You are on page 1of 73

1

Introduction
ZUENG-SANG CHEN
and to produce good quality of composts. The value of organic wastes as soil conditioners can be estimated in a number of ways, depending on the ultimate objectives of their use. Inoculants of mixed cultures of beneficial microorganisms have considerable potential for controlling the soil microbiological equilibrium, thus providing a more favorable environment for plant growth and protection. Organic matter content is usually used as an index of soil fertility. In their general review of the effects of organic matter, many researchers indicated that it influences the soil in three ways, such as, physically, chemically, and biologically. The fibrous portion of organic matter plays an important role in improving soil physical properties. It promotes soil aggregation and improves permeability and aeration of clayey soils. Its high moistureabsorbing power and high carbon for growth of microbial mycelia may help in the granulation of sandy soils to improve their nutrient- and water-holding capacity. Organic matter accounts for at least half the cation exchange capacity (CEC) of soils. Thus, it is very important not only in retaining nutrients from fertilizers applied but also in increasing the buffering capacity of soils, enabling crops to better cope with such stresses as soil acidity and nutrient excess. It helps increase availability of many nutrient elements. By itself, organic matter is a source of nitrogen (N), phosphorus (P), potassium (K), sulfur (S), and other major and secondary nutrient elements. The management of soil organic matter is important in maintaining soil productivity, reducing soil erosion, keeping the soil structure and nutrient pools, and controlling the water balance for sustainable soil management system. This book discusses composting, which has been found to be an efficient means of managing soil organic matter.

FOR MANY CENTURIES, the key to a permanent and sustainable agriculture had been the regular and extensive recycling of organic materials on soils as soil conditioners. Such materials included animal manure, green manure, crop residues, wood ashes, tree leaves, weeds from canals, wild grasses, urban sewage, and street refuse. Many of these materials were composted to destroy weed seeds and potential human and plant pathogens; to enhance their nutrient availability; and to facilitate their storage, transport, and application to land. These practices enabled farmers to maintain their soil quality and to maximize their crop production with negligible soil erosion and nutrient runoff. However, agriculture in the Asian and Pacific region has changed considerably, and farmers are now facing problems of rapid soil degradation and soil productivity loss. It is readily apparent that alternative agricultural practices and the ultimate goal of a long-term sustainable agriculture depend largely on regular additions of various organic amendments to soils. The quality and acceptability of many organic wastes, from both on-farm and off-farm sources, can be greatly enhanced through composting. Composting is allowing organic materials to decompose under more or less controlled conditions to produce a product that can be used as a fertilizer and/ or soil conditioner. In more recent technology of composting, forced aeration, mechanical shredding, mixing, grinding, drying, and even inoculation with microbial decomposers have been introduced. Composting is basically a microbial bio-oxidative process. Its purpose is to change the properties of an organic material or a mixture of organic biomass into a material that is safe to apply to crops as fertilizer or soil conditioner. The critical factors that affect composting and their interrelationships must be thoroughly understood to ensure optimum composting conditions

2
2.1 Introduction

Benefits and drawbacks of composting


plant production, help save money, reduce the use of chemical fertilizers, and conserve natural resources. Compost provides a stable organic matter that improves the physical, chemical, and biological properties of soils, thereby enhancing soil quality and crop production. When correctly applied, compost has the following beneficial effects on soil properties, thus creating suitable conditions for root development and consequently promoting higher yield and higher quality of crops (Figs. 1-3):

JEN-HSHUAN CHEN and JENG-TZUNG WU

Composting is the natural process of decomposing and recycling organic materials into a humus-rich soil amendment by the successive action of bacteria, fungi, actinomycetes, or earthworms. Many common materials can be composted on-site, including food wastes, leaves, grass clippings, plant trimmings, straw, shredded paper, animal manure, and municipal solid wastes. The final product is a stable dark-brown or black humus material with an earthy smell. Like other recycling efforts, composting has many benefits to agriculture, the environment, the economy, and the society. However, composts have some disadvantages to agriculture and the environment if they were processed or used under incorrect conditions. This chapter provides a brief introduction to the benefits and drawbacks of using composts.

2.3.1 Improves the physical properties of soils


!

2.2 Potential users and uses of finished compost


Prepared composts can be used by the following users: ! Agricultural and residential user group: used as soil amendment, fertilizer supplement, top dressing for pasture and hay crop maintenance, fertilizer substitute, mulch for fruit trees. ! Commercial user group: used as soil amendment for turf establishment, landscape planting and beds, potting mix component, peat substitute, topsoil substitute, mulch, fertilizer supplement. ! Municipal user group: used as landfill cover materials, topsoil for road and construction work, soil amendment, mulch for landscape planting.

2.3 The benefits of using composts to agriculture


Compost has been considered as a valuable soil amendment for centuries. Most people are aware that using composts is an effective way to increase healthy
!

Reduces the soil bulk density and improves the soil structure directly by loosening heavy soils with organic matter, and indirectly by means of aggregate-stabilizing humus contained in composts. Incorporating composts into compacted soils improves root penetration and turf establishment. Increases the water-holding capacity of the soil directly by binding water to organic matter, and indirectly by improving the soil structure, thus improving the absorption and movement of water into the soil. Therefore, water requirement and irrigation will be reduced. Protects the surface soil from water and wind erosion by reducing the soil-dispersion action of beating raindrops, increasing infiltration, reducing water runoff, and increasing surface wetness. Preventing erosion is essential for protecting waterways and maintaining the quality and productivity of the soil. Helps bind the soil particles into crumbs by the fungi or actinomycetes mycelia contained in the compost and stimulated in the soil by its application, generally increasing the stability of the soil against wind and water erosion. Improves soil aeration and thus supplies enough oxygen to the roots and escapes excess carbon dioxide from the root space.

Fig. 1.

Spinach grown in the field applied with compost

Fig. 2.

Maize grown in the field applied with compost

Fig. 3.

Lettuce grown in the field applied with compost

Increases the soil temperature directly by its dark color, which increases heat absorption by the soil, and indirectly by the improved soil structure. Helps moderate soil temperature and prevents rapid fluctuations of soil temperature, hence, providing a better environment for root growth. This is especially true of compost used as a surface mulch.
!

2.3.2 Enhances the chemical properties of soils


!

1) successful competition for nutrients by beneficial microorganisms; 2) antibiotic production by beneficial microorganisms; 3) successful predation against pathogens by beneficial microorganisms; 4) activation of disease-resistant genes in plants by composts; and 5) high temperatures that result from composting kill pathogens. Reduces and kills weed seeds by a combination of factors including the heat of the compost pile, rotting, and premature germination.

Enables soils to hold more plant nutrients and increases the cation exchange capacity (CEC), anion exchange capacity (AEC), and buffering capacity of soils for longer periods of time after composts are applied to soils. This is important mainly for soils containing little clay and organic matter. Builds up nutrients in the soil. Composts contain the major nutrients required by all plants [N,P,K, calcium (Ca), magnesium(Mg), and S] plus essential micronutrients or trace elements, such as copper (Cu), zinc (Zn), iron (Fe), manganese (Mn), boron (B), and molybdenum (Mb). The nutrients from mature composts are released to the plants slowly and steadily. The benefits will last for more than one season. Stabilizes the volatile nitrogen of raw materials into large protein particles during composting, thereby reducing N losses. Provides active agents, such as growth substances, which may be beneficial mainly to germinating plants. Adds organic matter and humus to regenerate poor soils. Buffers the soil against rapid changes due to acidity, alkalinity, salinity, pesticides, and toxic heavy metals.

2.4 Benefits of using composts to the environment


2.4.1 Pollution remediation
!

Absorbs odors and degrades volatile organic compounds. Binds heavy metals and prevents them from migrating to water resources or being absorbed by plants. Degrades and, in some cases, completely eliminates wood preservatives, petroleum products, pesticides, and both chlorinated and nonchlorinated hydrocarbons in contaminated soils.

2.4.2 Pollution prevention


!

2.3.3 Improves the biological properties of soils


!

! !

Supplies food and encourages the growth of beneficial microorganisms and earthworms. Helps suppress certain plant diseases, soilborne diseases, and parasites. Research has shown that composts can help control plant diseases (e.g. Pythium root rot, Rhizoctonia root rot, chili wilt, and parasitic nematode) and reduce crop losses. A major California fruit and vegetable grower was able to cut pesticide use by 80% after three years of compost applications as part of an organic matter management system. Research has also indicated that some composts, particularly those prepared from tree barks, release chemicals that inhibit some plant pathogens (Hoitink and Fahy 1986). Disease control with compost has been attributed to four possible mechanisms:

Avoids methane production and leachate formation in landfills by diverting organics for composting. Absorbs odors and degrades volatile organic compounds. Prevents pollutants in storm water runoff from reaching water resources, and protects groundwater quality. Prevents erosion and turf loss on roadsides, hillsides, playing fields, and golf courses. Minimizes odors from agricultural areas. Composting raw manure can minimize any potential environmental or nuisance problems. Raw manure is one of the primary culprits in the pollution of waterways, and odor from farms is considered an increasing problem in the rural areas.

2.5 Economic and social benefits of composting


The economic and social benefits of composting include the following: ! Brings higher prices for organically grown crops. ! Composting can offer several potential economic benefits to communities: - Extends current landfill longevity and delays the construction of a more expensive replacement landfill or incinerator. - Reduces or avoids landfill or combustor

tipping fees, and reduces waste disposal fees and long-distance transportation costs. - Offers environmental benefits from reduced landfill and combustion use. - Creates new jobs for citizens. - Produces marketable products and a less-cost alternative to standard landfill cover, artificial soil amendments, and conventional bioremediation techniques. Provides a source of plant nutrients and improves soil fertility; results in significant cost savings by reducing the need for water, pesticides, fungicides, herbicides, and nematodes. Used as an alternative to natural topsoil in new construction, landscape renovations, and container gardens. Using composts in these types of applications is not only less expensive than purchasing topsoil, but it can also often produce better results when establishing a healthy vegetative cover. Used as mulch for trees, orchards, landscapes, lawns, gardens, and makes an excellent potting mix. Placed over the roots of plants, compost

mulch conserves water and stabilizes soil temperatures. In addition, it keeps plants healthy by controlling weeds, providing a slow release of nutrients, and preventing soil loss through erosion.

2.6 Drawbacks of using composts


Agricultural use of composts remains low for several reasons: ! The product is weighty and bulky, making it expensive to transport. ! The nutrient value of compost is low compared with that of chemical fertilizers, and the rate of nutrient release is slow so that it cannot usually meet the nutrient requirement of crops in a short time, thus resulting in some nutrient deficiency (Figs. 4 and 5). ! The nutrient composition of compost is highly variable compared to chemical fertilizers. ! Agricultural users might have concerns regarding potential levels of heavy metals and other possible contaminants in compost,

Fig. 4.

Mg deficiency induced by applying compost with low Mg content.

Fig. 5.

Ca deficiency of cabbage grown in strongly acidic soil applied with compost.

particularly mixed municipal solid wastes. The potential for contamination becomes an important issue when compost is used on food crops. Long-term and/or heavy application of composts to agricultural soils has been found to result in salt, nutrient, or heavy metal accumulation and may adversely affect plant growth, soil organisms, water quality, and animal and human health (Figs. 6 and 7).

2.7 Why are so many farmers in Asia buying compost instead of making it by themselves?
Whatever composting system is used, farmers must have adequate land area and suitable equipment to manage a composting operation, and stable sources of raw materials for composting must be available. In addition, it usually takes a long time perhaps years to produce a stable compost product. Therefore, the requirements of raw materials, space, and equipment as well as the length of time required for composting usually discourage farmers from making compost by themselves.

A farmer's composting system costs include the annual fixed and variable costs attributable to the system. Capital investments include all composting system structures and equipment. Variable costs include labor, fuel, electricity, and maintenance charges. It is less economical and more expensive for farmers to make compost by themselves than to buy it from the market. A successful composting operation should have appropriate background in and techniques for producing high-quality composts without creating odor and other environmental problems. Some negative effects on agriculture and the environment have been found when immature or low-quality composts, usually produced under an inadequate composting process, were added to the soil (Fig. 8). It is better to use highquality composts bought from the market than to use immature ones produced by farmers if their knowledge of and/or techniques for composting are not proper. If not carefully and properly controlled, the composting process can create a number of environmental concerns such as air and water pollution, odor, noise, vectors, fires, and litter that can be a cause of complaints from neighbors or nearby residents. In addition, the potential worker's health and

Fig. 6.

Poor growth of tomato in soil with high EC value due to heavy application of animal compost.

Fig. 7.

Poor seed generation of cabbage in soil with salt accumulation due to heavy application of animal compost.
6

Fig. 8. Use of immature compost generally reduces vegetable production.

safety problems must be dealt with and solved. It may be too difficult for farmers to manage these problems.

2.8 References
Black, R. J and G. L. Miller. 1998. Benefits of using compost and mulch in Florida roadside planting. ENH-126. Environmental Horticulture Department, Florida Cooperative Extension Service, Institute of Food and Agricultural Sciences, University of Florida. Finck, A. 1982. Fertilizers and fertilization: Introduction and practical guide to crop fertilization. Verlag Chemie. GimbH, Weinheim. Federal Republic of Germany.

Hoitink, H. A. J. and P. C. Fahy. 1986. Basis for the control of plant pathogens with compost. Annual Review of Phytopathology, 24: 93-114. Mark Risse and Britt Faucette. 2000. Food waste composting, institutional and industrial application. Cooperative Extension Service, the University of Georgia College of Agricultural and Environmental Sciences. http://www.ces.uga.edu/pubcd/ B1189.htm U.S. Environmental Protection Agency (USEPA). 1993. Markets for compost. EPA1530-SW-90-073b. Washington, D.C.: Office of Policy, Planning and Evaluation; Office of Solid Waste and Emergency Response.

3
3.1 Introduction

What happens during composting?


as worms, mites, snails, beetles, centipedes, and millipedes are mainly the physical decomposers. Bacteria and worms are the powerhouse of chemical decomposers and physical decomposers, respectively. Microbes and invertebrates carry out decomposition of organic litter by utilizing its carbon and nitrogen contents as the energy source with oxygen and water, resulting in the production of carbon dioxide, heat, water, and soil-enriching compost (Fig. 1). Most organisms preferred for composting are aerobic (requiring oxygen) as they provide rapid and complete composting (Figs. 2 and 3). Other organisms can operate without oxygen (anaerobic conditions), and this process is sometimes called fermentation and usually occurs more slowly. They utilize nitrate, sulfate, carbonate, and ferric ions to oxidize organic compounds (Fig. 4). However, the greatest disadvantage of anaerobic process is the offensive odors produced during the process. It also produces organic acids, alcohols, methane and other gases, which may be harmful to the plants.

CHIU-CHUNG YOUNG, P.D. REKHA, and A.B. ARUN

Composting is the science of converting organic matter to useful products by the action of various organisms. Decomposition as a process occurs in nature at various levels. To attain the goal of having quality end products, various modifications have been applied to this natural process with a careful monitoring of the process. Composting is associated with the reclamation, recycling, treatment, and disposal of wastes. Reclamation and recycling are means of saving and reusing natural resources. Disposal has become a less desirable option because of environmental concerns. The composting process mainly involves a battery of actions carried out by the interplay of various organisms that form a web of life. Composting is generally defined as the biological oxidative decomposition of organic constituents in wastes of almost any nature under controlled conditions (Sharma et al. 1997). In this process, the organic substances are reduced from large volumes of rapidly decomposable materials to small volumes that continue to decompose slowly. The process brings the ratio of carbon to other elements into a balance, thus providing nutrients to plants in the absorbable state (Fig. 1). To understand the science of composting, a basic understanding of the various organisms involved is necessary. Based on their functions, these organisms have been classified as first-order consumers, which feed directly on the dead plant or animal materials; second-order consumers, which feed primarily on the first-order consumers or on the produce of these consumers; and third-order consumers, which feed on the second-order consumers. This system keeps different populations in check and maintains a healthy and balanced system. Further, the vast array of organisms found in the compost pile can be classified based on their functions as chemical and physical decomposers. Microscopic organisms such as bacteria, fungi, actinomycetes, and protozoa are the chemical decomposers, while larger organisms such

3.2 Composting principles


As stated earlier, composting is essentially a mass of interdependent bioprocesses carried out by an array of micro- and macroorganisms resulting in the decomposition of organic matter. Soil microbes oxidize organic compounds, and release essential minerals such as nitrogen, phosphorus, and sulfur, which plants need. This oxidation process is also called respiration, wherein carbon dioxide, water, and energy are produced followed by the release of minerals that are essential for the growth of plant and other soil organisms (Figs. 2-4). Carbon dioxide escapes to the atmosphere. The breakdown of organic matter is a dynamic process accomplished by a succession of microorganisms, with each group reaching its peak population when conditions have become optimum for its activity. Mesofauna such as mites, sow bugs, worms, springtails, ants, nematodes, and beetles do

Heat Water Vapor RAW MATERIALS Minerals Water Fresh organic matter CO2 Gas

Compost Pile or Windrow Microorganisms

Finished Compost

O2

Fig. 1.

Generalized representation of the composting process.

Plant residues Organic wastes

CO2

O2

O2 reduction Easily decomposable Organic matter O2

#
Humification (Polymerization)

# " !

#
Respiration

!
Slowly decomposable

"

CO2

H2O

#
Microbial Biomass Energy

Fig. 2.

Organic matter decomposition pathways for aerobic respiration. (Adapted from Reddy et al. 1986, with modifications)

Plant residues Organic wastes

NO-3 NO-3

Zone-I Aerobic Zone II & III Anaerobic Easily decomposable

Nitrification

! " NH4
Carbohydrates Fatty acids Organic acids NO
4

"
#

#
NO-2 N2

" Amino acids

#
Organic matter

" Dissimilatory

! " N2O

nitrate & nitrite reduction

Slowly decomposable Zone II- NO3 - reduction Mn+ reduction Zone III- Fe3+ reduction

!
CO2 Energy H2O Mn4+Reduction

Microbial

!!
Fe2O3

! ! ! !
!

" Fe3+Reduction ! #
Fe 2+

Mn O2

! #
Mn2+

Fig. 3.

Organic matter decomposition pathways for facultative respiration. (Adapted from Reddy et al. 1986, with modifications)

Plant residues Organic wastes Zone I Aerobic SO42-

CO2

$
SO 42-

Sulfide oxidation

Methane oxidation

Zone II & III Anaerobic

Organic matter

Easily decomposable

"

!
Slowly decomposable

"
#
#

Amino acids Carbohydrates Fatty acids Organic acids

Sulfate

" HS 2
Energy CO2 CH 4

#
H 2S

#
Acid fermentation

Microbial biomass Zone II- Sulfate reduction Zone III- Methane formation

" #
FeS

Short Chain Fatty Acids CO2, H2

Methane formation

Fig. 4.

Organic matter decomposition pathways for anaerobic respiration. (Adapted from Reddy et al. 1986, with modifications)
10

most of the initial mechanical breakdown of the materials into smaller particles. Mesophilic bacteria, fungi, actinomycetes, and protozoa (microorganisms that function at temperatures between 10o and 45oC) initiate the composting process. As temperature increases as a result of oxidation of carbon compounds, thermophiles (microorganisms that function at temperatures between 45o and 70oC) take over. Temperature in a compost pile typically follows a pattern of rapid increase to 49o-60oC within 24-72 hours of pile formation and is maintained for several weeks. This is the active phase of composting. The process involves the degradation of easily degradable compounds under aerobic conditions. The increased temperature kills pathogens, weed seeds, and phytotoxins. During this phase, oxygen must be supplied by either mixing, forced aeration, or turning the compost pile. As the active composting phase subsides, temperature gradually declines to around 38oC. Mesophilic organisms recolonize and the curing phase begins. During curing, organic materials continue to decompose and are converted to biologically stable humic substances (mature or finished composts). The maturing phase requires minimum oxygen and the biological processes/ activities become very slow. Considering the various stages of compost formation, the bio-oxidation phase used for the degradation of organic substances can be identified. In contrast, the synthesis phase of humic substances, started during the first phase of composting, develops and finally will be completed in the mature phase of the compost. Curing is a very critical stage and should be kept for 1-4 months. Compost is considered finished or stable after the temperature of the pile core reaches near-ambient levels. The processes are accomplished by different phases and are discussed below. ! Initial phase, during which readily degradable components are decomposed; ! Thermophilic phase, during which cellulose and similar materials are degraded by the high biooxidative activity of microorganisms; ! Maturation and stabilization phase. The processes can also be explained in terms of two well-defined phases, namely, mineralization and humification. The former is an intensive process involving the degradation of readily fermentable organic substances like carbohydrates, amino acids, proteins, and lipids. The degradation involves high microbial activities and generates heat, carbon dioxide, and water in addition to a partially transformed and stable organic residue. When the assimilable organic fraction is utilized, some of the cells undergo decay by auto-oxidation, which provides energy for the remaining cells (Fig. 5). The transformation process of the organic substances is completed in the second phase under less oxidative conditions, thus allowing the formation of the humic-character substance and eliminating the dense toxic compost, eventually formed during the first

phase. The humification phase is carried out by specific microbes, which synthesize the complex polymers that create the energy substratum for future microbial activities.

3.3 Conditions and components of the process


The following components are necessary for the process to progress smoothly and to obtain good quality compost as the end product. Under controlled conditions, natural decomposition progresses faster and yields a quality product. The rate of composting, like the rate of plant or animal growth, can be affected by many factors. The key factors are nutrient balance (C:N ratio), moisture content, temperature, and aeration. In addition, the organic substance selected for composting should be free of any toxic compounds such as detergents, surfactants, phenolics, and pharmaceutics, which will pose health risks either directly or through their metabolic, degraded products. It is therefore very important to minimize such materials in the input source.

3.3.1 Temperature
Temperature is an important parameter affecting microbial activity, and variations in temperature affect the various phases of composting (Epstein 1997; McKinley et al. 1985). Temperature is produced during the composting process, resulting from the breakdown of organic materials by microbes. The organisms in composting systems can be divided into three classes: cryophiles or psychrophiles (0o-25oC); mesophiles (25o-45oC); and thermophiles (>45oC). Cryophiles are found only during winter composting. Mesophiles, in association with thermophiles, generally predominate commercial composting systems. The temperature can range from near freezing to 70oC. Starting at ambient temperature when the components are mixed, the compost can reach 40o-60oC in less than two days depending on the composition and environmental conditions. Hence, heat is generated from within the compost medium, and applying external temperature is not necessary unless ambient temperature is far below freezing. Temperature is also a good indicator of the various stages of the composting process. The process is divided into four phases based on temperature. The first stage is the mesophilic stage, where mesophilic organisms generate large quantities of metabolic heat and energy due to availability of abundant nutrients, but gradually this will pave the way for the dominance of thermophiles. With depletion of food sources, overall microbial activity decreases and temperature falls to ambient, leading to the second mesophilic stage, where microbial growth will be slower as readily available food is consumed. Finally, compost material enters the maturation stage, which might take some months.

11

3.3.2 Composting microorganisms


The degradation of organic wastes is a natural process and begins almost as soon as the wastes are generated. Composting is a means of controlling and accelerating the decomposition process. Compost is normally populated by three general categories of microorganisms, namely, bacteria, actinomycetes, and fungi (Table 1). Although considered bacteria, actinomycetes are effectively intermediate between bacteria and fungi because they look similar to fungi and have similar nutritional preferences and growth habits. They tend to be more commonly found in the later stages of the composting process, and are generally thought to follow the thermophilic bacteria in succession. They, in turn, are followed predominantly by fungi during the last stages of the composting process (Fig. 5). The typical composting process is explained below. Initial degradation involves endophytes and epiphytes associated with the plant, and microbes from the air and within raw materials. These are rapidly followed by heat-tolerant bacteria and yeasts, which probably come from airborne sources. The bacteria and other organisms take up soluble sugars and amino acids from the plant materials first. Starch is then broken down and absorbed. Subsequently, pectin and cellulose are digested. Finally, fungi digest lignin and the waxes. The most common complex carbohydrate available in the environment is cellulose. In the absence of glucose, fungi specifically target the

breakdown of the cellulose in their environment, and do not waste energy on the unnecessary formation of enzymes for degradation of molecules that may not be present. Cellulose is a polymer of glucose, which is digested by a variety of enzymes. In simple terms, the enzymes may either cleave glucose molecules from the end of the polymer (exocellulase), or fragment the cellulose polymer into smaller molecules by internal digestion (endocellulase). Cellulases are especially common in soil- and plant-inhabiting fungi. Many fungi in the Ascomycotina and Basidiomycotina are able to digest cellulose. The necessary enzymes are less common in members of the Zygomycotina. Lignin is commonly found in plants. Lignin is a polymer of phenyl-propanoid units (C6C3), with a variety of carbon-carbon and carbon-oxygen linkages resulting in a complex chemistry and structure. Various enzymes are needed to completely degrade lignin. These can be classified into two functional groups: lignin peroxidases and manganese peroxidases. The enzymes are only induced in the absence of readily available nutrients. Thus, degradation of lignin is delayed and slow. Lignin molecules are commonly found associated with cellulose. Fungi with ligninases usually digest cellulose. Fungi with ligninolytic potential are more common in the Basidiomycotina than in any other group. Generally, lignin is broken down slowly because only few uncommon fungi are able to degrade the organic material when the environment is highly competitive.

COMPOSTING MATERIALS ORGANISMS OF UPPER TROPHIC LEVEL

Exhaust air Death of microorganisms

! "
White rot fungi

HUMIC SUBSTRATES

!
Mineral substances

Bacteria

Actinomycetes

Brown rot fungi

"

!
Metabolism of microorganisms

!
Metabolic products Water phase

"

CO2

" "

Water vapor

Carbohydrates, lipids, and proteins

Hemi cellulose

Cellulose, pectin

Lignin

Fig. 5. Generalized food-web scheme of the composting ecosystem. (Adapted from Kaiser 1996, with modifications)
12

Table 1. List of major microorganisms present in compost.


Actinomycetes Actinobifida ahromogena Microbispora bispora Micorpolyspora faeni Nocardia sp. Pseudocardia thermophilia Streptomyces rectus S. thermofuscus S. theromviolaceus S. thermovulgaris S. violaceus-ruber Thermoactinomyces sacchari T. vulgaris Thermomonospora curvata T. viridis Fungi Aspergillus fumigatus Humicola grisea H. insolens H. lanuginosa Malbranchea pulchella Myriococcum thermophilum Paecilomyces variotti Papulospora thermophilia Scytalidium thermophilim Sporotrichum thermophile Bacteria Alcaligenes faecalis Bacillus brevis B. circulans complex B. coagulans type A B. coagulans type B B. licheniformis B. megaterium B. pumilus B. sphaericus B. stearothermphilus B. subtilis Clostridium thermocellum Escherichia coli Flavobacterium sp. Pseudomonas sp. Serratia sp. Thermus sp.

Source: Palmisano, A.C. and Bartaz, M.A. (1996) Microbiology of solid waste, pp. 125-127. CRC Press, Inc. 2000. Corporate Bld. N.W. Boca Raton. FL 33431 USA.

3.4 Chemical processes


3.4.1 Nutrient balancecarbon/nitrogen ratio
Nutrient balance is very much dependent on the type of feed materials being processed. Carbon provides the preliminary energy source and nitrogen quantity determines the microbial population growth. Hence, maintaining the correct C:N ratio is important to obtain good quality compost. Bacteria, actinomycetes, and fungi require carbon and nitrogen for growth. These microbes use 30 parts of carbon to 1 part of nitrogen. Composting is usually successful when the mixture of organic materials consists of 20-40 parts of carbon to 1 part of nitrogen. However, as the ratio exceeds 30, the rate of composting decreases. Further, as the ratio decreases below 25, excess nitrogen is converted to ammonia. This is released into the atmosphere and results in undesirable odor (Pace et al. 1995). During bioconversion of the materials, concentration of carbon will be reduced while that of nitrogen will be increased, resulting in the reduction of C:N ratio at the end of the composting process. The reduction can be attributed to the loss in total dry mass due to losses of C as CO2 (Hamoda et al. 1998). Ammonium-N (NH4-N) and nitrate-N (NO3-N) will also undergo some changes. NH3 levels were increasing in the initial stages but declining towards the end (Liao et al. 1995). In several instances, NO3 concentrations were less during the initial phases but gradually increased towards the end (Neto et al. 1987) and, in some instances, remained unchanged (Palmisano et al. 1993). Maintaining NH3 concentration is important to avoid excess nitrogen losses and production of bad

odor. Maintaining C:N ratio after composting is also important to determine the value of finished compost as soil amendment for crops. The final C:N ratio of 15 to 20 will be expected and the value of more than 20 might have a negative impact and will damage the crop and seed germination. The value of 10 has been suggested as ideal.

3.4.2 Phosphorus
Levels of P along with N and K will be important to determine the quality of compost, as P is also one of the essential nutrients for plant growth. A C:P ratio of 100 to 200 is desirable (Howe and Coker 1992). Phosphorus is not lost by volatilization or lixiviation during the composting process, but P concentration might increase as composting proceeds (Warman and Termeer 1996).

3.4.3 Sulfur
Presence of S in sufficient quantities can lead to the production of volatile, odorous compounds (Day et al. 1998). The major sources of S are two amino acids, namely, cysteine and methionine. Under well-aerated conditions, the sulfides are oxidized to sulfates, but under anaerobic conditions, they are converted to volatile organic sulfides or to H2S, leading to a bad odor. Some compounds like carbon disulfide, carbonyl sulfide, methyl mercaptum, diethyl sulfide, dimethyl sulfide, and dimethyl disulfide might also lead to bad odors.

13

3.5 Physical processes


3.5.1 Moisture content
Moisture in compost comes from either the initial feedstock or the metabolic water produced by microbial action (0.6-0.8 g/g), but, during aerobic composting, 1 g of organic matter releases about 25 kJ of heat energy, which is enough to vaporize 10.2 g of water (Finstein et al. 1986). This will be further coupled with losses due to aeration (Naylor 1996), resulting in water loss during composting. Hence, moisture is an important factor to be controlled during composting as it influences the structural and thermal properties of the material, as well as the rate of biodegradation and metabolic process of the microbes. The moisture content of compost should be 60% after organic wastes have been mixed. Depending on the components of the mixture, initial moisture content can range from 55%-70%. However, if this exceeds 60%, the structural strength of the compost deteriorates, oxygen movement is inhibited, and the process tends to be anaerobic. Low C:N ratio materials (e.g., meat wastes) putrefy when anaerobic, while high ratio materials ferment. Both these processes produce odor, leach nutrients, increase pathogens, and block air passages in the pile, hence they must be avoided. As the moisture content decreases below 50%, the rate of decomposition decreases rapidly. Excessive moisture in the compost will prevent O2 diffusion to the organisms. Reduction in the moisture content below 30%-35% must be avoided since it causes a marked reduction in the microbiological activity. Moisture can be controlled either directly by adding water or indirectly by changing the operating temperature or the aeration regime. Feedstock with different moistureholding capacities can be blended to achieve an ideal moisture content.

3.5.3 Particle size


Decomposition and microbial activity will be rapid near the surfaces as oxygen diffusion is very high. Small particles have more surface area and can degrade more quickly. Haug (1993) suggested that, for particles larger than 1 mm, oxygen diffusion would limit in the central part of the particles, thus the interior parts of the larger particles will be anaerobic with a slower rate of decomposition. Particle size also affects moisture retention as well as free air space and porosity of the compost mixture (Nalyor 1996). Smaller particle size results in reduced air space and less porosity. Aerobic decomposition increases with smaller particle size; however, smaller particle size reduces the effectiveness of the oxygen supply. By turning regularly, this problem can be solved. The preferable size is 3 mm-50 mm diameter. Compaction can also influence the free air space. By employing grinding and sieving equipment, such problems can be avoided. At the end of the process, the bulk density of the compost would be expected to increase due to breakdown in the particle size of the material, resulting in more compact compost. But in some composting systems, where water evaporation and water loss are high, the bulk density might decrease as the materials will be dried during the composting period (Day et al. 1998).

3.6 Biological changes


3.6.1 Fate of pathogens during composting
In addition to the already discussed microbes, there will be many human, animal, and plant pathogens. It is not only the heat of the compost that destroys all these pathogens; it is a combination of factors including: ! competition for food from compost microorganisms; ! inhibition and antagonism by compost microorganisms; ! consumption by compost microorganisms; ! biological heat generated by compost microorganisms; and ! antibiotics produced by compost microorganisms. There is no doubt that the heat produced by thermophilic bacteria kills pathogenic microorganisms, viruses, bacteria, protozoa, worms, and eggs that may inhabit humans. A temperature of 50C (122F), if maintained for 24 h, is sufficient to kill all the pathogens, according to some sources. A lower temperature will take longer to kill the pathogens. A temperature of 46C (115F) may take nearly a week to kill the pathogens completely; a higher temperature may take only minutes. What we have yet to determine is how low those temperatures can be and still achieve satisfactory pathogen elimination.

3.5.2 Oxygen and aeration


Aeration is a key element in composting, especially in aerobic composting, as a large amount of oxygen is consumed during initial stages. Aeration provides oxygen to the aerobic organisms necessary for composting. Proper aeration is needed to control the environment required for biological processes to thrive with optimum efficiency. Oxygen is not only necessary for aerobic metabolism of microorganisms, but also for oxidizing various organic molecules present in the composting mass. It also has the important function of controlling temperature as well as of removing excess moisture and gases. If the oxygen supply is limited, the composting process might turn anaerobic, which is a much slower and odorous process. A minimum oxygen concentration of 5% is necessary to avoid an anaerobic situation. Turning the pile regularly or by mechanical agitation will ensure sufficient oxygen supply.

14

A compost pile that is too hot can destroy its own biological community. It can also leave a mass of organic material that must be repopulated to continue the necessary conversion of organic matter into humus. Such sterilized compost is more likely to be colonized by unwanted microorganisms, such as Salmonella. Researchers have shown that the biodiversity of compost acts as a barrier to colonization by such unwanted microorganisms as Salmonella. Without a biodiverse indigenous flora, such as what happens through sterilization, Salmonella are able to regrow. The microbial biodiversity of compost is also important because it aids in the breakdown of the organic material. For example, in high-temperature compost (80C), only about 10% of sewage sludge solids could be decomposed in three weeks, whereas at 50-60C, 40% of the sludge solids were decomposed in only seven days. The lower temperatures apparently allowed for a richer diversity of living things, which, in turn, had a greater effect on the degradation of the organic matter. Even if every speck of the composting material is not subjected to the high internal temperatures of the compost pile, the process of thermophilic composting nevertheless contributes immensely to the creation of a sanitary organic material. Or, in the words of one group of composting professionals: The high temperatures achieved during composting, assisted by the competition and antagonism among the microorganisms (i.e., biodiversity), considerably reduce the number of plant and animal pathogens. While some resistant pathogenic organisms may survive and others may persist in cooler sections of the pile, the disease risk is, nevertheless, greatly reduced.

more than 20 different types of volatile organic compounds and their intermediates were recovered from the municipal solid waste composting facility (Komilis et al. 2004). The major phytotoxic compounds include either phenolic compounds or short chain fatty acids (Young and Chou 2003). Some of the phenolics are vanillic, trans-p-coumaric, cis-p-coumaric, phydroxybenzoic, ferulic, and o-hydroxyphenylacetic acids; short chain fatty acids include acetic acid, propionic acid, and butyric acid. The amount of these compounds varies with the composting method and feedstock.

3.7.2 Process control


Composting, being a microbial process, can be proceeded with a desired efficiency when the environmental requirements for decomposition are met at their optimal levels. To attain this, it is necessary to control the treat process. The important control parameters such as pH, humidity, and C:N ratio can serve as indicators for expected process failure. It is necessary to monitor the pH and maintain it between 6 and 7.5, which is an optimum range. It is well understood that during the process, this parameter undergoes considerable change from an initial pH of 5-6 due to the formation of carbon dioxide and organic acids. As the process progresses, the value will rise to 8-8.5, which is due to the decomposition of proteins and elimination of carbon dioxide. In a practical operation, very little evidence exists that pH should be artificially adjusted. The microorganisms that produce the acids can also utilize them as food after higher oxygen concentrations are established. This typically occurs within a few days after the most readily biodegradable substances have been destroyed. The net effect is that the pH begins to rise after a few days. The rise continues until a level of 7.5-9.0 is reached, and the mass becomes alkaline. Attempts to control pH with sulphur compounds are often difficult to justify because of the cost involved. As discussed earlier, the temperature change during the process has a profound influence on the efficiency of the process. As microorganisms decompose (oxidize) organic matter, heat is generated and the temperature of the compost is raised a few degrees as a result. The temperature is increased to 60o-65oC in the second phase and the thermophilic digestion takes over. Thermophilic treatment has advantage because of the increased organic removal efficiency, improved solid-liquid separation, and destruction of pathogens. Above 60oC, the thermophilic fungus flora dies while continuing the actinomycetes activities. The process stops when readily biodegradable material is fully consumed. The temperature then gradually decreases, which activates the reinvasion of the thermophilic fungus flora, which attacks the cellulose materials. On the completion of the digestion, the temperature returns to the ambient. The increase in the temperature favors saprophytic

3.7 Chemical changes


During composting, around 50% of the organic matter will be fully mineralized, producing CO2 and water. Protein, cellulose, and hemicelluloses are easily degradable. Many of these compounds produce organic residues, referred to as humic matter. A great deal of work has been recently conducted on humic matter from various sources. The amount of humic acid increases during the process. Increase in aromatic structures, phenolic structures, and carboxylic structures was also evidenced, whereas decrease in O-alkyl structures, polysaccharides, and amino acids was recorded with no changes in alkyl structures and carbohydrates (Chefetz et al. 1998).

3.7.1 Toxic intermediates


Many phytotoxic chemicals will also be produced during composting that might significantly impact on germination, plant growth, and also plant pathogens (Young and Chou 2003). In many instances, composting can also be a source of zenobiotic and hazardous volatile organic compounds. Recently,

15

activities that cause the transformation of the material in composting. Most composting should include temperature in the thermophilic range. At these temperatures, the rate of organic matter decomposition is maximum, and weed seeds and most pathogenic microbes cannot survive. It is also very important to mix the composting substances so as to ensure that all parts are exposed to high temperatures.

will be dependent on the feedstock, influenced by aeration, temperature control, and nutrient levels.

3.8.1 Respiratory rates (O2 uptake and CO2 formation)


To ensure sufficient aeration in the compost pile, levels of oxygen consumption and carbon dioxide formation should be monitored regularly during the entire process. A 1:1 ratio (oxygen/ carbon dioxide) will be an indication of a good composting process. Usually during the process, the oxygen concentration will reflect the changes in the CO2 evolution and temperature curves. The oxygen will decrease from its initial value of 21% to a value of 10% over the first few days as the temperature increases and the CO2 evolution increases, but gradually the oxygen level increases and returns to the 21% level as the temperature reaches ambient. The relation between CO2 evolution and oxygen consumption is called respiratory quotient (RQ). The RQ value of a good composting process will be about 0.9 (Atkinson et al. 1996).

3.8 Mineralization
The end products of any composting process are water, organic and inorganic matter that can be used as soil amendment to supply essential nutrients to the plants, in addition to the buffering action and to increase water-holding capacity (Fig. 6). During the composting process, the ash or inorganic component increases due to the loss of 2 organic fraction or volatile solids as CO . Values of volatile solids present in feedstock are between 65% and 99%. About one-third (20%) of the organic 2 material is decomposed into water and CO , but this

Plant & animal residues

Carbohydrates (soluble), Cellulose Hemicellulose, Proteins, Lipids, etc. CO2 H20, "NH3 H2S, etc.

Lignins, tannins, etc.

#
Microbial decomposition

Microbial decomposition & modification

"

H2O, CO2 NH3 etc.

#
Microbial protoplasm & metabolic byproducts Modified lignin & aromatics

#
Aromatic Structures

Aminoacids, Proteins, etc.

Mineralization

#
Humus

Mineralization

Fig. 6.

Organic matter decomposition and the formation of humic substances. (Adapted from Bear 1964, with modifications)
16

3.9 Addition of bulking agents, shredding the substrates, and mixing


Generally, mixing of bulking agents such as woodchips, yard trimmings, bark, rice hulls, municipal solid wastes or previously composted materials is used to add a source of carbon, lower the moisture content, provide structural support, increase porosity, and favor aeration. The composting method involves the use of substrates that are fairly coarsely shredded to obtain biomasses with interstitial spaces (homogeneous empty spaces) that account for more than 25% of the total volume of the biomass to be bio-oxidized. This is done in an attempt to overcome the problem, commonly encountered with agglomerated biomasses, of anaerobic fermentation occurring during the biooxidation stage. If the substrates per se are extremely fragmented and/or have a tendency to agglomerate, it is common practice to mix them with materials comprising large pieces (bulking agents). This is to create the interstitial spaces necessary to enable the air to flow and diffuse uniformly during the bio-oxidation stage. Another solution is to use systems that enable the biomass to be turned fairly frequently so as to break up the lumps and expose the resulting fragments to the air. In all these cases, the substrates to be biotransformed are prepared in such a way as to prevent the formation of agglomerates, since damaging anaerobic fermentation would inevitably occur within these. The principal limitations of all composting systems currently in use are the following: ! Poor, uneven aeration of the biomass; ! Fluctuation of the temperature of the biomass during the bio-oxidation stage; ! Unsuitability of the system to the use of mycelial microorganisms; ! Few active contacts between microorganisms/ enzymes and substrate; ! Little, if any, protection of the enzymes/ microorganisms from external agents; ! Limited use of the capacity of the bioreactors. The reasons for these limitations have been identified and studied, as follows: 1. The poor and uneven aeration of the biomass is due mainly to the fact that during the biooxidative process, the structure of the solid substrates loses its original characteristics. As a result, the substrates tend to collapse and fall in on themselves or, in the case of rigid substrates, tend to become compacted. Consequently, areas develop where the substrates become compacted, reducing and/or eliminating the interstitial spaces. The airflow is then reduced or blocked in these areas. 2. The temperature fluctuations during the biooxidative stage are due to the moisture initially present in the biomass evaporating during turning (designed to break up the aggregated mass and aerate it at the same time) and to

3.

4.

5.

6.

conductive and convective phenomena developing in the biomass. The unsuitability of the system to the use of mycelial microorganisms is due to the fact that the mycelium is damaged when the biomass is turned and so prevented from developing to the optimal degree on the surface of the substrate to be used and converted to useful biomass and/or to a particular product. The low number of active contacts between microorganisms/enzymes and substrate is due mainly to the limited surface area of the substrate. The lack of protection for the enzymes/ microorganisms from external agents is due mainly to the virtually non-existent porosity of the solid substrates. The limited use of the capacity of the bioreactors is linked to the need to mix the biomass with bulking agents and/or to turn or stir the biomass.

3.10 Odor management


Odor is the major problem associated with composting. Adopting proper management options can solve this problem. Odor is usually produced because of anaerobic conditions. Sources of anaerobic odors include a wide range of compounds, mainly ammonia, hydrogen sulphide, dimethyl disulphide, methanethiol, volatile fatty acids, amines, and several aromatic compounds. Odor usually originates from the site where it is stored and its storage condition prior to composting. Once the ingredients are incorporated into the composting system, subsequent odor problems are associated with the anaerobic conditions. Hence, it will be very essential to bring them back to aerobic conditions. The best way is to combine ingredients with coarse, dry bulking agents to increase porosity and to allow sufficient oxygen penetration. Subsequent turning and forced aeration systems can also provide sufficient oxygen. In addition to these conventional systems, oxidizing chemicals like hydrogen peroxide, potassium permanganate, and chlorine can be used to control the odor, but care should be taken not to kill the composting microorganisms. In situ biological oxidation or biofilteration is also an effective method of controlling the bad odors. Commercial enzyme catalysts and different biofiltering units, which can effectively reduce the odors, are available in the market.

3.11 Composting accelerants


There are several commercially available accelerants, which are added to the composting pile with anticipated results. Here are examples of such commercially available accelerants:

17

CBCT is a proprietary blend of beneficial microorganisms (BM), macro- and micronutrients, amino acids, enzymes, proteins, vitamins, and minerals. The microorganisms in CBCT are selected both for their effectiveness in degrading organic matter and for their ability to grow synergistically to high concentrations. They are among the most effective decomposers in the composting process. CBCT produces odorless, hygienic, mature compost that can be safely applied to the land for improved soil structure, moisture retention, and addition of a wide range of nutrients. CBCT initiates and accelerates the composting process. When CBCT is activated in an environment in which essential nutrients (organic materials) are present, the CBCT microbes rapidly grow to high concentrations and become the dominant organisms. These microbes provide optimized degradation for the biodegradable component in organic wastes. The end product of the process is a 100% organic fertilizer containing primary nutrients as well as trace minerals, humus, and humic acids. The by-products of the process are carbon dioxide and water. CBCT microorganisms are unaltered microbes originally derived from the soil, which utilize organic matter as a food source. The cultures are safe for the environment and are not harmful to animals, plants, and humans. Their other benefits are the following: ! They control composting. ! They are convenient to use. ! CBCT is safe for the environment. ! CBCT is non-toxic to animals, plants and humans. ! They control flies and insects by creating a poor breeding substrate. ! The resulting compost is odorless and hygienic and can be safely applied to any soil. Compost Treat is a scientifically developed combination of selected microbials and nutrients designed to initiate and accelerate the composting process. Compost Treat assists natural composting of organic matter and produces a more consistent, mature compost. The bacteria in Compost Treat are a selected mixture of mesophiles and thermophiles. Mesophiles grow and metabolize well at medium temperatures (70-115F); thermophiles do well at higher temperatures (95-140F). These types of bacteria are the most effective decomposers in the composting process. The viable bacteria in Compost Treat also produce several enzymes. The highly active enzymes assist in the decomposition of plant cell walls and other organic materials. Protease, amylase, xylanase, and pectinase all work on hard-to-digest components of the plant cell wall. Benefits from these include more rapid heat production; controlled, optimum composting; and convenient to use.

3.12 Conclusion
The process of composting is complex and can also happen naturally. The human wit has been successfully put in use in order to apply this technique of decomposition to convert the organic litter to useful compost, which in turn is eco-friendly. As the types of wastes utilized for the process vary in their qualities, modifications are incorporated day by day in order to achieve quality product. Further, to meet the complex nature of wastes, which contain toxic, hazardous substances that affect the end product, composting needs to be handled carefully. Since compost products vary significantly in terms of biological, chemical, or physical contaminants, the quality level of a compost product must be suited to the intended use of the product. When the process is managed efficiently, composting ensures that the finished product can be safely returned to the environment.

3.13 References
Atkinson, C. F., D.D. Jones and J.J. Gauthier. 1996. Biodegradabilities and microbial activities during composting of municipal solid waste in benchscale reactors. Compost Science and Utilization. 4,4: 14-23. Bear, F.E. 1964. Chemistry of the soil, ACS Monograph series No. 160, P. 258. Chefetz, B., F. Adani, P. Genevini, F. Tambone, Y. Hadar, and Y. Chen. 1998. Humic acid transformation during composting of municipal solid waste. Journal of Environmental Quality 27: 794-800. Day, D.L., M. Krzymien, K. Shaw, W.R. Zaremba, C. Wilson, C. Botden, and B. Thomas. 1998. An investigation of the chemical and physical changes occurring during commercial composting. Compost Science and Utilization 6 (2): 44-66. Epstein E. 1997. The science of composting. Technomic Publishing, Inc., Lancaster, Pennsylvania, p. 83. Finstein , M. S., F.C. Miller, P.F. Strom. 1986. Waste treatment composting as a controlled system. pp. 363-398. In: W. Schenborn (ed). Biotechnology. Vol. 8-Microbial degradations. VCH Verlaqsgedellschaft (German Chemical Society): Weinheim F.R.G. Hamoda, M. F., H.A. Abu Qdais and J. Newham. 1998. Evaluation of municipal solid waste composting kinetics. Resources, Conservation and Recycling 23: 209-223. Haug, R. T. 1993. The practical handbook of compost engineering. Lewis publishers, Boca Raton. Florida. 717 p. Howe, C.A. and C.S. Coker. 1992. Co-composting municipal sewage sludge with leaves, yard wastes and other recyclables a case study. In: Air Waste

18

Management Association. 85th Annual Meeting and Exhibition, Kansas City, Missouri, 21-26 June 1992. Kaiser, J.. 1996. Modeling composting as a microbial ecosystem: a simulation approach. Ecological Modeling, 91 25-37. Komilis, D. P., R.K. Ham and J.K. Park. 2004. Emission of volatile organic compounds during composting of municipal solid wastes. Water Research 38: 1707-1714. Liao, P. H., May, A. C. and Chieng S. T. 1995. Monitoring process efficiency of full-scale invessel system for composting fisheries wastes. Bioresource Technology 54: 159-163. McKinley V.L., and J.R. Vestal. 1984. Biokinetic analyses of adaptation and succession: Microbial activity in composting municipal sewage sludge. Applied and Environmental Microbiology. 47 (5). pp.933-941 Mc Kinley, V. L., J.R. Vestal and A.E. Eralp. 1985. Microbial activity in composting. Biocycle 26 (10): 47-50. Naylor, L. M. 1996. Composting. Environmental and Science and Pollution series 18 (69): 193-269. Neto, J. T. P., E.I. Stentiford and D.D. Mara. 1987. Comparative survival of pathogenic indicators in windrow and static pile. pp. 276-295. In: M.de Bertoldi, M. P. Ferranti, P. L Hermite and F. Zucconi (eds.). Compost: Production, Quality and Use. Elsevier Applied Science, London, United Kingdom.

Pace, M.G., B.E. Miller and K.L. Farrel-Poe. 1995. The Composting Process October 1995. Extension, Utah State University. AG- WM 01 Palmisano, A C and M.A. Bartaz. 1996. pp.125-127. In: Microbiology of solid waste. CRC Press.Inc. 2000. Corporate Bld. N.W. Boca Raton. FL 33431 USA. Palmisano, A. C., D.A. Maruscik, C.J. Ritchie, B.S. Schwab, S.R. Harper and R.A. Rapaport. 1993. A novel bioreactor simulating composting of municipal solid waste. Journal of Microbiological Methods 56:135-140. Reddy, K. R., T.C. Feijtel and W.H. Patrick. 1986. Effect of soil redox conditions on microbial oxidation of organic matter. pp. 117-153. In: Y. Chen and Y. Avnimelech (eds.). The Role of Organic Matter in Modern Agriculture. Nijhoff, Dordrecht. Sharma, V.K., M. Canditelli, F. Fortuna and Cornacchia. 1997. Processing of urban and agroindustrial residues by aerobic composting: review. Energy Conversion and Management 38 (5): 453478. Warman, P. R. and W.C. Termeer. 1996. Composting and evaluation of racetrack manure, grass clippings and sewage sludge. Bioresource Technology 55: 95-101. Young, C. C and C.H. Chou. 2003. Allelopathy, plant pathogen and crop productivity. pp. 89-105. In: H. C. Huang and S. N. Acharya (eds.). Advances in Plant Disease Management. Research Signpost, Trivandrum, Kerala, India.

19

4
4.1 Introduction
Categories

Raw materials used for composting


nutrients for plants reduces importation of chemical fertilizers and their manufacture. The Taiwanese government has heightened its advocacy on recycling organic wastes, including livestock manure, in agricultural fields. However, composting organic wastes in a farmhouse has been difficult because of shortage of labor and raw materials (Fig. 1). Commercialized composting is quite popular in Taiwan, and the material resources are mainly from livestock manure. This paper discusses the various raw materials for composting, their benefits and drawbacks, and the techniques of composting, mixing together different raw materials to achieve a better balance.

CHONG-HO WANG, YU-WEN LIN, WEI-TEN HUANG, and LI-RONG CHIU

Large amounts of agricultural by-products or livestock wastes are produced annually in Taiwan (Table 1). Hogs excrete about 73 x 104 mt/year; chickens, 186 x 104 mt/year; and cattle, 30 x 104 mt/year. Annually, byproducts from bagasse amount to 31 x 104 mt; straw, 31 x 104 mt; rice husk, 1.7 x 104 mt; bark, 5 x 104 mt; fruit and vegetable, 7 x 104 mt; mushroom, 5 x 104 mt; and coconut, 0.5 x 104 mt (Huang and Lin 2001). Composting agricultural wastes and recycling them on arable lands have been widely adopted to lessen the pressure on landfills and conserve natural resources. Thus, the use of organic materials containing essential

Table 1. Estimated amounts of organic wastes produced in Taiwan, 1999.


Wet weight (104 t) 317 532 166 98 266 c 60 6d 70 10 e 27 3.3 f 1555.3 Moisture (%) 77 b 65 82 68 Dry weight (104 t) 73 186 30 31 31 1.7 5 4 7 5 5 0.5 379.2

Hog manurea Chicken manure Cattle manure Bagasse Rice straw Rice husk Bark compost Peanut husk Wastes from fruit/vegetable markets Sawdust waste from mushroom farming Waste from fish market Coconut shell Total

30 90 50 80 85

Note: a The total amount of animal manure produced in 1999 was calculated from the number of cattle, hogs, and chickens at the end of that year and the amount of their daily excretion. b Estimated values. c The amount of straw produced was about 6 t/ha, and most of it was incorporated directly. d Peanut husk produced was 900 kg/ha. e Each mushroom culture bag weighed 800 g. f Water content in coconut fruit was 25%. Source: Huang and Lin 2001

20

Fig. 1. Composting organic fertilizers in a farmhouse is difficult because of shortage in labor and raw materials.

4.2 Kinds and sources of main organic materials


Organic materials in Taiwan may be classified into eight categories on the basis of their origin, as follows: crop residues, green manure, common compost, mushroom compost, animal manure (cattle, swine, and poultry), municipal refuse, residues after oil extraction, and residues from processing animal products. The nutrient contents of these materials differ greatly. The quantity of each organic waste category is shown in Table 2 (Hsieh and Hsieh 1990). The total amount of these major organic wastes is estimated up to 1,878 104 mt per year (Table 2). Some wastes such as rice straw, with a high C:N ratio of more than 20, can be considered a good source of carbon, but other wastes like animal manure with a low C:N ratio of less than 15 can be considered a primary source of nitrogen (Wang 1989). The daily average excretion of cattle (27.5 kg/ head) was the highest, followed by swine (0.85 kg/ head), then poultry (0.12 kg/head) (Table 3). According to Yen (1989), the nutrient content of poultry manure is much higher than that of swine or cattle manure (Table 4). Assuming the N, P2O5, and K2O content of daily excretion of livestock as shown in Table 5, the total amounts of nutrients in the manure of swine, cattle, and poultry in Taiwan were estimated to be 108,555 mt of N; 121,113 mt of P2O5; and 87,711 mt of K2O. These quantities were equivalent to 45.5% N; 173.7% P2O5; and 90.1% K2O of the chemical fertilizers used in 1987 in Taiwan (Table 6). Chang (1995) also reported that assuming the N, P2O5, and K2O contents of the daily excretions of livestock as shown in Table 5, the total amounts of nutrients in the manure of swine, cattle, and poultry in Taiwan were estimated to be 15.20 x 104 mt of N; 17.19 x 104 mt of P2O5; and 11.57 x 104 mt of K2O. These quantities were equivalent to 58% N, 233% P2O5, and 110% K2O of the chemical fertilizers used in 1987 in Taiwan (Table 7).

4.3 Chemical composition of organic wastes


Samples of crop residues; used mushroom compost; common compost; green manure; municipal wastes; swine, cattle, and poultry manure; residues after oil extraction; and animal by-products in Taiwan were analyzed for the chemical composition of their nutrient contents (Tables 8-14).

4.3.1 Crop residues


Rice straw, rice hull, and other straws of graminaceous crops with abundant fibrous materials usually have a high C:N ratio, with a low nitrogen content but fairly high potassium and silica contents (Table 8). Potassium and silica help improve the resistance of crops to disease and lodging, and fibrous materials provide an energy source for soil microorganisms as well as improve and condition soil physical properties (Fig. 2). Crop residues are used as mulches to cover the surface of the soil and help maintain favorable soil moisture content and temperature as well as prevent the accumulation of salts or the multiplication of weeds on the soil surface (Fig. 3). These materials can well be combined with swine or poultry manure that has a high nitrogen content to make better compost for crops (Lin et al.) (Fig. 4).

4.3.2 Green manure


Leguminous green manure crops are an important source of natural nitrogen. They fix nitrogen from the air and at flowering stage are usually incorporated into the soil, about ten days before planting the main crop. In extensively cropped areas, green manure crops are of great value to farmers since they reduce fertilizer costs. In intensively cropped areas, they may compete with the main crop for land. However, even in areas

21

Table 2. Estimated quantity of major organic wastes produced in Taiwan.


Properties during fermentation Type of organic waste WaterProduction holding (1,000 mt/year) capacity 2,600 600 2,100 600 6,000 2,500 200 4,040 90 50 18,780 Medium Low Medium Medium Medium Medium Medium Medium Decomposition Easy Difficult Difficult Easy Easy Easy Difficult Easy Easy Difficult Bulk density Low Very low Low Medium Low Low Low Low C:N ratio Large Large Large Small Small Large Large Large Large (>20) (>20) (>20) (<15) (>20) (>20) (>20) (>20)

Rice straw Rice hull Sugarcane leaves Wastes from food processing factories Swine manure Manure from other animals Corn cobs Municipal refuse Wastes from fruit/vegetable markets Bark waste Total
Source: Wang 1989

Table 3. Daily average excretion of swine, poultry, and cattle.


Swine Feces (kg/head/day) Urine (L/head/day) 0.85 2.70 Chicken 0.12 Cattle 27.5 13.5

Note: Swine: 50-kg body weight; poultry: adult chickens; cattle: 500-kg body weight. Source: Yen 1989

Table 4. Average content of N, P2O5, and K2O in fresh manure of swine, poultry, and cattle.
Moisture Swine feces Urine Chicken feces Cattle feces Cattle urine
Source: Yen 1989

Unit: %

N 0.63 0.48 1.66 0.43 0.47

P2O5 0.92 0.07 2.92 0.38 0.14

K2O 0.28 0.16 1.79 0.29 1.32

76.6 98.0 65.4 81.9 99.3

Table 5. Daily excretions of N, P2O5, and K2O of swine, poultry, and cattle (Unit: g/head/day).
N Swine (ratio) Poultry (ratio) Cattle (ratio) 18 (1) 2 (1) 182 (1) P2O5 9.7 (0.54) 3.5 (1.75) 123.0 (0.68) K2O 6.7 (0.37) 2.1 (1.05) 258.0 (1.42)

Note: Swine: 50-kg body weight; Poultry: adult chickens; Cattle: 500-kg body weight. Source: Yen 1989

22

Table 6. Total amount of N, P2O5, and K2O in swine, poultry, and cattle manure in Taiwan, 1987.
N No. of head Swine Poultry Cattle Total Equivalent to % annual chemical fertilizer consumption
Source: Yen 1989

P2O5

K2O

mt/year 46,830 50,334 11,391 121,113 173.7 25,258 88,119 7,736 87,711 90.1 17,447 54,093 16,171

7,129,034 68,978,000 171,759 108,555 45.5

Table 7.
Sources

Estimated environmental loadings of N, P2O5, and K2O coming from chemical fertilizer consumptions and animal excretions in Taiwan, 1994 (10 kton).
N 26.29 (100) 7.36 6.74 1.10 15.20 (58) 41.49 P2O5 7.37 (100) 4.64 11.80 0.75 17.19 (233) 24.56 K2O 10.53 (100) 2.77 7.24 1.56 11.57 (110) 22.10

Chemical fertilizers (Index) Hogs Chickens Cattle Total (Index) Sum


Source: Chang 1995

with very intensive multiple cropping systems like in Taiwan, some farmers are still growing Sesbania sesban or Crotalaria juncea as a green manure crop in summer, and Berseem clover, milk vetch, and rape in the winter fallow season before planting the main crop like rice, corn, and sorghum. The chemical composition analysis of these green manure crops is shown in Table 9. These green manure crops that have low C:N ratio (lower than 20 at vegetative stage) can be considered primary sources of nitrogen (Fig. 5).

compost should be combined with a proper amount of high-nitrogen manure such as swine or poultry manure or oil extraction residues and be well fermented to kill the mycelia, before applying to the soil (Fig. 8).

4.3.4 Animal manure (cattle, goat, swine, chicken)


The nutrient content of swine manure is slightly higher than that of cattle manure, but with a higher copper content (Table 11) and lower content of fibrous material, discouraging repeated, long-term applications of this manure (Fig. 9). It is best to dilute this manure by mixing it with rice hull, sawdust, rice straw, and similar fibrous materials and fermenting it before use. The nutrient content of chicken manure is much higher than that of swine manure (Table 11). However, its higher content of zinc and antibiotics and lower content of fibrous material discourage direct applications of fresh poultry manure to the soil. The best way to utilize this manure is to mix it with cattle and swine manure, rice straw, rice hull, sawdust, and other fibrous materials, and ferment it thoroughly before use. Cattle manure has a reasonably high content of nitrogen, potassium, and fibrous materials. It is good animal manure because it does not have heavy metals and antibiotics in it (Table 11). Repeated applications

4.3.3 Used mushroom compost


With the rapid development of the mushroom industry, used mushroom compost has become a good source of organic manure in Taiwan in recent years. Such compost consists mainly of sawdust (Fig. 6) and added with materials such as limestone and rice bran. Used mushroom compost has low potassium content as a result of leaching losses during mushroom culture, but the phosphorus, calcium, and C:N ratio and organic matter contents remain high (Table 10 and Fig. 7). Also, used mushroom compost has a high fibrous material content which improves soil physical properties and biological activity. However, the remnant mycelia in these materials may sometimes have a harmful effect on the roots of some crops. Therefore, it is recommended that used mushroom

23

Table 8. Chemical analysis of crop residues (dry matter basis).


C:N ratio
Rice straw Rice hull Rice bran Corn stalks Sorghum stalks Soybean stems Peanut stems Peanut hull Coconut shell 78-88 70-106 18-22 68 73 40 30 28 37 96 67-78

OM

P2O5
0.05-0.11 0.11-0.46 3.60-4.47 0.37 0.25 0.14 0.37 0.37 0.18

K2O
2.0-2.1 0.28-1.3 1.43-2.45 1.61 1.94 1.63 1.31 1.27 0.50

CaO

MgO

SiO2 (%)
4.9a 12.7 4.1 3.9 2.9 2.5 1.8

54-56 39-52 50-55 55 53 51 42 49 53

0.64-0.69 0.48-0.70 2.0-2.4 0.81 0.73 1.28 1.30 1.73 1.43

0.42-1.2 0.21-0.34 0.13-0.35 0.35 0.60 0.18 1.97 1.96 0.36

0.3-0.52 0.09-0.4 1.11-1.78 0.48 0.62 0.15 1.15 0.77 0.20

a Hsieh and Hsieh 1990. Source: Fertilizer guideline 2001

Table 9. Chemical analysis of green manure crops (dry matter basis).


Crops Sesbania sesban Crotalaria juncea Berseem clover Rape Ricia dasy carpa C:N Ratio 23.5 30.2 18.8 22.6 12.4 C % 38-56 53.7 46.3 38-42.7 46.0 N % 1.3-4.0 1.78 2.46 1.89-2.7 2.4-3.70 P2O5 % 0.28-1.0 0.46 0.78 0.71-0.94 0.54-0.92 K2O % 0.85-2.83 2.28 3.00 2.1-3.60 3.8-4.80 CaO % 1.61 1.40 1.83 2.23 1.15 MgO % 0.32 0.50 0.63 0.78 0.68

Source: Fertilizer guideline 2001

Fig. 2. Crop residues usually have low nitrogen content, but fairly high potassium and silica content, and a high C/N ratio: a) rice straw; b) crushed rice hull; c) peanut hull; d) coconut shell.
24

Fig. 3. Crop resides are used to cover the surface of the soil; these materials are good mulches which help maintain a favorable soil moisture content and temperature, and prevent the accumulation of salts or the multiplication of weeds on the soil surface.

Fig. 4.

Crop residues can well be combined with swine or poultry manure with high nitrogen content, to make better compost.

Fig. 5. Green manure which has a low C/N ratio of lower than 20 at vegetative stage, can be considered primarily as a source of N: a) Sesbania sesban; b) Crotalaria juncea; c) Berseem clover; d) Rape.
25

Fig. 6. Mushroom compost made up mainly of sawdust.

Table 10. Results of chemical analysis of mushroom culture wastes (dry matter basis).
C:N Ratio
White mushroom compost Shitake mushroom compost Jews ear mushroom compost Golden mushroom waste Mushroom waste sawdust

OM %

C %

N %

P2O5 %

K 2O %

CaO %

MgO

44 27-42 75 58-90 57

75

39 37-57 56 0.94-3.5 40

0.88 0.56-1.35 0.57 0.76-4.4 0.70

0.53 0.89-2.04 0.69 0.45-1.21 0.30

0.20 0.10-0.69 0.08 0.8-3.3 0.72

4.09 2.27-5.78 4.96 0.37-1.9

0.67 0.47-0.93 0.42

58 71

Source: Fertilizer guideline 2001

Fig. 7. Used mushroom compost has low potassium content as a result of leaching losses during mushroom culture, but phosphorus, calcium, and C/N ratio and organic matter contents remain high.

26

Fig. 8. Used mushroom compost should be combined with a proper amount of high-nitrogen manure such as swine or poultry manure or residues after oil extraction, and should be well fermented to kill the mycelia, before it is applied to the soil.

Fig. 9.

The direct application of fresh poultry manure to the soil is discouraged due to its high content of zinc and antibiotics, and low content of fibrous material.

Table 11. Results of chemical analysis of animal manure (dry matter basis).
C:N ratio Cattle Goat Swine Egg chicken Meat chicken 19-28 16-21 17-31 9-14 11-28 C (%) 25-40 36-48 4-54 27-32 25-47 N (%) 0.89-2.1 1.6-2.4 1.6-2.9 0.6-2.9 1.8-2.5 P2O5 (%) 0.55-4.81 1.5-5.27 1.0-7.1 1.4-6.8 2.11-6.6 K2O (%) 1.6-3.5 1.9-4.0 0.16-1.93 0.77-3.8 1.41-3.6 CaO (%) 0.20-2.0 1.3-5.4 0.8-9.0 0.73-8.2 1.57-21 MgO (%) 0.83-2.1 0.7-1.40 0.15-1.7 0.3-1.8 0.5-1.5 Cu Zn ---mg/kg--20 510 80 122 624 724

Sources: Fertilizer guidelines 2001 Hsieh and Hsieh 1990

27

of this manure to the soil can be recommended, but phosphorus should be supplied from other sources to make up for its shortage in this manure. Nutrient content of goat manure is slightly higher than that of cattle manure (Table 11).

these seed residues with rice hull, sawdust, mushroom compost, bone meal, oyster shell, among other things, and fully ferment the compost before use.

4.3.6 Residues from processing animal products


The nutrient contents of animal residues differ greatly according to the type of residue (Table 13). Animal blood, meat, horn, feet, wool, and feathers can all be used as a source of nitrogen fertilizer since they all have very high nitrogen content. Oyster shell and eggshell are good sources of calcium and bone meal can be a good source of phosphorus. However, all of them are very low in potassium. Fur should not be used in composting because of its high chromium content that can easily accumulate in the soil, causing toxicity in crops.

4.3.5 Residues from oil extraction


Oil extraction residues from oil seeds generally have high nitrogen content and low level of carbonaceous material (Table 12 and Fig. 10). Liberal applications of this material to the soil may greatly promote the growth of a crops vegetative parts. However, crops given this treatment are usually weak and easily attacked and damaged by plant pests and environmental stresses. As well, applying these residues to the soil when they are still fresh often attracts large numbers of soil-borne insects, which may also harm the crop. It is best to mix

Table 12. Chemical analysis of residues left after oil extraction from various oil seeds (dry matter basis).
Sample Sesame cake Soybean cake Soybean meal Cottonseed meal Castor bean meal Rapeseed meal Coconut meal Rice bran
Source: Hsieh and Hsieh 1990

N (%) 6.20 4.72 5.89 4.47 4.68 4.55 3.12 1.95

P2O5 (%) 1.26 1.85 1.81 0.80 0.89 0.87 0.57 4.37

K2O (%) 0.70 1.66 1.94 1.52 1.25 1.39 2.23 1.50

CaO (%) 3.02 0.39 0.38 0.36 0.98 1.18 0.25 0.20

MgO (%) 1.13 0.51 0.50 0.83 0.93 0.86 0.66 1.39

Fig. 10. Residues from oil seeds after oil extraction generally have high nitrogen content and low level of carbonaceous material.

28

Table 13. Chemical analysis of residues from processing animal products (dry matter basis).
Samples N (%) Animal blood Snail meal Fur meal Oyster shell Eggshell Bone meal 12.6 3.9 4.5 0.2 1.0 4.7 P2O5 (%) 0.16 3.39 0.11 0.14 0.21 21.98 K2O (%) 0.22 0.54 0.02 0.02 0.15 0.27 CaO (%) 0.28 20.30 0.56 28.84 30.52 17.22 MgO (%) 0.17 1.46 0.18 0.75 0.65 0.63 Cr (%) 2 18349 5 7

Source: Hsieh and Hsieh 1990

4.3.7 Reasons for composting agricultural waste for fertilizer use


!

4.4 Conclusion
Because of their multiple roles in improving the physical, chemical, and biological properties of soil, organic materials are very important in maintaining soil fertility. However, organic materials per se cannot give the full range of soil benefits. Some carbonaceous organic materials such as rice straw, corn stalk, rice hull, and sawdust are very useful in improving the physical and biological properties of soil, but they are very slow in releasing nutrients like nitrogen, phosphorus, and potassium. Some nitrogenous organic materials such as residues left after oil extraction and swine and poultry manure are high in nitrogen, phosphorus, and potassium but have little impact on improving soil physical properties. A proper combination of carbonaceous and nitrogenous organic materials makes an ideal compost or an effective and complete fertilizer. Furthermore, because some organic materials contain harmful mycelia, antibiotics, plant pests, and excessive level of heavy metals, proper mixing and composting, leading to dilution and sterilization, may greatly promote the quality of organic materials.

Composting improves the physical characteristics of agricultural wastes. It lowers the C:N ratio, thus avoiding the nutrient competition between plants and microorganisms. Because agricultural wastes contain relatively less nitrogen, they inhibit the growth and reproduction of nitrogen-loving microbes. It lowers the volume of waste by four-fifths its original volume. It sterilizes, because of high temperatures during composting, weed seeds, germs, and pests in agricultural wastes, reducing the cost of production and disease and pest control. It minimizes poor aeration problems. When directly applied without composting, agricultural wastes may exude toxic substances such as H2S, organic acids, and phenolic compounds and gas of methane and N2O.

4.3.8 Advantages of composting agricultural wastes


!

Composting lessens pollution impact on environments. It boosts soil fertility, improving both the biological and physicochemical properties of the organic material. It allows the utilization of essential nutrients from agricultural wastes in growing crops. It aids the use of slow-release fertilizers, particularly its nitrogen content, which after onethird is used, becomes slow-release humic nitrogen. It has a growth-promoting humic substance or phyto-hormone that accelerates root development. It increases and diversifies the microbe phase, reducing pathological and pest incidence. It minimizes nutrient loss as negatively charged organic material maintains and holds nutrients. It produces humic substances (humates) with high buffering capacity for better soil management.

4.5 References
Chang S.S. 1995. Research and development in the appropriate use of organic materials for crop productions current status and perspective. pp. 114. In: Taiwan Agricultural Research Institute (ed.), Proceedings of the Conference on Rational Application of Organic Fertilizer. Huang S.N. and C.C. Lin. 2001. Current of organic materials recycling in southern Taiwan. pp. 14-24. In: Food and Fertilizer Technology Center for Asian and Pacific Region (ed.), Proceedings of the International Workshop on Recent Technologies of Composting and Their Applications. Hsieh, S.C. and C.F. Hsieh. 1990. The use of organic matter in crop production. 315:1-19. Taiwan, ROC: Taichung District Agricultural Improvement Station. Lin, Y. W., T. S. Liu, and C. H. Wang. 2003. Study on nitrogen mineralization characteristics of organic materials. J. Agric. Res. China 53:178-190.

29

Wang, H.H. 1989. Utilization of agricultural wastes in organic farming. Organic Farming, special publication 16:217-227. Taiwan, ROC: Taichung District Agricultural Improvement Station. (In Chinese). Yen, S.C. 1989. Utilization of animal wastes in organic agriculture. Organic Farming, special publication 16:229-242. Taiwan, ROC: Taichung District Agricultural Improvement Station. (In Chinese).

30

5
5.1 Introduction

Equipment for different scales of composting


5.3 Pile composting
Organic materials, mixed at reasonable proportions to have C:N values between 20 and 30 and to have suitable moistures, were piled outside or inside of a composting shed at least 2 m3 with a height of 1-2 m. Outside windrow composting is only adapted to the seasons or regions with few rainy days. The length of frontage and depth of a composting shed (Fig. 1) is 1227 m and 4-9 m, respectively, in Japan (Harada 1995; Ibuki 1996), and 12-18 m and 9-18 m, respectively, in Taiwan (Sheng et al. 1995). The height of the eaves should be more than 3.5 m for the operation of the bucket loader. The height of the breast wall is usually 1.8-2.5 m. Partition walls are always settled in a composting shed. It is advisable to have a plastic roof that is pervious to light to avail of solar energy in temperate countries such as Japan (Harada 1995). But the roof of a composting shed is impervious to light in subtropical and tropical countries such as Taiwan.

SHIUAN-YUH CHIEN AND MING-HUEI CHANG

In Taiwan, lots of livestock manure and agricultural wastes are produced each year (Lin 1999; Chien 1999). If the wastes are not treated properly, they will cause serious environmental pollution. In the past, there were five methods of dealing with organic wastes. These were: throwing them away directly into country roadsides, streams, and brooks; burning them; using them as feed; incorporating them directly into farmland; and composting them (Chien 1999). The first two methods are not environment-friendly. The third method cannot be adapted for all organic wastes. The fourth method, though economical, may pose some sanitary problems, and may cause the soil to become too reductive, thus retarding the normal growth of crops. Only the fifth method is generally accepted by the people and the government. Through the composting process, the malodorous and unstable organic wastes are converted to organic fertilizers and soil conditioners. Using composts increases soil fertility and saves on chemical fertilizer costs.

5.3.1 Static piling composting


During composting, there is no further mixing or agitation. It may produce much stinky odor because there is no turning, thus causing an anaerobic condition. This option is adopted for less amounts of organic wastes, yard or household wastes (Fig. 2). Mixing organic materials is made by a manual shovel or bucket loader. With this composting method, the time needed to stabilize the organic materials may be one year or more (Lin 1999; Barkdoll et al. 2002; Harada 1995).

5.2 Methods and equipment for different scales of composting


Based on studies by researchers and work experiences of farmers in Taiwan, there are several composting methods for dealing with agricultural wastes. They are pile composting, box chamber composting, open furrow composting with turning and aeration, and enclosed vessel composting with mechanical agitation and aeration (Lin 1999). Generally speaking, less capital investments in equipment mean less capacity to treat wasted organic materials. These materials also need longer composting periods to reach maturity. In contrast, more capital investments in equipment mean more capacity and efficiency for composting organic materials (Barkdoll et al. 2002; Lin 1999; Ibuki 1996; Harada 1995; Fabian 1993). The above composting approaches are discussed in the following sections.

5.3.2 Piling along with turning composting


Organic materials are well mixed and then piled on the floor inside or outside of the composting shed. Composting materials are turned regularly with a manual shovel or bucket loader (Fig. 3). This is done in order to mix the composting materials uniformly and to improve the air permeability (Barkdoll et al. 2002;

31

4,000 - 9,000 Roof 12,000 - 27,000


1,500 4,000 - 9,000 >3,800
4,000 - 9,000 10

Compost pile

Slope 1/50

Fig. 1.

Sketch plan/profile (Harada 1995) and picture of a composting shed.

Fig. 2.

Organic materials are statically piled in a unit block volume.

Fig. 3.

Piling materials are turned regularly with a bucket loader.


32

1,200 800 3,500

Harada 1995). They are turned frequently during the initial period of high oxygen demand and heat generation and may be turned less frequently as the composting process proceeds. They may need to be turned several times per week, depending on the material being composted. Piling materials in a unit block volume surrounded by partition walls (Fig. 4) are 4-6 m (length) x 3-5 m (width) x 1.0-1.8 m (height) in Taiwan. A bucket loader (1-3 m3) is frequently used as a turner to agitate the materials once a week. The time for composting the materials to be stabilized is about two to three months. If the piling materials are agitated with a bucket loader once a week along with pumping with air, the time needed for organic materials to be stabilized as compost is about 28-35 days. Materials added with some maturity composts should be degraded more quickly than those without. A forced aeration system is placed under the piles (Fig. 5) to maintain a minimum oxygen level throughout the composting mass. This aeration system usually consists of a series of perforated pipes or floors running underneath the pile connected to a pump that blows air (positive pressure) through the pile to provide 100-300 L/min.m3 gas flux with 320 mm Hg aeration pressure. Aeration time is set

to pump air 10 min and stop 50 min intermittently. The capacity of the system to treat wasted organic materials can reach 40 t per day. It will produce 5,000-12,000 t of compost a year in southern Taiwan (Lin 1999).

5.4 Box chamber composting


The box chamber has three walls which are made of concrete block. The length of frontage, depth, height, and volume of a chamber are 3-5 m, 5-6 m, 1.6-2.5 m, and 8-50 m3, respectively, in Japan (Harada 1995), and 1.8-3.6 m, 1.8-5.6 m, 1.8-2.6 m, and 5-45 m3, respectively, in Taiwan. The aeration pipes are arranged on the floor of box chambers, and the gas flux is the same as the above description. Usually, several chambers are joined in one line (Fig. 6). Organic materials, adjusted to proper C:N values (2030) and moisture contents (55-65%) and mixed well, are then put into the chambers by using a bucket loader, belt conveyor (Fig. 7) or screw conveyor. The time needed for composting the organic materials to become mature will be more than two months (Lin 1999).

Fig. 4. Piled materials in a unit block volume surrounded by partition walls.

Fig. 5. Forced aeration pipe system is placed under the piling materials.
33

Fig. 6. Composting box chambers are joined in one line.

Screw conveyer

Animal waste Bulking agent Wastewater tank

Fig. 7. Composting box chambers filled with organic materials by using a screw conveyer (Harada 1995).

5.5 Open strip furrow with scoop type or rotary type turner composting
A strip furrow (Fig. 8) is 3-6 m wide, 1.5-2.0 m deep, and 50-80 m long (Lin 1999). Organic materials are composted in the strip furrow with turning machines and aeration. A turning machine is supported on a set of rails equipped over the strip furrow, and moves automatically without a manual operator. The wellmixed organic materials are placed at the front end of the strip furrow by using a bucket loader or conveyor. As the turning machine moves forward, 0.5-2 m/day, on the rails from another end of the furrow, it mixes and transfers the composting materials behind (Lin 1999; Ibuki 1996; Harada 1995). They are agitated automatically with the turner once or twice a day. At the bottom of the furrow, there is an aeration system containing aeration pipes and blowers (Fig. 9). Two types of turning machines, the scoop type stirrer (Fig. 10) and the rotary stirrer (Fig. 11), are usually used for agitating the composting materials.

When the strip furrow is equipped with a scoop type stirrer or rotary type stirrer, the suitable material stacking height is 1-1.5 m or 1 m, respectively. The composting period is dependent on the length of strip furrow, the frequency of turning, and the transfer distance of composting materials at each turning. Usually, the machine is operated once a day. It moves the materials 1-2 m at each turning, and if the strip furrow is 60 m long, the composting period is 30-60 days. In Taiwan, the composting capacity is enough to treat the feces produced from 1,200 heads of cow or 130,000 heads of layers to produce 1,100 t of mature compost each year (Lin 1999).

5.6 Open strip furrow with hung axle (crane) type turner composting
This composting system is basically modified from the above-mentioned open strip furrow with scoop type or rotary type turner composting (Lin 1999; Ibuki 1996). In this system, the hung axle (crane) type turner (Fig. 12)

34

Fig. 8. Composting strip furrow filled with organic material.

Fig. 9. At the bottom of composting strip furrow, there is a pipe system aerated with blowers.

is equipped, and the width of the strip furrow is larger. Usually, it may have 10-20 m in width. The turner automatically carries composting materials from one place to another in the furrow. Forced aeration is also provided in the composting system, which can treat the feces produced from 150,000 heads of hogs to produce 5,000 t of compost each year in Taiwan (Lin 1999).

5.7 Open elliptical furrow with turner composting


The lengths of the horizontal axis and the vertical axis of the elliptical furrow are 50-100 m and 6-10 m, respectively (Lin 1995; Harada 1995). The elliptical furrow is divided into two sides. Raw organic materials are put at the end of one side. The turning machine

moves round from the end of the other side, while the materials are transferred in an opposite direction in the elliptical furrow (Fig. 13). The scoop type or rotary type turning machine system and an aeration pipe system (the principles of operation are essentially the same as above) are equipped in this system. The materials are matured and dried after one round. The composting capacity and composting time are the same as those of the above open strip furrow with turner composting.

5.8 Open circular column furrow composting


The diameter and depth of the open circular column furrow (Fig. 14) are 6-9 m and 2-3 m, respectively (Lin 1999; Ibuki 1996; Harada 1995). At the bottom of the circular column furrow, there is an aeration pipe

35

Fig. 10. Sketch (Harada 1995) and picture of scoop type turner on strip furrows.

36

Fig. 11. Sketch (Harada 1995) and picture of rotary type turner on strip furrows.

Fig. 12. Crane type turner hung on a steel beam.

37

7,500 . 9,500 . 10,500

Turntable 180oC

Turntable 180oC Hopper Screw conveyer 100M

Screw conveyer Raw materials

Hopper

Wall

Wall

Wall

Fig. 13. Sketch (Harada 1995) and picture of elliptical composting furrow equipped with turners.

38

Deodorizing apparatus Odor Raw materials Blower Blower Screw conveyer

Deodorized air

Fig. 14. Sketch (Harada 1995) and picture of circular column furrows equipped with turners.

system, which is the same as that mentioned above. Raw composting materials are put into the circular column furrow from the circumference with a conveyor or bucket loader. The materials are mixed and transferred to the center of the column by using the scoop type stirrer (Ibuki 1996; Harada 1995). The mature compost is moved from the bottom of the center of the furrow. The characteristic of this composting method is that the scoop type stirrer is equipped almost vertically (Fig. 15). Raw organic materials are usually piled to a height of 1.5-2.5 m. This system may be adopted for composting in a cold region because the higher stacking of raw materials can be kept warm (Ibuki 1996; Harada 1995). In Taiwan, the composting capacity is enough to treat the feces produced from 88,000 heads of layers to produce 720 t of mature compost each year. The time needed for stabilizing the organic materials is 30-45 days in Taiwan (Lin 1999).

5.9 Enclosed vertical column composting


In enclosed vertical column compsting (Fig. 16), raw organic materials are put from the top of the column by a conveyor and agitated by a moving paddle. Air is blown up from the moving paddle through the composting materials by an aeration system (Fig. 17). The composted material is removed from the bottom of the column. The volume of the vertical composting column is 16.8 m3. Each day, one-third column volume of organic material (about 4 t) is loaded (Lin 1999). After three days in the column to get rid of stinky odors, the composted organic material is moved out and conveyed to the composting shed for further composting to reach maturity. The retention time of the material is very short, because the volume of the

39

Fig. 15. Sketch (Harada 1995) and picture of scoop type stirrer equipped vertically on circular column composting.

Fig. 16. Enclosed vertical column composting.

40

Heater

Blower

Exhaust gas

Moving paddle Air Moving paddle

"

$ %

& !

" Material inlet # Air outlet $ Air filter % Reduction gear & Outlet ! Hopper

Blower

Fig. 17. Aeration system of enclosed vertical column composting (Harada 1995).

reactor is smaller than those of the other compost bins and the composting shed. Mainly used for poultry wastes, this system has also been used for swine and cattle wastes recently (Ibuki 1996; Harada 1995).

5.10 References
Barkdoll, A. W., R. A. Nordsedt, and D. J. Mithchell. 2002. Large-scale utilization and composting of yard waste. CIR 1027:1-15. Agricultural and Biological Engineering Department, Florida Cooperative Extension Service, Institute of Food and Agricultural Sciences, University of Florida. Chien, S. Y. and T. C. Juang. 1999. Developing technology for producing compost from wasted mushroom sawdust. Composting Technology. pp. 91-106. In: Special Bulletin No.88 of Agricultural Research Institute, COA, Republic of China (In Chinese). Fabian, E E., T. L. Richard, D. Kay, D. Allee and J. Regenstei. 1993. Agricultural composting: A feasibility study for New York farms. Cornell University, USA: Cornell Composting. pp.1-50. Harada, Y. 1995. Practical aspects of animal waste composting. pp. 64-86. In: Lecture of international training course on microbial

fertilizers and composting. Rural Development Administration (Republic of Korea) and Food and Fertilizer Technology Center for Asian and Pacific Region. Ibuki, T. 1996. Examples of dairy manure composting in Japan. pp. 142-153. In: Proceedings of international training workshop on microbial fertilizers and composting. Taiwan Agricultural Research Institute and Food and Fertilizer Technology Center for Asian and Pacific Region. Lin, C. W. 1999. Composting of livestock feces. Composting technology. pp. 107-141. In: Special Bulletin No.88 of Taiwan Agricultural Research Institute, COA, Republic of China. (In Chinese). Lin, C. W. and S. Y. Chien. 1995. The research of using agricultural waste to produce compost. Proceedings of the techniques of reasonably using organic fertilizers. Tauoyang District Agricultural Improvement station, Republic of China. pp. 43-58. (In Chinese). Shen, S. Y., Lin, C. W., Hong, C. M. and M. D. Kao. 1995. Manual of using poultry feces as organic resources. pp. 68-70. In: Special Bulletin No.34 of Taiwan Livestock Research Institute, COA, Republic of China. (In Chinese).

41

6
6.1 Introduction

Composting methods
YUH-MING HUANG
!

Composting does not fully decompose all degradable organic materials. Because of various conditions, composting depends on the purpose, scale of production, materials, site, and climate. But whether it is for a small farm or a commercial plant, the first consideration in composting is processing the farmyard manure and crop wastes into a safe compost. The composting method should be able to inactivate the weed seeds and animal and plant pathogens. Accordingly, maintaining a ceiling temperature is very important during the process. The starting materials consist of readily degradable, slowly degradable, and resistant components in suitable proportions. For onfarm scale, the procedures designed and the investment required to process composting should be as simple as possible, especially for remote areas.

Residues of leguminous plants and green manure are also dried before being stored for composting and kept away from the rain. These materials have high C and N contents, so they can be used as animal feed. The animal manure can be used as a composting material. Manure from an animal feeding farm should not be exposed to the rain and its water content must remain low before composting to avoid N loss and sanitation problem. Wet manure from the farm should be put directly into the composting process; otherwise, it must be kept covered with 10-cm layer of mature compost, rice husk, straw, or sawdust as biofilter to prevent the release of odorous gases.

6.4 Ingredient calculation


Almost all organic wastes, except those containing high toxins, can be used as composting materials. There are no set ingredients for making compost, but the most important principle is that the ingredients are decided based on what purpose the compost produced is for. The easiest way is to calculate the C:N ratio of the mixed materials, as follows: ! To increase the soil organic matter or make a growth media, materials with high C:N ratio will be the main ingredient. The materials should have high contents of cellulose and lignin, especially the latter compound. The starting C:N ratio of the mixed materials should be kept at around 40-50. Under this formula, the materials have to be composted for about 10-15 weeks. The final C:N ratio of about 30 is good enough for application into soil and as growth media. The product is also a good base fertilizer for fruit trees and root crops, but chemical fertilizers or other organic fertilizers with high N content are also needed as supplemental fertilizers. ! To grow leafy vegetables, the starting C:N ratio of the mixed materials should be around 25-30. Under this formula, the materials have to be

6.2 Site selection and preparation


In choosing a place for composting wastes on-farm or off-farm, the first thing to consider is a downwind place where there is no flooding during the rainy season and no potential to pollute surface water and groundwater. A shelter is necessary in an area with high rainfall. A windbreak of slates or trees is also needed to counter strong winds and snow drifts. A concrete floor or at least a hardened soil surface is needed to make it easily workable manually or by machines, and to prevent wastewater from leaching into groundwater.

6.3 Pretreatment
Before composting, the raw materials, depending on their type, are stored and pretreated as follows: ! Straw and husk of grains/cereals (e.g., rice and wheat), corn stalk, bagasse, sawdust, and other materials (e.g., tree trunks and branches) with a high C:N ratio used as main component for soil organic matter additives should be dried and stored in a nearby place.

42

composted for about 6-10 weeks. The produced compost is ideal for organic farming of short-term vegetables. It is also a good base fertilizer for vegetables and fruit trees. ! To serve as supplemental fertilizer for vegetables (including leaf, fruit, stem, and root crops) and fruit trees, the starting C:N ratio of the mixed materials should be around 20-25. Under this formula, the materials have to be composted for about 4-6 weeks. The height of piles for this formula should be lower than the general suggestion. Calculating the amount of ingredients is the most important step in the whole composting process. Rather than buying materials outside of the farm, composting wastes available on-farm is the better alternative to develop the required composted mixture of materials. Besides, because they are bulky, the transport cost is quite high for most organic materials.

6.6.1 Static piles


The least equipment is required in managing this system. After the raw materials are wet with suitable water content (60-75% depending on the material used), they are mixed thoroughly then piled up as a round type (Fig. 1), the bottom part measuring about 2.5-3.5 m in diameter and 1.5-1.8 m in height. A square-type pile measuring 2.5-3.5 m x 2.5-3.5 m x 1.8-1.8 m is also acceptable. The scale of piles depends on the season: in hot season, a smaller scale is better for diffusing extra heat, while a larger pile is more suitable for the cold season. The pile can also be covered with a plastic sheet or organic material to keep the heat under reasonable levels. Piles are generally turned manually with fork or by machines like a frontend loader or bulldozer equipped with a bucket, rake, or blade. The most effective way of turning these piles is moving them by bucket from its original site to an adjacent area. Turning using a rake can aerate the pile on-site.

6.5 Size reduction


To shorten the compost time, some raw materials should be chopped or ground into chips to reduce their sizes. The straws and green manure should be cut into lengths of 5-10 cm. The wood and shoots of fruit trees and vines should be cut with shredders, grinders, or chippers.

6.6.2

Windrow system

For the windrow system, its bottom width is around 2-3 m with a height of 1.5-1.8 m, but its length depends on the scale of the place. A multi-windrow system placed side by side (Fig. 2) is also allowed. The preparation process is the same as that for static piles.

6.6 Composting systems


Different composting systems are widely used. Among these are static piles, turned windrows, forced-aeration static piles, in-vessel, and close-housing systems. Some systems require high investments in housing and equipment, but to save on costs for small farms, a simpler design is best. Static piles and windrow systems are the first two choices, while the forcedaeration static piles can be used either throughout the process or only for the first two weeks of composting raw materials with low C:N ratio.

6.7 Turning
Turning composting materials gives the benefit of forcing aeration, homogenizing materials, releasing excess heat, and adjusting water content. Since a high quantity of heat may be released to the media (23 MJ/ kg dry weight of volatile solids) during the decomposition of organic matter, the rise in temperature is inevitable. As such, the faster and higher the temperature rises, the lower is the C:N ratio of the materials formulated. To inactivate the

1.8 m 1.2

2.5 - 3.5m

Fig. 1. Wet raw materials piled up as a round type, the bottom part measuring about 2.5-3.5 m in diameter and 1.5-1.8 m in height.

43

pathogens, the ceiling temperature must be kept at 60-70C for the first two weeks. After which, the ceiling temperature can be kept at 45-55C for maximum biodegradation. Therefore, at the different stages, especially when the temperature rises higher than the ceiling temperature, the composting materials must be turned over. Most soil-borne plant viruses are more heatresistant than other pathogens. Materials infected by the tobacco rattle virus, for instance, require a ceiling temperature of 70C. Otherwise, a ceiling of 55-60C is suggested.

are acceptable for the test. When the germination rate is higher than 80% that of the control, then the composting process is ready to be terminated.

6.9 After composting


If the produced compost is not used immediately, it should be kept with water content lower than 35%. This can be done by spreading the compost over a drying surface to avoid continuous decomposition during storage.

6.8 Terminating the composting process


Much literature has been written on determining the maturity of compost using chemical, physical, and biological methods. The following are key factors that can help determine when to end the composting process: ! The compost no longer has odorous gases. ! The temperature of the composting materials no longer rises after turning. ! The compost is easily crushed by hand. ! The compost is already of dark color. ! The extracts of the compost already allow a high seed germination rate. Taiwans Pak-choi seeds

6.10 Adding microbial compounds


Microorganisms are present everywhere, and it is organic materials that nourish them. Usually, it is not necessary to add commercial microorganisms in composting materials. For initial materials with a high C:N ratio, adding some readily decomposable components such as rice bran, soybean meal, or young green manure is enough. Sometimes, applying urea stimulates microorganism growth. Adding 1-2% of surface soil or old compost to introduce microorganisms may also be done.

Fig. 2. A multi-windrow system.

44

7
7.1 Introduction

Management of composting
SHANG-SHYNG YANG
anaerobic methods. However, the aerobic method is generally preferred, since it proceeds more rapidly and provides greater pathogen reduction because higher temperatures are attained.

Proper and regular additions of on-farm organic wastes such as animal manure and crop residues are of utmost importance in maintaining the tilth, fertility, and productivity of agricultural soils; in protecting them from wind and water erosion; and in preventing nutrient losses through runoff and leaching. The restoration and rehabilitation of degraded soils to an acceptable level of productivity can be enhanced by using various off-farm sources of organic wastes such as sewage sludge, municipal solid wastes, crop wastes, and agricultural and industrial processing wastes. However, the quality and acceptability of these materials as soil amendments can be greatly improved through composting, or even by co-composting them with available on-farm wastes. Sustainable agriculture is increasingly viewed as a long-term goal that seeks to overcome problems and constraints on the economic viability, environmental soundness, and social acceptance of agricultural production systems worldwide. Soil quality has focused on soil productivity, food safety and quality, human and animal health, and environmental quality. It may also play a major role in plant health and in the nutritional quality of the food that is produced. Composting is a viable means of transforming various organic wastes into products that can be used safely and beneficially as biofertilizers and soil conditioners. It can resolve a number of problems associated with the use of raw and unstable organic wastes as soil amendments such as malodors, human pathogens, and undesirable chemical and physical properties. During the composting process, organic wastes are decomposed, plant nutrients are mineralized into plant-available forms, pathogens are destroyed, and malodors are abated (Parr and Hornick 1992; Yang 1997b). Composting is a microbiological process that depends on the growth and activity of mixed populations of bacteria, actinomycetes, and fungi that are indigenous to the wastes being composted. It can be done by either aerobic or

7.2 Management of composting materials


Many biomass materials show active decomposition accompanied by elevated temperature. However, some of these materials can be used for high-value utilization and are considered unsuitable for composting. For example, grass, straw, and foliage are generally more valuable as feeding materials for livestock. Therefore, other by-products and disposable products are generally considered for composting. These materials range from animal wastes, with a high fertilizer value, to straw and husk, with a minimal fertilizer content (Yang 1994, 1997a). The materials presently used for composting in Taiwan are listed in Table 1. These materials are sometimes processed separately, but more often they are processed as mixed materials to adjust the C:N ratio, bulk density, and moisture content. Selection of composting materials is important since it directly influences composting quality. There is less possibility of increasing harmful materials in composting of rice straw, but increasing use of livestock manure and industrial and municipal wastes creates concerns for composting. Therefore, selection of good raw materials for composting is crucial in quality control of composting. Hog manure has low C:N ratio, high bulk density, and bad aeration. Adding straw and corncob to hog manure will improve the C:N ratio and bulk density. Adding nitrogen and phosphate will also improve the composting. The C:N ratios of the composts of chicken manure, hog wastes, soybean oil extract residues with mushroom growth medium wastes decreased from 30 to 15 during composting. Rice hull contained SiO2 (11.20-14.10%), nitrogen (0.48-0.50%), and potassium (0.31-0.68%); the contents of other elements were low.

45

Table 1.
Materials

Materials suitable for composting in Taiwan.


Characteristic properties Porous, low moisture content, not easily fermented without pretreatment Porous, cellulosic materials Sloppy, high moisture content, easily fermented, offensive smell Cellulosic materials Not easily fermented; needs separation of inorganic materials, glasses, plastics

Straw, husk, sawdust, pulp, bark, corncob, bagasse, tea residue, coconut pulp Waste mushroom media Animal wastes, animal by-products, fishery by-products, sludge, vegetable-market wastes, household wastes, sewage Green manure Municipal refuse

Table 2.
Crop waste

Chemical analysis of crop wastes.


C:N N P K % Na Ca Mg SiO Fe Mn Zn Cu Ni Cr Cd Pb

mg/kg
0.16 0.36 0.20 0.46 0.14 0.05 0.20 0.07 0.12 0.25 0.15 0.43 0.15 1.41 0.10 0.97 0.48 0.17 1.69 0.25 2.49 0.26 11.39 0.26 0.39 0.03 0.05 0.29 0.37 0.69 0.63 0.15 0.21 1.02 0.68 8.5 10.3 11.2 14.1 3.0 3.8 3.1 3.4 1.8 1.2 4.8 1.8 406 455 186 160 183 268 672 610 388 2885 281 146 418 347 110 108 23 30 45 27 45 47 63 188 34 31 40 43 35 40 33 17 18 175 70 156 8 5 7 6 6 10 15 4 10 8 16 17 20 24 18 21 5 26 26 2 3 4 6 3 4 4 4 9 22 8 4 0.52 0.45 0.28 0.22 0.41 0.49 0.43 0.35 1.16 1.74 2.65 3 5 3 2 3 3 8 4 27 6 9

Rice straw (Japonica) 107 Rice straw (Indica) 85 Rice hull (Japonica) 116 Rice hull (Indica) 151 Corn stalk 61 Sorghum stalk 74 Soybean stem 39 Peanut stem 40 Peanut hull Bark Tobacco leaves 11 Tobacco factory waste 39

0.48 0.67 0.48 0.50 0.91 0.73 1.36 1.33 0.70 1.58 3.50 1.12

0.07 0.09 0.05 0.07 0.16 0.11 0.16 0.11 0.12 0.05 0.14 0.21

1.44 1.41 0.31 0.68 1.34 1.61 1.09 0.91 0.46 0.60 2.54 0.30

Source: Hsieh and Hsu 1993

Corn and sorghum stalks contained potassium (1.341.61%) and nitrogen (0.73-0.91%). Soybean and peanut stems had high contents of nitrogen, potassium, calcium, and magnesium. Although bark had high contents of nitrogen, calcium, and potassium, it also had high content of phenolic compound. Therefore, composting of bark is necessary. Tobacco leaves and tobacco factory wastes had high contents of nitrogen, potassium, calcium, magnesium, and nicotine (Hsieh and Hsu 1993) (Table 2). Vegetable wastes had high crude protein and crude fat contents, while bamboo shoot wastes had high cellulose and hemicellulose contents. Food wastes contained high total carbon and total nitrogen contents, while C:N ratio was low. Cow feces contained high crude fiber and ash contents, while total nitrogen, crude protein, phosphate, and potassium contents were low. Hog feces had average total nitrogen, crude protein, phosphate, and potassium contents; while chicken feces had high ash, total nitrogen, crude protein, phosphate, and potassium contents (Yang et al. 1991; Tsai and Yang 2004) (Table 3). Biosolids, municipal solid wastes, yard wastes, and food wastes contain pathogens. In yard wastes, the major source of pathogens is domestic animal

feces; in food wastes, eggs, chicken parts, and other contaminated sources can result in significant levels of pathogens. Composting, if carried out properly, is very effective in destroying pathogens. This is primarily the result of temperature-time relationships. However, other factors contribute to the demise of pathogens such as antagonistic organisms and ammonia. Regrowth of salmonellae is not a serious problem as these organisms die very quickly during curing and storage. Knoll (1961) described several experiments on different Salmonella strains subjected to composting temperature at the composting plant. After 14 days of reactor time with temperatures of 55o-60oC and a moisture content of 40-60%, the product did not contain pathogens. Pig manure has a considerable copper content coming from feed materials. Sewage sludge has a relatively high nutrient content and can be used as an organic fertilizer. However, it has a high content of heavy metals and, possibly, harmful organic materials. Human manure waste sludge had high phosphate and lead contents (Um and Lee 2001). Food wastes mainly come from agricultural products and can be used as raw materials for composting. The problem with food wastes was high salt (4.10%) and fat (3.53-8.02%) contents (Tsai and Yang 2004). According to company

46

Table 3. Compositions of organic raw materials for composting.


Item Total carbon 44.2 41.1 39.1 40.1 42.2 40.6 50.4 66.5 43.12 37.3957.06 49.8756.35 Total nitrogen % 7.0 0.9 0.7 0.5 0.3 0.3 0.2 0.1 2.03 1.543.25 2.152.63 0.30 0.60 1.60 C:N Ash Carbohydrate 11.6 22.8 25.0 16.3 21.6 6.8 10.9 0.5611.72 1.933.82 Cellulose Lignin % 7.5 31.8 37.0 41.9 48.2 55.2 48.2 8.6120.38 14.7338.11 13.6 17.1 11.2 20.6 15.5 15.3 30.5 Crude protein 43.6 5.6 4.1 3.4 2.1 1.8 1.3 14.0549.28 17.9321.10 Crude fat Hemicellulose

Sludge Rice root Rice straw Rice hull Wheat straw Paper waste Sawdust Lignin reagent Vegetable waste Bamboo shoot waste Food waste Food waste in college dormitory Food waste in national apt. Cow feces Hog feces Chicken feces

6.3 45.9 60.2 74.1 126.0 140.0 242.0 923.0 9.00 11.5328.42 21.4623.26

14.9 15.5 12.8 18.6 10.9 18.1 1.3 14.7326.38 4.488.30

2.47- 1.2810.49 5.06 0.84- 9.090.89 10.66

9.5310.34 8.4210.32 7.426.129.7 41.3 16.0 44.0 15.012.7-

12.918.7 20.0 28.0-

Source: Yang et al. 1991; Tsai and Yang 2004

Table 4. Nutrient contents of sludge from different sources.


Sludge T-C T-N % Sludge from water purifying system Sewage sludge Human manure waste Textile sludge Food sludge Dairy sludge Paper sludge Alcohol sludge Beverage sludge Oil sludge
Source: Um and Lee 2001

P2O5

K2O

Cu

Cr mg kg
-1

Cd

Pb

20.13 26.86 32.26 30.83 49.98 43.29 30.67 38.43 41.75 37.14

0.91 2.05 2.72 3.73 3.51 5.86 0.48 4.28 4.05 1.47

0.59 2.58 7.31 1.51 1.52 4.68 0.17 1.18 2.03 0.70

0.52 0.38 0.35 0.29 0.54 0.55 0.30 0.99 0.56 0.23

218 1,015 138 269 103 72 111 128 163 43

270 43 411 49 28 42 24 89 117

3 3 1 8 0.4 3 0.4 17 19

67 35 65 9 42 67 148 191

type, harmful materials in sludge are different. Sludge from food processing companies has a lead content and that from dairy industries has a low metal content. Sludge from oil industries has chromium and lead contents, and some harmful effects on plant growth were reported (Table 4). Sludge from industrial wastewater treatment process used as raw materials for composting needs safety examination. Some say that mixing harmful material with sound material can dilute adverse effects. But it is necessary to exclude such contaminant in the selection of raw materials. There are specific inspection standards for the preinspection of required usable materials. These standards include effective constituents of classified

harmful materials. In Korea, the standards of composting materials are as follows: organic material, >60%; As, <50 mg kg-1; Hg, <2 mg kg-1; Pb, <150 mg kg-1; Cd, <5 mg kg-1; Cu, <500 mg kg-1; Cr, <300 mg kg-1; Zn, <900 mg kg-1; and Ni, < 300 mg kg-1 (Um and Lee 2001). Co-composting is a waste treatment method where different types of wastes are treated together. As an attractive method of resource recovery and waste disposal, it offers many advantages. Co-composting is expected to cost less than separate treatment systems, mainly due to the low cost per volume treated at large treatment plants. Co-composting of potato starch sludge and piggery manure as a substitute for sawdust

47

was established. Composting by the aerated static pile method for 10-14 days and curing process with stacking the composts in 10-12 bags high can obtain a good product (Yang et al. 2001). Municipal solid wastes contain 74-84% organic matter, 1.4% nitrogen, 0.1% phosphorus, and 0.1% potassium in Jakarta, Indonesia. The addition of nitrogenous compounds to adjust the C:N ratio can often accelerate the process by alleviating the microbial demand for nitrogen. Urea addition accelerated the composting process (47 days for urea addition and 54 days for control) (Prihatini and Kurnia 2001).

composts and vegetable-market waste composts are the major composts in Taiwan. However, some companies provide sewage sludge as raw materials for composting. These materials are mixed with livestock manure for composting and have a high phosphate content (Table 5). Therefore, compost is applied with a mixture of chemical fertilizers to get appropriate ratios in N, P, and K.

7.3.3 Immature compost


Composting is minimizing the damage from intermediate compounds in the degradation of organic materials. It takes about 6 months in a natural environment and about 80 days in an industrial facility. Some immature composts are being sold in the market. Biological indicators or regulations should be established for maturity determination.

7.3 Problems in the use of composts


7.3.1 High price
Average price of commercial compost is about $2 for a 20-kg package in Korea and about $3 in Taiwan. When compared with nitrogen supply only, the price of compost is 20 times higher than that of chemical fertilizer. Although most farmers agree on the positive effects of composts, some low-grade composts are very popular for their low price and some crop damages are attributed to industrial wastes used as raw materials.

7.3.4 Salt accumulation and environmental pollution


When livestock manure compost is applied as N requirement to crops, phosphate input becomes an excess of 200-954% over standard input. Food waste compost has high concentrations of salt and fat. Excess application of such compost damages crops by salt accumulation in the soil. Generally, such soil shows unbalanced excess accumulation of phosphate, potassium, and nitrate (Um and Lee 2001; Tsai and Yang 2004).

7.3.2 Unbalanced nutrient content


Livestock manure composts are the major commercial composts in Korea, while both livestock manure

Table 5. Composition of composts.


Item pH Moisture content Organic matter Total organic carbon Total nitrogen C:N P K Ca Mg

% Hog waste compost Hog waste and rice hull compost Cow waste, chicken waste, sawdust, and bagasse compost Cow waste and rice hull compost Chicken waste compost Chicken waste and rice hull compost Sheep waste compost Manure compost Bagasse compost Bagasse and hog waste compost Bark compost Sawdust compost Sludge and rice hull compost Vegetable waste compost 6.9 7.98.3 7.5 7.1 7.7 7.1 9.3 8.1 6.9 41.1 25.636.8 42.0 36.2 29.6 22.9 57.2 60.0 68.0 67.0 19.4 60.4 85. 71.677.6 65.0 63.2 50.0 55.2 76.1 73.0 58.0 37.6 42.2 53.9 63.8

% 2.1 1.42.4 1.9 2.3 2.2 2.0 2.7 0.3 1.08 16 1523 15 12 10 12 13 36 30 21 10 34.4 1.0 3.85.5 2.3 2.9 9.2 6.3 2.1 0.1 0.2 0.25 6.6 0.76 0.4 1.73.1 1.7 4.2 3.34.3 5.5 0.5 0.92.4 1.3 2.3 1.8 1.6 1.2 -

38.2 32.234.9 29.3 28.4 22.5 24.8 34.2 10.8 33.0

5.0 3.0 4.6 14.3 4.5 4.2 0.04 0.57 0.48 0.7 1.13 8.6 5.9 -

26.0 1.22 2.0 189.3 1.9 22.0 24.3 37.1 2.4 1.08

8.9 1.1 3.67 0.27

48

7.3.5 Greenhouse gases emission


Unsuitable handling or treatment of livestock wastes has caused the emission of harmful gases such as ammonia from livestock production systems. In addition, livestock wastes contribute significantly to the emissions of CO2, CH4, and N2O, which are greenhouse gases. Some efforts have already been made to quantify the emissions from livestock waste stores and treatment systems. One cubic meter of swine dung heap compost emitted 400-970 g of ammonia, 0.6-386 g methane, and 1.9-71.9 g nitrous oxide. During composting for eight days, each cubic meter of compost produced 2 g of CO2, 239 mg CH4, and 660 mg NH3. Each ton of hog wastes, cow wastes, and chicken wastes produced 22.1 g, 16.3 g, and 12.9 g of methane, respectively, during a year of composting. Methane emission rates from the compost of the mixture of hog wastes, chicken wastes, and sawdust ranged from 0.12 mg-2 h-1 to 707 mg-2 h-1 depending on the stage of maturity. Nitrous oxide emission rates were between 0.15 mg m-2 h-1 and 25.0 mg m-2 h-1, and those of carbon dioxide ranged from 1,036 mg m-2 h-1 to 19,558 mg m-2 h-1 (Osada et al. 2000; Tang and Wang 2000; Osada 2001; Chen et al. 2003; Lai et al. 2003).

7.3.6 Quality control


Compost manufacturing standards are based on the supply of organic materials to the soil. There are no regulations on nutrient content. The commercial organic fertilizer and commercial compost companies are small in scale compared with other manufacturing industries. They thus have difficulties in keeping their high-quality manpower and in maintaining their composting facilities. Special fresh industrial wastes that are supplied by the waste disposal companies have many serious problems. Most commercial organic products and composts give the composition on the package label (Table 6). But there is no such information for organic products or composts in bulk, so it is difficult to get practical information for using them.

7.3.7 Heavy metal content


There are some regulations on heavy metals in composts in order to protect soil environment. Allowable heavy metal concentrations on a dry weight basis are used in Taiwan, Europe, the USA, and Japan, and on a fresh weight basis in Korea (Table 7). Although sewage sludge contains considerable

Table 6. Standards of compost in Taiwan.


Item Moisture content Organic matter Total nitrogen % Compost (general) Layer waste compost Mixed compost <35 <35 <35 60 40 40 0.6 2.0 0.3 0.3 2.0 0.3 0.3 1.0 0.2 0.01 0.01 0.00 0.08 0.08 0.08 Total phosphate Total K2O Cu Zn

Table 7. Allowance of heavy metal content in composts.


Country USA Japan Austria Belgium Colombia Italy Holland Canada Spain Switzerland Korea Taiwan Cd 225 5 4 5 2.6 10 0.72 34 40 3 5 5 Cr 1000 (1.5) 150 150200 210 3500(Cr+3) 10(Cr+6) 50200 50 750 150 150 Cu 4501000 400 100500 100 600 25300 60100 1750 150 150 Hg 510 2 4 5 0.8 10 0.22 0.150.5 3 2 2 Ni 50200 100 50100 50 200 1050 60 400 50 25 Pb mg/kg 2501000 (3) 500 6001000 150 500 65200 150500 1200 150 150 150 9002500 1000 10001500 315 2500 75900 500 4000 500 500 50 13 525 1020 50 50 26 25 60 10 5 23 20 Zn As Co Mo

49

amounts of plant nutrients such as organic matter, N, P, and K, it is not used in agricultural practices because of its high concentration of heavy metals. The Cu contents are generally high in poultry and hog composts from 9 mg kg-1 to 394 mg kg-1 and from 6 mg kg-1 to 301 mg kg-1, respectively. The Zn contents are between 56 mg kg-1 and 1,147 mg kg-1 and between 35 mg kg-1 and 623 mg kg-1, respectively. The Cr contents are often high in organic manure of animal origin, in hog compost, and in compound organic manure. The contents are 2 mg kg-1-54,324 mg kg-1, 3 mg kg-1-8,486 mg kg-1, and 2 mg kg-1-9,196 mg kg-1, respectively (Table 8). The product of hog compost contains nitrogen, 2.8%; P2O5, 1.7%; K2O, 0.23%; organic matter, 86%; Zn, 349 mg kg-1; and Cu, 140 mg kg-1. Rice straw contains potassium (1.41-1.44%) and nitrogen (0.48-0.67%); the contents of other components are low. The high content of Cr in the manure is conjectured to be related to the admixing of animal skin powder which is the by-product of the leather industry (Lian and Lee 1994). The sludge of final clarifier has the highest Cu content, the primary clarifier of aerobic treatment comes next, and the anaerobic treatment has the lowest. However, the sludge of anaerobic treatment has the highest Fe and Mn contents (Fu and Chen 1990). The heavy metal contents of sludge depend on the feed supplement.

7.3.8 Organic compounds


The presence of toxic organics in compost depends on the type of feedstock involved. Biosolids can contain organic compounds as a result of the disposal of industrial, commercial, and household wastes. Pesticides can be found in yard wastes and food wastes. Pthalates are found in plastics along with other organic dynes and compounds. Household wastes discharged into the municipal solid waste stream contain oils, solvents, pesticides, and many other toxic organic compounds. Paper products may contain toxic organics as a result of printing inks and ash discharged from incinerators or boilers may contain dioxins. Compost has been shown to be an effective method of microbial degradation of many toxic organics. Thus, many organisms are capable of decomposing toxic organics and use the C as an energy source. Many of the studies were conducted under laboratory conditions. Therefore, research is needed to develop techniques for cost-effective composting of recalcitrant organic compounds. These may include inoculating specific organisms and providing the environment that sustains them.

Table 8. Heavy metal contents of commercial organic fertilizers.


Organic fertilizer Oil extract residue Plant residue Animal residue Chicken compost Hog compost Cow compost Bark compost Cu 9-161 (24)* 1.4-30 (15) 1-561 (84) 9-394 (99) 6-301 (101) 14-225 (82) 22-52 (34) Zn 37-372 (73) 10-127 (53) 42-2,827 (424) 56-1,147 (286) 35-623 (232) 51-308 (152) 77-216 (127) 5-814 (163) Cd 0-2.5 (0.27) 0-0.2 (0.03) 0-4.8 (0.68) 0-6.2 (1.4) 0-7.0 (1.9) 0-3.7 (1.36) 2-5.6 (3.4) 0-5.6 (1.2) Cr 0.2-24 (5.7) 2.1-15 (6.7) 2-54,324 (6,252) 1-282 (23) 3-8,486 (898) 3.1-314 (39) 21-46 (28) 2-9,196 (469) Ni 1.4-15 (6.3) 4.0-5.7 (4.9) 0.7-415 (47) 1.1-30 (12) 3.5-176 (32) 4.3-18.6 (11) 12-31 (23) 2.4-92 (18) Pb 0.3-27 (4.5) 0.3-5 (3.0) 1.4-37 (7.2) 0-183 (14) 0.5-60 (17) 1.6-19 (7.4) 13-34 (23) 0-150 (11) As 0.7-13 (2.6) 1 6-100 (36) 0-66 (9.7) 0-27.3 (9.6) 1.5-113 (18) 0.3-24 (13) 0.9-65 (25) 0-57 (13) mg/kg (dry weight basis)

Compound compost 2.4-234 (38)


*Mean value Source: Lian and Lee 1994

50

7.3.9 Distribution and use of composts


The major reasons for non-adoption of rice straw composting include time-consuming interference with various farming activities, difficulty in applying, and burning as a convenient method of disposal. In addition, there are some problems in both quality and handling of standard animal waste composts, namely: content of fertilizer nutrients and fertilizer efficiency are unclear and their seasonal measurement fluctuates notably; maturity of compost is indefinite; special spreader machine and large storage space are necessary, as well as high forwarding charges, because the specific gravity of compost is small and the moisture content is usually high. To solve these problems, the following criteria must be set (Hara 2001): ! Production of high-quality animal waste composts that fit the needs of the cultivating farms by guaranteeing the quality and fertilizer efficiency, and adjustment of the content of plant nutrients of the compost produced for the cultivating plant. ! Production of animal waste composts which can withstand distribution over distances by decreasing the volume and moisture content of produced compost. ! Production of animal waste composts which can be spread easily. ! Creation of manuals of animal waste compost fertilization. ! The handling of animal waste compost, which is produced in granular or pellet form, is better than that of conventional compost.

7.3.10 Shelf life


The major consideration under this item is how to prolong the shelf life of the compost for agricultural applications by the inoculation of biological microbes.

7.4 Inoculation of microbes


Microbes can dissolve insoluble phosphate compounds in soils, fix nitrogen to biomass from the air, mineralize organic compounds to release nutrients, and secrete plant growth-promoting substances. Inoculation of appropriate microbes into composts could improve the quality of composts and shorten the period of maturity. Inoculation of phosphatesolubilizing Klebsiella pneumoniae subsp. and nitrogen-fixing Azospirillum brasilense into the compost of livestock wastes and mushroom waste media increased the soil phosphate and nitrogen concentrations, and the growth of kale, cabbage, and corn. Inoculation of Rhizobium japonicum and Bradyrhizobium japonicum for nitrogen fixation supported the growth of food crops and legumes, and increased the yields of soybean, peanut, and alfalfa by 6-21%, 25-69%, and 123-179%, respectively (Kang and Ha 1994; Chien 2001).

The application of microbial inoculants as microbiofertilizers can accelerate the decomposition of organic residues in various processes with a concomitant release of plant nutrients through mineralization; facilitate the uptake of plant nutrients (e.g., mycorrhizal association); increase the nitrogen content of plants through symbiosis (e.g., Rhizobium) and other associative N2-fixing systems (e.g., Azospirillum); and improve plant growth and vigor by providing plant growth-promoting substances. The well-nodulated legumes reduce soil erosion and increase soil fertility. The young rubber trees also benefit from the mineralizable nitrogen released during the decomposition of the legume residue. Nitrogen fixation by Azorhizobium, Azospirillum in association with several plants had been well documented: maize, kallar grass, sorghum, millet, sweet potato, and oil palm. It increased the number, yield, and nitrogen concentration of the crops. Inoculation with VAM (vesicular-arbuscular mycorrhiza) had also been found in rubber, cocoa, Calopogonium caeruleum, chili, and groundnut. Bacillus pumilus could decompose cellulose and grow at high temperature (Shamsuddin 1994). In Taiwan, nitrogen-fixing microbes, Rhizobium, had been applied in lupins, alfalfa, peanuts, crotelarias, and soybeans to increase crop yields; phosphate-solubilizing bacteria, Bacillus, had been used to improve the phosphate availability in the soil; VA-mycorrhizal fungi, Glomus, had been inoculated to enhance the nutrition of the host plant (Chang and Young 1992; Tseng et al. 1993; Young 1994; Chang et al. 2001). For the inoculation of microbes for biofertilizer preparation, it is necessary to investigate the following items: 1) selection of effective and competitive biofertilizers for the crops; 2) inoculant production and application to the field to ensure the benefits of plant-microorganisms symbiosis; 3) study of microbial persistence of biofertilizers in soil environments including stresses on production; 4) agronomic, soil, and economic evaluation of biofertilizers in agricultural production; and 5) study of rhizosphere environment and microbial interaction. The quality of the municipal solid waste compost could be improved by inoculation with beneficial microorganisms such as antagonists. Inoculation of non-pathogenic Pseudomonas spp. to control Pythium and Sclerotium in tomatoes reduced the damage intensity by the pathogen from 63.0% to 38.9% and from 41.7% to 16.7%, respectively (Iswandi 1994a, b). Antagonistic fungi Gliocladium frimbiatum or Trichoderma hamatum inoculation into municipal solid waste compost inhibited the Rhizoctonia, Sclerotium, Pythium, and Fusarium diseases in peanuts (Sinaga 1992). Inoculation of non-pathogenic Pseudomonas spp. LIES and LD reduced the infection of peanuts by Sclerotium rolfsii (Iswandi 1994b). Inoculation of Trichoderma sp. SS33 into the compost hastened the decomposition of rice strawchicken manure mixture, Azotobacter sp. H1BFA 4b

51

promoted the biological nitrogen fixation and enriched the nitrogen content, and Trichoderma-Azotobacter increased the decomposition rate and nitrogen fixation activity. The biofertilizer prepared with the inoculation increased the soil nitrogen, phosphate, potassium, organic matter, and rice yield (Espiritu and dela Torre 2001).

7.6 Effects of composts on crop yields and soil qualities


Application of compost to the soil improved the soil organic matter, P, and Zn contents. The average organic matter, P, and Zn contents of compost application were 2.85%, 8.14 ppm, and 3.53 ppm, respectively; while the values of farmers practice were 2.18%, 6.14 ppm, and 2.62 ppm, respectively. In addition, rice plant with compost application contained higher N, P, and Zn than that of farmers practice. More recently, soil quality includes soil productivity, food safety and quality, human and animal health, and environmental quality (Parr and Hornick 1992; Evangelista 2001). The yields of cabbage in a cattle compost plot with 1.0, 2.0, and 3.0 N application rates were 55-61%, 7582%, and 84-91% as those with chemical fertilizer application, respectively; while the yields of rice at first crop season were 72-74%, 80-83%, and 88-91% as those with chemical fertilizer application, respectively (Huang and Lin 2001). The relative N efficiencies of compost for cabbage and rice were 0.219-0.329 and 0.254-0.292, respectively. N-value in compost of cattle manure had about 25% of chemical fertilizer. The relative N-value of chicken manure was the highest with 0.37 of chemical fertilizer in rice cultivation, cattle manure (0.34) was the second, and hog manure (0.32) was the lowest. In cabbage cultivation, hog manure (0.37) was the highest, chicken manure (0.29) was the second, and cattle manure (0.25) was the lowest. In peanut production, cattle manure (0.21) was the highest, hog manure (0.19) was the second, and chicken manure (0.11) was the lowest (Table 10). Mineralization rate and amount of N-release varied with soil texture and kind of manure. However, N-value of compost increased gradually year by year in continuous application conditions. Soil Zn contents in all compost plots were higher than those in chemical fertilizer plots. The hog manure compost plot had a higher Cu accumulation (Huang and Lin 2001). When composts and residues are applied to soils, many of the introduced microorganisms can function as biocontrol agents by controlling or suppressing soil-

7.5 Values of composts


Composts are considered substitutes for chemical fertilizers. Compost amendments increased cation exchange capacity, buffering capacity, chelating capacity, soil aggregation, aggregate stability, waterholding capacity, soil porosity, water infiltration, water percolation, nutrient availability, and earthworm population; they decreased soil crusting, bulk density, and plant pathogens. In addition, compost application alleviated acidic and alkaline conditions, and stimulated beneficial microorganisms to produce polysaccharides and/or antibiotics (Parr et al. 1994). Among the other values of composts are animal feed, water gain, soil conservation, and soil-carbon sequestration. The values of four common composts (i.e., cattle manure, crop residues, sewage sludge, and municipal solid wastes) based on their average macronutrient contents are shown in Table 9. The important consideration is good quality composts and residues. Biofertilizers and soil conditioners have far greater values than just their macronutrient contents. These materials have a much greater residual effect on soil tilth and fertility than most chemical fertilizers because of the slow-release character of their nitrogen and phosphorus components. Colacicco (1982) estimated that the cumulative agronomic and economic value of some organic materials applied to agricultural soils could be more than five times greater in the postapplication period than the value realized during the year of application. Application of compost improved the development of root systems, increased the diversity of root fungal flora, promoted the growth of plants, reduced the incidence of soil-borne diseases, and depressed the propagation of pathogens (Nitta 1994).

Table 9. Values of some organic wastes based on their macronutrient contents. Organic waste Cattle manure Crop residues Sewage sludge Municipal solid wastes N 4.4 1.1 4.0 0.7 Nutrients P % 1.1 0.2 2.0 0.2 K 2.4 2.0 0.4 0.3

52

Table 10. Relative N efficiency of cattle manure compost.


Year Crop 0 1995 1996 1997 1998 1995 1996 1998 1995 1996 1997 autumn cabbage autumn cabbage autumn cabbage autumn cabbage spring paddy rice spring paddy rice spring paddy rice summer paddy rice summer paddy rice summer paddy rice 0 0 0 0 0 0 0 0 0 0 Rate of compost-N application (kg N/ha) 240 (160)* 0.2917 0.3208 0.3542 0.3625 0.2438 0.3375 0.2688 0.3063 0.3688 0.3813 480(320) 0.3021 0.3438 0.3417 0.3229 0.1969 0.2688 0.2781 0.3344 0.3781 0.3156 720(480) 0.2472 0.2847 0.2750 0.2764 0.1958 0.2583 0.2500 0.3021 0.3188 0.2583

*Value in parenthesis present for paddy soils.

Table 11. Fertilizer N efficiency. Nitrogen source Azolla Fresh rice straw Azolla compost Rice straw compost 1/2 Azolla + 1/2 urea 1/2 fresh rice straw + 1/2 urea 1/2 rice straw compost + 1/2 urea Supergranulated urea Urea Fertilizer efficiency (kg rice kg-1 N applied) 3 2 3 7 9 5 3 10 10

borne plant pathogens through their competitive and antagonistic activities (Kloepper et al. 1989). The possible mechanisms are the shifts in soil microbiological equilibrium following the addition of microbial inoculants and organic amendments. These include antibiosis, competition, parasitism, detoxification, and inhibition.

7.7 Microbiology of composting


Various microorganisms participate in composting, and the microflora population changes successively at different stages of composting. In the early stages of composting, mesophilic bacteria and fungi are dominant and consume sugar, starch, and protein. As the temperature rises above 40oC, these are replaced by thermophilic bacteria, actinomycetes, and thermophilic fungi. At this stage, lipids, hemicellulose, and cellulose are decomposed. Finally, as the temperature falls, mesophilic bacteria and fungi reappear. In other words, sugar and starch are the first to decompose during composting, followed by hemicellulose, and, finally, lignin. During the composting of sewage sludge, the number of mesophilic actinomycetes and fungi was around 102 cells g-1 of dry compost (Kubota and Nakasaki 1994). Inoculation or seeding has been reported to be effective in some composting procedures. On the other

hand, some negative results have also been reported about inoculation. The rate of composting of sewage sludge is mainly controlled by the degradability of the solid substrate, and not by the kinds of microorganisms inhabiting the compost. The seeding of compost products at the beginning of sewage sludge composting is unlikely to have any appreciable effect on the rate of composting or on the quality of the final product. However, the contribution of thermophilic bacteria and thermophilic actinomycetes to the CO2 evaluation rate depends heavily on the amount of inoculum added. In the fertilizer N efficiency, Azolla and Azolla compost score 3 kg rice per kg N applied as compared with 10 for urea and supergranulated urea (Table 11). However, 1/2 Azolla + 1/2 the recommended rate for urea increased fertilizer N efficiency to 9, which is comparable to that of urea (dela Cruz 1994).

7.8 Odor control


7.8.1 Ventilation
Many of the materials used for composting are liable to become putrid and stinky. These materials are readily decomposed by microorganisms. Ventilation should provide enough air to avert anaerobic conditions during the temperature build-up stage, but should not

53

remove so much heat that temperatures cannot rise to biologically inhibitive levels. Vacuum has been used as an odor-control measure, so that the exhaust gas can be vented through a scrubber pile. This involves placing a water condenser trap between the composting mass and the scrubber pile. In tests which applied this method, materials composted at the highest temperature gave off the most unpleasant odor, while materials composted at the lowest temperature were judged the least unpleasant. Temperatures which maximized the decomposition rate gave a final product with the least odor.

7.8.2 Biofiltration
Biofiltration involves passing an odorous airflow through a layer of filter material (compost, filamentous peat, etc.), followed by the biodegradation of the captured odor components. The odor components are transferred from the gas to the liquid and solid particles in the filter material. The microbial degradation of the odor components then takes place on these particles. The pressure drop across the biofilters depends not only on the applied air load, but also on the nature and the composition of the filter material.

7.9 Evaluation of compost maturity


Compost maturity is important in assessing both the quality of the end product and its possible uses. Some uses require very mature compost (horticulture, market gardening, plant nurseries, etc.). Others require fresh compost (hot beds in the greenhouse and for mushroom production, etc.). Fresh compost generates heat, while mature compost prevents any adverse effects on plants from the heat of decomposition, a high C:N ratio, or phytotoxic compounds. Several parameters are available which indicate compost maturity. These include temperature; odor; texture; C:N ratio; pH value; gas production; cation exchange capacity; level of ammonium, nitrate, and immobilized nitrogen; total organic carbon; level of hydrogen sulfide; polysaccharide; adenosine triphosphate; chromatographic tests; colorimetric tests (after extraction of humic components); polymerization of humic substances; hydrolase activity; respiratory activity test; behavior of earthworms; and phytotoxicity tests (germination and growth tests) (Table 12). Although many methods have been proposed, none of these have been pursued far enough to allow full appreciation of their potential value.

only 12-70 L, the maximum temperature was only 56o58oC, and temperatures higher than 50oC lasted only two to six days. A temperature higher than 70oC does not favor the growth of fungi in the compost and reduces the rate of decomposition. Therefore, the volume of the different substrates used for composting should be adjusted to give a maximum temperature of around 70oC and temperatures higher than 50oC for more than two weeks. With appropriate aeration, the temperature of swine manure increased on the second day and remained at 50o-70oC for more than two weeks. After 17 days of composting, the volume of the substrate had fallen by 30%. If the compost was not aerated, the maximum temperature was lower than 60oC, and temperatures mainly ranged from 40o to 50oC. Temperatures in the upper layer were higher than those in the middle and bottom layers. Temperature fluctuations in different parts of the substrate were not significantly different in either case. The maximum temperature of pig farm sludge during composting exceeded 68oC. After 30 days incubation, the fermented sludge had become goodquality compost (Fu and Chen 1990). The maximum temperature of a mixture of bagasse and swine manure was 63oC, while temperatures remained between 50o and 60oC for more than one week. In the case of a mixture of bagasse and alcoholic slops, the maximum temperature was 60oC, while temperatures remained above 50oC for only two or three days (Huang et al. 1994). Straw undergoes rapid decomposition when temperatures reach around 70oC, although not as rapidly as animal wastes. When straw had been chopped and placed in a pile for 55 days, temperatures reached their maximum level on the third day, and remained at 60o-65oC during incubation. Temperatures rose again when the substrate was turned, but over time, the maximum temperature gradually decreased. When pig farm sludge was composted with rice hull, the temperature increased to 62oC on the second day, eventually rising to a maximum of 68oC. When the same sludge was mixed with composted pig manure and made into compost, the maximum temperature was 74oC on the second day. In both cases, temperatures remained above 55oC for two weeks. The temperature of compost with mushroom growth medium and chicken manure, soybean residues, bone meal, swine manure, urea, or calcium superphosphate is 50o-60oC for three weeks during incubation. Turning the compost will stimulate the decomposition of organic matter and produce heat from the fermentation (Lin 1994).

7.9.1 Temperature
A temperature rise is necessary not only to accelerate the decomposition of organic constituents, but also to inactivate harmful organisms. Swine manure reached a maximum temperature of 74oC and had an average temperature of 50o-70oC over more than two weeks in a container with a volume of 120 L. In a container of

7.9.2 Odor
Mature composts should smell like forest soil (typical soil odor is caused by actinomycetes). Soil smell is primarily the result of two gases, geosmin and 2methylisobornol, which are by-products of fungi and actinomycetes. If these two gases are present in

54

Table 12. Evaluation of the change in composting.


Parameter Fermentation condition Temperature pH Thermogram Microscopy Direct count Image analysis Microbes Composition COD Soluble COD Soluble BOD C:N ratio Initial N/Final N TOC/TON in aqueous Immobilized nitrogen Ammonia ATP Ash Organic acid Cellulose, hemicellulose Reactive-C Microbial or enzyme activity Respiration rate Condition for stability Stable Alkali (anaerobic, 55oC, 24 h) Stable Biomass Biomass, residue Decrease, stable (thermophilic) COD < 700 mg/g dry compost Stable Stable < 20 < 0.75 5-6 <1.56% (dry weight basis) Absence Decrease, then stable Increase, then stable Stable Stable Stable < 10 mg CO2/g compost (7 days) < 7.5 mg CO2/g compost (7 days) < 2 mg CO2/g compost/d (very stable) 2 mg-5 mg CO2/g compost/d (stable) 5 mg-10 mg CO2/g compost/d (moderately stable) 0 mg-0.5 mg O2/g VS/h (very stable) 0.5 mg-1.0 mg O2/g VS/h (stable) 1.0 mg-1.5 mg O2/g VS/h (moderately stable) Decrease, then stable Darkish brown, 1<Y<13 Earthy Present Stable Nitrification start Disappear Disappear 30 mg-50 mg glucosides/g weight < 35% I. D. *<2.4 stable, while I. D.< 27 unstable >5% > 110 mg/g total organic substance Dark in center, light in surround, and irregular margin Absorbance of alkali extract is stable High molecular weight Correlating to total carbon, total ash, nitrogen, cellulose, hemicellulose, CEC, and lignin contents, and germination index > 60 cmol(+)/kg ash free matter Increase very slow and stable G. I. **>50 Reference Stickelberger 1975 Jann et al. 1959 Owa 1994 Jone and Mollison 1948, 1990 Citenesi and de Bertoldi 1979 Lossin 1971 Yang et al. 1993 Yang et al. 1993 Juste 1980 Juste 1980 Chanyasak et al. 1982 Wang 1978 Spohn 1978 Colin 1977 Yang et al. 1993 Owa 1994 Inoko et al. 1979 Zhang et al. 1992 Morel et al. 1979 Germon et al. 1980 Epstein 1997

Methane emission Color Odor Geosmin, 2-methylisoborneol Headspace gas NO2-/NO3Reductant Organic acid Sugar Easy hydrolyze polysaccharides Reducing sugar/total sugar Total organic carbon-soluble sugar-fermentation time Humic substances Humic carbon/total carbon Alkali soluble humic substance Filter paper method UV spectrophotometry Gel chromatography Near infrared spectroscopy

Chen et al. 2003 Sugahara et al. 1982 Chanyasak et al. 1982 Becker 1995 Wang and Tzeng 1986 Finstein and Miller 1985 Spohn 1978 Chanyasak et al. 1982 Morel et al. 1979 Inoko et al. 1979 Morel et al. 1979 Watanabe and Kurihara 1982 Witt 1982 Hertelendy 1974 Morel 1982 Kubota and Nakasaki 1994 Harada 1995

Cation exchange capacity Hydrolase Bioassay Germination test

Harada et al. 1971 Colin 1977 Zucconi et al. 1981

*I.D. = 3.166-(0.111 AGE)+(0.059 TOC)+(0.832 pHs), AGE=Day of fermentation, TOC=Total organic carbon, PHs=Hot water soluble sugar **G.I. = Germination index.

55

compost, it is possible that they could be used to determine maturity (Becker 1995).

7.9.3 Microcalorimeter analysis


The promotion effect in the decomposition of organic matter was evaluated by analyzing the thermogram. The decomposition of compost was promoted by the addition of microbial materials (Owa 1994).

components. Apart from the loss of total organic carbon and easily degradable components, an increase in highly polymerized organic products may be found. Total organic carbon, C:N ratio, polysaccharide content, and state of humic substances can all be considered suitable measurements of compost maturity. In addition, cation exchange capacity, waterholding capacity, and ash content may be useful parameters for determining maturity.

7.9.4 Biomass production


Three kinds of parameters may be used for determining the maturity of composts: respirometry, analysis of the biodegradable constituents (total organic carbon and polysaccharides), and biochemical activity (ATP and enzyme activity).

1.

Total organic carbon and C:N ratio

1.

Respirometric methods

Respirometry measures the oxygen uptake or carbon dioxide emission, either in pure compost or in compost mixed with soil. It indicates a marked drop in respiratory activity as the compost matures. Compost is considered relatively mature when the respiration rate is less than 5 mg CO2-C/g compost C. Rates over this amount reflect different stages of immaturity. Oxygen uptake and composting time have a significant negative correlation. Oxygen uptake less than 1.0 mg O2/g VS/h is considered relatively mature (Owa 1994; E&A Environmental Consultants 1994). Methane emission from compost below 0.4 mg m-2 h-1 can be used as the index of maturity (Chen et al. 2003).

2.

Degradable organic substances

A number of physico-chemical parameters are useful indicators because of their high correlation with compost degradability. The correlation between the respiration of a soil and compost mixture, and the dry matter production of ryegrass in a similar mixture, allows us to define two classes of maturity. Compost with a value of less than 2.4, corresponding to a low rate of respiration in the soil and compost mixture, may be considered mature. At the other end of the scale, a value of more than 2.7 shows that the compost is still highly degradable and will stunt vegetable growth, so, therefore, is not mature (Morel et al. 1979).

3. Biochemical parameters
The concentration of ATP and the hydrolytic enzyme activity (proteolytic, amylolytic, and cellulolytic activities) can be used as indications of compost maturity.

As composting proceeds, the carbon content falls, while the nitrogen content increases, so that the C:N ratio falls. The C:N ratio changes with the ageing of compost, finally reaching values which are characteristic of a stable organic material. The maturity of compost made from urban refuse could be defined in two classes according to the organic matter/nitrogen ratio: semi-mature urban compost (OM:N < 60) and mature urban compost (OM:N < 50). Organic matter content was 20 g/100 g dry matter (Owa 1994). As Juste and Pommel (1977) indicated, it is preferable to keep a constant check on changes in the C:N ratio during fermentation rather than measure these occasionally. The C:N ratio and the final C:N/ initial C:N ratio provide good indications of the maturity of compost. Swine manure has a low C:N ratio, a high bulk density, and poor aeration. The addition of straw and corncob to the swine manure will improve all three factors, while the addition of carbon and phosphate will improve the composting process. The C:N ratio of compost made from chicken manure, swine manure, soybean cake, and mushroom growth waste medium fell from 30 to 15 during composting (Lin et al. 1994). Composted swine manure contained 2.8% N, 1.7% P2O5, 0.23% K2O, 86% organic matter, 349 mg kg-1 Zn, and 140 mg kg-1 Cu (Wang et al. 1992). Compost made from husk and bark has a C:N ratio of around 35, and may have certain quality problems. Although bark contains a high level of N, Ca, and K, it also has a high level of phenolic compounds (Table 2). Tobacco leaves and tobacco factory wastes had a high content of nicotine as well as of N, K, Ca, and Mg. Composting to reduce the nicotine content was therefore necessary (Hsieh and Hsu 1993). The bulk density and water-holding capacity, and the levels of ash, total N, and 1.0 N HCI insoluble nitrogen increased in the course of composting corncob, while the total C and C:N ratio decreased (Chung et al. 1993). The bulk density of soil amendments and composts was 20 g/100 g dry matter (Owa 1994).

7.9.5 Chemical analysis


During composting, the organic constituents undergo transformations, which lead to biologically more stable

2.

Nitrate content

In the earlier stages of composting, ammonium is produced by the decomposition of nitrogenous compounds such as protein. As the compost matures,

56

the ammonium is oxidized into nitrate by the action of ammonium-oxidizing bacteria and nitrate-oxidizing bacteria. Consequently, nitrate accumulates in the mature compost. The presence of nitrate can be detected with diphenylamine. A diphenylamine solution dissolved in concentrated sulfuric acid is added to water extracted from the compost. If nitrate is contained in the extract, the solution turns blue. This method can be used to test the maturity of cattle manure compost but not of swine manure compost and poultry manure compost, which produce only a very small quantity of nitrate even when they are mature.

Wt = Wo (1 - Ao%/At%) where Wt = weight of substrate at time t Wo = weight of substrate at time zero At = ash content of substrate at time t Ao = ash content of substrate at time zero (1 - Ao%/At%) = the degree of maturity

6.

Cation exchange capacity (CEC)

3.

Organic matter content

During composting, the organic matter content decreases, while the ash content increases. The organic matter content remains constant when the compost is mature and stable. Swine manure and cattle manure take three to four weeks to become stable, while stabilization of chicken manure takes two weeks. The decomposition of organic matter during the composting process is characterized by the changes in residual rate (i.e., the percentage of organic matter which remains compared with the original amount). Poultry wastes are more easily decomposed than those of cattle and pig. The decomposition rate of cattle wastes is similar to that of pig wastes.

The negatively charged CEC of organic matter increases as compost matures. Conventional CEC tests used for soil are not suitable for compost samples. Harada (1995) had developed a suitable and simple method to determine the CEC of compost. The CEC of composted cattle manure increased to 110 cmol(+)/kg over 4-5 weeks, and thereafter remained constant. A highly significant correlation was observed between the CEC and the C:N ratio (r = -0.992), total carbon (r = -0.968), total nitrogen (r = 0.995), and ash content (r = 0.992) in composted cattle wastes. Thus, since the CEC reflects the changes in the constituents during maturation, it is a useful parameter for estimating the degree of maturity of the compost. The CEC of organic amendments and composts was <50 cmol(+)/kg dry matter (Owa 1994).

7.

Organic acid generation

4. Polysaccharides
When composting begins, the simple polysaccharide content of urban refuse is high (20% of the total organic material). However, this falls to only 410% after 240 days. Water-soluble polysaccharides undergo a similar change. Simple polysaccharide components, therefore, are progressively decomposed by microflora during the different stages of thermogenesis and maturation. Water-soluble sugars, consisting mainly of mono- or di-saccharides, disappear much faster than hydrolysable sugars. Genuine polysaccharides, however, take far longer to decompose because of their structure. At the end of the composting period, the quantity of polysaccharides extractable in acid is relatively high. This seems to correspond with newly synthesized stabilized microbial components rather than fractions, which are not yet decomposed.

Acetic acid, propionic acid, and butyric acid generated with decomposition of organic matter were measured. The concentration of acetic acid in soil solution was increased by the addition of biological soil amendments and composts.

8.

pH

Changes in pH have been noted to occur during the composting period and, therefore, have been considered as a possible indicator of biological activity. Generally, the pH drops during the very early stages of composting and then increases to a range of 6.5-7.5. Acid pH values indicate a lack of maturity due to short composting time or occurrence of anaerobic conditions.

9.

Organic chemical constituents

5.

Ash content

The ash content of the substrate is a constant value during composting. Organic matter decomposition leads to a relative increase in the ash content. Therefore, the ash content of the substrate in relation to the total weight of the substrate can be used as a parameter of the maturity of the compost. Compost maturity based on ash content is calculated as follows:

Cellulose content yields a good index of the degree of maturity of the compost. Cellulose decreased with duration of composting. Inoko et al. (1979) reported that hemicellulose and cellulose (reducing sugars) decreased from about 36% of the total dry weight to about 20% after 60 days. Amino acids and low fatty acids greatly decreased during the composting of refuse and garbage. Zhang et al. (1992) suggested that reactive carbon decreased from 33 mg/ 100 mg sample to about 11 mg/100 mg for 160 days of composting.

57

7.9.6 Chromatography test


The chromatography test uses humic substances, which are extracted from the compost and isolated using a filter paper. The slightly polymerized components move towards the periphery of the filter paper, while the highly polymerized components stay at the center. Chromatograms are read by the shape and color of three separate zones (Table 13). This technique is essentially qualitative, and gives a quick indication of the state of maturity of the compost. However, the interpretation of chromatograms may sometimes be difficult. It is quite simple to classify compost from the same plant, so the chromatography test is very suitable for compost control when the raw material is known to be homogeneous. Paper chromatograms showed that the time needed for compost to reach maturity was around 10 days in the case of chicken manure, 15 days for swine manure, and 14-21 days for sludge from pig farms (Lin 1994). Gel chromatography, monitored by UV absorbance at 280 nm of the water extracted from solid samples during composting, can also be used for the evaluation of maturity. Compounds with a relatively high molecular weight correspond to the peak with the least-elution volume. During the composting of sludge, organic compounds in the water extract changed to those with a high molecular weight (Fig. 1). Gel chromatograms, using Sephadex G-15 gel and a compost of mixed urban waste and bark, gave different results. The UV 280 nm absorbance of the relatively high molecular weight compounds, corresponding to the peak with the least-elution volume, increased considerably in one-week-old compost, indicating that the organic compounds in the water extract had shifted to a higher molecular weight. On the other hand, the differences between compost samples after 1 week, 2 months, and 3 months of composting were small (Kubota and Nakasaki 1994).

7.9.8 Colorimetric method


Mature composts should be dark-brown to black regardless of the feedstock. Alkaline extraction of the organic material from compost has a brown hue in a range of shades, depending on the age of the compost. The hue is light at the beginning of composting, and tends to darken as the compost matures. These color variations provide the basis for a test to determine maturity, using periodic measurements of the optical density of alkaline extracts from the product. This technique does not require sophisticated equipment, and can easily be used by composting plants. The compost has reached maturity when the optical density of the alkaline extract remains a stable tint. The 2N KCl extract of composts at different stages was eluted with TSK HW- 55(F). The absorbance at wavelengths of 235 nm, 280 nm, and 490 nm is shown in Fig. 2. The absorbance of raw substrate was very low in all tests, while the absorbance of compost after 17 days incubation at wavelengths of 235 nm and 280 nm was high. The molecular weight of composted pig manure was around 100,000, while it was only around 10,000 for the control. Molecular weight with the absorption at wavelengths of 235 nm and 280 nm increased as composting continued. There was no significant difference between a supplement of 5% sludge in swine manure, and one of 10%. The odor was the same in both treatments. The color changed from yellowish to darkish brown after 40 days incubation. The absorbance at wavelength 490 nm represents the humic substances. The absorption of the extract of compost at 490 nm was very weak before 17 days incubation, while the absorption peak was high after one-month incubation. Therefore, absorption at a wavelength of 490 nm can be used as a parameter for the presence of humic substances and the degree of stability of compost (Sheen and Wang 1994).

7.9.7 Humification parameters


Humification ratio was the percentage of total extractable humic-C as related to the total organic-C. Inbar (1979) reported that the total content of humic material extracted from separated cattle manure compost increased from 377 g/kg to 710 g/kg organic matter. Humic material increased rapidly during the first 60 days, and from 60 to 140 days, a very gradual change was noted.

7.9.9 Near infrared spectroscopy analysis


A rapid estimate of the quality of cattle waste compost can be obtained by using near infrared spectroscopy analysis. The levels of total C, total N, ash, cellulose, hemicellulose, and lignin of the compost, and its CEC, can be measured by this method. Its biological activity, in terms of inhibition of seed germination, was assayed using Brassica rapa seeds, and an attempt was made to estimate this by near infrared spectroscopy.

Table 13. Chromatograms of organic wastes during composting.


Chromatogram Center Transition zone Periphery Source: Morel et al. 1985 Fresh compost White to pink Rings Brown Mature compost Red to violet Irregular outline Clear with jagged edges

58

Fig. 1. Change in gel chromatograms for urban waste. U-1: Inlet of composting chamber; raw materials were bark and garbage in 1:5 ratio (volume); U-2: Outlet of the composting chamber after one week in the chamber; U-3: Second stage fermentation after about 2 months under windrow conditions; and U-4: Second stage fermentation at about 3 months (Kubota and Nakasaki 1994).

Fig. 2. Absorption spectra of the extract of compost at wavelength 280 nm: a) Pig waste and 10% sewage sludge; b) Pig wastes, 5% sewage sludge and 5% water; c) Pig wastes and 10% water (Sheen and Wang 1994).

59

Although the accuracy of the estimation was not high, at least severely inhibitory activity could be detected by near infrared spectroscopy (Harada 1995).

7.9.10 Volatile gases


During the composting of swine manure, the concentration of volatile gases is highest in the upper part of the pile and lowest at the bottom of the pile. After 40 days incubation, the gas concentration remained constant, showing that the compost had stabilized (Sheen and Wang 1994).

7.9.11 Phytotoxicity test


This test shows the presence of phytotoxic properties in composted urban wastes. Two kinds of test plants, maize and bean, are grown in the compost, and their growth compared with that of plants growing in nonenriched peat with a pH raised by the addition of calcium carbonate. E&A Environmental Consultants, Inc. (1990) conducted a study using compost produced from municipal biosolids with eucalyptus or sawdust. Rooted plant cuttings of Pittosporium tovira were transplanted into pots containing compost soil mixtures.

process, the salinity of compost made from wood shavings, and composted urban wastes and grape marc (skins and seeds left after pressing grapes for wine) compost was very high. Levels diminished progressively as composting proceeded. The level of electrical conductivity remained stable at a very low level in sewage sludge and bark compost. Changes in the conductivity do not always follow the same pattern in different piles, but in cases of excessive salinity at the beginning of composting, it defines the point after which the compost may be used. This parameter, which is easy to calculate, seems therefore worth taking into account during the preparation of compost. The electrical conductivity and cation exchange capacity of sewage sludge compost increased and the pH gradually decreased during incubation (Chung et al. 1992).

7.10 Conclusions
It can be seen that there are many methods available for assessing compost maturity. Some of these, such as the measurements of biomass activity, require complex laboratory equipment and are not suitable for small compost plants or farms. Others, such as the seed germination tests, are easily carried out with simple equipment and a minimum of scientific training. A major problem in the quality control of composts is the range of variability in the raw materials. Many of the testing methods available are suited to some materials, but not to others. Very few, if any, of the testing techniques described can be applied with equal success to the whole range of materials used in modern commercial compost production.

7.9.12 Germination test


The analysis of metabolic toxicity may supply a wide range of information on the stabilization process, thus providing a useful instrument for industrial, agricultural, and environmental analyses. Germination tests are a simple and rapid type of bioassay, which can be quickly carried out with simple instrumentation. Cress (Lepidium sativum L.) is often selected as the test seed because of its rapid response (24 h). According to one study, maximum sensitivity is at 27oC and after 24 h incubation. This index has proved to be a most sensitive parameter, able to account for both low toxicity (10-7M, or 25 ppb, abscisic acid) which affects root growth, and toxicity which affects germination (Zucconi et al. 1985). Germination of cabbage seed was 50-80% in a composted mushroom growth medium, rising to 8090% when the medium was composted with a mixture of chicken manure, swine manure, and soybean residues (Lin et al. 1994). The germination rate of cabbage grown in composted swine wastes was 88% (Fu and Chen 1990). Germination of radish seed on chicken manure before composting was 4.2%, rising to 91.6% after 10 days of composting, and 95.6% after 32 days of composting. This shows that composted animal wastes or crop residues can be used directly in the field without any damage to growth.

7.11 References
Becker, H. 1995. Good earth. Agriculture Research 42(6): 16. Chang, C. H., C.Y. Hsieh, and S.S. Yang. 2001. Effect of cultural media on the phosphate-solubilizing activity of thermo-tolerant bacteria. Journal of the Biomass Energy Society of China 20: 79-90. Chang, F. P. and C.C. Young. 1992. Effects of VA mycorrhizal fungi and phosphprus-solubilizing bacteria inoculated on growth of tea cuttings in plastic bag. Taiwan Tea Research Bulletin 11: 7989. Chen, I. C., C.H. Chang and S.S. Yang. 2003. Greenhouse gases emissions and microbial populations during the composting. pp. 157-178. In: Yang, S. S. (ed.). Flux and Mitigation of Greenhouse Gases Emissions (IV). Global Change Research Center and Department of Agricultural Chemistry of National Taiwan University, and Institute of Biotechnology of National Pingtung University of Science and Technology, Taiwan. Chien, S. Y.. 2001. Using microorganisms to promote compost quality. In: Proceedings of the

7.9.13 Conductivity
The level of electrical conductivity shows the overall salinity of the substrate. Early in the composting

60

International Workshop on Recent Technologies of Composting and Their Application. Food and Fertilizer Technology Center for the Asian and Pacific Region and Department of Soil and Environmental Sciences of National Chung-Hsing University, Taiwan. pp. 11-111-22. Chung, R. S., M.Y. Yap, H.C. Lin and T.C. Chang. 1993. Effect of composting of corncob on the availability of rock phosphate and the physicochemical properties of the composts. Journal of the Chinese Agricultural Chemical Society 31: 220-228. Chung, R. S., S.Y. Yuan, Y.S. Wang and H.C. Lin. 1992. Fertilizer characteristics of cabbage and city refuse compost. Journal of the Biomass Energy of China 11: 21-30. Colacicco, D.. 1982. Economic aspects of composting. BioCycle 23(5): 26-30. Daudin, D. and Michelot, P. 1985. Experimental use of organic by-products as culture substrates. In: Gasser, J.K.R. (ed.). Composting of Agricultural and Other Wastes. Elsevier Applied Science Publishers, London, pp. 216-226. dela Cruz, R. E.. 1994. Status of microbial fertilizers in the Philippines. In: International Seminar on the Use of Micro Organic Fertilizers for Agricultural Production. Rural Development Administration of Korea and Food and Fertilizer Technology Center, Suwon, Korea. pp. 279-306. E & A Environmental Consultants, Inc. 1990. Results of a growth trial examining the pytotoxicity of eucalyptus-swedge sludge compost and sawdustsewage sludge. Report to Las Virgenes Municipal WaterDistrict. Las Virgenes, CA. E & A Environmental Consultants, Inc. 1994. In-house data. Espiritu, B. M. and dela Torre, J. C. 2001. Combination of microbial inoculants and compost in rice production. pp. 12-112-16. In: Proceedings of the International Workshop on Recent Technologies of Composting and Their Application. Food and Fertilizer Technology Center for the Asian and Pacific Region and Department of Soil and Environmental Sciences of National Chung-Hsing University, Taiwan. Evangelista, P. 2001. Organic-based fertilization for rice production in the Philippines. pp. 9-19-21. In: Proceedings of the International Workshop on Recent Technologies of Composting and Their Application. Food and Fertilizer Technology Center for the Asian and Pacific Region and Department of Soil and Environmental Sciences of National Chung-Hsing University, Taiwan. Fu, C. M. and Chen, S. Y. 1990. Characteristic and utilization of excess sludge compost on the pig farm. Journal of the Biomass Energy Society of China 9: 137-143. Hara, M. 2001. Improvement of physical and chemical properties of composts and its application. pp. 1519-13. In: Proceedings of the International

Workshop on Recent Technologies of Composting and Their Application. Food and Fertilizer Technology Center for the Asian and Pacific Region and Department of Soil and Environmental Sciences of National Chung-Hsing University, Taiwan. Harada, Y. 1995. The Composting of Animal Wastes. Extension Bulletin No. 408, Food and Fertilizer Technology Center for the Asian and Pacific Region, Taipei, Taiwan. Hsieh, C. F. and K.N. Hsu. 1993. A survey on the mineral nutrient content of the different kinds of organic material in Taiwan. In: Research Report No. 302. Taichung District Agricultural Improvement Station, Changhua, Taiwan. Huang, C. M., W.C. Chen, H.K. Sheen, S.J. Chang, S.W. Lee and L.H. Wang. 1994. Mass production of bagasse compost in an economic way. pp. 122130. In: Proceedings of Symposium on Technology and Utilization of Composting. The Biomass Energy Society of China. Taipei, Taiwan. Huang, C. M., Lee, S. W. and Chen, W. C. 1991. New method for hog waste composting. pp. 77-90. In: Proceedings of Symposium on Pig Waste Treatment and Composting. II. The Biomass Energy Society of China. Taipei, Taiwan. Huang, S. N. and Lin, C. C. 2001. Current status of organic materials recycling in southern Taiwan. pp. 14-19-25. In: Proceedings of the International Workshop on Recent Technologies of Composting and Their Application. Food and Fertilizer Technology Center for the Asian and Pacific Region and Department of Soil and Environmental Sciences of National Chung-Hsing University, Taiwan. Husted, S. 1994. Seasonal variation in methane emission from stored slurry and solid wastes. Journal of Environmental Quality 23: 585-592. Inbar, Y. 1979. Formation of humic substances during the composting of agricultural wastes and characterization of their physicochemical properties. Ph. D. Dissertation. Hebrew University of Jerusalem, Isreal. Inoko, A., K. Miyamatsu, K. Sugahara and Y. Harada. 1979. On some organic constituents of city refuse composts produced in Japan. Soil Science and Plant Nutrition 25: 225-234. Iswandi, A.. 1994a. Biofertilizers and their application in Indonesia. pp. 85-116. In: International Seminar on the Use of Micro Organic Fertilizers for Agricultural Production. Rural Development Administration of Korea and Food and Fertilizer Technology Center, Suwon, Korea. Iswandi, A.. 1994b. Potential of municipal solid waste compost for agriculture in Indonesia. pp. 117-126. In: International Seminar on the Use of Micro Organic Fertilizers for Agricultural Production. Rural Development Administration of Korea and Food and Fertilizer Technology Center, Suwon, Korea.

61

Juste, C. and B. Pommel. 1977. La Valorisation Agricole de Dechets: 1. Le Compost Urbain. Misistre de la Culture et de lEnvironnement, Ministre de lAgricole, France. pp. 75. Kang, U. G. and Ha, H. S. 1994. Selection and use of effective strains of Rhizobium in Korea. pp. 127145. In: International Seminar on the Use of Micro Organic Fertilizers for Agricultural Production. Rural Development Administration of Korea and Food and Fertilizer Technology Center, Suwon, Korea. Kloepper, J. W., Lifshitz, R. and Zablotowicz, R. M. 1989. Free-living bacterial inocula for enhancing crop productivity. Trends in Biotechnology 7: 3944. Knoll, K. H. (1961). Public health and refuse disposal. Compost Science 2: 35-40. Kubota, H. and Nakasaki, K. 1994. Our composting research since 1975: Composting reaction and product evalution. pp. 1-34. In: Proceedings of Symposium on Technology and Utilization of Composting. The Biomass Energy Society of China. Taipei, Taiwan. Lai, C. M., Feng, C. F. and Yang, S. S. 2003. Flux and mitigation of nitrous oxide from composting of livestock wastes. pp. 179-198. In: Yang, S. S. (ed.) Flux and Mitigation of Greenhouse Gases Emissions (IV). Global Change Research Center and Department of Agricultural Chemistry of National Taiwan University, and Institute of Biotechnology of National Pingtung University of Science and Technology, Taiwan. Lian, S. and Lee, Y. C. 1994. Metal contents of organic manures in Taiwan and the current criteria of regulation. pp. 158-173. In: Proceedings of the Symposium on Soil and Fertilizer Pollution. Chinese Soil and Fertilizer Society, Taichung, Taiwan. Lin, J. K. 1994. Effect of hog waste compost on the soil microflora and activity in different environmental conditions. In: Final Report of Soil and Fertilizer Research 1993. Department of Agriculture and Forest, Taiwan Provincial Government, Nan-tou, Taiwan. Morel, J. L., Jacquin, F., Guckert, A. and Barthel, C. 1979. Contribution a la determination de tests de la maturite des compost urbains. Compte-rendu de fin de contrat no 75 124. Ministere de lEnvironnement et du Cadre de Vie. ENSAIA. Nancy, France. pp. 32. Nitta, T. 1994. Roles of organic fertilizers in promoting soil microbiological fertility. pp. 69-84. In: International Seminar on the Use of Micro Organic Fertilizers for Agricultural Production. Rural Development Administration of Korea and Food and Fertilizer Technology Center, Suwon, Korea. Osada, T. 2001. Gas emission from composting animal waste. pp. 4-14-15. In: Proceedings of the International Workshop on Recent Technologies of Composting and Their Application. Food and

Fertilizer Technology Center for the Asian and Pacific Region and Department of Soil and Environmental Sciences of National Chung-Hsing University, Taiwan. Osada, T., Kuroda, K. and Yonada, M. 2000. Nitrous oxide, methane and ammonia emissions from composting process of swine waste. The Japanese Society of Wate Management Experts 2: 51-56. Owa, N. 1994. Microbial material and soil amendment in Japan. pp. 167-174. In: International Seminar on the Use of Micro Organic Fertilizers for Agricultural Production. Rural Development Administration of Korea and Food and Fertilizer Technology Center, Suwon, Korea. Parr, J. F. and Hornick, S. B. 1992. Utilization of municipal wastes. In: Metting, F. B. (ed.) Soil Microbial Ecology: Applications in Agricultural and Environmental Management. Marcel Dekker Inc., New York. pp. 545-559. Parr, J. F., Hornick, S. B. and Kaufman, D. D. 1994. Use of microbial inoculants and organic fertilizers in agricultural production. pp. 1-40. In: International Seminar on the Use of Micro Organic Fertilizers for Agricultural Production. Rural Development Administration of Korea and Food and Fertilizer Technology Center, Suwon, Korea. Prihatini, T. and Kurnia, U. 2001. Organic waste management by means of environment friendly technology. pp. 6-16-13. In: Proceedings of the International Workshop on Recent Technologies of Composting and Their Application. Food and Fertilizer Technology Center for the Asian and Pacific Region and Department of Soil and Environmental Sciences of National Chung-Hsing University, Taiwan. Shamsuddin, Z. H. 1994. Application of microbial inoculants for crop production. pp. 219-231. In: International Seminar on the Use of Micro Organic Fertilizers for Agricultural Production. Rural Development Administration of Korea and Food and Fertilizer Technology Center, Suwon, Korea. Sheen, S. Y. and Wang, H. H. 1994. Study on stabilization parameters of hog waste compost. pp. 1-34. In: Proceedings of Symposium on Technology and Utilization of Composting. TheBiomass Energy Society of China. Taipei, Taiwan. Sinaga, M. S. 1992. Evaluation of compost to control crop diseases. Agricultural Potential of Citywaste Compost. Research Collaboration between Bogor Agricultural University with Center for Policy and Implementation Studies, Jakarta. Sommer, S. G., Sibbesen, E., Nielsen, T., Schjoerring, J. K. and Olsen, J. E. 1996. A passive flux samplers for ammonia volatilization from manure storage facilities. Journal of Environmental Quality 25: 241-247. Tang, C. C. and Wang, Y. P. 2000. Methane emission and mitigation from the composting and

62

application of livestock wastes. pp. 95-109. In: Yang, S. S. (ed.) Flux and Mitigation of Greenhouse Gases Emissions (II). Global Change Research Center, Department of Agricultural Chemistry and Agricultural Exhibition Hall of National Taiwan University, Taipei, Taiwan. Tsai, H. S. and S.S. Yang. 2004. Microbial conversion of food waste for biofertilizer preparation with thermophilic and lipolytic microbes. Renewable Energy (in press). Tseng, C. N., J.J. Hua and C.C. Young. 1993. Studies on gentic relationships among isolates of Rhizopus from Taiwan by Multilocus enzyme electrophoresis. Journal of Chinese Agricultural Chemical Society 31: 120-129. Um, M. H. and Lee, Y. 2001. Evaluation of organic wastes for composting and quality control of commercial composts in Korea. pp. 8-18-18. In: Proceeding of the International Workshop on Recent Technologies of Composting and Their Application. Food and Fertilizer Technology Center for the Asian and Pacific Region and Department of Soil and Environmental Sciences of National Chung-Hsing University, Taiwan. Wang, L. F., S.W. Lee, C.M. Huang, Y.D. Chung, S.J. Chang and W.C. Chen. 1992. Secondary treatment of hog waste and manufacture of rapid stabilization compost. pp. 119-127. In: Experiment Report of Taiwan Sugar Company 1991-1992. Research Institute of Taiwan Sugar Company, Tainan, Taiwan. Yang, S. S. 1994. Composts and agricultural production in Taiwan. Soils and Fertilizers in Taiwan, 1994: 29-62.

Yang, S. S. 1997a. Preparation and characterization of compost. Journal of the Biomass Energy of China, 16: 47-62. Yang, S. S. 1997b. Preparation of compost and evaluating its maturity. Food and Fertilizer Technology Center Extension Bulletin 445: 1-23. Yang, S. S., C.B. Wei, K. Koo and S.S. Tsai. 1991. Food and agricultural wastes produced in Taiwan area. Journal of the Biomass Energy Society of China 10: 70-87. Yang, S. S., H.T. Yang, W.R. Chen and C.B. Wei. 2001. Properties of livestock and poultry wastes during composting in the farms of northern Taiwan. Journal of the Chinese Environmental Society 24: 8-25. Young, C. C. 1994. Selection and application of biofertilizers in Taiwan Agriculture. pp. 147-165. In: International Seminar on the Use of Micro Organic Fertilizers for Agricultural Production. Rural Development Administration of Korea and Food and Fertilizer Technology Center, Suwon, Korea. Zhang, L., F. Chung and M.A. Cole. 1992. A simple chemical assay for estimating compost maturity. ASA Annual Meeting, Minneapolis, MN. Zucconi, F., A. Monaco and M. Forte. 1985. Phytotoxins during the stabilization of organic matter. pp. 73-86. In: Gasser, J. K. R. (ed.) Composting of Agricultural and Other Wastes. Elsevier Applied Science Publishers, London.

63

8
8.1 Introduction

Using organic fertilizers


REN-SHIH CHONG
One of the reasons for the unsustainability of cultivated soils is the decline in soil organic matter content. Adequate amount of soil organic matter also greatly reduces the difficulties of good crop production (Figs. 1-4) (Allison 1978). Therefore, restoring and maintaining a high soil organic matter content is the principal strategy for attaining economic progress and improving environmental quality. Increases in soil biomass, biological abundance, and diversity are directly related to increased levels of organic matter and good management practices, which, in turn, positively influence soil structure, nutrient cycling and availability, buffering capacity, and pest and disease control in cultivation systems. There is also a close relationship between the nutrient status of soils and the organic matter content. Researches have shown that under long-term treatments, adding farmyard manure has raised soil fertility and yields to levels greater than those under synthetic fertilizer treatments. In addition to directly supplying nutrients from the mineralization of organic matter, the mechanisms of higher availability of nutrients with soil of higher organic matter contents are multiple. Parsa and Wallace (1979) showed that both dog manure and sewage sludge at lower rates were very effective in correcting the Fe deficiency of sorghum in calcareous soil by significantly increasing the dry matter yield and the uptake of Fe, Zn, Cu, and Mn. Benefits of compost amendments to soil also include pH stabilization and faster water infiltration rate due to enhanced soil aggregation (Stamatoados et al. 1999). Soils applied with compost initially had a lower soil pH than those applied with synthetic fertilizers, but over time soil pH increased to higher levels in soils with compost than those with synthetic fertilizers (Figs. 5 & 6) (Bulluck et al. 2002). The levels of mycorrhizal colonization were greater under organic treatments than under the conventional. Organic matter increased the available phosphorus in the soil through the organic anion, preventing P fixation and replacing the P bound to the soil (Swenson et al. 1949; Nagarajak et al. 1970; Kafkafi et al. 1998).

Sustainability of agriculture has become a major global concern since the 1980s. Soil organic matter is very important in the functions of soil inasmuch as it is a good indicator of soil quality because it mediates many of the chemical, physical, and biological processes controlling the capacity of a soil to perform successfully. A comparison of cultivated and uncultivated soils has demonstrated a reduction in soil organic matter with cultivation (Mann 1986). Soil organic matter properties (e.g., C:N ratio and macroorganic matter) have been proposed as diagnostic criteria for soil health and performance. However, the importance of organic matter to crop production receives less emphasis, and its proper use in soil management is sometimes neglected or even forgotten. Moreover, understanding nutrient supply or agricultural systems is essential for maintaining longterm productivity. Yields of crops grown in organic and conventional production systems can be the same (Drinkwater et al. 1995; Stamatoados et al. 1999). However, agriculture or agroindustries produce high quantities of organic wastes that are typically rich in nutrients, which can well be used in agriculture to conserve nutrients as well as reduce waste discharge and the use of chemical fertilizers.

8.2 Beneficial effects of organic fertilizer


Incorporating moderate amounts of animal manure and other organic materials into the field is an established agricultural practice generally recognized to have beneficial effects on the soils physical, chemical, and microbiological properties. For example, the use of properly composted organic soil amendments has been associated with desirable soil properties. These properties include greater plant water-holding and cation exchange capacity, lower bulk density of soils, and inducer of beneficial microorganisms (Lin et al. 1973; Parr et al. 1986; Chao et al. 1996).

64

Fig. 1.

Vigorous growth of Brassica campertris L. ssp. rapifera var. laciniifolia (Kitam.) under appropriate application of compost.

Fig. 2. Vigorous growth of Lactuca sativa var. cispa L. under appropriate application of compost.

65

Fig. 3.

Vigorous growth of Oryza sativa var. taikeng No. 5 under appropriate application of compost Left: With chemical fertilizer; right: with compost on the basis of two-fold of N as chemical fertilizer.

Fig. 4.

Growth of tea plant (Camellia sinensis L. var. TTES No. 12) in an Oxisol with chemical and organic fertilizers.
66

Fig. 5.

Growth of Lactuca sativa var. cispa L. in an Oxisol (pH 4.39) with different kinds of compost on the basis of the same amount of N. From left to right: Hog dung compost; sawdust-cattle dung compost; sugarcane residue-cattle dung compost; liming with calcium carbonate to pH 6.5 and applied with chemical N, P, and K fertilizer; pea seedling residue-rice hull compost; and control (with only chemical fertilizer).

Fig. 6. Growth of Lactuca sativa L. cv. tsueyhwa in an Oxisol with different rates of compost and lime Top: from left to right: B) control only with chemical N, P, and K fertilizers (final soil pH 3.85); SI, SII, and SIII) rates of compsot applied were 5, 10, and 20 g/kg and with the final soil pH 3.99, 4.32, and 4.44, respectively. Bottom: from left to right: B) control only with chemical N, P, and K fertilizers; LCI, LII, and LIII) rates of calcium carbonate applied were 1, 2, and 4 cmolg/kg with final soil pH (1:1) 4.34, 4.99, and 6.24, respectively.
67

It has been shown that microbial activity and biomass are higher in fields with organic amendments than fields with conventional fertilizers (Drinkwater et al. 1995). Soils with compost application have higher propagule densities of Trichoderma species than soils amended with synthetic fertilizers regardless of their production system history (Bulluck et al. 2002). The supply of organic manure allows the direct uptake by plants of specific chemicals needed for the development of their immune system. Therefore, the application of organic manure also makes a direct contribution to the anti-phytopathogenic potential of soils (Fig. 7). This is particularly important in the case of the fungal damping-off diseases such as Rhizotinin, Fusarium, and Pythium (Lampkin 1990). The most important mechanism is the antagonism of soil microorganisms toward each other, which may take the form of producing toxins and antibiotics, competing for nutrients and energy, and/or parasitism (Lampkin 1990). The buildup of soil organic matter and maintenance of a protective surface cover under organic and minimum tillage systems favor a reduction in soil loss and its associated problems.

8.4 Fundamentals of organic fertilizer application


Just like chemical fertilizers, adequate organic fertilization programs supply the amount of plant nutrients needed to maximize crop production and net return. Essentially, fertilization management makes certain that soil fertility is not a limiting factor in crop production. The major factors that affect the selection of the kind, rate, and placement of organic fertilizers are fertilizer characteristics, crop characteristics, soil characteristics and management, fertilizer placement, and carryover effects. Manure application in excess of crop needs can cause a significant buildup of P, N, other ions, and salts in the soil. Dormar and Chang (1995) showed that cattle feedlot manure application for 20 years resulted in a significant increase in soil P levels (from 9 mg/kg to 1,200 mg/kg). The research of Hao et al. (2003) showed that the high application rate of manure resulted in considerable nitrate N accumulation, reaching 80-100 mg/kg. An ideal agricultural soil can contain as much as 5% organic matter by weight. However, an increase in soil organic matter content increases the risks of nutrient losses to the environment. Organic matter losses can be as great as 1.5% annually (Raymond 1986). Darwish et al. (1995) showed that at least 95% of total applied annual manure over 15 years was degraded. However, soil organic content can still be increased and maintained by the application of compost. Insufficient knowledge of specific fertilizer values and inadequate application rates can result in under or over use (Conacher and Conacher 1998). The ultimate purpose of applying organic fertilizer is to establish the suitable soil organic matter content. High initial applications to build up the organic pool and cut back in subsequent years would be appropriate.

8.3 Adverse effects of organic fertilizer


Some of the organic matter can be added to soil without any risk while others can produce toxic and depressant effects on plants and the microbial community. Among the possible negative effects of compost application to cropland is the potential environmental release of toxic heavy metals into the environment, as mentioned in earlier chapters, and the transfer of these elements from the soil into the food chain (Petrzzelli et al. 1989). Therefore, a prime concern in the use of organic matter is the presence of heavy metals in the amendments. Industrial and municipal wastes used as soil amendments in agricultural lands may result in undesirable levels of heavy metals. More than 10% of fruit and vegetable samples like carrots, potatoes, silverbeet, and safflower monitored in Victoria, Australia, in 1992-1993 contained cadmium residues in excess of maximum permitted concentrations (Conacher and Conacher 1998). Kirchmann and Tengsved (1991) showed that compounds such as phenol, cresol, and dibutyl phthalate were present in significant concentrations in pig slurry and high levels of phenol, cresol, nonylphenol, and di2-ethylhexyl phthalae (DEPH) were found in sewage sludge. Nitrate pollution of surface water and groundwater, especially potable water, is an increasingly serious environmental problem. Although soluble N fertilizers are considered major culprits, compost is a significant contributor to this problem. In the long term, any waste disposal system may not be sustainable because of the harmful accumulation of ions in the soil and/or groundwater.

8.5 Characteristics of organic fertilizer


The requirement for an organic manure to supply inorganic N synchronously with crop demand is in conflict with the requirement for a soil conditioner to provide persistent, stable organic matter. It is important to quantify the effect of compost on 1) the nutritional and 2) the conditioning value of organic materials, to enable one function to be maximized over the other, or a balance to be found between the two functions, according to the requirements of the situation (Robertson and Morgan 1995). Composting is a biochemical process converting various components in organic wastes into relatively stable humus-like substances that can be used as a soil amendment or organic fertilizer (Jeong and Kim 2001). It helps improve the physical and chemical properties of the waste and reduce its phytotoxicity (Marchain et al. 1991). Composting is also considered one of the most suitable ways of disposing of unpleasant wastes and of increasing the amount of

68

organic matter that can be used to restore and preserve the environment (Stentiford 1987). The finished compost was rated as stable with minimum impact on soil C and N dynamics. A good compost should be tolerated readily by growing crops and should not interfere with root growth and development in the way which fresh manure can do. Composted organic materials, therefore, can act as slow-release sources of plant-available N. Therefore, mature compost is the first choice (Fig. 8). Nutrient contents can vary widely according to manure type (Titiloye et al. 1985) or compost materials (Tables 1-4). Organic matters added to soils contain a wide range of C compounds that vary in rates of decomposition. The biological breakdown of the added organic matter depends on the rate of degradation on each of the C-containing materials present in the sample (Reddy et al. 1980). Ajwa and Tabataba (1994) showed that the amount of CO2-C releases increased rapidly initially, but the pattern differed among the organic materials used. Gilbertson et al. (1979) showed that the annual mineralization rate of organic N in animal manure was positively correlated with the N content of waste. Variation in environmental factors, however, may cause a change in the decomposition rates of organic materials in soils. Of these factors, moisture content, temperature, soil pH, aeration and soil structure and texture, agricultural practices (e.g., cultivation), substrate specificity, and available minerals have been reported to be most important (Broadbent et al. 1964; Kowalenko et al. 1978; Clark and Gilmour 1983). Most of the N found in a composting mixture is organic, principally as part of the structure of proteins and simple peptides. The proportion of added organic matter that is mineralized after compost application ranks from several up to a hundred percent, depending on experimental conditions and compost types. Hadas and Pornoy (1994) reported that the mineralization constant for composted manure was commonly 5% to 10% per year. Bitzer and Sims (1988) found that an average of 66% of the organic N in poultry manure was mineralized in the first year. Cabrera et al. (1994) confirmed this rapid mineralization from poultry manure, estimating that 35% to 50% of organic N could be mineralized within 14 days of incorporation into soil. Griffin et al. (2000) reported that the amount of N in manure mineralized in a cropping season varied with the different manures: cattle manure, 25%; dairy manure, 35%; poultry manure, 60%; and swine manure, 50%. Traditionally, manure has been applied to farmlands to increase soil fertility on the basis of crop N requirement. Organic matter applied, therefore, should be calculated based on its mineralization rate. For example, the application of cattle manure is 20,000 kg/ha at a rate of 100 kg N/ha. Compost is a source of fertilizer N in varying degrees. Thus, understanding the factors that control mineralization will make compost more valuable for agricultural and horticultural uses (Sikora and Szmide 2001).

However, owing to the limited soil in a pot culture, the recommended application rate of manure is 3-5% of the soil by weight, depending on the N content of the manure. Chicken manure is one of the most economically efficient types of manure. This efficiency is due to its high pH, low organic C, high inorganic N, and low C:N ratio compared with the other types of manure. In practice, in order to increase soil fertility on the basis of crop N requirement, organic materials of low C:N ratio such as green manure, seed cake, or poultry manure are the better choice. In this purpose, the optimum C:N ratio for finished compost is about 15:1. On the other hand, when the significant organic C accumulation is more important, the organic manure of higher C:N ratio such as bark compost, compost with a higher proportion of woody materials, or cattle manure would be better.

8.6 Crop characteristics and growth stage


The amounts of nutrients, especially N, required by different crops are different. The amounts of nutrients absorbed by a crop also differ depending on its growth stages, owing to differing biomass accumulation. A synchrony between crop demand and nutrient mineralization is desirable. Therefore, in programming fertilization with organic materials for different crops, the N-uptake dynamic is equally or more important than the total N uptake. This is to say that the timing of nutrient release from manure is important and needs to be known for manure of different quality. For annual crops, manure with sufficient nutrients supplied in the short growth stage to meet the crop requirement is essential. Because N is required during the whole growth period of vegetables, manure with high mineralization rate is superior to that of low mineralization rate. On the contrary, for perennial crops, the application of manure is quite different from that of annual crops. Generally, the organic matter content of fruit-cultivated soils is low. Cultivated soils with less than 2% organic matter content are considered unfavorable for sustainable agricultural production. A corrective dressing (applied before planting) in fruit production systems is the application of soil organic amendments in quantities that provide a sufficient reserve of organic matter for several years at the recommended application rates of 50-100 t/ha (Pinamonti and Siche 2001). On the other hand, maintenance dressing at the recommended application rates of 40-60 t/ha every two to three years keeps an adequate vegetative production balance. Manure with low mineralization rates is a better choice for both dressings for perennial crops.

69

Fig. 7. A stable compost resulted in healthy soil; Cucumis melo L. var. chito was not affected by powdery mildew (Sphaerotheca fusca) (right), as compared with the affected plant (left) (courtesy of Dr. C.H. Liao).

Fig. 8. Incorrect application of fresh crop residue to fruit trees.


70

Table 1. Selected chemical composition of some raw plant materials for compost (g/kg).
Source Rice straw Rice husk Rice bran Sorghum straw Corn straw Soybean plant Peanut plant Peanut husk Peanut cake Rape cake Sesame cake Cotton seed cake Cocoa nut cake Soybean cake Sugar cane trash Tea residue Sunflower seed cake Castor seed cake Tea seed cake Saw dust C 540-560 390-520 500 530 550 520 490 530 260-510 480 550 260-540 100 270-510 330-530 460-510 450-460 480-510 400-580 N 6.4-6.9 3.6-7.0 2.04-3.37 7.3 8.1 17.6 17.3 16.0 44.0-75.0 34.0-59.6 50.0-58.0 39.0-58.0 34.0 440-84.6 2.6-10.8 25.2-36.4 46.9-57.9 10.0-17.5 5.6-14.8 C:N ratio 78-88 74-108 22 73 68 30 28 33 5.9-6.8 9 10 6-11 3 4-7 49-126 14-18 52.0 8-10 27-41 27-102 P 0.2-0.5 0.3-2.0 10.7-19.5 1.1 1.6 1.6 1.6 1.5 3.1-12.0 4.2-25.0 9.8-32.0 7.9-11.8 9.9 5.2-7.6 0.3-.09 1.9-2.2 17.0 7.9-8.6 1.1-2.1 1.3-13.4 K 16.6-17.4 2.3-10.8 11.9-20.3 16.7 13.4 10.9 10.5 10.7 7.5-15.4 8.3-14.3 11.1-19.0 9.1-12.9 17.9 16.6-34.9 1.9-4.7 4.1-5.4 14.0 8.3-12.0 7.3-19.0 0.8-10.0 Ca 3.0-8.6 1.1-2.4 1.3-2.5 4.3 2.5 14.1 14.0 4.7 1.7-6.9 3.0-9.8 40.2 3.9-10.0 16.2 1.9-20.7 2.9 3.3-4.4 5.9-15.6 0.1-11.0 3.5-41.3 Mg 1.8-3.1 0.3-2.4 4.5-10.7 3.7 2.9 6.9 4.6 2.0 2.8-4.9 3.5-5.6 6.6 4.6-5.2 2.8 0.2-0.4 0.7 1.6 5.5 1.1-4.0 2.2-5.6

Table 2. Selected chemical compositions of animal dung for compost (g/kg).


Source Cattle dung (dry) Swine dung (dry) Chicken dung (dry) Leghorn dung (dry) Goat dung (dry) C 250-400 40-540 250-470 270-320 360-480 N 18.9-2.35 16.0-47.8 6.0-410 6.0-29.0 16.0-24.0 C/N ratio 15-28 14-31 8-28 9-14 18-23 P 2.1-2.4 4.4-51.1 6.1-30.5 6.1-29.7 6.5-23.0 K 6.1-29.0 1.3-16.0 6.4-35.0 5.9-31.5 15.8-33.2 Ca 1.4-14.3 5.7-10.4 11.2-90.0 5.2-58.6 9.3-38.6 Mg 3.6-12.6 0.9-10.2 3.0-12.1 1.8-10.8 4.2-8.4

Table 3. Selected chemical compositions of some animal source materials for compost (g/kg).
Source Poultry feather Fish bone powder Fish refuse Leather meal Bone meal C 380-390 N 88.5-136.2 28.2-78.8 47.6-7.0 45.0-90.3 83.5 C:N ratio 86-87 P 1.2-2.0 8.6-9.1 18.8-29.8 0.1-4.5 11.5 K 0.4-0.9 8.0-8.3 1.4-7.9 0.2-4.4 3.9 Ca 2.2-4.7 3.9-55.1 1.0-7.7 50.2 Mg 0.3-1.7 2.8-41.2 0.1-2.8 1.8

Table 4. Selected chemical compositions of compost derived from different materials (g/kg).
Compost Cattle dung compost (1) Cattle dung compost (2) Swine dung-saw dust Chicken dung-sawdust Chicken dung Cattle dung-tea residue Molasses compost Pea residue-rice hull Sugar cane residue Sugar cane residue C 104-370 150-290 360-385 280-310 170-500 377-387 224 360 240-348 120-480 N C:N ratio P K Ca Mg

10.2-31.9 10.0-20.0 3.5-10.6 9.9-143 12.0-39.0 4-24 3.5-28.8 3.3-5.0 20.9-70.5 10-16 4.7-7.6 4.9-181 22.5-24.6 11-14 17.4-18.9 2.3-28.8 4.0-57.0 6-94 0.9-91.7 2.5-54.8 32.0-34.2 11-12 4.7-11.9 10.1-16.6 17.5 13 2.6 35.1 280-30.0 12 1.6-4.6 10.1-14.9 16.2-29.0 8-21 2.3-7.0 10.4-35.8 8.5-35.4 9-38 0.8-11.5 3.5-41.0

1.6-2.3 5.3-8.9 13.6-164.3 4.2-22.8 26.1-39.7 4.5-6.1 62.1-68.8 0.9-1.1 13.6-235.7 1.8-30. 17.9 4.9 24.5 3.2 55.1 6.9 14.2-16.0 6.5-8.2

71

8.7 Soil characteristics and management


The properties of the soils significantly affect the mineralization rate of organic N. Under similar climatic conditions, the organic C accumulation varies possibly due to different soil types and the quality and quantity of litter accumulation. The important soil factors that influence the mineralization of compost includes pH, salts, presence of toxic quantities of inorganic or organic compounds, moisture, and temperature (Broadbent et al. 1964; Kowalenko et al. 1978; Clark and Gilmour 1983). The decomposition rates of organic manure are generally slow in clay soil, minimal tillage, sod culture, and organic mulch. The estimated annual dry organic matter decomposition is 3-4 t/ha in cold-temperate regions (Pinamonti and Sichee 2001). On the contrary, fast decomposition of organic manure occurs in sandy and stony soils and in soils with persistent solar radiation. The estimated annual dry organic matter decomposition is 7 t/ha in cold-temperate regions (Pinamonti and Sichee 2001).

an adequate level of soil organic matter, a critical component of soil fertility and productivity. Organic manure is considered as slow-release N fertilizer because it releases or mineralizes only a fraction of its total N content during the application season. High initial applications to build up the organic pool and cut back in subsequent years would be appropriate. In supplying the nutrient requirements, the amount of manure applied can be calculated based on the rate of N applied and the rate of organic N mineralization in the application season.

8.11 References
Allison, __. 1978. Soil organic matter and its role in crop production. Elsevier. Amsterdam, the Netherlands. Ajwa, H. A. and M. A. Tabataba. 1994. Decomposition of different organic materials in soils. Biol. Fertil., Soils, 18: 175-183. Bitzer, C. C. and J. T. Sims. 1988. Estimating the availability of N in poultry manure through laboratory and field studies. J. Environ. Qual., 17:47-54. Broadbent, E. F., R. H. Jackman, and J. McNicoll. 1964. Mineralization of carbon and N in some New Zealand allophonic soils. Soil Sci., 98:118-128. Bulluck, L. R., M. Brosius, G. K. Evanylo and J. B. Ristaino. 2002. Organic and synthetic fertility amendments influence soil microbial, physical and chemical properties on organic and conventional farms. Appl. Soil Ecol., 19: 147-160. Cabrera, M.L., S. C. Tyson, T. R. Kelley, O. C. Pancarbo, W. C. Merka, and S. A. Thompson. 1994. Nitrogen mineralization and ammonia volatilization from fractionated poultry letter. Soil Sci. Soc. Am. J., 58:367-372. Chao, W. L., H. J. Tu, and C. C. Chao. 1996. Nitrogen transformations in tropical soils under conventional and sustainable farming systems. Biol. Fertil. Soils, 21: 252-256. Clark, M. D. and J. T. Gilmour. 1983. The effect of temperature on decomposition at optimum and saturated soil water contents. Soil Sci. Soc. Am. J., 47: 927-927. Conacher, J. and A. Conacher. 1998. Organic farming and the environment, with particular reference to Australia: A review. Biol. Agric. Horti., 16: 145-171. Darwish, O. H., N. Persaud, and D. C. Martens. 1995. Effect of long-term application of animal manure on physical properties of three soils. Plant Soil, 176: 289-295. Dormar, J. F. and C. Chang. 1995. Effect of 20 annual application of excess feedlot manure on labile soil phosphorus. Can. J. Soil Sci., 75: 507-512. Drinkwater, L. E., D. K. Letourneau, F. Workneh, A. C. H.van Bruggen, and C. Shennan. 1995. Fundamental differences between conventional and organic tomato agroecosystems in California. Ecol. Appl., 5: 1098-1112.

8.8 Fertilizer placement


Generally, the organic manure for annual crops is surface-applied and should be mixed with topsoil by conventional tillage. The organic manure for the maintenance dressing of perennial crops is surfaceapplied or placed on the under-row area, with or without mixing with topsoil by conventional tillage. The organic manure can also be applied directly into the planting hole near the root system at the time of planting to improve the edaphic environment of the root system. For pot culture, the organic manure is generally completely mixed with the soil so that the root system can develop vigorously in the whole pot soil.

8.9 Carryover effects


The application of composted manure to provide for crop N requirements may greatly increase the levels of P and other ions in the soil. Long-term or heavymanure application increases microbial activity and the potential for mineralization of soil organic matter. Consequently, it may result in nitrate accumulation deep into the soil that may induce the transformation of soil Po fractions to Pi fractions. Long-term application of organic manure may also result in an imbalance in base composition. It is important to alternate the application of organic manure with different composition or properties.

8.10 Conclusion
The use of organic manure to fertilize agricultural lands is positive from the perspective of a recycling economy. Application of organic matter to soils directly maintains

72

Gilbertson, C. B., F. A. Norstadt, A. C. Mathure, R. F. Holt, A. P. Barnett, T. M. McCalle, C. A. Onstad, and R. A. Young. 1979. Animal waste utilization on cropland and pastureland: a manual for evaluation agronomic and environmental effects. USDA Utilization Research Report No. 6 and EPA600/2-79-059. Washington, DC. Griffin, T. S. and C. W. Honeycutt. 2000. Using growing degree days to predict nitrogen availability from livestock manures. Soil Sci. Soc. Am. J., 64: 1876-1882. Hao, X., C. Chabg, G. R. Travia and F. Zhang. 2003. Soil and carbon and nitrogen response to 25 annual cattle manure applications . J. Plant Nutr. Soil Sci., 166: 239.-245. Hadas, A. and R. Pornoy. 1994. Nitrogen and carbon mineralization rates of composted manures incubated in soils. J. Environ. Qual., 23: 11841189. Jeong, Y. K. and J. S. Kim. 2001. A new method for conservation of nitrogen in aerobic composting processes. Bioresource Technol., 79: 129-133. Kafkafi, U., B. Bar-Yosef, R. Rosenbery, and G. Sposito. 1988. Phosphorus adsorption by kaolinite ad montmorillonite: II. Organic anion competition. Soil Sci. Soc. Am. J., 52: 1585-1589. Kirchmann, H. and A. Tengsved. 1991. Organic pollutants in sewage sludge. Swedish J. Agric. Res., 21: 115-119. Kowalenko, C. G., K. C. Ivarson, and D. R. Cameron. 1978. Effects of moisture content, temperature, and nitrogen fertilization on carbon dioxide evolution from field soils. Soil Biol. Biochem., 10: 417-423. Lampkin, N. 1990. Organic farming. Ipswich, UK: Farming Press. pp. 214-271. Lin, C. F., T. S. Wang, A. H. Chang, and C. Y. Cheng. 1973. Effects of some long-term fertilizer treatments on the chemical properties of soil and yield of rice. J. Taiwan Agric. Res., 22: 241-292. Mann, L. K. 1986. Changes in soil carbon storage after cultivation. Soil Sci., 142: 29-288. Marchain, U., D. Levanon, O. Danai, S. Musaphy. 1991. A suggested solution for slaughter wastes: uses of the residual materials after digestion. Bioresource Technol., 37: 127-134. Nagarajak, S., A. M. Posner, and J. P. Quick. 1970. Competitive adsorption of phosphate with polygalacturonate and other organic anions on kaolinite and oxide surfaces. Nature 228: 83-85. London. Parr, J. F., R. I. Papendick, and D. Colacicco. 1986. Recycling of organic wastes for a sustainable

agriculture. pp. 29-43. In: Lopez-Real J. M. and R. D. Hodges (eds.) The role of microorganisms in a sustainable agriculture. London, UK: Academic Press. Parsa, A. A. and A. Wallace. 1979. Organic solid wastes from urban environment as iron sources for sorghum. Plant Soil, 53: 453-461. Petrzzelli, G., L. Lubrano, and G. Guidi. 1989. Uptake by corn and chemical extractability of heavy metals from a four-year compost treated coil. Plant Soil, 116: 23-27. Pinamonti, F. and L. Siche. 2001. Compost utilization in fruit production systems. pp. 177-200. In: Stoffella P. J. and B. A. Kahn (eds.) Compost utilization in horticultural cropping systems. New York, USA: Lewis Publishers. Raymond P. Poincelot. 1986. Toward a more sustainable agriculture. Westport, Connecticut, USA: Avi Publishing Company, Inc. pp. 116-161. Reddy, K.R., M.R. Overcash, R. Khaleel, and P.W. Westerman. 1980. Phosphorus adsorptiondesorption characteristics of two soils utilized for disposal of animal wastes. J. Environ. Qual. 9:8692. Robertson, F. A. and W. C. Morgan. 1995. Mineralization of C and N in organic materials as affected by duration of composting. Aust. J. Soil Res., 33: 511-524. Sikora, L. J. and R. A. K. Szmide. 2001. Nitrogen sources, mineralization rates, and nitrogen nutrition benefits to plant form compost. pp. 287305. In: P. J. Stoffella and B. A. Kahn (eds.) Compost utilization in horticultural cropping system. New York, USA: Lewis Publishers. Stamatoados, S., M. Werner, and M. Buchanam. 1999. Field assessment of soil quality as affected by compost and fertilizer application in a broccoli field (San Benito County, California). Appl. Soil Ecol., 12: 217-225. Stentiford, E. I. 1987. Recent developments in composting. pp 52-60. In: de Bertildi M., Ferrani, M., LHermite P., and Zucconi F. (eds.) Compost, production, quality and use. Elsevier. London. Swenson, R. M., C. V. Cole, and D. H. Sieling. 1949. Fixation of phosphorus by iron and aluminium and replacement by organic and inorganic anions. Soil Sci., 67: 3-22. Titiloye, E. O., E. O. Lucas, and A. A. Agboola. 1985. Evaluation of fertilizer value of organic waste materials in South Western Nigeria. Biol.Agric. Horti., 3: 25-37.

73

You might also like