Vous êtes sur la page 1sur 120

ENERGY FROM WASTEWATER

A FEASIBILITY STUDY
TECHNICAL REPORT

Report to the
Water Research Commission
by
S Burton, B Cohen, S Harrison, S Pather-Elias,
W Stafford, R van Hille & H von Blottnitz
Department of Chemical Engineering
University of Cape Town

WRC Report No. 1732/1/09


ISBN 978-1-77005-849-1
Set No. 978-1-77005-846-0
JULY 2009

DISCLAIMER
This report has been reviewed by the Water Research Commission (WRC) and approved for
publication. Approval does not signify that the contents necessarily reflect the views and
policies of the WRC, nor does mention of trade names or commercial products constitute
endorsement or recommendation for use.

ii

EXECUTIVE SUMMARY
The current view of wastewaters is that they generally represent a burden and necessarily
incur energy costs in processing before they can safely be released into the environment. The
opportunity exists to improve the current wastewater treatment processes by applying new
solutions and technologies that can also reduce energy inputs and/or generate energy for
other processes. This study explored the various waste streams and the appropriate
technologies that could be used to generate energy.
A survey of the quality and quantity of wastewaters in South Africa identified the top three
sectors having the greatest potential energy recovery as the formal and informal animal
husbandry sector (cows, pigs and chickens), fruit and beverage industries (distillery, brewery,
winery, fruit juicing and canning) and domestic blackwater (sewage). An estimated 10 000
MWth can be recovered from the wastewaters in the whole of South Africa, representing 7%
of the current Eskom electrical power supply*. However, since most of the waste streams are
widely distributed, the energy from wastewater is best viewed as on-site power.
The most appropriate technologies and their limitations are partly determined by the value of
the required energy product (heat, electricity, combined heat and power or fuel) and the
driving market forces that determine how this energy product can be used with our current
technology. Furthermore, the ease of separation of the energy product from water can be key
to the feasibility of the process (e.g. biogas separates easy from wastewater by natural
partitioning whereas bioethanol requires energy intensive distillation). Anaerobic digestion
(AD) is the most commonly recognised technology and has been applied to wastewaters of
different characteristics at both small and large scales. AD is suitable for use with domestic
sewage (particularly since 40% of South Africans are currently not serviced with waterborne
sewage), as well as in the industrial and agricultural sectors. Bioethanol production by
fermentation is suited to concentrated, high carbohydrate wastewaters and has potential in
the fruit industry where sugar-rich wastewaters are generated in large volumes. Similarly,
combustion and gasification are restricted to applications with concentrated waste streams
(containing <40% water) due to the energy expended in de-watering and are most appropriate
in the treatment of dewatered and solar-dried (or previously stockpiled) wastes. In contrast to
these technologies, the growth of plant biomass for combustion/gasification and algal
biomass for biodiesel production is suited to dilute waste streams. The sequestration of
carbon dioxide and facilitated wastewater-treatment by photosynthetic oxygenation are added
benefits. However, there are no large scale algal biodiesel production processes operating at
present, in spite of published claims that the technology is technically an economically
feasible and the growing incentives due to the recent increase in diesel prices. Microbial fuel
cells (MFC) are an emerging technology that can also operate with dilute waste streams while
producing electricity directly. MFCs are suited to applications in remote/rural sites with no
*

Approximately 140 000 MWth or 42 000 MWe (Eskom data tables 2007). Where MWth and MWe refer the thermal
6
and electrical power in megawatts (10 W), respectively

iii

infrastructure, but the technology is still in early development. A single waste stream or
technology may not be suitable for achieving efficient energy recovery and the integration of
technologies and/or waste streams may be required to realise the maximum energy from
wastewater potential. There are also frequently missed opportunities for reducing energy
needs by the recovery of waste heat. The greatest potentials are realised when an industrial
ecology approach is used to integrate process or waste heat from several industries, with preplanning and the formation of industrial parks.

The net energy generated, reduction in pollution (wastewater treatment), and water
reclamation are the main costs and benefits considered in assessing the feasibility of an
energy from wastewater project. However, additional benefits such as certified emission
reductions (CERs), fertiliser production, or the production of other secondary products could
tip the balance of economic feasibility. For the implementation of energy from wastewater
technologies, essential services (WWTP operation, schools, and hospitals) and the needs of
communities not serviced by sewage and electrical infrastructure should preferentially be
targeted.
Several risks, barriers and drivers to developing an energy from wastewater project were
identified. There is a general lack of research capacity and skills, and a greater need for
research collaboration and information-sharing between research groups, government
agencies and municipal practitioners. There is also no incentive for the generation of clean,
renewable energy such as feed-in tariffs, green energy tariffs or peak tariffs.
The use of wastewater as a renewable energy resource can improve energy security while
reducing the environmental burdens of waste disposal. Energy from wastewater therefore
facilitates the integration of water, waste and energy management within a model of
sustainable development.

iv

CONTENTS:

Page

1. TECHNICAL REPORT: Appropriate energy from wastewater technologies and a


review of literature and national and international practice
1.1 Anaerobic Digestion (AD) to produce biogas
1.1.1
Introduction
1.1.2.
Advantages and constraints associated with AD
1.1.3
Application and practice of AD technology
1.1.4.
Implementation of AD
1.1.5
Costs and benefits
1.2 Combustion and gasification for the production of steam
1.2.1
Introduction
1.2.2
Application of combustion / gasification
1.2.3
Costs and benefits of combustion / gasification
1.3 Utilisation of waste heat
1.3.1
Introduction
1.3.2
Implementation of utilisation of waste heat
1.3.3
Costs and benefits in utilisation of waste heat
1.4 Fermentation to produce biomass and secondary products
1.4.1
Introduction
1.4.2
Implementation of fermentation of wastewaters
1.4.3
Cost and benefits in fermentation of wastewaters
1.5 Algal biomass for biodiesel production
1.5.1
Introduction
1.5.2
Implementation of biodiesel production
1.5.3
Costs and benefits in biodiesel production from wastewaters
1.6 Fermentation for production of bioethanol
1.6.1
Introduction
1.6.2
Implementation of bioethanol production
1.6.3
Costs and benefits in bioethanol production from wastewaters
1.7 Microbial fuel cells
1.7.1
Introduction
1.7.2
Implementation of MFCs
1.7.3
Costs and benefits in the application of MFCs to wastewaters
1.8 References

2. APPENDIX: Surveys, case studies and workshops


2.1
2.2
2.3
2.4
2.5
2.6
2.7
2.8
2.9
2.10

Survey of WWTP wastewaters


Survey of brewery wastewater
Survey of pulp and paper wastewaters
Survey of petrochemical wastewaters
Survey of Animal husbandry wastewaters
Case study of energy from fruit wastewaters
Case study of combustion versus gasification of waste stream
Case study of Cape Flats WWTP: Best practice?
Case study of the performance of household anaerobic digesters
Industrial stakeholder workshop

20

23

28

31

36

40

45
53

Acknowledgements:
Certain studies that have been incorporated into this project were undertaken by students at
UCT. These are acknowledged as follows:

Thermal gasification or anaerobic digestion of biomass for use in a distributed


renewable energy network (Sarah Dowling)

Optimising algal growth and lipid production (Melinda Griffiths)

Energy potential of the wastewaters from the paper and pulp and the brewing industry
(Andrew Krige)

Energy potential of wastewaters of the petroleum and animal husbandry industries


(Deleki Zuko and Sehoana Kabelo Albert)

Deriving energy from fruit wastes (Andrew de Beer and Peter Dawson)

Carbon dioxide sequestration and fixation by algae (Nick Langley and Bertus Louw)

Cape Flats WWTP: Best practice? (Loice Badza)

vi

1:

APPROPRIATE ENERGY FROM WASTEWATER TECHNOLOGIES AND NATIONAL AND


INTERNATIONAL PRACTICE

This report comprises a review of the literature with respect to energy produced from wastewater, and
a review of information available on the practice of energy generation from wastewater occurring in
South Africa and internationally.
This report constitutes the technical section of the final report for this research project, WRC
K5/1732. The report is divided into sections, each dealing with one technology, and providing, with
respect to the respective technology: Introduction on the theory of the technology; information on its
application and the energy potential.
SECTION 1.1 Anaerobic digestion to produce biogas
1.1.1

Introduction

Anaerobic digestion leads to the ultimate gasification of organic carbon containing wastewaters and
slurries to carbon dioxide, methane and hydrogen. It has been successfully used in the wastewater
treatment industry to reduce both the volume and COD of waste sludge. Anaerobic digestion of
wastewaters and sludges is a complex process, involving a number of different microorganisms. A
schematic representation of the process, typically consisting of three metabolic stages, is shown in
the process flow diagram in the Figure 1, below (GATE, 1999).

Figure 1. Schematic showing the stages and microorganisms involved in anaerobic digestion for biogas production.

Initially, particulate organic matter such as cellulose, hemicellulose, pectin and lignin are solubilised in
a hydrolysis step catalysed by extracellular enzymes. The soluble products of this stage are converted

into organic acids, alcohol, hydrogen and carbon dioxide by acidogenic microorganisms. The simple
organic substrates are further metabolised by acetogenic microorganisms to produce acetate, the
primary organic substrate for methanogenesis, hydrogen and carbon dioxide. Aside from acetate, CO2
and hydrogen, methanogenic microorganisms may also directly ferment other substrates, of which
formic acid and methanol are the most important (Bouallagui et al., 2007).
The AD digester can be designed as a cylindrical, cube shaped, or egg-shaped vessel and is
composed of a waterproof material with an inlet into which the fuel is introduced, normally in slurry
form. In small to medium scale facilities, the gas holder is typically an airtight steel container which
covers the top of the digester, excluding air and collecting the gas produced (Amigum and Von
Blottnitz, 2007). For industrial scale plants or ponding digesters the gas storage component is
appropriately designed for the application. The simplest reactor type consists of a single digester,
operated as a fed-batch or continuously stirred tank reactor. All the processes associated with
anaerobic digestion occur within the single unit, which may lead to process inefficiency, given the
differences in metabolism of the microbes responsible for the different phases of digestion. The typical
influent substrate concentration is in the range of 3 to 8% total solids (Gunaseelan, 1997), although
the solids loading can be as high as 25%. Retention times within these reactors are typically in excess
of six days. To overcome the problem of biomass washout while reducing the residence time, a
number of reactor types have been developed including anaerobic filter or fluidised bed systems
(where biomass is encouraged to attach to solid support material) and the upflow anaerobic sludge
blanket (UASB) reactor (certain microorganisms induce floc formation thereby retaining biomass in
the absence of a solid support). The use of two-stage digesters, that physically separate the
hydrolysis, acidogensis and acetogenesis stage form the methanogenic stage (Ghosh et al., 1975;
Gunaseelan, 1997; Niishio et al., 2007), often result in significantly reduced total retention times and
increased biogas production as the conditions for both sets of organisms can be independently
optimised.
There are several sources of waste with different characteristics: municipal (sewage), agriculture
(waste from animals, crop residues, and food processing) and industries (paper and textiles,
petrochemicals). Anaerobic digesters have been successfully employed for the generation of energy
from a wide range of these wastes. The table below illustrates the range of substrates that have been
tested at laboratory or pilot scale. Typical biogas yields from the digestion of primary sewage sludge in
an egg-shaped digester, common on sewage treatment plants, range from 0.4-0.9 l/g organic dry
substrate.
The yields are typically expressed in terms of gas production per unit of COD or volatile solid (VS),
3
and are in the order of 0.2 to 0.45 m /kg VS (Gunaseelan, 1997). The energy yield from methane is

calculated based from a value of 890.3 kJ per mole of methane (or 11.04 kWh/m3 of methane) (Duerr
et al., 2007).

Table 1: Methane and hydrogen production potential from agricultural wastes utilising a range of anaerobic digestion
technologies
Feed

Temp.
(C)
55

HRT (days)
-

Organic
loading rate
29 g-wet wt/L.d

50 (H2)
37 (CH4)

Fruit and vegetable waste

47

1.06 g VS/L.d

Potential energy
recovery
3
145 m H2/d (214
kwh)
3
514 m CH4/d
2.2 l H2/L (24 kJ/L)
2.2 l CH4/L (79
kJ/L)
0.16 l CH4/g VS

Fruit and vegetable waste

32

0.9 g VS/L.d

0.26 l CH4/g VS

Bouallagui et al., 2005

Fruit and vegetable waste

20

1.6 g VS/L.d

0.47 l CH4/g VS

Bouallagui et al., 2005

Fruit and vegetable waste

23

3.6 g VS/L.d

0.37 l CH4/g VS

Bouallagui et al., 2005

Fruit and vegetable waste

20

2.8 g VS/L.d

0.45 l CH4/g VS

Bouallagui et al., 2005

Bread waste (100 g/L) +


SS (10% w/v)
Beer waste (pressed
filtrate from lauter tun)

Reference
Nishio & Nakashimada,
2007
Nishio & Nakashimada,
2007
Bouallagui et al., 2005

Fruit and vegetable waste

2.5

6.8 g VS/L.d

0.35 l CH4/g VS

Bouallagui et al., 2005

Fruit and vegetable waste

7 + 10

4.4 g VS/L.d

0.34 l CH4/g VS

Bouallagui et al., 2005

Fruit and vegetable waste

2 + 2.3

5.65 g VS/L.d

0.42 l CH4/g VS

Bouallagui et al., 2005

Brewery slurry

55

26

3.23 g COD/L.d

0.37 l CH4/g COD

Zupancic et al., 2007

Brewery slurry

55

17.5

5.40 g COD/L.d

0.38 l CH4/g COD

Zupancic et al., 2007

Brewery slurry

55

16

6.27 g COD/L.d

0.42 l CH4/g COD

Zupancic et al., 2007

Brewery slurry

55

13.5

8.57 g COD/L.d

0.37 l CH4/g COD

Zupancic et al., 2007

Artificial garbage slurry

55

0.5 + 0.54

75 g/L.d

199 mmol H2/L.d


442 mmol CH4/L.d

Ueno et al., 2007

Fruit and veg waste +


manure

35

21

3.19 g VS/L.d

0.45 l CH4/g VS

Callaghan et al., 2002

Barley waste + SS

37

0.025 l CH4/g VS

Neves et al., 2006

Hydrolysed barley waste


+ SS
Cheese whey

37

0.22 l CH4/g VS

Neves et al., 2006

37

1+4

19.78 g COD/L.d

0.30 l CH4/g COD

Saddoud et al., 2007

Olive mill water + solids

55

36

3.62 g COD/L.d

46.2 l CH4/L
OMW.d

Fezzani & Ben Cheikh,


2007

Olive mill water + solids

55

24

5.58 g COD/L.d

38.2 l CH4/L
OMW.d

Fezzani & Ben Cheikh,


2007

Olive mill water + solids

55

24

1.79 g COD/L.d

29.6 l CH4/L
OMW.d

Fezzani & Ben Cheikh,


2007

Distillery effluent

35

14

2.74 kg/Cu m

4.44 l CH4/L

Banerjee & Biswas,


2004

Distillery effluent

55

14

2.74 kg/Cu m

5.40 l CH4/L

Banerjee & Biswas,


2004

Molasses

35

7 g COD/L

0.23 l CH4/g COD

Jimnez et al., 2004

Fermented molasses

35

7 g COD/L

0.31 l CH4/g COD

Jimnez et al., 2004

Algal sludge

35

10

4 g VS/L.day

0.14 l CH4/g VS

Yen & Brune, 2007

Algal sludge + wastepaper

35

10

4 g VS/L.day

0.29 l CH4/g VS

Yen & Brune, 2007

40 g COD/L

0.24 l H2/g COD

12

33 g VS/L

0.40 l H2/g VS

Vijayaraghavan et al.,
2006
Vijayaraghavan et al.,
2006

Palm oil mill effluent


Jackfruit peel

1.1.2

Advantages and constraints associated with AD

Anaerobic digestion offers a number of significant advantages over competing technologies for the
treatment of carbonaceous municipal, industrial and agricultural effluents. These include low energy
requirements, reduced sludge production and the economic recovery of energy. Thermophilic1
anaerobic digestion has the additional benefits of increased waste stabilisation and sludge
dewatering. This is particularly relevant to the treatment of sewage sludge, where the elevated
temperature leads to increased destruction of viral and bacterial pathogens (Chen et al., 2007 and the
study comparing combustion with AD and enhanced AD summarised in the K5/1732 Essence report),
but may incur large energy input costs.
In terms of constraints on the feasibility of anaerobic digestion technology, inhibition of the component
microbial processes and the resulting decrease in process efficiency can be a major concern.
Common reasons for poor performance are mixing problems, or the inability to maintain the correct
balance between acid forming and methane forming microbial communities. These two microbial
groups differ widely in their physiology, nutritional requirements, growth kinetics and susceptibility to
environmental perturbations (Chen et al., 2007). Inhibitory substances may be present at high
concentrations in wastewaters and sludges, or may be formed during the initial phases of anaerobic
digestion. Inhibition is typically identified by a decrease in steady state methane production and an
accumulation of organic acids. The most significant inhibitors are discussed briefly below.
Ammonia. Ammonia is produced by the biological degradation of nitrogen-containing compounds,
most commonly urea and proteins (Kayhanian, 1999). Its aqueous speciation is pH dependent, with
+
the ammonium ion (NH4 ) dominating at low pH and free ammonia (NH3) under alkaline conditions.

The free form is freely membrane permeable and has been suggested to be the primary source of
inhibition, resulting in a proton imbalance or potassium deficiency (Gallert et al., 1998). Of the
organisms involved in anaerobic digestion, the methanogens are most susceptible to inhibition by
ammonia (Kayhanian, 1994). The extent of inhibition is mediated by a number of factors, such as pH,
temperature, the presence of antagonistic cations (Na+, Ca2+, Mg2+) and acclimation, to the extent that
literature values for 50% reduction in methane production vary from 1.7-14.0 g/L total ammonia
nitrogen (Chen et al., 2007).
Sulphide. Sulphate is a common constituent of many industrial and agricultural wastewaters. Under
anaerobic conditions it is reduced to sulphide by sulphate reducing bacteria (SRB) (Hilton and
Oleszkiewicz, 1988). Inhibition occurs at two levels, initially through competition for substrate and
secondarily as a result of sulphide toxicity. SRB are not able to utilise complex organic carbon sources
so do not compete with hydrolytic or acidogenic microbes. Competition between SRB and acetogenic
and methanogenic bacteria has been observed, but results are often contradictory. The extent of
1

Thermophilic microorganisms grow optimally at temperatures above 40C. Mesophilic microorganisms grow optimally at

temperatures between 25 and 40C; and psychrophilic microorganisms grow optimally below 5C.

competition appears related to the initial microbial concentrations and the COD/SO42- ratio, with SRB
dominating at COD/SO42- ratios below 1.7, MPB above 2.7 and active competition occurring between
the two (Choi and Rim, 1991). Temperature has also been shown to have an effect with SRB more
likely to dominate at 37C and MPB at 55C (Colleran and Pender, 2002). With respect to sulphide
toxicity there are considerable discrepancies in the published literature with respect to the nature of
sulphide toxicity and the relative toxicity of dissociated and undissociated sulphide. Mechanistically,
sulphide inhibits cellular activity by denaturing native proteins, interfering with coenzyme sulphide
linkages and affecting assimilatory sulphide metabolism (Speece 1983; Vogels et al., 1988).
Unionised sulphide appears to play the dominant roles between pH 6.4 and 7.2, with total sulphide
concentrations becoming more important above pH 7.8. Fermentative microbes are less susceptible
to sulphide toxicity than acetogenic or SRB, with the MPB being the most susceptible (McCartnery
and Oleszkiewicz, 1991). For MPB the published IC50 values range between 50 and 250 mg/L,
depending on pH (Parkin et al., 1983; Oleszkiewicz et al., 1989; McCartnery and Oleszkiewicz, 1993).
Sulphide removal and acclimation of MPB have been shown to improve AD performance.
Light metal ions. The effect of aluminium, calcium, magnesium, potassium and sodium on
methanogenesis has been extensively studied (Chen et al., 2007). These ions may be present in the
wastewater, be released by digestion of organic matter or be added for pH control. In all cases
moderate amounts of the ions are required for growth, but excessive amounts are inhibitory.
Excessive amounts of Ca and Mg affect the system primarily through the precipitation of carbonate
and phosphate, resulting in the loss of essential nutrients and buffering capacity as well as scaling of
the microbes (Keenan et al., 1993; Van Langerak et al., 1998). Potassium and sodium are more
acutely toxic, given their role in maintaining membrane potential among other functions (Jarrell et al.,
1984; Soto et al., 1991). The IC50 values reported in literature differ significantly due to matrix
composition and degree of acclimation, but range between 0.15-0.74 m for potassium (Mouneimne et
al., 2003) and 5.6-53 g/L for sodium (Soto et al., 1993; Kim et al., 2000; Vallero et al., 2003a, b).
Heavy metal ions. Heavy metals may be present in significant concentrations in municipal sewage
and sludge, as well as a number of industrial effluents. While these elements form essential
components of enzymes that drive many anaerobic reactions they become acutely toxic at elevated
concentrations (Sterrit and Lester, 1980). Significantly, they are not biodegradable and although
present below toxic concentrations may be accumulated to inhibitory levels within the cells over time.
The toxicity of heavy metals is affected by a number of factors such as form of the metal, pH and
redox potential, but once again the methanogenic microbes are more susceptible than acetogens
(Zayed and Winter, 2000). Published IC50 values vary considerably and appear particularly affected by
solids content, which has a mitigating effect. However, the relative sensitivity of acidogenesis and
methanogensis to heavy metals is Cu>Zn>Cr>Cd>Ni>Pb and Cd>Cu>Cr>Zn>Pb>Ni respectively (Lin,
1992; 1993).

Organics. A wide range of organic compounds have been found to be inhibitory to anaerobic
processes. These are predominantly non-polar compounds which accumulate in bacterial
membranes, resulting in swelling of the membranes, leakage of cellular components, disruption of
ionic gradients and eventually cell lysis (Sikkema et al., 1994; Chen et al., 2007). The inhibition
concentrations vary for specific compounds and are affected by the concentration of the compound,
biomass concentration, exposure time, cell age, acclimation and temperature. Some of the major
classes of inhibitory organic compounds are chlorophenols, halogenated aliphatics, N-substituted
aromiatics, long chain fatty acids and lignin and lignin related compounds (Chen et al., 2007). A
number of these groups form important fractions of wastewaters and can have a significant impact on
efficiency of biogas formation.
1.1.3

Application and practice of AD technology

The application of anaerobic digestion technology appears to fall into two distinct categories, based
largely on the development status of the location. In developing countries, particularly in Africa and
Asia, the technology is applied at small or institutional scale and the biogas generated is used
primarily for heating or cooking, particularly in areas not serviced by the electrical grid. In these
situations the level of technology employed is kept low, such that the plant is robust and does not
require careful monitoring or technically skilled maintenance staff. Ideally the construction of such
units can be achieved by relatively unskilled labour.
In developed countries the installations tend to larger and employ more sophisticated technology. The
drivers for implementing such technology are not the provision of basic services to isolated
communities, but rather generation of renewable, environmentally benign energy. Developed
countries, particularly in Europe, often have an existing natural gas infrastructure for heating,
electricity generation and motor vehicle fuel. For boilers and stationary combined heat and power
(CHP) engines the biogas can be utilised with little or no modification. However, prior to utilisation as
motor vehicle fuel or incorporation into the natural gas grid the biogas needs to be purified to a high (>
95%) methane content. This process is uncommon in developing countries.
Based on the distinctly different technologies and drivers, it is not possible to apply uniform criteria to
assess changes in cost and efficiency with changes in scale. However, a recent study by Amigun and
Von Blottnitz (2007) investigated scale economies of 21 biogas installations across the African
3
continent. The installations varied in capacity from four to 100 m . Plant capital costs were converted

from local currency to US$ equivalents and a statistical assessment performed, correlating the plant
capacity to capital cost. The relationship between plant capacity and installation cost is not
proportional and has been represented by the empirical expression:
C = kQn
where k is the dimensioned proportionality factor, C is the cost, Q the capacity and the exponent n the
cost capacity factor. For chemical plants an n value of 0.6 is typical, although this has been shown to
increase to 0.85 where the process involves significant solids handling. The study found a statistically

significant cost capacity factor of 1.20 for small scale biogas installations (Amigun and Von Blottnitz,
2007). This is a significant finding and implies that rather than reducing capital costs (6/10 rule) scale
up of small scale biogas installations significantly increases capital investment costs. The outcome of
this study contrasts with a similar study performed in India by Singh and Sooch (2004). They
investigated three different models of small scale biogas plants and concluded that installation and
operating cost increased proportionally with capacity, implying an n value of 1. However, the capacity
of the plants used in their study did not exceed 6 m3. In addition, each model was assessed
individually as the construction costs varied. Despite these discrepancies, both studies indicate that
small scale biogas installation deviated significantly from the 6/10 rule.
1.1.4

Implementation of AD

The application of biogas, generated by the anaerobic digestion of organic material, for heating,
cooking and energy generation has been widely employed for many years. China began a program of
mass adoption of household biogas in 1975 and within a few years units were being constructed at a
rate of 1.6 million per year. The units were poorly designed and of low quality and by the 1980s many
of these were no longer in use. By 1992, only 5 million units remained operational and many of these
had been redesigned due to leakage. This pattern of rapid introduction and failure of biogas units was
repeated in India, Nepal, Vietnam and Sri Lanka. However, as technology and research improved the
units became cheaper and more robust, particularly following the development of polyethylene
digesters (Ho, 2005). There are currently over 2 million family sized units in operation in India, and
over 200 000 families a year are switching from the traditional fireplace to biogas for cooking and
heating (Singh and Sooch, 2004; Ho, 2005).
Biogas utilisation in Nepal is increasing rapidly and the country has recently overtaken India and
China as the nation with the highest number of biogas plants per capita. Its program is regarded as a
model for successful application of alternative energy in the developing world. There are currently
over 120 000 biogas plants in Nepal, with the Biogas Support Program (BSP) targeting in excess of
300 000 plants by 2009 (Acharya et al., 2005). The BSP is facilitating this by ensuring uniform plant
design and quality, continued R & D and importantly through subsidisation (up to 50% of capital cost)
and stimulation of financial support mechanisms (Acharya et al., 2005).
The production of biogas using AD in a decentralised WWTP in South Africa is explored further as a
case study in the current project. (This will be described in the final report).
In developed countries, on-farm and centralised rural biogas facilities are still significant, but the
majority of biogas generation is performed at a larger scale and is driven by environmental and
economic concerns and the burgeoning carbon economy. Many of the applications of the biogas
generated in this manner require improvement of the biogas quality. The composition of biogas, the
applications and the techniques employed for upgrading biogas in the developed world are discussed
below. The biogas produced by anaerobic digestion is composed primarily of methane (CH4) with

significant amounts of carbon dioxide and nitrogen and a number of constituents at trace
concentrations. The typical biogas composition is shown in Table 2:
Table 2: Typical biogas composition
Component

Typical range (%)

Methane (CH4)

50-80

Carbon dioxide (CO2)

25-50

Nitrogen (N2)

0-10

Hydrogen sulphide (H2S)

0-3

Hydrogen (H2)

0-1

Oxygen (O2)

0-2

Ammonia (NH3)

0-3

Siloxanes

Trace

Methane, also known as natural gas, has a heat of combustion of 802 kJ.mol-1. This value is the
lowest of the hydrocarbon fuels, but on a per unit mass basis methane is more attractive than
complex hydrocarbons. Methane is considered to have an energy content of 39 MJ.m-3. Based on the
typical range of methane concentrations in biogas, a heating value of 18.6-26.04 MJ.m-3 is standard
(Amigun and Von Blottnitz, 2007). The characteristics of biogas are somewhere between natural gas
and town gas (derived from cracking of cokes). The Wobbe index is a measure of the
interchangeability of fuel gases and is defined as the higher heating value over the square root of the
specific gravity (Table 3).
Table 3: Characteristics of different fuel gases
Parameter

Unit

Natural gas

Town gas

Biogas
(70% CH4, 30% CO2)

Calorific value (lower)


Density

kg.m

Wobbe index (lower)

Max CO2-conc in stack gas


Dew point

36.14

16.1

21.48

0.82

0.51

1.21

-3

MJ.m

39.9

22.5

19.5

-1

0.39

0.70

0.25

m air/m gas

9.53

3.83

5.71

vol %

11.9

13.1

17.8

59

60

60-160

Max. ignition value


Theor. Air requirement

-3

-3

MJ.m

m.s
3

Biogas can be applied to all applications designed for natural gas, although for certain applications
the biogas needs to be upgraded to remove non-methane components. Table 4 indicates the gaseous
components which need to be removed in order to upgrade biogas for particular applications.

Table 4: Gaseous components requiring removal prior to application of biogas


Application
Gas heater (boiler)
Kitchen stove

H2S

CO2

H2O

< 1000 ppm

no

no

yes

no

no

< 1000 ppm

no

no condensation

Vehicle fuel

yes

recommended

yes

Natural gas grid

yes

yes

yes

Stationary engine (CHP)

Gas boilers do not require high quality gas, although it is recommended that H2S be removed to below
1000 ppm in order to maintain the dew point around 150C. Where significant amounts of H2S are
present, the condensate will contain sulphurous acid which can corrode chimneys and heat
exchangers, particularly if they are composed of mild materials. Biogas has been used as a fuel for
combined heat and power (CHP) engines for many years, typically at sewage works, landfill sites and
biogas installations. Engine sizes range from 45 kW to several MW depending on the size of the
installation. The gas quality requirements are similar to those for a gas boiler, with the exception that
H2S concentration should be reduced to guarantee a reasonable operating life for the engine. Petrol
engines are more susceptible to H2S corrosion than diesel engines. As a result diesel engines are
almost exclusively used for large scale (> 60 kWel) applications.
The development of modern gas turbines for electricity generation is a significant new development
as these robust engines have comparable efficiency to internal combustion engines. The turbines
facilitate heat recovery in the form of steam. Efficient small-scale gas turbines have not been
developed, with the majority of engines being larger than 800 kWel.
The operation of motor vehicles on natural gas is a growing worldwide phenomenon. There are
currently in excess of 1 million gas powered vehicles worldwide. Biogas has been used to power
these vehicles, but the gas quality demands are high. Therefore, purification of the biogas is required
to increase the calorific value, remove corrosive components (H2S, NH3 and H2O) and particulates,
specifically siloxanes (Dewil et al., 2006). The upgraded biogas has a methane content in excess of
95%. A number of European studies have confirmed that upgraded biogas rated as the cleanest fuel
with respect to impact on the environment, climate change and human health.
A wide variety of technologies exist for the upgrading of biogas to vehicle fuel standards. These are
summarised in Table 5, below.

Table 5: Summary of current technologies for upgrading biogas


Component removed
Carbon dioxide

Technology
Water scrubbing
Polyethylene glycol
scrubbing
(Selexol)
Carbon molecular
sieves
Membranes
systems

Hydrogen sulphide

Biodesulphurisation
Biological filters
FeCl3 dosing to
digester slurry
Iron oxide

Siloxanes
Oxygen and nitrogen

Impregnated
activated carbon
Solvent scrubbing
Solvent extraction
Membranes or
molecular sieves

Notes
Counter-current packed bed, can remove some H2S. Potable water not
needed.
Similar to water scrubbing, but more efficient reduced solvent demand and
0
pumping costs. Solvent stripped with steam or inert gas to minimise S
formation.
Loose adsorption of molecules in pores selectivity by pore size and gas
pressure differences. Typically 4 units in series. Removes CO2 and water
vapour.
Two main technologies: High pressure (gas-gas). Operated at 36 bar. Good
removal of CO2, H2S. Less efficient N2 removal. Membrane life 3 yrs. Hollow
fibre membranes allow compact units.
Low pressure (gas-liquid). Uses microporous hydrophopic membrane at 1
bar. Liquid adsorbent NaOH or amine. Efficient removal of H2S and CO2, both
0
can be recovered as saleable products (S and industrial CO2)
0

Typically done at digester stage. Use Acidithiobacillus sp. to form S and


2some SO4 . Stoichiomteric O2 dosing critical.
Combined with water scrubbing in large applications. Provides increased
surface area for reaction. 4-6% air added to biogas to facilitate oxidation.
Precipitation of insoluble iron sulphide. Reduces high H2S levels, but further
reduction required for vehicle fuel quality.
H2S reacts with iron oxides or hydroxides to form iron sulphides. Endothermic
reaction so TC > 12C required. Iron oxide regenerated by reoxidation,
0
yielding S . Regeneration in separate bed. Iron oxide surface area increased
by coating wood chips or using pellets.
In association with CO2 removal by molecular sieves. H2S oxidation catalysed
by impregnating AC with potassium iodide. Best operation at 7-8 bar, 50-70C.
As described above for CO2, using water, Selexol or NaOH.
Siloxanes (organic Si compounds) are oxidised to SiO2 in combustion
engines. Can cause major damage and significantly reduce engine life.
Removed by absorption into specific hydrocarbon mixture.
Introduced form air through inefficient harvesting of biogas. High O2 increases
explosion risk. Removal is expensive prevention by efficient biogas
collection advocated.

Biogas utilisation in EU countries is extensive, as can be seen Table 6. However, the majority is still
derived from landfill sites rather than liquid wastes or sewage sludge. A relatively small portion of the
gas is currently used in electricity generation.

10

Table 6: European Union energy from biogas data for 2006. Data includes substrate source for anaerobic digestion
and the proportion of the energy applied to electricity generation (adapted from Biogas barometer, 2007).
Country

Total biogas

% for electricity

Biogas per AD substrate source (GWh)

(GWh)

Landfill

Sludge

Other

Germany

22 370

33

6 670

4 300

11 400

UK

19 720

25

17 620

2 100

Italy

4 110

30

3 610

10

490

Spain

3 890

17

2 930

660

300

France

2 640

19

1 720

870

50

Netherlands

1 380

21

450

590

340

Austria

1 370

30

130

40

1 200

Denmark

1 100

26

170

270

660

Poland

1 090

26

320

770

Belgium

970

22

590

290

90

Greece

810

71

630

180

Finland

740

590

150

Czech Republic

700

25

300

360

40

Ireland

400

27

290

60

50

Sweden

390

14

130

250

10

Hungary

120

18

90

40

Portugal

110

30

110

Luxembourg

100

33

100

Slovenia

100

32

80

10

10

Slovakia

60

50

10

Estonia

10

70

10

Malta

In many European countries where intensive agriculture, particularly animal farming, is practiced
eutrophication of aquatic environments due to influx of animal waste became a significant
environmental issue. In Sweden, during the 1990s, there was a move towards regional biogas plants
to deal with this issue. Two examples are in operation in the cities of Laholm and Linkping. The
inputs and outputs to these plants are shown below:
Table 7: Input and output data from sample Swedish biogas plants, 2004 data (IEA Bioenergy, 2007)
Input (tonnes/year)

Laholm plant

Linkping plant

Manure from pigs and cattle

28 000

2 000

Abattoir waste

10 000

30 000

Industrial organic waste

3 000

6 000

Household waste

1 000

250

Other

6 000

7 000

Total

48 000

45 000

28 000

52 000

Output (tonnes/year)
Biofertiliser for farming
Other

15 000

Total

43 000

52 000

Both plants are operated as conventional stirred tank reactors (2 250 m3 and 3 700 m3 respectively)
with residence times of 25-30 days. In both cases the biogas is upgraded prior to use. In the case of

11

Laholm the upgraded biogas is supplemented with 5-10% propane to increase the Wobbe number to
that of natural gas. A portion (25% annual average) is directed to the district heating plant, while the
remainder (18 000 MWh/year) is added to the natural gas grid. The Linkping plant produces an
equivalent of 48 000 MWh/year of biogas, over 95% of which is upgraded to vehicle fuel quality. Since
2005 all public transport vehicles in the city (> 60 buses) have been converted to run on biomethane
and in 2005 the first biomethane powered train was unveiled. This has made it possible to reduce CO2
emissions from urban transport by over 9 000 tonnes per year, while reducing local emissions of dust,
sulphur and nitrous oxides (IEA Bioenergy, 2007).
Management of bioenergy facilities is an important factor in ensuring efficient energy recovery. A five
and a half year study on a bioelectricity plant in Greece, generating energy from biogas derived from
sewage sludge digestion, highlighted this issue. Technical issues and the fact that plant operation was
synchronised with the local supply network resulted in the unit only covering about 16% of the energy
requirements of the facility, despite having the capacity to cover almost 40%. Even at a highly
subsidised electricity price the payback on capital costs has been estimated at 16 years, based on
outputs of the last 2 000 days (Tsagarakis and Papadogiannis, 2006).
In South Africa the first biogas to electricity plant, funded by carbon credits generated under the Clean
Development Mechanism (CDM) of the Kyoto Protocol, was launched in October 2007. The plant is
situated near Mossel Bay and utilises process wastewater generated during the operation of
PetroSAs gas to liquid plant at Duinzicht as the substrate for anaerobic digestion. Historically, the
biogas generated was flared and has led to an estimated loss of at least 1 300 GWh of gross heat
value since the commissioning of the plant. The plant was developed and financed by MethCap, a
company owned by the international environmental and engineering firm WSP, and will make use of
three GE Jenbacher biogas generator sets, each with an output of 1.4 MW, giving the plant a total
output of 4.2 MW. The plant is expected to generate 33 000 certified emissions reductions (CEMs)
annually.
1.1.5

Costs and Benefits

3
Capital costs for the construction of small scale (6 m ) biogas units have been shown to vary from

around US$300 to US$900, depending on location, plant design and construction material (Amigun
and Von Blottnitz, 2007; Singh and Sooch, 2004). The cost capacity factor governing increase in
construction cost appears to be regionally dependent, with a value of 1.2 being calculated in Africa,
but only 1.0 in Southeast Asia. Annual operating costs for such digesters are in the range of US$230,
with the significant majority being associated with the acquisition of substrate manure. This is
consistent with European studies (IEA Bioenergy, 2007) which highlight acquisition and transport of
substrate as the major operating expense.
The anaerobic digestion of organic wastes has a number of associated benefits, in addition to the
treatment of a waste source and the generation of energy. Methane is a significant greenhouse gas

12

with a global warming impact considerably greater than that of carbon dioxide according to the IPCC
(Ishikawa et al., 2006). Waste organic material degrades naturally, often under anaerobic conditions,
with the associated release of methane. The controlled digestion of this material with the collection
and utilisation of the biogas significantly reduces uncontrolled methane discharge with the associated
reduction of the greenhouse effect.
Upon completion of the anaerobic digestion process a number of by-products remain. The liquid
effluent resulting from the process typically requires an additional aerobic treatment prior to discharge,
in order to reduce COD and BOD to acceptable levels. In addition to the wastewater a solid residue,
termed digestate, remains. The composition of the digestate depends on the initial feedstock and
influences its potential applications (Singh and Sooch, 2004; Ho, 2005).The typical composition is
predominantly lignin, cellulose, biomass sludge and inorganic components, including ammonia,
nitrates and phosphates. In the absence of pathogens or potentially toxic elements the digestate
makes an ideal biofertiliser, often after a period of aerobic treatment to degrade lignin and oxidise
ammonia to nitrate. If the digestate is not suitable for fertiliser it can be incorporated into building
materials such as bricks or fibreboard (IEA Bioenergy, 2007).
As noted above, biogas is a particularly clean burning fuel. When used as a cooking and heating fuel
in rural households this results in associated health benefits, largely due to the reduction in particulate
material. A survey conducted by the BSP in Nepal indicated a significant reduction in occurrences of
respiratory problems, asthma, eye infections and chronic headaches among respondents who had
switched to biogas from dirtier fuels. Additional benefits were derived from no longer having to collect
and carry firewood (Acharya et al., 2005).
Table 8: Total suspended particle concentrations in flue gas of common cooking fuels (Acharya et al., 2005).
3

Fuel

Total suspended particles (mg/m )

Biogas

0.25

LPG

0.32

Kerosene

0.48

Crop residue

5.74

An additional monetary benefit derived from the combustion of biogas exists in the form of carbon
credits. For example, each of the functioning digesters in Nepal prevents five tonnes of carbon dioxide
equivalents from being pumped into the atmosphere annually. These credits can be sold to
developed countries to offset their excessive emissions and are worth in excess of US$ 5 million.
Despite the encouraging models, current anaerobic digestion technology is not yet sufficiently efficient
to compete with fossil fuel technologies on a purely economic level. For this to occur significant
improvements are required in process technology, or additional benefits such as waste reduction or
greenhouse gas mitigation need to be taken into consideration.

13

1.2

Combustion and gasification for the production of steam

1.2.1

Introduction

Combustion:
The direct combustion of biomass in the presence of excess air results in the formation of hot flue
gases that are typically used to produce steam to drive the electric turbine, with an efficiency of 33 to
45% (Bain and Craig, 1988). Two approaches for combustion are presented: mono-incineration,
where the sewage or biomass sludge is the only energy source and co-combustion, where these are
combusted in the presence of conventional fuels such as coal. In the combustion of sludges, the first
step focuses on dewatering and thermal decomposition with release of volatiles (devolatilisation or
pyrolysis) resulting in a gaseous stream containing H2 (2 to 5%), CO2 (7 to 24%), CO (28 to 66%) and
hydrocarbons (16 to 33%). As the operating temperature increases, the CO and H2 components
increase, and are subsequently oxidised in the second stage to carbon dioxide and water (Werther
and Ogada, 1999).
Gasification:
Gasification is a thermal process yielding a combustible gas (or producer gas or fuel gas) as a
product. Gasification occurs in an atmosphere of sub-stoichiometric oxygen. Biomass particles are
heated up and release their volatile components in a process termed pyrolysis or devolatilisation and
when the oxygen concentration is limited, syngas rich in CO and H2 is produced.
Either sequentially or simultaneously, depending on the heating rate and the supply of oxygen, the
carbon skeleton is consumed via combustion and gasification reactions leaving residual ash and tar.
The reactions involved and the heat enthalpy (H0,rxn) are shown in the reaction scheme below (2):

Figure 2. Reaction scheme for the complete combustion of biomass.

The gas produced is characterised as having a low (5 MJ/Nm3) to medium (15 MJ/Nm3) energy
content, depending on whether air or steam is used as the gasifying agent (Scott, 2004; Prins et al.,
2007; McKendry, 2002). Bubbling bed or fluidised-bed gasifiers and fixed bed downdraft gasifiers are
two types which have found application in small scale production. They display good mixing properties

14

and can effectively crack tars if the bed is catalytically active (Min, 2005).
Table 9: Estimation of biomass resources in SA for conversion to energy. The amounts of biomass, coal fines or
sewage required for a medium size power plant (2 mW electric output or 5 mW heat output) is tabulated.

a) Total biomass available = agriculture residues + forestry residues + 5% of available land dedicated to energy crops (invasive
plant excluded)
b) Total RSA coal production for 2003: 238 million tonnes (www.bullion.org.za).
c) Waste: 0.1 kg/person/day dry weight. South Africa population: 48.5 million.
d) Higher heating values or HHV does not account for the formation of water vapour in the flue gas stream. Hence the need for
adjusting the calc according to the moisture content of the feedstock
e) Adapted from reference (Kavalov and Peteves, 2005)
f) Coal fines moisture content as fired: 64%; Coal fines ash content: 47%; Coal fines HHV: 16 mJ/kg
g) Assuming a thermal efficiency of 40%

1.2.2

Application of combustion / gasification

There are limited examples of this application in South Africa (some application has been reported for
the fruit industry e.g. juicers in Ceres and Grabouw, and for bagasse in the sugar industry). It is widely
used in Europe in order to meet the regulations on material sent to landfill (Lurgi and Philip Conoco
are main designers and suppliers of incinerators). BRI Energy (USA) has developed a unique process
in which a range of waste materials undergo thermal gasification at high temperature into CO, CO2
and H2 (http://techmonitor.net/techmon/05jul_aug/nce/nce_waste.htm).
1.2.3

Costs and benefits of combustion / gasification

When considering the potential for combustion or gasification of biomass from wastewater, it is
essential to consider the potential operational problems of this approach. Specifically, moisture
content, tar formation, mineral content and bed agglomeration are highlighted. The feedstock to a
gasifier has to be relatively dry, with a moisture content of 50 to 60%. Several gasifier designs have
been proposed to include a dryer upstream of the combustor, but there is a clear trade off between
the amount of energy available in the feedstock to the amount of energy expended on drying. Tar
formation lowers the overall conversion of biomass to gas. To convert tars, either in-bed technology or
downstream reforming is required. Primary tars can be thermally cracked at 800C, while secondary

15

tars can be cracked at 1000C. In addition, the ash that remains may find application in the brickwork
industry or may need to be disposed of into the environment. There are several negative
environmental effects as a result of the gaseous and solid phase pollutants, (heavy metals, dioxin,
furans and N2O and NOx). However, evidence suggests that the heavy metals, dioxin and furan
emissions can be controlled and that NOx emissions remain below 5% while N2O and CO emissions
are prevented by high temperature and methods such as Activated Carbon Facilitated Oxidation (ACFOX) (Brain, 2007).

16

1.3 Utilisation of waste heat


1.3.1

Background

Many processes give rise to liquid streams with elevated temperatures. Depending on the source and
fate of these streams (i.e. waste or product streams), the need may exist for reducing the temperature
of the streams prior to discharge or utilisation. This cooling can be achieved either by dedicated
cooling systems or through heat exchange with other streams. The use of the energy inherent in
these streams in other applications which require heat (rather than cooling or discharging without
energy recovery) has the potential to offset energy demand for other heating applications from other
(non-renewable) resources. The process industry has had decades of experience with heat
integration and the literature, analytical tools and practice are well established in this area. The main
motivation has been financial, but there are also growing incentives to improve energy efficiency to
lower greenhouse gas emissions. Therefore, neither on-site process heat recovery nor energy
recovery from hot gaseous streams are considered in this review. There are opportunities to recover
waste heat from processes where onsite recovery is not already widely practiced. The main
motivations for exploring such opportunities are the potential financial savings, and environmental and
sustainability considerations associated with the supply of energy from fossil fuels.
Here we discuss the heat contained in wastewaters which are to be discharged from industrial
processes. This represents one element of overall process heat integration where waste process
heat, including that contained in process streams which are at a certain temperature and require
cooling prior to being fed into another part of the process, is used elsewhere in the process. The
distinction is made in the context of the current study which is focused on energy from wastewater, in
reality this cannot be considered in isolation. An understanding of fundamental thermodynamics
suggests that the recovery of waste heat can be implemented at any scale. Providing there is a heat
gradient between the two streams (i.e. one is hotter than the other), and that they are separated by a
good heat conductor, energy will flow from the hot stream to the cooler stream. It is noted that if the
temperature difference between the two streams is not great enough the heat transfer will be
ineffective. In addition, the heat capacities of the liquid affect their suitability for heat exchange.
Specific heat capacity (expressed in J/g.K) measures the number of joules of energy required to
change the temperature of one gram of the substance by one Kelvin.
The range of applications for waste heat are a function of the quality of the heat high grade heat has
a wider range of industrial applications than lower grade heat, although the latter can be used
applications such as pre-heating of feed streams. Thus finding the correct match between available
energy supply and energy requirements is essential.
Given that heat exchange is usually conducted over an equipment boundary (such as a heat
exchanger), the chemical composition of the wastewater is relatively unimportant, providing the
construction materials of the heat exchanger are suitably matched to the wastewater composition.

17

The chemical composition will, however, affect the specific heat capacity of the waste stream and this
needs to be matched with the rate of flow of wastewater so that optimal heat transfer down the
temperature gradient can occur. This application may, however, be inappropriate for safety reasons
when a wastewater stream is used directly in space heating there may be risks of circulating
hazardous wastewaters through offices and warehouses.
The theory surrounding heat integration is well established, with various optimisation tools such as
pinch analysis being available to design energy recovery systems to maximise the amount of energy
recovered. Integrated heat recovery usually simultaneously considers recovery from both liquid and
gaseous streams. Heat recovery may also be integrated with other system optimisations. For example
a WRC funded study by Fraser et al (2006) explored simultaneous optimisation of recovery of heat
and water in industrial processes using thermal pinch analysis and water pinch analysis. One of the
key findings from the study was the significant benefits in doing heat exchanger network design early
in the design process.
Despite the theory being relatively well established, it is identified that the practice of on-site heat
integration in industry is not as extensive as would be expected. Apart from the impact of fundamental
stream properties, and the requirement for finding matching heat streams, other factors which may
limit the realisation of opportunities for heat integration include ensuring sufficient energy availability
on a site, the fact that heat cannot be transported over long distances, and the availability of suitably
qualified personnel (preferably with postgraduate experience in the field) for opportunity identification,
communication, design and implementation. Anecdotal evidence suggests that the latter is a
significant barrier to implementation of opportunities (PI, 1999).
In addition to on-site waste heat recovery, with the emergence of the meta-discipline of Industrial
Ecology in recent years there has been a regional sharing of process heat, outside of specific plant
boundaries. These applications suffer the similar limitations to implementation described from on-site
usage. Some examples are, however, presented below.
1.3.2

Implementation of utilisation of waste heat

Although a priori planning for recovery of process heat is preferred, retrofitting of reticulation for use of
process heat is possible, depending on the particular application and plant configuration. Furthermore,
the financial viability of retrofitting heat integration is increased if an additional savings such as water
usage is simultaneously realised. Thus, if retrofitting is desired, its feasibility needs to be evaluated on
a case-by-case basis.
The level of monitoring and maintenance for applications of process heat are typically fairly low once
they have been optimised, although once again this will depend on the application. Many industrial
processes need a consistent temperature for operation, hence monitoring is required to ensure that
this is maintained by the supplied heat. At the same time, however, the heat source is likely to be

18

consistent (if it originates from another steady state process), and hence the need for constant reoptimisation of the system is likely to be limited.
A wide variety of examples of overall process heat integration can be found from around the world,
although as mentioned previously experience suggests that these are not exploited as extensively as
they could be. It is further noted that within the context of this study it is difficult to isolate examples of
energy recovery from wastewater rather than process streams. Furthermore, isolation of case studies
which focus on liquid process heat streams only is difficult as total plant heat optimisation focuses on
a combination of liquid and gaseous streams or gaseous streams alone.
Natural Resources Canada (2004) identifies various projects which may be suitable for application of
process integration, and classifies them on the basis of payback periods. Some examples of these are
shown in Table 10.
Table 10: Example projects suitable for heat integration (Natural Resources Canada, 2004)
Quick Wins

Operational improvements to
refrigeration systems, e.g.
adjusting
compressor
pressure
to
ambient
temperature;
Hot
water
management
improvements,
e.g.
balancing hot water supplies
with demands;
Enhanced
evaporator
performance, e.g. improved
non-condensable venting.

Short-to
medium-payback
projects,
typically with a one-to three-year payback

Instrumentation
modifications,
e.g.
improved control loops;

Optimisation of refrigeration systems,


e.g. increasing refrigerant pressures
within operational constraints;

Evaporator feed preheat by improving


heat recovery;

Rearranging existing heat exchangers to


make better use of them, and addition of
new heat exchangers;

Pasteuriser heat recovery, e.g. use of


heat storage;

Recovering steam condensate to boiler


house;

CIP water preheat by heat recovery;

Long-term payback projects, typically


with a three-to six-year payback:

Indirect heat recovery via a hot


water loop;

Kettle vapour heat recovery;

Cogeneration (CHP) via steam


turbine, gas turbine, or gas/diesel
engine;

Boiler economiser (payback will be


lower for large steam boilers);

Mechanical vapour recompression


(heat pumps);

Increased number of evaporator


effects.

The success of integrated applications for waste heat utilization has been varied since it depends
upon fostering collaborations for sharing energy and other resources. A good example of regional,
integrated waste heat utilisation is the Asns power station in Kalundborg, Denmark which was not
included in planning during construction of the power station (http://www.symbiosis. dk/, 2007). Asns
is Denmarks largest power station with an installed capacity of 1 037 MW. The excess heat from the
power station provides process steam for the local refinery, pharmaceutical and enzyme
manufacturing factories, and supplies heating for over four thousand households in the town. In
addition to energy integration, other benefits include reduced water usage, waste recovery, and
additional employment creation. At the Bruce Energy Centre in Canada some forward planning to
integrate energy usage was undertaken. The Centre is situated adjacent to a nuclear Ontario Power
station enabling a supply of steam and electricity at a low cost. The electricity and steam is used by
an alcohol distillation operation, a company which dehydrates crops to produce nutrient rich feeds for
livestock, a food manufacturer which concentrates fruit and vegetables, a plastic manufacturer and for
heating a greenhouse.

19

A number of further examples of the use of waste heat for district household heating in colder climates
can be found. The city of Gtheborg (Sweden) uses process heat from industries to supply a
proportion of the district heating system. The heat is from waste heat from oil refineries in the vicinity
of the city, and from a waste-fired CHP plant. Heat is also recovered via heat pumps that utilise
sewage water (Holmgren, 2007). Heat integration is also feasible for low-level industrial waste heat as
shown by a study in Delft (Netherlands). The heat generated by a pharmaceutical plant in the North of
the city at a temperature of between 25 and 35C is upgraded by heat pumps before being taken to
the central heating grid (Ajah et al., 2007). Similarly, low grade waste heat from power stations and
industrial applications has been used for water heating in aquaculture applications. The Asns trout
fish farm (Denmark) and the Happy Shrimp company (Netherlands) make use of waste heat from a
power plant to heat aquaculture ponds. Similar applications are seen the Czech Republic, Canada,
France, the US and others.
Other examples of applications include:

Various Exxon Mobils refineries have achieved significant savings through heat integration
(Exxon Mobil, 2007),

Through heat integration, a pulp and paperboard mill in Canada achieved savings on
purchased thermal energy of 15% and reduced the temperature of effluents by 3C, with a
simple payback period of less than 10 months for most projects (Natural Resources Canada,
2007).

The city of Vancouver has established heat recovery from municipal sewage. Raw sewage is
passed through a heat exchanger where thermal energy is captured. Electrically-powered
heat pumps are used to boost temperatures from the 10 to 20o C sewage temperatures, to
the 65 to 90o C range for the hot water distribution system. The heat pump will produce
roughly three units of heat energy for every one unit of electricity consumed an efficiency
rating of 300 per cent. Southeast False Creek will be the first sewage wastewater heat
recovery plant in North America, and could set a precedent for the development of this
technology

in

other

locations

within

Vancouver,

and

in

other

urban

areas.

(http://vancouver.ca/sustainability/documents/sefc-factsheetheatplant-back.pdf)
Heavy industry in South Africa lends itself favourably to heat integration within and between industrial
plants. However, the limitations presented above, particularly those surrounding human resource
capacity and integrated (pre-) planning have played a role in not realising further opportunities thus
far. It is further identified that the potential for use of low grade heat for district heating is limited due to
the milder weather than in many parts of Europe where such opportunities are realised.
1.3.3

Assessment of costs and benefits in utilisation of waste heat

The primary costs associated with heat integration relate to any additional piping, pumping, insulation
and equipment requirements for heat exchange. On a process plant these are likely lower than when
sharing of process heat is outside the plant boundaries. At the same time, use of process heat can

20

allow for the downsizing or elimination of the need for other energy sources such as boilers. The
financial benefit and payback periods on investment will vary widely depending on the situation. In
general a retrofit will be more expensive than incorporating integration into the original project design.
Operating expenditure will be relatively low, and will be limited to pumping costs and general
maintenance.
Unlike many of the other technologies presented in this report, use of process heat achieves no
wastewater treatment benefits (apart from removing the need for cooling prior to discharge which may
otherwise be required). Furthermore, there is no added benefit associated with the production of
secondary products or wastes which may require management elsewhere. Wastewater containing
pollutants from which heat energy is recovered will still require treatment prior to discharge.

21

1.4 Fermentation to produce biomass and secondary products


1.4.1

Background on fermentation of wastewater

Wastewater streams often contain organic compounds which may be useable as carbon and nutrient
sources for growth of microorganisms. Such compounds would be referred to as fermentable
substrates, and these fermentations can generate fuels such as bioethanol and biogas (see sections
1.1 and 1.6) directly, but also:

Serve as pre-treatments for energy from wastewater technologies (such as enhancing


bioethanol or biogas production from certain wastewaters).

Provide a nutrient medium to support the growth of biomass (microorganisms, algae or plant).
This biomass can subsequently be combusted or used in bioethanol, biogas or biodiesel
production technologies.

The applications of fermentation include treating wastewaters from olive-mills dairies, breweries, wine
distilleries, whiskey distilleries, slaughter-houses, potato waste, and the tomato canning industry.
Examples of South African wastewaters containing potentially fermentable substrates with estimates
of volume and load are shown in Table 11, below.
Table 11: Examples of South African wastewaters containing fermentable substrates
Wastewater

COD (g/L)

Sewage

0.8 -1.2
Av= 0.86
1.5-9.2
Av= 5.3
11-21
Av= 16
55 201
Av= 100
5-15
Av=10
25 45
Av=30
35
6
3
0.1 -2.5
Av=1
0.7
0.2-0.9
Av= 0.7

Dairy*
Red meat and poultry
abattoirs
Olive production
Fruit processing
Distilleries:
-Grain and grape
-Sugar cane molasses
Winery
Brewery
Textile Industry
Pulp and Paper
Petrochemical waste

VOLUME
(ML per year)
2 766 400

Load
(Mg/ year)
2 379 104

6346

33 637

11 000-31 000

336 000

89

8 900

14 000

140 000

Grain: 63
Grape: 342
3500-4000
1000
28 000
25 000

12 150
131 250
6 000
23 533
25 000

80 000
56 000
1140 (crude);
2 939
3048 (synthetic);
2 to 11(re-refinery)
*Only the formal dairy industry sector considered here. Other animal husbandry sectors (cattle for beef, pigs and chicken are
not shown (case studies in the Appendix for further details).

Important issues which need to be considered in developing energy-yielding bioconversions from


wastewaters include:

The volume, variability and/or characteristics of the wastewater from any one industrial
activity may be too low to provide suitable media for the fermentation. However, combining or
integrating of wastes may be feasible or necessary.

22

The concentration of carbonaceous components may be very low and need de-watering in
order for efficient fermentation to be achieved. This dewatering process will have its own
energy input requirement.

Many industrial wastes may contain other components which are inhibitory to microbial
growth. These may need to be removed by, for example, solvent extraction or precipitation
before the waster is used for fermentation. This extraction may yield economically valuable
secondary products (as in the case of olive wastewater).

1.4.2

Implementation of fermentation of wastewaters

The basic fermentation technologies and knowledge base are well established. The requirements for
sterility, cooling/ heat transfer and extent of aeration are important issues in the scale-up of
operations. Wastes would be introduced into the bioreactor or separation/extraction equipment and
products would undergo downstream processing that is separate from the main plant and therefore
there should be little problem with retrofitting into existing infrastructure. Bioconversion is a wellestablished laboratory technology that generally requires constant monitoring of several parameters
for the control and optimisation for a particular waste stream. Many parameters (e.g. mixing, pH, dO2,
nutrient levels) can readily be automated and controlled, but in some cases may require skilled
operators.
The most abundant agricultural wastes are rich in complex polymeric molecules such as cellulose that
require significant energy for breakdown. In nature, several hydrolytic and oxidative enzymes
produced by a variety of fungi and bacteria work in synergy to degrade cellulose, hemicellulose and
lignin (Gosh and Gosh, 1992). Fermentations can provide biological methods of pre-treatment where
the addition of water may be required for processing of the waste. The pre-treatment of corn stover
with fungi such as Cyathus stercoreus and Phanerochaete chrysosporium has been shown to
effectively improve the cellulose digestibility three- to five-fold (Keller et al 2003), thereby increasing
the yields in the subsequent ethanol fermentation. The fungus Trichoderma reesei was shown to
improve the acid hydrolysis of cotton waste with a subsequent increase in the yield from ethanol
fermentation (Rajesh et al., 2008). Similarly, the biological conversion of wheat and rice straw to
ethanol has been show to be feasible with a pre-treatment of Aspergillus niger or Aspergillus awamori
followed by standard Saccharomyces cerrevisaeia fermentation to yield 2.5 g/L ethanol (Seema et al.,
2007).
The two-stage or multi-stage anaerobic digester systems also consist of fermentation pre-treatments
(see section 1.1). The different digestion vessels are optimised to bring maximum control over the
microbial communities so that the initial stages of hydrolysis, acetogenesis and acidogenesis occur in
separate bioreactor(s) to the methanogenesis that produces methane-rich biogas. The two stage
system has been shown to increase efficiency by reducing the residence time while increasing the
yields of biogas attained for a range of organic substrates (Schfer et al., 1999; Gunaseelan, 1997).
This has been successfully demonstrated for municipal waste and represents 10% of the anaerobic

23

treatment capacity in Europe (de Baere 2000; Pavan et al., 2000). The two-stage or multi-stage
digestion systems have particular application to wastewaters rich in complex polysaccharides, lipids,
and proteins that are inaccessible for other energy from wastewater technologies such as bioethanol
production (Zeeman and Sanders 2001, Ahring 2001). For example, the biogas yields from celluloserich wastes can be greatly improved through pre-treatment using a twostage system (Gijzen et al.,
1998) and this has been successfully been applied to the water hyacinth, Eichhornia crassipes, that
often invades watercourses in South Africa (Kivaisi and Mtila, 1997). Similarly, the two stage system
has been shown to be effective for biogas production from the lipid and protein rich wastewaters of
slaughterhouses and fish processing factories (Saddoud and Sayadi 2007; Ahring 2001).
Fermentations have also been used for the production of other valuable secondary products. These
include the cultivation of biomass for animal feed (Ugwuanyi et al., 2007), growth of the valuable
biocontrol fungi, Trichoderma viride, (Verma et al., 2006), and the production of valuable secondary
metabolites such as antibiotics and carotenoids (Ho, 2005; and http://www.ctre.iastate.edu/research).
1.4.3

Assessment of cost and benefits in fermentation of wastewaters

An assessment of the costs and benefits is complex and specific to the waste stream and context.
Costs include the capital costs of the bioreactor and associated equipment, operational costs for
trained staff to maintain and optimise cultures, mixing, aeration (where required), sterilisation and
separation costs. The benefits include water bioremediation through removal of COD (waste disposal
reduction) the energy generated from wastewater and the value of any additional secondary products.
Fermentations can be used to directly produce energy products such as bioethanol and biogas
(discussed elsewhere section 1.1 and 1.2 and 1.6), as a pre-treatment step, or to generate biomass
from the nutrients in the waste stream. The generated biomass may later be used as an energy
resource for a secondary energy generating technology (i.e. combustion/gasification, bioethanol,
biogas or biodiesel production) and/or it may contain valuable secondary products (e.g. plant
fertilisers, animal feeds or neutraceuticals). The feasibility will depend on integration an assessment
of these complex factors, but will primarily depend on the characteristics of the feedstock and its
supply. In the chain of collection and transport, wastes often become complex wastewaters simply
because of dilution; thereby rendering the wastewaters unsuitable for generating energy using a
number technologies. Where this cannot be avoided, the use of dilute wastestreams to grow biomass
or the combining of waste streams may enable an energy from wastewater process to become
feasible.

24

1.5 Algal biomass for the production of biodiesel


1.5.1

Background on production of biodiesel from wastewaters

In order to utilise the dissolved solutes within wastewater for energy generation, it is valuable to
convert these to an energy rich raw material that can be removed from the dilute wastewater
environment through a physical separation of phases (i.e. to create the energy raw material in a
phase other than the liquid phase.) One approach is the formation of biomass for subsequent
processing. An extension of this approach is the accumulation of hydrocarbons of low oxidation state
as oils in the biomass, typically using algae. In this environment, the wastewater body provides
nutrients to support algal growth. The algae synthesise cellular components, including oils, using
energy from sunlight (photosynthesis). Additionally, the major carbon source,CO2, can be provided
from an industrial off-gas. This approach has been successfully demonstrated by using Botryococcus
braunii to produce hydrocarbons from a secondary sewage effluent (Banerjee, 2007). Due to their
simple cellular structure, microalgae have higher rates of growth, photosynthesis and productivity than
conventional crops. The cultivated algae could be used in several technologies such as anaerobic
digestion or the extraction of oil for biodiesel production.
Certain species of algae can produce large quantities of vegetable oil, up to 80% dry weight, as a
storage product (Becker 1994). Hence, algae can be up to 23 times more productive per unit area
than the best oil-seed crop (Table 12). Other advantages are that algae can be grown on wastewaters
(saline and other) and consume less fresh water than conventional crops since many species can
tolerate brackish or saline waters (Tsukahara and Sawayama 2005).
The conversion of the oil through transesterification is shown below (Figure 3). The transesterification
reaction converts triglycerides (oil) to alkyl esters (biodiesel) by the addition of an alcohol such as
methanol in the presence of a catalyst. Glycerol is a by-product (Harding, 2007).

Figure 3. Transesterification of oil to produce biodiesel

25

Table 12: Average productivities of common oil crops compared to algae

Most algal species with high lipid contents considered for biodiesel production are usually either
green

algae

(Chlorophyceae)

or

diatoms

(Bacillariophyceae)

(http://www1.eere.energy.

gov/biomass/pdfs/biodiesel_from_algae.pdf.). The main components of algae are proteins (typical 40


to 60%), carbohydrates (typically 15 to 30 %), nucleic acids (<5%) and lipids. Maximum reported lipid
contents for algal species noted for oil production are from 20 to 80% (as % dry weight). The type of
lipid varies with species, but generally the hydrocarbon chain is in the C16 to C20 range which is
suitable for making biodiesel (Ma and Hanna, 1999; Becker, 1994).
Algae have typically been grown in open raceway ponds. This has been by far the most common
method of algae production as such structures are fairly easy and cheap to construct or existing water
bodies may be used. The 5000 to 6000t per annum dry biomass of algae produced worldwide is
generated mainly in open systems. The most effective type of open system used is the raceway, an
oblong looped pond mixed by a paddlewheel. These raceways usually have water depths of 15 to
20cm with biomass concentrations of up to 1 g/L and productivities of 50 to 100 mg/L/day. The main
advantages of open systems are their low cost, use of natural illumination and lack of oxygenation.
Disadvantages include problems with pollution and contamination, evaporation of water and
susceptibility to environmental conditions such as temperature and dilution by rain (Ma et al., 1999;
Pulz, 2001). In contrast, for closed photobioreactor systems, nearly all parameters can be controlled
by the operator. However, photobioreactors are generally much more costly to build and operate than
open systems and are constrained by the need to achieve optimal light penetration. They often utilise
artificial lighting, which decreases the sustainability of the system, but they can achieve higher
productivities 800 to 1300 mg/L/day (Pulz, 2001).
Algal growth rates and productivities at laboratory scale are very difficult to replicate on a larger-scale
as the area receiving sufficient sunlight becomes limiting at large scale. However, once pilot scale is
achieved, most algal ponds/reactors (e.g. raceways) are modular in design, so scaling up of algal
growth does not change the cost or efficiency significantly. For harvesting, the concentration of algal
biomass is an important determinant. Certain unit operations (straining, filtration, flocculation) become
more economical at larger scale. Centrifugation, in particular, benefits from increased scale since it
requires large capital investment. Processing of harvested algal cells would benefit from economy of
scale (particularly usage time of equipment), but small-scale oil extraction and biodiesel production
are also feasible as the equipment is relatively simple and can be modularised. Large-scale

26

production may make it economical to purify the glycerol by-product of biodiesel production.
A variety of waste streams can be used for algal oil production, but must have sufficient
concentrations of inorganic nitrogen and phosphate to enable the support of algal growth. The optimal
range of nutrients to support algal growth is 0.2 to1.0 g/L nitrogen (as ammonia or nitrate) and 0.05 to
20 mg/L phosphate (Banerjee et al., 2002; Tsukahara and Sawayama, 2005). Organic carbon often
promotes growth of algae, but high levels of organic carbon can be inhibitory to growth and promote
the growth of heterotrophic bacteria. Sewage wastewaters are a particularly good medium and algae
help to remove the COD, nitrogen and phosphorous load, as well as many odours. Further, the
photosynthetic algae generate enough oxygen to allow bacterial oxidation of organic waste. This
releases nutrients (CO2, ammonia, phosphate) for conversion into additional algal biomass
(Benemann et al., 1977). It should be noted that nitrogen limitation is often used to increase the oil
yield of a variety of species. This is only possible in a high-N content waste stream if algal culture
continues until the nitrate, ammonia and urea are depleted and then nitrogen limitation occurs. In
secondary sewage treatment, it took 9 days for B braunii to reduce nitrate levels to below 0.01 mg/L
(Tsukahara and Sawayama, 2005). Studies at UCT are underway to relate nitrogen availability to lipid
productivity.
Benemann et al. (1977) reported that almost all sewage and wastewater oxidation ponds are of the
facultative type. These are 1 to 3 m deep and arranged in cells of up to 50ha, mixed only by wind
action and recirculation of effluent. The bottom of these ponds is anoxic, resulting in fermentation of
settled sludge and algae. This design is optimal for waste-treatment but not algal growth. To maximise
algal biomass, ponds must be shallow (30-50cm) and have short retention times (2-4 days in
summer) and be mechanically mixed. These are called high-rate algal ponds and can produce over
100 mg/L algal dry weight. This requires a retrofitting of ponds for algal growth for either biomass
(linked to combustion or fermentation) or lipid production.
Harvesting involves concentrating the algae from their dilute suspension, which is a considerable
challenge (Benemann et al., 1977). Settling or chemical flocculation can be economical, but can also
take considerable pond-time (settling can take 2-3weeks). Oil extraction techniques are similar to
those used in the food and vegetable oil industry, but require a purpose-built extraction plant close to
the algal ponds, for integrated operation. Transesterification requires unique equipment, but does not
have to be on-site and extracted oil may be transport to a central transesterification facility.
The components are relatively simple and consist of ponds, pumps and some form of harvesting and
oil extraction process that can be operated by trained, but not highly skilled personnel.
Transesterification is fairly simple but strict production standards are required in order to meet
specifications. Further, chemical handling requires a level of skill, but this operation can be
centralised.

27

1.5.2

Implementation of biodiesel production

No large-scale production of biodiesel from algal oil has been reported. Biodiesel has been made
from algal oil at the lab scale and it has been shown to be technically feasible (Miao and Wu, 2006)
and there are several examples.

Piggery wastewater in rural Korea used to grow Botryococcus braunii for hydrocarbons (An et
al., 2003). The growth rate of B braunii in secondarily treated sewage was 0.35 g/L/week and
algal concentration could be maintained at 400 mg/L after 11 days. (Tsukahara and
Sawayama, 2005)

A group of 7000 gallon ponds and one 1.6 acre pond in California showed that algae could
probably be produced for anaerobic digestion at less than $0.01/pound (1960 $) in open,
shallow, sewage-fertilised ponds with recycle of digester residue and water and the recovery
of heat and CO2 from the power plant for use in the growth process (Oswald and Golueke,
1955). Used algal species found naturally in sewage water (mostly green algae such as
Chlorella and Scenedesmus (together constitute 95-99% of wild sewage algal population in
California),

Chlorogonium,

Chlamydomonas,

Euglena

and

Microactinium).

Chlorella

pyrenoidosa could be maintained at 250 mg/L. A 2-3 day retention time was optimal and a
mixing velocity of at least 1 foot/sec required.

Solix biofuels in Colorado (http://www.solixbiofuels.com/) builds and tests photobioreacors


and a second generation prototype has been launched.

Some parts of South Africa (e.g. in the Upington area, N Cape and near Messina, Limpopo) have
ideal conditions for growth of high-oil content algae: long sunlight hours during summer, relatively high
temperatures. South Africa also has more open land (compared to Europe), and a relatively large
demand for transport fuels such as diesel (8.7 billion litres used in 2006). Several saline waste-water
streams relating to the mining and desalination industries could also potentially be used by salttolerant algal species, supplemented by seawater in coastal areas.
South Africa does not have significant large-scale algal farming experience. The biofuels industry in
South Africa is also still in the formative stage. There are a number of fine chemicals and nutraceutical
algal production, including NCSA, Cape Carotene, BioDelta.
The most recent development is a New Zealand company called Aquaflow Bionomic producing
biodiesel from wild algae grown in sewage ponds. This is believed to be the world's first commercial
production of bio-diesel from "wild" algae. The company expects to produce biodiesel at the rate of at
least one million litres of the fuel each year. They propose to harvest algae directly from the settling or
stabilisation ponds of standard WWTP and other nutrient-rich wastewaters (http://www.nzherald.co.nz/
and http://www.aquaflowgroup.com/)

28

1.5.3

Costs and benefits of biodiesel production from wastewaters

The costs vary widely depending on whether open ponds or closed bioreactors are used, the source
of the water, power used, nutrients and CO2 and the use of wastewater, flue gas, recycled heat and
power. There have been multiple pilot-scale trials of algal growth for biodiesel. The Aquatic Species
Programme (funded from 1978-1996) was a DOE program focussed on production of biodiesel from
high-lipid content algae in open ponds, using waste CO2 from flue gas. Great advances were made in
algal culture engineering, but the process, although technically feasible, was found to be
uneconomical at the time due to high cost of algae production. This technology has not yet been
successfully implemented anywhere on a large scale, mostly due to low algal yields making the
process uneconomical. The main factors were (1996):
* Low cost of conventional fuel (diesel prices were hovering around $1.10 per gallon).
* No monetary value for carbon mitigation capability of biodiesel.
* Higher than expected cost of the production system.
* Lower than expected productivity of outdoor open pond system.
A decade later, the world is different. Diesel is selling at or above the $2.50 mark; while the cost of
petroleum oil continues to increase. There are also new incentive s from the Clean development
Mechanism and Certified Emission Reductions (carbon credits) for algal biodiesel. A robust market
for renewable biofuels is emerging.

29

1.6 Fermentation for production of bioethanol


1.6.1

Background on bioethanol production from wastewaters

The production of bioethanol as a renewable liquid fuel is well established. Bioethanol can either be
used on its own or blended with conventional liquid fuels to form either Gasohol or Diesohol (as an
alternative to butanol). Factors driving its position as a biofuel include ease of transport as it is a
liquid, a heating value of about 67% of gasoline, ability to blend it directly to comprise some 10% of
gasoline with no modification of engines required and ability to enhance the octane rating (Bailey and
Ollis, 1986). While most development of this fuel has focused on its production from agricultural
materials, potential exists to utilise the organic fractions in wastewater and wastewater sludges for
conversion to ethanol (Bailey and Ollis, 1986).
Typically, bioethanol is formed by fermentation of simple sugars such as glucose and fructose to
ethanol under anaerobic conditions. Alternatively bioethanol may be formed by fermentation of the
monosaccharide to acetic acid and subsequent enzymic conversion to ethanol (Shuler and Kargi,
(2002; Nielsen and Villadsen, 1994). The typical batch industrial process has a fermentation time of
approximately 50 hours at 20 to 30 C. A yield of ethanol from sugar is approximately 90% of the
theoretical value, at an ethanol concentration of 10 to 16% (v/v) (Bailey and Ollis 1986).
Specific drawbacks of the conventional ethanol fermentation process include the dominant
contribution of raw materials to cost and environmental impact. The slow reaction rates result in large
fermenter volumes that contribute to capital cost, water usage and to the subsequent volume of
wastewater generated, all of which influence production cost. It is estimated that 3 to 10 litres of
wastewater are generated per litre of ethanol produced. Indeed this wastewater becomes a further
resource for energy generation. The high water content (typical ethanol concentrations achieved are
10 to 16 % (v/v)) implies that subsequent distillation is energy intensive. Conventional ethanol plants
may expend more than 30% of the bioethanol heating value in distillation. Key considerations for
efficient ethanol production include the use of continuous processes with cell retention or recycle,
allowing productivities to be increased 10 to 20 fold over batch processes (comparative operating
costs are estimated to decrease by 1.6 to 1.7 fold) (Bailley et al., 1986). Further, the use of bacteria
such as Zymomonas mobilis or other novel species enable enhanced productivity through increased
glucose uptake rates and ethanol formation rates. The use of thermophilic microorganisms would
further enable increased production rates and allow partial distillation of the ethanol during the
fermentation.
1.6.2

Implementation of bioethanol production

Various examples of the use of wastewater and wastewater sludges for the production of bioethanol
have been reported. The VTT Technical Research Centre of Finland has developed technology for the
distributed production of ethanol by fermentation of food processing wastewaters. This technology
enables production even at a small scale and is estimated to have potential to meet 2% of the total

30

volume of petrol sold in Finland and is currently being commercialised by St1 Biofuels Oy
(http://www.vtt fi/?lang=en.62).
In the FlexFuel project, developed by the University of Southern Denmark, the core biogas plant has
been expanded to include a pre-treatment plant for household waste, a pre-treatment plant for steam
explosion for treating solid waste to release monomeric units, and a unit for fermentation to ethanol
and its subsequent distillation (http://websrv5.sdu.dk/bio/flexfuel.). The overall process uses a wide
ranging combination of raw and waste materials and allows recovery of both bioethanol and biogas
for processing to combined heat and power, as well as N- and P-rich products for agriculture
(Figure 4).

Figure 4. Illustration of varied feedstocks to the ethanol biogas process (University of Southern Denmark).

There are several reports on the increased ethanol production following the processing of wastewater
from ethanol fermentations. CSR Distilleries in Australia implemented a continuous fermentation of the
distillery dunder, previously discarded to sea, resulting in the further production of ethanol, a reduction
in water consumption by 70%, and the production of a biodunder secondary-product and is sold as a
potassium rich agricultural fertiliser (http://www.p2pays.org/ref/04/03358.htm.; Rao et al., 2007).
Ethanol and hydrogen (H2) can also be produced by fermentation of the glycerol containing
wastewaters from the biodiesel production process. Despite a high salt content of the wastewaters,

31

productivities of 30 to 65 mmol/L/h and yields 0.80 to 0.85 mol ethanol per mol glycerol were been
achieved using the bacteria Enterobacter aerogenes (Ito et al., 2005).
BRI Energy (USA) has developed a unique process in which a range of waste materials undergo
thermal

gasification

at

high

temperature

into

CO,

CO2

and

H2

(http://techmonitor.net/techmon/05jul_aug/nce/nce_waste.htm.). The CO is then transformed into


ethanol by the acetogenic bacterium Clostridium ljungdahlii and the ethanol distilled. Typical yields are
340 litres per tonne bio-solids (municipal solid waste, biomass waste, animal wastes, etc.), increasing
to 680 litres per tonne as the oxidation state of the bio-solids decreases (used tyres, hydrocarbons).
1.6.3

Costs and benefits in bioethanol production from wastewaters

While the potential of ethanol production from wastewaters and wastewater sludges has been clearly
demonstrated through the selection of examples given above, a number of potential drawbacks must
be addressed in the implementing of this technology (Von Blottnitz and Curran, 2007). Firstly, the
ethanol fermentation is a very dilute process using large quantities of wastewater and large reactor
volumes. The process intensification is a key factor and may require concentration of the waste
organics where these occur as dilute solutes, suggesting a capital cost and energy penalty. Where
sterile or hygienic processes necessitate the use of heat treatment, the energy requirement is a
function of volume. Further, the ethanol product needs to be continuously extracted to prevent
inhibition of the process and requires energy intensive conventional distillation. The typical ethanol
producing microorganisms do not utilise a broad range of organic materials, hence enzymatic pretreatments or microorganisms with a broad substrate range are sought. Finally, co-production of
ethanol and H2 is of increasing interest, allowing effective utilisation of both the carbon and H2
fractions, compared to the alternative production of H2 and CO2.
The production of ethanol is currently limited to the use of carbohydrates (typically hexose
monosaccharides such as glucose and fructose) that are used in fermentations with yeasts
(Saccharomyces spp.) and, more recently, bacteria (Zymomonas spp).

These wastewaters are

readily available in the fruit processing industry, which are therefore appropriate sites for generating
energy from wastewaters using bioethanol fermentations.

However, the majority of agricultural

processing wastes consist of lignocellulose (combination of cellulose, hemicelulose and lignin) and
the required pre-treatment is a major challenge to the development of cost effective cellulose to
ethanol technology (Mosier et al., 2005). Pre-treatment can be carried out in several ways including
mechanical (Cadoche and Lopez) 1989), steam explosion (Gregg and Saddler, 1996) ammonia fibre
explosion (Kim et al., 2003), acid or alkaline pre-treatment (Damaso et al., 2004; Kuhad et al., 1997)
and biological treatment (Keller, et al., 2003). A number of thermochemical pre-treatment methods
have been developed to improve digestibility (Wyman et al., 2005), and these can also serve as a
source of process heat energy that can be recovered (Sheehan et al., 2003). The pre-treatment may
also involve the selective removal of compounds that are inhibitory to the downstream fermentation
processes.

32

All biomass for bioethanol goes through several steps that require resources (including energy inputs)
and infrastructure: the biomass needs to be grown, collected, dried, fermented, and the biofuel
combusted. The energy balance (total net energy gain; which is the energy output from combustion
the resulting ethanol fuel minus the energy input required to obtain and process the biomass) of
bioethanol from crops is thought to be marginally net positive with current farming practices (e.g.
energy balance of 1.3 for corn in the USA while and up to 8 for sugarcane in Brazil,
http://ngm.nationalgeographic.com). For the traditional corn (maize) fermentation up to 67 % of the
energy inputs are for the fermentation and distillation process, while the remaining 23% are energy
costs associated with the biomass production (agricultural crop) (Chandel et al., 2007 and
http://www.carbohydrateeconomy.org/). The use of wastewaters for bioethanol production reduces
the significant biomass production costs as well as the economic and social conflicts of food crops
being used for fuel production. There are also several developments that have optimised bioethanol
fermentations for improved yields and efficient ethanol distillation (Oh et al., 2000; Iraj et al., 2002). In
particular, membrane distillation (molecular sieve technology) has been shown to be the most efficient
and cost effective option among the available distillation processes (Binat and Simandl 1999 and
http://spie.org/x14497.xml) and it can significantly reduce the energy inputs involved in traditional
ethanol distillation.

33

1.7 Microbial fuel cells


1.7.1

Background on microbial fuel cells

Bacteria can be used to generate electricity that can be harvested in microbial fuel cells (MFCs)
(Davis and Higson, 2007). In a MFC, bacteria that oxidise a substrate are kept physically separated
from the electron acceptor by a proton exchange membrane. Electrons pass from the bacteria to the
anode in the same chamber and then via a circuit to the cathode where they combine with protons
and oxygen to form water. The difference in the potential coupled to electron flow produces electricity,
Figure 5 (below).

Figure 5. Diagram of a dual chamber microbial fuel cell (MFC). Organic material in wastewater entering the anode
chamber, is oxidised by microorganisms. The electrons (e-) can be transferred to the anode and via a wire to the
cathode where oxygen is reduced to water. A proton exchange membrane allows protons (H+) to flow between the
chambers. The flow of electrons (e-) is electrical current (electricity).

A number of studies have been conducted on MFCs operated with both pure cultures (e.g.
Shewanella putrefaciens, Pseudomonas aeruginosa, Geobacter sp., Rhodoferax ferrireducens and
the thermophilic Bacillus licheniformis or Bacillus thermoglucosidasius) (Kim et al., 2002; Rabaey and
Verstraete, 2005); and mixed cultures that were enriched either from sediment or activated sludge
from wastewater treatment plants (Rabaey et al., 2005; Lowy, 2006; Logan and Regan, 2006). MFCs
with mixed bacterial cultures have some important advantages compared to MFCs operated with pure
cultures, including a higher resistance to process disturbances, greater substrate versatility and a
higher power output (Rabaey et al., 2005). MFCs have recently been modified for hydrogen
production (Ditzig et al., 2007 and Liu et al., 2005). This bioelectrochemically assisted microbial
reactor, BEAMR, results in the production hydrogen at the cathode with high yields and a limited
surplus energy investment.

34

1.7.2 Implementation of MFCs


No large-scale MFC have been reported. Several aspects needed for an efficient MFC are hampering
scale up:
- The influent needs to reach the whole anode matrix sufficiently

- Protons need rapid diffusion towards the membrane

- Sufficient electrical contact needs to be established between bacteria in suspension and the
anode

- Sufficient voltage needs to be reached over the MFC to have a useful power

- Instatement of an aeration device should be avoided

To overcome these bottlenecks, several techniques can be copied from conventional fuel cells, such
as the use of fuel cell stacks, constructed from several plate shaped fuel cells (Larminie and Dicks
2000). MFCs can enhance the growth of bioelectrochemically active microbes during wastewater
treatment thus they have good operational stabilities. Continuous flow and single-compartment MFCs
and membrane-less MFCs are favored for wastewater treatment due to concerns in scale-up (Jang et
al., 2004; Moon et al., 2005; He et al., 2005). No large scale/ commercial plants are in operation.
Wastewaters of very different characteristics can be used. Sanitary wastes, food processing
wastewater, dairy manure, swine wastewater and corn stover are all suitable biomass sources for
MFCs because they are rich in organic matters (Suzuki et al., 1978; Liu et al., 2004; Oh and Logan,
2005; Min et al., 2005b; Zuo et al., 2006). Up to 80% of the COD can be removed in some cases (Liu
et al.,2004; Min et al., 2005b) and a Coulombic efficiency as high as 80% has been reported (Kim et
al., 2005). This technology can use bacteria already present in wastewater as catalysts to generate
electricity while simultaneously treating wastewater (Lui et al., 2004; Min and Logan, 2004).
The technology can be retrofitted into existing architecture. For example it may replace or supplement
a stage in the WWTP and provide additional source of power without conversion (i.e. produces
electricity directly). They can be employed in remote locations lacking electrical infrastructure or
critical applications that should not rely on external power inputs (i.e. biosensors). The design of the
equipment is still in development and may require skilled technical staff for design, installation and
initial support for the user. The equipment can be constructed with basic skills and understanding of
electronics. The systems are largely in the development stage and therefore may require considerable
optimisation. Although many operational difficulties have been to be overcome before full scale up and
implementation, a limited degree of monitoring is necessary. As long as there is constancy in the
supply and characteristics of the wastewater, the systems can operate robustly and efficiently with
very little operations and maintenance coats.
There are no commercialised applications of MFCs for energy production from wastewaters. There
are a myriad of small electronic remote devices that have favoured the development of MFCs
(Wilkinson, 2000). The robot, EcoBot-II, solely powers itself by MFCs to perform some behavior
including motion, sensing, computing and communication (Leropoulos et al., 2003b; Leropoulos et al.,

35

2004; Melhuish et al., 2006). Additional applications of MFCs are in biosensor and remote sensor
development (Liu et al., 2003); such as BOD sensors (Chang et al., 2004). MFCs are especially
suitable for powering small telemetry systems and wireless sensors that have only low power
requirements to transmit signals such as temperature to receivers in remote locations (Leropoulos et
al., 2005c; Shantaram et al., 2005). MFCs themselves can serve as distributed power systems for
local uses, especially in underdeveloped regions of the world. Locally supplied biomass can be used
to provide renewable power for local consumption. The MFC technology is particularly favoured for
sustainable long-term power applications. The technology is suited to developing countries with a
large proportion of the population not having access to electrical and water-borne/sanitation services,
since MFCs can generate electricity directly and in remote locations without significant costs on
infrastructure.
1.7.3

Costs and benefits in application of MFCs

Capex is the greatest costs with the electrodes and construction of the MFC chambers being
significant costs. The majority of cost in a MFC is attributed to the platinum catalysed carbon paper
used as the cathode component. The electrode costs have been limiting, but there have been several
improvements in design and are continual advancements in this area such as membrane-less MFCs
(Ghangrekar and Shinde, 2007) and alternative, cheaper electrodes (graphite, carbon felt, foam or
packed-granules, and platinum mesh) (Logan and Regan 2006).
To examine the potential for electricity generation, consider a EU food processing plant producing
7500 kg/day COD of effluent wastewater (Logan and Regan 2006). This represents a potential for 950
kW of power or 330 kW assuming 30% efficiency. A MFC (power output 1 kW/m3) with a volume of
350 m3 is therefore needed; costing approximately 2.6 m Euros (Rabaey and Verstraete, 2005) The
produced energy, calculated on the basis of 0.1 Euros per kWh, is worth about 0.3 M Euros per year,
providing a ten year payback without other cost-benefit considerations (i.e. inflation, operation,
maintenance costs, environmental offsets and compliance with local legislation for wastewater
discharge).
MFCs can purify waste to reduce the COD from 40-80% of the input material. The direct conversion
into electricity (coulombic efficiency) is typically more than 90%. There is little information on the
survival of pathogens following MFCs, but one would predict that the growth of a specific bacterial
consortium in the MFC will limit the presence of human enterobacterial pathogens (Salmonella, E.
coli, Ascaris, etc.). More research in this area is needed. There are benefits that need to be taken into
consideration such as the carbon dioxide mitigation due to both the reduction in electricity (coalgenerated) and also the reduced methane emission that would occur if the wastewater was treated by
conventional means. In terms of benefits, the direct conversion to electricity is a distinct advantage
compared to other technologies that generate heat which must then be converted to electricity (at
typical efficiency of 25%).

36

Some sludge will be generated by microbial biomass and non-biodegradable solids in the MFC. The
volume of microbial biomass generated is in between that of aerobic and anaerobic processes. The
sludge may prove useful as a fertiliser (as has sludge from aerobic and anaerobic digesters), or may
provide a biomass resource for energy generation by combustion or gasification. Similarly, effluent
wastewater from the MFC will still contain organic load (COD and nutrient) and may require further
purification, potentially with concomitant power generation via methane, or find application an
agricultural fertiliser.
MFCs are capable of converting the chemical energy stored in the chemical compounds in a biomass
to electrical energy with the aid of microorganisms. The chemical energy from the oxidisation of fuel
molecules is converted directly into electricity instead of heat, the Carnot cycle with a limited thermal
efficiency is avoided and theoretically higher conversion efficiency can be achieved. Efficiencies of 8097% have been reported (Chaudhury and Lovley, 2003) (Rabaey et al., 2003) (Rosenbaum et al.,
2006). However, MFC power generation is still very low (Tender et al., 2002; Delong and Chandler,
2002), that is the rate of electron abstraction is very low. One feasible way to solve this problem is to
store the electricity in capacitors or rechargeable devices and then distribute the electricity to endusers (Leropoulos et al., 2003a). The energy output is of the order 0.18 to 18 kW/m3 of the MFC and
the power output for the removed COD is 200-300 mW/g COD removed with loading rate of 0.574 kg
COD/m3 per day (Mohan et al., 2007).
The distinct advantages of MFCs are:

Able to utilise a variety of substrates at low concentrations

Direct conversion of substrate energy to electricity enabling high conversion efficiency.

Efficient operation at ambient (and below 200C) temperatures.

Elimination of gas treatment requirements because the off-gases of MFCs are mostly carbon
dioxide

Do not need energy input for aeration

Can be employed in remote locations lacking electrical infrastructure.

Furthermore, MFCs have recently been modified for hydrogen production instead of electricity
production (Ditzig et al., 2007). This bioelectrochemically assisted microbial reactor, BEAMR, results
in the production hydrogen at the cathode with high yields and a limited surplus energy investment.
This may be a more favourable route so that the energy generated can be stored and supply a
growing hydrogen economy.
The current high material costs mean that feasibility expenditure (Capex) for energy recovery with a
MFC is estimated to be at a level 10 times that of an anaerobic digester (Rabaey et al., 2005).
However, waste-driven MFCs produce less excess biomass than aerobic wastewater treatment
facilities (Rabaey et al., 2005). Since the sludge treatment cost of wastewater treatment facilities can
be considerable, this can have important implications on the economic feasibility of the process.

37

Furthermore, MFCs are a sustainable platform technology applicable in diverse fields without
substantial modification, because they can use a wide array of electron donors to efficiently generate
energy even at low-moderate temperatures, with low concentrations of electron donors in the
wastewater. The application of MFCs is therefore currently still in development and limited mainly by
material cost of the electrodes and the proton exchange membrane, but there are continual
advancements in this area such as membrane-less MFCs (Ghangrekar and Shinde, 2007) and
alternative, cheaper electrodes (graphite, carbon felt, foam or packed-granules, and platinum mesh)
(Logan et al., 2006). Therefore, current obstacles are the R&D and that should be pursued to develop
this technology sufficiently to implement engineering scale trials in South Africa.

38

1.8

REFERENCES

Acharya, J., Bajgain, M.S. and Subedi, P.S. (2005). Scaling up biogas in Nepal: what else is needed? Boiling
Point, 50: 2-4.
Ahring, B.K (2001) Biomethanation I . Advances in Biochemical Engineering/Biotechnology Vol 81. ISBN# 978-3540-44322-3
Ajah, A. N., Herder, P. M., and J.Grievink (2007). Integrated conceptual design of a robust and reliable wasteheat district heating system, Applied Thermal Engineering 27, 1158-1164.
Akunna, J. C. and Clark, M. (2000). Performance of a Granular-bed Anaerobic Baffled Reactor (GRABBR)
Treating Whisky Distillery Wastewater., Bioresource Technology 74, 257-261.
Amigun, B. and Von Blottnitz, H. (2007). Investigation of scale economies for African biogas installations. Energy
Conservation and Management, 48: 3090-3094.
An, J.Y., Sim, S.J, Lee, J.S. and Kim, B.W. (2003). Hydrocarbon production from secondarily treated piggery
wastewater by the green algae Botryococcus braunii. Journal of applied phycology. 15: 185-191
Bailey, J. and Ollis, D. (1986). Biochemical Engineering Fundamentals, (2nd ed). McGraw-Hill
Bain RL, Overend RP and Craig KR. (1998). Biomass-fired power generation. Fuel Processing Technology, 54, 116.
Baloch, M. I., Akunna, J. C., and Collier, P. J. (2007). The performance of a phase separated granular bed
bioreactor treating brewery wastewater, Bioresource. Technology 98, 1849-1855.
Banata F.A. and Simand, J. (1999) Membrane distillation for dilute ethanol Separation from aqueous streams
Journal of Membrane Science. Volume 163, Issue 2, 1 November 1999, Pages 333-348
Banerjee, Anirban; Sharma, R.; Chisti, Y.; and Banerjee, U. C. (2002). Botryococcus braunii: A Renewable
Source of Hydrocarbons and Other Chemicals. Critical Reviews in Biotechnology 0738-8551 vol: 22 (3) p:245279
Banerjee, S. and Biswas, G. K. (2004). Studies on biomethanation of distillery wastes and its mathematical
analysis, Chemical. Engineering. Journal 102, 193-201.
Bouallagui, H., Touhami, Y., Ben Cheikh, R. and Hamdi, M. (2005). Bioreactor performance in anaerobic digestion
of fruit and vegetable wastes. Process Biochem. 40: 989-995.
Brain, M. (2007). http://science.howstuffworks.com/ac-fox1.htm.
Cadoche L, Lopez GD (1989). Assessment of size reduction as a preliminary step in the production of ethanol
from lignocellulosic wastes. Biol. Wastes 30: 153157.
Callaghan, F. J., Wase, D. A. J., Thayanithy, K., and Forster, C. F. (2002). Continuous co-digestion of cattle slurry
with fruit and vegetable wastes and chicken manure, Biomass. and Bioenergy. 22, 71-77
Chandel A.K., Chan E., Rudravaram R. Lakshmi M., Narasu, L., Rao, V. and Pogaku Ravindra, P. (2007).
Economics and environmental impact of bioethanol production technologies: an appraisal. Biotechnology and
Molecular Biology Review Vol. 2 (1), pp. 014-032.
Chang, I. S., Jang, J. K., Gil, G. C., Kim, M., Kim, H. J., Cho, B. W., and Kim, B. H. (2004). Continuous
determination of biochemical oxygen demand using microbial fuel cell type biosensor, Biosensors and
Bioelectronics. 19, 607-613.

39

Chaudhuri, S.K., Lovley, D.R. (2003). Electricity generation by direct oxidation of glucose in mediatorless
microbial fuel cells. Nat. Biotechn. 21(10):1229-1232.
Chen, Y., Cheng, J.J. and Creamer, K.S. (2007). Inhibition of anaerobic digestion process: A review. Bioresource
Technology (in press). doi:10.1016/j.biortech. 2007.01.057
Choi, E. and Rim, J.M. (1991). Competition and Inhibition of sulphate reducers and methane producers in
anaerobic treatment. Water Science and Tech. 23:1259-1264
Colleran, E., and Pender, S. (2002). Mesophilic and thermophilic anaerobic digestion of sulphate-containing
wastewaters. Water Science and Technology, 45: 231-235.
Damaso MCT, Castro Mde, Castro RM, Andrade MC, Pereira N (2004). Application of xylanase from
Thermomyces lanuginosus IOC-4145 for enzymatic hydrolysis of corncob and sugarcane bagasse. Appl.
Biochem. Biotechnol. 115: 1003-1012.
Davis, F. and Higson, S. P. J. (2007) Biofuel cells--Recent advances and applications, Biosensors. and
Bioelectronics. 22, 1224-1235.
De Baere,L. (2000) Anaerobic digestion of solid waste: state-of-the-art. Water Science and Technology 41 (3),
283-290
Delong, E.F., Chandler, P., (2002). Power from the deep:microbially powered fuel cells tap into an
abundantecosystem energy circuit. Nat. Biotechnol. 20, 788789
Dewil, R., Appels, L. and Baeyens, J. (2006). Energy use of biogas hampered by the presence of siloxanes.
Energy Conservation and Management, 47: 1711-1722.
Ditzig, J., Liu, H., and Logan, B. E. (2007) Production of hydrogen from domestic wastewater using a
bioelectrochemically assisted microbial reactor (BEAMR), International. Journal of Hydrogen Energy Volume 32,
Issue 13, September 2007, Pages 2296-2304
Duerr, M., Gair, S., Cruden, A., and McDonald, J. (2007). Hydrogen and electrical energy from organic waste
treatment, International. Journal of Hydrogen Energy, 32, 705-709. Electrochemistry Communications. 6, 955958.
Fezzani, B. and Cheikh, R. B. (2007). Thermophilic anaerobic co-digestion of olive mill wastewater with olive mill
solid wastes in a tubular digester, Chemical. Engineering. Journal 132, 195-203.
Fraser, K., Ndwandwe, P., Basnal, A., Isafiade, N.S., Nyathi and Majosi, T. Water Conservation Through Energy
Conservation, Final Report to the Water Research Commission, October 2006.
GATE (Deutsches Zentrum fur Entwicklungstechnologien)(1999):Biogas Digest" Volume I-IV, German
Appropriate Technology Exchange (http://www5.gtz.de/gate/id/publications.htm),
Ghangrekar, M. M. and Shinde, V. B. (2007). Performance of membrane-less microbial fuel cell treating
wastewater and effect of electrode distance and area on electricity production, Bioresource Technology 98, 28792885.
Ghosh, S., Conrad, J. R., and Klass, D. L. 1975. Anaerobic acidogenesis of wastewater sludge. Journal of the
Water Pollution Control Federation 47, 30-45.
Gijzen, H.J., Zwart, K.B, Verhagen, F.J.M., Vogels, G.D. (1988.High-rate two-phase process for the anaerobic
degradation of cellulose, employing rumen microorganisms for an efficient acidogenesis Biotechnol. Bioeng.
Vol/Issue: 31:5 Pages: 418-425

40

Gohil, A. and Nakhla, G. (2006) Treatment of tomato processing wastewater by an upflow anaerobic sludge
blanket-anoxic-aerobic system, Bioresource Technology 97, 2141-2152.
Gosh, B.K., Gosh, A. (1992) Degradation of Cellulose by Fungal Cellullase. Microbial Degradation of Natural
Products, ed. G. Winkelmann, VCH Publishers, Inc., New York, pp.84-126.
Gregg DJ, Saddler JN (1996) Factors affecting cellulose hydrolysis and the potential of enzyme recycle to
enhance the efficiency of an integrated wood to ethanol process. Biotechnol Bioeng. 51: 375-383.
Gunaseelan, V. N. (1997) Anaerobic digestion of biomass for methane production: A review., Biomass and
Bioenergy 13, 83-114.
Harding, K.G., Dennis, J.S., Von Blottnitz, H. and Harrison, S.T.L. (2007). A life-cycle comparison between
inorganic or biological catalytic process routes for the production of biodiesel. Under review: Journal of Cleaner
Production
He, Z., Minteer, S.D., Angenent, L. (2005). Electricity generation from artificial wastewater using an upflow
microbial fuel cell. Environ. Sci. Technol. 39:52627.
Ho, M-W. (2005). Biogas bonanza for third world development. ISIS press release
http://www.i-sis.org.uk/BiogasBonanza.php.
Holmgren, K. (2007). Role of a district-heating network as a user of waste-heat supply from various sources the
case of Gteborg, Applied Energy 83, 1351-1360.
http://biro-net.aber.ac.uk/Symposium_2004/abstracts_2004_symposium/EcoBot%20II.pdf.
http://spie.org/x14497.xml. SPIE: Small steps toward membrane distillation commercialization.
http://www.carbohydrateeconomy.org/library/admin/uploadedfiles/How_Much_Energy_Does_it_Take_to_Make_a
_Gallon_.html. How Much Energy Does It Take to Make a Gallon of Ethanol? David Lorenz and David Morris.
August 1995. Institute for Local-Self Reliance (ILSR)
http://www.gastechnology.org/webroot/app/xn/xd.aspx?it=enweb&xd=4reportspubs\4_8focus\himetfocus.xml.
HIMET-A Two-Stage Anaerobic Digestion Process for Converting Waste to Energy
http://www.ist-world.org/ProjectDetails.aspx? Demonstration of a flexible plant processing organic waste, manure
and/or energycrops to bio-ethanol and biogas for transport.
Huang, L.P., Jin, B. Lant, P., Zhou, J. (2005). Simultaneous saccharification and fermentation of potato starch
wastewater to lactic acid by Rhizopus oryzae and Rhizopus arrhizus. Biochem. Engin. J. 23 265276
IEA Bioenergy, Tasks 24 and 37 http://www.ieabioenergy.com/OurWork.aspx
Iraj N, Giti E, Lila A (2002). Isolation of flocculating Saccharomyces cerevisiae and investigation of its
performance in the fermentation of beet molasses to ethanol. Biomass Bioenergy 23: 481486
Ishikawa, S., Hoshiba, S., Hinata, T., Hishinuma, T. and Morita, S. (2006). Evaluation of a biogas plant from life
cycle assessment (LCA). International Congress Series, 1293: 230-233.
Ito T., Nakashimada, Y., Senba K, Matsui, T. and Nishio N. (2005). Hydrogen and Ethanol Production from
Glycerol-Containing Wastes Discharged after Biodiesel Manufacturing Process". J.Biosci.Bioeng., 100, 260-265
Jang, J.K., Pham, T.H., Chang, I.S., Kang, K.H., Moon, H., Cho, K.S. (2004). Construction and operation of a
novel mediator-and membraneless microbial fuel cell. Process Biochem 2004;39:100712.

41

Jarrell, K.F., Sprott, G.D. and Matheson, A.D. (1984). Intracellular potassium concentration and relative acidity of
the ribosomal proteins of methanogenic bacteria. Canadian Journal of Microbiology, 30: 663-668.
Jimnez, A. M., Borja, R., and Martn, A. (2004). A comparative kinetic evaluation of the anaerobic digestion of
untreated molasses and molasses previously fermented with Penicillium decumbens in batch reactors.
Biochemical Engineering Journal 18, 121-132. Journal of Hydrogen Energy In Press, Corrected Proof, -166.
Karim M. I. A. and SIsrunk. W. A. (1985) Treatment of Potato Processing Wastewater with Coagulating and
Polymeric Flocculating Agents. Journal of Food Science 50:6, 16571661
Kavalov, B. and Peteves, S D. (2005). Status and Perspectives of biomass-to-liquid fuels in the European Union,
European Commission, Directorate General Joint Research Centre (DG JRC) Institute for Energy.
Kayhanian, M. (1994). Performance of high-solids anaerobic digestion process under various ammonia
concentrations. J. Chem. Tech and Biotech. 59: 349-352
Keenan, P.J., Isa, J. and Switzenbaum, M.S. (1993). Inorganic solids development in a pilot-scale anaerobic
reactor treating municipal solid waste landfill leachate. Water and Environmental Research, 65: 181-188.
Keller F.A., Hamilton J.E., Nguyen Q.A. (2003). Microbial pre-treatment of biomass: potential for reducing severity
of thermochemical biomass pre-treatment. Appl. Biochem. Biotechnol. 27-41: 105-108.
Kim HT, Kim JS, Sunwoo C, Lee YY (2003). Pre-treatment of corn stover by aqueous ammonia. Biores Technol.
90: 39-47.
Kim, H. J., Park, H. S., Hyun, M. S., Chang, I. S., Kim, M., and Kim, B. H. (2002). A mediator-less microbial fuel
cell using a metal reducing bacterium, Shewanella putrefaciens, Enzyme and Microbial. Technology 30, 145-152
Kim, I.S., Kim, D.H. and Hyun, S.-H. (2000). Effect of particle size and sodium ion concentration on anaerobic
thermophilic food waste digestion. Water Science and Technology, 41: 67-73.
Kivaisi A.K. and Mtila M. (1997). Production of biogas from water hyacinth (Eichhornia crassipes) (Mart) (Solms)
in a two-stage bioreactor,. World Journal of Microbiology and Biotechnology, Volume 14, Number 1, pp. 125131(7)
Kuhad RC, Singh A (1993). Lignocellulose biotechnology: current and future prospects. Crit Rev Biotechnol 13:
151172.
Larminie, J. and Dicks, A. (2000). Fuel Cell Systems Explained. John Wiley and Sons, New York.
Leropoulos, I. A., Greenman, J., Melhuish, C., Hart, J. (2005). Comparative study of three types of microbial fuel
cell. Enzyme Microb. Tech. 37:23845
Leropoulos, I., Greenman, J., Melhuish, C., Hart, J. (2005). Energy accumulation and improved performance in
microbial fuel cells. J. Power Sources. 145:2536
Leropoulos, I., Melhuish, C., Greenman, J. (2003). Artificial Metabolism: Towards True Energetic Autonomy in
Artificial Life. In: Banzhaf W, Christaller T, Dittrich P, Kim JT, Ziegler J, editors. Advances in Artificial Life.
Proceedings of the 7 European Conference on Artificial Life; 2003 Sep 14-17; Germany. Germany: SpringerVerlag. p 792-799.
Leropoulos, I., Melhuish, C., Greenman, J. (2004). Energetically autonomous robots. In: Groen F, et al., editor.
Intelligent autonomous systems, vol. 8. Amsterdam: IOS Press; p. 12835. (March, 2004).

42

Lin, C.Y. (1992). Effect of heavy metals on volatile fatty acid degradation in anaerobic digestion. Water Research,
26: 177-183.
Lin, C.Y. (1993). Effect of heavy metals on acidogenesis in anaerobic digestion. Water Research, 27: 147-152.
Liu, H, Grot, S and . Logan B.E (2005), Electrochemically assisted microbial production of hydrogen from
acetate, Environ. Sci. Technol. 2005, 39, 43174320.
Liu, H., Ramnarayanan, R., Logan, B.E. (2004). Production of electricity during wastewater treatment using a
single chamber microbial fuel cell. Environ. Sci. Technol. 28:22815.
Liu, J., Olsson, G., and Mattiasson, B. (2003). Monitoring of two-stage anaerobic biodegradation using a BOD
biosensor, Journal of Biotechnology 100, 261-265.
Logan, B. E. & Regan, J. M. (2006). Electricity-producing bacterial communities in microbial fuel cells. Trends in
Microbiology, 14, 512-518.
Lowy, D. A., Tender, L. M., Zeikus, J. G., Park, D. H., and Lovley, D. R. (2006). Harvesting energy from the marine
sediment-water interface II: Kinetic activity of anode materials, Biosensors and Bioelectronics. 21, 2058-2063.
Ma, F. and Hanna, A. (1999). Biodiesel production: a review. Bioresource technology, 70, 1-15.
McCartney, D.M and Oleszkiewicz, J.A. (1991). Sulfide inhibition of anaerobic digestion of lactate and acetate.
Water Research, 25: 203-209.
McCartney, D.M and Oleszkiewicz, J.A. (1993). Competition between methanogens and sulphate reducers:
Effect of COD: sulphate ratio and acclimatization. Water and Environmental Research, 65: 655-664.
Melhuish, C., Leropoulos, I., Greenman, J., Horsfield, I. (2006). Energetically autonomous robots: food for
thought. Auton. Robot. 21:18798
Min, T., Yoshikawa, K., and Murakami, K. 2005b. Distributed gasification and power generation from solid
wastes", Energy 30. 2219-2228.
Moon, H., Chang, I.S., Jang J.K., Kim, B.H. (2005). Residence time distribution in microbial fuel cell and its
influence on COD removal with electricity generation. Biochem. Eng. J. 27:5965.
Mosier, N., Wymanb, C. Dalec, B. Elanderd, R . Lee, Y, Holtzapplef, M. and Ladisch M. (2005). Features of
promising technologies for pre-treatment of lignocellulosic biomass. Bioresource Technology Volume 96, Issue 6,
April 2005, Pages 673-686
Mouneimne, A.H., Carrre, H., Bernet, N. and Delgens, J.P. (2003). Effect of saponification on the anaerobic
digestion of solid fatty residues. Bioresource Technology, 90: 89-94.
Natural Resources Canada (2004), Industrial Processes: Food and Drink: Process Integration, Typical Projects,
accessed online at http://cetc-varennes.nrcan.gc.ca/en/indus/agroa_fd/ip_pi/pt_tp.html, October 2007.
Neves, L., Ribeiro, R., Oliveira, R., and Alves, M. M. (2006). Enhancement of methane production from barley
waste, Biomass. and Bioenergy. 30, 599-603.
Nielsen, J . and Villadsen, J (1994). Bioreaction Engineering Principles, Plenum Press.
Nishio, N. and Nakashimada, Y. (2007). Recent development of anaerobic digestion processes for energy
recovery from wastes, Journal of Bioscience. And Bioengineering. 103, 105-112.

43

Oh KK, Kim SW, Jeong YS, Hong SI (2000). Bioconversion of cellulose into ethanol by nonisothermal
simultaneous saccharification and fermentation. Appl. Biochem. Biotechnol. 89: 153.
Oh, S.E., Logan, B.E. (2005). Hydrogen and electricity production from a food processing wastewater using
fermentation and microbial fuel cell technologies. Water Res. 39:467382.
Oleszkiewicz, J.A., Marstaller, T. and McCartney, D.M. (1989). Effects of pH on sulfide toxicity to anaerobic
processes. Environmental Technology Letters, 10: 815-822.
Oswald and Golueke, (1960). Biological transformation of solar energy. Adv. in App. Micro. 11:233-242.
Parawira, W., Murto, M., Zvauya, R., and Mattiasson, B. (2004). Anaerobic batch digestion of solid potato waste
alone and in combination with sugar beet leaves, Renewable. Energy 29, 1811-1823.
Parkin, G.F., Speece, R.E., Yang, C.H.J. and Kocher, W.M. (1983). Response of methane fermentation to
industrial toxicants. Journal of the Water Pollution Control Federation, 55: 44-53.
Pavan P, Battistoni P, Cecchi F, Mata-Alvarez J. Two-phase anaerobic digestion of source sorted OFMSW
(organic fraction of municipal solid waste): performance and kinetic study. Water Sci Technol. 2000;41(3):111-8.
PI. (1999). Conference Proceedings: Volume III, Summary. International Conference on Process Integration,
Copenhagen, Denmark, 7-10 March 1999.
Pizarro, C.X., Mulbry III, W.W., Blersch, D., Kangas, P. (2006). An economic assessment of algal turf scrubber
technology for treatment of dairy wastewater. Ecological Engineering. 26:321-327.
Pulz, O. (2001). Photobioreactors: production systems for phototrophic microorganisms. Appl Microbiol
Biotechnol 287-293.
Rabaey, K. and Verstraete, W. (2005). Microbial fuel cells: novel biotechnology for energy generation. Trends in
Biotechnol. 23: 291-2
Rabaey, K., Lissens, G., Siciliano, S.D. and Verstraete, W. (2003). A microbial fuel cell capable of converting
glucose to electricity at high rate and efficiency. Biotechn. Lett. 25:1531-1535.
Rajesh A. R, R., E.M. Rajesh, R. Rajendran, S. Jeyachandran (2008) Production of bioethanol from cellulosic
cotton waste through microbial extracellular enzymic hydrolysis and fermentation

EJEAFChe, 7 (6) pg 2984-

2992.
Ren, Z., Ward, T. E., and Regan, J. M. (2007) Electricity Production from Cellulose in microbial electricity
generation a bacterial fuel cell operating on starch, microbial fuel cell. Process Biochem. 39:100712.
Rosenbaum, M, Schroder, U., Scholz, F. (2006). Investigation of the electrocatalytic oxidation of formate and
ethanol at platinum black under microbial fuel cell conditions. J. Solid State Electrochem. 10:8728.
Saddoud A, and Sayadi S. (2007) Application of acidogenic fixed-bed reactor prior to anaerobic membrane
bioreactor for sustainable slaughterhouse wastewater treatment. J Hazard Mater. 2007 Nov 19;149(3):700-6.
Saddoud, A., Hassairi, I. and Sayadi, S. (2007). Anaerobic membrane reactor with phase separation for the
treatment of cheese whey, Bioresource. Technology 98, 2102-2108.
Schfer J., Ulrike Schmid-Staiger and Walter Trsch (1999). One and two-stage digestion of solid organic waste.
Water Research Volume 33, Issue 3, Pages 854-860 Gabriele Schober,
Scott, S. A. (2004). The gasification and combustion of sewage sludge in a fluidized bed, PhD dissertation,
University of Cambridge.

44

Seema J. Patel, R. Onkarappa, K. S. Shobha: Study Of Ethanol Production From Fungal Pretreated Wheat And
Rice Straw. The Internet Journal of Microbiology. 2007. Volume 4 Number 1
Shantaram, A., Beyenal, H. Veluchamy, R.R.R.A. and Lewandowski, Z. (2005). Wireless sensors powered by
microbial fuel cells.Environ. Sci. Technol. 39 (13), pp. 50375042.
Sheehan J, Aden A, Paustian K, Killian K, Brenner J, Walsh M, Nelso R (2003). Energy and environmental
aspects of using corn stover for fuel ethanol. J. Ind. Ecol. 7: 117-146.
Shuler, M. L. and Kargi, F. (2002). Bioprocess Engineering. Prentice Hall.
Sikkema, J., De Bont, J.A.M. and Poolman, B. (1994). Interaction of cyclic hydrocarbons with biological
membranes. Journal of Biological Chemistry, 26: 8022-8028.
Singh, K.J. and Sooch, S.S. (2004). Comparative study of economics of different models of family size biogas
plants for state of Punjab, India. Energy Conservation and Management, 45: 1329-1341.
Soto, M., Mendz, R. and Lema, J.M. (1991). Biodegradability and toxicity in the anaerobic treatment of fish
canning wastewaters. Environmental Technology, 12: 669-677.
Speece, R.E. (1993). Anaerobic biotechnology for industrial waste treatment. Water Science and Tech. 17: A416A427.
Sterrit, R.M. and Lester, J.M. (1980). Interaction of heavy metals with bacteria. Science of the Total Environment,
14: 5-17.
Suzuki, S., Karube, I., Matsunaga, T. (1978). Application of a biochemical fuel cell to wastewater. Biotechnol.
Bioeng. Symp. 1978;8:50111.
Tender, L.M., Reimers, C.E., Stecher, H.A., Holmes, D.E., Bond, D.R., Lowy, D.A., Pilobello, K., Fertig, S.J.,
Lovley, D.R. (2002). Harnessing microbially generated power on the seafloor. Nat. Biotechnol. 20(8):821-5
Tsagarakis, K.P. and Papadogiannis, C. (2006). Technical and economic evaluation of the biogas utilization for
energy production at Iraklio Municipality, Greece. Energy Conservation and Management, 47: 844-857.
Tsukahara, K. and Sawayama, S. (2005) Liquid fuel production using microalgae. Journal of the Japan Petroleum
Institute 48, 251-259.
Ueno, Y., Kawai, T., Sato, S., Otsuka, S., and Morimoto, M. (2007). Biological production of hydrogen from
cellulose by natural anaerobic microflora., Journal of Fermentation and Bioengineering 79, 194-197.
Ugwuanyi, J.O., Harvey, L.M. and McNeil, B. (2007). Linamarase activities in Bacillus spp. responsible for
thermophilic aerobic digestion of agricultural wastes for animal nutrition. Waste Management, 27: (11), 1501-1508
Vallero, M.V.G., Hulshoff Pol, L.W., Lettinga, G. and Lens, P.N.L. (2003b). Long-term adaptation of methanol-fed
thermophilic (55C) sulphate-reducing reactors to NaCl. Journal of Industrial Microbiology and Biotechnology, 30:
375-382.
Van Langerak, E.P.A., Gonzales-Gil, G., Van Elst, A., Van Lier, J.B., Hamerlers, H.V.M. and Lettinga, G. (1998).
Effects of high calcium concentrations on the development of methanogenic sludge in upflow anaerobic sludge
blanket (UASB) reactors. Water Research, 34: 1255-1263.
Verma, M., Brar, S.K., Tyagi, R.D., Surampalli, R.Y. and Valro, J.R. (2007). Starch industry wastewater as a
substrate for antagonist, Trichoderma viride production. Biores. Techn. 98 (11):, 2154-2162

45

Vijayaraghavan, K. and Ahmad, D. (2006). Biohydrogen generation from palm oil mill effluent using anaerobic
contact filter, International. Journal of Hydrogen Energy 31, 1284-1291.
Vogels, G.D., Kejtjens, J.T. and Van der Drift, C. (1988). Biochemistry of methane production. In: Zehnder, A.J.B.
(Ed.), Biology of Anaerobic Microorganisms. John Wiley and Sons, New York.
Von Blottnitz, H. and Curran, M. A. (2007). A review of assessments conducted on bio-ethanol as a transportation
fuel from a net energy, greenhouse gas, and environmental life cycle perspective. Journal of Cleaner Production,
15, 607-619.
Werther, J. and Ogada, T. (1999). Sewage sludge combustion, Progress. in Energy and Combustion. Science 25,
55-116.
Wilkinson, S. (2000). Gastrobots benefits and challenges of microbial fuel cells in food powered robot
applications. Auton. Robot. 9:99111.
Wyman, CE. (1996). Ethanol production from lignocellulosic biomass: Overview. In Handbook on Bioethanol,
Production and Utilization (Wyman, C.E., ed.), pp.1-18, Taylor & Francis .
Yen, H.-W. and Brune, D. E. (2007). Anaerobic co-digestion of algal sludge and waste paper to produce
methane., Bioresource Technology 98, 134.
Yu, H. Q., Zhao, Q. B., and Tang, Y. (2006). Anaerobic treatment of winery wastewater using laboratory-scale
multi- and single-fed filters at ambient temperatures, Process Biochemistry 41, 2477-2481.
Zayed, G. and Winter, J. (2000). Inhibition of methane production from whey by heavy metals protective effect
of sulfide. Applied Microbiology and Biotechnology, 53: 726-731.
Zeeman G., and Sanders W. (2001)

Potential of anaerobic digestion of complex waste(water).; Water Sci

Technol. 44(8):115-122.
Zuo, Y., Maness, P.C., Logan, B.E. (2006) Electricity production from steamexploded corn stover biomass. Energ
Fuel. 20:171621.
Zupancic, G. D., Straziscar, M., and Ros, M. (2007). Treatment of brewery slurry in thermophilic anaerobic
sequencing batch reactor. Bioresource Technology 98, 2714-2722.

46

2:

APPENDIX: SURVEYS, CASE STUDIES AND WORKSHOPS

The following are the findings from surveys, case studies and workshops of the energy from
wastewater project:
2.1 Municipal wastewater treatment plants (WWTP)
The characteristics of the wastewater discharges will vary from location to location depending upon
the population and industrial sector served, land uses, groundwater levels, and degree of separation
between storm water and domestic wastes. The domestic wastewater includes both the blackwater
(faeces and urine from toilets) and greywater (washing and food preparation). The industrial loads will
have characteristics depending upon their nature and will contribute differently to the individual
WWTP.
Wastewater Energy potentials at WWTP
The potential for energy can be calculated for the entire population as well as that currently serviced
by municipal WWTP infrastructure at present.
(1) Total potentials. It is estimated that human activities generate 200 L/day wastewater per person
(http://www.dwaf.gov.za/dir_ws/content/lids/PDF/Q&A.pdf.). Since the population of South Africa is
48.5 million then 9700412600 L per day is generated. With the COD of 0.860 g/L (see typical COD of
Cape Town WWTP, below) this amounts to 8342354836 g per day. As 1 day= 86400s, this is 96555
g/s or 96.555 kg/s. Since the energy content is 15 mJ/kg this amounts to 1488 MJ/s or 1488 MW. This
includes the domestic blackwater and greywater loads.
(2) Potential from current municipal WWTP:
7600 mL/day at WWTP. With a COD of 0.860 g/L this amounts to 6536000000 g per day. As 1 day=
86400s, then 75.648 kg/s is produced. Since the energy content is 15 mJ/kg this amounts to 1134
MJ/s or 1134 MW.
If the WWTP are operating at 75% of their capacity the amount is 850 MW.
Note: The current municipal WWTP service approx. 60% of the population presently (see 4)
(3) Total domestic blackwater load
48 502 063 people generating 100 g (dry weight) of faeces per day with an energy of 15 MJ/kg. This
represents 48 502 06 kg per day or 56.136 kg/s. Since energy value is 15 MJ/kg, this represents 842
MJ/s or 842 MW
(4) Total captured domestic blackwater (human faeces component): the serviced population:
Currently

60.4%

of

the

population

(48

502

063)

have

flush

toilets

(http://www.statssa.gov.za/community_new/content.asp Community survey 2007) that the WWTP can


capture and this therefore amounts to 29 295 246 persons. Therefore 29 295 246 people with flush

47

toilets generating 100 g (dry weight) of faeces per day. This represents 29 295 25 kg per day or
33.907 kg/s. Since energy value is 15 MJ/kg, this represents 509 MJ/s or 509 MW.
Table 2.1.1: The capacity of WWTP in South Africa is shown below:
Plant size category
Estimated number of
Median wastewater flow
Total volume of wastewater
1
2
(M/day)
plants (#)
treated (M/day)
treated (M/day)
< 0.5
488
0.25
122
0.5 2
108
1.25
134
2 10
208
6
1248
10-25
93
17.5
1625
25-100
71
62.5
4460
Estimated total volume of wastewater treated in South Africa
7589
1
Estimated number of plants obtained from the DWAF data base corrected for the plant category sizes used in this report.
2
The actual plant sizes in each plant size category was available.
Data: Golder and Associates study of WWTP commissioned by the WRC

The Western Capes spread of wastewater treatment plant sizes is similar to the national
situation; small (0.5 to 2.5 mL/day) and micro size plants (<0.5 mL/day) constitute the majority of
WWTPs in South Africa. Activated sludge process technology appears to be the dominant
treatment process.
Western Cape (73)
North West (69)
Micro plants <0.5 Ml/day
Small plants 0.5-2.5 Ml/day
Small plants 1-5 Ml/day
20-100 Ml/day
Large plants
7.5-25 Ml/day
Medium plants
5-20 Ml/day
Medium plants
2.5-7.5 Ml/day
Medium plants
1-5 Ml/day
Small plants

Medium plants 2.5-7.5 Ml/day


Medium plants 5-20 Ml/day
Medium plants 7.5-25 Ml/day

<0.5 Ml/day
Micro plants

Large plants 20-100 Ml/day

Fig. 2.1.1: Spread of WWTP sizes in Western Cape and North West Province
City WWTP
The WWTP in Cape Town are shown in the map below (Fig 2.2). The WWTP are distributed
throughout Cape Town but the density does not always reflect the area served or population density
(e.g. Cape Flats serves an exceptionally large area). More detailed data for the Cape Town WWTP is
shown below.

48

Fig. 2.1.2: WWTPs in Western Cape

49

92.2

155.3

152.8

734

473

802

586

278

1 329

(141)

834

1 073

681

599

859

1 149

784

716
Av. COD =837
mg/L

Kraaifontein

Llandudno

Macassar

Melkbosstrand

Millers Point

Mitchells Plain

Oudekraal

Parow

Potsdam

Scottsdene

Simons Town
Wesfleur
Domestic
Wesfleur
Industrial

Wildevolvlei

Zandvliet

121.4

1 219.8
Total [v] =
16523

260.2

414.1

830.1

233.3

nd

779.0

nd

290.3

964.4

6.6

305.4

nd

1 207

803.0

(1 310)

614

Green Point

80.7

Klipheuwel

663

Gordons Bay

4 419.8

63.9

1 662.3

Hout Bay

881

Cape Flats

917.3

1 649

745

Camps Bay

nd

(818)

Bellville Orbal
Borcherds
Quarry

1 245

Bellville DA

2 751.4

789

Athlone

Volumes
(ML/month)
[y]

Influent
COD
(mg/L) [x]

Wastewater
treatment
plant

[x].[y]=
1582946000

873377000

203997000

175567000

133403000

55228000

281934000

890697000

194572000

1035291000

170116000

773449000

3122000

223870000

146530000

493042000

53504000

3893844000

476055000

1512133000

2069190000

2170539000

Load
Kg/month
3
([x].[y])10

COD (mg/L)

50

Table 2.1.2: COD values for Cape Town WWTPs

WWTP

t in
l
i
t
l
l
l
t
y
s
d
c
y
n
ne - DA rba rry Ba l at B ay oi n B a we tei n dno s ar ran oi n Pl a raa row dam ene ow sti t ri a l vl e v li e
a s
lo
P
e
P
k
s
s
u
e
F
t
u
d
n
O
t
s
a
T
u
s
h
u vo nd
a ss rs l ls de P ot tts s
e
e d
o
m
t
u
n
l
p
e
f
n
d
c
l
h
n
Q
i
i
A lv
p
o ree Ho l ip
a l a Ma bo i ll e he Ou
P c o on Do In lde Za
l
il le ds Cam Ca rd
k M i tc
K Kra L
S i m ur eur Wi
o
G
Be el lv her
el
G
S fl e sfl
M
M
B rc
es e
Bo
W W

5,000
4,500
4,000
3,500
3,000
2,500
2,000
1,500
1,000
500
0

COD values for Cape Town Wastewater


treatment plants

ANODISING
BAKERY
BEVERAGES
BOTTLING
BUILDING
CHEMICALS
CONSTRUCTION
CONTAINER
COSMETICS
DAIRY

WINERY
VEH/WASH
TRANSPORT
TISSUE
TEXTILE
TANNERY
RAGS
RADIATOR
PRINTING
PLATING
PLASTIC
PHARM ACEUT
PAPER
PAINT
PACKAGING
M OTOR
M ETAL
M EAT
LAUNDRY

ENGINEERING
ENZIMES
FOOD
FOOD/FISH
FOOD/SPICE
FUEL
GALVANISING
GAS
LABORATORY
LAUNDRY

LABORATORY
GAS
GALVANISING
FUEL
FOOD/SPICE
FOOD/FISH
FOOD
ENZIM ES
ENGINEERING
DAIRY
COSM ETICS
CONTAINER
CONSTRUCTION

MEAT
METAL
MOTOR
PACKAGING
PAINT
PAPER
PHARMACEUT
PLASTIC
PLATING
PRINTING
RADIATOR
RAGS

CHEM ICALS
BUILDING
BOTTLING
BEVERAGES
BAKERY
ANODISING

TANNERY
TEXTILE
TISSUE
TRANSPORT
VEH/WASH

500000

1000000

WINERY

5000000

10000000

15000000

Load (kg/month)

Fig.2.1.3: Contribution of wastewater from industry in Cape Town

51

20000000

The characteristics of wastewater (more detailed analysis of nutrients) for two Cape Town WWTPs
are shown in the table below (data 2007).
Table 2.1.3: Chemical characteristics of the wastewater in some Cape Town WWTPs
CAPE
FLATS

2/8

2/15

2/22

3/1

3/8

3/15

3/22

3/29

4/5

4/12

4/19

4/26

COD mg/L

1,266

937

835

1,099

1,326

788

817

547

2,629

846

827

1,393

SS mg/L

920

595

455

705

766

430

463

368

1,840

480

263

880

N mg/L

103.2

73.5

66

90

94.6

77

76.1

56.6

173.4

84.4

79.1

103.9

P mg/L

27.3

17.3

16.3

22.3

26.9

15.8

20

12.8

61.7

15.9

17.7

28.6

ATHLONE

2/8

2/15

2/22

3/1

3/8

3/15

3/22

3/29

4/5

4/12

4/19

4/26

COD mg/L

915

890

751

866

863

873

667

810

990

901

855

921

SS mg/L

392

336

300

318

290

312

254

308

357

370

420

280

N mg/L

61.4

59.3

55.3

61.2

55.6

57.2

53.1

52.4

59.2

60.9

72.4

55.2

P mg/L

10.9

18.5

9.3

10.7

9.4

10.8

9.6

10.8

16.5

11.7

11.6

8.6

COD: Chem Oxygen Demand. SS: Suspended Solids. N: Total Kjeldahl Nitrogen. P: Total Phosphorus

From this data it can be seen that the loads are relatively constant for the majority of plants although
some exceptional intermittent high loads can be seen. This may be due to seasonal or unexpected
loads or operational failure (power failure or other) and seasonal rainfall (resulting in water ingress
into the sewage system). The current loads to the Cape Town municipal WWTP have a typical
average COD of 860 mg/L and a total volume of 16523 ML/month or 551 ML/day. The approximate
values for suspended solids 400 mg/L, total nitrogen of 50-100 mg/L and total phosphorous of 10-30
mg/L. The loads are not evenly distributed at the various WWTP (presumably reflects capacity and
locality/wastewaters served). For, example that of Borcherds quarry is twice that of Athlone. Further
analysis of the characteristics of the wastewater at municipal WWTP is also shown for two WWTP in
Cape Town Cape Flats and Athlone.
It is obvious that the food, beverages and textile industries (in order of deceasing contribution)
contribute the greatest loads to Cape Town WWTPs. Of the remaining minor contributors, dairy and
fish food followed by paper and packaging and the meat industries dominate. The total load for
industry is 35139581 kg/month and the Cape Town WWTPs currently treats 1582946000 kg/month
(see above). Therefore, the industry contributes 22% of the total load to the Cape Town WWTP. Note
that the actual total industry wastewaters are considerable larger since industries may carry out
primary treatment on-site before discharging to the WWTP or they may discharge to land, rivers or
sea.
The industrial loads are difficult to capture since many are received as mixed streams. In Cape Town
only Westefleur operate separate streams for industry and domestic wastewater. There may also be
fluctuations in the characteristics, as can be seen with Cape Flats in the fluctuations in COD nitrate,
and phosphorous. This may be due to performance or fluctuation in the wastewaters received since
many industries may generate seasonal waste streams. In contrast, the characteristics of Athlone are
more constant.
52

In the case of Durban Municipality, the loads and characteristics are fairly similar to Cape Town, but
there are several exceptions where industrial activity results in loads and characteristics very different
from that of domestic sewage. The data on distribution of wastewater (industrial as a % of the total
WWTP) show that wastewaters have selected characteristics due to their locality and degree of
industrial loads. For example, Hammarsdale which is 95% chicken and textile with COD values 4
times that of the typical domestic WWTP (Table below). Many of the industries in Durban do not
discharge to municipality WWTP but directly to sea with very high loads: COD 500-140 000 mg/L (e.g.
Frametex, Sappi, CG Smith, Drum services). Other industries such as SAB have on-site treatment
(AD) and then discharge to the municipal WWTP. The energy potential opportunities are therefore a
lot greater (and the environmental burdens not taken into consideration) and the application of a
universally efficient method/process for energy recovery from city municipal wastewaters with the
present infrastructure will not be easily attainable. The municipal WWTP are currently being
interrogated through a survey.

53

528.12

30%

30%

30% (textile)

5%

40% (textile, food, paper)

KWAMASHU

NORTHERN

NEW GERMANY

PHOENIX

VERULAM

782.58
2806.67

MPUMALANGA

HAMMARSDALE
731.72

837.42

UMBILO

50% (textile)

926.44

KWANDENGEZI

UMHLATUZANA

683.18

DASSENHOEK

827.06

671.80

UMHLANGA

HILLCREST

410.92

829.83

648.98

951.04

UMDLOTI

95% (textile + chicken)

660.40

UMKOMAAS
561.95

618.51

ISIPINGO

822.08

AMANZIMTOTI

50% (brewery, chemicals)

623.36

KINGSBURGH

768.53
769.19

12%

CENTRAL

COD (AVG) mg/L

CRAIGIE BURN

% Industrial

WORKS

186.6369

31.0845

265.0754

573.8493

433.2646

231.4788

17.10069

25.32393

39.7086

48.9282

1983.138

43.7124

1569.088

356.2091

737.1186

133.3515

452.8992

1896.161

328.1469

5761.461

ML/month

Load (kg/month)

54

Fig. 2.1.4: COD values (load) for Durban Municipality WWTP

I
I
N H I
N
T
K
X
AL UR G OT GO AS HU ER ANY NI AM O T G A E EZ I LO G A AL E ES NA
TR B UR IMT IP IN MA A S TH M OE UL DL A N HO NG MB AN SD CR ZA
N
N
B
L
E
R
U
L
E
R
H
I
M
M
R
U A R LL T
Z IS KO A
E
A I A
O GE P VE U MH SE ND
C IG G S AN
M
A
UM M H HL
U AS A
U KW N W
R IN M
P M
D KW
E
M
C K A
M HA
N
U

500

1000

1500

2000

2500

3000

3500

4000

4500

5000

14840

2830

7160

9010

Saldana

Hopefield

Shelly Point

St. Helena

Langebaan

Paternoster

Activated
Sludge
Activated
Sludge
Sequence
batch reactor
Activated
Sludge
Activated
Sludge
Ponding

Primary
method
5000
904
200
1256
1400
225

1056
226
659
1440
-

Capacity
Volume
m3/d

1366

Aeration
Kwh/d

175

1511

1096

129

285

2324

Actual
volume
m3/d

154

901

716

45.9

283

1484

Actual
COD
Kg/d

455

671

780

356

956

688

Average
COD
mg/L

14.8

70.3

83.5

45.3

95.4

83.7

TKNN
Mg/L

59

45.9

56.4

23.5

41.5

51.83

NH3
Mg/L

0.09

0.10

0.11

0.13

0.10

0.12

TKN/COD

8.1

7.0

7.6

7.4

8.1

7.6

pH

52

88

94

76

93

87

%COD
removal

>1800

Nil

200

>1800

nil

100

Final effluent
Coliform/100
mL

55

4. In many cases the WWTP are operating poorly with discharge failing to meet the required governmental regulations.

municipality indicate 0.6-3.7 kWh/m3 wastewater treated).

3. Activated sludge appears to be preferred methods for municipal WWTP, with consequent large energy inputs for aeration (data from Saldana

WWTP, the potential for energy generation is a lot greater that that predicted from the current WWTP operation and loading capacity.

2. Since approx 40% of the population do not have flush toilets and many industries treat-on site and do not discharge their loads to the municipal

industry in a given area the loads and characteristics can vary widely.

ammonia. The total nitrogen/COD is 0.1 and the total phosphorous is approx 10 to 20 mg/L. However depending upon the locality and degree of

1. Wastewaters with COD values of 700-800 mg/L, pH 7-8. The total nitrogen is 50 to100 mg/L; with approximately half the total nitrogen present as

From this data it can be summarised that:

Data: Abott & Abott

Population

WWTP

Table 2.1.4: Characteristics of wastewater at WWTPs in Saldanha, Western Cape

Several smaller WWTP are operating on capacities of <5 ML/day (<150 Ml/month). An example is shown for Saldana municipality, below.

Town WWTP

Peri-urban/Rural WWTP and decentralised wastewater treatment systems (DEWATS)


In peri-urban and rural settings wastewater is often treated locally. This may range from the
bucket system, French drains, septic tanks, pit-latrines and no toilet at all. In 2007 a little more
than 60% of households in South Africa had access to a flush toilet. Gauteng, Free State,
North West and Western Cape were the only provinces where more than 50.0% of
households that own a flush toilet. Free State had the highest number of households still
using the bucket system. More than half of households in Limpopo (56.3%) used a pit latrine
without ventilation, and 25.2% of all households in the Eastern Cape had no toilet at all
(http://www.statssa.gov.za/community_new/content.asp)
Table 2.1.5: Percentage of sewage systems in provinces in South Africa

There are many reasons for the failure to implement water-borne sewage (flush toilets) and
centralised WWTP, but a major factor is the infrastructural costs the cost of implementing
WWTP for peri-urban and rural households the lower density of the population means high
cost of piping. There are therefore huge opportunities to be gained in decentralised approach
that brings local benefits decentralised wastewater treatment system (DEWATS). The
application of AD systems in these rural or remote settings where the human and animal
faeces are captured and harnessed in an AD digester for the production of biogas fuel for
cooking and heating. The benefits also include water conservation, reduced fertiliser inputs
since the partially purified wastewaters can be used for irrigation (in compliance with
reductions limits of COD and coliforms as determined by Water Act and DWAF). These are
explored in the case study of the performance of household anaerobic digesters.

56

2.2 : Brewing Industry


Introduction and Data Sources
This case study explored the potential for energy recovery from wastewaters generated by
the brewing industry in South Africa. The sector is considered in terms of formal and informal
operations. SAB produces approximately 98% of the clear beer brewed in South Africa (Cole
2007). However, clear beer constitutes only about 60% of total beer production the
remainder is comprised of sorghum beer, where, informal, small-scale production dominates
(although some industrial scale installations do exist (such as those of United National
Breweries). The wastewaters of this informal production are not deemed to be readily
quantifiable. Therefore it was decided to define the wastewaters produced by the operations
of SAB and scale up accordingly to provide a value for industry as a whole.
To obtain the desired data from SABs seven breweries, a detailed characterisation of the
various wastewaters was requested. However, similar to the requests originally submitted to
the pulp and paper industry (Sec. 2.1.1) these were met with reluctance. Therefore, the
necessary data on the volumes and carbonaceous content of combined wastewater streams
was provided by Cole (2007). Supporting information was obtained from SAB (2006) and from
the company website. The work of Cohen (2006) provided data collected from SABs
Newlands brewery.

Results of the Study


The technologies and wastewater characteristics at each of the seven SAB breweries are
fairly uniform. Approximately 3.2 L of combined wastewater (from packaging, the cellars and
the brew house), with a COD concentration of about 3000 mg/L prior to any on-site treatment,
is generated per L of beer produced (Cole 2007). A total beer production of 2.605 107 hL
(SAB 2006), together with knowledge of the relative capacities of the breweries, then allows
the calculation of the individual beer and wastewater volumes presented in Table 1. Untreated
brewery wastewater typically has a temperature in the region of 40C and a pH of around 5.5
(Cohen 2006).

Current production and wastewater treatment figures


Five of the SAB breweries include their own wastewater treatment plants where the majority
of the COD is removed prior to disposal to sewer. Alrode, Rosslyn, Prospecton, Newlands and
Ibhayi each have an anaerobic digestion unit, whilst the wastewater at Ibhayi undergoes
further aerobic treatment before passing through sand and carbon filters. Information on the
COD removal performance of the anaerobic digesters was gained from Cole (2007) and is
3
shown in Table 1. Methane production figures are based on a yield of 0.35 m CH4 per kg

COD digested (McCarty 1964), a ratio which corresponds well with that estimated by Cohen

57

(2006) in his study of SABs Newlands brewery. The energy associated with this methane is
simply calculated using a heat energy value for methane of 39.74 MJ.m-3. The methane-rich
biogas captured from the anaerobic digesters is currently flared. However, it is the companys
intention to explore opportunities of utilising this renewable energy source in the generation of
process heat and/or electricity across its breweries (Cole 2007). In this regard, Cohen (2006)
notes that, although fluctuations in biogas flow and composition may present technological
design challenges, the overall efficiency of fuel use in cogeneration (combined heat and
power) can be as high as 85%.

Table 2.2.1 Annual combined wastewater and wastewater treatment data for SABs breweries
Beer
Brewery

Wastewater

COD

COD post-AD

COD removed

CH4
3

Energy
3

x10 (hL)

(Ml)

(mg/L)

(mg/L)

(tonne)

x10 (m )

(GJ)

Alrode

6,885

2,203

3,000

400

5,728

2,005

79,673

Rosslyn

5,972

1,911

3,000

1,500

2,867

1,003

39,873

Prospecton

4,396

1,407

3,000

400

3,658

1,280

50,875

Newlands

3,733

1,194

3,000

400

3,106

1,087

43,196

Ibhayi

1,991

637

3,000

400

1,656

580

23,038

Chamdor

1,825

584

3,000

n/a

n/a

n/a

n/a

Polokwane

1,244

398

3,000

n/a

n/a

n/a

n/a

TOTAL

26,046

8,334

17,015

5,955

236,655

Total anaerobic digestion potential


The study of Cohen (2006) found that a relatively constant COD reduction rate of about 90%
was achievable through the anaerobic digestion of SABs brewery wastewaters. This results
in the total potential figures for energy recovery presented in Table 2. To provide some
perspective, the potential thermal energy of this methane production, at each of the
breweries, amounts to around 7.6% of the sites total energy requirements (157.4 MJ.hL-1
(SAB 2006)). Efficiency of conversion from thermal to electrical energy is conservatively
estimated at 25%.

58

Table 2.2.2 Total potential figures for energy recovery from SABs wastewaters using anaerobic digestion
COD removed
Per

year

CH4 Produced per year


3

Energy

produced

Thermal power

(tonne)

x10 (m )

per year. (GJ)

(MW)

Alrode

5,948

2,082

82,737

2.63

Rosslyn

5,160

1,806

71,772

2.28

Prospecton

3,798

1,329

52,832

1.68

Newlands

3,225

1,129

44,857

1.42

Ibhayi

1,720

602

23,924

0.76

Chamdor

1,577

552

21,930

0.70

Polokwane

1,075

376

14,952

0.48

Sub-total

22,504

7,876

313,005

9.92

(combined)

1,030

360

14,321

0.46

Total

23,533

8,237

327,325

10.38

Breweries

Maltings

plants

SABs two malting plants at Caledon and Alrode have been included in the analysis of Table
4. They have a combined capacity of about 220 000 tonnes per annum and have a specific
wastewater production of approximately 2.6 kL per tonne (SAB 2006) at a COD concentration
of just less than 2000 mg/L (Cole 2007).
Scale-up to entire sector
Taking into consideration the proportion of South Africas total beer production for which SAB
is responsible (ca. 60%) , the data in Table 2 can be broadly extrapolated to yield a total
theoretical potential of about 17.3 MW of thermal power recoverable from the wastewaters of
the brewing industry sector as a whole. However, due to the informal nature of a large
proportion of the extrapolated production and the resultant uncertainties surrounding the
comparability of the nature and quantity of wastewaters, the accuracy of this total figure is
unclear. Similarly, the extent to which these informal wastewaters are accessible for energy
recovery is of concern.
It should also be noted that, although anaerobic treatment of brewery wastewater is a proven,
well-established process, with many systems operational worldwide, alternative technologies
incorporating energy recovery, such as microbial fuel cells (Associated Press 2007) and
bioethanol fermentation (Rao et al., 2007), may also be considered.

59

Concluding remarks Brewing Sector


The wastewaters of the brewing sector are generally suitable for biological energy recovery.
However, the sector is dominated by multinational organisations with sufficient resources and
motivation to explore options for energy recovery from their wastewaters themselves, on a
rigorous, site-specific basis. The impression was gained that such investigations have been,
and continue to be, conducted and that viable technologies will be implemented. Furthermore,
biological treatment of a large proportion of the industrys wastewaters is already currently
performed, albeit not necessarily incorporating energy recovery.

60

2.3 Survey of Paper and Pulp wastewaters


Introduction and Data Sources
Initial investigations revealed that the South African pulp and paper industry is dominated by
Sappi and Mondi who account for 100% of virgin pulp production and about 88% of all paper
and board production (Chamberlain et al., 2005). Furthermore, only a limited number of
installations exist within the sector. As a result, it was initially decided to compile detailed data
for each of the principal installations, as opposed to largely aggregated sector data. Water
circuit data detailing separate wastewater flows (their volumes, carbonaceous content,
temperature, pH, concentration of potential inhibitors, etc.) would constitute a valuable future
resource for the project, enabling identification of certain individual sites or even streams of
particular interest for potential intervention and assisting in establishing the most suitable
energy recovery technology.

The various process technologies and product ranges in place at South Africas different mills
differ markedly and consequently so too do the wastewater volumes and characteristics. Even
if only aggregated sector results are sought, it is not sufficient to obtain wastewater data from
a single sample installation and apply it across the entire industry. Ideally, mill data from each
of the principal sites is required. For this study, it was felt that the most straightforward means
of obtaining data was to directly approach the relevant personnel in the pulp and paper
industry. Being unable to assist with data themselves, recommended contacts at its member
companies were provided by PAMSA, an industry association representing all South Africas
major pulp & paper producers (Hunt 2007). Considering their dominance of the industry and
the resultant importance of their co-operation, Sappi and Mondi were then approached with
requests for detailed information on wastewater volumes and characteristics at each of the
mills, in line with the objective of this study. However, due to the environmental pressures
placed on the industry, such data is considered sensitive and the requests were viewed with
some suspicion. Information on combined wastewaters prior to any on-site treatment was
received for the Mondi-related operations at Richards Bay (flow, COD, pH, temperature)
(Naylor 2007) and those at Felixton, Piet Retief and Springs (flow, COD) (Scheckle 2007).
Anecdotal reports were also provided by Van der Westhuizen (2007) and Scheckle (2007).

Further information was gathered through the open literature. Macdonald (2004) documents
the water consumption figures for each of the Mondi, Sappi, Nampak and Kimberly-Clark
operations in South Africa. These can be used to provide a measure of wastewater flows
through broad application of the assumption that typical mills expel about 85% of water
consumed as wastewater (Macdonald 2007). In the absence of mill data, the first step
towards estimating the desired wastewater information is determining the various pulp and
paper production figures at each installation. These can then be employed in conjunction with

61

literature values of specific (i.e. per unit production) wastewater flow associated with the
production of individual pulp or paper products in order to quantify the combined wastewater.
Similarly, values of specific COD load to wastewater can be used to describe the character of
these wastewaters.

A full breakdown of pulp and paper production at the Mondi-related sites at Felixton, Piet
Retief and Springs was provided by Scheckle (2007). The remaining data had to be sourced
from available literature where, a high degree of product specificity was required. The virgin
pulp was characterised by bleach status (bleached or unbleached), fibre source (wood or
bagasse) and method of pulping (groundwood, thermo-mechanical, neutral sulphite semichemical, kraft, soda, Ca-sulphite or Mg-sulphite). Likewise, paper was specified as coated
fine paper, uncoated fine paper, newsprint, linerboard, fluting, cartonboard, other packaging
paper or tissue. A further distinction necessary was the proportion of these paper production
figures (if any) that had as their fibre source an on-site recycled fibre plant. This data was
ultimately compiled through a synthesis of total industry production figures (PAMSA 2007) and
the information obtained from Macdonald (2004), Chamberlain et al. (2005), company
websites and various online engineering articles. As a result of inconsistencies between
sources (due to, for example, recent expansion or lack of distinction between capacity and
production) and incomplete or aggregated data, this exercise was not straightforward and
some educated guesswork was required in reconciling this information.

Specific wastewater data


The primary source for the data covering specific wastewater flows and specific COD loads to
wastewater was the European Commission report (2001), whilst other important sources
include the Food and Agriculture Organisation of the United Nations (1996), Rintala and
Puhakka (1994), Pokhrel and Viraraghavan (2004) and Thompson et al. (2001). Since it is
common industry practice to perform wastewater treatment prior to release to the
environment, the focus is commonly on regulatory compliance and, as such, data on
untreated wastewater can be difficult to source. Such information was nevertheless obtained
for the production of as many as possible of the different pulp and paper products
manufactured in South Africa. Care is also required in attaching appropriate data to integrated
pulp and paper mills or to non-integrated mills producing only either pulp or paper products
integrated mills have, for example, opportunity for greater recycling, and the effects of
additional processes/products on wastewaters are not necessarily simply additive.

It is important to recognise that this specific wastewater data acquired is for a combined
stream comprising such effluents as cooking and evaporator condensates, washing losses,
accidental spills and those from wood handling, the bleach plant, stock cleaning and the

62

paper machine. As previously mentioned, it was the original objective to provide a more
detailed survey enumerating and characterising the separate available wastewater streams to
whatever extent possible. Some of these individual streams can have very high COD
concentrations, although these are typically reduced through stripping or extraction of organic
by-products. It is also important to recognise that the pulp and paper industry already
recovers a great deal of energy from its process (as opposed to wastewater) streams and is
one of the largest producers of energy from renewable resources in the country (FSA 2004).
Indeed, the recovery of energy and pulping chemicals through the incineration of the
concentrates resulting from the evaporation of spent cooking liquor is essential to the cost
effectiveness of most pulping operations. The fuel value of the recoverable black liquor from
kraft pulp mills, for example, is normally enough to make them at least self-sufficient in terms
of heat and electrical energy.

Results of the Study


Annual data for combined wastewater prior to any on-site treatment is presented in Table
2.3.1 for each of the principal mills in South Africa. The Wastewater column contains the
data extracted from the work of Macdonald (2004), or, in the case of the Mondi-related
operations at Richards Bay, Felixton, Piet Retief and Springs, data received from the mills
themselves. The Est wastewater column contains the values calculated using the mills pulp
and paper production figures and the various specific wastewater flow figures gained from the
literature. The ranges of these specific flow values are typically substantial and the selection
of a single value for each of the products was guided to an extent by the Wastewater
numbers. These single values tend to fall towards the lower end of the ranges, a result which
is expected since, as large consumers of water operating in a water scarce country, South
African mills do generally by necessity exercise efficient water use strategies (FSA 2004). The
Est wastewater figures are included primarily as a gauge of the potential accuracy of the
technique. Whilst the exceptions hint at the difficulties in making generalisations in this
industry, it can be seen that the estimates are accurate to within 15% for twelve out of
seventeen mills.

For the Richards Bay, Felixton, Piet Retief and Springs installations, the COD data presented
in Table1 are those provided by the mills themselves. Otherwise, the values are those gained
using the mills pulp and paper production figures and the various ranges of specific COD load
to wastewater sourced from the literature. Considering the vastness of the literature ranges of
COD evident in Table 1, it is anticipated that the mill data would fall within these calculated
ranges. Indeed, this is the case for Richards Bay, Piet Retief and Springs (all towards the
lower end). The notable exception, however, is Felixton, where even the upper end of the
calculated COD range (722-3122 ppm) is less than 14% of the actual value. This certainly
underlines the variability within the industry and the potential errors in this analysis.

63

Table 2.3.1 Annual combined wastewater data (prior to any on-site treatment) for the South African pulp and
paper industry sector
Wastewater

Est wastewater

Temperature

COD
pH

Mill

(C)

(ML)

(ML)

(mg/L)

Merebank

10264

10085

470-1659

Richards Bay

21361

21300

1399

8.24

44.38

Felixton

1933

2000

22842

Piet Retief

566

1750

6021

Springs

1046

1008

1940

Saiccor

33320

32582

615-3073

Stanger

6248

3760

319-1175

Enstra

7586

6227

578-1929

Adamas

506

462

848-3221

Ngodwana

10413

13996

1219-4607

Tugela

15470

6387

358-1305

Cape Kraft

428

408

595-4167

Bellville

655

576

733-2443

Kliprivier

506

432

711-2372

Riverview

208

180

721-2404

Rosslyn

298

320

671-4698

803

864

897-2989

360

625-4375

1210

789-3116

Mondi

Sappi

Nampak

Kimberly-Clark
Enstra
New Era
Gayatri
Other

It may be expected that the temperature of the combined wastewaters at each of the mills
would be of approximately the same order as that at Richards Bay (Macdonald 2007). The
pH, however, may differ markedly, particularly at Sappis Saiccor mill due to the acidic nature
of their pulping operation.

64

1.

Anaerobic digestion potential

The application of anaerobic digestion technology for the treatment of wastewaters from the
pulp and paper industry is well-researched (Rintala and Puhakka 1994; Pokhrel and
Viraraghavan 2004; Thompson et al., 2001; Lee Jr et al., 2006). Although wastewaters from
particular sources within a mill may be more degradable than those from others, most
contaminants are readily degraded (European Commission 2001), and a relatively constant
COD removal efficiency of 80% is achievable through anaerobic digestion. Residual COD
levels may however necessitate additional treatment prior to release to the environment
(Thompson et al., 2001).
Assuming an 80% reduction in COD and a methane yield of 0.35 m3 CH4 per kg COD
digested (McCarty 1964), the industrys total energy recovery potential using anaerobic
digestion can be calculated (Table 2). A heat energy value for methane of 39.74 MJ.m-3 is
used.

65

Table 2.3.2 Total potential figures for energy recovery from South Africas pulp & paper industry sectors
wastewaters using anaerobic digestion
Energy

Electric power

X10 (m )

(GJ)

(MW)

3862-13620

1352-4767

53712-189446

1.7-6.0

Richards Bay

23914

8370

332621

10.55

Felixton

35329

12365

491397

15.58

Piet Retief

2728

954

37942

1.2

Springs

1623

568

22575

0.72

Saiccor

16384-81920

5734-28672

227885-1139425

7.23-36.13

Stanger

1593-5872

557-2055

22152-81680

0.7-2.59

Enstra

3506-11707

1227-4097

48764-162834

1.55-5.16

Adamas

343-1304

120-456

4777-18136

0.15-0.58

Ngodwana

10152-38375

3553-13431

141211-533763

4.48-16.92

Tugela

4437-16145

1553-5651

61711-224563

1.96-7.12

Cape Kraft

204-1427

71-499

2835-19843

0.10-0.63

Bellville

384-1280

134-448

5341-17804

0.17-0.56

Kliprivier

288-960

101-336

4006-13353

0.13-0.42

Riverview

120-400

42-140

1669-5564

0.05-0.18

Rosslyn

160-1120

56-392

2225-15578

0.07-0.49

576-1920

202-672

8012-26705

0.26-0.85

180-1260

63-441

2504-17525

0.08-0.56

Other

764-3016

267-1056

10626-41950

0.34-1.33

Total

106547-243921

37292-85372

1481965-3392703

47.0-107.6

COD removed

CH4

(tonne)

Merebank

Mill

Mondi

Sappi

Nampak

Kimberly-Clark
Enstra
New Era
Gayatri

Within the context of this study, the wastewater of Mondis Felixton mill is obviously of
particular interest due to its elevated COD concentration and associated potential recoverable
energy. However, the competitiveness of the industry dictates that these wastewaters are
understandably also of interest to Mondi and the implementation of various energy recovery
technologies is currently being investigated (Scheckle 2007). Similarly, Sappi have thoroughly

66

researched the possibility of recovering energy from wastewater streams using an array of
technologies at each of their installations (Van der Westhuizen 2007). Currently none is
economically viable, although the steadily increasing trend of fossil fuel prices demands that
the situation be regularly reassessed.
It is also pertinent to note that many of the plants already have in place primary and
secondary wastewater treatment facilities, from which the effluent is released to the
environment. However, anaerobic treatment is not common, as these systems are relatively
sensitive to disturbances. Available information suggests that no energy recovery is
performed.

Concluding remarks
The wastewaters from the pulp and paper sector are theoretically suitable for biological
energy recovery, providing some of the operational considerations surrounding system
robustness are able to be managed. The theoretical potential thermal power from the pulp
and paper sector as a whole, using anaerobic digestion to generate biogas is 45-100 mW.
In conclusion, it is noted that the sector is dominated by multinational organisations with
sufficient resources and motivation to explore options for energy recovery from their
wastewaters themselves, and this on a rigorous, site-specific basis. The impression was
gained that such investigations have been, and continue to be, conducted and that viable
technologies will be implemented. Furthermore, biological treatment of a large proportion of
the industrys wastewaters is already currently performed, albeit not necessarily incorporating
energy recovery.

67

2.4: Survey of petrochemical refineries


Wastewater Generation in Refineries
Petroleum refineries generate wastewater during operations including desalting of crude oil,
steam stripping operations, product fractionators, reflux drum drains and boiler blowdown [1].
The crude oil desalter unit is the major source of contaminated wastewater in the crude oil
refinery with the production rates ranging between 40-90 litres of wastewater per tonne of
crude processed. This wastewater contains phenols, oil, chlorides and organic matter, and
suspended solids are also generated during desalting due to the presence of sediments in
crude oil. Sour wastewater is the aqueous waste coming from crude or resulting from
stripping operations and distillation/fractionation overhead condensates; it contains phenols,
sulphides, ammonia and in some cases oil depending on the position of unit source in the
process [2].
Typically the wastewater streams are mixed, thus giving rise to streams that are more difficult
to treat and invariably lead to higher processing costs. These streams are typically treated at
on-site WWTP together with surface water runoff and domestic wastewaters (sewage) from
the site [3, 4]. Different refineries make use of different process routes and feedstocks, and
hence defining a typical wastewater stream is difficult. Attempts to obtain wastewater
compositions from local refineries were unsuccessful. A range of literature values was thus
collected for use in this study, as shown in the Table below.

Table 2.4.1: Refinery wastewater BOD and COD concentrations in (mg/L)


COD

BOD

Heavy Metals

Oil

Phenol

Reference

300-600

150-250

0.1-100

100-300

20-200

300-800

150-350

0.1-100

<3000

20-200

236-490

109-275

25-266

6-270

It is noted that the compositions presented in the Table above are representative of those
from crude oil refineries. The composition of wastewater from coal-to-liquid and gas-to-liquid
refineries will be expected to differ significantly. This is discussed further below. Oil and
solvents are the main wastewater pollutants in petroleum refining operations which enter the
feedstream [4]. Other contaminants include such as ammonia, sulphur compounds and spent
acids. Heavy metals (copper, zinc, chromium, nickel, lead, mercury, cadmium, selenium,
arsenic, silver and barium) and cyanide are also present which potentially impact on the
treatment processes which may be used for these streams, depending on their concentrations
in the stream.
In terms of wastewater volumes, anything from 0.1 to 5 m3 of wastewater is generated per
tonne of crude oil processed [2, 7].

68

Refineries in South Africa


South Africa has four crude oil refineries, as well as the Sasol and PetroSA coal-to-liquid and
gas-to-liquid operations. The four crude oil refineries, along with their crude processing
capacity, are shown in the Table below. Also shown is the outer limits of the wastewater
volumes generated from these refineries.

Table 2.4.2 Crude Oil Processing and Wastewater Generation from Refineries in South Africa (2002 Figures)
[8]
Crude

m of wastewater per

Processing

m of wastewater per
3

Refinery

Capacity (Barrels/day)

Crude (tonne/day)*

day (0.1 m /tonne)

year (5 m /tonne)

Engen

115 000

15 640

1 564

78 200

Sapref

180 000

24 327

2 432

121 635

Natref

108 000

11 623

1 162

58 115

Caltex

100 000

13 515

1 351

67 575

TOTAL

503 000

65 105

6 510

325 525

* At a density of 0.85 kg/L

The total volume of wastewater generator by the four refineries is thus between 118,816,000
and 2,376,000 ML per year.

Comment on the Energy Recovery Potential from Refinery Wastewater


Consideration of the COD and BOD values presented in Table 2.4.1 suggests limited potential
for recovery of energy from wastewater from crude oil refineries. In addition, the presence of
various contaminants have the potential to adversely affect any biological processes which
may be used for energy recovery.

A review was conducted to determine best practice in wastewater treatment from crude oil
refineries around the world, and whether energy recovery from wastewater is practiced. The
primary option for energy recovery is separation of oil out of wastewater streams with high
organic contents and then burning the oil or recycling it back into the process. Recovery
would thus be performed on certain streams, prior to combination with other lower organic
content streams.
The World Bank and International Finance Corporation [7] propose a list of the most common
and appropriate approaches for treating refinery wastewater which include source
segregation and pre-treatment of concentrated wastewater streams. Typical wastewater
treatment steps include: grease traps, skimmers, dissolved air floatation or oil water
separators for separation of oils and floatable solids; filtration for separation of filterable solids;
flow and load equalisation; sedimentation for suspended solids reduction using clarifiers;

69

biological treatment, typically aerobic treatment, for reduction of soluble organic matter
(BOD); chemical or biological nutrient removal for reduction in nitrogen and phosphorus;
chlorination of effluent when disinfection is required; dewatering and disposal of residuals in
designated hazardous waste landfills. Thus none of the technologies presented in the EfWW
study are considered as suitable in this context.
When one considers the coal-to-liquid and gas-to-liquid refineries in South Africa more
potential for energy recovery from wastewater through traditional EfWW technologies is
identified. At PetroSA gas-to-liquids refinery in Mossel Bay an anaerobic digester with a 4.2
MWe (or approximately 12 MW thermal) biogas to electricity plant has been installed to
recover energy from their wastewater stream. It is likely that similar installations can be
considered for the Sasol coal-to-liquid plants. It is noted that the Sasol synfuels production is
approximately three times greater than that of PetroSA. If it is assumed that the energy
recovery potential is proportional to refinery output, the energy potential from Sasol
operations could be as high as 36 MW thermal.

Summary of the Case Study Outcomes


In summary, it was concluded that limited potential exists for the recovery of energy from
mixed wastewater streams from crude oil refineries in South Africa due to the dilute streams
encountered. Some potential might exist for stripping oil from the more concentrated streams
and recycling or burning, but that was not considered in the context of this study. The
potential for energy recovery from wastewater has already been demonstrated at a gas-toliquid refinery (PetroSA), and is likely to have a similar application at coal-liquids-refineries.
The thermal energy generated from PetroSA is 12 mW and the national potential could be as
high as 48 MW.

70

References
1. UAEU, (2006), United Arab Emirates University, Removal of Phenol from Refinery
Wastewater,

accessed

online

at

http://eclsun.uaeu.ac.ae/edc/archive/edc2006/tasks/task3.htm, March 2008.


2. Pearce, K. and Whyte, D. (2005), Water and Wastewater Management in the Oil Refining
and Re-refining Industry, WRC Report TT 180/05.
3. Al Zarooni M and Elshorbagy W., (2006), Characterization and assessment of Al Ruwais
refinery wastewater, Journal of Hazardous Materials A136, pp. 398405.
4. Alva-Argaez A., Kokossis A.C., Smith R., (2007), The design of water-using systems in
petroleum refining using a water-pinch decomposition, Chemical Engineering Journal 128, pp.
3346
5. Lenntech Petrochemical Company, Refinery Process Water, accessed online at
http://www.lenntech.com/petrochemical.htm, March 2008.
6. Beychok M.R., (1973), Aqueous Wastes from Petroleum and Petrochemical Plants, John
Wiley & Sons, London, pp. 33-112 & 156-266
7. World Bank and International Finance Corporation (2007), Environmental, Health, and
Safety

Guidelines

for

Petroleum

Refining,

accessed

online

at

http://www.ifc.org/ifcext/enviro.nsf/Content/EnvironmentalGuidelines, March 2008.


8. SAPIA

(2003),

Annual

Report,

Appendix

9,

accessed

online

www.sapia.co.za/pubs/2003_ARep/files/Sapia_2003_appendices.pdf, March 2008.

71

at

2.5: Survey of Animal Feedlots and Processing


Introduction and Approach
Animal rearing and processing operations were identified as being of relevance to this study
due to the high organic load wastes which are generated from these operations. The use of
organic wastes from this sector for energy recovery is commonplace around the world, both
on a rural or single household scale (such as in China) and for large scale animal processing
operations (such as in the United States). In the US alone, it is estimated that in 2007 alone,
approximately 215 GWh equivalent of energy was generated from liquid and solid wastes
collected at animal feedlots alone [1].
A combination of organic solid and liquid wastes are typically managed together for energy
recovery, and the combined waste stream available at a single location will likely allow for the
full potential for energy generation from this sector to be realised. Anaerobic digestion, the
energy recovery technology most likely to be used in this sector, will handle the combined
waste stream. The diagram below indicates the relationship between manure type,
characteristics and management options as a function of solids content.

Figure 2.5 Manure Characteristics that Influence Management Options [2]

Within the context of this study, which has a focus on energy from wastewater, a decision had
to be made as to whether to explore only the liquid/slurry proportion of this stream in terms of
its energy potential, or whether to extend this to include solid animal manure as well, given
that liquid and solid waste streams will likely be managed together. It was decided to first
express energy recovery potential of the wastewater stream on its own where information was
available (to be consistent with the project scope), and then, where relevant, extend this to
include animal manure.
In order to achieve the task of determining the overall potential for energy recovery from this
sector, the following information requirements were identified:

72

1.

Composition of wastewaters in terms of organic matter (COD, BOD and/or VS)

2.

Total wastewater generation from each of the components of the sector, calculated

through quantification of the number of animals held/processed in each component and the
wastewater generation per animal
3.

Solid waste generation per animal (where relevant)

4.

Potential for energy recovery from liquid and solid waste streams.

5.

Number of different types of animals in South Africa

One of the challenges in obtaining an accurate indication of the potential from this sector is
the variability in information presented in the literature. The following generic observations are
offered:

A number of factors influence the composition and volume of wastewater generated

in the sector. These include differences in farming/processing approaches, animal densities,


location, washing of areas, wastewater management, etc. Differences between waste
management at various sites give order of magnitude differences in composition.

Some literature sources report information on the waste generated by the animals

only (i.e. urine and faeces), others report the compositions of the waste stream after
management (e.g. washing of floors, etc. which has been diluted with wash water, and also
includes waste feed and other solids). This affects both volume and composition data.

The ability to capture waste and wastewater will be vastly different in feedlots, dairies,

etc. and in rural areas.

Even at individual locations the literature identifies variability in wastewater

compositions. At any one location, wastewater compositions there is seasonal variation in


compositions due to changing animal densities (for example due to calving or slaughtering)
and seasonal rainfall, as well as the time of day at which samples are taken to name a few.

Waste and wastewater production and composition of an animal are affected by, for

example if an animal is gestating or lactating.

Information on the number of animals in South Africa was difficult to obtain, with

different sources offering significantly different data. One reason offered for this observation is
that some estimates include rural animals and others do not.
In addition, there is a wide variability in information with respect to the potential for recovery of
organic matter from these streams for energy.
Given the variability in data, the approach taken in this case study was to use a selection of
approaches and data to provide the outer bounds of energy potential in the sector. More
detailed, in depth studies would help to provide more accurate data in this regard.
In the study the following components of the animal feedlots and processing sector were
explored:

73

Feedlots

Dairies

Piggeries

Chicken farming

Red meat abattoirs

Poultry abattoirs

Wastewater Composition
The table below presents the range of wastewater concentrations established through the
literature survey for the five sectors identified above.

Table 2.5.1 Range of COD, BOD, VS and TSS for Wastewaters from Various Animal Feedlots and Processing
Operations (mg/L)
COD

BOD

VS

TSS

References

Feedlots

No data found on wastewater compositions, combined solid and liquid waste used

Dairies

1500-9200

350-1600

255-830

250-600

5, 6, 7, 8

Piggeries

7141-14600

702-7790

3019-13051

5137-16965

9, 10, 11, 12

Red meat abattoirs

1100-21485

600-15510

3100-5600

13, 14

Poultry abattoirs

2360-11600

600-8700

640-1213

15, 16, 17, 18

600-1480

Although much data was found in the open literature on poultry litter from poultry breeding
facilities, limited data on the wastewater component of this stream in isolation could be found.
The reason offered for this is that breeding facilities are often lined with bedding material that
absorbs liquid and is periodically removed, or alternatively poultry manure falls onto the floor
where it is dried by the ventilation systems and removed as dry solid matter. Where cleaning
is done through flushing with water, the resultant wastewater stream is very dilute [2]. Hence
poultry breeding was considered in terms of energy potential from solid waste only.

74

Table 2.5.2 presents solid waste or manure compositions from animals in feedlots, piggeries
and poultry breeding facilities

Table 2.5.2 Organic Composition of Selected Solid Wastes (kg/kg wet solids)

Feedlots Cattle

BOD

VS

References

0.02667

0.1-0.106

19, 20

0.076-0.092

20

0.187-0.259

Piggeries
3

Poultry

Wastewater and Solid Waste Generation per Animal


The information provided above provides some guidance on the composition of wastes on a
weight basis. In order to translate this into overall availability of wastes in South Africa,
information on wastewater and solid waste generation per animal is required. Once again, a
wide range of values is found in the literature, a summary of this information appears in Table
2.5.3 and Table 2.5.4.
Table 2.5.3 Wastewater Generation per Animal
Animal

Wastewater production

Reference

(l/animal/day)
4

Dairies (l/animal/day)

11-45.3

2, 21

Piggeries (l/animal/day)

5.4-9

12, 23

Red meat abattoirs (l/animal)

90-1250

24,

Poultry abattoirs (l/animal)

19-38

28

In terms of solid waste generation, the amount of waste generated depends on the type and
the weight of animal Table 2.5.4

Assumed to be the same for dairies


Includes broilers and layers
4
Milk room + milking parlor + holding area, including manure washed out
3

75

Table 2.5.4 Solid Waste Generation per animal


Animal

Wet manure production

Reference

(kg/day)
Cattle

19, 20, 22

average

40-50

220 kg

13.2-18.7

300 kg

26-27

450 kg

36-38

Pigs

Poultry

1.7-26.2

20, 22

0.1

22

Number of Animals Reared and Slaughtered


The table below provides an estimate of the number of animals reared and slaughtered in
South Africa in the sectors studied.

Table 2.5.5 Animals Reared and Slaughtered Annually


Animal

Number (million)

Reference year

Reference

Cattle (includes beef and dairy cattle)

13.911

2007

32

0.621

1991

Pigs

1.65

2007

32

Chicken farming

559

2003

33

8.62

2003

33

545

2003

33

Dairy cattle

Cattle, sheep, lambs and pigs slaughtered (by


commercial operations)
Chickens slaughtered (by commercial operations)

The following further information and assumptions are made regarding cattle in estimating
energy recovery potential:
1.

The Department of Agriculture [35] estimates that 420 000 cattle are kept in feedlots

where the majority of their waste is likely to be able to be recovered. The remainder would
roam free in the veld in the day, and hence their wastes can only be captured at night when
5

Typically, the manure is 85 % to 90 % water [19]


Production by pigs is affected by whether they are lactating, gestating, nursing etc.
The reference source suggests that there are 136,716 dairy cattle in SA in the National Milk Recording and Progeny
Testing Scheme. They also estimate about 22% of cows participate in the scheme, so there are 621,436 dairy cattle
in the 5 main breeds in SA.
8
The reference source suggests that this was the total number of broilers slaughtered and layers in the industry as a
whole.
9
Note that the wastewater from rearing sheep and lambs was excluded from the study as they are typically in the
fields and collection of these wastes is thus impossible. Many will, however, be slaughtered in abattoirs, hence
their inclusion here.
6
7

76

they are placed in holding pens. It is assumed that half of the solid wastes from these cattle is
recovered, and that no liquid waste is recovered from cattle not in feedlots.
2.

It is assumed that the waste from dairy cattle can be recovered at night, as well as

during milking and holding in milking houses. For the remainder of the day they are allowed to
graze freely when waste cannot be recovered.
3.

Recovery of biogas from free range cattle (see below) is likely to be lower than that

from wastes from cattle in feedlots. In the absence of better information, however, the
recovery rates are assumed to be similar. This will likely result in an over estimate of this
information.

Recovery of Biogas
In this case study recovery of energy via Anaerobic Digestion is considered as for the
foreseeable future, this is the technology most likely to be used. As mentioned above, a
variety of information is available on energy recovery potentials. Typically, energy recovery is
reported on the basis of VS. The Table below presents a selection of the literature findings on
energy recovery potentials as used in this current study.

Table 2.5.6 Energy Recovery Potential from Wastewaters


Wastewater
Stream

Recovery Potential

Feedlot

0.2-0.3 m kg VS of biogas with a


methane content of 55 to 75%

-1

Dairy
Piggery

Comment

Reference source

Based on cow slurries


with a VS of 75-85% total
solids

26

Values
used.
-1

(i) 0.25-0.5 m kg VS of biogas with


a methane content of 70-80%,

for

feedlots

are

(i) VS of 70-80% of total


solids

(i) 26

(i) VS of 70-80% of total


solids

(i) 26

(ii) 30

(ii) 0.24 m CH4/kg COD


Poultry farming

-1

(i) 0.35-0.6 m kg VS of biogas with


a methane content of 60-80%
3

(ii) 31

(ii) 0.2 to 0.3 m methane per kg VS


3

Red Meat Abattoirs

0.18 m methane per kg COD in


wastewater

Poultry Abattoir

0.14 to 0.26 m methane/kg VS

Banks
(1994)
referenced in 31
Methane content assumed
to be the same as for
feedlots.

as

Based on a review of a
wide range of literature
provided by 31

Results and Conclusions


Using the data as compiled above, the ranges of energy recovery potentials shown in Table
2.5.7 were calculated for the wastewater and solid waste streams.

77

Table 2.5.7 Energy Recovery Potential from Wastewater, Solid Waste and Mixed Wastes
Stream

Potential

from

Liquid

Wastes (MWth)

Potential

from

Solid

Wastes (MWth)

Potential from Mixed


Waste Slurry (MWth)

Cattle in Feedlots (liquid +


solid)

79.1-214.5

Rural cattle

1270.9-3444.5

Dairies

0.1-2.3

116.8-118.4

116.9-120.7

Piggeries

2.0-33.2

16-681.3

18.0-714.5

Red meat abattoirs

0.2-48.9

Poultry farms
Poultry abattoirs

940.1-2976.1
0.6-7.0

A number of observations can be made from these results. Firstly, in the context of this
current study, energy recovery from wastewater in isolation from that recovered from solid
waste in the sector (as shown in the first column of the table) is not likely to make a significant
contribution to national or even local energy demand in dairies and poultry abattoirs. Solid
and liquid waste is likely to be managed together in most technologies.
Potential for energy recovery exists in centralized operations, including feedlots, dairies,
piggeries and abattoirs. On-site energy recovery systems on small to medium feedlots, dairies
and piggeries are likely to provide at a minimum onsite energy, and larger operations will have
the potential to export energy.
Although solid waste collected from corralled rural cattle is suggested to represent the most
significant potential for energy recovery in this sector, the practicalities of energy recovery
from this source are significant. This is both in terms of collection (manual collection of the
solid wastes from the kraals would be required in the mornings), and in terms of requirement
for a rollout of small scale digesters close to the source of the wastes. Hence the proportion of
this energy potential which could realistically be recovered would be significantly lower than
that shown in the table.

78

References
1.

United States Environmental Protection Agency (US EPA), AgSTAR Program,

accessed online at http://www.epa.gov/agstar/, February 2008.


2.

US EPA, Development Document for the Proposed Revisions to the National

Pollutant Discharge Elimination System Regulation and the Effluent Guidelines for
Concentrated Animal Feeding Operations. Chapter 6: Wastewater Characterization and
Manure Characteristics, accessed online at www.epa.gov/guide/cafo/pdf/DDChapters5-7.pdf,
February 2008.
3.

J.L. Hatfield, B.A. Stewart, Animal Waste Utilization: Effective Use of Manure as a Soil

Resource, CRC Press 1998


4.

J.M. Sweeton, Manure and Wastewater Management for Feedlots, Reviews of

Environmental Contamination and Toxicology, 167:121-153


5.

http://www.ctahr.hawaii.edu/wwm/waste.asp, accessed February 2008.

6.

N. Schwarzenbeck, J. M. Borges and P. A. Wilderer, Treatment of dairy effluents in an

aerobic granular sludge sequencing batch reactor, Volume 66, Number 6, March, 2005
7.

B. Demirel, O. Yenigun, and T.T. Onay, Anaerobic treatment of dairy wastewaters: a

review, Process Biochemistry Volume 40, Issue 8, July 2005, Pages 2583-2595
8.

B. Sarkar, P.P. Chakrabarti, A. Vijaykumar, Vijay Kale, Wastewater treatment in dairy

industries possibility of reuse, Desalination 195 (2006) 141152


9.

J-H Shina, S-M Leeb, J-Y Junga, Y-C Chunga and S-H Noh, Enhanced COD and

nitrogen removals for the treatment of swine wastewater by combining submerged membrane
bioreactor (MBR) and anaerobic upflow bed filter (AUBF) reactor, Process Biochemistry
Volume 40, Issue 12, December 2005, Pages 3769-3776
10.

E. Snchez, R. Borja,L. Travieso, A. Martn and M. F. Colmenarejo, Effect of influent

substrate concentration and hydraulic retention time on the performance of down-flow


anaerobic fixed bed reactors treating piggery wastewater in a tropical climate, Process
Biochemistry, Volume 40, Issue 2, February 2005, Pages 817-829
11.

J. A. Oleszkiewicz, A comparison of anaerobic treatments of low concentration

piggery wastewaters, Agricultural Wastes 8 (1983) 215-231


12.

M. B. Vanotti, A.A. Szogi, P.G. Hunt, P.D. Millner and F.J. Humenik, Development of

environmentally superior treatment system to replace anaerobic swine lagoons in the USA,
Bioresource Technology Volume 98, Issue 17, December 2007, Pages 3184-3194
13.

O.S. Amuda, A. Alade, Coagulation/flocculation process in the treatment of abattoir

wastewater, Desalination 196 (2006) 2231


14.

NT Manjunath, I Mehrotra, RP Mathur, Treatment of wastewater from slaughterhouse

by DAF-UASB system, Wat. Res. Vol. 34, No. 6, pp. 1930-1936, 2000

79

15.

U. Tezel, J.A. Pierson, S.G. Pavlostathis, Effect of polyelectrolytes and quaternary

ammonium compounds on the anaerobic biological treatment of poultry processing


wastewater, Water Research 41 (2007 ) 1334-1342
16.

R. del Pozo, V. Diez and S. Beltrn, Anaerobic pre-treatment of slaughterhouse

wastewater using fixed-film reactors, Bioresource Technology Volume 71, Issue 2, January
2000, Pages 143-149
17.

C. Chvez P., R. Castillo L., L. Dendooven and E.M. Escamilla-Silva, Poultry

slaughter wastewater treatment with an up-flow anaerobic sludge blanket (UASB) reactor,
Bioresource Technology, Volume 96, Issue 15, October 2005, Pages 1730-1736
18.

V. Del Nery, I.R. de Nardi, M.H.R.Z. Damianovic, E. Pozzi, A.K.B. Amorim and M.

Zaiat, Long-term operating performance of a poultry slaughterhouse wastewater treatment


plant, Resources, Conservation and Recycling, Volume 50, Issue 1, March 2007, Pages 102114
19.

Queensland Department of Primary Industries and Fisheries, Feedlot Waste

Management

1.

Manure

Production

Data,

accessed

online

at

http://www2.dpi.qld.gov.au/environment/5166.html, February 2008.


20.

US EPA AgSTAR Program, FarmWare 3.0 Users Manual Appendix C, accessed

online at http://www.epa.gov/agstar/pdf/handbook/appendixc.pdf, February 2008.


21.

S.H. Christopherson, D.Schmidt, K. Janni , Evaluation Of Aerobic Treatment Units In

Treating

High

Strength

Waste

From

Dairy

Milk

Houses,

accessed

online

at

http://www.pirana.biz/downloads/Minnesota%20Evaluation%20Of%20Aerobic%20Treatment
%20Units%20In%20Treating%20High%20Stren.doc, March 2008.
22.

Scorecard the Pollution Information Site, Estimating Waste Produced by Livestock

Operations, accessed online at http://www.scorecard.org/env-releases/aw/us.tcl, February


2008.
23.

J.A. Lory, J. Zulovich and C. Fulhage, Hog Manure and Domestic Wastewater

Management Objectives, accessed online at http://www.thepigsite.com/articles/8/biosecuritydisinfection/1938/hog-manure-and-domestic-wastewater-management-objectives,

March

2008.
24.

D.I. Mass and L. Masse, Characterization of wastewater from hog slaughterhouses

in Eastern Canada and evaluation of their in-plant wastewater treatment systems, Canadian
Agricultural Engineering Vol. 42, No. 3 July/August/September 2000
25.

L.K. Wang, Waste Treatment in the Food Processing Industry, CRC Press, 2006.

26.

R. Steffen,; O. Szolar, and R. Braun, Feedstocks for Anaerobic Digestion, University

of

Agricultural

Sciences

Vienna,

accessed

online

http://homepage2.nifty.com/biogas/cnt/refdoc/whrefdoc/d8feed.pdf, March 2008.

80

at

27.

K. Krich, D. Augenstein, JP Batmale, J. Benemann, B. Rutledge and D. Salour,

Biomethane from Dairy Waste: A Sourcebook for the Production and Use of Renewable
Natural Gas in California, Report prepared for Western United Dairymen, July 2005 accessed
online

at

http://www.westernuniteddairymen.com/Biogas%20Fuel%20Report/Biomethane%20sourcebo
ok.pdf, March 2008.
28.

Kiepper B.H., (2003), Characterization of poultry processing operations, wastewater

generation, and wastewater treatment using mail survey and nutrient discharge monitoring
methods,

Masters

Thesis,

University

of

Georgia,

accessed

online

at

March

http://outreach.engineering.uga.edu/publications/kiepper_brian_h_200308_ms.pdf,
2008.
29.

B.S. Magbanua, Jr., T.T. Adams and P. Johnston, Anaerobic codigestion of hog and

poultry waste, Bioresource Technology, Volume 76, Issue 2, January 2001, Pages 165-168
30.

J Cheng, B Liu, Swine Wastewater Treatment in Anaerobic Digesters with Floating

Medium, Transactions of the ASAE, Vol. 45(3): 799805, 2002


31.

E. Salminen and J. Rintala, Anaerobic digestion of organic solid poultry

slaughterhouse waste a review, Bioresource Technology Volume 83, Issue 1, May 2002,
Pages 13-26.
32.

Directorate: Agricultural Statistics, Monthly Food Security Bulletin, November-

December

2007,

accessed

online

at

http://www.nda.agric.za/docs/Cropsestimates/RSA%20Food%20Security%20Bulletin%20Nov
-%20Dec%2007.Final.doc, March 2008.
33.

Department of Agriculture, Animal Production Trends,

accessed

online

at

http://www.nda.agric.za/docs/trends2003/animal_production.pdf, March 2008.


34.

accessed

online

at

http://agriculture.kzntl.gov.za/portal/Publications/ProductionGuidelines/DairyinginKwaZuluNat
al/BreedsofDairyCattle/tabid/239/Default.aspx, March 2008.
35.

National Department of Agriculture (2007), The Value Chain for Red Meat: Chapter 4,

accessed online at http://www.nda.agric.za/docs/fpmc/Vol4_Chap4.pdf, March 2008.

81

2.6 ENERGY FROM THE FRUIT INDUSTRY WASTEWATER


Introduction and background
Energy recovery from fruit wastewater in South Africa has, to date, been relatively poorly
exploited and presents a potential opportunity for a source of renewable energy. The
development of suitable technologies for the conversion of waste to energy in the industrial
fruit processing sector can significantly benefit the industry. The aim of the case study was to
investigate the potential for energy recovery from wastewater streams generated by the
industrial fruit processing sector. The study was based in Cape Town and the main focus of
the project was fruit processing activity in the Western Cape where there are a number of
major fruit production and processing companies.
The study included a review of the fruit processing sector in South Africa focusing on the
Western Cape, and a review of techniques currently used to treat the wastewaters from fruit
processing. Technologies available for treatment were also reviewed and the two most
suitable were selected for experimental investigation. Samples of wastewaters from a
selection of local fruit processing plants were obtained and characterised to provide data for
the determination of energy recovery potential. Finally, technical and economic feasibility
were considered.

Overview of fruit production and processing in the Western Cape


The Western Cape agricultural sector generates about 23% of the total value added to the
agriculture sector of South Africa. The province has ideal conditions for both deciduous and
citrus fruit growing and fruit farming in the province was valued at R2.4 billion in 2005. In
2002/2003 approximately 1.46 million tonnes of deciduous fruit was produced, of which
approximately 20% was processed into fruit juice and about 30% was canned and dried.
(Nationally, the annual production was approximately 1 900 000 tonnes). In 2003/2004 the
Western Cape produced approximately 370 000 tonnes of citrus fruit, of which approximately
20 % was processed into juice. (Annual national production of citrus fruit in 2005/2006 was 2
100 000 tonnes).
The major types of fruit processing in the province are canning, juicing, wine making, and fruit
drying. Approximately 205 000 tonnes of deciduous fruit were processed by the canning
industry in 1999/2000. The typical water consumption for canned fruit processing and fruit
3
3
juicing is approximately 7 m /t and 10.7 m /t of raw produce, respectively. The wastewater

from these processes usually contains suspended solids, particulate organics, as well as
various cleaning solutions and softening or surface-active additives (e.g., sodium hydroxide,
nitric and phosphoric acids at low concentrations).
Fruit juicing typically yields solid residues to approximately 50 %wt of fruit used, and large
volumes of water (up to 10 m3/tonne raw fruit) are used largely for cleaning. This water

82

contains organics including carbohydrates, and can, of course, be mixed with the solid
residues for energy recovery purposes.
In the wine industry it is estimated that water usage of between 0.7 and 3.8 m3/tonne of
grapes processed (0.8-4.4 L/L of wine produced) is typical. The wastewater from the wine and
spirits industry usually contains a high organic load, suspended solids and dissolved solids.

Current treatment methods


The wastewaters from fruit canning and juicing are usually treated by means of land irrigation,
lagooning or treatment in an activated sludge process. With respect to lagoon technologies,
anaerobic lagoons require an input of approximately 100 g BOD/m3/d (Sigge, 2005) and must
operate at temperatures above 10C. The methane generated can be utilised for energy.
Facultative lagoons (stabilisation ponds) can be used to treat wastewaters with BOD levels in
the order of 5 g BOD/m3/d. Activated sludge treatment, may be problematic due high
carbohydrate loadings of fruit wastewaters, and the resulting in high microbial growth and
poor settling (Sigge, 2005; Casey, 1999). Anaerobic digestion (AD) is common and
successful as a treatment method; this has been widely investigated
Land irrigation is feasible for low COD wastewaters, or, after dilution, higher COD
wastewaters. More than 95% of South African wineries currently irrigate their wastewater onto
land by means of sprinkler systems, which, as with cannery wastewater, is an effective means
of lowering the COD of the effluent. Composting and landfill disposal are also common but do
not offer any opportunities for energy recovery.

Estimation of energy potential from fruit processing wastewater


A feasibility study was carried out to assess the potential for energy recovery from wastewater
generated by the industrial fruit processors in the Western Cape. The waste streams
generated from specific fruit processors in the province were characterised and classified,
and the potential for ethanol production by means of fermentation and the production of
biogas by means of anaerobic digestion were investigated using the data obtained from the
waste stream classification as the basis for determining energy recovery. Fruit juicing and
canning operations are usually seasonal, operating for about one third of the year, starting in
February/March. The wastewater streams from the juicing processes typically contain a sugar
content of about 10 g/L at a COD of 15000 mg/L. Fruit pomace from the juicing processes,
containing about 70% moisture, is also generated during production and also provides a
potential source of fermentable carbohydrates. The pomace typically contains 5 to 6 wt%
sugars. The wastewater streams from a typical canning operation have a sugar content of
about 3 g/L at a COD of 4400 mg/L.
An experimental investigation into the feasibility of ethanol production by means of yeast
fermentation showed that approximately 6 L or 4.7 kg of ethanol can be produced per ML of

83

wastewater which is equivalent to 101 MJ of energy per ML of wastewater (de Beer and
Dawson, 2007). The biogas yield from anaerobic digestion (AD) of the wastewater is
approximately 3750 L or 2.5 kg of methane per ML of wastewater which has an energy value
of 115 MJ per ML. The fermentation process converts the sugars present in the wastewater
streams into ethanol, while anaerobic digestion converts a greater fraction of the organic
material in the stream into biogas and thus yields more energy per volume than the
fermentation process. Synthetic apple waste was used as the starting solution for the
fermentations, for purposes of simplification. In comparing fermentation with anaerobic
digestion, substrate solutions with a high sugar (8-9 wt%) and COD (250 000 mg/L)
concentration, and a low sugar (0.8-1 wt%) and COD (15000 mg/L) concentration,
respectively, were used. The dilute fruit waste samples used here were similar to the most
concentrated wastewater streams produced by two local fruit processing plants. In anaerobic
and aerobic fermentations yeast (Saccharomyces cerevisiae) was used and in anaerobic
digestion, sludge was used as the inocculum. From the results it is clear that aerobic
fermentation has a larger COD reducing potential than that of anaerobic fermentation. The
greatest ethanol production (5.8 wt% ethanol) was obtained by the anaerobic fermentation of
the concentrate fruit waste sample. The sludge fermentations performed better than was
expected but still only produced 2.2% ethanol which is considerably less than the yeast
fermentation and not considered viable. It was concluded from the experimental results that
some form of waste stream concentration would have to be performed to increase the sugars
concentration to make ethanol production more viable.

Consideration of feasibility
Although both fermentation and AD technologies are suitable for the recovery of energy from
fruit processing wastewater, the viability of each is very case specific and factors such as
waste COD, sugar concentration, and volumes need to be considered. It is also important to
consider that biofuel production from fruit waste must compete commercially, with not only
other methods of well established bioethanol production, but also producers of regular
petroleum products, which currently operate at large production capacities and typically have
a well established market share. However, further developments and increased interest in
biofuel production could lead to the commercialisation of more large-scale plants, with higher
production capacities. Economies of scale would thus lower capital expenditure and increase
profitability per volume of biofuel produced. The implementation of new technologies, such
membrane concentration of wastewater streams, could also lead to increased economic
feasibility of biofuel production.

84

2.7

Thermal gasification or anaerobic digestion of biomass for use in a distributed

renewable energy network.


This study compares thermal (gasification) and biological (anaerobic digestion) process
routes for producing a combustible gas from biomass for use in a distributed, renewable
energy network. The global hypothesis of the study states that based on the composition and
physical properties of a particular biomass, so one route will be favourable over the other.
Furthermore, the following objectives have been identified. a.) Identify operational problems
or technical barriers in each process b.) Identify threshold limits, based on composition, to
guide a decision maker as which process to consider c.) Develop process designs using the
above algorithm d.) Evaluate the different designs on the basis of energy yield and exergy
The feedstocks which will be evaluated are: corn stover, wood chips, indigenous grasses and
algal or sewerage sludge. The feedstocks provide a range of compositional variation and
some are predisposed to gasification (wood chips), others to anaerobic digestion (the
sludges) and choice of process for the corn stover and grass is not immediately clear.
Gasification occurs in an atmosphere of sub-stoichiometric oxygen requirement. Biomass
particles are heated up and release volatile their volatile components. This is known as
pyrolysis or devolatilisation. Occurring sequentially, or simultaneously, depending on heating
rate, the carbon in the carbon/ash skeleton is consumed via combustion and gasification
reactions leaving the ash behind. The gas that is produced is a low (5 MJ/Nm3) to medium (15
MJ/Nm3) calorific gas depending on whether air or steam is used as a gasifying agent.
Anaerobic digestion is the conversion of organic material to a mixture of mainly methane and
carbon dioxide via the activity of three classes of micro-organisms, viz. aerobic hydrolytic
fungi or anaerobic hydrolytic bacteria, anaerobic acetogenic bacteria and finally anaerobic
methanogenic Archaea. The operational problems in gasification were identified as bed
agglomeration and tar formation. Simple methods for predicting biomass ash fusion
temperature, as has been done for coal gasification, were not found. Tar formation is not
necessarily a function of the biomass composition, but rather depends on reactor design. Inbed and post-bed tar reforming methods, especially when both used, can successfully crack
tars to produce a gas usable in gas turbines.
The success of anaerobic digestion of lignocellulosic materials is dependant on an effective
pre-treatment. Characteristics of a potential process must improve the formation of sugars or
the ability to subsequently form sugars by enzymatic hydrolysis; avoid the degradation or loss
of carbohydrate; avoid the formation of byproducts inhibitory to the subsequent hydrolysis and
fermentation processes and must be cost-effective. Processes fitting these criteria include:
dilute acid hydrolysis, hot water hydrolysis, steam explosion, Ammonia Fibre Explosion
(AFEX), Ammonia Recycle Percolation (ARP), Soaking in Aqueous Ammonia (SAA) and lime.

85

Lime and air pretreated biomass exhibits up to 93% conversion from cellulose to glucose and
operates at ambient conditions. The only caveat is its slow rate (2 weeks to a month).
The methodology for this project will be as follows:
1. Identify, from the literature and from first principles, critical limits of operation
2. Develop If, then, else algorithm using these critical limits
3. Use the algorithm to create a process flowsheet
4. Assess the flowsheet using exergy analysis
5. Validate the choice of unit operation experimentally
Some aspects of the decision algorithm have already been investigated, viz. moisture
content, lignin content and ash content. It has been shown that a moisture content higher than
60% precludes the biomass feedstock from gasification as the energy obtained from the fuel
gas is less than the energy consumption of a drier. An ash content of 1 kg alkali/GJ was
identified as a threshold limit for sound gasification operation. Preliminary work on the
threshold limit for lignin content did not yield usable results as the relationship was linear due
to the assumed constant cellulose conversion. Further investigation into how to represent the
effect of increased lignin content on cellulose conversion is needed.
In conclusion, the applicability of gasification and anaerobic digestion are dependent on the
qualities of the feedstock. This report sets the basis for further work to gain a deeper
understanding of the interaction of the various properties of biomass feedstock and ultimately
propose a decision making framework.

86

2.8. The sludge drying process at Cape Flats Treatment Works (CFTW): Is it best
practice?

87

Introduction
The Cape Flats Treatment Works (CFTW), owned and operated by the City of Cape Town, is
the only such works in the country to have installed a sludge drying process. The primary
stabilised sludge from the anaerobic digesters are sintered into pellets at temperatures of
500C using biogas from the anaerobic digesters as fuel. Since these pellets can be used as
an alternative/additional fuel in the cement industries, this illustrates two instances of energy
capture and use from wastewater.
With the kind support of the Wastewater Department of the City of Cape Town, Ms. Loice
Badza, a final year Chemical Engineering student, undertook her research project on a
technical, environmental and institutional evaluation of the CFTW sludge drying process.

Hypothesis and Methodology


The research was guided by the following hypothesis: The choice of using the biogas for
sludge drying and subsequently using the sludge pellets as a combustion fuel at PPC cement
works represents best practice for the Cape Flats WWTP.
Technical, environmental and institutional evaluations were carried out. A flowsheet analysis
and a material and energy balance completed the technical evaluation. Aspects of the mass
and energy balance investigated include: CH4 production output, sludge production, pellet
production and fuel use. The environmental assessment considered indicators such as GHG
emissions, the cost of dumping waste and electricity and heat generation and .was based on
qualitative data gathering from interviews and quantitative data from the mass and energy
balance. A set of interviews with plant managers at the plant provided information to complete
the institutional evaluation.

Results
The CFTW sludge drying project was commissioned in 1999 and is run in the form of a public
private partnership; with the sludge drying plant built and operated by a joint venture of
Biwater with Murray and Roberts. One of the major reasons for its construction was that the
practice of on-site sludge burial was causing the seepage of nutrients into the adjacent
Zeekoevlei wetlands causing eutrophication. The drying plant uses the Swiss Combi Drying
system (www.swisscombi.com). A block flow diagram of the drying plant as it fits into the
treatment works is shown in Figure 1.

88

a) Technical evaluation
The flow diagram covers CFTWs sludge handling process only. The solid lines represent
material flow and the dotted lines, heat networks.
Biogas

Gas
Holder

Primary sludge
P-40
Anaerobic
Digesters

Thickening
centrifuges

Heat exchanger

A
Gas
boiler
C

DAF sludge
water

Stabilized sludge

Dewatering
centrifuges

Gas
burner

Rotary drier

Pellets

water

Sludge cake to landfill

Figure 2.8.1 : Simplified diagram of the Energy from wastewater project at CFTW

The heat network system is described below.


A: This is the main use of biogas currently produced. There are three anaerobic digesters,
each were designed to produce 400 m3/h of biogas at 65% methane. Gas flow-meters
commissioned in September 2007 reported gas flow rates of 140 m3/h, and the biannual
methane content of the biogas was reported as 55% methane (June 2007).
B: Because the plant is currently not producing enough biogas, no biogas is used to fuel the
gas burner. However, in the past, this path was viable.
C: For optimum digester performance, pre-heating the sludge is carried in double pipe heat
exchanger, with sludge on the inside and warm water (heated by the boiler) on the outside.
D: Potential for heat recovery from the sludge drying process involves use of warm water (~
70oC) to pre-heat the sludge.
E: Diesel fuel is currently used in the burner. Between October 2006 and September 2007, no
biogas was used for drying and a total of 493 kilolitres of diesel were used.
Figure 2 gives an idea as to how functional the drying plant has been for the period of
October 2006 to September 2007. Sludge from the dewatering centrifuges is about 20%
solids, whilst the pellets are 95% solids, thus every cubic meter of sludge would reduce to

89

roughly 0.2 m3 of pellets. It can be seen that during this review period, there was always a
portion of the sludge that could not be dried and that both the overall amount of sludge to be
handled, and the fraction actually dried, show significant variability.
Sludge handling overview
sludge cake

1000

Pellets

800

Volume, m /mth

1200

600
400
200
0
Oct-06 Nov-06 Dec-06 Jan-07 Feb-07 Mar-07 Apr-07 May-07 Jun-07 Jul-07 Aug-07 Sep-07
time, month
Figure 2.8.2 : Amounts of dewatered sludge and dried pellets disposed

In order to run a plant optimally and efficiently, measurements of key operating variables are
required. The table below reports what is actually measured at the CFTW related to the
sludge drying operation.
1. Primary sludge

Flowmeters working
Compositions measured via sampling analysis.
o

COD measured is for supernatant return to 2 treatment only


2. DAF sludge

No flowmeters present.
Compositions measured via sampling analysis

3. Thickened centrate

Flowmeters working; DS and TSS content measured

4. Digester 1 feed

Flowmeters working; DS content measured; COD not measured

5. Digester 2 feed

Flowmeters working; DS content measured; COD not measured

6. Digester 3 feed

Out of commission

7. Biogas to gas holder

Flowmeters installed September 2007; methane content measured biannually

8. Dewatered centrate

Flowmeters working; DS and TSS measures

9. Mixed dewatered stream

No flowmeters

10. Sludge cake

Volume estimated using truck maximum capacity

12. Pellets product

Volume estimated using truck maximum capacity and maximum bag volume

The COD measured at the plant are the influent (plant load), primary treatment liquid effluent
(to 2o treatment), secondary treatment effluent and final effluent (to maturation ponds).

90

b) Environmental evaluation
The sludge drying plant was built to protect the local environment, esp. Zeekoeivlei, from
leachate generated from sludge buried on-site. The operation of the CFTW sludge drying
facility reduces this environmental burden, although there are months in which the drying
plant is not sufficiently operational to dry all of the sludge.

This primary environmental benefit does however require significant energy usage. At best,
this energy is provided from the wastewater treatment process in the form of biogas from the
anaerobic digesters, in the process transforming methane, a potent greenhouse gas, to the
less potent greenhouse gas carbon dioxide. The project was, however, not designed to utilise
all methane gas, leaving a significant portion (400 m3/h) of biogas, to be vented. In recent
practice, however, since insufficient quantities of biogas are currently being produced (140
m3/h), all of this was used for digester preheating. This meant that the drying plant was run on
a fuel known as LO 10, supplied by FFS Refineries, at about 50% of the price of Diesel). As
this is also an energy product recovered from waste materials, the resulting environmental
issues of resource depletion and greenhouse gas emissions are of secondary importance.

c) Institutional evaluation
The Cape Flats Wastewater Treatment Works was commissioned in 1980. The new
technology, sludge dewatering and drying plant, came on line in 1999. Biwater-Murray and
Roberts JV was awarded the tender to build the plant and also to operate and maintain it. The
plant is owned by the City of Cape Town.

The main recipient of the pellets was a large cement manufacturing company, PPC, who used
the pellets as an energy source to replace coal in their combustion kilns. However, the lack of
a formal contract between the CFTW and PPC confirming consistent supply of pellets was a
drawback to both institutions. Since the CFTW was not obliged to give away the pellets to
PPC only, they also supplied farmers in the Philippi area with the pellets to use as fertiliser
(This is not an agreed or contractual practice but more of an ad hoc basis and only if there are
stockpiles on site). PPCs level of commitment was so high that they were prepared to pay for
guaranteed supply of the fuel pellets, but the income from such secondary industries are not
permitted (as determined by governmental regulations). PPC stopped using pellets to fire their
cement kilns because the quantities were erratic and the volumes insufficient for their needs
and CFTW presently have a contract with Wasteman to remove all sludge produces to the
Vissershok landfill site. The sludge still requires drying to about 95% solids content in
accordance with the Biwaters contract with the WWTP.

91

Conclusions
This investigation of the sludge drying practice at the Cape Flats Treatment Works from an
energy from wastewater perspective has yielded the following insights:
1. Although the CFTW process can be regarded as a case of energy from wastewater since it
can utilise both the biogas from the anaerobic digesters and the dried sludge pellets as fuel,
the project was not conceived as an energy project but as an environmental protection
project.
2. The energy and climate change aspects of the sludge drying process can still be
significantly improved.

The CFTW was designed to produce 400 m3/h of biogas at 65% methane, but recent

analysis indicated only 140 m3/h at 55% methane (June 2007). This indicates that the system
is operating at only about 30% efficiency.

The sludge drying operation is significantly cheaper when it can harness biogas fuel

from well-functioning digesters which was not the case for the review period, where almost
half a million litres of diesel fuel were used over a 12 month period.

If the AD system were functioning optimally, the biogas generated could be used to

supply the CFTW with the majority of its energy requirements (2.45 gWh).
3. The opportunity of harnessing the dried sludge as a replacement energy product was lost
because the plant is not permitted to generate revenue from pellet production.

92

2.9: Case study of household AD biogas systems


Introduction
The standard, centralised water-borne sewage system for the disposal of domestic
wastewaters is that they are unsustainable and represents lost opportunities in recovering the
energy and water. This system flushes pathogenic bacteria out of the residential area, using
large amounts of water and thereby transfers a concentrated domestic health problem into a
diffuse health problem for the entire settlement and/or region. Energy input is required to treat
the wastewater at the centralised wastewater treatment plant, where the cost of treatment
increases as the volume of wastewater increases. The centralised municipal wastewater
treatment plants are also particularly costly to deploy in rural or peri-urban settings, making
the provision of wastewater treatment services to the South Africans who do not have access
to water-borne services (approx 40% of the population) a huge challenge

History of household AD biogas


Theres evidence that biogas was used to a heat bath water in Assyria during 10 BC, but the
first recorded digester to produce biogas from wastes was built in a leper colony India in
1859. China began a mass adoption of biogas in 1975 under the slogan biogas for every
household. Within the first few years, 1.6 million digesters were constructed annually, but
these were of low quality and by 1980, half of all digesters were not in use and the rate of
adoption had slowed. By 1992, only 5 million family sized plants were still operating, many of
them redesigned to avoid leakage. The technology has continued to be developed and
implemented on a large scale and in 2005 China had 17 million digesters with annual
3
production of 6.5 billion m biogas. This supplied approximately 50 million people biogas fuel

for sufficient for their daily household needs. The government aims to reach annual
production of biogas is projected to reach an annual production of 25 billion m3 biogas by
2020. Importantly, biogas provides energy to one quarter of households in rural areas. Similar
developments took place in India and by 2000 more than 2 million biogas plants were built.
However, there were several problems with poor design, lack of maintenance skills and
insufficient capacity to deal with the problems meaning that 1/3 were non-operational.
Lessons have been learnt from these early years the government is aiding a new
development and implementation program. There are also recent success in countries such
as Nepal, Sri Lanka, Philippines and Vietnam.

There is a great opportunity for developing countries since there is a constant available
supply raw material (animal dung and human domestic wastewaters), the technology is
relatively simple and robust, and it requires relatively small investments. The technical viability
of biogas technology has been repeatedly proven in many field tests and pilot projects but
numerous problems arose with mass implementation (Karekezi, 1994) and biogas has been

93

implemented only to a limited extent in most African countries. The reasons for the lack of
development of this technology are that the investment cost of even the smallest of the biogas
units is prohibitive for most rural households of sub-Saharan Africa (e.g., a family size plant
cost US$1667 beyond the means of rural farmers), immature technical properties of plants
themselves, a lack of government commitment (policy) and dissemination strategy
(inadequate user training and follow-up services), and a limited private sector input because
of low profit incentive (Amigun and Von Blottnitz, 2007).

Benefits of biogas technology.


Individuals judge the profitability of biogas plants primarily from the monetary surplus gained
from utilising biogas and biofertiliser in relation to the cost of the plants. In terms of energy
value, 1 m3 Biogas (Approx. 6 kWh/m3) is equivalent to:
Diesel, Kerosene (approx. 12 kWh/kg) 0.5 kg
Wood (approx. 4.5 kWh/kg) 1.3 kg
Cow dung (approx. 5 kWh/kg dry matter) 1.2 kg
Plant residues (approx. 4.5 kWh/kg d.m.) 1.3 kg
Hard coal (approx. 8.5 kWh/kg) 0.7 kg
City gas (approx. 5.3 kWh/m3) 1.1 m

Propane (approx. 25 kWh/m3) 0.24 m

In practice, 1 m3 of biogas can cook three meals for a family, generate 1.25 kWh of electricity,
or power a 1hp internal combustion motor for 2 hours.
Biofertiliser slurry can be quantified as a monetary benefit, but this depends largely on the
previous practice of the raw material to be digested. In agricultural settings the dung and
fodder residues are often heaped in the open, leading to heavy losses of minerals through
sun and rain.
The reduced disposal costs can be counted as benefits of a biogas system where the
disposal of waste and wastewater is regulated by law and where disposal opportunities exist.
In rural households, human faeces are collected in pit latrines and once the pit latrines are full
(every two years), they are filled with soil and a new pit is dug. Excavation costs and
subsequent disposal costs can be saved and calculated as benefit. Similarly, if a septic tank is
used, the emptying cost can be counted as benefit. In the city or town setting, the reduced
loads discharged to the municipal WWTP (and the reduction in costs incurred) could also be
counted as benefits if the appropriate levies existed.
The main environmental benefits are that a renewable source of energy is captured. The use
of a renewable energy reduces the CO2 emissions through a reduction of the demand for
fossil fuels or firewood. There are also important benefits of methane mitigation since
methane is the second most important greenhouse gas contributing to global warming.

94

Lastly, applying AD biogas technology to the treatment of industrial or municipal wastewater,


can be protect water resources (rivers, sea, ground and drinking water resources). The
effluent wastewater can often be reused after appropriate polishing, (e.g. as process water in
industry or as irrigation water in agriculture) so the costs avoided in using potable water can
be directly translated into benefits.

Technical performance
Several anaerobic digesters (AD) for biogas production have been installed by
(http://www.agama.co.za/). The aim of this study was to assess the performance of three
small (<10 m3) anaerobic digesters that are being used in households in the Western Cape.
Key performance indicators were the biogas produced (flow rate and %methane) and the
reduction in wastewater organic load (COD, nutrients, and coliform bacteria).
A

Fig. 2.9.1: Structure of a typical Anaerobic Digester

AD at Noordhoek (Cape Town) and Stanford (Overberg) were assessed. The sizes and
loading capacity of the ADs are shown below, together with the predicted biogas production
(m3/day).

Table 2.9.1: Characteristics of ADs


AD digester

AD volume

Type and description of load

Predicted

Predicted
optimal

(m3)

biogas

production

energy

optimal
produced

kWh per day *

(m3/day)
Noordhoek

(Thor)

Domestic (3 persons) greywater and

3.0

18

2.3

14

3.8

23

blackwater + horse manure 2 wheel


barrows

Noordhoek

(Richmond)
Stanford Valley

Domestic

greywater

and

blackwater(2-6 persons domestic)


10

Domestic backwater (6-20 persons)


3

* The predicted biogas production is 0.38 m /day/m biodigester. Biogas has a heat energy value of approx. 6 kWh/m

95

Three separate site visits were carried out in order to take samples of wastewater and biogas.
These were 5 November, 11 December 2007 and 25 January 2008. The time interval was
chosen to accommodate the large residence times in the digester (20-40 days) and the
seasonal fluctuations (may affect performance through changes in temperature) as well as to
reliably asses performance over a time period. The experiments assessed the performance of
the AD digester alone within the complete decentralised wastewater treatment system
(DEWATS). Other components of the system for the further treatment of the wastewater
include an anaerobic baffled reactor (ABR), a constructed wetland (plant reed-bed) and an
algal polishing pond. At Stanford Valley, the effluent wastewater from the AD is passed onto a
reed bed and then onto an algal pond before it is discharged into the fields to irrigate pasture.
A more recent installation in the town of Stanford, uses an anaerobic baffled reactor prior to
the plant reed bed, but this system was not assessed since it has only recently been installed
(September 2007).

Methods
A syringe (50 ml sample) of biogas and 200 mL sample of wastewater were taken and
transported back to the laboratory for analysis. The biogas was analysed by isothermal gas
chromatography (GC) at 140C. A standard of 9.7% methane was used for calibration and the
% methane in the biogas determined from relative peak areas. The biogas flow rate was
estimated in a rudimentary fashion using a simple 1 kPa pressure gauge. Owners were
instructed to place the gauge on the gas outlet after a period of cooking (usually at the end of
the day), measure the pressure (P1) and then measure the pressure again (P2) after a time
period of several hours. The flow rate m3/day was calculate using time interval and
P1V1=P2V2. To determine the number of coliform bacteria in the provided wastewater sample.
The test was carried out as described by McCarthy, Delaney and Grasso and in accordance
with the American Public Health Association methods for testing drinking water and the U. S.
Environmental Protection Agency for testing water using a two-step, delayed incubation,
membrane filtration method. M-Endo Agar LES is used in the standard membrane filtration
technique whereby E. coli bacteria produce a metallic-red colony within 24 hours incubation at
35C. A sterile 3 mM a membrane filter absorbent pad (Whatmann) was placed inside the
cover of a Petri dish and 2.0 mL Lauryl Tryptose Broth added. The wastewater sample (100
mL) was filtered through a sterile membrane (0.2m Whatman nitrocellulose) using a vacuum
manifold, placed onto the pad containing Lauryl Tryptose Broth and incubated at 35C for 2
hours. The membrane was then transferred from the pad to the surface of the M-Endo Agar
LES medium (keeping the side on which the bacteria have been collected facing upward) and
incubated at 35C for 24 hours. As a positive control approx. 100 E. coli were seeded into 100
mL sterile water and as a negative control 100 mL of sterile water was used: both were
treated in the same way as the wastewater sample.
Standard colorimetric methods were used for the testing of total nitrogen, total phosphorus,

96

nitrate, nitrite, ammonia and phosphate (AQUANAL reagents, Sigma) and pH.
Results
The data is summarised in the table below. Values are an average with the standard error
(SE) in parenthesis for three determinations from the three separate site visits. Inflow and
outflow refer to samples taken at the inlet riser and outlet riser, respectively (labelled as A and
B in the diagram above).

Table 2.9.2: Characteristics of inflow and outflow of ADs visited


Stanford Valley

Noordhoek_Thor

Noodhoek_Richmond

Biogas %methane

68 (4)

61 (3)

69 (6)

Pre-and Post-AD:

Inflow

Outflow

Inflow

Outflow

Inflow

Outflow

pH of AD

6.6 (0.4)

6.9 (0.2)

5.8 (0.4)

6.7 (0.6)

6.9 (0.1)

7.1 (0.2)

COD (mg/L)

1834

1252

967

919

35 (9)

311 (89)

123 (21)

38 (18)

27 (11)

Total Nitrogen
Nitrate
Ammonia
Total Phosphorous
Total coliforms per 105 (14)
x105/100 mL

Stanford Valley outflow from algal pond 2x105cfu/ml (3)

97

AD biogas performance survey


A questionnaire directed to the owner and user of the household biogas system was also
used to survey the performance, acceptability and benefits of the AD biogas systems. Below

Stanford Valley
Biogas %methane

Noordhoek_Thor

66 (4)

61 (3)

Noodhoek_Richmond
62 (6)

Pre-and Post-AD: Inflow

Outflow

Inflow

Outflow

Inflow

Outflow

pH of AD

6.6 (0.4)

6.9 (0.2)

5.8 (0.4)

6.7 (0.5)

6.9 (0.1)

7.1 (0.2)

COD (mg/L)

1834 (175) 1012 (122)

3910 (401) 2338 (389)

967 (96)

719 (165)

Total Nitrogen

45 (12)

34 (18)

76 (11)

64 (16)

23 (13)

22 (9)

Ammonia

27 (5)

18 (4)

68 (4)

61 (7)

21 (6)

20 (4)

Total Phosphorous 22 (4)

19 (3)

45 (5)

44 (13)

12 (7)

7 (4)

35 (19)

211 (99)

83 (31)

38 (18)

19 (11)

E.

coli

x10 /100 105 (14)

mL

are the responses from the three household AD units assessed above as well as the recently
installed system in the Town of Stanford (Peter Byrsshe):
__________________________________________________________________________
STANFORD VALLEY FARM AND CONFERENCE CENTRE: Contact: Steve Castle.
email: info@stanfordvalley.co.za
Energy from wastewater questionnaire: Household AD biogas:
The following questionnaire is part of the Energy from Wastewater feasibility study, carried out at
University of Cape Town (funded by the Water Research Council).
OWNER
Why did you decide on a biogas system?
It is with our sustainable energy use goals
Where/How did you hear about system?
Research. AGAMA Energy
How much did the system cost? (ZAR and date)
ZAR 60 250 for the system without the labour costs (supplied by Stanford Valley). Completed
07/2006
What are the loads that you supply to the anaerobic digester (AD)?
Kitchen waste, household black water. Input varies depending on occupancy (20 max).
What are happens to the wastewater outflow?
Polished water currently used to flood irrigate pasture
How much does the biogas system save you on cooking costs? (gas or money saved (ZAR))
Difficult to gauge as REAL volumes produced are unknown.
Any other important benefits you can identify? No
Is the system reliable? Do you have backup fuel?

98

We find the system reliable. We do have backup LPG gas and electricity.
Why dont more people have biogas systems?
Perhaps the upfront costs as well as the unit footprint might be a deterrent.
OPERATOR (user of biogas)
How does the biogas burn compared to ordinary LPG gas?
Better SameX Worse
How does the biogas smell compared to ordinary LPG gas?
Better Same WorseX
How much gas or cooking time do you get per day?
Approximately 6 hours per day.
Any other problems or issues with biogas systems?
Lack of accurate measurement tools.
__________________________________________________________________________
STANFORD Village. Contact: Peter Byrsshe.
email:peter@bysshe.co.za
Energy from wastewater questionnaire: Household AD biogas:
The following questionnaire is part of the Energy from Wastewater feasibility study, carried out at
University of Cape Town (funded by the Water Research Council).
OWNER
Why did you decide on a biogas system?
Recycle water and sort out problem with French drain as well as capture energy
Where/How did you hear about system?
Eric at Stanford Valley farm
How much did the system cost? (ZAR and date)
ZAR 45000 for the system AD, baffled reactor and reed bed. Completed 11/2007
What are the loads that you supply to the anaerobic digester (AD)?
Kitchen waste, household blackwater and greywater. Poo from 2 dogs and kitchen waste from
the restaurant they run in town (20 L tub per day)
What are happens to the wastewater outflow?
Pumped to 13 000 L storage tank for garden irrigation
How much does the biogas system save you on cooking costs? (gas or money saved (ZAR))
All household cooking for a family of 4. Approximately ZAR 150 per month saved
Any other important benefits you can identify?
Water usage reduced.
Reduced

Increased energy efficiency and reduced carbon footprint reduced.

burden on the Municipal WWTP and costs incurred (ZAR100 per month to empty a septic tank
system
or charges of approximately ZAR35 per month if connected to the municipal
sewage system however no
mechanism to recover these costs)
Is the system reliable? Do you have backup fuel?
Reliable and supplies all the cooking needs for the family of 4. Backup LPG gas not required
Why dont more people have biogas systems?

99

The mindset or awareness. This is increasing with the energy crisis. The initial costs mean it is
unaffordable for many.
OPERATOR (user of biogas)
How does the biogas burn compared to ordinary LPG gas?
Better SameX Worse
How does the biogas smell compared to ordinary LPG gas?
Better SameX Worse
How much gas or cooking time do you get per day?
Approximately 3 hours per day.
Any other problems or issues with biogas systems?
Some maintenance required occasional stirring. Need cheaper pre-cast units
__________________________________________________________________________
NOORDHOEK Richmond : Carol Richmond
email:richmoc@lancet.co.za
Energy from wastewater questionnaire: Household AD biogas:
The following questionnaire is part of the Energy from Wastewater feasibility study, carried out at
University of Cape Town (funded by the Water Research Council).
OWNER
Why did you decide on a biogas system?
To decrease dependence on coal-based energy (electricity)
Where/How did you hear about system?
At a sustainability conference in Pretoria 2005
How much did the system cost? (ZAR and date)
ZAR35 000 without labour costs Completed 07/2007
What are the loads that you supply to the anaerobic digester (AD)?
Kitchen waste, household black water and greywater.
wetland (natural swimming pool )

Also plant material from constructed

What are happens to the wastewater outflow?


Goes into the ground. Plan to have a polishing pond and then irrigate with it (garden)
How much does the biogas system save you on cooking costs? (gas or money saved (ZAR))
Saves 2 hrs LPG gas per day.
Any other important benefits you can identify?
Free energy. Save on fertilizer requirements (effluent and sludge), reduces global warming by
reducing methane emissions
Is the system reliable? Do you have backup fuel?
Yes reliable and used for cooking needs. We have backup wood and anthracite burning stove,
solar water heaters and a solar pump , LPG gas and mainline grid electricity.
Why dont more people have biogas systems?
High capital outlay. Perception it is labour intensive, smelly and/or still experimental
OPERATOR (user of biogas)
How does the biogas burn compared to ordinary LPG gas?

100

Better SameX Worse


How does the biogas smell compared to ordinary LPG gas?
Better Same X Worse
How much gas or cooking time do you get per day?
Approximately 2 hours per day.
Any other problems or issues with biogas systems?
Does not work well with Eurogas stoves. Causes thermocouple spark to malfunction? Works
well with simple cast iron stoves (match lit).
__________________________________________________________________________
NOORDHOEK Thor. Contact Anthea Thor.
Energy from wastewater questionnaire: Household AD biogas:
The following questionnaire is part of the Energy from Wastewater feasibility study, carried out at
University of Cape Town (funded by the Water Research Council).
OWNER
Why did you decide on a biogas system?
Energy saving and reducing methane emission
Where/How did you hear about system?
Biophile magazine and AGAMA Energy
How much did the system cost? (ZAR and date)
Unsure. Completed 2007
What are the loads that you supply to the anaerobic digester (AD)?
Household of 1-3 persons. Kitchen waste, household blackwater and greywater and manure from
two horses (approx half a wheelbarrow per day.
What are happens to the wastewater outflow?
Polished water used to irrigate fruit trees. Uses a solar pump
How much does the biogas system save you on cooking costs? (gas or money saved (ZAR))
Approx 9 kg per month (ZAR140).
Any other important benefits you can identify?
Close the loop of water cycle and fertilizer use
Is the system reliable? Do you have backup fuel?
We find the system reliable ad always sufficient
Why dont more people have biogas systems?
Exactly! especially when that from several houses could be combined. SA not resourceful and
uncommon/unheard of technology
OPERATOR (user of biogas)
How does the biogas burn compared to ordinary LPG gas?
Better SameX Worse
How does the biogas smell compared to ordinary LPG gas?
Better SameX Worse
How much gas or cooking time do you get per day?
Sufficient for cooking requirements

101

Any other problems or issues with biogas systems?


Takes about 3 months to establish the biogas production. Needs some maintenance. Only put
biodegradable material in the waste stream and no sand
__________________________________________________________________________

Conclusions

The biogas generated has a methane content of 58-72% and the quantity is sufficient for the
households family cooking requirements.

The AD is capable of reducing the COD considerably (42-68%), but the levels of other nutrients
(nitrogen and phosphorous) are relatively unaffected. Similarly. the levels of coliform bacteria
is reduced a mere 3-5 fold. The values for pH were not always in the range considered
optimal for AD (pH 6.8-7.5). This may be a reflection of both the acidic waster source and the
loads the acid nature of the Stanford (Thor) AD in particular may be a result of the heavy
loading with horse manure.

In general, the quality of the AD effluent exceeds the levels permitted by government
regulations for discharge into environment. This wastewater effluent needs further treatment
systems (reed bed algal pond and/or baffled reactor), and these are in place at the three
locations, but this study only assessed the performance of the AD.

The owners and operators of the AD biogas systems were satisfied with the performance and
generated sufficient biogas for their household cooking needs. Two of the owners wished to
further capture energy and are investigating a generator for electricity production or a biogas
boiler for hot water and under-floor heating.

The following were identified as issues or

obstacles to widespread adoption of household biogas systems:

Initial capital expenditure is considerable. Estimated ZAR10000 to ZAR 100000 for a


6-10 m3 AD, depending on other components of the DEWATS system.

There is a lack of government incentives for either clean renewable energy or


reduction in the loads discharged to the municipal WWTP.

There are some problems with using the biogas in more sophisticated gas stoves.
The systems work well with the rudimentary cast iron gas units with minor
modification to adjust the air flow.

Many of the owners were unsure if they were loading their systems properly and
optimally and had no measure of biogas production. Additionally, there was some
uncertainty if the final wastewater discharge was safe for irrigation crops for human
consumption and met the standards set by the government. Presently, the owners
have chosen to irrigate either pasture for cattle grazing or ornamental plants and fruit
and nut trees.

Lack of awareness of the need for clean and sustainable energy and the view of
waste as a costly burden are obstacles to widespread uptake of AD biogas systems

102

Initial capital expenditure is considerable. Estimated R10 000 to R100 000 for a 6-10
m3 AD, depending on other components of the DEWATS system.

There is a lack of government incentives for either clean renewable energy or


reduction in the loads discharged to the municipal WWTP.

There are some problems with using the biogas in more sophisticated gas stoves.
The systems work well with the rudimentary cast iron gas units with minor
modification to adjust the air flow.

Many of the owners were unsure if they were loading their systems properly and
optimally and had no measure of biogas production. Additionally, there was some
uncertainty if the final wastewater discharge was safe for irrigation crops for human
consumption and met the standards set by the government. Presently, the owners
have chosen to irrigate either pasture for cattle grazing or ornamental plants and fruit
and nut trees.

A lack of awareness of the need for clean and sustainable energy, and the view that
waste is a costly burden are obstacles to the widespread uptake of AD biogas
systems.

103

SECTION 2.10 Energy from Wastewater Industrial Stakeholder Workshop

14 March 2008 9:15-12:30


Studio 2, Dept. of Chemical Engineering, University of Cape Town, Rondebosch
Followed by visit to the SAB Brewery in Newlands

Attendees:
Peter King City of Cape Town
Nokuzola Lujiza City of Cape Town
Anton Laubscher Distell
Kevin Sampson City of Cape Town
Kevin Fawcitt City of Cape Town
William Soekoe City of Cape Town
Greg Austin Agama Energy
Neil Parker Agama Energy
Melumzi Nontanga City of Cape Town
Eleonore Schmollgruber
Harro von Blottnitz UCT
Peter Hoffman SAB
Johan van den Berg CDM Africa
Brett Cohen UCT
Simisha Pather-Elias UCT
William Stafford UCT
Sue Harrison UCT
Stephanie Burton UCT
Rob van Hille UCT

104

1. Energy from wastewater: Project aims, definitions, international practice, technology


(William Stafford, UCT)
William Stafford introduced the aims of the project, the approach and the outcomes of the
project. A brief review of local and international practice was included.
Q&A:

What is the profitability of the biodiesel from algae in terms of carbon

sequestration/emissions and financial payback? There have been studies in Europe which
show a return in capital investment of less than 5 years. There have been studies in other
countries, but each country needs to do its own study due to different costs for building,
electricity, transport, etc. In order to determine the carbon sequestration, a full life cycle
assessment needs to be done, and because this takes such a long time, there havent been
very many studies. Environmental factors such as the amount of sunlight and land area for
ponding are also an important consideration.

What is the efficiency of making biofuels because of the energy intensiveness and

there are many steps to producing the final product? Algal biodiesel can produce secondary
products such as carotene which are profitable. The waste material (cake) can be used as
animal feed, fertiliser, etc. There is great value in the algae cake that remains post processing
(waste) and this can tip the balance of economic feasibility. Peter King then suggested using
the Western Tannery Plant, which also grows Spirulina as a nutritional supplement, as a case
study to determine the economic feasibility and the energy efficiency. Nobody knew if the
plant was still operational since Spirulina that was grown on a tannery waste and used as a
health food supplement resulted in human illness associated with pathogens in the Spirulina
product. Similarly, algae grown on wastewater containing a high heavy metal content is
unsuitable for use as an agricultural fertiliser. Wastewater in Cape Town has very low heavy
metal content, and in that respect is suitable for algal cultivation. Johannesburg has high
levels of heavy metals in their wastewater due to more industry and mining.

2. Energy from wastewater Our Experiences: SAB Newlands Brewery (Peter Hofman,
SAB)
Peter Hofman gave a presentation on their anaerobic digester at the Newlands plant. Due to
new environmental legislation and the rising the costs of discharging the wastewaters to
municipality for treatment SAB decided to treat wastewater on site and plans to harness the
biogas energy.

SAB has made a commitment to reduce its energy consumption by 15% by 2012 and
increase the energy consumption from renewables by 4% in the next few years.

105

70% of SABs breweries have wastewaster treatment plants with AD biogas and 60%

of the biogas is used for heat (steam generation). SAB has calculated an 8%
reduction in energy usage by utilising t AD biogas.
The AD at SAB is of the upflow type (UASB) and has a life span of about 5 years

(time before it needs to be emptied and cleaned of accumulating sludge), and the
residence time is about 1.5 days. The max. COD of the influent is about 8000 mg/L
which is reduced to 350 mg/L at the effluent. The flow rate is 3000 m3/day. Biogas
flow rate 3000-6000 m3/day, about 70% methane and 25% CO2. The biogas energy is
127.4 gJ/day or 5.3 gJ/hr output and can be used for combined heat and power. The
appropriate technology selected for utilising the biogas depends on various factors
which include economic viability, maintenance, quality of gas, noise, safety and
sustainability, supply fluctuations, reticulation, future expansion. These technologies
include:
Hot water generation through water re-circulation
Boiler de-aerating water heater
Steam generation using methane
Absorption chilling for cooling
Waste heat recovery from turbines
Hydrogen fuel cell technology too early to consider
At SAB Newlands, using the biogas for steam generation was most favoured option
due to ease of adaptation, support and maintenance. This system is typically 83%
3
3
efficient and produces 4700 m /day of steam and 27 MJ/m . However, SAB is

currently contractually tied to steam generation by electrical boilers for the next few
years. When this contract terminates they hope to pursue this option. The recovery of
capital expenditure is estimated to be 3-7 years.
Q&A:

Is SAB under pressure to do secondary treatment of its wastewater? No, they


discharge wastewater at 350 mg/L to the municipal WWTP so they do not pay any
excess for treatment charges. However, recently council charges have increased by
125% on a volume basis!

The electricity price has increased 3 fold in the last 10 years. Will SAB revisit the
economic feasibility of alternate energy and AD when the electricity price increases?
Yes, we constantly re-evaluate the feasibility in relation to fuel, electricity and WWTP
tariffs for treatment. However, the main consideration is the availability of appropriate
technology.

106

3. Energy from wastewater Our Experiences: PetroSA (Johan van den Berg, WSP)
Johan van den Berg discussed the financial barriers and implications of CDM at

PetroSAs GTL plant in Mossel Bay.


Discussed the Kyoto Protocol in brief. Developed countries need to decrease carbon

emissions by 5%. Developed countries need to determine the best practices for doing
this and then transfer this information and technology to developing countries.
The CDM and costs for attaining certified emissions reductions (CER) are USD 40 k-

80 k excluding the EIA process. Therefore only large operations are considered viable
(require more than 15 000 tonnes of CO2 equivalent emissions reduction per annum).
There is a growing carbon market or economy value of carbon market estimated at

US $30 billion in 2006

PetroSA has invested USD 4 million, and expects 33 000 CERs per annum

Risks: Too expensive to hedge forex risk on CERs that are euro-denominated and
inflation differentials should see the rate improve in favour of the project
Other barriers include: Time, Licensing and regulations and Commodity price
fluctuations

Q&A:

What is the payback time? Payback on PetroSAs CDM project is estimated to be 5-7
years.

Do the drivers of the CDM projects need to be from first world countries? No, in this
case WSP was a South African company and was bought over by a European
company

Can a company approach WSP? Can a company in a developing country approach a


company in a developed country to finance a project? Yes, it can be initiated from
both sides

4. Energy from wastewater in SA on industrial scale An Estimate of Potential (William


Stafford and Brett Cohen, UCT)
Bret Cohen provided a summary of the energy potentials of wastewaters in South Africa and
this set the scene for the small group discussions which were held.

5. Small Group Discussions


The attendees were broken up into three groups and asked to reflect on the following nine
questions:

107

1. What are the technology barriers to greater uptake of Energy from wastewater in SA?
2. What are the perception barriers to greater uptake of Energy from wastewater in SA?
3. What are the funding/financial barriers to greater uptake of Energy from wastewater in
SA?
4. What do you perceive as the institutional barriers to greater uptake of Energy from
wastewater in SA?
5. What additional benefits do you see in greater uptake?
6. What advantages would there be for Energy from wastewater to be done centrally or at
point of wastewater generation?
7. Should individual industries be driven to greater uptake, or is it just a nice to have and
not core business?
8. Given potentials which were presented here today, relative to national demand, should
Energy from wastewater be a key national focus area for energy recovery, or are there
other places where larger gains can be had?
9. What is required to remove the barriers

The responses to each of these questions are detailed below. One general observation which
was made repeatedly is that the vast majority of the discussions were focussed around AD for
biogas. This in itself demonstrated the need for greater education and the necessity for on the
ground demonstration of some of the other technologies that form the basis of this project.

1.

What are the technology barriers to greater uptake of Energy from wastewater

in SA?

Variability of influent wastewaters many are cyclical

Gas holding requirements

Requirement for integration of several technologies one is seldom sufficient

Need to consider that quality of water after energy-directed treatment may not be as
high as if water quality was the main objective

Need for sulfur removal, and phosphate accumulation

Small companies not having enough resources to pursue Energy from wastewater
need cooperation mechanisms

Lack of skills for design, implementation and operation

108

Technology is too expensive, both in terms of capital outlay and the skills/expertise
required for maintenance. E.g. need to fly in technicians from abroad to service gas
turbines, etc.

Scalability of new technologies (fuel cells, etc.) not proven

Problems of service provision at energy generation level relates to qualified


personal above.

Already problems with maintenance of existing capacity need for public/private


cooperation.

Some of the above issue have resulted in a move away from technologies with the
potential to generate energy (e.g. AD) to those that dont (in reference to sewage
treatment).

Much of the equipment needs to be imported and there are long delays in acquisition.
Maintenance (skills and equipment) is lacking.

Human resource expertise is lacking

Design is not always suited to local context of a developing country like SA

2.

What are the perception barriers to greater uptake of Energy from wastewater in

SA?

No one wants to fund it

Risk perceptions

Perception that its very complex

Problems working in the public sector

Perceived as a wealth generation mechanism rather than a benefit for the community

Whats in it for me attitude

Lack of concern about GCC in SA at present

Distaste for dealing with wastewater esp. sewage

Requirement for proven technology which is successful locally, not just in developed
countries e.g. not enough SA case studies of AD

Newer technology requires significant capital spend

Institutional lethargy

Perception that novel technologies are too expensive (CapEx and OpEx)

Centralized energy is best

109

SA has enough cheap electricity

Government will ensure that there is enough power for all

Cost of WWTP is reflected in rates only

Focus of wastewater treatment is on effluent quality not energy generation.

Despite the fact that high WWTP loads coincide with peak energy demand, there is
little focus on utilising locally generated energy.

3.

What are the funding/financial barriers to greater uptake of Energy from

wastewater in SA?

Wastewater is usually very low on the budget priority list wastewater does not win
votes

Huge capital expenditure required

Govt tenders not awarded for more than 3 years, problem when payback time >3
years

Projects need to show definite financial benefits within a reasonable timescale to


justify funding. Need to payback time to be short

Not always financially dependent in industry

Need for cooperation between companies

Industry needs to be incentivised

It will take pressure of regulation to force industry to comply need regulation

Ratepayers need to see benefit

SA has more pressing problems that need fixing before Energy from wastewater

Municipalities stretched financially to maintain existing infrastructure.

Benefit to city

Corporate farmers important barrier

Push when electricity price increases

Length of EIA fast tracking?

Public private partnership

CDM and carbon credits difficult, implementation model CDM = bad model

Carbon credit market not accessible to smaller scale operations.

The budget for WWTP is often available but there is lack of capacity for
implementation and spending

110

The renewable energy from wastewater treatment competes with cheap coal-derived
power and there is no feed in tariff or rate for peak power

Funding from DME or DSM/Eskom possible but process extremely slow

No feed in tariff for renewable energy

?MIG funding possible

4.

What do you perceive as the institutional barriers to greater uptake of Energy

from wastewater in SA?

Lack of skills / qualified staff

Inefficiency in government departments

Too much red tape and need for streamlining the process of getting permissions

Slow and difficult to obtain power generation license (NER)

Length/cost of EIAs for smaller operations and well-established technologies

Institutional lethargy -> finances

Municipal management finance act problematic (contract between municipalities and


private enterprise must be limited to <3 years or projects must be put out to tender).
Longer contracts subject to public participation. 5-10 years is required to make
Energy from wastewater successful.

Municipalities often need go to open tender on projects so good ideas often do not
progress. Therefore, universities and research institutions are better positioned to
make unsolicited approaches.

No FIT or green energy tariff (3-5 years for WWTP)

DWAF is toothless and does not enforce legislation for discharge (neither does

municipality). Legislation is high and compliance low

5.

What additional benefits do you see in greater uptake?

Increasing awareness of GCC

Clean water as well as energy

Opportunities for re-engineering equipment and processes and development of new


plants

Opportunities for distributed energy generation

111

Demo projects will overcome perception barriers

Finance benefit

CDM credits

Decentralised electricity production

Decentralised energy generation leading to improved grid stability.

Electricity security

Energy security for individual sites.

Sanitary/health issues

Better fertiliser at a local level

Food security at a local level

Financial benefits from proven technologies.

Overcoming public and institutional perception barriers.

Reduction of pressure on primary resources, particularly in remote communities


(wood for cooking)

Enhanced production of organic fertilisers (although given the current outcry relating
to sludge fertilisation in Philadelphia this may not be sustainable)

Avoid wastewater discharge tariffs

Improve quality of environment (rivers and water table) as well as gas emissions
(avoid coal power)

6.

What advantages would there be for Energy from wastewater to be done

centrally or at point of wastewater generation?

Smaller plants are good for cyclical production of wastewater

Industrial park-type organisation would be useful

Problem of inconsistent feed to sewage works if industries do not supply their waste

Potential that wastewater becomes an asset with monetary value rather than a costly
liability

Centralised systems will provide less visible benefits

Localized people get electricity esp. for agriculture/industry

Transportation of load from WWTP to another WWTP to be able to generate enough


electricity > may work for municipality

112

More beneficial to build 1 big unit than many smaller units but many smaller units
will work better for rural areas

Onsite treatment reduces impact on municipal WWTPs.

In some cases economy of scale may necessitate centralised energy generation

De-centralization would reduce pumping/transport costs of disposal and increase


efficiency

possibility of pre-planning and industrial ecology more possible (govt. limited by


legislative hurdles)

Mixed waste streams a problem at municipal WWTP (varying and unknown


characteristics)

On-site industrial treatment means that the govt. legislation is avoided (i.e..
secondary industry possible). Industry is also more nimble and efficient at
implementation for servicing and maintenance.

7. Should individual industries be driven to greater uptake, or is it just a nice to have


and not core business?

Industries need to be incentivised

Freeing up of COD capacity

Visible brownie points

Incentives for secondary treatment

Penalize

Policies: Too long, need to be incentives rather than penalise

Needs to be a carrot and stick approach including both incentives for compliant
industries and significant penalties for non-compliance. New DWAF costing system
for penalties may partially address this

Industry should be driven to greater uptake. Partly through enforcement of legislation


(DWAF) Need to comply with legislation for discharge (Water Act) as well as protect
the environment at large: the polluter pays (NEMA).

The local legislation is inconsistent since each municipality has by-laws for discharge
tariffs some industries will move to an area with more favourable by-laws

113

8. Given the potentials which were presented here today, relative to national demand,
should Energy from wastewater be a key national focus area for energy recovery, or
are there other places where larger gains can be had?

Still need clean water as a priority

Should be major focus but info needs to be disseminated

Will we refer to Energy water instead of wastewater in the future?

If energy is the driver (and not wastewater treatment), the technologies and
processes will be different

Rural areas sludge easy disposal

Need to look at co-benefits

There are easier ways to generate energy (solar, wind)?

The wastewater treatment plant needs to be adapted for energy generation (focus
always on water quality)

Energy from wastewater will not supply all of SA energy needs but could fill the
current missing demand and make waste treatment processes sustainable

9. What is required to remove the barriers?

International investment

Large funding agencies (DST, DME, etc.) need to make seed funding available for
these projects.

Incentives

Right technologies available and awareness of them

New national mindset

Effective legislation

Need for different departments to come together for the common goal is this
feasible?

Demonstration plants are needed

Operation of successful demonstration plants to overcome negative perceptions.

Eliminate replication of research and development efforts

Major enhancement of skills is required at all levels more engineers and qualified
plant operators

114

Vous aimerez peut-être aussi