Sei sulla pagina 1di 262

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/304957255

Design of anaerobic bio-reactors for the simulation of biogas production from


municipal solid waste

Thesis · January 2006


DOI: 10.13140/RG.2.1.4152.4722

CITATIONS READS
13 446

1 author:

Asinyetogha Hilkiah Igoni


Rivers State University
38 PUBLICATIONS   404 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Exploitation of the Potential of Municipal Solid Waste in Low Income Economies in Sub-Saharan Africa View project

Effect of Raw Sewage on Surface Water Quality in Opobo Town, Nigeria View project

All content following this page was uploaded by Asinyetogha Hilkiah Igoni on 07 July 2016.

The user has requested enhancement of the downloaded file.


CHAPTER 1

INTRODUCTION

1.1 BACKGROUND INFORMATION

In the general study of the behaviour of matter the whole concept of

energy is usually described in relation to the capacity of a person or

machine to do work; and work must continuously be done for the

sustenance of human activities/systems. Therefore, energy is fundamental

to the continued survival of a people, their socio-economic and indeed

political development, and prosperity for future growth; including the rate of

growth of infrastructural facilities, level of agricultural and industrial

production, state and level of delivery of social services, etc. Evidently the

energy requirement of these different societal structures/components is

strictly dependent on population, that is, as population increases more

energy should be available for use. However, the reverse of this expectation

seems the case in Nigeria, as Ilori (2000) states that “in Nigeria, domestic

demand for energy is growing while total domestic production is declining as

a result of poor performance and low output of our refineries and power

generating plants”. This energy demand–supply imbalance (with large

population chasing little amount of energy) has expectedly resulted in very

high cost of energy in the country, which has recently been exacerbated by

the 2003 progressive deregulation policy and the current petroleum tax of

the Obasanjo Government, putting independent marketers at liberty to sell


2

their petroleum (energy) product at whatever price suits them, with the

concomitant distress on both the country and her people. Whereas this

remark is by no means aimed at appraising the merits or otherwise of

government policy on the energy sector, it is intended only to highlight the

price response to an over-dependence on a static supply regime in line with

the trite economic concept of increasing price for a high demand with low

supply.

The energy stress in Nigeria is the result of an over-dependence on

the use of fossil fuels, particularly petroleum products, which, despite that

the country does not even have the relevant and necessary technology and

infrastructure to harness, is also largely depleting and thus classified as

non-renewable. FAO (1997) reports the interview with Gustapo Best, its

Senior Energy Coordinator, where he said that fossil fuels would only last

another 40 – 50 years. The issue of the depletion or exhaustibility of

petroleum products has been contentious and confusing, and for its

relevance in this background presentation, a cursory review is imperative.

One major area of confusion or misunderstanding in the depletion saga is

the issue of ‘reserve’ duration. C. J. Campbell (1995, 1997) cited in Worika

(2001) explains that an R/P (reserve/production) ratio of 42 years does not

mean that current production can be sustained for 42 years because all

oilfields decline as their reserves are depleted and that this depletion rate

varies from field to field. Even though the Russian – Ukrainian theory of

abyssal abiotic energy, opposes the notion of exhaustibility of energy

deposit, against the background that new and better technologies would
3

always facilitate the exploitation of the energy resource, it is indeed not

clear whether their improved technologies would also create the energy

sources. In fact, in this connection Worika (2001) refers to the views of

Mathew Simmons, when he says, “any one who thinks that technology has

won the battle over depletion should review the oilfields in the UK sector of

the North Sea, where every scrap of technology has been used.

When fossil fuels deplete, considering their mode of formation, it

takes a considerable period (in the region of thousands of years) for their

primary source to be replenished, which is quite un-appealing for energy –

thirsty people. This is why there is a shift in focus from the present use of

petroleum products to alternative energy sources that are renewable,

particularly when the environmental pollution effects and unfriendliness of

the mode of exploitation of fossil fuels are added to these other

considerations.

1.2 THE CONCEPT OF BIOGAS

When organic wastes decompose, they do so in the presence or

absence of air (oxygen) referred to as aerobic or anaerobic decomposition

respectively. This decomposition could be naturally occurring or artificially

induced, under controlled conditions. In either case of decomposition there

are several by-products as shown in the chart of Figure 1.1.


4

Organic
Wastes

Anaerobic Decay Aerobic Decay


(without oxygen) (with oxygen)

natural artificial natural artificial

decay airtight plant wastes open compost piles


under guts of digester animal bodies
water animals dung piles

peat manure biogas sludge ammonia humus ammonia


carbon carbon
dioxide dioxide

Figure1.1: End Products of Organic Decay


Adapted from Steadman, 1975

From the chart it is depicted that one of the end products of

anaerobic decay is biogas, which could be produced naturally from decay

under water or the guts of animal, and artificially in an airtight digester. It is

based on this mode of production that biogas has been severally defined.
5

Itodo and Philips, 2001 describe biogas as “a methane-rich gas that

is produced from the anaerobic digestion of organic materials in a biological

– engineering structure called the digester”. This definition simply suggests

that biogas is only produced artificially; but from what we now know, this is

not the case. It is not clear what would have limited the scope of this

definition, as it were, but from the conceptual framework of the study of the

authors it is believed that perhaps their comparison of artificial production

processes may be responsible for ignoring the naturality of the occurrence

of biogas. Itodo & Philips are actually not alone in this style of defining

biogas. GEMEY, 2000 says biogas is “Gas rich in methane, which is

produced by the fermentation of animal dung, human sewage or crop

residues in an air-tight container”. It is, therefore, possible to say that this

style of definition probably derives from the fact that it is through the

controlled (artificial) production that usable biogas is generated. However,

GEMEY’s definition is strictly associated with only a set of organic materials,

which are probably the ones in use for biogas generation at the time of his

definition, not realizing the potency of the organic matter component of

municipal solid waste (MSW).

At this point it is important to state that the decomposition of organic

matter in the absence of air could be elicited by the use of physical or

chemical processes at high temperature and/or pressure, or the use of

microorganisms at low temperature and atmospheric pressure, the

preferred method being dependent on their relative polluting impacts on the

environment. However, despite the method used, gas is produced, but is


6

referred to as biogas if generated as a result of the action of

microorganisms on the organic wastes (Hobson et al, 1981). This is why

biogas is now defined as “a by-product of the biological breakdown, under

oxygen-free conditions of organic wastes such as plants, crop residues,

wood and bark residues, and human and animal manure, --- and is known

by such other names as swamp gas, marsh gas, ‘will o’ the wisp’ and gobar

gas” (Mattocks, 1984); and as Deep Green Energy (SeedTree, 2003), and

digestion gas (Oregon State Department of Energy, OSDE, 2002), and

natural gas (Harris, 2003), and landfill gas (LFG) & sewage gas (Xuereb,

1997).

Biogas is a colourless, relatively odourless and inflammable gas, with

the following composition (Madu and Sodeinde, 2001).

Table 1.1: Composition of Biogas

Constituents % Composition

Methane (CH4) 55 – 75%

Carbon dioxide (CO2) 30 – 45%

Hydrogen Sulphide (H2S) 1 – 2%

Nitrogen (N2) 0 – 1%

Hydrogen (H2) 0 – 1%

Carbon Monoxide (CO) Traces

Oxygen (O2) Traces


7

They explain that biogas burns with a blue flame and has a heat value of

4500 – 5000 kcal/m2 when its methane content is in the range of 60 – 70%.

It is also stable and non-toxic.

1.2.1 Uses of Biogas

"A Car That Runs on Waste


A farm owner in Finland now has a car that runs on the gas
given off by decomposing waste. "The car is fueled by biogas,
which is produced from wastes that are cleaned and pressurized
in a biogas reactor located on the car owner's farm," reports the
Finnish magazine Suomen Iuonto. Biogas is the cleanest-burning
vehicle fuel in use today, and since it can be produced during the
recycling of garbage, it is very ecologically friendly. In fact, one of
the by-products of biogas production is valuable agricultural
fertilizer. Cars already designed to use natural gas - about two
million worldwide - can also run on biogas. In Sweden many city
buses are powered by biogas, and some
gas stations there already offer
biogas in addition to other
fuels. The article notes a
final advantage: " Biogas
is much cheaper than
gasoline or diesel fuel.""

Culled from AWAKE magazine of November 8, 2003

All the world over, biogas has been variously used for heating

purposes and/or electricity generation, for example, in the UK, Xuereb


8

(1997) reported that although the use of biogas for electricity generation

was still at an experimental stage, it already accounted for about 0.5% of

the total electricity output; and biogas fuels about 1% of US electrical

generating, while giving climate change benefit equivalent to reducing CO 2

emissions in the electricity sector by more than 10%. Biogas is also

presently used in India, China, Taiwan, Brazil, Singapore, etc.

Tchobanoglous and Burton (1991) state that in large plants, digester gas

may be used as fuel for boiler and internal combustion engines, which are in

turn used for pumping wastewater, operating blowers, and generating

electricity. Despite the heating and electricity uses of biogas, the residues of

biogas production are used as a low-grade fertilizer.

In addition to being a cheap source of energy, Xuereb (1997)

enumerates other advantages of biogas as follows.

- Biogas is inflammable, and hence potentially explosive (this could

be used for cooking)

- Biogas is one of the ‘greenhouse effect’ producing agents; hence

its use helps to reduce the amount released in the atmosphere.

- Although on burning biogas, carbon dioxide is released in a

similar way to the burning of fossil fuels, the carbon dioxide

released is not considered as a net contributor to the global

carbon dioxide level since it originated from plants, which have

absorbed it from the atmosphere. Hence the carbon dioxide

released also does not contribute to the ‘greenhouse effect’


9

- The use of biogas also helps to minimize the unpleasant

decomposition smells produced in landfills since otherwise these

gases would have been released directly into the atmosphere.

Hence, especially where landfills are situated close to inhabited

areas, the use of LFG makes the landfills slightly more socially

acceptable.

Zenith Corporate Services Private Limited (2003) further expounded the

benefits of using biogas even for cooking as:

- Replaces firewood and kerosene

- Reduces smoke & provides cleaner environment

- Prevents drudgery of women

- Improves rural sanitation

- Helps in minimizing deforestation and improves ecological

balance

1.2.2 Sources of Biogas

In the past, the generation of biogas has been by the use of such

feedstock as “livestock farm waste (various manure, slurry and waste

waters) and agro-industrial waste (abattoirs, wineries, vegetable processing

plants, etc) - - -” (Vassiliou, 1997). This is why biogas is also described as

“the fuel produced through anaerobic fermentation of manure and vegetable

matter in digesters“; the fermentation of animal dung, human sewage or

crop residues in an airtight container (GEMET, 2000). So, the general belief

is that “liquid manure systems work best for anaerobic digestion" in the
10

production of biogas. However, this is hardly the case, excepting that the

generation of biogas was indeed first associated with liquid wastes and

sludge; hence Kiely (1998) explained that “the unit treatment process of

anaerobic digestion is used world-wide for the treatment of industrial,

agricultural and municipal waste waters and sludge”; but also noted that “in

recent years it is also being applied to the treatment of municipal solid

wastes”. Little wonder then that Vassiliou (1997), after successfully

generating biogas from “waste of large livestock farms (raw manure + wash

water), and the waste of the food and drinks industry” explained that the

second stage of the project is to generate biogas from “the organic

component of source-separated municipal solid waste (MSW). What then is

municipal solid waste?

1.3 MUNICIPAL SOLID WASTE DEFINED

Generally, waste is regarded as a useless material that is unwanted

and therefore discarded. The New Edition Concise English Dictionary

(1999) explains that waste is “anything or anyone rejected as useless,

worthless, or in excess of what is required”. But Byrne (1997) was more

generous with his description of waste, when he said that waste is material,

which has no direct value to the producer and so must be disposed of. This

could be why Bailie et al (1997) insist that “for practical purposes, the term

waste includes any material that enters the waste management system”. A

waste management system, according to them, being organized programs

and central facilities established not only for final disposal of waste but also
11

for recycling, reuse, composting and incineration. They say “materials enter

a waste management system when no one who has the opportunity to

retain them wishes to do so”. It is then possible, from this definition, to say

that the material that is disposed by the person to whom it has no direct

value, could be valuable to any other person(s)

Waste is usually classified according to the established three states

of matter as gaseous, liquid, and solid waste in which it occurs. This is why

Sincero and Sincero (1999) explained that a gas that is wasted is a gas

waste, such as polluted air from a process that is vented into the

atmosphere. A liquid that is wasted, such as the polluted water of

wastewater, is called a liquid waste. Finally, a solid that is wasted is a solid

waste. Hence, Ogunbiyi (2001) states that solid waste is a non-fluid type of

waste, which makes its handling and management relatively difficult,

compared to the types of waste that can flow from one location to the other,

or even vaporize. This is in line with the perspective of Bailie et al (1997)

when they said solid waste refers to all waste materials except hazardous

waste, liquid waste and atmospheric emissions. However, Kiely (1998) says

solid wastes are those wastes from human and animal activities including

liquid wastes like paints, old medicines, spent oils etc. Thus, it is therefore

possible to have solid waste intermixed with liquid waste. Nonetheless, in

whichever manner the waste occurs, this study considers solid waste as

largely non-flowing.

By its peculiar non-flowing nature, solid waste continually remains at

the site of generation or deposition, until it is physically removed for


12

effective disposal. Consequently, solid waste poses so many problems to

the environment and indeed to the managers. Such problems as offensive

odours, obstruction of traffic flow, blocking of waterways/drain channels,

leading to flooding of an area, destruction of environmental aesthetics, and

pollution of the air we breathe with the concomitant unfavorable effects on

public health, can all be caused by solid waste. Yet, a common way of

describing the relative toxicity of solid waste and indeed providing solution

to the solid waste menace is in terms of its place of generation or point of

origin. Hence municipal solid waste (MSW) is defined as “all waste

collected by private and public authorities from domestic, commercial and

some industrial (non-hazardous) sources” Kiely (1998); and Bailie et al

(1997) say MSW comprises small and moderately sized solid waste items

from houses, businesses, and institutions. However, an apt definition of

MSW that seems to properly align with the basic inclination of this work is

that given by Byrne (1997), when he says that “municipal waste is the waste

generated from urban areas, particularly houses and shops"(emphasis

mine).

1.3.1 Management of Municipal Solid Waste

Like many other cities in Nigeria (Ekenta, 2001), large volumes of

refuse are generated on a daily basis in Port Harcourt and also improperly

discarded by residents, along the streets, and mainly at road junctions, and

even along major roads and in waterways and drain channels. Even though

there are official locations where residents should discard these wastes,
13

perhaps for reasons of proximity, unofficial locations have been created

everywhere. Unfortunately, both the official and unofficial locations are still

improperly organised/managed.

For proper management of MSW, Ogunbiyi (2001) quoting former

General Manager of the Lagos Waste Management Agency, Engr. G.O.

Saka, described management of solid waste as “the storage, collection,

treatment and disposal of waste in such a manner as to render them

harmless to human and animal life, as well as the environment”. The

functions of collection, processing and disposal, are referred to as the “three

basic functional elements of solid waste management” (Sincero and

Sincero, 1999). All these clearly show that the management of solid waste

could not be described as effective if, from the beginning, the waste is not

properly discarded for collection in a manner that would not impair the

quality of the environment. The concern of the enhancement of the

environmental quality as an integral component of solid waste management

was further elucidated by Ogunbiyi (2001) with the Federal Environmental

Protection Agency (FEPA) blue-print on MSW management in Nigeria,

which emphasized “waste treatment (minimization recycling and re-use) as

the main principle in waste management.

The position of FEPA was probably informed by global developments

in solid waste management, as, all the world over, particularly in developed

countries, the trend in this regard is the development of management

practices that would be ecologically friendly and protect public health. For

instance, the European Union (EU) Landfill Directive (1995) specifies the
14

type of waste acceptable at landfills. The same is told of the United States

(US) solid Waste Act of 1965, and the California Assembly Bill 939 of 1993

(Kiely, 1998). All these led to the concept of integrated management of solid

waste, which involves the adoption and application of suitable waste

management practices to achieve specific waste management objectives

and goals.

In Nigeria, the solid waste management practices adopted include

basically collection and disposal. Even so, in the execution of these two

fundamental elements, there are a lot of flaws, laxities and inefficiencies.

For instance, the function of collection could be after several days or weeks

when the wastes would have caused all their havoc; and the disposal could

be at a site no better than the location of collection. In the Port Harcourt

metropolis David–West (2001) made a more graphic presentation when he

said that “even where government has contracted out the clearing and

disposal of city refuse, the contractors have the irresponsible bad habit of

transporting the refuse in full loaded OPEN trucks and thereby littering the

streets with these refuse”. Akintola (1978) studied the overall refuse

management procedure in Ibadan city and concluded that the slow rate of

clearance of refuse sites had led to deliberate littering of the cities by

depositors. Presently the government of Rivers State has commissioned

some refuse disposal agents to contend with the solid waste collection and

disposal assignment; but the high waste load and the absence of the waste

treatment function seem to render these efforts unyielding.


15

Betty (1967) studied the refuse disposal situation in North America

and asserted that the three methods of refuse disposal at the time, namely

sanitary landfills, on-site disposal and incineration were inadequate. He

indeed wondered aloud at how to contend with the situation and suggested

that “the only way to solve the problem of garbage is to blast them to the

moon”. Sandback (1980) says that refuse should be recycled, as it will

provide a better protection to the environment and health. In discussing the

subject of resource recovery, Oyinlola (2001) explains that “solid waste

disposal can be viewed not only as a natural problem but as an opportunity

for recovery and utilization of potentially valuable materials”, insisting that

resource recovery should be incorporated as a component of municipal

solid waste management system. He further explains that “resource

recovery from municipal solid waste includes the conversion of solid waste

components to energy, fuels and other valuable products, as well as the

recovery or salvage of existing materials”.

1.3.2 Energy Potential of Municipal Solid Waste

One of the prominent features associated with urbanization is rapid

waste generation, because as population increases more and more

volumes of wastes are generated. According to Pickford (1977) every one

generates refuse when we prepare our foods, or when we sweep our

homes and factories, although the quantity and type of refuse vary. Oyinlola

(2001) cites the 1997 appraisal report of the Urban Development Bank of

Nigeria Plc; and states that the estimated average per capita waste
16

generation for the country is 0.45kg/day, and that for Port Harcourt

metropolis is 0.33kg/day.

Out of the total MSW generated Byrne (1997) says “Organic waste,

ranging from garden wastes to food scraps is still the main component”.

Also from the report of the Urban Development Bank of Nigeria Plc; it was

found that the organic components account for about 76% of total MSW. In

a study of the effect of household size, income level and food consumption

pattern on solid waste generation in Enugu Urban of Nigeria, Oluka (2001)

found that the organic component of the waste is about 91.67% of total

MSW. Little wonder then that the Oregon State Department of Energy

(2002) stated “municipal solid waste contains a large volume of biomass”;

where ”biomass resources are any plant or derived organic matter available

on a renewable basis”. This includes agricultural and forestry crops, and

animal and municipal wastes. Biomass resources can be burned for heat,

produce electricity, or converted to liquid or gas fuels (Suave, 2002).

In developing what is now referred to as the best environmentally

practicable option (BEPO) in solid waste management the concept of

recycling and re-use potential of MSW has become prominent. This involves

conversion and transformation of the waste into a form that will reduce its

volume and especially generate useful by-products in the process. This

corroborates the view of Cipolla (1962) in Hughes (1976) when he states

that “the more successfully man can use his own energy output to control

and put to use other forms of energy, the more he acquires control over his
17

environment and achieves goals other than those strictly related to animal

existence”.

Particularly, because of the organic nature of a large percentage of

the components of MSW, the waste could therefore be converted into some

form of energy, in line with the views of Oyinlola (2001). This transformation

could be accomplished by the use of physical or chemical processes at high

temperatures and/or pressure, or the use of microorganisms at low

temperature and atmospheric pressure, the preferred method being

dependent on their relative polluting impacts on the environment. When

MSW is subjected to these processes, one significant by-product is

methane gas, which is referred to as biogas if it is generated as a result of

the action of microorganisms on the waste in an airtight enclosure.

It is now obvious that MSW has a component that can be converted

to energy, even though Kottner (2002) says “that energy potential of waste

from (the food processing) industry is difficult to estimate, because the

digestibility of some substrates is not yet examined enough”. However, it is

important to state, as Pickford (1977) had explained, that “the quality and

type of waste vary from location to location, as well as from people to

people”, and therefore “the solutions are location – specific” (Oluka 2001).

Kiely (1998) puts it more aptly when he says that “solid waste is non-

standard and typically no two wastes are the same. Even domestic waste

from a single house will vary from week to week and from season to

season”. In fact MSW in Nigeria has been described as a complex mixture

of biodegradable and non-biodegradable substances, which is why


18

Ogumola (1989) said “Nigeria has no records showing the types of

materials and quantities of their waste. Consequently, it has not been easy

for anybody to evolve an effective refuse collection and disposal system for

the country”

1.4 OBJECTIVES OF THE STUDY

From the exposition thus far on the subject, the objectives of this

work can be categorized as follows:

1.4.1 General Objective

- To develop and evaluate appropriate digester models for the

optimization of biogas production from the anaerobic digestion of

Municipal Solid Waste.

1.4.2 Specific Objectives

To achieve this broad-based objective, the following specific

objectives were pursued.

(i) Determination of the rate of MSW generation in Port Harcourt

metropolis

(ii) Determination of the basic characters and composition of the

waste

(iii) Investigation of the bio-kinetics of anaerobic digestion of MSW

(iv) Determination of the potential of generating biogas from MSW

in Port Harcourt metropolis


19

(v) Development of appropriate models for the design of the

anaerobic digester, and

(vi) Simulation and evaluation of anaerobic digestion of MSW in

batch and continuous-flow processes, with a view to

optimizing process performance.

1.5 SCOPE OF THE STUDY

This work involves the experimental determination of kinetic

parameters for the development of engineering models, by material balance

analysis, for the design of an anaerobic digester using solid wastes from

domestic and commercial centres in the Port Harcourt metropolis aimed at

generating biogas.

1.6 LIMITATIONS OF THE STUDY

The paucity of information on the anaerobic digestion of MSW,

particularly using the batch reactor, was a major constrain, as almost all

existing digesters use sewage sludge or animal dung in complete-mix

digesters. A deficiency of appropriate facilities for field and laboratory

experimentation was also of serious concern, as most of the materials were

improvised.

1.7 JUSTIFICATION OF THE STUDY

A very wide gap exists between the energy supply (and demand) of

primitive times and that of civilized societies. This is possibly attributable to


20

the growing world population, diversification of the activities of man and

indeed to advances in science and technology. The energy supply–demand

imbalance of contemporary societies is such that, although the energy

requirements of primitive societies were easily and readily provided by the

use of human or animal power, even with improved technologies, the

energy requirements of modern societies are hardly available. Yet energy

has remained an indispensable requirement of all ecosystems.

The unavailability of energy in the desired proportion is attributable in

the main to an over-dependence on non-renewable fossil fuels, which are

largely depleting and cause high-level degradation of environmental quality.

But there is an abundance of renewable sources in the country from which

alternative energy could be generated. Illoeje (1998) defines renewable

energy as that energy derived from resources which when consumed is

replaceable within a relatively short time.

In Nigeria, the high-energy demand and the sole dependence on

petroleum products have led to a continual upward review of the prices of

petroleum products, hence making energy use very expensive. This is so

because as Ilori (2001) puts it “in Nigeria, domestic production is declining

as a result of poor performance and low output of our refineries and power

generating plants”.

In April 2001 the United Nations Commission on sustainable

development in its 9th session, expresses the view that current pattern of

energy production, distribution and utilization are unsuitable (The Guardian,


21

May 7, 2001). The Commission further identifies three options available for

the institutionalization of sustainable energy, namely:

(i) Improving energy efficiency to get more out of existing resources.

(ii) Expediting the use and decentralization of renewable energy,

which currently makes up 2% of total energy consumption, and

(iii) Accelerating the diffusion of cleaner advanced fossil fuel

technologies while phasing out polluting unsuitable fossil fuels.

When this is added to the high-energy demands of Nigerians, and the

abundance of renewable energy resources in the country, it becomes

imperative to explore alternative sources of energy production for the

country. But, despite this seeming inevitability of an alternative energy

resources base, Esan (2001) alerts that there has not been any special

infrastructure put in place for renewable energy production in Nigeria.

To develop the desired infrastructure for the exploitation of the

abundant renewable energy resources available in the country, Ekenta

(2001) advises on the generation, adoption and dissemination of

“environmentally sound technologies (ESTs)”, defined by Agenda 21 as

those technologies which protect the environment, are less polluting, use all

resources in a more suitable manner, and recycle more of their wastes in a

more acceptable manner than the technologies for which they are

substitutes. So, the waste-to-energy project will indeed elicit sustainable

development, which is defined by the Brundtland report as “Development

that meets the needs of the present without compromising the ability of

future generations to meet their own goals” (Worika, 2001).


22

Fundamentally, MSW is waste, which in our system is usually

discarded without any further processing. First, the disposal of the waste

without treatment causes serious environmental pollution problems,

particularly with serious implication on health. Additionally, the process of

discarding the waste requires land, finance, time and complex technologies.

In Port Harcourt, all of these requirements are relatively scarce, and

sometimes unavailable. So, if instead of investing all these expensive efforts

in discarding the wastes, the efforts are channeled to a transformation of the

MSW in the metropolis into a useful biogas, achieving the discarding (or at

least a considerable reduction in waste volume), the treatment, and a useful

by-product, then it can easily be asserted that this study is worthwhile. But

even then, Hobson et al (1981) say that “burning and pyrolysis are not

always feasible in the waste-to-energy conversion, as the dry waste of the

household (papers, etc.) is mixed with wet vegetable matter and other

kitchen scraps. So, it is then obvious that the anaerobic digestion process is

more suitable.
23

CHAPTER 2

THEORY AND REVIEW OF LITERATURE

2.1 BASIC PROCESSES OF BIOGAS PRODUCTION

Biogas is derived from the microbial degradation of organic waste

stored in the absence of air; a process called anaerobic digestion and

variously defined as “a biological decomposition of organic waste done in

the absence of air (Sincero and Sincero, 1999) and “a biochemical process

in which particular kinds of bacteria digest biomass in an oxygen–free

environment” (Oregon State Department of Energy, 2002). Several different

types of bacteria work together to break down complex organic wastes in

stages, resulting in the production of biogas. This is why Chawla (1985)

says anaerobic digestion is “a bioreactor in which organic matter is

progressively degraded in the absence of oxygen by a process known as

methanogenesis”. Anaerobic digestion has been widely used for the

treatment of industrial, agricultural and municipal waters and sludge, which

accounts for why Hobson et al (1981) define it as a method of stabilizing,

and thus reducing pollution from the sewage sludge produced in several

treatment works; and Reynolds and Richards (1996) say it is the “biological

oxidation of degradable organic sludge by microbes under anaerobic

conditions”. However, Kiely (1998) explains that recently anaerobic

digestion is also being applied to the treatment of municipal solid waste

(MSW) and thus offers a more holistic definition when he says that
24

anaerobic digestion is “the use of microbial organisms in the absence of

oxygen, for the stabilization of organic materials by conversion to methane

and inorganic products, including carbon dioxide”. So, anaerobic digestion

evolved originally and primarily as a waste treatment process, with biogas

generated only as a “waste product”

2.2 MICROBIOLOGY OF THE ANAEROBIC PROCESS

The essential components of the anaerobic digestion (AD) process

are the organic waste and the bacteria, interacting in an airtight enclosure

called anaerobic digester. The organic waste constitutes the ‘food source’

for the bacteria, which convert it into the various end products and by-

products. On the other hand, the bacteria involved in the process are

usually facultative anaerobes, described as obligate anaerobes during

methanogenesis.

Anaerobic decomposition is a complex process, occurring in three

basic stages because of the activities of the variety of microorganisms.

Initially, a set of microorganisms converts organic material to a form that a

second set of organisms utilizes to form organic acids. Then in the final

stage, methanogenic anaerobic bacteria utilize these acids to complete the

decomposition process and give off biogas.

Reynolds and Richards (1996) enumerate these three stages

involved in anaerobic digestion as:

(i) liquefaction of solids

(ii) disgestion of soluble solids, and


25

(iii) gas production

In describing these three stages of the anaerobic process, Kiely

(1998) however noted that four different trophic microbiological groups

(bacteria) are recognized in AD, and that it is the cumulative effect of all

these groups that ensures process continuity and stability. He then explains

the three stages thus:

- Hydrolysis: - the breakdown of high molecular compounds by

hydrolytic and fermentative bacteria to low molecular compounds,

as in lipids to fatty acids polysaccharides to monosaccharides,

proteins to amino acids, etc.

- Acidogenesis: - where the lower molecular components of fatty

acids, amino acids and monosaccharides are converted by

acetogenic (acid forming) bacteria to lower molecular

intermediate compounds such as propionate, butyrate, formate,

methanol and acetate; and

- Methanogenesis: - this is the final stage of methane production

from hydrogen by hydrogenophillic methanogens, and from

acetate by aceticlastic methanogens.

To prosecute these stages the complex organic substrates such as

carbohydrates, proteins, fats and lipids will be hydrolyzed into simpler

soluble products, which are further converted into acetic acid, hydrogen and

carbon dioxide. Kiely (1998) cites Gujer and Zender (1983), as

enumerating seven sub-processes of the anaerobic process thus:

- Hydrolyses of complex particulate organic matter


26

- Fermentation of amino acid and sugars

- Anaerobic oxidation of long chain fatty acids and alcohol

- Anaerobic oxidation of intermediary products

- Acetate production from CO2 and H2

- Methane production by hydrogenophillic methanogenes using

CO2 and H2O.

These sub-processes are diagramatized in the flowchart below.

Complex Polymers

Proteins Carbohydrates Lipids

Hydrolytic Hydrolysis
Bacteria Process
Amino Acids, Sugars Fatty Acids, Alcohols

Intermediate Products:
Acidogenic Propionate, Butyrate, Acidogenesis
Bacteria Valerate Process

Acetate H2, CO2 Methanogenesis


Process

Aceticlastic Methane, Hydrogenophilic


Methanogens CO2 Methanogens

Figure 2.1: Stages in methane production from organic waste


Adapted from Kiely (1998).
27

Eckenfelder (2000) further explained that during hydrolysis, there is

no reduction of the chemical oxygen demand (COD); but during the

conversion of the monomers to volatile fatty acids (VFAs) there is minimal

reduction of COD. Eventually when the organic acids are broken down to

CH4 and CO2, there is considerable reduction of COD. A schematic of the

carbon and hydrogen flow in the anaerobic digestion process is shown in

Figure 2.2.

4%
H2

24% 28%

Higher
76%
Complex organic CH4
organics acids

52% 72%

20%
Acetic
acid

Hydrolysis Acetogenesis Methanogenesis

and fermentation and dehydrogenation


Stage 1 Stage 2 Stage 3

Figure 2.2: Carbon and Hydrogen Flow in Anaerobic Digestion Process


Adapted from Eckenfelder (2000)
28

The breakdown of carbohydrates, nitrogenous compounds and fats

can simply be expressed using chemical formula as follows:

C6H12O6 + 2H2O 2C2H4O2 + 2CO2 + 4H2 (2.1)

From the acetic acid and hydrogen products of the above reaction, methane

would be produced thus.

2C2H4O2 2CH4 + 2CO2 (2.2)

4H2 + CO2 CH4 + 2H2O (2.3)

When these expressions are combined, the generalized equation for the

anaerobic digestion process is obtained as follows:


Organic Combined Anaerobic New Energy Other end
+ + + CH 4 + C02 +
matter water microbes cells for cells products
(2.4)

2.2.1 Benefits of Anaerobic Digestion

Anaerobic digestion has generally been accepted as the basic

process of treatment of organic waste in a hermetically closed reactor or

tank. In this process, the breakdown of the organic matter by anaerobic

microorganisms and the release of biogas, lead to a considerable reduction,

if not complete elimination of the odour of the waste. Vassiliou (1997) says

that “one of the distinctive benefits of biogas plants based on anaerobic

treatment of waste is that, as a by-product of their operation and, as if by

magic the bad smell disappears and the pollution load (as measured by the

BOD-Biological Oxygen Demand Index) is virtually eliminated. Hence,

Biogas plants and technologies are not only, organic recycling plants. They

also serve as plants for mitigating pollution and controlling the environment”.
29

Hence, the Oregon State Department of Energy (2002) states that “in recent

years increasing awareness that anaerobic digesters can help control

animal waste odour and disposal has stimulated renewed interest in the

technology”.

Kottner (2002) adds that the advantages of anaerobic digestion

technology is also the ‘CO2 – neutral generation of energy’ and the

avoidance of ‘methane and nitrous oxide emissions, the saving of fertilizer

and chemical sprays, the reduction of landfill area and the protection of

groundwater’.

Clearly, all these postulations tend to suggest, like the OSDE (2002)

explains, that ‘it is often the environmental reasons – rather than the

digester’s electrical and thermal energy generation potential – that motivate

farmers to use digester technology’. They were, however, quick to add that

this is especially true in areas where electric power costs are low. They said

that anaerobic digester systems can reduce fecal coliform bacteria in

manure by more than 99 percent, virtually eliminating a major source of

water pollution.

However, in developing countries, like Nigeria, the twin by-products

of anaerobic digestion, namely the environmental pollution control and the

energy generation are significantly essential to the lives of the people; and

particularly the energy component, at a time when the country is solely

reliant on fossil fuels, namely crude oil products. This is why it is easy to

rather enunciate the position of Hobson et al (1981) when they said, “the

emphasis at the moment is largely on energy production, but in general . . .


30

most effective gas production from waste usually gives most effective

pollution control”.

In discussing the benefits of AD, Stafford et al (1980) and Kiely

(1998) have a consensus on the following points:

(i) Production and conservation of energy

(ii) Reduction of pollution potential of waste

(iii) Enhancement in fertilizer/fuel value of waste product

(iv) Elimination of pathogens and weed seeds (if mesophilic or

thermophilic)

(v) Reduction in the quantity (or volume) of solids or nuisance

value.

Tchobanoglous et al (2003) in what they called ‘the rationale for

anaerobic treatment’ attempted to compare the anaerobic treatment

process to aerobic process, and states that in addition to the anaerobic

treatment process being a net energy producer, it results in lower biomass

production, requires fewer nutrients, and has higher volumetric organic

loads.

Agunwamba (2001) adds importantly amongst other things, that AD

can utilize different kinds of wastes.

2.3 TYPES OF ANAEROBIC DIGESTERS

Over the years, digesters for anaerobic process have been

developed into different configurations. Existing types of digesters are

classified into conventional or low-rate digesters, high-rate digesters, batch


31

digesters, continuous digesters, fixed-cover digesters and floating cover

digesters, with many variations, including high solids digesters and two-

stage digesters. It is important to note, as Mattocks (1984) explained that

whereas a batch digester, which can be any suitably sized container or tank

is loaded sealed and after a period of gas collection, emptied, a continuous

feed digester receives substrate on a continuous or daily basis with a

roughly equivalent amount of effluent removed. Whatever type or variation

of digester there is, they all can trap methane, and reduce feed coliform

bacteria but differ in cost, climate sustainability and the concentration of

manure solids they can digest. A typical anaerobic digester process is

shown in Figure 2.3.


32

Gas Gas
Abstraction Abstraction

Gas 5 - 15% Gas


Tank Vol
Sludge
Scum Influent
Sludge
Influent
Supernatant Active
Supernatant
Abstraction Mixing
Actively Digesting
Sludge Zone

Stabilized
Sludge

Digested Sludge
Sludge Withdrawal

(a) Low Rate (b) High


Rate

Figure 2.3: Schematic of a typical AD process


Adapted from Kiely (1998)

These various digester configurations have been described by

Reynolds and Richards (1996). The conventional or low-rate digesters have

intermittent mixing, intermittent sludge feeding and intermittent sludge

withdrawal. They explain that when mixing is not being done, the digester

contents are stratified (as in Figure 2.3a). On the other hand, high rate

digesters are characterized by continuous mixing and continuous or

intermittent sludge feeding and sludge removal. Despite these

classifications, Steadman (1975) says, “the simplest type of methane


33

digester is just a closed container such as a drum, tank or pit in the ground

into which the digestible material is loaded. The container is then sealed to

the air. Some means must be found for trapping the gas which is given off”.

Kiely (1998) classified reactors into first and second-generation

types:

(i) In the first generation type, comprising of batch digesters, plug-

flow digesters and continuously stirred tank reactor (CSTR) and

anaerobic contact reactor (Figure 2.4) the hydraulic retention time

() is equal to the solid retention time (c); while.

(ii) In the second generation type of upflow-downflow anaerobic filter,

downflow stationary fixed film reactor, fluidized bed reactor,

upflow anaerobic sludge blanket reactor, and the hybrid

anaerobic sludge reactor (Figure 2.5), the solids retention time

(c) is greater than the hydraulic retention time ().

Schematic representations of these various reactor classifications

are shown in Figures 2.4 and 2.5.


34

Biogas Biogas

Influent Effluent Influent Effluent


Plug

Sludge recycle
(a) Batch digester
(b) Plug flow digester with recycle
Gas
headspace
Biogas Biogas

Effluent
Influent Effluent Influent Effluent
Aerobic
Anaerobic

Sludge recycle

(c) CFSTR (d) Anaerobic contact reactor or


contact stirred reactor (CSTR)
Figure 2.4: First Generation Digesters
Adapted from Kiely (1998)

Gas
Gas

Influent
Influent or
Effluent
Fixed media

Influent or
Effluent
(a) Upflow/downflow anaerobic filter (b) Downflow stationary fixed film
Gas Gas
Gas
Effluent
Effluent Effluent
Media
No media
Recycle

Sludge Influent
Influent
Influent (e) Hybrid anaerobic sludge reactor
(d) Upflow anaerobic sludge
(c) Fluidized bed blanket (UASB)

Figure 2.5: Second Generation Digesters


Adapted from Kiely (1998)
35

The first generation reactors of Kiely’s classification, particularly the

batch reactor, plug-flow reactor and continuously stirred tank reactor, have

featured prominently in anaerobic treatment of organic waste. Reynolds and

Richards (1996), explain that whereas in a batch reactor, the reactor is

charged with the reactants, the contents well mixed and left to react, and

then the resulting mixture discharged; in continuous flow reactors, including

the plug-flow and CSTR, the feed to the reactor and the discharge from it

are continuous.

However, these different types of digesters are employed for use

based on the type of waste involved; as Kayhanian et al (1991) cited in

Kiely (1998) state that high solids digesters are particularly suited to the

treatment of municipal solid waste. Inspite of all these, Hobson et al (1981)

state that the AD process is an empirical science dependent on practical

test. They explain that because of the peculiarity of waste to different

situations and locality, and particularly the complexity of substrates and the

reactions involved in their breakdown, it is difficult to predict the behaviour

of waste in digesters. Therefore, they insist that “the only way is to do trial

digestions” using the waste in question.

2.3.1 Description of Batch and Continuous-flow Digesters

A batch reactor is characteristically a system in which a given

quantity of material is fed in at specified intervals and also withdrawn after a

given time. It is easily described as a non-flow system, for which reason the

composition of a batch digester is not uniform; and it is suited for use in high
36

solids digestion processes. A schematic of a batch digester is shown in

Figure 2.6.

Composition changes with time

[So]

Figure 2.6: Schematic of a batch digester

A continuously stirred tank reactor (CSTR) is one in which the

reactants are continuously fed into the reactor at one point, and the

products of partially reacted materials, also continuously discharged from

another point of a well-mixed vessel. For these reasons the CSTR is also

frequently referred to as either “continuous-feed digesters” or “completely

mixed reactors”. Because of the thorough mixing in the CSTR, the contents

are assumed uniform in concentration with no concentration gradients,

hence equal to the effluent concentration.


37

Q, [So] Q, [X], [S]

V, [X] [S]

Figure 2.7: Schematic of a CSTR

Also, because of the flow characteristics of materials in a CSTR the

reactor is principally adapted to the digestion of low-solids materials.

Tchobanoglous et al, 1993 define low-solids anaerobic digestion as “a

biological process in which organic wastes are fermented at solids

concentrations equal to or less than 4 to 8 percent”. The Oregon State

Department of Energy (2002) says, “a complete mix digester is suitable for

manure that is 2 percent to 10 percent solids”; and Steadman (1975) says,

“the recommended percentage of solids in the digesting mixture is between

7% and 9%.

2.3.2 Previous Anaerobic Digesters

In the 19th century Pasteur (Steadman, 1975) had written about the

possibilities of methane from organic decay. He explains that anaerobic

digesters probably started in the form of covered sewage lagoons, where


38

the Chinese were known to have collected methane for power uses. At a

time, a number of street lamps in London were run on gas from

underground sewers.

Tchobanoglous and Burton (1991) state that anaerobic digestion is

among the oldest forms of biological wastewater treatment, and that its

history can be traced from the 1850s with the development of the first tank

designed to separate and retain solids.

However, Steadman (1975) explains that it seems that the first

specially built plant for producing methane from manure, was at Bombay in

1900. “Experiments carried out at the Indian Agricultural Research Institute

in New Delhi in 1939, resulted in small prototype gas plants which were

installed for testing in ten villages around Delhi”. During the years 1956-7

the two tons of manure produced daily by the 1,000 animals owned by one

L. John Fry (Steadman, 1975) posed a serious disposal problem. Fry then

employed the principle of methane production using a series of digesters

made from 50-gallon oil drums. Later, he built one very successful plant in

South Africa in the early 1960’s.

Tchobanoglous and Burton (1991) also explained that one of the first

installations in the United States using separate digestion tanks was the

wastewater treatment plant in Baltimore, Maryland. But the Oregon State

Department of Energy (2002) says that “anaerobic digestion and power

generation at the farm level began in the United States in the early 1970s, .

. . and in 1978, Cornell University built an early plug-flow digester designed

with a capacity to digest the manure from 60 cows”.


39

Clearly, the variation in anaerobic digester designs is contingent

upon the postulation, as previously enunciated, that the solution to refuse

disposal and indeed refuse treatment problems is ‘location–specific’, which

is also here corroborated by the position of the International Development

Research Center (IDRC), when they said that “the viability of a particular

biogas plant design depends on the particular environment in which it

operates. Therefore, the research problem becomes one of providing a

structure in which technologists, economists and users of the technology

can combine to produce both the appropriate hardware for various

situations and the infrastructure that is necessary to ensure that the

hardware is widely used”.

In the use of the anaerobic digester, Agunwanba (2001) explains the

“Valorga Process”, which he says is a technique for processing

biodegradable organic wastes with the main aim of maximizing the value of

the organic matter and the protection of the environment. The

mechanization unit of the Valorga process is shown in Figure 2.8.


40

Figure 2.8: The Valorga Process


Adapted from Agunwamba, 2001

Generally, digesters vary in size, scale and complexity, depending on

the purpose, type of waste, and economic and technological advancement

of the people. Digesters of 4,500m3 have been built for very large cattle

feedlot units in America (Hobson et al, 1981). A frequently cited case study

of the anaerobic digester is the complete mix digester located at Colorado

Pork LLC near Lamar, Colorado (McNiel, 2000). The Colorado Pork LLC

digester ‘began operations in September 1999 and is fed with 20,000

gallons per day of combine manure and flush water from the commercial

swine operations in Colorado, which produces about 60kW of electricity. A

schematic of the complete mix digester is shown in Figure 2.9.


41

Figure 2.9: Schematic of the Colorado Complete Mix Digester


Adapted from McNiel (2000)

Small-scale digestion units have been built in developing countries.

Hobson et al (1981) state that the greatest number of small digesters have

been built in India, China and Taiwan. They explain that the Indian digester

is based on the design instituted by J. J. Patel, and is referred to as the

‘Gobar digester’ shown in Figure 2.10. Steadman (1975) explains that

‘Gobar’ is Hindi for cow dung. It is also referred to as the Khadi design, and

is based on the principle that gas produced will lift a bell-shaped dome

located above the digestion vat (Mattocks, 1984). Substrate enters one side

of the digester and displaces effluent out of the other side.


42

Characteristically, its feedstock is usually cattle manure plus ‘night soil’, the

human excrement. (Hobson et al, 1981).

Figure 2.10: Indian Gobar Digester


Adapted from Hobson et al (1981)

In Taiwan, the sunken tank, floating gas dome type digester, similar

to the gobar digester, has been designed (Chung Po, 1975, cited in

Hobson, et al, 1981). Here pigs’ excreta are used.


43

Figure 2.11: Taiwan Model Anaerobic Digester


Adapted from Hobson et al (1981)

The Chinese have developed small, cheap digesters using animal

and vegetable wastes. Unlike the Indian design, the Chinese is completely

inside the ground.

Figure 2.12: Chinese-Type Anaerobic Digester


Adapted from Hobson et al (1981)
44

Notably, all these digesters, from the Colorado complete mix to the

Indian design, to the Taiwan and Chinese models, use one type of animal

excreta, manure or another; as Hobson et al, (1981) and Kiely (1998) noted

that investigations into the use of MSW are still at their preliminary stages.

Indeed Vassiliou (1997) while discussing what he described as the “Biogas

Project” of Cyprus, stressed the “use of state-of-the-art technology to

produce biogas, electricity, and organic fertilizer from organic waste –

initially from livestock farm waste (various manure, slurry and waste waters)

and agro-industrial waste (abattoirs, wineries, vegetable processing plants

etc); and, in due course, from source-separated MSW.

However, a prominently featured ‘dry solids’ anaerobic processing

system was reported by Kiely (1998). It is the ‘Dranco’ process for organic

MSW systems in Belgium, which uses a solids concentration of 30 to 35

percent. Dranco means dry anaerobic composting. Although the mode of

operation of the Dranco system was not elaborately stated, Kiely explains

that the process has a cycle time of 16 to 21 days, and biogas production is

5 to 8m3/m3 of reactor. The gas content is 55 percent CH4, and energy

production is 140 to 200m3 biogas per tonne of raw organic waste at 40 to

60 per cent dry solids. The flow chart in Figure 2.13 shows the Dranco

process of treating organic MSW.


45

Organic MSW

55% DS 100 l/d


Biogas
Dry anaerobic fermentor Gas
3000 M3 engine Generator
160M3/L 1.25 MW
Recycle DS
filtrate

Press Heat

Dryer

50 l/d Stable
60 - 80 % DS compost

Figure 2.13: High solids Dranco process for organic MSW


Adapted from Kiely (1998)

Also in Nigeria all efforts have been concentrated in the area of

generating biogas from livestock wastes and manure (Itodo et. al., 1997;

Itodo and Awulu, 1999; Itodo and Phillips, 2001, etc). The closest research

effort to the generation of biogas from MSW is Bitrus (2001), who

purportedly developed a mathematical/computer model for the design of a

biological-reactor for generating methane gas from MSW. However a critical

review of his work was an exhibition of, a theoretical synthesis of existing

mathematical models, without any empirical experimentation as to the

applicability of the model to the reality of the solid waste situation in our

environment, and without any consideration for the economics and process
46

effectiveness and efficiency of digesting a high solid feedstock at a 4% total

solids (TS) concentration.

Tchobanoglous et al (2000) enlist some of the processes that are

either in use or in contemplation for use for the anaerobic digestion of the

organic fraction of MSW, as presented in Table 2.1 below.

Table 2.1: Summary of anaerobic digestion processes and technologies


for treatment of the organic fraction of MSW a

Anaerobic Country Status Description


digestion
process
Sequential batch USA Experim- The SEBAC is a batch anaerobic three-stage
anaerobic ental process. In the first stage, a bed of coarsely-
composting stage shredded feedstock is inoculated by recycling
(SEBAC) leachate from the third-stage reactor in the final
stages of digestion. Volatile acids and other
fermentation products generated during startup
are removed from the first-stage reactor to the
second-stage reactor for conversion to
methane.
High-solids USA Under The high-solids anaerobic digestion/aerobic
anaerobic develop- composting is a two-stage process. The first
digestion/aerobic ment stage involves the dry digestion (solid content of
composting 25 to 32 percent) to convert the organic fraction
process of MSW to methane. The second stage involves
aerobic composting of the anaerobically
digested solids to produce a fine humus-like
material that can be used as a fuel or soil
amendment.
Semi-solid Italy Under The semi-solid anaerobic digestion/aerobic
anaerobic develop- composting is a two-stage process. The first
digestion/aerobic ment stage involves the semi-dry digestion (solids
composting content of 15 to 22 percent) to convert the
process organic fraction of MSW to energy. The second
stage involves the aerobic composting of the
commingled anaerobically digested solids and
the biodegradable fraction of organic MSW to
produce humus-like material.
Leach-bed two- UK Experim- The leach-bed two-phase anaerobic digestion
phase anaerobic ental involves the rapid bio-leaching of organic matter
digestion process stage from the putrescible material in MSW in
specially engineered and prepared landfill. To
accelerate the process, the resulting leachate is
recirculated through the solid material.
47

Two-step Germany Experime The two-step anaerobic digestion process is


anaerobic ntal used to treat organic substrates at low C/N
digestion stage ratios and high loading rates. The process is
based on a sequential biochemical conversion
for the organic solids, which allows for better
process control. This process is carried out in a
semi-liquid phase in the mesophilic temperature
range.
Bio-waste Denmark under The bio-waste anaerobic treatment system is
process develop- designed to treat the source-separated
ment household solid waste along with industrial and
agricultural waste. the complete-mix digester is
operated in the thermophilic temperature range.
Kampogas Switzerland Under
process develop- The kampogas is a new anaerobic digestion
ment process treat grit, yard waste, and vegetable
waste. The digester is cylindrical and positioned
horizontally. The digester, equipped with a
hydraulically driven stirrer, is operated at high
solids concentrations in the thermophilic
temperature range.

Dranco process Belgium Develop- The dranco process is used for the conversion
ment of the organic fraction of MSW to produce
energy and humus-like product, called humutex.
The digestion process is carried out in vertical
plug-flow reactor with no mechanical mixing, but
leachate from the bottom of the reactor is
recirculated. The dranco digester is operated at
high-solids concentrations and in the mesophilic
temperature range.

BTA process Germany Develop- The BTA process is developed especially to


ment treat organic fraction of MSW. The BTA
treatment process includes: (1) pretreatment of
incoming waste by mechanical, thermal and
chemical means; (2) separation of dissolved
and un-dissolved biogenous solids; (3)
anaerobic hydrolysis of biodegradable solids;
and (4) methanization occurs at low solids and
mesophilic temperature ranges. After
dewatering, the non-degraded solids, with a
total solids concentration of 35 percent, are
used as a compost-like material.
Valorga process France Develop- The valorga process is comprised of sorting
ment unit, methane producing unit, and refining unit.
the anaerobic fermenter operates at high-solids
concentrations and in the mesophillic
temperature range. The mixing of the organic
matter in the reactor is achieved by re-
circulation of biogas under pressure at the
bottom of the digester.
48

Bio-cell process Nether- Under The biocell process is batch system developed
Lands develop- to treat source-separated MSW (fruit, yard
ment wastes, and vegetable wastes and agricultural
wastes. The digester used was circular in
shape, 11.25m in diameter and 4.5m in height.
The digester feedstock, at a total solids
concentration of 30 percent, was obtained by
mixing the incoming source-separated organic
MSW with digested solids from previous
digestion run.

Adapted from Tchobanoglous et al (2000)

From this presentation, it is evident that the process of the anaerobic

treatment of MSW is still largely at the experimental stage.

2.4 BIO-KINETICS OF ANAEROBIC DIGESTION

Principally anaerobic digesters are described as “microbiological

production plants” which makes knowledge of the bacteria and their

reactions, and their limitations of growth, and factors affecting their growth,

imperative in the design and running of digesters (Hobson et al, 1981).

Thus, in the design of an anaerobic digester for the generation of useful

quantities of valuable energy a very important consideration is usually the

rate of growth or otherwise of the microbes required for the process. This

factor is relevant in the design of process systems, as it aids the

determination of the rate at which various components of the organic matter

are decomposed, and also the rate of biomass sludge production. All these

variables affect reactor size and configuration; and also the residence

(detention) time of organic matter in the reactor.

The batch process (culture) is frequently prescribed for the digestion

of the kind of waste (MSW) under consideration, because the high solid
49

content and usual large sizes of the waste predispose it to longer periods of

biochemical reactions, as a result of the length of time required for microbial

growth, and then substrate digestibility. Hobson et al, 1981, state that “the

time course and the kinetic model of the ‘dry digester’ is essentially that of

the batch culture”. However, a lot more possibilities exist for the treatment of

the waste, including its processing in a continuous-flow digester, where the

waste would be shredded to fine particles and diluted to meet the desired

total solids concentration for CSTRs.

Despite a relative increased volume of operation per unit time in a

continuous-flow processing, the process is also accompanied by good

product quality control. Thus Levenspiel (1999) insists that “the steady state

flow reactor is ideal for industrial purposes when large quantities of material

are to be processed, and when the rate of reaction is fairly high to extremely

high.

In biological processes, the theoretical basis for the kinetics of

bacterial growth, decay and substrate utilization reactions according to

Wong (1982) is derived from an understanding of autocatalytic reactions

first characterized and explained by Michaelis–Menten (1913) in enzyme-

catalysed fermentation reactions. He explains that this was further

elaborated in the form of mathematical models of continuous growth pure

culture microbial systems by Monod (1949) and Novick and Szilard (1950).

In these considerations, two basic methods feature prominently in the

determination of the rate at which the substrate reduces or is degraded.

Reynolds and Richards (1996) state that whereas the first approach uses a
50

modification of chemical kinetics, which equations are derivable from the

Michaelis-Menten relationship for substrate utilization; the second approach

is based on the Monod relationship for microbial growth kinetics.

Basically in batch cultures for both old and fresh charges, there is

usually a lag period after inoculation, during which the bacterial inoculums

adjust to the new medium, or indeed for the regeneration of bacteria

particularly for fresh charges. Under these conditions, the substrate

utilization is described as zero order immediately after the inoculation when

substrate concentration is still relatively high, and later as first order when

substrate concentration becomes low.

In describing the kinetics of the microbial process, it is usually

convenient to distinguish between the kinetics of growth of microbes, and

the kinetics of substrate utilization. However, different reaction orders (as

previously noted) occur for a variety of organisms, substrates and

environmental conditions.

2.4.1 Modeling Anaerobic Digestion as First Order Reaction

When the rate of reaction of a process is directly proportional to the

concentration of the reactant, the reaction is described as first order; then

the conversion of a single reactant (A) to a single product (P) could be

described mathematically as:

d S 
  k S  (2.5)
dt
51

where k is the first-order reaction rate constant, and S, the concentration at

any time, t.

S 
Integrating In o   k1t (2.6a)
 Se 

S  kt
or log  o   1
 (2.6b)
 Se  2.3

The time taken for the reaction to attain 50% completion or half its initial

concentration, which is a useful measure of reaction performance, could be

obtained using the following relationship:

 S 
In  o   k t 12 (2.7)
 So / 2 

In 2 0.69
Therefore t 12   (2.8)
k k

(which is also called the half saturation constant)

The relationship between the rate of reaction and substrate

concentration for first order reactions is graphically shown on the next page.
52

Co

Concentration
Slope
as log C

Time (t)

Figure 2.14: Substrate concentration influence on rate of reaction

Adapted from Kiely (1998)

2.4.2 Anaerobic Digestion Reaction Mechanism

Several processes take place in the course of the anaerobic

degradation of organic materials. The mechanism of anaerobic digestion

reaction has been variously described in terms mainly of the analytic

evaluation of the process operation.

2.4.2.1 The Digestion Process Rate Equation

For the anaerobic digestion process, particularly for mixed cultures

the biomass degradation rather than the number of organism is of essence.

If ‘X’ represents the mixed population of microorganisms utilizing the

organic waste, then the rate of increase in biomass, which is proportional to

the initial biomass concentration, is normally modeled as a first-order


53

process (Kiely, 1998; Sincero and Sincero, 1999). The rate equation

expressing this first order relationship is of the form:

d X 
rx    X  (2.9)
dt

where
d X 
rx = = growth rate of biomass, mg/l/day
dt

X = concentration of biomass, mg/l

 = specific growth rate of the mixed population


(days-1)

= mass of cells produced / mass of cells present


per unit time

t = time (days)

By first-order kinetics, as had been stated, if Xo represents the

biomass at time, t = 0, then

d X 
x t
integrating: x X    0 dt (2.10a)
o

In X   InX o   t (2.10b)

 [X ] 
In    t (2.10c)
 o 
[ X ]

[X ] = [Xo] exp t  (2.10d)

However, a growth rate that follows this expression, called the

exponential growth rate may not always be the case, particularly for the

batch culture where environmental conditions change during its lifetime.


54

Introducing a conversion parameter called the fractional conversion

(), which refers to the ‘fraction of the reactant converted to the product’

(Levenspiel, 1999), such that if Xo be the initial concentration of the

reactant, and X is the concentration of the reactants at any point in time, t,

then the conversion of the reactants in a constant-volume system will be:

Xo  X X
  1 (2.11)
Xo Xo

dX
and d  (2.12)
Xo

Considering the rate equation in terms of the fractional conversion gives

d
  1    (2.13)
dt

which upon rearrangement and integration becomes:


d
t

0 1     0 dt (2.14a)

or  ln1     t (2.14b)

 X 
and a plot of ln1    or ln  against time is expected to produce a
 o
X

straight line through the origin.

2.4.3 Bacteria Growth Pattern in Batch Culture

Monod (1949) cited in various literatures (Sincero and Sincero, 1999.

Kiely, 1998; Reynolds and Richards, 1996, and Hobson et al, 1981, etc.),
55

was the first to identify that in pure cultures,  is a function of or limited by

the concentration(s) of a limiting substrate and then developed the empirical

equation:

   max
S  (2.15)
K s  S 

where max = maximum growth rate, days-1

S = concentration of limiting substrate, mg/l

Ks = half saturation constant (i.e concentration of S


when  = max/2 (mg/l)

Therefore, substituting for ‘’ in equation (2.14b) now relates the fractional

conversion to both the maximum growth rate of biomass and the substrate

concentration thus:

 ln1      max
S  t (2.16)
K s  S 

The Figure 2.15 describes the growth pattern of bacteria in batch

culture. It shows that after an initial lag period for the bacteria to adapt to

their new environment, there is an exponential increase in the number of

viable cells, which is facilitated by the availability of excess organic matter,

and limited only by the ability or otherwise of the microorganism to process

the substrate. On the other hand, the declining growth phase arises from a

shortage of substrate; and this continues until the number of viable bacteria
56

becomes stationary, when the rate of reproduction is equal to the rate of

death.
Concentration of biomass, X

Exponential Declining Stationary Endogenous


growth growth phase phase
phase phase
Xm

Xo

So = Smax Sin = 0
Time, t

Figure 2.15: Schematic of Microbial Growth Pattern in Batch Culture

Adapted from Viessman and Hammer (1999)

Considering the dynamics of death of microbes in process operation

as a decay of microbial population kd[X], where kd is the rate of decay or the

endogenous decay coefficient, then the model for the rate of increase of the

mixed population of microbes [X] becomes

d X  S  X   k X 
  max (2.17)
K s  S 
d
dt
57

Relating this to the rate equation, then

rx   max
S  X   k X  (2.18)
K s  S 
d

2.4.4 Bacteria Growth Pattern in Continuous Culture

Hobson et al (1981) state that the simplest form of continuous culture

also called the chemostat culture is a tank with inflow of medium and

outflow of bacteria, metabolic products and depleted medium. They explain

that the theoretical model formulation for continuous culture was originally

proposed by Herbert et al (1956) based on the Monod description of

bacterial growth in pure cultures.

As the fresh medium flows into the charged tank, there is an

interaction between the incoming medium and the residual charge

described by a dilution rate (D) of the culture, which represents the inverse

of the detention time of medium and bacteria in the system (Hobson et al,

1981) and defined by Bailey and Ollis (1986) as the number of tank volumes

which pass through the vessel per unit time; and this corresponds to the

concept of space-velocity presented by Levenspiel (1999).

Therefore

Q 1
D =  (2.19)
V 

where D = dilution rate h-1

Q = flow rate of medium

V = volume of the culture


58

d X 
From equation (2.9)   X  , it has been shown that the bacteria are
dt

growing in the culture at a rate of [X].

Considering the dilution rate, then, the bacteria are also flowing out

of the culture at a rate of D[X]

This means that the rate of change of bacterial concentration will be

given as:

d X 
    X   D X  (2.20)
dt

 d S  
Also the rate of change of the limiting substrate concentration   in the
 dt 

culture will be

d S   X 
 DSo   DS   (2.21)
dt Y

where So = incoming substrate concentration

S = out flowing substrate concentration

Y = yield factor

dSo = input rate

dS = output rate

 X 
= rate of substrate consumption by bacteria
Y

Substituting for '' in equation (2.20) from equation (2.15), and neglecting

cell death gives

d X   S  
  max X    DX  (2.22)
dt  K S  S  
59

and substituting for '' in equation (2.21) from equation (2.15) gives;

d S  X    S  
 DSo   DS   max   (2.23)
dt Y  K S  S  

Under steady – state conditions, where the concentration of either the

biomass or substrate does not change with time,

d X  d S 
  0 (2.24)
dt dt

Therefore, from equation (2.22), the steady state mass balance on

the biomass becomes.


 max X 
S    DX   0 (2.25)

 K S  S  

which gives

 DK S
 S     D  (2.26)
max

where S - substrate concentration of a sterile feed X o  0 and from

equation (2.23), the steady state mass balance on the substrate becomes

DSo   DS  
X    S  
   0 (2.27)
Y
max
 s
K  S  

from where,

X  
 Y S o   K S 
D 
 (2.28)
  max  D 
60

X - biomass concentration of a sterile feed (Xo = 0).

Equations (2.26) and (2.28) show that the parameters S and X are clearly

 Q
dependent on the flow rate  D   at steady state conditions.
 V

From equation (2.20), considering steady state conditions, then

X    D  0 (2.29)

which shows that at steady state  = D, implying that the rate of growth of

biomass equilibrates the rate at which the medium passes through the

system, indicating that a non-zero cell population can only be maintained at

steady state.

Hobson et al (1981) further define a critical or washout dilution rate

Dc as

 S o  
DC   max   (2.30)
 K S  S o  

As ‘D’ tends to ‘Dc’, the culture will washout because [X] will drop drastically

since no substrate will be used for biomass growth because  and [S] are at

their highest possible values. So, the critical dilution rate describes the level

of dilution at which the culture will washout before there was any digestion.

The diagram in Figure 2.16 is a graphical representation of bacterial

growth pattern in a continuous culture with one growth limiting substrate.


61

Figure 2.16: Bacterial growth pattern in continuous culture


Adapted from Hobson et al, 1981

2.4.5 Substrate Kinetics

Substrate kinetics is founded on the premise that as organisms grow,

substrates are consumed; so that the rate of decrease in substrate

concentration is proportional to the rate of increase of the concentration of

the organism (Reynolds and Richards, 1996). So, in terms of the net growth

rate of biomass, equation (2.16) becomes


 ln1       max
S   k  t
d (2.31)
 K s  S  
62

Now, assuming that all substrate could be converted into biomass,

the rate of decrease of the substrate will be described as:

d S  d  X 
  (2.32)
dt dt

But practically this idealization is not feasible due to inefficiencies in the

conversion process. However, introducing proportionality constant, U, gives

the following relationship, which is of a more practical relevance.

d S  S  X   k X 
  U max (2.33)
K S  S 
d
dt

or 
d[S ] 1
  max
S  X   k X  (2.34)
K S  S 
d
dt Y

where U = specific substrate utilization rate

and Y = specific yield of organisms, mg/l

These formulations are depicted graphically as shown in Figure 2.17.


63

Growth Rate Constant, , time -1


max

½max

Ks Substrate Concentration, S

Figure 2.17: Relationship between the growth constant,  , and the


substrate concentration, S
Adapted from Reynolds and Richards (1996)

2.4.6 Process Inhibition

The Monod function is usually not valid for substrates, such as

volatile acids, which limit growth at low concentrations and are inhibitory to

the organisms at higher concentrations Andrews (1978). Viessman and

Hammer (1993) were more apt when they said that methane bacteria are

strict anaerobes that are very sensitive to the conditions of their

environment. They explained that these types of bacteria could be

adversely affected by excess concentrations of oxidized compounds,

volatile acids, soluble salts, and metal cations. Following the limitations of

the Monod function to accommodate the effect of these inhibitory

developments, Andrews (1978) presented the inhibition function proposed


64

by Andrews (1968) which encompasses the effects of these inhibitions on

process performance, thus:

1
   max (2.35)
K [S ]
1 S 
[S ] Ki

where Ki = inhibition coefficient, moles/litre

So, when the anaerobic digestion process is put in perspective, a more

elaborate form of the rate of substrate utilization becomes:

 k X 
d[S ] 1 1
   max
K s S  d
(2.36)
dt Y 1 
S  Ki

The inhibition function is useful in the development of a model, which can

predict process failure due to organic overloading as well as, hydraulic over

loading.

In these considerations, Viessman and Hammer (1993) further state

that pending failure of the anaerobic digestion process is evidenced by a

decrease in gas production, a lowering in the percentage of methane gas

produced, an increase in the volatile acids concentration, and an eventual

drop in pH.
65

2.5 DEFICIENCIES OF PREVIOUS WORKS

Waste-to-energy conversion processes and indeed waste treatment

generally are usually a function of the type of waste. Previous works that

are somewhat related to this study are seriously deficient in several regards,

including the following.

(i) There has been no empirical determination and consideration of the

peculiar physical, chemical and biological characteristics of the MSW

in Port Harcourt metropolis for the design of anaerobic digesters,

considering that according to established literatures previously stated

in this work, ‘the treatment of every waste is location-specific’.

(ii) No investigation had been conducted to determine the kinetic

parameters or coefficients that describe the behaviour of the MSW in

anaerobic digestion processes.

(iii) The exothermicity of the anaerobic digestion had previously been

neglected in related works, thus resulting in erroneous heat balance

models, which do not satisfy the isothermal requirements of the

process.

(iv) As a consequence of the foregoing no appropriate mathematical

models had been developed for the description of the bio-kinetics of

MSW, and the design of anaerobic digesters for the digestion

process.

(v) In Nigeria there had been no design or attempt that considered the

use of the batch process for the digestion of MSW.


66

2.6 WORK CARRIED OUT IN THIS STUDY

Deriving from the inadequacy of previous studies in this area, as

highlighted above, the following work was carried out in this work.

(i) Establishment of data about the physical, chemical, and biological

characteristics of MSW.

(ii) Determination and proper evaluation of the kinetic parameters in the

anaerobic digestion of MSW.

(iii) Determination of the microbial heat generated during the process,

and consequently the net heat required for the desired isothermal

digester condition.

(iv) Development of models for the digestion process and design of the

anaerobic digesters

(v) Design and evaluation of batch and continuous-flow digesters for the

digestion of MSW for biogas generation

(vi) Systems simulation to determine optimum performance parameters.


67

CHAPTER 3

DESIGN MODELS FORMULATION

A variety of models have been developed for the design of digesters,

and evaluation of digester operation and performance. However, these

models are generalized, and their use for the description of any digester

depends on the type, characteristics, and location of the waste. Hence the

necessity to formulate appropriate models for the design and assessment of

the digesters using MSW generated in Port Harcourt metropolis.

3.1 MATERIAL BALANCE OF THE ANAEROBIC


DIGESTION PROCESS

The basic design approach for an anaerobic bioreactor is the

formulation of models for the reactor process, based on an analysis of

material balances during the processes. Several literatures (Tchobanoglous

et al, 2003; Agunwamba, 2001; Kiely, 1998; Reynolds and Richards, 1996;

Tchobanoglous and Burton, 1991; Andrews, 1978, etc) state the general

form of a material balance expression as follows:

Rate of
Rate of
Rate of Appearance or Rate of material
Accumulation
 material flow  Disapperarance  flow out of
of materialin
int o Re actor of material due Re actor
Re actor
to Re action

(3.1a)

This is further simplified as:


68

Accumulation = Inflow + Net growth - Outflow (3.1b)

and considering the anaerobic digestion process, this expression can be

symbolically represented as:

d[ X ]
Vr  Q[ X o ]  Vr  net  Q[ X ] (3.2)
dt

d[ X ]
where = rate of change of microorganism
dt
concentration in the reactor measured in terms
of mass (mixed liquor volatile suspended solids),
massMLVSS/unit volume. Time

Vr = volume of reactor

Q = flowrate, volume / time

Xo = concentration of microorganisms in
influent, mass MLVSS/unit volume

X = concentration of microorganism in reactor,


mass MLVSS/unit volume

net = net rate of microorganisms growth, mass


MLVSS/unit volume time

[S ]
since net =  max [ X ]  kd [ X ] (3.3)
K s  [S ]

d[ X ]  [S ] 
then, Vr  Q[ X o ]  Vr   max [ X ]  kd [ X ]   Q[ X ] (3.4a)
dt  K s  [S ] 

 
Vr  Q [ X o ]  [ X ]  Vr   max
d[ X ] [S ]
or [ X ]  kd [ X ]  (3.4b)
dt  K s  [S ] 

This represents the general model for the anaerobic digestion process.
69

3.2 Application of the Model to Batch Reactor Processes

Applying the general form of the material balance expression to a

batch process where there is no flow (i.e. Q = 0), the first term of the right

hand side of the equation becomes zero;

i.e. Q ([Xo] – [X]) = 0 (3.5)

(a) Material Balance for Mass of Microorganism

d[ X ]  [ S ][ X ]
 Vbd  m Vbd  kd [ X ]Vbd (3.6a)
dt K s  [S ]

or (eliminating Vbd)

d[ X ]  [ S ][ X ]
 m  kd [ X ] (3.6b)
dt K s  [S ]

and this represents the mass balance for the mass of microorganisms in the

batch reactor.

(b) Material Balance for Total Substrate Utilization

The material balance for the total substrate utilization in a batch

process is equally given as;

d[S ] k[ S ][ X ]
Vbd   Vbd (3.7a)
dt K s  [S ]

or (eliminating Vbd)

d[S ] k[ S ][ X ]
 (3.7b)
dt K S  [S ]

where

k - maximum rate of substrate utilization per unit mass of cells


70

produced (mass/mass, time)

m
and k (3.8)
Y

The solution of equation (3.7b) is as follows:

dt K  [S ]
 s (3.9a)
d[S ] k[ S ][ X ]

dt Ks 1
  (3.9b)
d[S ] k[ S ][ X ] k[ X ]

K s dS dS
dt    (3.9c)
k[ S ][ X ] k[ X ]

t t Se Se
K s dS dS
t  0dt  S k[S ][ X ]   k[ X ]
So
(3.9d)
o

Ks  S  [S ]  [Se ]
so that t  ln o   o (3.10)
k[ X ]  S e  k[ X ]

This equation expresses the time, t, required to achieve a given fractional

conversion (), which is also called the time for batch digestion, obtained

from a computer solution of the equation using Simpson’s numerical

approximation, thus:

Let 'N' represent the range of integration, such that N = 10

( Se  So )
Then interval h (3.11a)
N

such that S(N) = So + (N x h) (3.11b)

Ks  S (N )
f (N )   (3.11c)
k[ X ]S ( N )  kd [ K s  S ( N )]
71

S(0) = f(0) + f(10)

S(1) = f(1) + f(3) + f(5) + f(7) + f(9) (3.11d)

S(2) = f(2) + f(4) + f(6) + f(8)

so that t
h
S (0)  4S (1)  2S (2) (3.12)
3

3.3 Models for the Batch Digester Design

The design of the batch digester involved the determination of the

following parameters (Kiely, 1998).

- Solids Retention Time (SRT), days

- Hydraulic Retention Time (HRT), days

- Volatile Solids Loading Rate, kg VS/m3/day

- Solids Production Rate, kg SS/m3/day

- Gas Production, m3 of CH4 of reactor/day

- Tank Configuration

- Mixing Systems

- Heating Systems

Kiely explains the necessity to provide for adequate solids residence times,

to include enabling the volatile solids to be fully degraded.

The design parameters are described below with the applicable

models where necessary.


72

3.3.1 Solids Retention Time (SRT)

The SRT is the average period of time required for the material to

remain in the system. It is usually regarded as the most critical parameter in

the system design; as it affects the treatment process performance, digester

volume, sludge production etc. (Tchobanoglous, et al, 2003). The SRT is

also frequently referred to as the mean cell residence time or vice versa

(Reynolds and Richards, 1996).

It is important to restate at this point that according to Kiely’s

classification of reactors, for the ‘first generation’ reactors of which the batch

reactor is a part, the SRT is equal to the HRT. He further states rather

emphatically that for digestion systems without recycle, the SRT is the same

as the HRT.

The SRT is defined mathematically as

mass of solids in tan ks , kg


SRT = (3.13a)
rate of solid removed kg / day

Reynolds and Richards (1996) state the symbolic representation as:

X
c  (3.13b)
X
where

c = mean cell residence time, (days)

X = amount of dry solids in the digester, (kg)


73

X = amount of dry solids produced per day in the digested sludge

(kg)

It is also important to note that with decreasing mean cell residence

time, c, there is a certain value of c below which the cells are washed from

the system faster than they can multiply, and therefore waste stabilization

does not occur. This critical value of c is called the minimum mean cell

residence time mc, and defined mathematically as:

1 YKS
  kd (3.14)
 m
C KS  S

where:

mc - minimum mean cell residence time

In many practical situations, S is much greater than K s that the equation

(3.14) can be rewritten as:

1 YKS
  kd (3.15)
 m
C KS  S

and the ratio of c to mc described as the process safety factor (SF), so

that

C
SF = (3.16)
 Cm

where:

SF - safety factor of the process

3.3.2 Volatile Solids (VS) Loading Rate (VSLR).


74

This is another very important factor relevant in the sizing of

digesters. It expresses principally the mass of volatile solids added per day

per unit volume of digester capacity.

i.e.

Volatile solids added daily , kg VS / day


VS loading rate = (3.17)
Working volume of the digester , m 3

The upper limit of volatile solids loading rates is typically determined

by the rate of accumulation of toxic materials, particularly ammonia or

washout of methane formers (WEF, 1998 in Tchobanoglous, et al, 2003).

This of course leads to very low methane production. On the other hand

excessively low volatile solids loading rates can result in designs that are

costly to build and are troublesome to operate.

Closely related to the VSLR is the volatile solids destruction, which

signifies the amount of volatile solids reduced during the digestion process.

Tchobanoglous et al (2003) explain that “the degree of stabilization obtained

is often measured by the percent reduction in volatile solids.

3.3.3 Digester Volume

The determination of the volume of a digester is usually dependent

on the volume of fresh sludge added daily, the volume of digested sludge

produced daily, and the required digestion time in days (Reynolds and

Richards, 1996). They state that additional volume provision must be made

for the supernatant liquor, gas storage, and storage of digested sludge.
75

To determine the volume of a batch digester in which the supernatant

is removed from the batch of digesting sludge as it is produced, it has been

found that the volume of the remaining digesting sludge versus the

digestion time (referred to as average volume of digesting sludge) is a

parabolic function (Reynolds and Richards, 1996). They also present the

parabolic function relating the various sludge-volumes as established by

Fiar et al (1968) thus:

2
V avg  V1  V1  V2  (3.18)
3

where

Vavg = average volume of digesting sludge m3/day

V1 = volume of fresh sludge added daily m3/day

V2 = volume of digested sludge produced daily, m3/days,

which is the same as the final sludge volume

t = retention time

Again the volume of total sludge in the digester (both digesting and digested

sludge) is given by

Vs = Vavg . td + V2 . ts (3.19)

where

Vs = total sludge volume, m3

td = time required for digestion, days

ts = time provided for sludge storage, days

Taking sludge volume to account for two-thirds (2/3) of the total volume

(Kiely, 1998), then the total volume of the batch digester will be:
76

Vbd = 3/2Vs (3.20)

where Vbd = total batch digester volume, m3

Substituting equations (3.18 and 3.19) into this equation (3.20) gives the

following relationships.

Vbd = 3/2(Vavg.td + V2.ts) (3.21a)

3   
V1  V1  V2 td  V2t s 
2
Vbd  (3.21b)
2  3  

3
Vbd   V1  2V2  td  V2ts  (3.21c)
2 3 

Vbd 
1
V1  2V2 td  3V2ts  (3.21d)
2

3.3.4 Digester Tank Configuration

Tank configuration primarily describes the shape of the digester, and

of course the relationship between the digester size parameters. There are

a number of different shapes adopted for the construction of anaerobic

digesters. Tchobanoglous and Burton (1991) say that “anaerobic digestion

tanks are either cylindrical, rectangular or egg-shaped”. Whereas these

different digester configurations have inherent advantages due mainly to

their efficiency or otherwise of mixing of the contents, cleaning of the inside

of the digester walls, and unloading the digester, eliciting their use in

different circumstances, Kiely (1998) states that anaerobic digester “tank

configurations are now primarily cylindrical with diameters of 5 to 50m and

heights of 3 to 25m”. These size-ranges clearly puts the digester diameter-


77

to-height ratio at 2:1, which is understandably so for increased top-surface

areas of digester contents for enhancement of decomposition.

This design adopts the cylindrical configuration, particularly against

the background of the type of waste, which has large particle sizes.

Therefore, with a known volume of digester tank the diameter and height of

the digester (as a cylindrical tank) could be determined as follows:

Vbd = Abd x Hbd (3.22)

where: Abd - cross - sectional area of batch digester

Hbd - height of batch digester

But A, the cross-section area of a cylinder is given as:

 Dbd 2
Abd  (3.23)
4

where: Dbd - diameter of the batch digester

Dbd
So, since H bd  it follows that the digester volume will now be.
2

 Dbd 2 Dbd
Vbd  . (3.24a)
4 2

or

 Dbd 3
Vbd  (3.24b)
8

8Vbd
from where Dbd  3 (3.25)

3
8Vbd
and therefore H bd   (3.26)
2
78

3.3.5 Model for Digester Mixing

Mechanical mixing using an auger, also referred to as a screw-

conveyor, has been prescribed for the batch design because of the

relatively high solids content (in terms of total solids concentration) of the

mixture. Writing about the screw-conveyor Raymus (1997) says “Almost any

degree of mixing can be achieved with screw-conveyor flights cut,. . .”. He

explains that the power required for the screw-conveyor is composed of two

components, namely: that necessary to drive the screw empty and that

necessary to move the material. Whereas the first component is a function

of conveyor length, speed of rotation, and friction in the conveyor bearings,

the second is dependent on the total weight of material conveyed per unit

time, conveyed length, and depth to which the trough is loaded. The latter

power item is in turn a function of the internal friction and friction on metal of

the conveyed material.

The Table A.IV in Appendix IV shows a wide range of capacities and

power requirement for various sizes of screws handling up to 801 kg/m 3 of

material of average conveyability. It is therefore necessary to determine the

capacity of the auger to be able to use the Table to select appropriate power

unit to drive the auger.

The productivity of an auger transporter is given by the following

equation (Evseev et al, 1972):


79

 D 2  d 2 
QAUG  60   S  n   (3.27)
4

where:

QAUG - capacity of the auger, tons/hr

D - outer diameter of auger, m

d - inner diameter of auger (ie diameter of shaft), m

S - pitch of auger, m

n - speed of auger, rpm

 - specific weight of material to be moved, tons/m3

 - coefficient of fill of material  0.25 to 0.40

Normally, S = (0.75 to 1.0)D and d = (0.25 to 0.35)D

When the capacity is determined with the appropriate auger

parameters, then the Table would be used.

3.4 Application of the Model to CSTR Processes

To apply the general form of the material balance expression to a

CSTR will actually be to adopt the form of the general model because the

overall characteristics of the CSTR including its flow regime was considered

in the development of the model.

(a) Material Balance for Mass of Microorganism

The mass balance for the mass of microorganisms in a complete-mix

reactor, will, therefore be:


80

d X   S  X  k X 
Vc  QX o   X   Vc   max  (3.28)
K S  S 
d
dt  

d X 
Now, considering steady state condition, i.e.  0 , and assuming that
dt

Xo is negligible at the commencement of the process, then equation (3.28)

becomes


QX   Vc   max
S  X   k X  (3.29a)
K S  S 
d 
 

or (eliminating [X])


Q  Vc   max
S   k  (3.29b)
d 
 K S  S  

Vc
But is defined as the mean cell residence time (c). Therefore
Q

1
  max
S   kd (3.30a)
C K S  S 

1  max S   k d ( K S  S )
or  (3.30b)
c K S  S 

(b) Material balance for total substrate utilization

The mass balance for substrate utilization in a CSTR will be given as

d S 
Vc  QSo   rSVc  Q S  (3.31a)
dt
81

d S 
Vc  QSo   S   rsVc (3.31b)
dt

where rs - rate of substrate utilization, and defined mathematically as

k S 
rs   X  (3.32)
K S  S 

So, substituting for ‘rs’ in equation (3.31b) from equation (3.32) and

 d S  
assuming steady – state condition  i.e.  0  gives
 dt 

k S 
Q SO   S   X Vc  0 (3.33a)
K S  S 

 V   k S  
or [ So ]  [ S ]   c   X  (3.33b)
 Q   K S  S  

k S 
 So   S    h X  (3.33c)
K S  S 

where h = hydraulic retention time which is the same as the mean

cell residence time (c) for no cell recycle anaerobic system, and describes

the digestion time for the CSTR, such that

([S o ]  S e )(K s  [ S e ]
 h  (3.34)
k[ S e ][ X ]
82

From equation (3.30a)

S  
1  1 
  kd  (3.35)
K S  S   max   C 

And from equation (3.33c),

S  
[ So ]  [ Se ]
(3.36)
K S  S  h k [ X ]

Therefore, combining equations (3.35) and (3.36) gives

S   [Se ] 
1  1 
  kd  (3.37a)
h k[ X ]  max   C 

h k 1 
So   S     kd X  (3.37b)
 max  c 

 max So   S  c
or X   (3.38a)
k 1  kd c h

Y So   S  c
or X   (3.38b)
1  kd c h

Also, rearranging equation (3.35) and substituting for max from

equation (3.8)

K S 1  k d  c 
S   (3.39)
 c Yk  k d   1

again from equation (3.37b)

So   S  
1  1  kd c 
  (3.40a)
X  h Y   c 
83

or
So   S  
1 k
 d (3.40b)
X  h Y c Y

and from combining equations (3.33c) and (3.40a), then

c Ks 1
  (3.41)
1  k d c S Yk Yk

Considering the fractional conversion () of the substrate (S), defined as:

S   S 
  o
(3.42a)
S  o

then S  
S  (3.42b)
o
1  

Substituting for [So] in equation (3.33c) gives

S   S    h
k S 
X  (3.43a)
1    K s  S 

S   S 1     h
k S 
X  (3.43b)
1    K s  S 

h 
K s  S S   S 1   
(3.44a)
k S X 1   

  K S  S S 
h 
1     k X S  
(3.44b)

  K s  S  
h 
1     k X  
or (3.44c)
84

3.5 Models for the CSTR Design

The development of models for the CSTR is strictly based on the

requirements of a high rate, low solids processing where there is

“continuous mixing and continuous or intermittent sludge feeding and sludge

withdrawal, and the contents are in a homogenous state (Reynolds and

Richards, 1996). The fixed cover variation will also be considered for the

design. Of course this latter consideration is even justified by Reynolds and

Richards (1996) when they stated that “high-rate digesters usually have

fixed covers”.

For the purpose of the comparative evaluation in this study, only

models different from those of the batch digester will be presented here.

3.5.1 CSTR Volume

Several literatures, Tchobanoglous et al; 2003, Kiely, 1998, Reynolds

and Richards 1996 etc; state that the volume of a CSTR is defined as the

product of the rate of flow of the medium and the hydraulic retention time.

This is stated mathematically as:

Vc = Qh (3.45)

where; Vc = overall volume of the continuous digester, m3

Q = influent sludge flow rate, m3/day

h = hydraulic retention time, days.

Substituting for h from equation (3.44c) in equation (3.45) gives


85

    K S  S  
Vc  Q     (3.46)
 1     k X  

3.5.2 CSTR Dimensions and Configuration

The configuration of tanks for continuous – flow anaerobic processes,

according to Kiely (1998), tends to be ‘circular or square in shape and rarely

rectangular’.

For the development of the CSTR dimensions in this work, a circular

cross-sectional configuration is adapted to guarantee effective and efficient

mixing as Tchobanoglous et al, 2003 state that “the conditions will depend

on the reactor geometry and the power input”.

3.5.2.1 Diameter and Height of CSTR

The diameter and height of the tank is related to the tank volume by

the following equation;

Vc = Ac x Hc (3.47)

where

Ac = cross-sectional area of the tank, and

Hc = height of the tank.

For a circular cross-section, the area of the cross-section is defined

mathematically as:

Dc 2
Ac  (3.48)
4
86

So, substituting for Ac in equation (3.77) gives

1
Vc  Dc 2 H c (3.49)
4

Therefore, adapting the established relation between the digester tank

diameter and height as being in the ratio of 2:1, so that D c = 2Hc or Hc =

Dc
, then
2

1
Vc  Dc
3
(3.50)
8

8Vc
or Dc  3 (3.51)

Substituting for Vc from equation (3.46) gives:

8     K S  S  
Dcf = 3 Q    (3.52)
  1     k X  

and

2Q     K S  S  
Hc  3    (3.53)
  1     k X  

3.5.3 Gas Diffusion Rates For CSTR Mixing

Stafford et al (1980) presents the formula for the rate of gas diffusion

as given by Morgan and Neuspiel, thus:

Q = K Ve3 Dc (3.54)

Where: Q - gas discharge rate (m3/s)

K - proportionality constant
87

Ve - representative velocity (m/s)

They also reported that a velocity value of 0.18 m/s was found

appropriate for effective mixing of digester content and prevention of scum

formation. Also K values between 10 and 13 gave satisfactory mixing

performance for conventional-size digesters.

3.6 GENERAL MODELS FOR BATCH AND CONTINUOUS


DIGESTERS

3.6.1 Models for Digester Heat Requirements

Generally in biological reactors heat may be added or removed from

the medium to achieve process stability. This is usually necessitated by the

following considerations during the reaction.

(i) Heat generated in substrate conversion may be inadequate to

maintain the desired level of temperature, in which case heat must

be added to the system,

(ii) On the other hand, the heat generated during substrate conversion

may be in excess of heat required to maintain the desired

temperature level, such that the system may need to be cooled.

Clearly these two considerations are contingent upon the heat of

reaction (H) defined as “the difference in energy required to break the

bonds in the reactants when compared to the energy required to break the

bonds in the products” (Willis, 2000). The heat of reaction is either

exothermic (when heat is given off by the reaction) or endothermic (when


88

heat is absorbed by the reaction). In this anaerobic digestion, the process

is considered exothermic since the microbes involved in the decomposition

exercise are living organisms that respire, and expected to give off heat in

the process. Bailey and Ollis (1986) put it more aptly when they said “the

microbial reactors involve a heat source arising directly as a result of

microbial metabolism”.

It is therefore imperative to ascertain the effect of the heat generated

on the temperature of the system, i.e. to know whether or not it can maintain

the desired temperature and hence whether heat would be added to the

system or not.

3.6.1.1 Energy Balance for the Digestion Process

The overall heat requirement for the digester is better appreciated

from the perspective of a general energy balance of the entire system, thus:

Rate of Rate of Rate at which Rate of change


Energy  Energy  heat is added  of energy within
flow in flow out due to reaction system

(3.55a)

This is represented mathematically as:

Q1C p Tin  Tdatum   Q2C p T  Tdatum    krVXH   MC p


dT
(3.55b)
dt

where

Q1 = inflow mass flow rate, kg/s

Q2 = outflow mass flow rate, kg/s

Cp = specific heat capacity of material


89

kr = reaction rate term

Vr = volume of reactor m3

X = concentration of material in reactor

H = exothermic heat of reaction

M = mass of material in the reactor, kg

Tin = temperature of inflow material oC

Tdatum = external reference temperature, oC

T = temperature of material in the reactor, oC

t = time, s

If Q1 = Q2, then the energy balance differential equation becomes

QC p Tin  T   kVXH  MC p
dT
(3.56a)
dt

 QC p Tin  T   kVXH
dT
or MC p (3.56b)
dt

dT
For steady state isothermal condition,  0
dt

:. QC p Tin  T   krVXH  0 (3.57a)

or QC p Tin  T   krVXH (3.57b)

The right hand side of the equation represents the rate of heat

generation due to the reaction, described by Bailey and Ollis (1986) as the

heat generation rate from cell growth, Qgr,


90

3.6.1.2 Microbial Heat Generation

For the batch reactor, the instantaneous microbial heat generation

rate is given by

1
Qgr  Vb [ X ] (3.58)
Y

while the corresponding equation for a continuous – flow isothermal reaction

at steady state is

Qgr  Vc [ X ]  [So ]  [S ]YS Q  [ X ]kd (3.59a)

Qgr
 [ S o ]  [ S ]YS D 
[ X ]k d
or (3.59b)
Vc Vc

These equations show that whereas Ys may depend on the culture

age in a batch reactor, it may depend on the dilution rate in a continuous-

flow system.

where

Qgr = heat generation rate from cell growth, kcal/hr

Vr = volume of reactor, m

 = specific growth rate of biomass, day-1

X = concentration of biomass, mg/l

Y = yield of cell mass per kilocalorie of heat evolved


g/kcal

YS
Y  (3.60)
H S  YS H C
91

where

Ys = yield of cell mass per mass of substrate

consumed, g cell/g substrate

Hs = heat of combustion of substrate, kcal/g

Hc = heat of combustion of cell mass, kcal/g

3.6.1.3 Digester Heat Requirements

Having determined the exothermic heat of reaction, in the event that

it is not enough to sustain the desired digester temperature, then additional

heat must be required by the digester. The purpose of the additional heat

for the digester is outlined in section 4.4.8.

(a) Heat Required to Raise Temperature of Incoming Sludge

The heat required to raise the temperature of the in-flowing material

to the temperature of the digester has been given by Reynolds and

Richards (1996) as:

x Td  TS  x
100 1
QS  P x Cp (3.61)
ps 24

where

Qs = heat required for incoming sludge, J/h

P = mass of fresh dry sludge solids added per day

ps = percent dry solids in the fresh sludge

Td = temperature in the digester oC

Ts = temperature of the fresh sludge oC


92

Cp = specific heat constant (4200J/kg-oC)

(a) Heat Required to Compensate for Heat Losses

During the process of digestion, heat would be lost through the

digester surfaces, viz the walls, floor and roof. These heat losses through

the boundaries of the digester are usually determined using the empirical

relationship developed by Fourier for one dimensional steady heat flow

through a composite wall and expressed by Tchobanoglous et al, 2003 as;

Qh = UAT (3.62)

where

Qh = heat required to compensate heat losses, J/s

U = heat transfer coefficient for heat loss surfaces,

J/m2.s.oC

A = area of heat loss surfaces,m2

T = change in temperature of heat exchange surfaces, oC

The components of the losses are as follows.

(i) Heat loss through the wall

Let the heat loss through the wall be denoted by Q w. It is defined

mathematically as:

Qw = UwAw(Td -To) (3.63)

where

Qw = heat loss through the wall J/s

Uw = wall heat transfer coefficient, J/m2.s.oC


93

Aw = area of the wall, m2

Td = inside temperature of digester oC

To = outside or external temperature, oC

(ii) Heat loss through the floor

This is given as

Q f  U f . A f .Td  Te  (3.64)

where

Qf = heat loss through the floor, J/s

Uf = floor heat transfer coefficient, J/m2.s.oC

Af = area of floor, m2

Te = temperature of earth beneath the floor,oC

(iii) Heat loss through the roof

This is given as

Qr = Ur.Ar.(Td - To) (3.65)

where

Qr = heat loss through the roof, J/s

Ur = roof heat transfer coefficient, J/m2.s.oC

Ar = area of roof, m2

Therefore heat required to compensate for heat losses will be

Qc = Q w + Q f + Q r (3.66)

3.3.1.3.1 Total external heat required

Total external heat required by the digester will be

QT = Qs + Qc (3.67)
94

3.3.1.3.2 Net heat required by digester


The digestion process is exothermic, given off a certain amount of

heat defined by equations (3.58) and (3.59b). Therefore, the desired

quantity of heat for the system described as the heat required by the

digester reduces to:

QN = QT - Qgr (3.68)

For the batch digester,

1
QN  QT  Vb [ X ] (3.69)
Y

and for the CSTR

QN  QT  Vc [ So ]  [ Se ]Ys D  [ X ]kd (3.70)

Defining the specific digester heat requirement as the heat required per unit

volume of the digester, then:

QN
 qt (3.71)
Vr

where

qt = total specific heat required by digester

3.6.1.4 Overall Heat Transfer Coefficient

The overall heat transfer coefficient comprises of film coefficients of

liquid inside the digester, the gas above the liquid, the thermal conductivity

of the wall of the digester and/or insulation material, and also the film

coefficient of the air outside of the digester wall, the earth beneath the floor

of the digester, and radiation heat transfer coefficient.


95

Rogers and Mayhew (1982) give the expression for the overall heat

transfer coefficient per unit area of tube wall as:

 r 
In i 
   i 1  
n
1 1 r 1
 (3.72)
U' 2ro ha i 1 2ki 2rn h6

where

U’ = overall heat transfer coefficient per unit length of

tube

ha = film coefficient of fluid inside the tube

ro = inside radius of tube

ri = subsequent radii of concentric tubes

ki = thermal conductivity of material of tube wall

rn = outside radius of tube

hb = film coefficient of fluid outside the tube

Relating the overall heat transfer coefficient to a typical shell and

tube heat exchanger, Kiely (1998) cites Coulson and Richardson (1991)

who proposed the following relationship, introducing the effect of fouling

factors thus:

1 1 1 d /d d d d d
   d o In o i  o i  o i (3.73)
U' hb hod 2k i hid ha

where,

hod = outside fouling factor, W/m2 oC

do = tube outside diameter, m


96

di = tube inside diameter, m

hid = inside fouling factor, W/m2 oC

Typical values of the film coefficients, particularly as influenced by the

fouling factors have been presented by Tchobanoglous et al, 2003.

3.6.1.4.1 Wall Heat Transfer Coefficient


This is determined with respect to the transfer of heat from both the

liquid body (sludge) and the gas inside the digester.

Therefore, let Uwl and Uwg be the heat transfer coefficients for the

liquid and gas portions of the digester wall, so that,

1 1 n
xi 1
U wl

hal
  i 1 ki

hb
(3.74)

and

1 1 1 n
xi 1
U wg

hag

hb
 
i 1 ki

hc
(3.75)

where,

hal = film coefficient of liquid in the digester

hag = film coefficient of gas in the digester

hc = coefficient of conductance of gas above liquid in


digester

3.6.1.4.2 Floor Heat Transfer Coefficient

This is defined mathematically as

1 1 n
xi
Uf

hal
 i 1 ki
(3.76)
97

Here, ki includes thermal conductivity of earth beneath the digester floor,

perhaps as an insulator.

3.6.1.4.3 Roof Heat Transfer Coefficient

This is given as:

1 1 1 n
x 1
   i  (3.77)
U r hag hb i 1 ki hC

So depending on the nature and type of lagging for different digester

surfaces, Uwg  Ur.

3.6.1.5 Heating of the Digester

Contemporary anaerobic digesters use external heaters where “the

sludge is pumped at high velocity through the tubes, while water circulates

at high velocity around the outside of the tubes” (Tchobanoglous et al,

2003). They explain that the circulation promotes high turbulence on both

sides of the heat transfer surface and results in higher heat transfer

coefficients and better heat transfer. Additionally, according to Reynolds

and Richards (1996), the use of an external heater, which heats the pumped

sludge outside the digester controls the problem of caking associated with

internally heated digesters, since any sludge caking will occur in the pipes in

the heat exchanger. It allows for de-caking of the pipes and enhances

digester maintenance and high heat transfer efficiency.

A schematic of an externally heated digester is shown below.


98

Gas dome

Gas

Hot Water Heat Exchanger


Heated Sludge

Heated Supernatant
Sludge Gas
Line Liquor Line

Hot Water Pump


Heater
Recycled
Sludge

Digested

Gas Furnace Flame Trap Sludge Line


Fresh
Sludge Waste Gas

(to flare)

Figure 3.1: Schematic of an externally heated digester

Adapted from Reynolds and Richards (1996)

3.6.1.5.1 Heater Requirements

Evidently the heater or heating source requires a given quantity of

water to be circulated around a portion of the surface of the digester.

(i) Required Quantity of Water for Heating

Generally, the specific digester heat requirement can be defined

mathematically as

qt  M w C w Td  To  (3.78)

where;

Mw = mass flow rate of hot water


99

Cw = specific heat of hot water

qt
 Mw  (3.79)
C w Td  To 

(ii) Required Surface Area for Heat Exchange

The heat flow across the walls of a tube has been given by equation

(3.62) as

QN = UAT

For the externally heated digester, ‘A’ in the above equation is the surface

area with which heat is exchanged with the incoming sludge stream to the

digester; and U = Uw.

Therefore,

QN
A = (3.80)
U Td  TS 

With a predetermined radius of heat exchanger pipe (or the pipe

through which the incoming sludge flows), the length of the pipe in the hot

water bath can be determined using the trite relationship of:

A = 2rl (3.81a)

A
or l = (3.81b)
2r

where

r = radius of heat exchanger pipe

l = length of pipe in water bath


100

3.6.2 Model for Digester Cost Estimation

Initial capital investment in the manufacture of anaerobic digester is

usually estimated as a function of an existing digester with similar

characteristics. This is why it is somewhat difficult to achieve adequate

estimation, as there are very few commercial digesters in operation.

However, Peter and Timmerhaus (1981) presented the power factor

equation, which gives a fairly accurate estimate. The equation is of the form:

0.6
C 
X  Y  2  (3.82)
 C1 

where: X = cost of proposed digester

Y = current cost of existing digester

C1 = capacity of existing digester with known cost

C2 = capacity of proposed digester

This empirical formulation considers the price index of the base year

of manufacture of the digester with known cost. So the actual capital cost of

the proposed digester would be obtained relative to the price index of the

current year, in this case 2005.

The first step in the estimation process is to use the relative price

indices of the respective years to determine what would have been the

actual cost of the digester with known cost and capacity in the current year.

Whitesides (2001) explains this when he said, “cost indices must be used

when basing the approximated cost on other than current prices”. He states

that the known cost of the digester must be multiplied by the ratio of the cost

index of the current year to that of the base year.


101

This translates mathematically thus:

I 
Y  Co   (3.83)
 Io 

where: Co - base cost of existing digester

I - current price index

Io - base price index

When this is related to the power factor equation, then, the estimated cost of

the proposed digester will be:

0.6
 I  C 
X  Co   2  (3.84)
 I o  C1 

Mosey (1980) in Hobson, et al (1981) reported a 230 m3 digester,

which cost 94,000.00 pounds as at 1980. So, using the cost indexes table of

Marshall and Swift in Peter and Timmerhaus (1981) (see Appendix V), and

extrapolating from 1980 to 1989, and up to 2005, the current price index for

2005 was obtained as 1490.56; such that Io = 560 and I = 1491.

 1491
Therefore Y  94000 
 560 

= £250,275.00

= N 56,311,875.00 (@ N 225.00 per pound)

3.6.3 Models for Methane and Biogas Production

The amount of methane (CH4) produced in the anaerobic

decomposition of organic materials is usually determined using the


102

stoichiometric equation developed by Buswell (1939) as presented by

Andrews (1976), which was also represented by Buswell and Mueller (1952)

in Kiely (1998) as:

 a b n a b n a b
C n H a Ob  n    H 2 O      CO2     CH 4
 4 2 2 8 4 2 8 4

(3.85)

It is important to state that this expression is actually just a balance,

which complies with conservation of mass without necessarily involving the

feasibility of the reaction or the speed with which it may take place

(Trambouze et al, 1988).

The prediction of the volume of biogas produced from the anaerobic

decomposition of organic substrates is simplified by first of all predicting the

volume of methane gas produced during the process, and after which the

stoichiometric relationship is used to determine the volume of biogas.

From the stoichiometric relationship expressed above, considering an

organic matter represented as C6H10O5, such that n = 6, a = 10 and b = 5,

then the following chemical relationship can be established thus;

C6H10O5 + H2O  3CO2 + 3CH4 (3.86)

Knowing that the molecular weights of C6H10O3 = 162, CO2 = 44, and CH4

= 16, then from equation (3.86)

1 mole of C6H10O5  3 moles CH4

1 kg C6H10O5  1000g  1000g/162g/mole


103

= 6.173 moles

6.173 moles C6H10O5 will yield (6.173 x 3) = 18.519 moles CH4

Relying on Avogadro’s postulation:

1 mole of CH4  0.0224 m3

Therefore, 18.519 moles CH4  18.519 x 0.0224

= 0.415 m3

This shows that 1 kg of the organic matter would yield 0.415 m 3 of CH4,

which accounts for the difference in concentration of the MSW as it is

digested from So to Se. Therefore, the unit of CH4 generated will be

represented as:

kg/m3 of CH4 produced = 0.415 (So – Se) (3.87a)

Then with a conversion factor (M = 1.43) of COD to VS the equation

becomes

Cm = 0.415M (So - Se) (3.87b)

where Cm - actual concentration of dissolved methane gas

3.6.3.1 Volume of Methane Produced

The weight of gas usually transferred from a liquid to a gas region

occurs across a liquid gas interface with the interfacial area in this case

equal to the cross-sectional area of the digester. Therefore, the rate of

diffusion in the gas-liquid contact system is easily described by the diffusion

equation, thus:

dCg
 KL
A
Cst  Cm  (3.88)
dt VL
104

Which upon integration becomes

Cg  K L
A
C st  Cm t (3.89)
VL

dC
Where - rate of diffusion, kg/m2 day
dt

KL - diffusion coefficient

A - interfacial gas transfer area

VL - volume of liquid

Cst - saturation concentration of gas in the liquid

Cg - concentration of methane gas in gas collector

The saturation concentration of the gas is given as:

Pm
C st  (3.90)
Hc

Where Pm = partial pressure of the methane gas

Hc – Henry’s constant

Graef and Andrews (1973) suggest that for mesophilic digester Henry’s

constant has been found as 3.25 x 10-5 mol/ (l/mmHg).

The mass of methane gas produced is given as:

Mm = Cg x Vg (3.91)

And the volume of methane produced becomes

Mm
Vm  (3.92)
m

Where Mm – mass of methane gas, kg

Vg - volume of gas collector, m3


105

m - density of methane gas, kg/m3

Vm - volume of methane gas produced, m3

Perry and Green (1997) give the density of methane as 0.644kg/m 3

3.6.3.2 Volume of biogas produced

Considering the percentage composition of the biogas product to be

60% methane and 40% other constituents, then the total volume of biogas

produced will be:

100
Vt  Vm  (3.93)
60

3.6.3.3 Volume of methane produced per unit volume of reactor

This is given as:


Vm
Vmr  (3.94)
Vb

3.6.3.4 Volume of methane per volatile solids added

This is given as:

Vm
Vmv  (3.95)
VS

3.6.3.5 Volume of methane per unit of biodegradable VS

This is given as:

Vm
Vmb  (3.96)
1  R VS
3.6.4 Characteristics of Waste Effluent Concentration

3.6.4.1 Effluent Volatile Solids Concentration


106

This is defined by the following relationship

Se = Seb + RSo (3.97)

where Se - effluent VS concentration, mg/l

Seb - effluent biodegradable substrate

concentration mg/l

Concentration of refractoryVS
R  Re fractory fraction  (3.98)
Concentration of inf luentVS

3.6.4.2 Effluent Biodegradable Substrate Concentration

This is obtained as follows:

Seb = Sib ( 1 - ) (3.99)

where Sib - influent biodegradable substrate concentration, mg/l

 - fractional conversion

3.6.4.3 Percentage Stabilization

The efficiency of the waste stabilization system for biogas generation

has been presented by Viessman and Hammer (1993) as:

E
So  Se 100 (3.100)
So

Where E - efficiency of removal of biodegradable waste load.


107

CHAPTER 4

EXPERIMENTAL AND COMPUTATIONAL WORKS

4.1 INVESTIGATION TECHNIQUES

This research involved investigations carried out both in the field,

which is the entire metropolis of Port Harcourt where the municipal solid

waste (MSW) is found, and the laboratory, using the ‘Reaction Kinetics

Laboratory’ of the Chemical/Petro-Chemical Engineering Department of

Rivers State University of Science and Technology.

4.1.1 Materials used for this study

The materials used for this study included those for field and

laboratory experimentations.

(a) Materials for Field Investigation: - The materials used for the

field investigations are as follows: standard refuse bins/bags, measuring

tape, weighing scale, compactors, wheelbarrows, rain boots, hand gloves,

gas masks, spades/shovels/forks, MSW, and knives.

(b) Materials for Laboratory Investigation: - The materials used

for the laboratory investigations included: weighing balance, moisture

content cans, oven dryer, digital pH meter, thermometers, batch reactors,

hose, graduated (measuring) cylinders, gas burettes, glue, retort stands,

and clamps.
108

4.1.2 Mode of Data Collection

Relevant data for the study were obtained using both primary and

secondary sources of information. The primary source includes mainly

participant observation, field surveys, laboratory and field experimentations,

and discussions with key actors, particularly in the area of solid waste

generation/management. Discussions were held with officials of the Rivers

State Environmental Sanitation Authority, which is responsible for refuse

disposal in the metropolis, Federal Office of Statistics, Port Harcourt Office,

National Population Commission, Port Harcourt office, Public Health

Department of the Rivers State Ministry of Health, and the Port Harcourt

City Local Government Council. On the other hand, the secondary source

included the use of literatures, maps, survey reports and public and private

records.

4.1.3 Method of Data Analysis

Field experimental data generated in this study were subjected to

simple engineering analysis, using some mathematical formulae as

presented in the course of this report, to establish the peculiar

characteristics of the municipal solid waste in Port Harcourt metropolis.

The laboratory experimental data were analysed using the differential

method of analysing reactor experimental data, where the data is plotted on

a curve and compared with the rate equation expected to describe the

process under investigation. With the data and the rate equations, based on
109

Monod kinetics, the kinetic coefficients describing the behaviour of the MSW

in processing were established.

On the overall the Microsoft visual basic programming software was

used to develop a programme for the simulation of the operation of the

digesters, incorporating the Simpson's mathematical formulation (otherwise

called the Simpson's rule); and appropriate mathematical model equations

for the optimization of the digesters’ operations were developed by

statistical evaluation of simulation results, using the Microsoft chart editor.

4.2 FIELD INVESTIGATION TECHNIQUES

4.2.1 Description of Study Area

The study area is the Port Harcourt metropolis. The map of Port

Harcourt (Appendix X) was used to delineate the study area so that waste

load within the metropolis was ascertained as accurately as possible.

Port Harcourt is the capital of Rivers State and the nerve center of

industrial activities, particularly oil and gas activities, in Nigeria. For this

reason, the metropolitan city is heavily populated. Records available at the

Federal Office of Statistics, 2003, put the current estimated population of

Port Harcourt at about 1,356,000.

Port Harcourt lies within latitude 05o 21’N and longitude 06o 57’E with

a mean annual rainfall of 2280mm and average minimum and maximum

temperatures of 25.1oC and 30.3oC respectively (Fubara–Manuel et al,

2000). So, temperatures and humidity are relatively high throughout the

year. The area is characterized by two distinct seasons viz, the wet and dry
110

seasons with 70% of the annual rains falling between April and August, and

22% spread over the three- (3) months of September to November. The

driest months are from December to March while the soil type consists

mainly of poorly drained silt clays mixed with sand (Ayotamuno et al, 2000).

4.2.2 Determination of Solid Waste Load in Port Harcourt

To determine the rate of waste generated in Port Harcourt

metropolis, the city was delineated into zones in line with the services of the

private refuse disposal companies, a procedure adapted from Ayotamuno

and Gobo (2004). The companies have different sizes of disposal trucks,

and indeed loaded them disproportionately. Field survey of the four official

dumpsites was undertaken three days (Monday, Wednesday and Saturday)

a week for four (4) weeks, noting the number of times of dumping of refuse

by the various companies and from the various designated zones. The sizes

of regular trucks used, while loaded, i.e. with the load configuration shown

in Figure 4.1, were also measured using a 30-meter tape. Because of the

overloading of the trucks and the eventual disconfiguration of the load out of

the trailers of the truck, the shape of the trailers with load is then as shown

below.

Waste Load

Trailer

Figure 4.1: Shape of Disposal Truck Trailer with load


111

With this shape, the regular volume formula of a product of length,

width and depth (V = L x b x d) no longer applied. So the waste load per

truck was then estimated using the Simpson’s rule for the estimation of

volumes, also called the Prismoidal formula. The rule states as follows:

Vl =
d
A1  4 A2  2 A3    4 An 1  An  (4.1)
3

d
which is also the same as saying (First area + Last area + 4 x sum of all
3
even areas + 2 x sum of all odd areas)

It is important to note that this rule applies only when there is an odd

number of cross-sections.

where: Vl - volume of waste load in refuse truck (m3)

d - distance between the ordinates formed (m)

A 1 – An - areas of each cross-sectional unit or strip(m2)

With this estimation, it was possible to determine the average daily

volume of waste generated in the city. When this was related to the

estimated population of the city, then the average per capita rate of refuse

generation in the metropolis was obtained.

4.2.3 Sources of the Waste

The wastes investigated are those generated from domestic and

commercial centers, in keeping with the scope of this research. This was

established by the operational schedule of the refuse disposal companies,

noting the various waste receptacles from where they collected wastes.
112

4.2.4 Composition of the Waste

The composition here refers to the physical components of the

wastes from the different sources. This became necessary since the wastes

occur in composite form, and the different components have varying organic

and inorganic nature; and particularly because only the organic component

of the waste is relevant to this work

To sort the waste into its components, standard refuse bins were

used to collect and measure the waste by volume, which was then put on a

weighing scale to determine the mass of the waste, after knowing the mass

of the refuse container.

One (1) container-waste-load of total volume 0.098m3 was collected

from each waste receptacle, for designated 10 sites randomly selected

throughout the metropolis, given a total of ten (10) container-waste-loads of

total volume 0.98m3.

When the wastes were collected with the bin, they were weighed in

their composite form as-discarded, and then the same mass of waste was

compacted with manual compactor until the change in volume became

constant. The measured wastes were then sorted into individual

components on the bases of their organic and inorganic character. After the

sorting, both the organic and inorganic components were measured by

volume and weight, both ‘as-discarded’ and ‘as-compacted’. The results

were used in the determination of the ‘as-discarded’ and ‘as-compacted’

densities, and the “ratio of the ‘as-compacted’ density (c) to the ‘as-
113

discarded’ density (d) is the compaction ratio (r) (Sincero and Sincero,

1999), which was employed in the design of the digester for the waste.

c
i.e. r= (4.2)
d

4.3 LABORATORY INVESTIGATION TECHNIQUES

4.3.1 Moisture Content Determination

The moisture content of the waste is relevant in the estimation of the

calorific value of the waste and reactor sizing (Kiely, 1998). Sincero and

Sincero (1999) explain that the moisture content of organic matter must be

put at the optimum to subject it to anaerobic digestion, hence knowledge of

the moisture content is important. The over-drying method of determination

of moisture content was used to determine the waste moisture content; and

this was done on ‘wet and dry bases’ using the formula proposed by

Sincero and Sincero (1999), thus.

W
Pd  (100) (4.3)
Sd

W
Pw  (100) (4.4)
Sw

where Sd - mass of dry solid wastes (kg)

Pw - moisture content of waste on wet basis (%)


114

W - mass of moisture (kg)

Sw - mass of wet solid wastes (kg)

Pd - moisture content of waste on dry basis (%)

The oven-drying experimentation was at a temperature of 77oC.

4.3.2 Characteristics of the Waste

The characterization of the waste was done in terms mainly of the

chemical and biological properties. The chemical properties of the waste

were determined using the standard methods of proximate and ultimate

analyses, as contained in the 16th edition of the ‘Official methods of

analysis, Association of Official Analytical Chemist’, AOAC (1995).

The following parameters were obtained from the proximate analysis.

- Moisture content by percentage weight

- Volatile matter

- Fixed carbon, and

- Non-combustible fraction (ash)

From the ultimate analysis, the important and relevant chemical

elements in the waste-to-energy transformation were determined.

The energy content of the waste was determined using the Khan, et

al (1991) equation for the estimation of the energy content of MSW,

presented by Kiely (1998) as:

Ec = 0.051 [F + 3.6 (CP)] + 0.352 (PLR) (4.5)

where: Ec - Energy content of MSW, MJ/kg


115

F - % of food by weight

CP - % of cardboard and paper by weight

PLR - % of plastic and rubber by weight

The level of total solids (TS), volatile solids (VS) biochemical oxygen

demand (BOD) and chemical oxygen demand (COD) were determined

according to the specifications in the Standard Methods, and were used to

estimate the biodegradability and pollution potential of the organic wastes.

The biodegradability of the waste is defined by Kiely (1998) as:

BF = 0.83 – 0.028 LC (4.6)

where: BF - Biodegradability fraction expressed on a volatile

solids (VS) basis

LC - Lignin content of VS, % of dry weight

4.3.3 Reactor Experimentation

To investigate the kinetics of municipal solid waste as a lead to

determining the rate constants and other kinetic parameters for the design

of the anaerobic digester, laboratory-scale batch digesters were set up in

the Department of Chemical/Petro-Chemical Engineering laboratory.

Five batch-wise anaerobic digesters each of 5 liters volume were set

up for this experiment. The schematic of the experimental design layout is

as shown in Figure 4.2 below:


116

Screw
Gas Hose

Clamp

Retort
Thermometer stand

Burette
Lagging
Material Graduated
cylinder
Digester Water
(with waste)
Work
bench

Figure 4.2: Schematic of Reactor Experimentation Design Layout

The anaerobic digesters were actually improvised with large

bournvita cans that have been cleaned of their previous content. This

improvisation consideration was not just aimed at cost effectiveness, but

actually derived substantial impetus from the postulation of Hobson et al

(1981), that “with a batch digester a smaller experimental system may be

suitable as the digester has only to be loaded once and may not even need

to be stirred. One or two litres could be big enough”. The containers were

properly lagged with wool material of about 25mm thickness, to minimize as

much as possible the interaction between the temperatures inside and

outside of the digesters.

Two perforations were made on the cover of the digester through

which the gas hose and thermometer were fitted. The hose extending from
117

the digester top was connected to the tail of a burette, which in turn was

then partly immersed in water in a graduated cylinder.

The waste materials were processed (shredded and mashed), and

the digesters were then loaded with 2kg of organic MSW, which was diluted

to a 26.7% total solids (TS) concentration after metals, glass and other non-

biodegradable materials had been removed.

The percentage total solids (TS) of the mixture was determined after

the moisture content had been determined by the oven–drying method

using equation (4.4).

The pH was measured from a digital pH meter, and the substrate and

biomass concentrations, were respectively determined in terms of the

chemical oxygen demand (COD), and the mixed liquor volatile suspended

solids (MLVSS). From an ultimate analysis the carbon, (C) and nitrogen (N)

content were known from which the C: N ratio was computed. While the

Carbon was determined using the Wackley-Black method, the Nitrogen was

determined with the usual macro-kjedahl method.

The thermometer, which passed into the headspace of the digester,

was used to measure the temperature inside the digester; and the ambient

temperature was measured from maximum and minimum thermometers at

the same time. These temperature measurements were taken at 0800

hours and 1400 hours, and aimed at determining temperature variation

within and outside the digester, in order that proper digester insulation

considering also construction materials is determined.


118

After these initial measurements from the waste replications, the

digesters were then made airtight with glue and other adhesives, and the

set-up allowed to run. Each of the digesters was dismantled at intervals of 5

days, which gave the experimentation a total lifespan of 25 days. At each

dismantling, substrate (COD) and microbial (MLVSS) concentration

measurements were repeated. The full view of the reactor experimentation

set-up is shown in the Plate 4.1.

Plate 4.1: Full view of digesters experimental set-up


119

4.3.4 Application of Batch Experiment Data to the


Design of Continuous Digesters

The investigation of anaerobic digestion processes is easily

conducted in batch systems basically because of its non-flow processing

and its ability to handle relatively small quantity of no-mix materials.

Hobson, et al (1981) explain that a batch digester makes it possible for

small experimental systems to be set – up, as the digester has only to be

loaded once, and may not even need to be stirred. This view was

corroborated by Bailey and Ollis (1986) when they said that the batch

reactor is simple, needs little supporting equipment and is therefore ideal for

small-scale studies on reaction kinetics. Also in this connection, Levenspiel

(1999) asserts that it is easy to interpret the result of the experimental batch

reactor.

But how would the data generated from a batch experiment be

applied to the design of continuous digesters? Bailey and Ollis (1986) made

an implicit assumption that there is a connection between batch kinetics and

the kinetics of continuous reaction processes. They established a qualitative

relationship between batch and continuous processing times. According to

them if tb and tc represent respectively the batch and continuous holding

times required to convert material at concentration Co into concentration Cf,

then;

Cf
dC
tb  
Co
f (C )
(4.7)
120

V C f  Co
and tc   (4.8)
F f (C f )

These equations are better interpreted geometrically for both

ordinary and autocatalytic kinetics. Notably while ordinary kinetics involves

mainly non-biological materials where f(C) is a decreasing function of C: the

more C present the slower its rate of formation; in autocatalytic reactions,

like cell growth, f(C) increases with C.

1
Geometrically, a plot of against C is used to depict the holding
f (C )

times for the processes. Figure (4.3a) below is a schematic representation

of this plot, where tb is the area under the curve from Co to Cf, and tc is the

area of the rectangle with sides of length Cf – Co and 1


f (C f )

1
f (c )

1
f (c f )

Co TB

Co Cf C

(a) Ordinary kinetics


121

1
f (c )

tb
1
f (c f )

tc

Co Cr C

(a) Auto catalytic kinetics

Figure 4.3: Graphical representations of mean residence time

Adopted from Bailey and Ollis (1986); Levenspiel (1999).

These representations are made for both ordinary and autocatalytic

kinetics, where it is evident that while tb is less than tc in ordinary kinetics, it

is greater in autocatalytic kinetics.

However, a quantitative relationship is based on the assumption that

the fermentation can be described by only one variable, cell concentration,

say; such that if for the batch system,

dc
 f (c) (4.9)
dt

The slope of the batch concentration time curve at any given

concentration will give the value of ‘f’ at that particular concentration, C.


122

dc
slope = = f(c1)
C1 dt
c =c1

Figure 4.4: Determination of f(c) from batch data

Adopted from Bailey and Ollis (1986)

Consequently, from the reduced batch data a plot of f(c) against c can be

generated;
f(C)
D(c-co) = net rate of c removal by
inlet, effluent streams

f(c) = rate of c formation by reaction

F
slope = D
V

C
Co C* = concentration in
steady – state CSTR
Figure 4.5: Use of f(c) for CSTR design

Adopted from Bailey and Ollis (1986)


123

such that if for the continuous system,

f(c*) = D (c* - co) (4.10)

then the steady-state concentration (c*) can be determined by drawing a

straight line, representing the right hand side of the equation, with slope, D,

through the c axis at c = co. The intercept of this straight line with the f(c)

plot gives the continuous processing effluent concentration. In its reverse

form, the dilution rate required to achieve a known concentration can also

be determined.

Generally, although there are clear limitations to the use of batch

experiment data to design continuous processing systems, Bailey and Ollis

insist that batch data are useful in a qualitative sense for choosing a

processing configuration for a continuous reactor. They also say that

preliminary quantitative design estimates can be made for continuous

processes based on batch reaction information.

4.4 DIGESTER DESIGN CONSIDERATIONS

In the design of a suitable digester for the biodegradation and indeed

stabilization of the municipal solid waste, with the attendant production of

biogas a number of factors were considered, particularly against the

background of the type of waste, the rate of waste generation and the local

environmental conditions, like the ambient temperature, etc. Some

assumptions were also made for ease of design computations. However,

these assumptions, whenever used, were strictly based on veritable


124

engineering and environmental principles, often deriving from established

experimental data in the literatures. The following are the design

considerations.

4.4.1 Types of Digesters

A variety of digester types of course exist for the anaerobic treatment

of organic wastes. The choice of the type of digester depends on several

operational factors, including particularly the nature of waste to be treated;

and this assessment has usually been done in terms of the solids content of

the waste. The Oregon State Department of Energy (2002) in its

classification of types of digesters explains that ‘a covered lagoon digester’

is used for liquid manure of less than 2 percent solids; ‘a complete mix

digester is suitable for manure that is 2 percent to 10 percent solids’, and

the ‘plug-flow digesters are suitable for ruminant animal measures having a

solids concentration of 11 percent to 13 percent”. The type and solids

content of the waste they considered are such that the wastes are capable

of flowing on their own, or forming slurries with water and eventually flowing,

so could be used in a continuous operation.

MSW is predominately solid and non-flowing, and Kayhanian (1991)

in Kiely (1998) had suggested the use of ‘high solids digester’, which was

expounded by Hobson et al (1981) when they said “the solid-state

digestion” though still largely theoretical with only some small units having

being built, “is always a batch process”. But the continuous-flow digesters,

which are by nature low-solids digesters have also been employed to


125

generate methane from human, animal, and agricultural waste, and from the

organic fraction of municipal solid waste (Tchobanoglous, et al (1993);

although this had always required much more water to be added to the

waste in order to bring it to the hydraulic regime of the system.

Hobson et al (1981) reported the experimental digestion of domestic

garbage diluted with sewage to a total solids concentration of between 5

and 7 percent. The diagram below shows the flow process in a low-solids

digester using MSW.

Gas
separator CH4

CO2

Ferrous Heavy To other energy


materials fraction conversion units Digester gas

Municipal Schredder Magnetic Air Light Digester


solid separator classifier 60 oC
waste fraction

Pump

Slurry
Solids to Vacuum
landfill filter
Blending-mixing tank

Filtrate
Liquid-air
separator
Chemical or
sewage sludge
feed
To atmosphere

Figure 4.6: Flow diagram for the low-solids anaerobic digestion process
for the organic fraction of MSW

Adapted from Tchobaroglous et al, 1999


126

Hobson et al (1981) says that stirred – tank digesters have to deal

with slurries of between about 3 and 10% total solids (TS), much of which

are suspended solids. In their analysis of what they described as “solid’

feedstock”, including green vegetable matter of about 20% solid

concentration, they explained that “if these materials are to be used as feed

for a stirred tank digester, then they will have to be made into slurries.

According to them “experience suggests that a slurry of about 10% TS is

the maximum thickness that can be pumped and piped even if the particle

size is small, and 7-8% TS may be the maximum which can be handled by

the smaller pumps and pipelines.

So it will then be proper in this work to determine which of these two

variations of digesters, viz: the high and low-solids digesters respectively, is

best suited for the generation of biogas from the organic component of

MSW.

To attempt to optimize the digestion of MSW using the low-solids

digestion process in this work, the upper limit solids concentration of 10%

would be adapted because of the uniquely very high solids concentration of

MSW in Port Harcourt metropolis.

4.4.2 Temperature Control

Temperature is a very important factor affecting the success of the

digestion process, as indeed the activities of the anaerobes causing waste

decomposition are dependent on this parameter. The optimum temperature

ranges are the mesophilic, 85 to 100oF (30 to 38oC), and the thermophilic
127

120 to 135oF (44 to 57oC) (Tchobanoglous and Burton, 1991); “but the rate

of decomposition and gas production is rather sensitive to temperature, and

in general the process goes faster at high temperatures” (Steadman, 1975).

Despite this ‘thermophilic benefit’; the digestion process “becomes

increasingly unstable with rising temperatures” (Kottner, 2002), and requires

higher energy for heating, and produces poorer quality supernatant

containing larger quantities of dissolved solids etc. (Tchobanoglous and

Burton, 1991). However, Kiely (1998) insists that “most digesters now

operate at mesophilic temperatures where good stability and gas production

results”, which is why Kottner (2002) says that “the process of anaerobic

digestion is at its optimum at a temperature range of 25 o – 38oC (mesophilic

conditions)”. Adding, Tchobanoglous et al (2003) say “Reactor temperatures

of 25 to 35oC are generally preferred to support more optimal biological

reaction rates and to provide more stable treatment”. Just in this same vein,

Mattocks (1984) had explained that, though operating temperature is

critical, stabilizing the temperature and keeping it stabilized are even more

important”. He explains that “variations of plus or minus 1[degree] C in a

day may force the methane producing organisms into periods of dormancy”.

But in contrast, according to Viessman and Hammer (1993), “The rate of

biological activity in the mesophilic range between 5oC and 35oC doubles for

every 10oC– 15o C temperature rise. This is possibly why Kiely (1998) and

Viessman and Hammer (1993) agree that thermophilic digestion has not

been successful in practice because thermophilic bacteria are very sensitive

to small temperature changes; and Eckenfelder (2000) adds that “the


128

maintenance of higher temperatures is usually not economically justifiable”,

apparently corroborating the view of Mattocks (1984), who after noting that

at an elevated temperature minor changes in system conditions could offset

digester efficiency or productivity, stressed that an additional source of

energy will likely be required to maintain the digester contents at a constant

higher temperature.

At this point, it is important to also state that psychrophilic

temperature ranges (say 20o C) are not suitable for anaerobic digestion, as

according to Tchobanoglous et al (2003), the degradation of long chain fatty

acids is often rate limiting. If long chain fatty acids accumulate, foaming may

occur in the reactor and inhibit process continuity.

Fortunately, the average ambient temperature in the area of study,

even at the time of this study, is 32oC; leading to an adoption of the

optimum mesophilic temperature of 35o C for this design work.

4.4.3 pH Control

The level and variation of pH in the digesting material, also affects

the anaerobic digestion process. Viessman and Hammer (1993) were very

emphatic when they said that the hydrogen ion concentration of the culture

medium has a direct influence on microbial growth. The “digestion is

inhibited by excessive acidity” Steadman (1975). He explains that “the

bacteria involved in anaerobic digestion have a pH range of 6 to 8 with

values close to 7 for optimum activity”. In the initial phase of the process,

the production of volatile fatty acids depress the pH, but the reaction of CO 2,
129

which is soluble in water, with hydroxide ions for bicarbonate ions, HCO 3-,

tend to restore the neutrality of the process pH, thus making the process

self-stabilizing or ‘well buffered’. Eckenfelder (2000) explains that when the

rate of acid formation exceeds the rate of breakdown to methane, a process

imbalance results in which the pH decreases, gas production falls off, and

the CO2 content of the gas increases. The overall effect, according to

Mattocks (1984), is “impeding the effectiveness of the whole biogas

system”. Sufficient alkalinity has to be available at all times, up to a level of

approximately 3000mg/l, for sufficient buffering to be maintained, to ensure

a high rate of methane production. Eckenfelder (2000) says lime is

commonly used to raise the pH of an anaerobic system when there is a

process imbalance. He, however, cautions that care must be taken not to

apply excess lime, since this will result in precipitation of calcium carbonate;

and says that sodium bicarbonate can alternately be used for pH

adjustment.

4.4.4 Carbon-Nitrogen Ratio

Carbon and Nitrogen are a set of another very important elements

determining the performance of the anaerobic digestion process, as one or

the other usually constitutes a limiting factor. Whereas carbon constitutes

the energy source for the microorganisms, nitrogen serves to enhance

microbial growth. If nitrogen is limiting, microbial populations will remain

small and it will take longer to decompose the available carbon. Excess

nitrogen, beyond the microbial requirements, is often lost from the process
130

as ammonia gas, (Richard, 1998b). It has variously been found that the

bacteria in the digestion process use up the carbon present 30 to 35 times

faster than the rate at which they convert nitrogen. So for optimum

operation, the ratio of the carbon to nitrogen should be about 30:1 in the raw

material. Richard (1998) says that usually, nitrogen is the limiting element

in the processing of MSW, and additives such as manure, clean sewage

sludge (biosolids), septage and urea can be used as a supplemental

nitrogen source.

4.4.5 Moisture Content

The moisture content of the waste in the digester is highly essential

for the activities of the waste decomposing anaerobes. Sufficient amount of

moisture is required for effective anaerobic digestion.

4.4.6 Waste Particle Size

The particle size of MSW is quite relevant for the biological

transformation of the waste, and also for equipment sizing for further

treatment. The relatively large particles sizes of MSW, which Kiely (1998)

puts at an average of 100mm, considerably slows down decomposition

activities, hence the necessity of the reduction of the particle size of the

MSW. Richard (1998b) says reducing particle size increases surface area,

which is desirable for the process operation, as optimum conditions for

decomposition occur on the surfaces of organic materials. The particle size

reduction of MSW could be achieved through shredding or grinding, and


131

Agunwamba (2001) says, “the ultimate objective is to improve the efficiency

of the operation”, by presenting a larger surface area of the material for

microbial activity.

4.4.7 Mixing

Mixing is another very important operation in anaerobic digestion. It

is desirable to maintain uniformity of substrate concentration, temperature

and other environmental factors; prevent scum formation and solids

deposition (Agunwamba, 2001). In fact, Tchobanoglous and Burton (1991)

say proper mixing is one of the most important considerations in achieving

optimum process performance. Despite the relevance of mixing in the

anaerobic digestion process, there is a consensus of opinion that the mixing

of the substrate is energy intensive. Of course, this energy intensity of

mixing is even in the consideration that all the substrates are sludge. For a

solid matter like the MSW, and therefore a ‘dry solid’ digestion process,

mixing becomes rather difficult and indeed economically tedious. It is

important to state that even the mixing method of gas re-circulation may

also entail loss of valuable gas for use. However, the mixing methods of

mechanical stirrers and gas re-circulation respectively will be considered for

the high and low-solids digesters, depending on the eventual total solids

(TS) concentration of the systems.


132

4.4.8 Digester Heat Requirements

There is a consensus of opinion by different authors (Tchobanoglous

et al, 2003; Kiely, 1998; Reynolds and Richards, 1996, etc) that the heat

requirements of digesters are comprised of the quantity:

(i) to raise the incoming sludge to digestion tank temperatures,

(ii) to compensate for the heat losses through the walls, floor, and

roof of the digester, and

(iii) to make up the losses that might occur in the piping between

the source of heat and the tank.

It is important to state that in most of these research and operational

areas where digesters require external heating, ambient temperatures fall

within the psychrophilic range ( 20oC), and since “optimum temperature for

maximum rates of reaction is around 35oC (mesophilic), Agunwamba, 2001,

they definitely would require external heating to get digester temperature up

to the mesophilic range. This is why Reynolds and Richards (1996) insist

that digester heating is necessary in moderate and cold climates, especially

during winter, to maintain the digester temperature within the desired range.

This immediately implies that external heating of the digester is not very

necessary. Therefore, since the anaerobic digestion process is an

exothermic one leading to variation in the internal and external

temperatures of the digester, the applicable heat considerations for the

digester are the microbial heat generation in the digester, and the heat flow

across the digester boundaries


133

4.4.9 Cost of the Digester

The cost of an anaerobic digester for processing of MSW for energy

will actually depend on the cost of obtaining the feedstock, preparing the

MSW for digestion, and constructing and maintaining the plant. In fact

Tchobanoglous and Burton (1991) say, “Process costs (both capital and

operating, and maintenance) is extremely important in selecting the type

and size of reactor”. In this regard the bio-kinetic and design models for the

reactor will directly affect the digester cost, particularly in terms of digester

and feedstock volumes to yield desired quantity of gas. The cost effective

consideration here is for a batch digestion. Steadman (1975) states that “the

simplest type of methane digester is just a closed container such as a drum,

tank or pit in the ground into which the digestible material is loaded”, - i.e. a

batch digester. The simplicity of the batch digester design clearly should

also extend to the process cost. The Oregon State Department of Energy

(2002), after describing three types of digesters, namely a covered lagoon

digester (i.e. a batch digester), a complete mix digester, and a plug flow

digester, states that the batch digester “is the least expensive of the three”.

4.4.10 Loading Rate

The loading rate of organic materials into the digester greatly affects

anaerobic digester design, particularly the volume of the digester, and

indeed the overall process performance. Mattocks (1984) explains that

“loading rate is an important parameter since it tells us the amount of

volatile solids to be fed into the digester each day”. Volatile solids being the
134

portion of organic material solids that can be digested, while the remainder

of the solids is fixed. The fixed solids and a portion of the volatile solids are

non-biodegradable. Mattocks further says that the actual loading rate

depends on the types of wastes fed into the digester. This is indeed easy to

conjecture since the type of waste used for the process determines the level

of biochemical activity in the digester.

Tchobanoglous et al (2003) state that wide variations in influent flow

and organic loads can upset the balance between acid fermentation and

methanogenesis in anaerobic processes. They explain that for soluble

easily degradable substrates, such as sugars and soluble starches, the

acidogenic reactions can be much faster at high loadings and may increase

the reactor volatile fatty acids (VFA) and hydrogen concentrations and

depress the pH.

It is suggested that loading rate also affects the food to microbes

(F/M) ratio. According to Kiely (1998), a system achieves equilibrium when

the food substrates and the microorganisms consuming it are in balance.

When out of balabce, he explains that there could be too much substrate or

too little substrate, or too many organisms or too little organisms. Any of

these situations would destabilize the process; and the equilibrium

parameter is the F/M ratio.

4.4.11 Pretreatment of the Waste

Pretreatment of waste is actually the preparation of the waste for the

anaerobic digestion process. For instance, it has been severally presented


135

that MSW, particularly in this part of the world, is a complex mixture of both

organic and inorganic materials. So, since the digestion process is for the

decomposition of the organic components, the necessary implication is that

the organic components would have to be sorted from the inorganic ones.

Again, and importantly, the particle sizes of the organic waste are such that

will require reduction. These and many other processes the waste may

undergo before being fed into the digester are all classified as waste

pretreatment.

4.5 DESIGN COMPUTATIONS

4.5.1 General Design Parameters

Basis of Design 0.415 m3 CH4/kg of MSW

Daily organic waste load 1,043,038 kg

Total Solids (TS) 80.1%

Moisture Content 19.9%

Volatile Solids (VS) 66.6% TS

Fixed Solids (FS) 13.5% TS


As-discarded density (d) of waste 171.4 kg/m3
As-compacted density (c) of waste 222.2 kg/m3
Compaction ratio (r) 1.30

As-discarded volume (Vd) of waste 8781.25 m3

As-compacted volume (Vc) of waste 6754.81 m3

Operational temperature of digester 35 oC

Ambient temperature 30 oC
136

Specific gravity of dry solids 1.251

Specific gravity of wet solids (batch) 1.057

Specific gravity of wet solids (CSTR) 1.02

Specific heat capacity 4200J/s /oC

Fuel value of methane gas 26 MJ/m3

Computational Total Solids concentration (batch) 26.7%

Computational Total solids concentration (CSTR) 10.0%

Fractional conversion 0.8

4.5.2 Batch Digester Design Computations

4.5.2.1 Calculation of Volume of Batch Digester

From equation (3.18),

Vavg = 1722.26 - 2/3 (1722.26 - 103.82)

= 1722.26 - 1078.99

= 643.30 m3

 Vs = 643.30 (9.075) + 103.82 (20)

= 5841.16 + 2076.40

= 7917.56 m3

Vbd = 3/2 x 7917.56

= 11876.34.99 m3

Therefore, diameter of batch digester will be

8  11876.34
Db  3

= 31.15 m
137

and height of batch digester equals

31.15
Hb 
2

= 15.58 m

So that area of batch digester becomes

  31.152
Ab 
4

= 762.19 m2

4.5.2.2 Calculation of Amount of Biogas Produced

4.5.2.2.1 Concentration of Dissolved Methane Gas

For the batch digester, the concentration of dissolved methane gas

would be

Cm = 0.296 x 1.43 (50.18 – 17.79)

= 13.71 kg/m3

Assuming a partial pressure of the gas CH4 as 11mmHg, then:

11mmHg
Cst 
30769.23mmHg / mol / l

= 0.0003575 mol/l

= 0.00572 kg/m3

762.19
 C g  0.0984  (0.00572  13.71)  9.08
7917.56

= 1.179 kg/m3
138

4.5.2.2.2 Mass of Methane Gas

Mass of methane gas produced in the batch digester would be:

Mm = 0.1.179 x 3958.78

= 4667.40 kg

4.5.2.2.3 Volume of Methane Gas

Volume of methane produced in the batch digester

4667.40
Vm 
0.644

= 7247.52 m3

4.5.2.2.4 Volume of Biogas

Total volume of biogas produced is

100
Vt   7247.52
60

= 12079.20 m3

4.5.2.2.5 Volume of methane produced per unit volume of digester

7247.52
Vmb 
11876.34

= 0.610 m3 CH4/m3 reactor/day

4.5.2.2.6 Volume of methane produced per volatile solids added

7247.52
Vmv 
86431.40

= 0.0839 m3 CH4/kg VS added

4.5.2.2.7 Volume of methane produced per unit of biodegradable

VS added
139

7247.52
Vms 
(1  0.193)  86431.40

= 0.1039 m3 CH4/kg BVS added

4.5.2.3 Heat Requirement

The fluid film coefficients suggested by Coulson and Richardson

(1991) were used for the computation of the overall heat transfer coefficient

for the system. The wall heat transfer coefficient was given as 4.9 W/m 2 oC;

roof heat transfer as 4.7 W/m2 oC; and floor heat transfer coefficient as 2.85

W/m2 oC.

4.5.2.3.1 Energy Content of MSW

Using equation (4.5), and the data from field investigation, the energy

content of the substrate was estimated thus:

Ec = 0.051[29.2 + 3.6(12.4)] + 0.352 (9.9)

= 7.251 MJ/kg

= 1.733 kcal/g

= Hs

As Hc = 40% of Hs.

 Hc = 0.4 x 1.733

= 0.693 kcal/g

0.126
Also Ys 
0.126  0.0475
0.3293

= 0.293 g cell/g MSW


140

0.293
and Y 
1.733  0.293(0.693)

= 0.192 g/kcal

4.5.2.3.2 Microbial Heat Generation

Using the above parameters, the microbial heat generation is calculated

thus:

1
Qgrb  7917.56  0.126  10.69 
0.192

= 55,544.16 kcal/day

= 232.40 MJ/day

= 9.68 MJ/hr

4.5.2.3.3 Heat required to raise temperature of incoming sludge

100 1
Qsb  129,776.87  (35  30)   4200
26.7 24

= 425,298,731.30 J/hr

= 425.30 MJ/hr

4.5.2.3.4 Heat Losses

4.5.2.3.4.1 Heat loss through the wall

Qwb  4.9  1524.87(35  30)

= 37359.32 W

= 134493552 J/hr

= 134.49 MJ/hr
141

4.5.2.3.4.2 Heat loss through the floor

Q fb  2.85  762.68(35  20)

= 32,604.57 W

= 117,376,452.00 J/hr

= 117.38 MJ/hr

4.5.2.3.4.3 Heat loss through the roof

Qrb  4.7  762.68(35  30)

= 17,922.98 W

= 64,522,728.00 J/hr

= 64.52 MJ/hr

Therefore heat required to compensate for heat losses equals

Qcb = Qw + Qf + Qr

= 134.49 + 117.28 + 64.52

= 316.39 MJ/hr

4.5.2.3.5 Total external heat required

Qeb = Qsb + Qcb

= 425.30 + 316.39

= 741.69 MJ/hr

4.5.2.3.6 Net external heat required

Qnb = Qeb - Qgrb

= 741.69 – 9.69

= 732.00 MJ/hr
142

4.5.2.3.7 Total specific heat required

732.00
qtb 
11876.34

= 0.0616 MJ/hr/m3

61635
 M wb 
4200(35  30)

= 2.94 kg/m3/hr

= 70.4 kg/m3/day

4.5.2.3.8 Heat Exchanger Surface Area

Required surface area for heat exchange is

732000000
Ahb 
17640(35  30)

= 8299.32 m2

and the length of heating pipe for a given radius of pipe (say 0.3 m) is

8299.32
lb 
2    0.3

= 4402.36 m

4.5.2.3.9 Capacity of Mixing Screw

From equation 3.27

 (15.582  4.672 )
QAUG  60x x13.63x60x0.241x0.325
4

= 332935.91 tons
143

4.5.2.3.10 Cost of Batch Digester

0.6
 11876.34 
X db  56,311,875 
 2    0.2303

= N 600,306,243.00

4.5.3 CSTR Design Computations

4.5.3.1 Time for CSTR Digestion

From equation 3.30,

(6.79  2.41)(.072  2.41)


h 
(0.382)(2.41)(1.476)

= 8.00 days

4.5.3.2 Calculation of Volume of CSTR

Vc = 24.95 x 8.00

= 199.6 m3

Therefore, diameter of CSTR will be

8  199.6
Dc  3

= 7.98 m

and height of CSTR equals

7.98
Hc 
2

= 3.99 m

So that area of CSTR becomes


144

  7.982
Ac 
4

= 50.02 m2

4.5.3.3 Calculation of Amount of Biogas Produced

4.5.3.3.1 Concentration of Dissolved Methane Gas

Here the concentration of dissolved methane gas would be

Cm = 0.296 x 1.43 (6.79 - 2.41)

= 1.854 kg/m3

Assuming a partial pressure of the gas CH4 as 11mmHg, then:

11mmHg
Cst 
30769.23mmHg / mol / l

= 0.0003575 mol/l

= 0.00572 kg/m3

50.02
 C g  0.0984  (0.00572  1.854)  8.00
199.6

= 0.365 kg/m3

4.5.3.3.2 Mass of Methane Gas

Mass of methane gas produced in the batch digester would be:

Mm = 0.365 x 199.6

= 72.85 kg

4.5.3.3.3 Volume of Methane Gas

Volume of methane produced in the batch digester


145

72.85
Vm 
0.644

= 113.12 m3

4.5.3.3.4 Volume of Biogas

Total volume of biogas produced is

100
Vt   113.12
60

= 188.53 m3

4.5.3.3.5 Volume of methane produced per unit volume of digester

113.12
Vmc 
199.6

= 0.567 m3 CH4/m3 reactor/day

4.5.3.3.6 Volume of methane produced per volatile solids added

113.12
Vmvc 
20770.43

= 0.0545 m3 CH4/kg VS added

4.5.3.3.7 Volume of methane produced per unit of biodegradable

VS added

113.12
Vmsc 
(1  0.193)  20770.43

= 0.00675 m3 CH4/kg BVS added

4.5.3.4 Heat Requirement

The fluid film coefficients suggested by Coulson and Richardson

(1991) were used for the computation of the overall heat transfer coefficient

for the system. The wall heat transfer coefficient was given as 4.9 W/m 2 oC;
146

roof heat transfer as 4.7 W/m2 oC; and floor heat transfer coefficient as 2.85

W/m2 oC.

4.5.3.4.1 Energy Content of MSW

Using equation (4.5), and the data from field investigation, the energy

content of the substrate was estimated thus:

Ec = 0.051[29.2 + 3.6(12.4)] + 0.352 (9.9)

= 7.251 MJ/kg

= 1.733 kcal/g

= Hs

As Hc = 40% of Hs.

 Hc = 0.4 x 1.733

= 0.693 kcal/g

0.126
Also Ys 
0.126  0.0475
0.3293

= 0.293 g cell/g MSW

0.293
and Y 
1.733  0.293(0.693)

= 0.192 g/kcal
147

4.5.3.4.2 Microbial Heat Generation

Using the above parameters, the microbial heat generation is calculated

thus:

Qgrc  (6.79  2.41)(0.293)(24.95)  (1.476)(0.0475)

= 31.95 kcal/day

= 0.134 MJ/day

= 0.00558 MJ/hr

4.5.3.4.3 Heat required to raise temperature of incoming sludge

100 1
Qsc  31186.84  (35  30)   4200
10 24

= 272,884,850.00 J/hr

= 272.88 MJ/hr

4.5.3.4.4 Heat Losses

4.5.3.4.4.1 Heat loss through the wall

Qwc  4.9  100.04(35  30)

= 2450.98 W

= 8,823,528.00 J/hr

= 8.82 MJ/hr

4.5.3.4.4.2 Heat loss through the floor

Q fc  2.85  50.02(35  20)

= 2138.36 W

= 7,698,096.00 J/hr

= 7.70 MJ/hr
148

4.5.3.4.4.3 Heat loss through the roof

Qrc  4.7  50.02(35  30)

= 1175.47 W

= 4,231,692.00 J/hr

= 4.23 MJ/hr

Therefore heat required to compensate for heat losses equals

Qcc = Qw + Qf + Qr

= 8.82 + 7.70 + 4.23

= 20.75 MJ/hr

4.5.3.4.5 Total external heat required

Qec = Qsc + Qcc

= 272.88 + 20.75

= 293.63 MJ/hr

4.5.3.4.6 Net external heat required

Qnc = Qec - Qgrc

= 293.63 – 0.00558

= 293.62 MJ/hr

4.5.3.4.7 Total specific heat required

293.62
qtc 
199.6

= 1.471 MJ/hr/m3

1471000
 M wc 
4200 (35  30)
149

= 70.05 kg/m3/hr

= 1681.20 kg/m3/day

4.5.3.4.8 Heat Exchanger Surface Area

Required surface area for heat exchange is

293620000
Ahc 
17640(35  30)

= 3329.02 m2

and the length of heating pipe for a given radius of pipe (say 0.3 m) is

3329.02
lc 
2    0. 3

= 1765.87 m

4.5.3.4.9 Gas Diffusion Rate for CSTR Mixing

From equation 3.54

Q = 11.5 x 0.183 x 7.98

= 0.535 m3/s

4.5.3.4.10 Cost of CSTR

0.6
 199.6 
X dc  56311875 
 230 

= N 51,720,130.11
150

CHAPTER 5

RESULTS AND DISCUSSION

5.1 RESULTS

5.1.1 FIELD RESULTS

Table 5.1: Solid Waste Generation in Port Harcourt Metropolis *

Zone Vicinity Daily Contractor for


Waste Evacuation
Load (kg)
1. Old Port Harcourt Lomo Enterprises
Township up to UTC Junction 250,215 Limited

2. Diobu, Mile 1 – 4
up to Wimpey Junction 390,000 Refcol Nigeria Limited

3. Port Harcourt Expression /


GRAs 1-3 up to Rumukuta 275,840 Akor and Fidlis Limited
Junction

4. Elekahia Village / Rumuomasi


upto Woji / Elelenwo Villages 183,200 Palisco Nigeria Limited

5. Ogbunabali / Nkpogu Villages


upto to Abuloma / Oginiba 130,721 P. T. K. Limited
Villages

6. Rumuokoro / Igwuruta Villages 152,630


upto to Eleme / Oyibo towns Manfil Nigeria

7. Waterlines / Presidential Estate Industrial Support


upto to Artillery / Mgbuoba 122,500 Services Limited
Village
Total 1,505,106

*Data obtained in May 2002


151

Table 5.2: Physical composition of MSW in Port Harcourt metropolis

Waste Component Mass Mass Volume: Density: Volume: Density: as-


(kg) (%) as- as- as- compacted,
discarded discarded compacted, (kg/m3)
(m3) (kg/m3) (m3)
ORGANIC

- Food Waste 43.5 29.2 0.14 310.7 0.12 362.5

- Wood/Leaves 12.7 8.4 0.05 254.0 0.04 317.5

- Paper 18.5 12.4 0.21 88.1 0.17 108.8

- Plastics 14.7 9.9 0.24 61.3 0.23 63.9

- Textiles/Rubber/ 11.4 7.6 0.08 142.5 0.06 190.0

Leather

- Miscellaneous 2.7 1.8 0.02 135.0 0.01 270.0

Organics

INORGANIC

- Glass 20.1 13.5 0.11 182.7 0.08 251.3

- Metals 25.6 17.2 0.13 196.9 0.12 213.3

TOTAL 149.2 100 0.98 171.4 (av) 0.83 222.2 (av)

5.1.2 LABORATORY RESULTS

Table 5.3: Proximate Analysis of Organic MSW in Port Harcourt


Metropolis

Waste Proximate Analysis (% by weight)


Component
Moisture Volatile Matter Fixed Carbon Ash
Food Waste (mixed) 65.2 26 4.0 4.8
Wood / Leaves 19.2 65 15 0.8
Paper 6.9 9.1 6.0
78
Plastics 0.3 95 2.4 2.3
Textiles / Rubber / Leather 7.8 69 16.2 7.0
Average Values 19.9 66.6 9.3 4.2
152

Table 5.4: Ultimate Analysis of Organic MSW in Port Harcourt

Waste Ultimate Analysis (% by weight-dry basis)


Component
Carbon Nitrogen Hydrogen Oxygen Sulphur Ash

Food Waste 51 2.3 5.0 39 0.3 2.4

Wood / Leaves 49 1.2 6.0 42 0.1 1.7

Paper 45 0.6 5.0 45 0.1 4.3

Plastic 56 - 6.0 26 - 12

Textiles / rubber 59 5.4 6.0 19 0.2 10.4


leather

Average Values 52 1.9 5.6 34.2 0.14 6.16

Table 5.5: Experimental Batch Digesters' Data

Digester Duration Initial MSW Effluent MSW Initial Microbial Effluent Microbial
no. of Concentration, Concentration, Concentration, Concentration, Xe
Digestion So (mg/l) Se (mg/l) Xo (mg/l) (mg/l)
, t (days)
- 0 462.12 - 32.05 -

1 5 462.12 328.77 32.05 114.31

2 10 462.12 78.71 32.05 206.45

3 15 462.12 26.49 32.05 137.45

4 20 462.12 13.19 32.05 127.89

5 25 462.12 5.43 32.05 12.87


153

Table 5.6: Reduced Data from Batch Experimentation for the


determination of Process Kinetic Parameters

Digester Duration of _ _
no. Digestion, t X Xt -In (Se/So) So - Se So - Se
/Xt In (So/Se)
/Xt
(days)
- 0 - 0 0 - - -

1 5 73.18 365.9 0.340 133.35 0.364 0.000929

2 10 119.25 1192.50 1.770 383.41 0.322 0.00148

3 15 84.75 1271.25 2.859 435.63 0.343 0.00225

4 20 79.97 1599.4 3.556 448.93 0.281 0.00222

5 25 22.46 561.5 4.444 456.69 0.813 0.00791

_
X = average cell mass concentration during the biochemical reaction-that is X = ½(Xo + Xt),
where Xo and Xt are the cell mass concentrations at the respective times t = 0 and t = t
(Reynolds and Richard, 1996)

Table 5.7: Reduced data from batch experimentation


for determination of CSTR parameters

Rate of Specific Rate of Mean cell


MSW rate of 1 growth of residenc 1 1
/U / /Se
utilization, MSW microbes, e time, ,
dS utilization, dX days
/dt, /dt,
mg/l/day U, day-1 mg/l/day
0 0 0 0 - - -

41.72 0.570 1.754 16.45 4.444 0.225 0.00304

25.58 0.215 4.662 17.44 6.849 0.146 0.0127

6.25 0.074 13.56 7.03 12.06 0.0829 0.0378

4.66 0.058 17.16 4.79 16.67 0.060 0.0758

0.56 0.025 40.11 (0.767) 29.41 0.034 0.1842


154

5.2 DISCUSSION

5.2.1 DISCUSSION OF FIELD RESULTS

The field investigation involved the determination of the rate at which

solid waste is generated in the Port Harcourt metropolis, from which the per

capita waste generation is ascertained, when the generation rate is related

to the population of the municipality. Also during the field investigation the

physical composition was determined. This involved the sorting of the

otherwise composite solid waste into its different components, based on

their percent composition by weight of the entire mixed waste.

5.2.1.1 Rate of Solid Waste Generation

The rate at which the MSW is generated was actually determined in

the first instance in terms of volume, using equation (4.1) and converted to

kilograms using the average density value obtained in Table 5.2.

Table 5.1 shows the rate of solid waste generation in Port Harcourt

metropolis. It gives an indication of the quantity of solid waste generated

daily, which is 1,505,106kg (one million five hundred and five thousand one

hundred and six kilograms). The rate of generation of solid waste is indeed

a very important parameter in the design of a solid waste management

system. When this amount of solid waste generated in the metropolis is

spread across the estimated population of Port Harcourt metropolis of about

1,356,000 (One million, three hundred and fifty six thousand) as per official

records of the Federal Office of Statistics (2003), then the kilogram per

capita per day generation of waste in the metropolis becomes 1.11. This
155

figure is useful in a solid waste management design considering an

increase or otherwise in population.

5.2.1.2 Physical Composition of MSW

It is seen from Table 5.2 that the percentage by weight of the organic

components of the MSW is 69.3. The various individual items of the organic

component have also been classified into their weight percentages. These

classifications are quite significant in reactor sizing – determination of

digester volume and indeed the volumetric organic loading rate. They show,

for instance, that in relation to Table 5.1, the design for the digester volume

and loading rate must consider a daily organic waste load of 69.3% of

1,505,106kg, which is 1,043,038kg. This is corroborative of the perspective

of Mattocks (1984) when he said, “it is important to remember that a

digester must be designed on the basis of the amount of waste that can be

collected and actually fed to the digester, not on the quantity of waste

produced”

But the volume and density of a given mass of MSW vary in their

“as–discarded” and “as-compacted” states, and these have also been

shown in Table 5.2. With the average density values and using equation

(4.2) the compaction ratio (rc) was calculated as:

222.2
 1.30
171.4

The necessary implication of the compaction ratio is the

determination of the actual digester volume, which is more closely related to


156

the “as-compacted” state of the waste. This is because as the MSW is

shredded for processing, they tend to become more compacted as their

particle sizes reduce.

So, with a daily “as-discarded” waste volume of 8781.25 m 3 then the

compacted volume will be

Vd
Vc =
rc

8781.25
and Vc = = 6754.81m3
1.30

Where Vc – “as-compacted” volume

Vd – “as–discarded” volume

5.2.2 DISCUSSION OF LABORATORY RESULTS

The laboratory results are those of the chemical characterization of

the waste, in terms of the determination of some surrogate parameters in

place of the true chemical content, and also some relevant chemical

elements that compose the waste. Laboratory results also include data

from reactor experimentation.

5.2.2.1 Proximate and Ultimate

The proximate analysis is relevant in the determination of the portion

of the organic matter that is actually useful for the process. For instance, it

indicates the portion that is volatile, i.e. vaporizable, and from which it is
157

possible to determine the biodegradable component. The Table 5.3 shows

that the volatile matter of the MSW is 66.6% by weight of the organic waste.

Therefore, the volatile matter fraction of the organic component of

municipal solid waste is:

VM = 0.666 (0.693 x 1505,106)

= 694, 663.6kg (i.e. by weight)

or VM = 0.666 (0.693 x 8781.21)

= 4052.88m3 (by volume)

where VM - volatile matter.

It is important to state that the volatile solids, as a major component

of the organic waste are very useful in the determination of digester volume.

It is in fact the substrate for the process microorganisms. So the

concentration of the volatile solids at different times of the process is clearly

an indication of the rate at which the process progresses and helps in the

determination of the solid retention time.

From Table 5.4, the average carbon content is 52%, and that of

nitrogen is 1.9%. These values give a carbon – nitrogen ratio of 27:1, which

is conducive for the activities of microorganisms, as it falls into the desired

range for efficient process operation. It is an indication that the waste has

the desired amount of nitrogen to support a microbial population adequate

for active decomposition. Mattocks (1984) says that “to act efficiently on the

substrates, microorganisms need a 20-30:1 ratio of carbon to nitrogen, with

the largest percentage of the carbon being readily degradable”.


158

5.2.2.2 Reactor Experimentation

The substrate and microbial concentrations respectively with time

were the fundamental subjects of this experimentation from where other

kinetic parameters were determined. The Table 5.5 shows the result of the

reactor experimentation.

5.2.2.3 Interpretation of Batch Experiment Data

The interpretation of the batch experiment data presented in Table

5.5 is done with Figures AI-1 and AI-2 in Appendix I

Figures AI-1 and AI-2 are curves, respectively for the degradation of

MSW and growth of biomass with time of digestion. It is seen that the

highest utilization of MSW occurred between days 5 and 10, before the

degradation became more and more sluggish up to day 25 when there was

a sufficient reduction in MSW concentration to constancy level. Although it

was expected that MSW utilization would have been highest between 0 and

5 days given its abundance as the limiting component, but this was not the

case; a situation that is easily attributable also to the sluggishness in the

growth of microbes during the same period. This is noteworthy because the

system was indeed self-generating in biomass, as there was no initial

inoculation, as seed to prime the system. This was however deliberate to

study the self-generating capacity of the waste. This comparison is better

appreciated from Figure AI-3, which is an imposition of Figures AI-1 and AI-

2.
159

However, from Figure AI-2 it is clear that microbial growth was at its

peak when a sustained concentration of the waste was available to provide

sufficient ingredients for the sustenance of the growth of microorganisms.

Evidently, this growth was continuing even up to the 25th day, although a

noticeable decline in growth rate started from the 15th day, which again

corresponds, to the period when the rate of decomposition of the growth

limiting waste has become fairly constant. It is possible to assert that at this

point the available biodegradable volatile solids may have been depleted,

leading to insufficiency of nutrients for microbial growth. Thus, it can be

deduced that indeed the rate of decomposition of the MSW is a function

mainly of the concentration of the waste, which itself generates the

degrading microorganisms. Hence the process follows first order kinetics.

5.2.2.4 Verification of Applicable Rate Equation for the Process

It is now trite knowledge that in the kinetics of MSW degradation, rate

expressions are required for the description of the activities of the various

components of the system. These expressions are classified into different

reaction orders; and the applicability of any of the rate expressions is a

function of the relationship between the limiting substrate and the digestion

time.

To validate the earlier description of the process as a first order

reaction, the integral method of analysis as proposed by Levenspiel (1999)

was employed. This method of analysis requires that a plot of the negative
160

of the natural logarithm of the ratio of effluent MSW to influent MSW, - ln

Se So against time, t, should yield a straight line, as in Figure AI-4.
 
The Figure AI-4 is a plot of - ln Se So against t, and it is evident that

the best fitting trend line of the resultant curve is a straight line, which

tended to pass through the origin. This, therefore, satisfies the condition for

the use of the first order rate equation.

5.2.2.5 Determination of Kinetic Parameters

The digestion of the MSW by microorganisms is a time dependent

process and characterized by certain rate parameters like the growth yield

of microbes (Y), the maximum rate of utilization of the waste (k), the

saturation constant (Ks), the endogenous decay coefficient (kd) and the

maximum growth rate of biomass (max).

With the confirmation of the use of the first order rate equation for the

process analysis, these kinetic parameters were determined from the

experimental results by a linearization of the appropriate equations.

From Figure AI-4, which was obtained by linearizing equation (2.5)

The slope of the curve represents the maximum rate of utilization of MSW

(k). The equation describing the curve was obtained from a statistical

analysis of the curve using the Microsoft chart editor, as;

 Se 
ln   = 0.1883 t – 0.1926
 So 

From where
161

k = 0.1883 day-1

This value of k is strictly for the batch processing, and suggests that

a relatively low amount of the organic matter is consumed in relation to the

growth of microbes. This tends to corroborate the earlier assertion that the

system was indeed self-generating in microbial content; hence only little

amount of the substrate was used in the course of the digestion. It is also

believed here that the composite nature of the MSW, incorporating some

non-biodegradable components, would have contributed in part to the low

level of degradation of the waste. Also of significance in this consideration is

the assessment of the concentration of the MSW on a COD basis, which

implies the inclusion of the non-biodegradable components in the

concentration assessment. Therefore, in this regard, a careful sorting of the

waste into biodegradable and non-biodegradable components is likely to

result in a higher ‘k’ value.

With the method of Bailey and Ollis (1981) in the use of batch

experiment data for the design of a CSTR, which result is presented in

Table 5.7, and the procedure for the determination of kinetic parameters

outlined by Viessman and Hammer (1999), the curves in Figures AI-5 to AI-

7 are obtained.

I 1
Figure AI-5 is a plot of against obtained by the linearization of
U Se

the steady state equation, and from the solution of the Microsoft chart

editor,
162

1 204.6
  2.6189
U Se

1 Ks
from where, intercept of curve = and slope =
k k

Therefore,

1
= 2.6189,
k

or k = 0.382 day-1

Ks
and = 204.6
k

Ks = 204.6 x k

= 78.16 mg/l =0.0782 kg/m3

These values of k and Ks particularly in relation to the influent

substrate (MSW) and microbial concentrations, suggest that the digesting

microbes require a much longer time to regenerate, and indeed sluggish in

their performance. This appears understandable because again the

experimental investigation did not incorporate an inoculation of the digester

feed with seed biomass to prime the process but rather depended on a self-

generation of biomass and subsequent regeneration. Little wonder then that

Figure AI-2 displayed a sluggish growth of microbes between 0 and 5 days

before an eventual accelerated growth up to day 10. This observation would

further suggest that actual or practical digesters for MSW would require an

inoculation of the feed stream with microbes, or perhaps an operating

system with cell recycle.


163

From Figure AI-7, the growth yield of microbes (Y) is obtained as the

slope of the curve and the intercept as the lyses coefficient (kd)

0.368
Slope = Y 
0.4708

 Y = 0.3293

This value of the growth yield coefficient gives the impression that more

cells are formed per unit of the MSW degraded. This is indeed traceable to

the use of the mixed liquor volatile suspended solids (MLVSS) as an index

for the assessment of the mass of biomass.

The specific rate of growth of microorganisms is also determined

using equation (3.8) as

max = 0.3293 x 0.382

= 0.126 day-1

5.2.2.6 Interpretation of Experimental Data

The analysis of the field experimental data shows that an organic

load of 1,043,038.00 kg/day expectedly generated by the entire populace of

Port Harcourt metropolis contained a percentage Total Solids (TS) of 80.1,

out of which 66.6% is volatile. The biodegradable fraction of the Total

Volatile Solids (TVS) is 80.7%, which shows that 53.7% of the TS is

biodegradable. So with a Carbon to Nitrogen (C: N) ratio of 27: 1, it is

evident that the MSW under consideration is highly biodegradable, and so

can be subjected to digestion by microorganism in a waste-to-energy

conversion process.
164

Considering the requirements of treating the waste, namely the limits

of the percentage Total Solids in both batch (30%) and CSTR (10%)

processes, and considering also the large volume of the waste generated

and collected, the percentage TS was fractionally reduced with a factor of

0.33 and 0.125 to obtain the manageable percentage TS of 26.7 and 10 for

the batch and CSTR processes respectively.

For the batch processing with TS of 26.7%, an initial waste

concentration of 50.16 kg/m3 was reduced to an effluent concentration of

17.79 kg/m3; and that of the CSTR from 6.79 kg/m 3 to 2.41kgm3, at an

experimentally determined maximum utilization rate (k) of 0.382 day -1 and

yield coefficient (Y) of 0.3293. This marginal reduction in the substrate

concentration, giving rise to a percentage stabilization of 64.54 is clearly the

result of the low yield coefficient, and also the low maximum rate of

utilization of the substrate solid, which could be improved upon by adequate

seeding of the process at commencement.

Also with the daily waste load of 26.7% and the sludge volume of

26.8% for the batch, and 10% and 26.8% respectively for the CSTR, the

volume of the digesters and their corresponding heights and diameters were

determined. The volumes of methane and biogas, and other operation

parameters were also determined.

5.2.2.7 Interpretation of Simulation Results

In an attempt to study the digester operation behaviour and enhance

its performance, a programme was developed for the process using the
165

Microsoft Visual Basic Version 6.0 programming language, with which the

model digester was simulated over a wide range of fractional conversion

factors (0.2 -0.8), and percentage total solids concentrations (4 – 10% for

the CSTR, and 10 – 30% for the batch digester). The results of the

simulation tabulated in Tables AlIl-1 - AIll-6, and AlV-1 – AlV-5 have been

presented graphically in Figures 5.1 – 5.19.

The simulation considered the effect of the fractional conversion on

the volume of the digester; the volumes of methane and biogas; the time

required for digestion; and the effluent substrate concentration and

microbial growth. The resulting curves from the above relationships were

analysed statistically using the Microsoft Chart Editor. From Figure 5.1, it is

observed that the volumes of biogas and methane produced were a linear

function of the fractional conversion, described by the following models:

Vb = 18432  - 89.725 (5.1)

and Vm = 11059  - 53.831 (5.2)

which are indicative of a proportionate increase of the volume of the

products with increasing fractional conversion, although with wider marginal

increment towards the upper boundary of the fractional conversion.

This relationship further confirms the rich organic and indeed

biodegradable fraction of the MSW. On the other hand, from Figure 5.2 the

volume of the digester increases as a logarithmic function of the fractional

conversion, thus:

Vbd = 115.84 ln  + 11695 (5.3)


166

This shows that increasing fractional conversion would result in an increase

in volume, up to a point when less and lesser volume would be required for

a given total solids concentration. This volume - fractional conversion

relationship is further proven by the logarithmic relationship between the

time of batch digestion and fractional conversion, generated from Figure 5.3

thus:

td = 0.1113 ln  + 9.1063 (5.4)

which shows that with increasing fractional conversion a point will also be

reached when less and lesser time would be used for the digestion of a

given amount of total solids.

Also Figure 5.4 shows that the effluent concentration reduces with

increasing fractional conversion. While, although the microbial growth

increases with fractional conversion, it tends to constancy at higher levels of

fractional conversion, which appears to be representative of the stationary

phase of microbial growth in pure cultures as the limiting substrate

concentration reduces.

This effect of fractional conversion on the effluent substrate and

microbial concentrations are described mathematically by the following

equations.

Se = - 22.389  + 27.753 (5.5)

Xe = 7.8864  - 0.1219 (5.6)

All these relationships suggest that an increased gas production and

digester volume would accompany a high percentage stabilization of the


167

waste, since there is a resultant reduction in effluent concentration. Relating

Figures 5.2 and 5.5 it is seen that as the volume of the digester increases,

the cost also increases in the same manner as the volume, and described

by the logarithmic function.

Xdb = 4x106 ln  + 6x108 (5.7)

The effect of microbial growth on effluent substrate concentration and

the volume of digester, as well as the effect of volume of digester on the

volume of methane and hence biogas were also analysed. The results

shown in Figures 5.6 – 5.8 resulted in the development of the following

models:

Se = - 3.037 Xe + 27.851 (5.8)

Vbd = 117.14 ln Xe + 114 (5.9)

Vmb = 9x10-40 e0.0085Vbd (5.10)

With these models it is possible to determine the volume of batch

digester required to achieve a desired level of effluent substrate

concentration with a given level of microbial concentration; and also the

associated volume of methane and biogas to be produced from the

digestion of municipal solid waste.

For the CSTR simulation, the trend in the relationship of process

parameters is almost the same, except for model constants. Thus the

following models generated statistically from Figures 5.9 – 5.15 describe the

relationship between the process parameters.


168

Vbc = 74.189  – 2.8304 (5.11)

Vmc = 104.03  – 3.910 (5.12)

Vcd = 10.909 ln  + 202.62 (5.13)

tcd = 0.4377 ln  + 8.1248 (5.14)

Se = 5.4829  + 6.7963 (5.15)

Xe = 1.805  + 0.0322 (5.16)

Xdc = 2x106 ln  + 5 x 107 (5.17)

Se = -3.0376 xe + 6.8941 (5.18)

Vcd = 11.412 ln xe + 195.79 (5.19)

5.2.2.8 Evaluation of Batch and Continuous Digesters

Bailey and Ollis (1981) say the relationship between batch and

continuous biological reactors is usually assessed on the bases of biomass

production, substrate utilization and product yield. For a successful

evaluation of the processes, the batch and continuous models were

analyzed for the same level of the total solids concentration (10%TS), and

the results are presented in Table 5.8.


169

Table 5.8: Comparison of Batch and Continuous Digesters’


Performance at 10% TS

 t Se Xe Vol. of Digester Vol. of Methane Cost


Batch CSTR Batch CSTR Batch CSTR Batch CSTR Batch CSTR Batch CSTR
0.2 8.45 7.40 5.70 5.70 0.39 0.39 13867.30 184.55 351.92 17.31 659 49
0.3 8.69 7.62 5.15 5.15 0.57 0.57 14243.89 189.99 554.40 27.34 669 50
0.4 8.82 7.74 4.60 4.60 0.75 0.75 14444.77 193.03 758.43 37.49 675 51
0.5 8.90 7.82 4.06 4.06 0.94 0.94 14574.41 195.12 963.58 47.75 679 51
0.6 8.96 7.89 3.51 3.51 1.12 1.12 14668.43 196.83 1169.85 58.17 681 51
0.7 9.01 7.96 2.96 2.96 1.30 1.30 14743.59 198.44 1377.45 68.82 683 52
0.8 9.05 8.03 2.41 2.41 1.48 1.48 14809.47 200.22 1586.85 79.86 685 52

From the data provided in the Table, taking a point fractional

conversion of 0.8, the biomass production per unit volume in the batch

digester is 9.994x10-5 kg/m3 (0.09994 mg/l) and is greatly lower than that of

0.007392 kg/m3 (7.392 mg/l) for the CSTR. This is also the case for the

amount of methane produced per unit volume of digester, which is 0.1072

m3 CH4 per m3 of digester for the batch, and 0.3989 m3 CH4 per m3 of

CSTR. Also the cost of the digester follows this trend, as the cost per unit

volume of batch digester is N0.046 and N0.26 for the CSTR. This shows

that the cost of a unit volume of CSTR is about 6 times more than that of a

batch digester; and when this is related to the amount of methane/biogas

per unit volume of digester, it shows that the cost of a unit volume of gas in

a batch digester is N0.43 against N0.65 for the CSTR, indicating that it cost

more to produce gas in the CSTR than in the batch digester.

Considering the time of digestion, it is observed that the time

required to achieve the same level of microbial growth in the CSTR is lower

than that of the batch, which is an indication that “the continuous process
170

always provides a greater yield of cells per unit volume of cultivator vessel

than a batch process does” (Bailey and Ollis, 1981).

At this point, it can easily be deduced that the initial cost investment

in a batch digester is higher than that of a CSTR, but the overall amount of

gas produced is higher in the batch.

It is important to state here that this comparison of the batch and

continuous digesters is based on the same level of total solids concentration

of 10%. However, it has been noted earlier that one of the major distinctions

between the two systems occasioned by their respective flow regimes is

that, whereas the batch digester is suitable for high solids processing, the

CSTR is suitable for low solids processing with an upper limit of 10%TS

concentration. Therefore, the batch digester can handle higher levels of

total solids concentration than the CSTR.

In order to ascertain the desirable level of percentage total solids

concentration of municipal solid waste in the batch digester, the model

digester was simulated over a range of total solids concentration (10% -

30%) at a particular fractional conversion of 0.8.

The results of this simulation are presented in Table All-6, and

analysed using Figures 5.16 – 5.19.

From these Figures, relationships were established respectively

between the percentage total solids (PTS) and volume of batch digester,

time of digestion and microbial concentration. Figure 5.17 shows that as the

PTS increases, the volume of digester reduces, apparently because of an

increase in the level of microbial concentration. This continues until after a


171

PTS of 20% when the volume starts increasing. When this trend is related

to the constancy of time of digestion from a PTS of 20% (Figure 5.16), it is

believed that this point, which can more or less be described as a point of

inflexion, could be considered the optimum PTS concentration for the

operation of a batch digester treating municipal solid waste for biogas

generation.
Volume of gas produced, cubic meter

16000
14000
12000
10000
8000 Vm

6000 Vt

4000 Linear
(Vm)
Linear
2000
(Vt)
Linear
0 (Vm)
0 0.2 0.4 0.6 0.8 1
Fractional conversion, dimensionless

Figure 5.1: Variation of Volume of Gas Produced in the Batch


Digester with Fractional Conversion
172

11680

11660

Volume of batch digester, cubic meter


11640

11620

11600

11580

11560

11540

11520

11500

11480
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9
Fractional conversion, dimensionless

Figure 5.2: Variation of Volume of Batch Digester with Fractional Conversion

9.1

9.08

9.06
Time of digestion, days

9.04

9.02

8.98

8.96

8.94

8.92

8.9
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9
Fractional conversion, dimensionless

Figure 5.3: Variation of Time of Batch Digestion with Fractional


Conversion
173

25

Effluent substrate & microbial concentration, kg/cubic


20

15
meter

10

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9
Fractional conversion, dimensionless

Figure 5.4: Variation of Effluent Substrate & Microbial Concentrations


with Fractional Conversion in Batch Digester

595000000

594000000
cost of batch digester, naira

593000000

592000000
y = 4E+06Ln(x) + 6E+08
591000000 R2 = 0.9843

590000000

589000000

588000000
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9
Fractional conversion, dimensionless

Figure 5.5: Variation of Cost of Batch Digester with Fractional Conversion


174

25

Effluent substrate concentration, kg/cubic meter


20

15

10

0
0 1 2 3 4 5 6 7
Microbial growth, kg/cubic meter

Figure 5.6: Effect of Microbial Growth on Batch Digester Effluent Substrate


Concentration

11680

11660
Volume of batch digester, cubic meter

11640

11620

11600

11580

11560

11540

11520

11500

11480
0 1 2 3 4 5 6 7
Microbial growth, kg/cubic meter

Figure 5.7: Effect of Microbial Growth on Volume of Batch Digester


175

16000

14000

Volume of gas, cubic meter


12000

10000

8000
Vm

6000
Vt

4000 Expon. (Vm)

2000 Expon. (Vt)

0
11450 11500 11550 11600 11650 11700
Volume of batch digester, cubic meter

Figure 5.8: Effect of Volume of Batch Digester on Gas Production


Volume of gas in CSTR, cubic meter

140

120

100

80

60
Vmc
40 Vtc
Linear (Vmc)
20
Linear (Vtc)
0
0 0.2 0.4 0.6 0.8 1
Fractional conversion, dimensionless

Figure 5.9: Variation of Volume of Gas in CSTR with Fractional Conversion


176

202
200

Volume of CSTR, cubic meter


198
196
194
192
190
188
186
184
182
0 0.2 0.4 0.6 0.8 1
Fractional conversion, dimensionless

Figure 5.10: Variation of Volume of CSTR with Fractional Conversion

8.1
Time of CSTR digestion, days

8
7.9
7.8
7.7
7.6
7.5
7.4
7.3
0 0.2 0.4 0.6 0.8 1
Fractional conversion, dimensionless

Figure 5.11: Variation of time of CSTR digestion with fractional conversion


177

concentration, kg/cubic meter


CSTR effluent and microbial
5 Sec
Xec
4

0
0 0.2 0.4 0.6 0.8 1
Fractional conversion, dimensionless

Figure 5.12: Variation of CSTR Effluent Substrate and Microbial Concentrations


with fractional conversion

52000000

51500000

51000000
Cost of CSTR, Naira

50500000

50000000

49500000

49000000
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9
Fractional conversion, dimensionless

Figure 5.13: Variation of Cost of CSTR with Fractional Conversion


178

Effluent substrate concentration, kg/cubic meter


5

0
0 0.5 1 1.5 2
Microbial growth, kg/cubic meter

Figure 5.14: Effect of Microbial Growth on CSTR Effluent Substrate


Concentration

202

200
Volume of CSTR digester, cubic meter

198

196

194

192

190

188

186

184

182
0 0.5 1 1.5 2
Microbial growth, kg/cubic meter

Figure 5.15: Effect of Microbial Growth on Volume of CSTR


179

9.12
9.1
time of digestion, days 9.08
9.06
9.04
9.02
9
8.98
8.96
8.94
0 10 20 30 40
percentage total solids (%)

Figure 5.16: Relationship between Percentage Total Solids and Time of Digestion

30000
volume of batch digester, cubic meter

25000

20000

15000

10000

5000

0
0 10 20 30 40
percentage total solids, (%)

Figure 5.17: Effect of Percentage Total Solids on volume of batch digester


180

25000

volume of gas, cubic meter


20000

15000

10000
Vm
Vt
5000
Power (Vt)

0
0 10 20 30 40
percentage total solids, (%)

Figure 5.18: Effect of Percentage Total Solids on volume of biogas and


methane produced in Batch Digester

60

50
Se
Xe
Substrate and Microbial

40 Power (Xe)
Concentration, mg/l

Power (Se)
30

20

10

0
0 10 20 30 40
percentage total solids, (%)

Figure 5.19: Effect of Percentage Total Solids on substrate and microbial


concentration
181

CHAPTER 6

CONCLUSIONS AND RECOMMENDATIONS

6.1 CONCLUSIONS

The management of the environment for enhanced environmental

quality has being an issue of serious concern to environmental managers.

An effective approach is the integrated management style, which includes

the treatment of MSW resulting in biogas when degraded with

microorganism in an anaerobic biological reactor.

The design of biological reactors is primarily aimed at determining

the size and type of reactor and method of operation that are best suited for

a given treatment process. This work originally set out to design a digester

for the treatment of municipal solid waste (MSW) with a resultant production

of biogas; but it became necessary first to investigate the energy potential of

MSW, especially as generated in the Port Harcourt metropolis, and second

to determine which process flow regime (in terms of whether batch or

continuous would produce optimum operation performance. The dilemma in

this consideration is that whereas the batch system is usually associated

with high solids wastes, and for low volume, the CSTR considers low-solids

wastes and for high volume; and the MSW in Port Harcourt metropolis is of

high solids concentration and also of high volume.

The high organic content of the MSW and the associated C : N ratio

as found in this work predisposes the waste as a veritable product for


182

biogas generation; so that it is possible to generate commercial quantity of

useful biogas from MSW.

Results obtained from field and laboratory experimentations reveal

that MSW is not easily degraded anaerobically, and its eventual degradation

is also not associated with high COD removal. This is easily attributable to

the composite nature of the waste and the associated level of refractory

organics. So, an effective conversion of MSW to energy in an anaerobic

degradation process is predicated upon a controlled source separation of

the waste into its various components in order that only the organic portion

is treated.

The results also facilitated the development of design models for

both batch and continuous digesters. Upon simulation of the system over a

range of percentage total solids concentration and fractional conversion, the

results of the batch digester showed that the time and volume required for

batch digestion can be decreased by increasing the initial concentration of

microbes and decreasing initial substrate concentration. On the other hand,

the CSTR would perform better in less time if it were operated with sludge

re-cycle. On the overall a comparative evaluation of the batch and

continuous system at the same total solids concentration showed that

although a larger volume is required by the batch digester to handle the

same amount of waste as the CSTR, and the initial cost of building a batch

digester is higher, the cost of a unit volume of CSTR is about 6 times more

than that of a batch digester; and the cost of producing the same quantity of

gas in a CSTR is about 2 times that of a batch digester.


183

When these finds are considered against the background of the

ability of the batch digester to handle higher levels of percentage total

solids concentration then it is very easy to state that the batch digester is

more appropriate for the digestion of MSW for biogas production.

6.2 RECOMMENDATIONS

In view of the finds and conclusions drawn from this work, the

following recommendations are prescribed.

- Environmental managers should begin to consider anaerobic

digestion as a veritable mode of treatment for municipal solid waste.

- The anaerobic digestion of MSW should be exploited as a means of

generating alternative renewable energy source for our country,

Nigeria.

- The digestion should be executed in a batch digester operating at not

more than 20% total solids concentration.

- Source separation of the MSW should be encouraged.

- A batch digester should be constructed in future researches to

validate and improve upon the models developed in this work.

- The digester should be operated with appropriate initial concentration

of biomass, and with cell re-cycle.

- The by-product of the digestion process should be investigated for

further re-use, perhaps as fertilizer.


184

CONTRIBUTION TO KNOWLEDGE

This work has added the following developments to existing knowledge

on the subject matter, while also modifying some hitherto erroneous wave of

thinking.

- Data on the properties of municipal solid waste (MSW) in Port

Harcourt metropolis has been provided.

- The kinetic parameters of the anaerobic biochemical degradation of

MSW have also been provided.

- Models that will ease the design of anaerobic batch digesters for the

production of biogas from MSW have been formulated.

- It has been found that, contrary to unverified postulations, the batch

digester is better suited for the anaerobic degradation of MSW in the

production of biogas, compared to the continuous-flow digester.

- Again, contrary to existing literature that high solids digestion should

be for total solids concentration of greater than or equal to 22%

(Tchobanoglous et al, 1993), it has been found that the digestion of

MSW in the production of biogas is optimal at a total solids

concentration of 20%.
185

REFERENCES

Abowei, M. F. N. and Ogoni, H. A. (1990): "Development of Design


Equations for CSTR Reverse Liquid Phase Reaction. International
Journal of Modelling, Simulation and Control. 20, 4, 31 - 43.

Aguwamba, J. C. (2001): “Waste: Engineering and Management Tools”.


Immaculate Publications Ltd., Enugu, Nigeria. 572p.

Akintola, O. A. (1978): “Aspects of Solid Waste Management in Ibadan


City”, In Ajaegbu And Morgans (Eds). Geography And Planning In
Nigeria. Nigerian Geographical Association; University of Jos. Pp
111-124.

American Public Health Association; American Water Works Association,


and Water Environment Federation (1995): Standard Methods for
Examination of Water and Wastewater, 19th Edition.

Andrews, J. F. (1978): “The Development of a Dynamic Model and Control


Strategies for the Anaerobic Digestion Process”. In James, A (Ed.)
Mathematical Models in Water Pollution Control; John Wiley & Sons,
Chichester. 420p

AOAC (1995): Official Methods of Analysis, 16th Edition; Association of


Official Analytical Chemist, Washington D. C.

DHV Consulting Engineers (2000): “Aspects of Rural Centre Planning”,


Volume 3, Energy, Amersfoot, USA. 47p.

AWAKE Magazine (November, 2003); Benin City, Nigeria, p28.

Ayotamuno, M. J.; Akor, A. J; Teme, S. C; Essiet, E. W. U; Isirimah, N. O.,


and Idike, F. I. (2000): “Relating Corn Yield to Water Use During the
Ery Season in Port Harcourt Area, Nigeria”. Agricultural
Mechanization in Asia, Africa and Latin America (AMA); Japan. Vol.
31, No. 4, pp 47 – 51.

Ayotamuno. M. J. and Gobo, A. E. (2004): "Municipal Solid Waste


Management in Port Harcourt, Nigeria: Obstacles and Prospects.
Management of Environmental Quality: An International Journal, Vol.
15 No. 4. Emarald Group Publishing Limited. P389 - 398.

Bailey, J. E. and Ollis, D.F. (1986): Biochemical Engineering Fundamentals;


McGraw-Hill Book Publishers, New York, 753p.

Bailie, R. C., Everett, J. W., Liptak, B. G., Liu, D. H. F., Rugg, F. M. and
Switzenbaum, M. S. (1996); “Solid Waste”, In Environmental
186

Engineers Handbook. (Eds) Liu, D. H. F., Liptak, B. G., and Bouis, P.


A., Lewis Publishers, New York. 1148 – 1248.

Baker, W. E. Westine, P. S. And Dodge, F. T. (1991): Similarity Method in


Engineering Dynamics, Theory and Practice of Scale Modeling;
Elsevier Science Publishing Company Incorporated, New York. 384p

Betty, L. (1967): Waste Disposal Systems. Rhoda Publication Series No.


38, New York, p98

Bitrus, R. (2001): “Design of an Anaerobic Bio-Reactor for the Production of


Methane from Solid Waste Organic Matter”, Unpublished M. Tech.
Thesis of Chemical/Petrochemical Engineering, Rivers State
University of Science and Technology, Port Harcourt. 109p.

Bushwell, A. M. and Mueller, H. F. (1952): “Mechanics of Methane


Fermentation”. Journal of Industrial Engineering Chemistry, 44(3),
550.

Byrne, K. (1997): Environmental Science, Thomas Nelson & Sons Ltd, Uk.
206 p

Chawler, O. P. (1986): Advances In Biogas Technology, India Corporation


of Agricultural Research. 28 p

Coulson, J. M. and Richardson, J. F. (1983): Chemical Engineering, Volume


6, 1st edition, Pergamon Press, Oxford.

Coulson, J. M. and Richardson, J. F. (1991): Chemical Engineering, Volume


2, 4th edition, Pergamon Press, Oxford.

David-West, T. (2001): “Sources And Control of Environmental Pollution in


the Niger Delta Region”, 6th Engr. Souza Okpofabiri Memorial
Lecture, Nigerian Society of Engineers, Port Harocurt. Pp. 9-15.

Eckenfelder, W. W. Jr (2000): Industrial Water Pollution Control. McGraw-


Hill Higher Education, Boston Burr Ridge Pp 394-411.

Ekenta, O. E. (2001): “Solid Waste Management in Awka Municipality”


Proceedings of the National Engineering Conference and Annual
General Meeting of the Nigerian Society of Engineers. Pp. 86-92.

Esan, A. A. (2001): “Sustainable National Energy Development”,


Proceedings of the National Engineering Conference and Annual
General Meeting of the Nigerian Society of Engineers. Pp. 140-156.
187

Evseev, M. K. (1972): Mechanization and Electrification of Livestock


Production. Kolos, Moscow. 415p

Eze, C. L. (2004): Alternative Energy Resources (With comments on


Nigeria's position). First Edition. Macmillan Nigeria Publishers
Limited, Lagos. 237 Pp.

Federal Environmental Protection Agency Guidelines (2000): Guidelines


and Standards for Environmental Pollution and Control in Nigeria.
P36

Federal Office of Statistics (2003): “General Demographic and Health


Survey of Nigeria”, Federal Office of Statistics Bulletin; Federal
Ministry of Information and National Orientation, Abuja. P16

Food and Agriculture Organization (FAO) of the United Nations (1997):


“Biomass Fuels and Future: Interview with Gustapo Best, FAO
Senior Energy Coordinator; http://www.fao.org/NEWS/1997/971202-
e.htm. 3p.

Fubara-Manuel, I; Ozogu, B. A. and Igoni, A. H. (2000): “Maize Yield


Response to Different Irrigation Intervals”. African Journal of
Agricultural Teacher Education. Vol IX (1 & 2) pp 127 – 134.

General Environmental Multilingual Thesaurus (GEMET, 2000): “Biomass


Resources”. http://glossary.eea.eu.int./EEAGlossary/B/Biogas

Graef, S. P. and Andrews, J. F. (1973): “Mathematical Modelling and


Control of Anaerobic Digestion”, in C. F. Bennett (Ed), Water,
Chemical Engineering Progress Symposium Series, No. 136, Vol. 79,
p 101.

Harris, P. (2003): Beginners Guide to Biogas:


http://www.ees.adelaide.edu.au/pharis/biogas/beginners/html. 11p

Herbert, D., Elsworth, R. and Telling, R. C. (1956): Journal of General


Microbiology, 14,601.

Hobson, P. N., Bousfield, S., and R. Summers (1981): Methane Production


from Agricultural and Domestic Wastes, Applied Science Publishers
Ltd., London. 269p

Hughes, W. L. (1976): “Energy For Rural Development – Renewable


Resources and Alternative Technologies for Developing Countries”,
(Edited) Report Submitted to National Academy of Sciences,
Washington. 307p.
188

Illoeje, O. C. (1998) “Rural Energy Needs And Fire Supply Technologies".


Energy Commission of Nigeria.

Ilori, T. A. (2001): “Sustainable Energy Development through Conservation”,


Proceedings of the National Engineering Conference and Annual
General Meeting of the Nigeria Society of Engineers. Pp. 212 – 219.

Itodo, I. N. and Awulu, J. O. (1999): “Effects of Total Solids Concentration of


Poultry, Cattle and Piggery Waste Slurries on Biogas Yield”.
Transactions of the American Society of Agricultural Engineers. Vol
42(6). Pp1853 – 1855.

Itodo, I. N. and Phillips, T. K. (2001): “Determination of Suitable Material for


Anaerobic Biogas Digesters”, Proceedings of the 2nd International
Conference and 23rd Annual General Meeting of the Nigerian
Institution of Agricultural Engineers. Vol. 23 Pp. 437-441.

Kiely, G. (1998): Environmental Engineering, International Edition, Irwin


Mcgraw-Hill, Boston. 979p

Kahaynian, M., Lindenauer, K., Hardy, S., and Tchobanoglous, G. (1991):


“Two-Stage Process Combines Anaerobic and Aerobic Methods”,
Biocycle, March.

Kottner, M. (2002): "Biogas In Agriculture And Industry: Potentials, Present


Use And Perspectives". Waste Management World; Issue 1. James
and James (Science Publishers) Ltd; London. Jackie.Jones@
Jxj.Com. Pp 8 –14

Levenspiel, O. (1999): Chemical Reaction Engineering, (3rd Edition). John


Wiley and Sons, Inc; New York. 668p.

Madu, C. And Sodeinde, O. A. (2001): “Relevance of Biomass in the


Sustainable Energy Development in Nigeria”, Proceedings of the
National Engineering Conference and Annual General Meeting of the
Nigerian Society of Engineers. Pp 220-227.

Mattocks, R. (1984): “Understanding Biogas Generation”, Technical Paper


No. 4. Volunteer in Technical Assistance, Virginia, USA. 13p

McNiel Technologies Incorporated (2000): "A Report on Assessment of


Biogas-to-Energy Generation Opportunities at Commercial Swine
Operations in Colorado”, Submitted to State Colorado Governor’s
Office of Energy Conservation and Management and Department of
Energy Western Regional Biomass Energy Programme. 8p

Michaelis, L. and Menten, M. L. (1913): Biochem, 49, 333.


189

Monod, J. (1949): “The Growth of Bacterial Cultures”, Annual Review of


Microbiology, Vol 111.

New Edition Concise English Dictionary (1999); “Waste” Harper Collins


Publishers, Glasgow; p1679

Novick, A. and Szilard, L. (1950): “Experiments with the Chemostat on


Spontaneous Mutations of Bacteria”; Proceedings of National
Academy of Science, USA. Pp 708 – 719.

Ogumola, S. A. (1989): “Housing Refuse”, African Review of Business


Technology, Kenya. Pp. 43.

Ogunbiyi, A. (2001): "Local Technology in Solid Waste Management in


Nigeria” Proceedings of the National Engineering Conference and
Annual General Meeting of the Nigerian Society of Engineers. Pp.
73-79.

Olaoye, J. O. (2001): “Utilization of Biomass as Renewable Energy Source


in Nigeria”, Proceedings of the 2nd International Conference and 23rd
Annual General Meeting of the Nigerian Institution of Agricultural
Engineers. Vol. 23. Pp.457-462.

Oluka, S. I. (2001): "Effects of Household Size, Income Level, and Food


Consumption Pattern on Solid Waste Generation in Enugu Urban of
Nigeria”, Proceedings of the 2nd International Conference and 23rd
Annual General Meeting of the Nigerian Institution of Agricultural
Engineers. Vol. 23. Pp. 367-373.

Oregon State Department of Energy (2002): "Biomass Energy Technology"


[Online]; Available; www.oregondoe.org. 19p

Oyinlola, A. K. (2001): “Recycling/Resource Recovery Technology Option


for a Sustainable Solid Waste Management System In Nigeria”,
Proceedings of the National Engineering Conference and Annual
General Meeting of The Nigerian Society of Engineers. P. 63-72.

Perry, H. R. and Green, W. D. (1997): Perry’s Chemical Engineering


Handbook. 7th Edition. McGraw-Hill Publishers, New York. P21 – 8.

Peter, M. S. and Timmerhaus, K. D. (1981): Plant Design and Economics.


McGraw-Hill Book Company, Tokyo. P537.

Pickford, J. (1977): “Solid Waste In Hot Climates” In Feachen, et al (Ed)


Wastes And Health In Hot Climates. John Wiley And Sons
Publishers, P36-38
190

Raymus, G. J. (1997): “Handling of Bulk Solids and Packaging of Solids and


Liquids”. In Perry, H. R., Green, W. D., and Maloney, J. O. (Eds.):
Perry’s Chemical Engineering Handbook. 7th Edition. McGraw-Hill
Publishers, New York. 78p.

Reynolds, T. D. And Richards, P. A. (1996): Units Operations And


Processes In Environmental Engineering, Second Edition, PWS
Publishing Company, Boston. 798p

Richard, T. L. (1998): Municipal Solid Waste Composting: Physical


Processing. www/dept/compost/MSW.Factsheets/msw.fs1.html. 9p

Richard, T. L. (1998b): Municipal Solid Waste Composting: Biological


Processing. www/dept/compost/MSW.Factsheets/msw.fs2.html. 8p.

Rogers, G. F. C. and Mayhew, Y. R. (1982): Engineering Thermodynamics


– Work and Heat Transfer. English Language Book Society and
Longman Group Limited, London. 667p.

Sandbach, F. (1980): Environmental Ideology And Policy, Basil Blackwell


Publishers, England. P183-195.

Seedtree Biogas Program: http:/www.seedtree.org/biogas.html. 6p

Sincero, A. P. and Sincero, G. A. (1999): Environmental Engineering – A


Design Approach. Prentice-Hall of India Private Limited, New Delhi.
795p.

Stafford, O. A., Hawks, O. L., And R. Horton (1980): Methane Production


From Waste Organic Matter, CRC Press Inc. London. 306p

Steadman, P. (1975): Energy, Environment And Building, A Report To The


Academy Of Natural Sciences Of Philadelphia, Cambridge University
Press, Cambridge, London. 287p.

Tata Energy Research Institute, Teri, New Delhi, India. (1994): “Biogas: A
Source Of Rural Employment”. Tata Publications, New Delhi, India.
22p.

Tchobanoglous, G; Burton, F. L. (1991): Wastewater Engineering:


Treatment, Disposal And Reuse. Third Edition. Mcgraw-Hill Inc., New
York. 1334p

Tchobanoglous, G; Burton, F. L. And Stensel, H. D (2003): Wastewater


Engineering: Treatment And Reuse. Tata Mcgraw-Hill Publishing
Company Limited, New Delhi. 1819p
191

Tchobanoglous, G; Theisen, H. And Vigil, S. (1993): Integrated Solid


Waste Management: Engineering Principles And Management
Issues. Mcgraw-Hill Inc. New York. 978p

The Guardian Newspaper ‘World Leader’s Chart Fresh Path Way For
Sustainable Energy, Transport". Monday 7th 2001. Pp 31 & 45.

Trambouze P., Landeghem, H. V. And Wauguier, J. P. (1988): Chemical


Reactors (Design/Engineering/Operation). Gulf Publishing Company,
USA; 608p.

Vassiliou, N. (1997): “The Biogas Project Of Cyprus”, 1st International


Conference Of Energy And The Environment, Limassol, Cyprus. Pp.
757-761.

Viessman, W. (Jr) And Hammer, M. J. (1993): Water Supply And Pollution


Control. Harper Collins College Publishers New York, Pp 513-679.

Whitesides, R. W. (2001): “Process Equipment Cost Estimating by Ratio


and Proportion – a PDH online Course. www.PDHcenter.com. 8p

Willis, M. J. (2000): Continuous Stirred Tank Reactor Models; Department


of Chemical and Process Engineering, University of New Castle.
mark.willis@ncl.ac.uk. 15p

Wong, K. K. (1982): “Biological Kinetics Of Gas Production From The


Anaerobic Digestion Of Palm Oil Mill Effluent” Non Conventional
Energy Sources. ACIF Series - Vol. 3, Edited By Furlan, G. et al,
Proceedings Of The First Latin American School And Third
International Symposium. Pp 543 – 561.

Worika, I. L. (2002): Environmental Law and Policy of Petroleum


Development. Anpez Centre for Environment and Development, Port
Harcourt, Nigeria. 383p.

World Commission On Environment And Development (WCED) “Our


Common Future”, Oxford: Oxford University Press, 1987. P. 43.

Xuereb, P. (1997): Biogas – A Fuel Produced from Waste;


http://www.synapse.net.mt/mirin/newsletter/3836.asp. 2p.

Yelebe, Z. R. (2003): “Design of Complete Mix Anaerobic Activated Sludge


Digester (for Sewage Treatment). M. Tech. Thesis, Chemical/Petro-
Chemical Engineering, Rivers State University of Science and
Technology, Port Harcourt. 88p.
192

APPENDIX I

REPREESENTATION OF BATCH EXPERIMENT DATA

Concentration of MSW, S, mg/l 500


450
400
350
300
250
200
150
100
50
0
0 5 10 15 20 25 30
Digestion time, t, days

Figure AI-1: Variation of MSW concentration with time of digestion


Concentration of microorganism, X, mg/l

250

200

150

100

50

0
0 5 10 15 20 25 30
Digestion time, t, days

Figure AI-2: Growth of microorganism with time of digestion


193

500

Concentrations of MSW, S, & microorganism, X,


450

400

350

300
MSW

mg/l
250 Micro.
200

150

100

50

0
0 5 10 15 20 25 30
Digestion time, t, days

Figure AI-3: Combined plot of MSW and microorganism


concentrations with time

3
- ln (Se/So)

0
0 5 10 15 20 25 30
Digestion time, t, days

Figure AI-4: Plot for the verification of the process reaction order
194

45

Inverse of specific rate of MSW utilization, 1/U, per


40

35

30

25
day
20

15

10

0
0 0.05 0.1 0.15 0.2
Inverse of effluent concentration, 1/Se, l/mg

Figure AI-5: Plot for the determination of ‘k’ and ‘Ks’

0.6
Specific rate of MSW utilization, U, per day

0.5

0.4

0.3

0.2

0.1

0
0 50 100 150 200 250 300 350
Effluent MSW concentration, Se, mg/l

Figure AI-6: Plot for verification of half saturation constant, Ks, value
195

0.25

Inverse of mean cell residence time, 1/0c


0.2

0.15

0.1

0.05

0
0 0.1 0.2 0.3 0.4 0.5 0.6
Specific rate of MSW utilization, U, per day

Figure AI-7: Plot for the determination of Y and kd


196

APPENDIX II

FIELD DATA COMPUTATIONS

Total daily organic waste load = 1,043,038kg

Total solid = 80.1% of total load = 835,473.44kg

For a batch digestion requiring 26.7% TS,

Then TS of 26.7% = 278,491.15kg

Remaining TS = 835,473.44 - 278491.15

= 556, 982.29 kg/day

So, remaining total organic load = 1,043,038 - 556,982.29

= 486,055.71kg/day

This now represents the mass fraction of the organic matter with 26.7% TS

and 73.3% moisture content.

Therefore with organic load of 486,055.71kg/day

TS of 26.7% or organic load = 129,776.87 kg/day

VS = 66.6% TS = 86,431.40 kg/day

FS = 13.5% TS = 17,519.88 kg/day

Using the maximum limit of the fractional conversion of 0.8, then:

Volatile solids (VS) destroyed = 0.8 x 86431.40

= 69,145.12 kg/day

And volatile solids in digested sludge = 86,431.40 – 69,145.12

= 17,286.28 kg/day

 TS in digested sludge = 17,286.28 + 17,519.88

= 34,806.16 kg/day
197

The specific gravity of dry sludge is determined thus:

250
S gd 
100  1.5(66.6)

= 1.251

Specific gravity of wet sludge is

100  1.251
S gw 
73.3(1.251)  (100  73.3)

= 1.057

Daily volume of waste of 26.7% TS becomes

100 1 1
486,055.71  
26.7 1.057 1000

= 1722.26 m3/day

Concentration of feed MSW,

86431.40
So 
1722.26

= 50.18 kg/m3

Percentage of TS in digested sludge is

34806.16
 100
129776.17

= 26.8%

 Daily volume of waste of 26.8% TS will be

100 1 1
34806.16   
26.8 1.251 1000

= 103.82 m3/day

Biodegradability of the waste, BF = 0.83 – 0.028(0.83)


198

= 0.807

Refractory portion of the waste, R = 1 – 0.807

= 0.193

Initial biodegradable substrate concentration

Sib = 50.18 – (0.193 x 50.18)

= 40.5 kg/m3

and effluent biodegradable substrate concentration

Seb = 40.5 (1 - 0.8)

= 8.1 kg/m3

Concentration of effluent, MSW

Se = 50.18 (1 - 0.8) + (0.8 x 50.18 x 0.193)

= 17.79 kg/m3

 50.18  17.79 
 Percentage Stabilization, Eb     100
 50.18 

= 64.5%

Similarly, for a TS of 10% of total organic waste load, i.e. 104,303.80

kg/day

Remaining TS = 835,473.44 – 104,303.80

= 731,169.64 kg/day

and remaining total organic load = 1,044,303.80 – 731,169.64

= 311,868.36 kg/day

Now, this represents the mass of organic matter with 10% TS and 90%

moisture content.

Therefore with organic load of 311,868.36kg/day


199

TS of 10% of organic load = 31,186.84kg/day

VS = 66.6% TS = 20,770.43kg/day

FS = 13.5% TS = 4,210.22kg/day

Volatile Solids (VS) destroyed = 0.8 x 20,770.43

= 16,616.34 kg/day

Volatile Solids (VS) in digested sludge = 20,770.43 – 16616.34.30

= 4154.09 kg/day

Total Solids (TS ) in digested sludge = 4210.22 + 4154.09

= 8364.31 kg/day

Here specific gravity of wet sludge is

100  1.251
S gwc 
90(1.251)  (100  90)

= 1.02

Daily volume of waste of 10% TS

100 1 1
311,868.36   
10 1.02 1000

= 3,057.53 m3/day

Concentration of influent MSW,

20,770.43
Soc 
3,057.53

= 6.79 kg/m3

Percentage TS in digested sludge

8364.31
=  100
31,186.84

= 26.8%
200

Daily volume of waste of 26.8% TS

100 1 1
= 8364.31   
26.8 1.251 1000

= 24.95 m3/day

Initial biodegradable substrate concentration

Sic = 6.79 – (0.193 x 6.79)

= 5.48 kg/m3

Effluent biodegradable substrate concentration

Sebs = 5.48 (1 - 0.8)

= 1.096 kg/m3

Concentration of effluent MSW

Sec = 6.79(1 - 0.8) + (0.8 x 6.79 x 0.193)

= 2.41 kg/m3

 6.79  2.41 
Percentage stabilization, Ec     100
 6.79 

= 64.5%
APPENDIX III

TABULATED SIMULATION RESULTS FOR THE BATCH DIGESTER

Table A III - 1: Summary of Batch Digester parameters at 10% TS (i.e. TS = 31186.84, and VS = 20770.43)
 Se Xe td Vbd Eb Vm VT Vmb Qnb Ogrb Xdb
0.2 5.70 0.393 8.45 13867.30 16.13 351.92 586.53 0.025 272.48 0.624 658,807,320.43
0.3 5.15 0.574 8.69 14243.09 24.20 554.40 924.00 0.039 272.17 0.935 669,461,828.60
0.4 4.60 0.754 8.82 14444.77 32.27 758.43 1264.05 0.053 271.86 1.246 675,133,534.83
0.5 4.05 0.935 8.90 14574.41 40.18 963.58 1605.97 0.066 271.55 1.558 678,762,575.71
0.6 3.507 1.115 8.96 14668.43 48.40 1169.85 1949.75 0.080 271.24 1.871 681,386,384.17
0.7 2.958 1.296 9.01 14743.59 56.47 1377.45 2295.75 0.093 270.93 2.185 683,478,988.98
0.8 2.41 1.476 9.05 14809.47 64.54 1586.85 2644.75 0.107 270.61 2.501 685,309,846.40

Table A III - 2: Summary of Batch Digester parameters at 15% TS (i.e. TS = 54603.04, and VS = 36365.62)
 Se Xe td Vbd Eb Vm VT Vmb Qnb Ogrb Xdb
0.2 12.957 0.853 8.79 12041.12 16.13 760.98 1268.29 0.063 317.54 1.175 605,289,757.08
0.3 11.710 1.263 8.91 12180.85 24.20 1166.72 1944.53 0.096 316.96 1.760 609,494,503.99
0.4 10.46 1.674 8.97 12254.28 32.27 1573.36 2622.27 0.128 316.37 2.346 611,696,331.81
0.5 9.22 2.084 9.00 12301.02 40.28 1980.79 3301.31 0.161 315.79 2.933 613,095,379.20
0.6 7.97 2.495 9.03 12334.81 48.40 2389.08 3981.80 0.194 315.20 3.520 614,105,071.19
0.7 6.725 2.905 9.05 12361.83 56.47 2798.47 4664.12 0.227 314.61 4.108 614,911,902.66
0.8 5.479 3.315 9.07 12385.61 64.54 3209.38 5348.96 0.259 314.02 4.697 615,621,371.01
202

Table A III - 3: Summary of Batch Digester parameters at 20% TS (i.e. TS = 83234.43, and VS = 55434.13)

 Se Xe td Vbd Eb Vm VT Vmb Qnb Ogrb Xdb


0.2 23.27 1.507 8.92 11494.00 16.13 1346.85 2244.75 0.117 362.36 1.981 588663903.04
0.3 21.04 2.244 8.98 11564.55 24.20 2044.72 3407.86 0.177 461.38 2.967 590714729.90
0.4 18.80 2.981 9.01 11597.87 32.27 2743.26 4572.10 0.237 360.39 3.955 591820835.21
0.5 16.56 3.718 9.03 11620.67 40.28 3442.46 5737.43 0.296 359.40 4.943 592518470.39
0.6 14.32 4.456 9.04 11637.00 48.40 4142.44 6904.06 0.356 358.42 5.931 593018131.80
0.7 12.08 5.193 9.07 11653.57 56.47 4843.42 8072.37 0.416 357.42 6.922 593524458.26
0.8 9.84 5.93 9.08 11655.17 64.54 5545.81 9243.01 0.476 356.43 7.913 593879083.26

Table A III - 4: Summary of Batch Digester parameters at 25% TS (i.e. TS = 117081.02, and VS = 77975.96)

 Se Xe td Vbd Eb Vm VT Vmb Qnb Ogrb Xdb


0.2 36.75 2.36 8.97 11633.92 16.13 2159.94 3599.89 0.186 406.84 3.141 592,923,879.07
0.3 33.22 3.52 9.02 11675.09 24.20 3264.13 5440.22 0.280 405.27 4.707 594,181,953.76
0.4 29.68 4.69 9.04 11696.46 32.27 4368.91 7281.52 0.374 403.71 6.273 594,834,104.30
0.5 26.14 5.85 9.05 11709.96 40.34 5474.30 9123.83 0.468 402.14 7.839 595,246,159.01
0.6 22.61 7.02 9.06 11719.68 48.40 6580.43 10967.38 0.562 400.57 9.406 595,542,453.40
0.7 19.07 8.18 9.07 11727.43 56.47 7687.52 12812.54 0.656 399.01 10.97 595,778,603.46
0.8 15.54 9.35 9.08 11734.23 64.54 8795.98 14659.97 0.750 397.44 12.54 595,985,852.57
203

Table A III - 5: Summary of Batch Digester parameters at 30% TS (i.e. TS = 156142.79, and VS = 103991.10)

 Se Xe td Vbd Eb Vm VT Vmb Qnb Ogrb Xdb


0.2 53.48 3.42 9.01 12213.57 16.13 3260.32 5433.87 0.267 450.84 4.78 610,476,461.00
0.3 48.34 5.11 9.04 12240.79 24.20 4914.98 8191.63 0.402 448.46 7.16 611,292,394.56
0.4 43.19 6.81 9.05 12254.89 32.27 6570.18 10950.29 0.536 446.07 9.54 611,714,594.21
0.5 38.05 8.50 9.06 12263.79 40.34 8225.96 13709.94 0.671 443.69 11.93 611,981,093.46
0.6 32.90 10.20 9.07 12270.18 48.40 9882.48 16470.80 0.805 441.31 14.31 612,172,601.24
0.7 27.76 11.89 9.07 12275.28 56.47 11539.97 19233.28 0.940 438.92 16.70 612,325,165.78
0.8 22.62 13.59 9.08 12279.75 64.54 13198.81 21998.02 1.075 436.54 19.08 612,459,013.28

Table AIII - 6: Batch Results at Various Percentage Total Solids (PTS) Concentrations

PTS TS VS Se Xe td Vd Vm Vt Vmb Qgrb Qnb Xdb


10 31186.84 20770.43 2.410 1.476 9.05 14809.47 1586.85 2644.75 0.107 2.501 270.61 685,309.846.40
15 54603.04 36365.62 5.479 3.315 9.07 12.385.61 3209.38 5348.96 0.259 4.697 314.02 615.621.371.01
20 83234.43 55434.13 9.84 5.93 9.08 11665.17 5545.81 9243.01 0.476 7.913 356.43 593.879.083.26
25 117081.02 77975.96 15.54 9.35 9.08 11734.23 8795.98 14659.97 0.750 12.543 397.44 595,985,852.57
30 156142.00 103991.10 53.48 13.59 9.08 12279.75 13198.81 21998.02 1.075 19.083 436.54 612.459.013.28
204

APPENDIX IV

TABULATED SIMULATION RESULTS FOR THE CSTR

Table AlV- 1: Summary of CSTR Parameters at 4% TS (TS = 9971.44, VS = 6640.98)


 Se Xe tc Vcd Ec Vmc Vtc Vmcs Qgrc Qnc Xdc
0.2 0.901 0.089 5.53 44.09 16.13 0.75 1.25 0.0170 0.00188 218.13 20,900,646.91
0.3 0.814 0.118 6.34 50.53 24.20 1.44 2.40 0.0285 0.00339 218.13 22,681,554.18
0.4 0.728 0.146 6.87 54.79 32.27 2.22 3.70 0.0405 0.00503 218.13 23,811,008.69
0.5 0.641 0.175 7.285 58.06 40.34 3.07 5.12 0.0529 0.00679 218.12 24,653,148.69
0.6 0.554 0.203 7.64 60.91 48.40 4.01 6.68 0.0658 0.00869 218.12 25,372,677.53
0.7 0.468 0.232 7.99 63.74 56.47 5.06 8.43 0.0793 0.0108 218.12 26,074,410.71
0.8 0.381 0.260 8.40 66.98 64.54 6.29 10.48 0.0939 0.0131 218.12 26,860,958.08

Table AIV - 2: Summary of CSTR Parameters at 6% TS (TS = 16208.81, VS = 10795.07)


 Se Xe tc Vcd Ec Vmc Vtc Vmcs Qgrc Qnc Xdc
0.2 2.035 0.161 6.61 85.66 16.13 3.26 5.43 0.0380 0.0103 236.38 31,134,422.23
0.3 1.839 0.226 7.11 92.12 24.20 5.56 9.27 0.0604 0.0170 236.37 33,333,839.43
0.4 1.644 0.290 7.41 95.99 32.27 7.97 13.29 0.0831 0.024 236.36 33,333,839.43
0.5 1.449 0.354 7.62 98.76 40.34 10.48 17.46 0.1061 0.0312 236.36 33,908,964.18
0.6 1.252 0.419 7.80 101.08 48.40 13.09 21.82 0.1295 0.0386 236.35 34,837,637.68
0.7 1.056 0.483 7.97 103.31 56.47 15.86 26.43 0.1535 0.0463 236.34 34,837,637.68
0.8 0.861 0.548 8.16 105.81 64.54 18.88 31.46 0.1784 0.0547 236.33 35,340,686.40
205

Table AIV - 3: Summary of CSTR Parameters at 8% TS (TS = 23280.61, VS = 15504.89)

 Se Xe tc Vcd Ec Vmc Vtc Vmcs Qgrc Qnc Xdc


0.2 3.633 0.262 7.13 132.67 16.13 8.49 14.14 0.0640 0.0330 254.61 40.477,926.30
0.3 3.283 0.377 7.45 138.62 24.20 13.76 22.93 0.0992 2524 254.59 41,558,169.21
0.4 2.934 0.492 7.63 142.02 32.27 19.14 31.91 0.1348 0.0722 254.57 42,167,611.96
0.5 2.585 0.607 7.76 144.40 40.34 24.64 41.06 0.1706 0.0923 254.55 42,590,407.96
0.6 2.235 0.723 7.86 146.36 48.40 30.26 50.44 0.2068 0.1128 254.53 42,934,912.98
0.7 1.886 0.838 7.96 148.21 56.47 36.08 60.13 0.2434 0.17338 254.51 43,261,049.16
0.8 1.536 0.953 8.07 150.28 64.54 42.22 70.36 0.2809 0.1558 254.49 43,621,260.04

Table AIV- 4: Summary of CSTR Parameters at 10% TS (TS = 31186.84, VS = 20770.43)

 Se Xe tc Vcd Ec Vmc Vtc Vmcs Qgrc Qnc Xdc


0.2 5.700 0.393 7.40 184.55 16.13 17.31 28.86 0.0938 0.0803 272.82 49,344,119.55
0.3 5.151 0.574 7.62 189.99 24.20 27.34 45.57 0.1439 0.1253 272.77 50,211,118.81
0.4 4.603 0.754 7.74 193.03 32.27 37.49 62.48 0.1942 0.1706 272.73 50,691,375.40
0.5 4.055 0.935 7.82 195.12 40.34 47.75 79.58 0.2447 0.2163 272.68 51,021,044.33
0.6 3.507 1.115 7.89 196.83 48.40 58.17 96.95 0.2955 0.2626 272.64 51,287,867.41
0.7 2.958 1.296 7.96 148.44 56.47 68.82 114.71 0.3468 0.3098 272.59 51,539,422.92
0.8 2.410 1.476 8.03 200.22 64.54 79.86 133.10 0.3989 0.3582 272.54 51,816,708.20
206

Table AIV -5: CSTR Results For Various Percentage Total Solids (PTS) Concentration
PTS TS VS Sc Xe tdc Vcd Vmc Vtc Vmcs Qgrc Qnc Xdc
4 9971.44 6640.98 0.381 0.2604 8.40 66.98 6.29 10.48 0.0939 0.0131 218.12 26860958.08
6 16208.81 10795.07 0.861 0.5479 8.16 105.81 18.88 31.46 0.1784 236.33 236.33 35340686.40
8 23280.61 15504.89 1.536 0.9527 8.07 150.28 42.22 70.36 0.2809 0.1558 254.49 43621260.04
10 31186.84 20770.43 2.410 1.4765 8.03 200.22 79.86 133.10 0.3989 0.358 272.54 51816708.20
207

APPENDIX V (POWER REQUIREMENT FOR MIXING)

Table A-V - Power Requirements For Screw Conveyor

Capacity Diam. Diam Diam. Hanger Max. Size of Lumps Max. Feed Hp At Motor Max.
of of of Centers Torque section capacity
Flights Pipe Shafts, ft Peed Capacity, diam., 15-ft. 30-ft 45-ft 60-ft 75-ft at
in in in r/min in lb in Max Max. Max. Max. Max speed
Length Length Length Length Length listed
All Lumps Lumps
Lumps 20 to 10%
25% or less
5 200 9 2½ 2 10 ¾ 1½ 2¼ 40 7,600 6 0.43 0.85 1.27 1.69 2.11 4.8
10 400 10 2½ 2 10 ¾ 1½ 2½ 55 7,600 9 0.85 1.69 2.25 3.00 3.75 6.6
15 600 10 2½ 2 10 ¾ 1½ 2½ 80 7,600 9 1.27 2.25 3.38 3.94 4.93 9.6
12 2½ 2 12 1 2 3 45 7,600 10 1.27 2.25 3.38 3.94 4.93 5.4
12 3½ 3 16,400 1.27 2.25 3.38 3.94 4.93 11.7

20 800 12 2½ 2 12 1 2 3 60 7,600 10 1.69 3.00 3.94 4.87 5.63 7.2


3½ 3 16,400 1.69 3.00 3.94 4.87 5.63 15.6
25 1000 12 2½ 2 12 1 1 2 3 75 7,600 10 2.12 3.75 4.93 5.63 6.55 9.0
14 3½ 3 ¼ 2½ 3½ 45 16,400 12 2.12 3.75 4.93 5.63 6.55 9.0
3½ 3 16,400 2.12 3.75 4.93 5.63 6.55 11.7
30 1200 14 3½ 3 12 1¼ 2½ 3½ 55 16,400 12 2.25 3.94 5.05 6.75 7.50 14.3

35 1400 14 3½ 3 12 1¼ 2½ 3½ 65 16,400 12 2.62 4.58 5.90 7.00 8.75 16.9


40 1600 16 3½ 3 12 1½ 3 4 50 16,400 14 3.00 4.50 6.75 8.00 10.00 13.0

Adapted from Raymus (1997)


208

APPENDIX VI (COST INDICES)

Table A-VI: Cost Indexes as Annual Averages

Chemical
Marshal and Swift Nelson –Farrar Engineering
Installed-Equipment Eng. News Record Refinery Plant Cost
Indexes, 1926 = 100 Construction Index Construction Index
Year All- Process 1913 1949 1967 Index 1957 – 1959
Industry Industry = 100 = 100 = 100 = 1946 = 100
= 100
1975 444 452 2412 464 207 576 182
1976 472 479 2401 503 224 616 192
1977 505 514 2576 540 241 653 204
1978 545 552 2776 582 259 701 219
1979 599 607 3003 630 281 757 239
1980 560 675 3237 679 303 823 261
1981 721 745 3535 741 330 904 297
1982 746 774 3825 802 357 977 314
1983 761 786 4066 852 380 1026 317
1984 780 806 4146 869 387 1061 323
1985 790 813 4195 879 392 1074 325
1986 798 817 4295 900 401 1090 318
1987 814 830 4406 924 412 1122 324
1988 852 870 4519 947 422 1165 343
1989 895 914 4606 965 429 1194 355
1990 (JAN) 904 924 4673 979 435 1203 356

Adapted from Peter and Timmerhaus (1981)


209

APPENDIX VlI

BATCH DIGESTER SIMULATION RESULTS


210
211
212
213
214
215
216
217
218
219
220
221
222
223
224
225
226
227
228
229
230

APPENDIX VIII

CSTR SIMULATION RESULTS


231
232
233
234
235
236
237
238
239
240
241
242
243
244
245
246
247

APPENDIX IX

MICROSOFT VISUAL BASIC PROGRAM FOR THE BATCH DIGESTER


248
249
250
251
252
253
254

APPENDIX X

MICROSOFT VISUAL BASIC PROGRAM FOR CSTR


255
256
257
258
259
260

APPENDIX XI

FLOWCHART FOR THE DIGESTER DESIGN


APPENDIX XI

MAP DESCRIBING STUDY AREA

View publication stats

Potrebbero piacerti anche