 8, 2428.
Klock-Moore, K.K. 2005. Uses of compost in potting mixes. HortTechnology, 15(1),
5860.
Klock-Moore, K.A., and Fitzpatrick, G.E. 1997. Growth of impatiens Accent Red in
three compost products. Compost Science and Utilization, 5, 2630.
Klopotek, A. 1962. Uber das Vorkommen und Verhalten von Schimmelpilzen bel
der kompostierung Stadtischer Abfallstoffe. Antonie van Leeuwenhoek Journal
of Microbiology, 28, 141160.
Koivula, T., Raikkonen, T., Urpilainen, S., Ranta, J., and Hanninen, K. 2004. Ash
in composting of source-separated catering waste, Bioesource Technology, 93,
291299.
Korboulewsky, N., Bonin, G., and Massiani, C. 2002a. Biological and ecopyhsiolog-
ical reactions of white wall rocket (Diplotaxis erucoides L.) grown on sewage
sludge compost, Environmental Pollution, 117, 365370.
Korboulewsky, N., Duouyet, S., and Bonin, G. 2002b. Environmental risks of ap-
plying sewage sludge compost to vineyards: Carbon, heavy metals, nitrogen
and phosphorous accumulation. Journal of Environmental Quality, 31, 1522
1527.
Korner, I., Braukmeier, J., Herrenklage, J., Leikam, K., Ritzkowski, M., Schlegelmilch,
M., and Stegmann, R. 2003. Investigation and optimization of composting
processesTest systems and practical examples, Waste Management, 23, 1726.
Kostov, O., Rankov, V., Atanacova, G., and Lynch, J.M. 1991. Decomposition of
sawdust and bark treated with cellulose-decomposing microorganisms. Biology
and Fertility of Soils, 11, 105110.
Kostov, O., Tzvetkov, Y., Petkova, G., and Lynch, J.M. 1996. Aerobic composting of
plant wastes and their effect on the yield of ryegrass and tomatoes. Biology and
Fertility of Soils, 23, 2025.
Kubler, H. 1982. Air convection in self-heating piles of wood chips. Tappi, 65, 7983.
Kulcu, R., and Yaldiz, O. 2004. Determination of aeration rate and kinetics of com-
posting some agricultural wastes. Bioesource Technology, 94, 4957.
Solid Waste Management by Composting 391

Kumar, C.A. 1994. State of the art report on vermiculture in India. Council for Ad-
vancement of Peoples Action and Rural Technology (CAPART), New Delhi,
India.
Kwon, S.H., and Lee, D.H. 2003. Evaluation of Korean food waste composting with
fed-batch operations II: Using properties of exhaust gas condensate.Process Bio-
chemistry, 39, 10471055.
Larney, F.J., Olson, A.F., Carcamo, A.A., and Chang, C. 2000. Physical changes during
active and passive composting of beef feedlot manure in winter and summer.
Bioresource Technology, 75, 139148.
Larney, F.J., and Hao, X. 2007. A review of composting as a management alterna-
tive for beef cattle feedlot manure in southern Alberta, Canada. Bioresource
Technology, 98, 32213227.
Lax, A., Roig, A., and Costa, F. 1986. A method for determining the cation-exchange
Downloaded by [McGill University Library] at 04:56 14 November 2012

capacity of organic materials. Plant and Soil, 94, 349355.


Lazzari, L., Sperni, L., Bertin, P., and Pavoni, B. 2000. Correlation between inorganic
(heavy metals) and organic (PCBs and PAHs) micropollutant concentrations dur-
ing sewage sludge composting processes. Chemosphere, 41, 427435.
Lee, J.J., Park, R.D., Kim, Y.W., Shim, J.H., Chae, D.H., Rim, Y.S., Sohn, B.K., Kim,
T.H., and Kim, K.Y. 2004. Effect of food waste compost on microbial population,
soil enzyme activity and lettuce growth. Bioresource Technology, 93, 2128.
Levy, J.S., and Taylor, B.R., 2003. Effects of pulp mill solids and three composts on
early growth of tomatoes. Bioresource Technology, 89, 297305.
Liang, C., Das, K.C., and McClendon, R.W. 2003. The influence of temperature and
moisture contents regeimes on the aerobic microbial activity of a biosoilds com-
posting blend. Bioresource Technology, 86, 131137.
Lin, Q., and Brookes, P.C. 1999. Arginine ammonification as a method to estimate
soil microbial biomass and microbial community structure. Soil Biology and
Biochemistry, 31, 19851997.
Lodha, S., Sharma, S.K., and Aggarwal, R.K. 2002. Inactivation of Macrophomina
phaseolina propagules during composting and effect of composts on dry root rot
severity and on seed yield of clusterbean. European Journal of Plant Pathology,
108, 253261.
Logsdon, G. 1994. Worldwide progress in vermicomposting. Biocycle, 35, 6365.
Lopez-Real, J., and Foster, M. 1985. Plant pathogens survival during composting of
agricultural organic wastes. In: Gasser, J.K.R. (Ed.), Composting of agricultural
and other wastes, Elsevier Applied Science, London, pp. 291299.
Lopez, M.J., Elorrieta, M.A., Vargas-Garcia, M.C., Suarez-Estrella, F., and Moreno, J.
2002. The effect of aeration of lignicellulosic waste by white-rot fungi. Biore-
source Technology, 81, 123129.
Lopez-Real, J., and Baptista, M. 1996. A preliminary comparative study of three
manure composting systems and their influence on process parameters and
methane emissions. Compost Science and Utilization, 4, 7182.
Luo, S., and Netravali, A.N. 2003. A study of physical and mechanical properties of
poly (hydroxybutyrate-co-hydroxyvalerate) during composting. Polymer Degra-
dation and Stability, 80, 5966.
Ma, Y., Zhang, J.Y., and Wong, M.H. 2003. Microbial activity during composting of
anthracene-contaminated soil. Chemosphere, 52, 15051513.
392 S. Gajalakshmi and S. A. Abbasi

MacGregor, S.T., Miller, F.C., Psarianos, K.M., and Finstein, M.S. 1981. Composting
process-control based on the interaction between microbial heat output and
temperature. Applied and Environmental Microbiology, 41, 13211330.
Madejon, E., Lopez, R., Murillo, J.M., and Cabrera, F. 2001a. Agricultural use of three
(sugar-beet) vinasse composts: Effect on crops amd chemical properties of a
cambisol soil in the Guadalquivir river valley (SW Spain). Agriculture, Ecosystems
and Environment, 84, 5565.
Madejon, E., Dyaz, M.J., Lopez, R., and Cabrera, F. 2001b. Co-composting of sugar-
beet vinasse: Influence of the organic matter nature of the bulking agents used,
Bioresource Technology, 76, 275278.
Manios, T., and Stentiford, E. 2003. Sanitary aspect of using partially treated land. II.
Leachate as a water source in green waste composting. Waste Management, 24,
107110.
Downloaded by [McGill University Library] at 04:56 14 November 2012

Manios, T. 2004. The composting potential of different organic solid wastes: Experi-
ence from the island of Crete. Environment International, 29, 10791089.
Manser, A.G.R., and Keeling, A.A. 1996. Practical book of processing and recycling
municipal waste, CRC Press, Boca Raton, FL.
Marambe, B., and Ando, T. 1990. Phenolic acids as potential seed germination-
inhibitors in animal-waste composts. Soil Science and Plant Nutrition, 38, 727
733.
Marche, T., Schnitzer, M., Dinel, H., Pare, T., Champagne, P., Schulten, H.R., and
Facey, G. 2003. Chemical changes during composting of a paper mill sludge-
hardwood sawdust mixture. Geoderma, 116, 345356.
Mari, I., Ehaliotis, C., Kotsou, M., Balis, C., and Georgakakis, D. 2003. Respiration
profiles in monitoring the composting of by-products from the olive oil agro-
industry. Bioresource Technology, 87, 331336.
Mathur, S.P. 1991. Composting process. In: A.M. Martin (ed.), Bioconversion of waste
materials to industrial products, Elsevier, London, pp. 147183.
Mathur, S.P., Owen, G., Dinel, H., and Schnitzer, M. 1993. Determination of compost
biomaturity, I. Literature review. Biological Agriculture and Horticulture, 10,
6585.
Mays, D.A., Terman, G.L., and Duggan, J.C. 1973. Municipal compost: Effects on crop
yields and soil properties. Journal of Environmental Quality, 2, 8992.
Mbuligwe, S.E., Kassenga, G.R., Kaseva, M.E., and Chaggu, E.J. 2002. Potential and
constraints of composting domestic solid waste in developing countries: Find-
ings from a pilot study near in Dar es Salaam, Tanzania. Resources, Conservation
and Recycling, 36, 4559.
McConnell, D.D., Shiralipour, A., and Smith, W.H., 1993. Compost application im-
proves soil properties. Biocycle, 34, 6163.
McKinley, V.L., Vestal, J.R., and Eralp, A.E. 1985. Microbial activity in composting:
Part I. Biocycle, 26, 3943.
Michel, C., Jr., Forney, L.J., Huang, A.J.F., Drew, S., Czuprenski, M., Lindeberg, J.D.,
and Reddy, C.A. 1996. Effects of turning frequency, leaves to grass mix ratio, and
windrow vs pile configuration on the composting of yard trimmings. Compost
Science and Utilization, 4, 2643.
Miller, F.C. 1992. Biodegradation of soild wastes by composting. In: Biological degra-
dation of wastes, Martin, A.M. (Ed.), Elsevier Applied Science, London, pp. 130.
Solid Waste Management by Composting 393

Misra, R.V., and Roy, R.N. 2007. On farm composting methods. FAO, Rome.
www.fao.org/organicag/doc.
Mondini, C., Contin, M., Leita, L., and Nobili, M.D. 2002. Response of microbial
biomass to air-drying and rewetting in soils and compost. Geoderma, 105, 111
124.
Mondini, C., Fornasier, F., and Sinicco, T. 2004. Enzymatic activity as a parameter for
the characterization of the composting process. Soil Biology & Biochemistry, 36,
15871594.
Morel, J.L., Guckert, A., Nicolardot, B., Benistant, D., Catrou, G., and Germon,
J.C. 1979. Etude de levolution des caracteristiques physico-chimiques et de la
stabilite biologique des ordures menageres au cours du compostege.
Agronomie, 6, 693701.
Morisaki, N., Phae, C.G., Nakasaki, K., Shoda, M., and Kubota, H. 1989. Nitrogen
Downloaded by [McGill University Library] at 04:56 14 November 2012

transformation during thermophilic composting. Journal of Fermentation and


Bioengineering, 67, 5761.
Mosher, D., and Anderson, R.K. 1977. Composting sewage sludge by high-rate suction
aeration techniquesThe process as conducted at Bangor, ME, and some guide-
lines of general applicability. Interim Report Number SW-614d. U.S. Government
Printing Office, Washington, DC.
Moubray, J.M. 1943. Compost making at Chipoli, Southern Rhodesia. In: Howard,
A., An agricultural testament, Oxford University Press, New York, 367
375.
Mustin, M. 1987. Le compost, gestion de la matiere organique, F. Dubuse, Paris.
Muthukumar, G., and Mahadevan, A. 1983. Microbial degradation of lignin. Journal
of Science and Industrial Research, 42, 518528.
Nair, J., Sekiozoic, V., and Anda, M. 2006. Effect of pre-composting of kitchen waste.
Bioresource Technology, 97, 20912095.
Nakasaki, K., Sasaki, M., Shoda, M., and Kubota, H. 1985. Change in micro-
bial numbers during thermophilic composting of sewage sludge with refer-
nce to CO2 evolution rate. Applied Environmental Microbiology, 49, 37
41.
Nakasaki, K., Yaguchi, H., Sasaki, M., and Kubota, H. 1993. Effects of pH control on
composting of garbage. Waste Management and Research, 11, 117125.
Nakasaki, K., Uehara, N., Kataoka, M., and Kubota, H. 1996. The use of Bacillus
licheniformis HAI to accelerate composting of organic waste. Compost Science
and Utilization, 4(4), 4751.
Nell, J.H., and Wiechers, S.G. 1978. High temperature composting. Water Science
Africa, 4, 203212.
Nyns, E.J. 1986. Can biomethanation be included in the processing of compost-
like materials? In: Compost: Production, quality and use, Bertoldi, M. de,
Ferranti, M.P., Hermite, P.I., and Zucconi, F. (Eds.), Elsevier, London, pp. 97
99.
Oades, L.M., and Jenkinson, D.S. 1979. Adenosine triphosphate content of the soil
microbial biomass. Soil Biology and Biochemistry, 11, 201204.
Ott, P. 1990. The composting of farmyard manure with mineral additives and under
forced aeration, and the utilization of FYM and FYM compost in crop production.
PhD thesis, Universitat des Landes, Hessen.
394 S. Gajalakshmi and S. A. Abbasi

Ozores-Hampton, M., and Bryan, H.H. 1993. Municipal solid waste (MSW) soil
amendments: Influence on growth and yield of snap beans. In: Proceedings
of Florida State Horticultural Society, 106, 208210.
Pandey, B.P. 1982. Taxonomy of angiosperms, 4th ed., S. Chand and Company,
New Delhi, India.
Parades, C., Bernal, M.P., Cegarra, J., Roig, A., and Navarro, A.F. 1996. Nitrogen trans-
formation during the composting of different organic wastes. In: Van Cleemput,
O., Hofman, G., and Vermoesen, A. (Eds.), Progress in nitrogen cycling studies,
Kluwer Academic Publishers, Dordrecht, pp. 121125.
Parades, C., Bernal, M.P., Cegarra, J., and Roig, A. 2002. Bio-degradation of olive mill
wastewater sludge by its co-composting with agricultural wastes. Bioresource
Technology, 85, 18.
Parades, C., Cegarra, J., Bernal, M.P., and Roig, A. 2005. Influence of olive mill
Downloaded by [McGill University Library] at 04:56 14 November 2012

wastewater in composting and impact of the compost on a Swiss chard crop


and soil properties, Environment International, 31, 305312.
Pare, T., Dinel, H., Schnitzer, M., and Dumontet, S., 1998. Transformations of carbon
and nitrogen during composting of animal manure and shredded paper, Biology
and Fertility of Soils, 26, 173178.
Parkin, T., and Berry, E. 1994. Nitrogen transformations associated with earthworm
casts. Soil Biology and Biochemistry, 26, 12331238.
Peigne, J., and Girardin, P. 2004. Environmental impacts of farm-scale composting
practices. Water, Air and Soil Pollution, 153, 4568.
Pelaez, C., Mejia, A., and Planas, A. 2004. Development of a solid phase kinetic assay
for determination of enzyme activities during composting. Process Biochemistry,
39, 971975.
Perez-Murcia, M.D., Moreno-Caselles, J., Moral, R., Perez-Espinosa, A., Parades, C.,
and Rufete, B. 2005. Use of composted sewage sludge as horticultural growth
media : Effects on germination and trace element extraction. Communications
in Soil Science and Plant Analyses, 36, 571582.
Pogue, D.J., and Amer, R. 1997. George Washington, the revolutionary farmer:
Americas first composter. www.cityfarmer.org/Washington.html
Poincelot, R.P. 1974. A scientific examination of the principles and practice of com-
posting. Composting Science, 15, 2431.
Poole, R.T., Conover, C.A., and Joiner, J.N. 1981. Soils and potting mixtures. In: J. N.
Joiner (ed.), Foliage plant production, Prentice Hall, Englewood Cliffs, NJ, pp.
179202.
Priest, P. 1984. Extracellular enzymes, American Society for Microbiology,
Washington, DC.
Prince, S.P.M., Sivakumar, W.S., Ravi, V., and Subburam, V. 2000. The effects of
coirpith compost on the growth and quality of leaves of the mulberry plant
Morus alba L. Bioresource Technology, 72, 9597.
Rajendran, S. 1994. Exploiting earthworms for fertilizers. Down to Earth, 2(21), 2223.
Raviv, M. 2005. Production of high-quality composts for horticultural purposes: A
mini-review. HortTechnology, 15(1), 5257.
Raviv, M., Oka, Y., Katan, J., Hadar, Y., Yogev, A., Medina, S., Krasnovsky, A., and
Ziadna, H. 2005. High-nitrogen compost as a medium for organic container-
grown crops. Bioresource Technology, 96, 419427.
Solid Waste Management by Composting 395

Raymond, D.A., Chong, C., and Voroney, R.D. 1998. Response of four container
grown woody ornamentals to immature composted media derived from waxed
corrugated cardboard. Compost Science and Utilization, 6, 6774.
Reddy, K.R., Kaleel, R., Overcash, M.R., and Westerman, P.W. 1979. A nonpoint
source model for land areas receiving animal wastes: Ammonia volatilization.
Transactions of the ASAE, 22, 13981405.
Reinecke, A.J., and Venter, J.M. 1987. Moisture preferences, growth and reproduction
of the compost worm Eisenia foetida (Oligochaeta). Biology and Fertility of Soils,
10, 135141.
Reinecke, A.J., Viljoen, S.A., and Saayman, R.J. 1992. The suitability of Eudrilus euge-
niae, Perionyx excavatus, and Eisenia fetida (Oligocheeta) for vermicomposting
in Southern Africa in terms of their temperature requirements. Soil Biology and
Biochemistry, 24, 12951307.
Downloaded by [McGill University Library] at 04:56 14 November 2012

Reuveni, R., Raviv, M., Krasnovsky, A., Freiman, L., Medina, S., Bar, A., and Orion, D.
2002. Compost induces protection against Fusarium oxysporum, in sweet basil.
Crop Protection, 21, 583587.
Ribeiro, H.M., Vasconcelos, E., and dos Santos, J.Q. 2000. Fertilization of potted
geranium with a municipal solid waste compost. Bioresource Technology, 73,
247249.
Robertsson, M. 2001. Effects of interrupted air supply of the composting process
composition of volatile organic acids. In: Proceedings, Microbiology of Compost-
ing, Innsbruck, Austria, October 1820, 2000, pp. 189202.
Rodriguez-Kabana, R., and Truelove, B. 1982. Effects of crop rotation and fertilization
on catalase activity in a soil of the south-eastern United States. Plant and Soil,
69, 97104.
Roig, A., Lax, A., Ceggara, J., Costa, F., and Hernandez, M.T. 1988. Cation exchange
capacity as a parameter for measuring the humification degree of manures, Soil
Science, 146, 311316.
Ros, M., Hernandez, M.T., Garcia, C., Bernal, A., and Pascual, J.A. 2005. Biopesti-
cide effect of green compost against Fusarium wilt on melon plants. Journal of
Applied Microbiology, 98, 845854.
Rotenberg, D., Cooperband, L., and Stone, A. 2005. Dynamic relationships between
soil properties and foliar disease as affected by annual additions of organic
amendment to a sandy-soil vegetable production system. Soil Biology & Bio-
chemistry, 37, 13431357.
Rynk, R., van de Kamp, M., Willson, G.B., Singley, M.E., Richard, T.L., Kolega, J.L.,
Gouin, F.R., Laliberty, L., Jr., Kay, D., Murphy, D.W., Hoitink, H.A.J., and Brinton,
W.F. 1992. On-farm composting handbook, Natural Resource, Agriculture and
Engineering Service, Cornell University, Ithaca, NY.
Sanchez-Monedero, M.A., Roig, A., Parades, C., and Bernal, M.P. 2001. Nitrogen trans-
formation during organic waste composting by the Rutgers system and its effect
on pH, EC and maturity of the composting mixtures. Bioresource Technology,
78, 301308.
Saviozzi, A., and Riffaldi, R. 1998. Maturity evaluation of organic waste. Biocycle, 29,
5456.
Saviozzi, A., Riffaldi, R., and Levi-Minzi, R. 1987. Compost maturity by water extract
analysis. In: Compost: Production, quality and use, Bertoldi, M. de, Ferranti,
396 S. Gajalakshmi and S. A. Abbasi

M.P., LHermite, P., and Zucconi, F. (Eds.), Elsevier Applied Science, London,
pp. 359367.
Schloss, P.D., and Walker, L.P. 2000. Measurement of process performance and
variability in inoculated composting reactors using ANOVA and power anal-
ysis.Process Biochemistry, 35, 931942.
Schnurer, J., and Rosswall, T. 1982. Fluorescein diacetate hydrolysis as a measure of
total microbial activity in soil and litter. Applied and Environmental Microbiol-
ogy, 43, 12561261.
Sharma, V.K., Canditelli, M., Fortuna, F., and Cornacchia, G. 1997. Processing of urban
and agro-industrial residues by aerobic composting: Review. Energy Conversion
and Management, 38, 453478.
Shiralipour, A., McConnell, D.B., and Smith, W.H. 1992. Uses and benefits of MSW
compost: A review and an assessment. Biomass Bioenergy, 3, 267279.
Downloaded by [McGill University Library] at 04:56 14 November 2012

Shiralipour, A., Faber, B., and Chrowstowski, M. 1996. Greenhouse broccoli and
lettuce growth using cocomposted biosolid. Compost Science & Utilization, 4,
3843.
Sikora, L.J., Tester, C.F., Taylor, J.M., and Parr, J.F. 1980. Fescue yield response to
sewage sludge compost amendments. Agronomy Journal, 72, 7984.
Singh, M.D.P., and Malik, K.K. 1993. Isolation of a few lignocellulose degrading
fungi.Indian Journal of Microbiology, 33, 265267.
Smars, S., Johnsson, H., Beck-Friis, B., and Kirchmann, H. 2001. An advanced exper-
imental composting reactor for systematic simulation studies. Journal of Agri-
cultural Engineering and Research, 78, 415422.
Smars, S., Gustafsson, L., Beck-Friis, B., and Johnsson, H. 2002. Improvement of
the composting time for household waste during an initial low pH phase by
mesophilic temperature control. Bioresource Technology, 84, 237241.
Smith, W.H. 1996. Utilizing composts in land management to recycle organics. In:
Bertoldi, M. de, Sequi, P., Lemmes, B., and Papi, T. (Eds.), The science of com-
posting, Part 1, Blackie, Glasgow, UK, pp. 413422.
Smith, D.C., and Hughes, J.C. 2004. Changes in maturity indicators during the degra-
dation of organic wastes subjected to simple procedures. Biology and Fertility
of Soils, 39, 280286.
Smith, D.C., and Hughes, J.C. 2001. A simple test to determine cellulolytic activity
as indicator of compost maturity. Communications in Soil Science and Plant
Analyses, 32, 17351749.
Smith, D.C., Beharee, V., and Hughes, J.C. 2001. The effects of composts produced
by a simple composting procedure on the yields of swiss chard (Beta vulgaris
L. var. flavescens) and common bean (Phaseolus vulgaris L. var nanus). Scientia
Horticulturae, 91, 393406.
Smith, M., Friend, D., and Johnson, H. 2007. Composting for the home-
owner, University of Illinois Extension, College of Agricultural, Consumer
and Environmental Sciences, University of Illinois at Urbana-Champaign.
web.extension.uiuc.edu/homecompost/history.html.
Solano, M.L., Iriarte, F., Ciria, P., and Negro, M.J. 2001. Performance characteristics of
three aeration systems in the composting of sheep manure and straw, Journal
of Agricultural Engineering Research, 79, 317329.
Sommer, S.G. 2001. Effect of composting on nutrient loss and nitrogen availability
of cattle deep litter. European Journal of Agronomy, 14, 123133.
Solid Waste Management by Composting 397

Soumare, M., Demeyer, A., Tack, F.M.G., and Verloo, M.G. 2002. Chemical charac-
teristics of Malian and Belgian solid waste composts. Bioresource Technology,
81, 97101.
Soumare, M., Tack, F.M.G., and Verloo, M.G. 2003. Characterisation of Malian and
Belgian solid waste composts with respect to fertility and suitability for land
application. Waste Management, 23, 517522.
Spaggiari, G.C., and Spigoni, G.I. 1986. Transformation of urban sludges mixed with
grape stalks into organic fertilizers. In: Compost: production and use, Bertoldi,
M. De, Ferranti, M.P., Hermite, P.I., and Zucconi, F. (Eds.), Elsevier, New York,
pp. 100107.
Steger, K., Jarvisa, A., Smars, S., and Sundha, I. 2003. Comparison of signature lipid
methods to determine microbial community structure in compost. Journal of
Microbiological Methods, 55, 371382.
Downloaded by [McGill University Library] at 04:56 14 November 2012

Stelmachowski, M., Jastrzebska, M., and Zarzycki, R. 2003. In-vessel composting for
utilizing of municipal sewage-sludge. Applied Energy, 75, 249256.
Stentiford, E.I. 1996. Composting control: principles and practice. In: De Bertoldi,
M., Sequi, P., Lemmes, B., and Papi, T., (Eds.), The science of composting, Part
I, Blackie Academic and Professional, Chapman and Hall, London, pp. 224
252.
Strom, P.F. 1985a. Effect of temperature on bacterial species diversity in thermophilic
solid waste composting. Applied Environmental Microbiology, 50, 899905.
Strom, P.F. 1985b. Identification of thermophilic bacteria in solid waste composting.
Applied Environmental Microbiology, 50, 907913.
Szczeck, M.M. 1999. Suppressiveness of vermicompost against Fusarium wilt
of tomato. Journal of PhytopathologyPhytopathologische Zeitschrift, 147,
155161.
Tabatabai, M.A. 1994. Soil enzymes. In: Weaver, R.W., Angle, J.S., and Bottomley,
P.S (Eds.), Methods of soil analysis. Part IIMicrobiological and biochemical
properties, Book series No. 5, SSSA, Madison, WI, pp. 775833.
Tang, J.C., Kanamori, T., Inoue, Y., Yasuta, T., Yoshida, S., and Katayama, A. 2004.
Changes in the microbial community structure during thermophilic composting
of manure as detected by the quinone profile method. Process Biochemistry, 39,
19992006.
Tate, R.L. III. 1995. Soil enzymes as indicators of ecosystem status. In: Soil microbi-
ology, Tate,R.L. III (Ed.), Wiley, New York, pp. 123146.
Tchobanoglous, G., Theisen, H., and Vigil, S. 1993. Integrated solid waste manage-
ment: Engineering principle and management issues, McGraw-Hill, New York.
Tester, C.F., Sikora, L.J., Taylor, J.M., and Parr, J.F. 1977. Decomposition of sewage
sludge compost in soil: I. Carbon, nitrogen and phosphorus transformations.
Journal of Environmental Quality, 6, 459463.
Theodore, M., and Toribio, J.A. 1995. Suppression of Pythium aphanidermatum
in composts prepared from sugarcane factory residues. Plant and Soil, 177,
219223.
Tiquia, S.M., and Tam, N.F.Y. 2000a. Microbiological and chemical parameters for
compost maturity evaluation of spent pig litter disposed from the pig-on-
litter system. In: Wannan, P.R., and Taylor, B.R. (Eds.), Proceedings of the
International Composting Symposium (ICS 1999), II, CBA Press, Nova Scotia,
Canada, pp. 648669.
398 S. Gajalakshmi and S. A. Abbasi

Tiquia, S.M., and Tam, N.F.Y. 2000b. Co-composting of spent pig litter and sludge
with forced aeration. Bioresource Technology, 72, 17.
Tiquia, S.M., and Tam, N.F.Y. 2002. Characterization and composting of poultry litter
in forced aeration piles.Process Biochemistry, 37, 869880.
Tiquia, S.M., Tam, N.F.Y., and Hodgkiss, I.J. 1996. Microbial activities during compost-
ing of spent pig-manure sawdust litter at different moisture contents. Bioresource
Technology, 55, 201206.
Tiquia, S.M., Wan, J.H.C., and Tam, N.F.Y 2001. Extracellular enzyme profiles during
co-composting of poultry manure and yard trimmings. Process Biochemistry, 36,
813820.
Tiquia, S.M., Richard, T.L., and Honeyman, M.S. 2002a. Carbon, nutrient,and mass
loss during composting. Nutrient Cycling in Agroecosystem, 62, 1524.
Tiquia, S.M., Wan, J.H.C., and Tam, N.F.Y. 2002b. Microbial population dynamics and
Downloaded by [McGill University Library] at 04:56 14 November 2012

enzyme activities during composting. Compost Science and Utilization, 10(2),


150161.
Tiquia, S.M., Wan, J.H.C., and Tam, N.F.Y. 2002c. Dynamics of yard trimmings com-
posting as determined by dehydrogenase activity, ATP content, arginine am-
monification, and nitrification potential. Process Biochemistry, 37, 10571065.
Tisdale, S.L., Nelson, W.L., and Beaton, J.D. 1985. Soil fertility and fertilizers, 4th
edition, Macmillan, New York.
Tognetti, C., Mazzarino, M.J., and Laos, F. 2007. Improving the quality of municipal
organic waste compost.Bioresource Technology, 98, 10671076.
Tuomela, M., Vikman, M., Hatakka, A., and Itavaara, M., 2000. Biodegradation of
lignin in a compost environment: A review. Bioresource Technology, 72(2), 169
183.
U.S. Environmental Protection Agency. 1999. Biosolids generation, use, and disposal
in the United States, U.S. Environmental Protection Agency, Municipal and In-
dustrial Solid Waste Division, Office of Solid Waste, EPA530-R-99-009, U.S. En-
vironmental Protection Agency, Washington, DC.
U.S. Environmental Protection Agency. 2000. Municipal solid wastes in the United
States. 2000 Facts and figures, EPA530-S-02-001, U.S. Environmental Protection
Agency, Washington, DC.
Vaughan, D., and Malcolm, R.E. 1985. Soil organic matter and biological activity. De-
velopment in plant and soil sciences, vol. 16, Dr W. Junk Publishers, Dordrecht,
The Netherlands.
Veeken, A., and Hamelers, B. 2002. Sources of Cd, Cu, Pb and Zn in biowaste. Science
of the Total Environment, 2, 8798.
Veeken, A., de Wilde, W., and Woelders, H. 2004. Characteristics and composting of
poultry litter in forced-aeration piles. Bioresource Technology, 92, 121131.
Vinneras, B., Bjorklund, A., and Jonsson, H., 2003. Thermal composting of faecal
matter as treatment and possible disinfection methodlab scale and pilot scale
studies. Bioresource Technology, 88, 4754.
Vuorinen, A.H. 1999. Phosphatases in horse and chicken manure composts, Compost
Science and Utilization, 7, 4754.
Vuorinen, A.H. 2000. Effect of bulking agent on acid and alkaline phophomo-
noesterase and D-glucosidase activities during manure composting. Bioresource
Technology, 75, 113138.
Solid Waste Management by Composting 399

Waksman, S.A., Umbreit, W.W., and Cordon, T.C. 1939a. Thermophilic actinomycetes
and fungi in soils and in composts. Soil Science, 47, 3761.
Waksman, S.A., Cordon, T.C., and Hulpoi, N. 1939b. Influence of temperature upon
the microbiological population and decomposition processes in composts of
stable manure, Soil Science, 47, 83114.
Wang, P., Changa, C.M., Watson, M.E., Dick, W.A., Chen, Y., and Hoitink, H.A.J. 2004.
Maturity indices for composted dairy and pig manures.Soil Biology & Biochem-
istry, 36, 767776.
Wei, Y., and Liu, Y. 2005. Effects of sewage sludge compost application on crops
and cropland in a 3-year field study.Chemsphere, 59, 12571265.
Weppen, P. 2001. Process calorimetry on composting of municipal organic wastes,
Biomass and Bioenergy, 21, 289299.
Witter, E., and Lopez-Real, J.M. 1987. The potential of sewage sludge and compost-
Downloaded by [McGill University Library] at 04:56 14 November 2012

ing in a nitrogen recycling strategy for agriculture. Biology, Agriculture and


Horticulture, 5, 123.
Witter, E., and Lopez-Real, J.M. 1988. Nitrogen losses during the composting of
sewage sludge, and the effectiveness of clay soil, zeolite, and compost in ad-
sorbing the volatilized ammonia. Biological Wastes, 23, 279294.
Wong, J.W.C., and Fang, M. 2000. Effects of lime addition on sewage sludge com-
posting process. Water Research, 34, 36913698.
Wong, J.W.C., Mak, K.F., Chan, N.W., Lam, A., Fang, M., Zhou, L.X., Wu, Q.T., and
Liao, X.D. 2001. Co-composting of soyabean residues and leaves in Hong Kong.
Bioresource Technology, 76, 99106.
Wright, M. 1998. Home composting: Real waste minimisation or just feel good fac-
tor? Wastes Management, The Journal of the Institute of Wastes Management,
September, 2728.
Wu, L., Ma, L.Q., and Martinez, G.A. 2000. Comparison of methods for evaluating
stability and maturity of biosolids compost. Journal of Environmental Quality,
29, 424429.
Ylimaki, A., Toiviainen, A., Kallio, H., and Tikanamaki, E. 1983. Survival of some plant
pathogens during industrial-scale composting of wastes from a food processing
plant. Annals Agricolae Fenniae, 22, 7785.
Yuen, G.Y., and Raabe, R.D., 1984. Effects of small-scale aerobic composting on
survival of some fungal plant pathogens. Plant Disease, 68, 134136.
Zhang, W., Han, D.Y., Dick, W.A., Davis, K.R., and Hoitink, H.A.J. 1998. Compost
and compost water extract-induced systematic acquired resistance in cucumber
and Arabidopsis. Phytopathology, 88, 450455.
Zinati, G.M. 2005. Compost in the 20th century: A tool to control plant diseases in
nursery and vegetable crops. HortTechnology, 15(1), 1520.
Zorpas, A.A., Arapoglou, D., and Panagiotis, K. 2003. Waste paper and clinop-
tilolite as a bulking material with dewatered anaerobically stabilized primary
sewage sludge (DASPSS) for compost production. Waste Management, 23, 27
35.
Zorpas, A.A., Vassilis, I., Loizidou, M., and Grigoropoulou, H. 2002. Particle
size effects on uptake of heavy metals from sewage sludge compost using
natural zeolite clinoptilolite. Journal of Colloid and Interface Science, 250,
14.
400 S. Gajalakshmi and S. A. Abbasi

Zorpas, A.A., Kapetanois, E., Zorpas, G.A.., Karlis, P., Vlyssides, A., Haralambous, I.,
and Loizidou, M. 2000a. Compost produced from organic fraction of municipal
solid waste, primary stabilized sewage sludge and natural zeolite. Journal of
Hazardous Materials, B77, 149159.
Zorpas, A.A., Conatantinides, T., Vlyssides, A., Haralambous, I., and Loizidou, M.
2000b. Heavy metal uptake by natural zeolite and metals partitioning in sewage
sludge compost. Bioresource Technology, 72, 113119.
Zucconi, F., and de Bertoldi, M. 1987. Compost specifications for the production and
characterization of compost from municipal solid wastes. In: Compost: Produc-
tion, quality and use, M. de Bertoldi, M.P. Ferranti, P.L. Hermite, and F.Zucconi
(Eds.), Elsevier Applied Science, London, pp. 3050.
Zucconi, F., Monaco, A., Forte, M., and de Bertoldi, M. 1985. Phytotoxins during the
stabilization of organic matter. In: Composting of agricultural and other waste,
Downloaded by [McGill University Library] at 04:56 14 November 2012

J.K.R. Gasser (Ed.), Elsevier Applied Science, London, pp. 7385.


Zucconi, F., Pera, M.A., Forte, M., and Bertoldi, M de. 1981a. Evaluating toxicity of
immature compost. Biocycle, 22, 5457.
Zucconi, F., Forte, M., Monaco, A., and Bertoldi, M de. 1981b. Biological evaluation
of compost maturity. Biocycle , 22, 2729.

You might also like