You are on page 1of 40

HOW ACTIVATED SLUDGE HAS BEEN TRANSFORMED FROM A

WASTE TO A RESOURCE, AND THE IMPLICATIONS OF THIS FOR THE


FUTURE OF THE ACTIVATED SLUDGE PROCESS
Smith, S.R. MCIWEM
Department of Civil and Environmental Engineering, Imperial College London, UK
Email s.r.smith@imperial.ac.uk

Abstract

The activated sludge (AS) process has had a profound impact on improving the
quality of the water environment, but the process has a high energy consumption
associated with the aeration requirement to support suspended microbial growth.
However, the unique physico-chemical and microbiological/ecological properties
of the process and residual sludge offer significant opportunities for resource
recovery that can offset the energy demand, and can also be exploited to produce
valuable industrial chemicals and feedstocks and other value added outputs. The
disintegration of AS by a pre-treatment technique, to increase biogas production
and energy recovery by optimised sludge anaerobic digestion (AD) processes, can
bring the overall wastewater treatment system close to energy self-sufficiency.
Wastewater treatment and nutrient capture in AS has a critical role in conserving
finite P reserves and the embedded energy from industrial N fixation used in fertiliser
manufacture. Expanding this capacity, by maximising: (a) nutrient transfer to the
solid phase, (b) recycling AS biosolids, and (c) the production of concentrated
mineral nutrients, such as struvite and ammonium sulphate, will become increasingly
necessary to create a circular economy for nutrients for sustainable food
production. The resource potential of AS has yet to be fully exploited and this will be
achieved through an integrated approach involving combinations of energy,
nutrient, industrial chemical and possibly other recovery techniques. Diversifying
resource recovery will also increase the security and resilience of sludge
management routes.

Keywords
Activated sludge, Energy, Industrial chemicals, Nutrients, Resource recovery.

Introduction

The activated sludge (AS) process forms an integral part of conventional urban
wastewater treatment systems. These technologies are designed to meet four main
water pollution control objectives relating to:

Health, to minimise the risk of disease transmission by the water-borne route;


Ecology, to minimise the risk to the natural ecological balance of receiving
waters;
Aesthetics, to maintain the value of water for recreational activities;
Economics, to provide an appropriate level of treatment to achieve
environmental protection at reasonable cost.

Within this framework, the systems employed for wastewater treatment have been
demonstrably successful at improving and protecting the water environment and
thereby benefitting human health and well-being. However, the overall cost and
particularly the energy inputs required to meet ever more stringent wastewater
treatment standards are receiving increasing scrutiny (Georges et al., 2009). The
Water Industry consumes up to 3% of the total energy used (Howe, 2009), is the
fourth most energy intensive sector (POST, 2007) and contributes approximately 1%
of national greenhouse gas (GHG) emissions in the UK (Water UK, 2009); specifically,
wastewater treatment contributes almost 60% of overall GHGs emitted by the
industry (Ainger et al., 2009). Indeed, whilst overall UK GHG emissions have declined,
GHG emissions by the Water Industry have increased by up to 30% over the past 10
years (McAdam et al., 2011). To fulfil its renewable energy and climate change
obligations, the UK Government has committed to achieving 15% of UK total energy
from renewable sources by 2020; in playing its part in contributing to these
objectives, the Water Industry has an aspirational target of 20% renewable energy
generation (Howe, 2009).

The AS process continues to make a critical contribution to the observed


improvement in surface water quality primarily through its effectiveness at the main
functions of wastewater treatment, which essentially are to reduce effluent
suspended solids and biological oxygen demand (BOD), and the nitrification of
ammonia-N, the main chemical contaminant in wastewater. There are a wide
range of technical and operational advantages to the process, which have
contributed to its success and extensive deployment as the main biological process
of choice internationally for secondary wastewater treatment. These benefits
include:

Unique physico-chemical properties of the biological floc facilitating the


rapid and effective sorption of metal ions and particulate and colloidal
material;
High suspended solids removal (>97%);
Dynamic, mixed microbial community;
Relatively robust and flexible microbial ecology that can be adapted for
biological nutrient removal (N and P);
Aerobic biological activity that is generally more effective at degrading
anthropogenic organic compounds compared to anaerobic metabolism;
Cost effective;
Small footprint (10 - 15%) compared to alternative filter-based systems;
Relatively simple engineering and maintenance.

Whilst fit-for-purpose in terms of achieving water quality objectives, the AS process


has a major drawback due to the amount of energy consumed for aeration, which
represents 60% of the total energy requirement of conventional municipal
wastewater treatment (Shi, 2011). Energy efficiency of the process can be increased
by up to 20% by improving aeration systems and process control (Shi, 2011).
However, to maximise overall efficiency, and to potentially convert wastewater
treatment into a neutral or net energy producing system, alternatives to AS have
been proposed to minimise or avoid aerobic treatment and the energy demands of
aeration altogether. Suggested and potential strategies involve, for example,
anaerobic treatment of settled sewage after conventional primary sedimentation or
following novel up-concentration processes, utilising membrane filtration
technologies for effluent treatment and stripping dissolved methane (Verstraete et
al., 2009; McAdam et al., 2011; McCarty et al., 2011; Tauseef et al., 2013).
Alternative wastewater treatments based on physico-chemical methods are also
possible, but are still at an early stage of development (Rulkens, 2004).

Installing anaerobic secondary treatment appears a practical possibility for the


future, although McCarty et al. (2011) suggest that retrofitting anaerobic
wastewater treatment to existing conventional aerobic plants may be a costly
option. Therefore, given the uncertainties and lack of experience in the operational
performance and efficiency of the alternatives, and current investment in existing
wastewater treatment infrastructure, the replacement of AS assets is an unlikely
proposition. Therefore, the main strategy to offset the resource consumption by the
AS process is to maximise recovery of energy and other marketable resources from
the sludge. This may entail a strategy to increase overall AS production by, for
example, minimising solids retention times in aeration basins; an additional benefit is
that this approach also reduces process oxygenation and energy requirements
(McCarty et al., 2011).

The earliest reviews of novel approaches to resource recovery from sewage sludge
were published in the late 1980s (Frost and Campbell, 1986; Webber et al., 1986) and
by Webber (1991). Since then, a proliferation of research and published material in
the area has occurred highlighting the importance of sludge in general, and AS in
particular, as a valuable resource. Indeed, the evidence demonstrates that AS is
currently under-utilised as a resource. Recent reviews by Tyagi and Lo (2013) and Mo
and Zhang (2013) provide particularly useful summaries of some of the recent
developments.

This paper provides an overview of current sludge management and recycling


practices internationally, and the options and opportunities for resource recovery
from AS. The paper highlights the recovery of energy and nutrients, as well as the
potential of extracting valuable industrial chemicals, from AS. The importance of a
greater integration of recovery options is emphasised to maximise: (a) the resource
efficiency potential of AS, and (b) the resilience and flexibility of sludge
management routes. The implications of increasing resource efficiency for the future
of the AS process are also considered.

Sewage Sludge Production and Management

Sewage sludge is an inevitable and necessary output from conventional


wastewater treatment that arises due to the concentration and removal of pollutant
materials from the wastewater by physico-chemical and biological water
purification processes. Therefore, a direct consequence of implementing more
stringent standards of wastewater treatment is that sludge production is inevitably
raised. For example, total sludge production has increased by approximately 50% in
European 15 countries from 1984 and the current total production is approximately
10 M t/year DS (Figure 1). Other regions have seen an even more dramatic rise, for
example, from a position of almost no sludge production in 1980, wastewater
treatment plant (WWTP) in China are expected to generate approximately 10 M t
DS of sludge/year by 2015 (Figure 2). For comparison, national sludge surveys in the
US (for 2004) and Australia (for 2013) indicate production in these countries is
equivalent to 6.5 M t/year DS (7.2 M US ton) and 0.33 M t/year DS, respectively
(NEBRA, 2007; ANZBP, 2013). Current UK output is equivalent to 1.4 M t/year DS,
decreasing from a maximum production of 1.83 Mt DS in 2007 (Eurostat, 2013).

The treatment and management of sewage sludge presents a range of technical


and perception challenges, and the costs associated with this represent more than
50% of total wastewater treatment costs (Kroiss, 2004). Therefore, sludge production
is often considered as a disadvantage and should be minimised as far as possible
(degaard, 2004). However, sludge produced from wastewater treatment presents
a range of resource recovery opportunities, some of which are well established,
such as the reuse of sludge biosolids as fertiliser products in agriculture, whereas
others have yet to be fully maximised, for example the generation of biogas and
renewable energy from sludge anaerobic digestion (AD). There are also other
resource recovery options which have yet to be realised and, taken together, all of
these can offset the resource demands of the AS process.

Recycling sludge by land application (for agricultural use or other purposes, such as
restoration), or disposal of sludge in landfill are the two principal, final destination
routes for sewage sludge management in Europe (Figure 3). Separate statistics are
collected on the incineration of sludge, but this is not a disposal route per se;
incineration is effectively a sludge treatment process that achieves the maximum
solids reduction by thermal conversion and the residual ash is disposed of in landfill.
Energy production from sludge incineration processes is consumed for water
evaporation and to meet the parasitic load of the process, therefore, incineration
does not usually contribute to improved energy management (Mininni et al., 1997;
Thierbach and Hanssen, 2002; Kroiss, 2004). In general, the disposal of whole sludge
in landfill is limited and is regarded as unsustainable in the long-term and inconsistent
with EU landfill policy to reduce disposal of biodegradable municipal waste by this
route. Nevertheless, it represents a significant disposal route in some cases. There is
also uncertainty about the final destinations for sludge in some European countries
as suggested by the large others category in some reports (Figure 3). The
assessment of sludge management routes demonstrates that the available options
are limited and potentially vulnerable to future barriers or restrictions. Therefore,
diversification would increase the resilience of sludge management and this can be
achieved by also maximising the resource recovery opportunities from sludge.
12000
Sludge production t/year DS
10000

8000
x1000

6000

4000

2000

0
1980 1985 1990 1995 2000 2005 2010 2015
Year
All EU EU15

Figure 1: Sewage sludge production in the European Union (post 2002 data
extracted and adapted from Eurostat, 2013)

Figure 2: Numbers of wastewater treatment plant and sludge production in China,


1980-2015 (GWI, 2012)
Figure 3: Sewage sludge management from urban wastewater treatment in
European countries, 2009 (% of total mass) (Eurostat, 2013)

Activated Sludge: Properties and Resources

Under typical conditions, AS, which is the biological sludge generated from the
metabolic transformation of biodegradable substrates in wastewater, comprises 50%
of overall sludge production at a WWTP. Activated sludge contains a range of
valuable resources, including a high total organic carbon content, which is a
potential energy source or can be used as a bulky soil improver; the major plant
nutrients: N, P and S; cellulosic fibres; polyhydroxyalkanoate (PHA) polymers, which
are fully biodegradable, renewable polyesters (poly-3-hydroxybutyrate (PHB) is the
most common type); as well as hydrolase enzymes that are produced by
microorganisms to hydrolyse and metabolise the high lipid and protein content in
wastewater. A summary of the chemical composition of AS is presented in (Table 1).

Sewage sludge also contains inorganic and organic contaminants and the source
control of these from industrial and domestic sources is critical to sludge quality for
the major current recycling routes; this is particularly the case for agriculture
application. However, the concentrations of potentially toxic elements (PTEs) and of
a majority of persistent organic pollutants (POPs) and other organic contaminants
(OCs) have decreased with time due to effective source control legislation and
action by the Water Industry (ICON 2001; Smith, 2009). In a time trend analysis,
Olofsson et al. (2012) observed statistically significant changes in the concentrations
of 18 compounds out of 77 examined, and 75% of these were declining. Against this
background, and the implementation of microbiological controls on sludge (ADAS,
2001), comprehensive and quantitative risk assessments of contaminants and
infectious microbes (Gale, 2005; Schowanek et al., 2007; Eriksen et al., 2009),
demonstrate the risks associated with land application of treated sludge are minimal
and provide the technical case and justification supporting the reuse of sludge in
agriculture and for other soil improvement purposes. In future, however, it may be
possible to recover refined nutrient products for agricultural use also with reduced
contaminant concentrations, and at the same time generate renewable energy
and other valuable industrial products by applying a series of recovery stages to the
sludge generated by wastewater treatment.

Table 1: Ranges of selected chemical properties of activated sludge compared to


primary sludge (values in brackets are typical values) (Collinge and Bruce 1981;
CIWEM, 1995a,b; Satoh et al., 1999; EC, 2001; Tchobanoglous et al., 2003; Chua et al.,
2003; Chen et al., 2007; Dewil et al., 2008)

Property Untreated Primary Untreated Activated

Dry solids (%) 3.0-7.0 (6.0) 1.0-3.0 (1.5)


Volatile solids (% DS) 60-80 (70) 60-90 (80)
Grease and fat (ether extractable) 7.0-35 (20) 5.0-12
(% DS)
Protein (% DS) 20-30 (25) 30-60
Total N (% DS) 1.5-4.0 (2.5) 2.4-7.6 (5.0)
Total P (% DS) 0.4-1.2 (0.7) 1.2->5.0 (2.2)
Total K (% DS) <0.0-0.8(0.3) 0.4-0.6
Total S (% DS) 1.5 0.5-1.0
Fibre (% DS) (19) 0.4-10
Sugar (% DS) (0.7) 11
Starch (% DS) (1.5) No data
pH 5.0-8.0 (6.0) 6.5-8.0
Organic acids (mg L-1 HAc) 200-2000 (500) 1100-1700
Energy content (MJ kg-1 DS) 23-29 (25) 19-23
Polyhydroxyalkanoates (PHAs) (% DS) No data 20-60

DS, dry solids; HAc, acetic acid

International Research on Resource Recovery from Activated


Sludge

The Web-of-Science database holds 60,700 records relating to activated sludge for
the period 1950-2014. The level of research activity focussed on resource recovery
from AS has increased markedly in the past 10 years (Figure 4). This remarkable trend
underlines the significance and importance of AS as a resource and the innovative
ways that the process and material can be used, adapted and applied to recover
valuable products and components and to perform critical, environmentally
beneficial functions. Internationally, the main research output is originating from
China (27%) and USA (14%). The UK Water Industry may not be taking full advantage
of the potential benefits that these opportunities offer reflected in the relatively small
volume of UK research output in this area.
(a) Number of publications by year

(b) Number of publications by country

Figure 4: Numbers of papers published in the international literature on resource


aspects of activated sludge: (a) by year between 1990-2013 and (b) for selected
countries for 1950-2014 (results of Web-of-Science search: Activated AND Sludge
AND Resource)

Appraisal and Status of Resource Recovery Options

Energy

Energy efficiency of wastewater treatment

Shi (2011) presents a comprehensive assessment of the mass flow and energy
efficiency of municipal WWTP, considering the various options for increasing the
overall energy efficiency of the process, and particularly to offset the aeration
demand of AS. For example, the application of conventional mesophilic AD
combined heat and power (CHP) systems (25 - 30% electricity conversion efficiency)
can increase energy efficiency up to 30%. Energy efficiency can be raised further by
up to 50% by a combination of the following options: (a) improving primary
sedimentation, (b) sludge pre-treatment, or (c) thermophilic digestion. Increasing
the efficiency of primary treatment leads to greater primary solids transfer to, and
biogas production from, conventional AD systems, and also reduces the aeration
demand and biomass production in secondary treatment. However, applying
sludge pre-treatments and improving AD performance to increase VS destruction to
60% in combination with other measures can raise the energy efficiency to 80%.
Research and operational experience demonstrate that, with the integrated
application of appropriate technologies and processes, energy efficiencies
equivalent to 80 - 100% are possible and that a positive energy balance is feasible in
practice.

For recent reviews concerning the recovery of energy from sewage sludge and a
comparison of options, see Rulkens and Bien (2004) and Rulkens (2008).

Anaerobic digestion

Approximately 70% of sewage sludge is treated by AD in the UK (EA, 2014) and in the
US about 50% of sludge production receives AD treatment (NEBRA, 2007). The
scientific literature on the potential contribution of AD to global bioenergy
production and on the application of AD for waste activated sludge (WAS)
treatment specifically was reviewed recently by Appels et al. (2008, 2011).
Producing energy from organic waste materials is regarded as one of the major
future sources of renewable energy and AD processes in particular are expected to
play a critical role in the renewable energy market.

Anaerobic digestion processes involve a series of microbial transformations of


complex organic substrates by hydrolysis, acidogenic, acetogenic and
methanogenic reactions. The basic outputs are a stabilised, solid residual material,
suitable for beneficial reuse applications on land, and biogas, which contains CO2
and CH4 in the approximate volumes: 25 - 30% and 60 - 70%, respectively, and other
trace gases, and is a valuable renewable energy source. Optimum sewage sludge
digestion coupled with energy recovery, usually by CHP, is critical to the overall
energy efficiency of wastewater treatment (Shi, 2011), which can approach self-
sufficiency including the energy demand for secondary biological treatment
(Jenicek et al., 2013).

Sewage sludge is one of the most effective substrates for AD and CH4 production
(Figure 5). Conventional mesophilic AD typically achieves ~40% VS destruction and
gas production is equivalent to 0.8-1.1 m3/kg VS destroyed (Tchobanoglous et al.,
2003). However, until recently, little attention was paid to AD process operation to
optimise VS destruction and biogas production, but this is now a priority area for
research and operational action to manage energy consumption and increase
renewable energy generation by the Water Industry. For example, recent studies
have examined the effect of digestion temperature and solids loading rate on VS
destruction and gas production rates by conventional mesophilic AD (Winter et al.,
2012, 2013).
Figure 5: Overview of the average methane yields obtained through anaerobic
digestion of different waste types (Appels et al., 2011)

Activated sludge is less susceptible to AD processes compared to primary sludge;


the VS destruction of WAS is generally considered to be approximately 25% smaller
compared to primary sludge (Shi, 2011). Bolzonella et al. (2005) reported that the
average VS destruction measured for mesophilic AD of WAS at four large Italian
WWTP was 18% (range: 13 27%) and that the specific gas production rate
decreased by more than 60% with increasing solids retention from 8 to 35 days
during secondary wastewater treatment. The relatively poorer digestibility of AS is
explained because microbial cells form a major fraction of the organic matter in
secondary sludge and the structural characteristics of cell walls provide protection
against biodegradation processes. Consequently, the cellular biomass is relatively
resistant to the first hydrolysis stage of the biochemical degradation process and this
therefore represents the main rate-limiting step to AD of AS (Appels et al., 2008).

However, the relatively slow anaerobic digestibility of WAS can be overcome by by-
passing the hydrolysis stage by applying a pre-treatment technique to disrupt the
biomass causing cell lysis and the release of readily degradable organic substrates
for the acidogenic population. Carrre et al. (2010) have reviewed the available
pre-treatment technologies to improve sludge anaerobic degradability. These
include: biological (eg two-stage thermophilic-mesophilic AD is the most common
type), thermal hydrolysis, mechanical (eg ultrasonic, lysis-centrifuge or liquid shear)
and chemical treatments (eg oxidation with ozone, or alkali treatment). Biological,
thermal hydrolysis and mechanical methods are the most widely applied
techniques.
Thermal hydrolysis of sludge is becoming well established for sludge AD pre-
treatment. For example, there are 23 Cambi thermal hydrolysis reference
installations at WWTP internationally and it is particularly pertinent to mention this in
the context of resource recovery from AS because the largest plant is soon to be
commissioned at the Davyhulme Sewage Treatment Works in Manchester
(http://www.cambi.no/wip4/plant.epl?cat=10643&id=463527) where the AS process
was originally developed. The main advantages of the process include: (a) an
increase in biogas production and VS destruction of approximately 60%, (b)
increase in dry solids to 30% after dewatering, and (c) increased solids loading to 5
6 kg VS/m3/day (Carrre et al., 2010). Although thermal hydrolysis increases overall
AD performance and biogas yield, the thermal energy consumption is significant,
therefore, the overall energy balance is neutral. Thermal hydrolysis is also more
capital intensive compared to mechanical options.

Mechanical treatments have smaller capital costs, compared to thermal hydrolysis,


and lead to moderate improvements in biogas production (eg 15-30% for lysis-
centrifuge (Dohanyos et al., 1997) and 10-45% for ultrasonic treatment (Tiehm et al.,
2001) and up to 60% for liquid shear processes). In general, however, the electrical
consumption is approximately 0.3 kWh/kg VS and increase in biogas yield is
equivalent to 0.5 kWh/kg VS. Therefore, assuming an electrical conversion efficiency
of 30%, the energy balance is generally negative (Carrre et al., 2010). Nevertheless,
Jenicek et al. (2013) considered that upgrading thickening centrifuges to lysate
thickening centrifuges for effective WAS disintegration to increase biogas
production was effective as part of a series of measures to improve the energy
efficiency of WWTP. Recent operational experience with ultrasonic treatment for cell
lysis of AS at the Melton WWTP operated by Western Water, Melbourne, Australia
achieved a modest increase in biogas production of 10%, however, it was
considered successful also because of improvements in digestion process stability
and dewaterability of digested sludge (pers. com. Rod Curtis, Team Leader Melton
Recycled Water Plant).

Other biofuels

Hydrogen production from sewage sludge appears to have significant potential as


an alternative renewable energy source compared to current approaches to
sludge AD treatment and associated CH4 production (Tyagi and Lo, 2013). The
reasons favouring hydrogen include no GHG contribution from combustion and the
energy value is much larger (121 kJ/g net calorific value (NCV)) than hydrocarbon
fuels (CH4 has a NCV equivalent to 50 kJ/g). Producing hydrogen by biological
fermentation may be of particular interest for sludge treatment because it is similar
to the familiar approach and technology involved in sludge AD (Mudhoo et al.,
2011) and it is less energy intensive and complex compared to thermochemical
methods of hydrogen production by pyrolysis or gasification (Tyagi and Lo, 2013).
Microbial fermentation reactions are less sensitive to process conditions compared
to methanogenesis, and fermentation-based processes potentially operate with
shorter retention times, compared to standard AD (Cheong et al., 2007; Liu et al.,
2008). As is the case with sludge AD, however, the hydrolysis stage of anaerobic
fermentation is rate limiting and it is necessary to pretreat WAS to lyse the cellular
material to maximise the hydrogen yield (Ting et al., 2004; Wang et al., 2010). The
reported maximum hydrogen yield from sludge fermentation is 39.2 mL H2/g
chemical oxygen demand (COD) removed (Mudhoo et al., 2011). Therefore, the
energy value from hydrogen production by anaerobic fermentation is considerably
smaller and less than 10% of that possible by AD, which has a typical methane
output of approximately 300 ml/g COD destroyed (Astals et al., 2013). Biological
hydrogen production from WAS may have considerable future potential; however,
further optimisation and development of anaerobic fermentation is necessary to
improve the process efficiency for operational application (Mudhoo et al., 2011).

Other strategies to recover liquid or other gaseous fuel resources from sludge
include production of bio-oil or syngas by pyrolysis or gasification processes (Manara
and Zabaniotou, 2012), or biodiesel (Siddiquee and Rohani, 2011). Energy can be
recovered from sludge by advanced oxidation processes, for example, supercritical
wet oxidation (>375 oC and 22.1 MPa) (Rulkens, 2008). The direct production of
electricity from sewage sludge and specifically from WAS using microbial fuel cells is
also technically feasible (Dentel et al., 2004; Linji et al., 2013). However, based on the
current evidence and status, these techniques struggle to compete with well
established AD technologies for the treatment and recovery of energy and other
resources (eg nutrients) from WAS in terms of operational experience, practicability,
cost and energy efficiency (Rulkens, 2008; Tyagi and Lo, 2013).

Nutrients

The fertiliser industry consumes approximately 1.2% of global energy and is


responsible for a similar proportion of global GHG emissions; >90% of the energy is
used in ammonia (NH3) production (Swaminathan and Sukalac, 2004 cited by IPCC,
2007). The production of P fertilisers consumes less energy, but, in contrast to mineral
N which is produced by a chemical manufacturing process by fixing atmospheric N,
P minerals are extracted primarily by large scale surface mining, which potentially
causes local environmental damage (UNEP/IFIA, 2001) and geological reserves of
this important element are finite; existing rock phosphate reserves could be
depleted within 50 - 100 years (Cordell et al., 2009). An estimated 80% of global
phosphate production is used in the manufacture of agricultural fertilisers (Evans,
2014). Peak P production is anticipated within the next 20 - 25 years (Figure 6) and is
therefore predicted to have a significant impact on global agricultural production
and economies (Cordell et al., 2009). Managing the future P supply situation will not
be a trivial issue. However, a more optimistic view is that, with improved recycling
and efficiency of P use, a sustainable plateau of phosphate production (Figure 6)
from geological and recovered sources may be possible and sustainable (Dolan,
2013). Either way, maximising P recovery is a critical issue requiring a central policy
direction and an action plan to conserve P resources as far as practicable to ensure
long-term food security. This is a priority particularly for European countries which are
almost entirely dependent on P imports. A critical effect of the energy consumption
or resource availability for N and P is that fertiliser costs are expected to continue to
rise, and this will be one of the main factors contributing to increasing food prices in
future. The current price of N and P fertilisers is approximately 800/t of N and 1400/t
of P (as ammonium nitrate at 34.5% N and triple superphosphate at 20% P) (DairyCo,
2014).
Figure 6: Historical and potential future trends in global phosphate production
(Dolan, 2013)

Approximately 14,000 MLD of domestic plus 12,000 MLD of industrial wastewater,


respectively, are discharged to the wastewater collection system in the UK (ETRAC,
1998). This volume contains a total mass of N equivalent to approximately 180,000 t
N/y (Lillywhite and Rahn, 2005) and corresponds to a total N content in wastewater
of approximately 20 mg/L (Tchobanoglous et al., 2003). Data for the total mass of P
in sewage could not be found, however, assuming a typical total P content in
wastewater of 12 mg/L (Morse et al., 1993; Tchobanoglous et al., 2003) the total
mass of P is equivalent to approximately 114,000 t P/y. For comparison, this quantity
of N represents 18% of the total amount of N applied in fertilisers to agricultural land
in the UK in 2011/12 (1.0 million t of N was applied to farmland in fertiliser in the UK in
2011/12; AIC, 2013). However, the quantity of P in wastewater exceeds the amount
of P applied to farmland in fertilisers by approximately 40% (82,000 t of P was applied
to farmland in mineral fertilisers in the UK in 2011/12; AIC, 2013).

Nitrogen and P mass balances show that conventional wastewater treatment


systems are extremely wasteful of these valuable resources and of N in particular.
Only 12% of the N entering the system is transferred to the sludge where it may be
recycled for crop fertilisation, representing 3% of the total regional mass flow of N
(Kroiss, 2004); 40% is discharged in the effluent and 48% is lost through atmospheric
emissions (Figure 7). Phosphorus recovery is better and >40% of the P in wastewater
may be captured in the sludge with the remainder discharged in treated effluent,
with AS contributing >60% of the P supplied to the sludge stabilisation process (Figure
7). In conventional wastewater treatment, sewage sludge can therefore account
for 8% of the regional flow of P (Kroiss, 2004). Nutrients in treated wastewater
represent a valuable resource in areas where reuse is practiced for agricultural
irrigation. However, discharges in effluents impact the quality of surface water
systems and nutrient removal at WWTP is increasingly necessary. Biological and
chemical P removal techniques increase the capture and transfer of P to sewage
sludge and can raise the efficiency of P recovery by wastewater treatment by >95%
(Horan, 1990). Modifying the AS process for enhanced biological P removal (EBPR),
in particular, provides significant opportunities to increase resource recovery from AS
by downstream processes to maximise P retention. If P-removal efficiency reaches
approximately 90%, sewage sludge can contribute up to 15% of the total regional
flow of P (Kroiss, 2004). However, the transfer of N to the sludge is largely controlled
by the biological requirements of microbial cells, which have a typical N:C ratio
equivalent to 0.23 and total N content equivalent to 12.4% (dry weight)
(Tchobanoglous et al., 2003). Approximately 50% of the total N supplied to sludge
stabilisation processes is therefore contained in WAS; see Figure 7). Thus, the
recovery of N in the solid phase is mainly controlled by the amount of sludge that is
generated and this also depends on the type and operation of the nutrient removal
process. Nitrogen is removed from wastewater by microbiologically converting
soluble N, which is lost from the system to the atmosphere in the gaseous phase.

Given the resource implications associated with the industrial production and
manufacture of conventional fertilisers, rather than using more energy to simply
remove and discard the nutrients contained in wastewater, a more sustainable
strategy would be to focus on the recovery and recycling of these valuable nutrients
as agricultural fertilisers. This approach would conserve both the embedded energy
within the nutrient life-cycle overall and geogenic resources, by assisting the closure
of the nutrient loop (Green Alliance, 2007; McCarty et al., 2011). In future, therefore,
a shift away from current N elimination strategies in wastewater treatment (eg Ahn,
2006; Paredes et al., 2007) may be anticipated with greater emphasis being placed
on the recovery of N as a resource. There are also significant financial advantages
to be gained from this strategy considering the high and increasing cost of fertiliser
materials. Indeed, the approximate value per annum of UK sludge in terms of its total
N and P contents is equivalent to 110 million (assuming UK sludge production is 1.4
Mt/year DS and contains 4.4% and 3.1% of total N and P, respectively (Defra, 2010)).

Nutrient recovery processes from sludge that employ thermal treatments and acidic
or alkaline conditions and/or precipitation reactions (for example: degaard et al.,
2002; Rulkens, 2004; Dichtl et al., 2007), may require high energy and/or chemical
inputs and are not currently considered as cost-effective options (Verstraete et al.,
2009). Nevertheless, further research in this area is necessary and should also focus
on increasing the efficiency of solid-phase N recovery by wastewater treatment.

Fertiliser value

Field and laboratory investigations have comprehensively quantified the fertiliser


value of the different types of sewage sludge produced by the UK Water Industry
(Morris et al., 2003). The maximum availability of N contained in anaerobically
digested, thermally dried or alkaline stabilised products is typically equivalent to 30
to 35% of total N. It is usual practice to combine the primary and WAS sludge
streams together prior to the sludge stabilisation process, so that AS contributes to
the overall N availability of the final treated product. However, direct measurements
of the specific N fertiliser value of AS have also been performed. In field
investigations (Figure 8 and 9) the reported N equivalency of AS was 30 45%
(ORiordan et al., 1987a; Smith and Hadley, 1988), and in closed laboratory
incubation and controlled environment conditions, recoveries of mineral N released
from AS up to 60% of total N have been measured (Smith and Hadley, 1989, 1990).

(a) Nitrogen

(b) Phosphorus

Figure 7: (a) Nitrogen and (b) phosphorus mass balances of conventional


wastewater treatment and sludge anaerobic digestion plant (Shi, 2011)
Figure 8: Systematically arranged field fertilser experiment showing the yield
response of summer cabbage to thermally dried activated sludge applied in
factorially combined base and top dressings (Base dressing rates applied right to
left: 0, 125, 250, 375, 500 kg N/ha; top dressing rates applied front to back: 0, 50, 100,
150, 200 kg N/ha)

(a) Ammonium nitrate (b) Activated sludge


Yield (g/plant Fresh weight)

Top dressing Base dressing Top dressing Base dressing


(kg N/ha) (kg N/ha) (kg N/ha) (kg N/ha)

Figure 9: Yield response of summer cabbage (g/plant, fresh weight) to (a)


ammonium nitrate and (b) thermally dried activated sludge applied in factorially
combined base and top dressings (Smith and Hadley, 1988)

Treated sewage sludge typically has a phosphate fertiliser value of 50% (Defra,
2010). However, modifying the AS process for EBPR increases the P content and
raises the availability of P in the residual mixed treated sludge to a level that is at
least as effective as triple superphosphate fertiliser (Smith et al., 2003; Oladeji et al.,
2008; Miller and OConnor, 2009). Indeed, field trials specifically investigating the P
fertiliser value of AS show that the sludge from conventional biological aerobic
wastewater treatment has a large available P content equivalent to 80% relative to
superphosphate fertiliser (ORiordan et al., 1987b).

Nutrient recovery

In conventional wastewater treatment or plant operating EBPR producing AS with


typical P contents in the range 2.5 to 3.5%, 20 to 30% of the P incoming to sludge AD
processes is returned in the dewatering liquor (Figure 7; also see Jardin and Ppel,
1994). However, EBPR can lead to very large total P contents in AS, equivalent to 6
7%. Under these circumstances, the P feedback can increase to 40%, due to the
solubilisation of a fraction of the stored polyphosphate released under anaerobic
conditions, increasing the P load on the wastewater treatment process leading to
operational problems due to struvite (MgNH4PO4.6H2O) precipitation and scaling
(Jardin and Ppel, 1994). However, effective struvite crystallisation techniques have
been developed and are commercially available (Evans, 2014). These may recover
>70% of the P in AD dewatering liquors and potential recoveries of >90% are possible
(Le Corre et al., 2009; Liu et al., 2012). Estimates suggest (Gaterell et al., 2000) that
29,000 t/year of P could be potentially recovered through controlled struvite
precipitation at WWTP in the UK. Across Europe as a whole, the potential recovery
could be 134,000 t P/year. A further advantage is that ammonium-N is also
simultaneously recovered with P and efficiencies of 70%, and potentially up to 98%,
are reported (Le Corre et al., 2009). This has the potential to increase the capture of
N by up to 20% of the overall mass flow of N entering WWTP (Figure 7). Struvite
contains 12.6% P, 5.7% N and 9.9% Mg (Liu et al., 2012) and, although field trials have
yet to be performed with the material, plant growth studies show that struvite is a
highly effective N-P-Mg fertiliser. Chemical analysis results also indicate the
concentrations of heavy metals are small and significantly below the quality limits
applied to conventional mineral fertilisers (Liu et al., 2012). Given these benefits and
the resource potential of AS there is a strong case to assess the wider application of
struvite recovery to wastewater and sludge treatment in the Water Industry.

Other techniques to recover nutrients from AS by down-stream processing include:


(a) P recovery from the ash produced by incinerating sludge (Petzet and Cornel,
2011; Donatello and Cheeseman, 2013), (b) thermal conditioning WAS and P
precipitation (Kurado et al., 2002; Takiguchi et al., 2004), and (c) integrating
sidestream gas stripping or dewatering liquor treatment with sludge AD processes for
N capture as ammonium sulphate (Dichtl et al., 2007; Shi, 2011; Walker et al., 2011;
Yenign and Demirel, 2013; Serna-Maza et al., 2014).

In principle, sewage sludge ash (SSA) could be applied directly to land as a fertiliser
if it meets the contaminant limits for agricultural fertilisers. For instance, Petzet and
Cornel (2011) suggest that 30% of SSA in Germany could be potentially used directly
for fertiliser manufacture without further processing to reduce contaminant
concentrations. Acid extraction of SSA is generally considered as the most practical
option for P recovery from SSA (for example, see Hong et al., 2005); this approach
allows complete dissolution of P and heavy metal impurities are removed by
precipitation by increasing the pH value, although other more advanced
technologies including: nanofiltration, liquid-liquid extraction, or ion exchange have
also been examined. However, these procedures are either at the developmental
stage or the energy and chemical requirements have prevented their application at
full-scale. Other alternative technologies have been developed for thermal
recovery of P from SSA, but require high temperature treatments in the range 1000
2000oC and therefore also have a significant energy requirement. Consequently,
there are currently few examples where P recovery from SSA is practiced at full-
scale (Stark, 2004; Ptezet and Cornell, 2011). Therefore, until the recovery of P from
SSA becomes economically feasible, the disposal of SSA in mono-landfill is proposed
as a intermediate measure to conserve this important P resource and there are
plans to implement this in Germany (Kroiss, 2004; Petzet and Cornel, 2011).

Kuroda et al. (2002) and Takiguchi et al. (2004) describe a simple method for P
recovery from WAS produced from EBPR processes by heating the sludge to 70oC for
approximately 1 hour and P precipitation with CaCl2, without pH adjustment. The
technique recovered approximately 75% of the total P in the sludge as a high P
content (16%) calcium phosphate mineral. Further advantages included reduced P
release and increased digestion efficiency and methane production during sludge
AD. The treatment conditions are similar to those operating in existing sludge pre-
pasteurisation systems, therefore, there could be potential advantages gained from
integrating a P recovery stage into thermal disinfection processes. However, this
may be less attractive due to the energy input required to heat unthickened liquid
AS compared to struvite crystallisation post AD treatment, which also improves the
overall process N balance.

Anaerobic digestion processes are sensitive to ammonia toxicity and the upper
critical concentration reported for sludge AD at which inhibitory effects may be
detected is approximately 3000 mg/l of total ammonia N (TAN) and 400 mg/l as free
ammonia N (FAN) (Yenign and Demirel, 2013), although, process inhibition has
been observed at TAN concentrations as low as 760 mg/l of TAN (Walker et al.,
2011). Ammonia toxicity is a significant problem for AD of source separated food
waste due to the high N content and biodegradability of the waste (Walker et al.,
2011). However, the expanding application of high intensity pre-treatments, eg
thermal hydrolysis, for biomass disruption to primarily increase digestibility and biogas
yield from AD of WAS could also place sludge AD at potential risk from ammonia
toxicity (Carrre et al., 2010). If the codigestion of sludge with foodwaste were
realised in the future, this would also increase ammonia concentrations in sludge
digesters. Recovery processes require high concentrations of ammonia to be
effective (up to several thousand mg/l) (Shi, 2011) and are therefore suitable for the
treatment of digestates or concentrated dewatering liquors. Ammonia stripping for
N recovery is already employed at full-scale to sludge dewatering liquors following
AD (Dichtl et al., 2007); steam stripping, distillation and reversible chemosorption
have also been proposed, although these require high energy comsumption (Shi,
2011). Removal by sidestream treatment using a biogas stripping technique is also
suggested as a feasible method to control ammonia concentrations in food waste
AD processes (Walker et al., 2011; Serna-Maza et al., 2014) and may also be
potentially applicable to sludge AD for ammonia control and N recovery.
Biofertilisers

Amino acid chelated trace elements (AACTE) is an effective fertiliser product that is
currently manufactured and marketed in China, but production is limited by the
availability of protein sources. Liu et al. (2009) describe a novel technique for the
preparation of AACTE from aerobic biological wastewater treatment sludge. The
technology involved several stages of extensive chemical processing of the sludge,
including protein extraction and hydrolysis to produce amino acids, amino acid
purification and chelation with trace elements, and drying. However, an economic
assessment showed that the treatment costs were favourable compared to the
market value of the fertiliser product. The residual sludge generated by the process
(representing approximately 50% of the input DS) was transformed by sintering with
coal fly ash and iron tailings to produce a bioceramic material, which was effective
at ammonium and phosphate sorption from wastewater, and was also
economically viable.

Specialist Chemicals

Biopolymers

Bacterial polyhydoxyalkanoates (PHAs) are a unique group of polymers that are


produced by a large number of prokaryote species for intracellular metabolic
carbon/energy storage (Laycock et al., 2013). These biodegradable polymer
materials have properties similar to polyethylene and polypropylene and have
considerable industrial potential for the manufacture of bioplastics (Gurieff and
Lant, 2007; Laycock et al., 2013). The microorganisms that accumulate PHAs are
abundant in AS and the AS process therefore provides conditions suitable for PHA
synthesis for commercial exploitation (Coats et al., 2007). Poly-3-hydroxybutyrate
(PHB) is the most abundant PHA and, in wastewater treatment, has an important
physiological role in EBPR processes (Horan, 1990). A proposed production system for
PHA using the AS process is shown in Figure 10.

An increasing volume of research has focussed on optimising PHA production from


AS. Polyhydroxyalkanoates accumulate in both conventional and anaerobic-
aerobic AS (Takabatake et al., 2002). Chua et al. (2003) reported the rate of PHA
synthesis in biomass from conventional anaerobicaerobic processes acclimatised
only with wastewater was 20% in the DS. However, accumulation rates greater than
80% are required for economic viability. Currently, maximum accumulations rates of
up to 60% have been obtained in AS supplied with a VFA source (eg by direct VFA
chemical dosing or fermentation of primary or secondary sludges, or other organic
waste sources) and with other process optimisations (Rhu et al., 2003; Mahapatra et
al., 2007; Menmeng et al., 2009; Chen et al., 2013). Neverthess, the production costs
of microbial PHAs are almost 10 times higher than for oil-based plastics (Tyagi and
Lo, 2013) and this is reflected in the market price; for example, the current market
price for PHB is 5000 /t (Bio-plastics, 2013) compared to 1,140 /t for low density
polyethylene (Platts, McGraw Hill Financial, 2014). The carbon source represents
approximately 50% of the costs of production of industrially produced microbial
PHAs, and whilst PHA production by biological wastewater treatment may provide
significant commercial advantages (Serafim et al., 2008), downstream extraction
and processing of the polymer have an important financial impact (Gurieff and
Lant, 2007). Interestingly, however, based on a detailed lifecycle assessment, Gurieff
and Lant (2007) found that, as the loading rate increased (at constant volume) the
PHA option became increasingly more favourable than the biogas option for
wastewater treatment because PHA was a more valuable product whereas most of
the income generated by the biogas option was derived based on financial
savings. The environmental impact of biopolymer production was also significantly
lower than polyethylene production.

Settled
Raw WW sewage Effluent
Primary Activated Secondary
sedimentation sludge process sedimentation

Return activated sludge

Waste
*Pretreatment activated
sludge

Two-stage AD VFA Concentrated VFA


liquor liquor
Sludge PHA
Struvite
fermentation production
recovery
bioreactor

Second stage
PHA-
enriched
biomass

Biogas Biosolids

Figure 10: Proposed integrated wastewater treatment scheme for production of PHAs
from activated sludge (adapted from Chua et al., 2003 and Coats et al., 2007); *If
activated sludge is optionally used to supply VFAs, the efficiency of fermentation is
increased by applying a pre-treatment process

Using AS as the mixed microbial culture and substrate source for PHA production has
many advantages, however, further optimisation is necessary to increase process
efficacy, organic loading rate, and biomass PHA enrichment, as well as to reduce
the energy consumption involved in downstream processing (Gurieff and Lant, 2007;
Serafim et al., 2008; Tyagi and Lo et al., 2013). Further work is also necessary on the
associated residuals management and nutrient resource recovery after PHA
extraction, and to develop simple polymer recovery processes and applications for
PHA where purity of the polymer is less critical (Serafim et al., 2008). Nevertheless,
predictions are optimistic about the potential of PHA production from AS (Chua et
al., 2003; Coats et al., 2007; Gurieff and Lant, 2007) and it is plausible that this
technology will become part of a more integrated strategy to maximise the
resource efficiency potential of AS.
Biopesticides

The value of the biopesticide market in Europe in 2011 was 325 M and microbial
biopesticides represented almost 14% of total sales (AGROW, 2013). The demand for
biopesticides has increased significantly since 2000 with an annual growth rate
equivalent to approximately 17% due to better scientific support and user
confidence, and pressure to adopt more sustainable agricultural practices. Overall,
biopesticides account for approximately 5% of the European agrochemical market
(AGROW, 2013). The UK retail price of commercial insecticides formulated with
Bacillus thuringiensis is approximately 75/kg.

Activated sludge contains a source of organic carbon and other essential nutrients
suitable for supporting microbial growth as a substrate and culture media for various
biological processes. An example of this is the culturing and growth of B. thuringiensis
(Bt), a naturally occurring and effective biopesticide agent that produces various
crystal proteins that have highly specific insecticidal activity (Andrews et al., 1987;
Shieh, 1988; Bravo et., 2011). The use of AS as a production system for B. thuringiensis
and the effects of process conditions on the stability and entomotoxicity have been
investigated by several authors (for example, Sachdeva et al., 2000; Vidyarthi et al.,
2002; Brar et al., 2004).

Bacillus thuringiensis is cultured in sterile liquid sludge (in some cases hydrolysed) to
produce a stable aqueous suspension with similar or better insecticidal properties
compared to standard commercial formulations (Sachdeva et al., 2000; Brar et al.,
2004). Furthermore, Sachdeva et al. (2000) estimated the costs of using sludge to
produce B thuringiensis were approximately 40% smaller compared to a standard
synthetic medium. Pre-treatment is simpler compared to using other organic
substrates and sludge provides a complete nutrient source without further
supplementation. The overall best performance and product efficacy was
obtained at a sludge solids concentration of 2.5% (Vidyarthi et al., 2002).

Strictly, agricultural application of sludge-based products would be regulated under


the Sludge Directive (CEC, 1986), which could present a barrier to the practicable
end-use of biopesticides produced from AS, unless specific end-of-waste criteria
could apply. Furthermore, it is unlikely that the Water Industry would be well
positioned to enter into industrial-scale biopesticide production from sludge.
However, this is an example of a potential opportunity to diversify the management
routes and markets available for sludge products. Consequently, this and other
specialist recovery opportunities could potentially be realised more effectively
through the formation of industrial partnerships to share resources, technologies and
expertise to maximise the value addition to and resource efficiency of AS.

Protein and enzymes

Using sludge directly as a feed for animals, poultry and fish has been proposed, but
the small content of digestible nutrients and metabolisable energy render the
material unsuitable as a constituent of high energy diets, and sludge may also
interfere with animal health and nutrition, and food quality for human consumption
(Webber, 1991). An alternative approach is to recover components of nutritional
value from sludge. For example, AS can contain more than 60% (DS) of microbial
proteins (Table 1) and reported recovery efficiencies for extracting intracellular
proteins from WAS are in the range 80 - 90% (Chishti et al., 1992; Hwang et al., 2008).
However, use in animal feeds is unlikely to be an acceptable or viable approach to
AS resource utilisation due to concerns about dietary impacts and also because
economical nutrient supplements are available that avoid potential dietary
contamination issues (Tyagi and Lo, 2013).

However, extracting enzymes from AS for use as industrial chemicals has significantly
more potential and also because they can be used to increase hydrolysis rates and
biogas production in AD processes (Nabarlatz et al., 2008). The global market for
industrial enzymes was estimated at 2 billion in 2010 and is expected to increase to
2.6 billion by 2015 (BCC Research, 2011). Protease and lipase represent two of the
main intracellular enzymes detected in AS and have been successfully extracted by
a combination of non-ionic detergent and/or cation exchange resin treatments and
agitation (Gessesse et al., 2003; Nabarlatz et al., 2008) and these enzymes are
amongst the most widely used by industry (Haki and Rakshit, 2003). Indeed,
proteases constitute more than 65% of the global industrial enzyme market and are
extensively used in the food, pharmaceutical, leather, textile and detergent
industries. Lipases are the most versatile enzymes and are applied in a range of
bioconversion reactions and they are extensively used in the dairy, oleochemical
(eg biodiesel), detergent, pulp, pharmaceutical, cosmetic and leather industries
(Hou, 2002). Culture media represent up to 40% of the production costs of industrial
enzymes, therefore, using AS for this purpose could represent a significant potential
beneficial saving (Tyagi and Lo, 2013).

Heavy metals

The extraction of heavy metals directly from sludge or incinerator ash has been
recognised as a technically feasible recovery option since the early 1980s (Webber,
1991) and was reviewed most recently by Tyagi and Lo (2013). For example,
recoveries of individual elements of >97% by ultrasonication assisted acid leaching
have been achieved at both laboratory and pilot scale with high metal industrial
sludges. Tyagi and Lo (2013) describe a commercial plant in China which has
applied this technology to recover metals from sludge produced by printed circuit
board factories. Industrial sludges from electronics and metal based industries
potentially contain large metal concentrations and consequently represent a
potential environmental hazard. Therefore, metal extraction has both a financial
and environmental benefit by recovering a valuable resource and also reducing the
potential hazardous nature of the sludge for disposal. Neither of these conditions
usually apply to AS from municipal WWTP, however, since the metal content is below
critical environmental thresholds; therefore recovery is less economical compared to
high-metal contaminated industrial sludges and the material can be recycled
directly for beneficial uses following appropriate controls and procedures. Metal
concentrations increase in the residual ash following sludge incineration,
nevertheless, although a technically feasible option, no examples of the recovery of
metals from sludge or incinerator ash could be found.
Nanochemicals

The feedstock material for nanochemical production represents a major cost in the
manufacturing process and large volumes of inexpensive precursors are required
(Yuvakkumara et al., 2014). Recent research has therefore focussed on the recovery
of high-value, specialised nano-chemicals, for example, high purity nano-SiO2, from
waste organic and mineral residues (Kim et al., 2010; Yuvakkumara et al., 2014). Zou
et al. (2012; 2013) recently demonstrated the potential synthesis of high purity nano-
SiO2 and nano-Al(OH)3 from carbonised AS by a chemical extraction procedure. The
efficiency of the recovery process depends on the total concentrations of the main
elements present in the sludge but, unfortunately, no data were provided on the
sludge chemical properties. By way of reference, Eriksson (2001) reported the mean
total concentrations of Si and Al in Swedish sewage sludges were 4.5 and 4.0% (DS),
respectively. Treatment and/or disposal of process effluents are not reconciled by
Zou et al. (2012; 2013) or Yuvakkumara et al. (2014) and the nano-chemical
recovery procedures from organic residues require significant energy inputs for high
temperature treatments. Nevertheless, they are considered to be more economical
and resource efficient compared to conventional production routes.

Nano-silica powder is used in a wide range of high-value applications including:


ceramics, chemicals, catalysis, chromatography, energy, electronics, coatings,
stabilisers, emulsifiers and biological sciences. Aluminium hydroxide is largely
converted to A12O3 for the production of Al metal, it is also used as a feedstock for
producing other Al compounds manufacture and as a fire retardant.

Fibre products

Significant quantities of cellulose fibres originating primarily from toilet paper


represent a major component of the COD load to WWTP (25 30%). Ruiken et al.
(2013) investigated the use of fine mesh sieves (<0.35 mm) as a pre-treatment for the
AS process for fibre recovery, to replace conventional sedimentation. Applying
sieving to dry weather flow only, reduced the overall energy consumption for
wastewater and sludge treatment by 40% and the pay-back period was 7 years.
Thermographic measurements indicated that cellulose represented 79% of the total
mass of material collected by sieving, and 84% of the organic matter, with inorganic
material corresponding to 6% of the total mass. The rate of suspended solids removal
from raw wastewater was 40 - 50% (similar to sedimentation), suggesting that 35 -
40% of the influent suspended solids originated from toilet paper. Primary sludge
from conventional sedimentation also contains cellulose, but the amounts are
considerably smaller (25 32% of total mass, and 32 38% of the organic fraction)
than sieved material. In contrast to sieving, therefore, <50% of the cellulose was
removed by primary sedimentation. Thus, settled sewage following conventional
primary sedimentation contains significantly more cellulose than the effluent from
the fine sieve. Cellulose biodegradation is relatively poor in secondary wastewater
treatment (Verachtert et al., 1982) and does not contribute significantly to the
biological aeration demand. However, it is strongly dependent on temperature and
retention time. In batch AD tests, for instance, Ruiken et al. (2013) measured the
complete biodegradation of different fibre sources at 24 - 30oC with retention times
of 8 - 20 days. Therefore, at typically higher AD operating temperatures (35 39oC)
and retention times (~15 days), it is likely that the conversion to biogas will be
extensive potentially weakening the financial and energy case for separate fibre
recovery. Indeed other investigations indicate increasing solids loading to AD is
beneficial to the overall energy balance of wastewater treatment (Shi et al., 2011).
Suggested end-uses for recovered cellulose fibres include application as an
agricultural soil conditioner, as a biomass fuel source and for reprocessing as toilet
tissue (Ruiken et al., 2013). All of these require additional processing for pathogen
reduction, dewatering or fibre processing recovery, respectively, with additional
resource consumption implications.

Other Beneficial Resource Recovery

Novel Microorganisms

The unique microbiological characteristics of AS sludge are a major reason for the
effectiveness of the process in wastewater treatment. The microbial biodiversity of
AS and in anaerobic sludge treatment is highly dynamic and provides a significant
resource capable of performing important biotransformation and bioaccumulation
activities, including P and PHA accumulation, for instance, and fermentation and
methanogenic reactions. Many studies demonstrate this diversity and the adaption
of AS bacteria to biodegradation of anthropogenic organic chemicals, for
example, pyrethroid pesticides; herbicides; industrial solvents, chemicals and
precursors, and endocrine disruptors (Chong and Chen, 2007; Porter and Hay, 2007;
Guo et al., 2009; Jeong et al., 2008; Wang et al., 2009).

Activated sludge bacteria are also a source of virus-binding proteins (VBPs) which
can be applied in new technologies for the potential control of environmental
sources of infectious viral diseases (Sano et al., 2004). The predatory activity of
bacteriophagous protozoa in the AS process is also an important mechanism for
eliminating pathogenic bacteria (Figure 11), as shown in the classical experiments of
Curds and Fey (1969). Although not a complete barrier to transmission, these
biological properties and mechanisms, coupled with secondary sludge
sedimentation, which also physically removes larger (oo)cysts and eggs of parasitic
protozoa and helminths (Feachem et al., 1983), explain why AS has reduced the risk
to human health from infectious microorganisms potentially present in urban
wastewater.

Construction Materials

Using sewage sludge directly or the ash from sludge incineration as a substitute for
standard minerals in construction materials has been a long-established alternative
approach to the main sludge management routes (Webber, 1991). Whilst extensive
research has been completed to develop and improve production technologies
(Donatello and Cheeseman, 2013), their industrial uptake remains extremely limited.
This is because most of the technologies are uneconomic compared to readily
available standard aggregate and other feedstock minerals (Tyagi and Lo, 2013).
Sludge incineration ash is used to manufacture bricks at modest industrial scale in
Japan (Tyagi and Lo, 2013) and there was some commercial interest shown in the
UK (Anderson et al., 1996). A novel construction product developed by Forth et al.
(2006) is the Bitublock, which is produced by combining a blend of different waste
materials, including sewage sludge incineration ash, recycled glass and steel slag
with a bitumen binder. However, due to the porosity of the ash, the proportion of
bitumen binder required increased from 5% without ash inclusion to 16% at an
incorporation rate equivalent to 37.5%; this may therefore constrain the maximum
amount of ash that may be potentially used. Nevertheless, Forth et al. (2006)
projected that, assuming a filler blending rate of 10%, all of the ash produced from
sludge incineration could be utilised in block manufacture to meet the demand for
house construction in the UK.

Figure 11: Effect of bacteriophagous protozoa on E. coli removal from wastewater


and activated sludge (Curds and Fey, 1969)

Adsorbents and Filtration

The use of AS specifically and sewage sludge in general, as adsorbents for municipal
and industrial wastewater filtration is another alternative application avenue that
has been extensively research (for a comprehensive review, see Smith et al., 2009).
Activated sludge has a high propensity for metal adsorption and this characteristic
has advantages for the treatment of industrial, metal laden wastewaters (Dobson
and Burgess, 2007; Hammaini et al., 2007). This characteristic is associated mainly
with the capacity of AS for biosorption of metals, which is a physico-chemical
process based on ion exchange, complexation and surface microprecipitation
reactions involving a large variety of binding sites of extracellular polymeric
substances (EPS) and bacterial cell surfaces (Pagnanelli et al., 2009). Biosorption
mechanisms are very effective at removing heavy metals from wastewater and can
achieve removal efficiencies of 75100% (Dobson and Burgess, 2007). Metal
precipitation reactions with AS are also potentially important and control the
solubility of other elements, such as Pb (Pagnanelli et al., 2009).

Alginate is a group of linear polysaccharides composed of -L-guluronic acid and -


D-mannuronic acid residues produced by seaweed and bacteria. Seaweed
aliginates have various pharmaceutical and industrial applications including the
biosorption of heavy metals. Activated sludge has a high bacterial cell content and
the potential of AS as an alternative source for alginate extraction was
demonstrated by Zhang et al. (2009). The potential yield of Ca-alginate from AS
measured by Zhang et al. (2009) was approximately 10% by dry weight. The AS-
alginate was similar to algal material and was significantly better than commercial
activated carbon or bentonite products at Cu2+ adsorption.

Dried waste treatment sludges and AS, and sludge carbonised and chemically
activated by acid treatment have been examined in various water filtration
applications as low-cost adsorbents for metals, organic and other industrial
chemical contaminants (Gupta and Suhas, 2009; Smith et al., 2009; Bhatnagar and
Sillanp, 2010). Activated carbons represent the most sophisticated filtration
materials produced from sludge involving high temperature (500 1000 oC) pyrolytic
carbonisation and chemical activation steps (Smith et al., 2009). The energy
demand for the process can also be met through the combustion of generated
syngas. Reducing the inorganic content of carbonised char by acid washing with
HCl significantly improves performance and also enables the potential recovery of
other valuable resources including phosphate minerals. Chemical activation utilising
alkali metal hydroxides (KOH) is considered the most effective method for producing
high surface area adsorbents from sludge (Smith et al., 2009). This technology
represents another example of the opportunities available for exploiting wastewater
treatment sludge as a resource, and where scale-up to industrial application has yet
to occur.

Conclusions and Forward Vision

The AS process is a highly successful approach to wastewater treatment and has


contributed significantly to improving the quality of the water environment.
However, it also represents a major sink for energy consumption by the Water
Industry.

In future, therefore, there is likely to be greater emphasis placed on production of


biorenewable energy from anaerobic wastewater treatment and reduced
dependency on AS, but the operation of the existing AS infrastructure and as an
aerobic polishing process is likely to continue.

The energy balance of the AS process can be improved in the short term through
better process operation, in the medium term by expanding the current interest
in recovering energy from the residual sludge and in the longer-term by the
recovery of other valuable industrial chemicals from AS.

Energy recovery from AS by AD will increase and maximum biogas yields will be
achieved through the improved understanding and management of the
digestion process and by applying biomass disruption techniques. A life cycle
assessment is necessary to determine the most efficient techniques that maximise
the energy yield with the least resource expenditure. A combination of efficiency
improvements in the AS process with measures to maximise energy recovery by
sludge AD can bring the overall wastewater treatment process close to energy
self-sufficiency.

Recycling sludge in agriculture and for other land application purposes is already
a well established practice for the beneficial recovery of major plant nutrients (N,
P and S) and using sludge as an agricultural fertiliser is widely regarded as the
best practicable environmental option for sludge management. The AS process
is potentially very efficient at capturing P and maximising P recovery by this route
is necessary to assist in establishing a circular economy for this critical and finite
nutrient resource to ensure long-term food security. The displacement of high-
energy mineral fertilisers by nutrients supplied in sludge also provides a positive
contribution to the energy and GHG balance of wastewater and sludge
treatment.

In contrast to P, the AS process is relatively inefficient at capturing N in


wastewater, which is derived mainly from the use of high-energy N fertilisers for
food production. The treatment strategy has been to utilise more energy to
remove the N in wastewater by transforming soluble N microbiologically to the
gaseous phase. Future research is necessary to investigate potential approaches
to conserve N in the solid phase to maximise the efficient recovery of N with high
embedded energy from wastewater as a resource.

The unique properties of the AS process and AS can also be exploited to


produce valuable industrial chemicals, to displace oil-based materials and/or
provide alternative cost-effective feedstocks, as supplementary resource
recovery options to the renewable energy and agricultural routes.

Producing biopolymers and enzymes could potentially offer significant


opportunities for the recovery of high value industrial resources from AS, however,
other options, such as biopesticides or recovering heavy metals may be less
attractive. These opportunities could potentially be most effectively realised by
developing strategic industrial partnerships to share resources, technologies and
expertise to maximise the value addition and resource recovery potential of AS.

A combination of integrated recovery methods is likely to be applied to


maximise energy and resource efficiency and is a practicable prospect for AS in
future. For example, the production of PHA-based biopolymers from AS, coupled
with the recovery of renewable energy by AD (methane) in the short term, or
fermentation (hydrogen) in the longer-term, and agricultural fertiliser production
(eg biosolids, concentrated nutrient products from thermal-precipitation, AD
sidestream and dewatering liquor treatment, such as calcium phosphate,
ammonium sulphate and struvite). The success and viability of these will depend
on the financial viability, energy consumption, and chemical quality, compared
to conventional production methods.

References

ADAS (2001) The Safe Sludge Matrix. Guidelines for the application of sewage sludge
to arable and horticultural crops. https://www.gov.uk/managing-sewage-sludge-
slurry-and-silage [accessed: 5 Feb 2014]
AGROW (2013) European biopesticide market worth $540 M in 2011. No 666, June
19th 2013 [on-line]. http://www.cplconsult.com/wp-
content/uploads/2012/09/Agrow-Article-Redact.pdf [accessed: 13 Feb 2014]

Ahn, Y-H. (2006) Sustainable nitrogen elimination biotechnologies: A review. Process


Biochemistry, 41, 1709-1721.

AIC; Agricultural Industries Federation (2013) Fertiliser Statistics 2013. Agricultural


Industries Federation, Peterborough.
http://www.agindustries.org.uk/sectors/fertiliser/uk-fertiliser-consumption-trends-and-
statistics/ [accessed: 6 Feb 2014]

Ainger, C., Butler, D., Caffor, I., Crawford-Brown, D., Helm, D. and Stephenson, T.
(2009) A Low Carbon Water Industry in 2050. Report: SC070010/R3. Environment
Agency, Bristol. http://www.environment-
agency.gov.uk/research/library/publications/114393.aspx [accessed: 25 Feb 2014]

Anderson, M., Skerratt, R.G., Thomas, J.P. and Clay, S.D. (1996) Case study involving
using fluidised bed incinerator sludge ash as a partial clay substitute in brick
manufacture. Water Science and Technology, 34, 507-515.

Andrews, R.E., Faust, R.M., Wabiko, H., and Raymond, K.C. (1987) The biotechnology
of Bacillus thuringiensis. CRC Critical Reviews in Biotechnology, 6, 163-232.

ANZBP; Australian and New Zealand Biosolids Partnership (2013) Biosolids Production
in Australia: The Australian and New Zealand Biosolids Partnership National Survey of
Biosolids Production and End Use 2013. Australian Water Association, Sydney,
Australia.

Appels, L., Baeyens, J., Degrve, J. and Dewil, R. (2008) Principles and potential of
the anaerobic digestion of waste-activated sludge. Progress in Energy and
Combustion Science, 34, 755-781

Appels, L., Lauwers, J., Degrve, J., Helsen, L., Lievens, B., Willems, K., Van Impe, J.
and Dewil, R. (2011) Anaerobic digestion in global bio-energy production: Potential
and research challenges. Renewable and Sustainable Energy Reviews, 15, 4295-
4301.

Astals, S., Esteban-Gutirrez, M., Fernndez-Arvalo, T., Aymerich, E., Garca-Heras,


J.L. and Mata-Alvarez, J. (2013) Anaerobic digestion of seven different sewage
sludges: A biodegradability and modelling study. Water Research, 47, 6033-6043.

BCC Research (2010) Enzymes in industrial applications: Global markets [on-line].


http://www.bccresearch.com/market-research/biotechnology/enzymes-industrial-
applications-bio030f.html [accessed: 14 Feb 2014]
Bhatnagar, A. and Sillanp, M. (2010) Utilization of agro-industrial and municipal
waste materials as potential adsorbents for water treatment A review. Chemical
Engineering Journal, 157, 277-296.

Bio-plastics (2013) Polyhydroxybutyrate (PHB) [on-line]. http://www.bio-


plastics.org/en/information--knowledge-a-market-know-how/bioplastic-
types/polyhydroxyalkanoates [accessed: 12 Feb 2014]

Bolzonella, D., Pavan, P., Battistoni, P. and Cecchi, F. (2005) Mesophilic anaerobic
digestion of waste activated sludge: influence of the solid retention time in the
wastewater treatment process. Process Biochemistry, 40, 1453-1460.

Brar, S.K., Verma, M., Tyagi, R.D., Valro, J.R., Surampalli, R.Y. and Banerji, S.K. (2004)
Development of sludge based stable aqueous Bacillus thuringiensis formulations.
Water Science and Technology, 50, 229-236.

Bravo, A., Likitvivatanavong, S, Gill, S.G. and Sobern, M. (2011) Bacillus thuringiensis:
A story of a successful bioinsecticide. Insect Biochemistry and Molecular Biology, 41,
423-431.

Carrre, H., Dumas, C., Battimelli, A., Batstone, D.J., Delgens, J.P., Steyer, J.P. and
Ferrer, I. (2010) Pretreatment methods to improve sludge anaerobic degradability: A
review. Journal of Hazardous Materials, 183, 1-15.

CEC; Council of the European Communities (1986) Council Directive of 12 June 1986
on the protection of the environment, and in particular of the soil, when sewage
sludge is used in agriculture (86/278/EEC). Official Journal of the European
Communities, No. L 181/6-12.

Chen, H., Meng, Z., Nie, Z. and Zhang, M. (2013) Polyhydroxyalkanoate production
from fermented volatile fatty acids: Effect of pH and feeding regimes. Bioresource
Technology, 128, 533-538.

Chen, Y.G., Jiang, S., Yuan, H.Y., Zhou, Q. and Gu, G.W. (2007) Hydrolysis and
acidification of waste activated sludge at different pHs. Water Research, 41, 683-
689.

Cheong, D-Y., Hansen, C.L. and Stevens, D.K. (2007) Production of bio-hydrogen by
mesophilic anaerobic fermentation in an acid-phase sequencing batch reactor.
Biotechnology and Bioengineering, 96, 421-432.

Chishti, S.S., Hasnain, S.N. and Khan, M.A. (1992) Studies on the recovery of sludge
protein. Water Research, 26, 241-248.

Chong, N-M. and Chen, Y-S. (2007) Activated sludge treatment of a xenobiotic with
or without a biogenic substrate during start-up and shocks. Bioresource Technology,
98, 3611-3616.
Chua, A.S.M., Takabatake, H., Satoh, H. and Mino, T. (2003) Production of
polyhydroxyalkanoates (PHA) by activated sludge treating municipal wastewater:
effect of pH, sludge retention time (SRT), and acetate concentration in influent.
Water Research, 37, 3602-3611.

CIWEM; Chartered Institution of Water and Environmental Management (1995a)


Sewage Sludge: Utilization and Disposal. Handbooks of UK Wastewater Practice.
CIWEM, London.

CIWEM; Chartered Institution of Water and Environmental Management (1995b)


Sewage Sludge: Introducing Treatment and Management. Handbooks of UK
Wastewater Practice. CIWEM, London.

Coats, E.R., Loge, F.J., Wolcott, M.P., Englund, K. and McDonald, A.G. (2007)
Synthesis of polyhydroxyalkanoates in municipal wastewater treatment. Water
Environment Research, 79, 2396-2403.

Collinge, V.K. and Bruce, A.M. (1981) Sewage Sludge Disposal: A Strategic Review
and Assessment of Research Needs. Technical Report TR 166. Water Research
Centre (WRc), Swindon.

Cordell, D., Drangert, J-O., White, S. (2009) The story of phosphorus: Global food
security and food for thought. Global Environmental Change 19, 292-305.

Curds, C.R. and Fey, G.J. (1969) The effect of ciliated protozoa on the fate of
Escherichia coli in the activated-sludge process. Water Research, 3, 853-867.

DairyCo (2014) Fertiliser prices [on line]. http://www.dairyco.org.uk/resources-


library/market-information/farm-expenses/fertiliser-prices/ [accessed: 5 Feb 2014]

Defra; Department of Food and Rural Affairs (2010) Fertiliser Manual (RB209). 8th
Edition. The Stationery Office, Norwich.

Dentel, S.K., Strogen, B. and Chiu, P. (2004) Direct generation of electricity from
sludges and other liquid wastes. Water Science and Technology, 50, 161-168.

Dewil, R., Baeyens, J., Roels, J. and Van De Steene, B. (2008) Distribution of sulphur
compounds in sewage sludge treatment. Environmental Engineering Science, 25,
879-886.

Dichtl, N., Rogge, S. and Bauerfeld, K. (2007) Review Novel Strategies in Sewage
Sludge Treatment. Clean, 35, 473 479.

Dobson, R.S. and Burgess, J.E. (2007) Biological treatment of precious metal refinery
wastewater: A review. Minerals Engineering, 20, 519-532.
Dohanyos, M., Zabranska, J. and Jenicek, P. (1997) Enhancement of sludge
anaerobic digestion by use of a special thickening centrifuge. Water Science and
Technology, 36, 145-153.

Dolan, E. (2013) Doomsday: Will peak phosphate get us before global warming?
[online] http://oilprice.com/Metals/Foodstuffs/Doomsday-Will-Peak-Phosphate-Get-
us-Before-Global-Warming.html [accessed: 5 Feb 2014]

Donatello, S. and Cheeseman, C.R. (2013) Recycling and recovery routes for
incinerated sewage sludge ash (ISSA): A review. Waste Management, 33, 2328-2340.

EA; Environment Agency (2014) Anaerobic digestion (biogas) [on line].


http://www.environment-agency.gov.uk/business/sectors/32601.aspx [accessed: 9
Feb 2014]

EC; European Commission (2001) Disposal and Recycling Routes for Sewage Sludge.
Part 3 Scientific and Technical Report. European Commission, Luxembourg.
http://ec.europa.eu/environment/waste/sludge/pdf/sludge_disposal3.pdf
[accessed: 4 Feb 2014]

Eriksen, G.S., Amundsen, C.E., Bernhoft, A., Eggen, T., Grave, K., Halling-Srensen, B.,
Kllqvist, T., Sogn, T. and Sverdrup, L. (2009) Risk Assessment of Contaminants in
Sewage Sludge Applied on Norwegian Soils. Opinion of the Panel on Contaminants
in the Norwegian Scientific Committee for Food Safety. Norwegian Scientific
Committee for Food Safetly (VKM), Oslo, Norway.

Eriksson, J. (2001) Concentrations of 61 Trace Elements in Sewage Sludge, Farmyard


Manure, Mineral Fertiliser, Precipitation and in Oil and Crops. Report No. 5159.
Swedish Environmental Protection Agency, Stockholm.

ETRAC; Environment, Transport and Regional Affairs Committee (1998) Sewage


Treatment and Disposal. Second Report. HC 266-I. The Stationery Office, London.

Eurostat (2013) Water Statistics.


http://epp.eurostat.ec.europa.eu/statistics_explained/index.php/Water_statistics
[accessed: 29 Jan 2014]

Evans, T. (2014) Phosphorus: time to act. The Chemical Engineer, 1st February.
Institution of Chemical Engineers, London.

Feachem, R.G., Bradley, D.J., Garelick, H. and Mara, D.D. (1983) Sanitation and
Disease: Health Aspects of Excreta and Wastewater Management. World Bank
Studies in Water Supply and Sanitation 3. John Wiley & Sons, Chichester.

Forth, J.P., Zoorob, S.E. and Thanaya, I.N.A. (2006) Development of bitumen-bound
waste aggregate building blocks. Proceedings of the Institution of Civil Engineers
Construction Materials, 159, 23-32.
Frost, R.C. and Campbell, H.W. (1986) Alternative uses of sewage sludge. In
LHermite, P.L. (ed) Processing and Use of Organic Sludge and Liquid Agricultural
Wastes, pp. 94-109. D. Reidel, Dordrecht.

Gale, P. (2005) Land application of treated sewage sludge: quantifying pathogen


risks from consumption of crops. Journal of Applied Microbiology 98, 380-396.

Gaterell, M.R., Gay, R., Wilson, R., Gochin, R.J. and Lester, J.N. (2000) An economic
and environmental evaluation of the opportunities for substituting phosphorus
recovered from wastewater treatment works in existing UK fertiliser markets.
Environmental Technology, 21, 1067-1084.

Georges, K., Thornton, A. and Sadler, R. (2009) Transforming Wastewater Treatment


to Reduce Carbon Emissions. Report: SC070010/R2. Environment Agency, Bristol.
http://www.environment-agency.gov.uk/research/library/publications/114393.aspx
[accessed: 29 Jan 2014]

Gessesse, A., Dueholm, T., Petersen, S.B. and Nielsen, P.H. (2003) Lipase and protease
extraction from activated sludge. Water Research, 37, 3652-3657.

Green Alliance (2007) The Nutrient Cycle: Closing the Loop. Green Alliance, London

Guo, P., Wang, B., Hang, B., Li, L., Ali, S.W., He, J. and Li, S. (2009) Pyrethroid-
degrading Sphingobium sp. JZ-2 and the purification and characterization of a
novel pyrethroid hydrolase. International Biodeterioration and Biodegradation, 63,
1107-1112.

Gupta, V.K. and Suhas, (2009) Application of low-cost adsorbents for dye removal
A review. Journal of Environmental Management, 90, 2313-2342.

Gurieff, N. and Lant, P. (2007) Comparative life cycle assessment and financial
analysis of mixed culture polyhydroxyalkanoate production. Bioresource
Technology, 98, 3393-3403.

GWI; Global Water Intelligence (2012) Sludge Management: Opportunities in


growing volumes, disposal restrictions and energy recovery.
http://www.globalwaterintel.com/market-intelligence-reports/sludge-management/
[accessed: 27 Jan 2014]

Haki, G.D. and Rakshit, S.K. (2003) Developments in industrially important


thermostable enzymes: a review. Bioresource Technology, 89, 17-34.

Hammaini, A., Gonzlez, F., Ballester, A., Blzquez, M.L. and Muoz. J.A. (2007)
Biosorption of heavy metals by activated sludge and their desorption
characteristics. Journal of Environmental Management, 84, 419-426.
Hong, K-J., Tarutani, N., Shinya, Y. and Kajiuchi, T. (2005) Study on the recovery of
phosphorus from waste-activated sludge incinerator ash. Journal of Environmental
Science and Health, 40, 617-631.

Hou, C.T. (2002) Industrial uses of lipase. In Kuo, T.M. and Gardner, H.W. (eds) Lipid
Biotechnology. Marcel Dekker, New York, 387-397.

Horan, N.J. (1990) Biological Wastewater Treatment Systems Theory and Operation.
John Wiley & Sons, Chichester.

Howe, A. (2009) Renewable Energy Potential for the Water Industry. Report:
SC070010/R5. Environment Agency, Bristol. http://www.environment-
agency.gov.uk/research/library/publications/114393.aspx [accessed: 29 Jan 2014]

Hwang, J., Zhang, L., Seo, S., Lee, Y.W. and Jhang, D. (2008) Protein recovery from
excess sludge for its use as animal feed. Bioresource Technology, 99, 8949-8954.

ICON; IC CONSULTANTS (2001) Pollutants in Urban Wastewater and Sewage Sludge


Final Report to Directorate General Environment, European Commission.
http://ec.europa.eu/environment/waste/sludge/ [accessed: 29 Feb 2014]

IPCC; Intergovernmental Panel on Climate Change (2007) Climate Change 2007:


Mitigation. Contribution of Working Group III to the Fourth Assessment Report of the
Intergovernmental Panel on Climate Change. Metz, B., Davidson, O.R., Bosch, P.R.,
Dave, R. and Meyer L.A. (eds). Cambridge University Press, Cambridge.
http://www.ipcc.ch/publications_and_data/publications_ipcc_fourth_assessment_r
eport_wg3_report_mitigation_of_climate_change.htm [accessed: 5 Feb 2014]

Jardin, N. and Ppel, H.J. (1994) Phosphate release of sludges from enhance
biological P-removal during digestion. Water Science and Technology, 30, 281-292.

Jenicek, P., Kutil, J., Benes, O., Todt, V., Zabranska, J. and Dohanyos, M. (2013)
Energy self-sufficient sewage wastewater treatment plants: is optimized anaerobic
sludge digestion the key? Water Science and Technology, 68, 1739-1744.

Jeong, E., Hirai, M. and Shoda, M. (2008) Removal of o-xylene using biofilter
inoculated with Rhodococcus sp. BT062. Journal of Hazardous Materials, 152, 140-
147.

Kim, S-J., Seo, S-G. and Jung, S-C.(2010) Preparation of high purity nano silica
particles from blast-furnace slag. Korean Journal of Chemical Engineering, 27, 1901-
1905

Kroiss, H. (2004) What is the potential for utilizing the resources in sludge? Water
Science and Technology, 49, 1-10.
Kuroda, A., Takiguchi, N., Gotanda, T., Nomura, K., Kato, J., Ikeda, T. and Ohtake, H.
(2002) Asimple method to release polyphosphate from activated sludge for
phosphorus reuse and recycling. Biotechnology and Bioengineering, 78, 333-338.

Laycock, B., Halley, P., Pratt, S., Werker, A. and Lant, P. (2013) The chemomechanicl
properties of microbial polyhydroxyalkanoates. Progress in Polymer Science, 38, 536-
583.

Le Corre, K.S., Valsami-Jones, E., Hobbs, P. and Parsons, S.A. (2009) Phosphorus
recovery from wastewater by struvite crystallization: A review. Critical Reviews in
Environmental Science and Technology, 39, 433-477.

Lillywhite, R. and Rahn, C. (2005) Nitrogen UK. Warwick HRI, Wellesbourne.

Linji, X., Wenzong, L., Yining, W., Aijie, W., Shuai, L. and Wei, J. (2013) Optimizing
external voltage for enhanced energy recovery from sludge fermentation liquid in
microbial electrylysis cell. International Journal of Hydrogen Energy, 38, 15801-15806.

Liu, D., Zeng, R.J. and Agelidaki, I. (2008) Effects of pH and hydraulic retention time
on hydrogen production versus methanogenesis during anaerobic fermentation or
organic household solid waste under extreme-thermophilic temperature (70oC).
Biotechnology and Bioengineering, 100, 1108-1114.

Liu, Y., Kong, S., Li, Y. and Zeng, H. (2009) Novel technology for sewage sludge
utilization: Preparation of amino acids chelated trace elements (AATCE) fertilizer.
Journal of Hazardous Materials, 171, 1159-1167.

Liu, Y., Kumar, S., Kwag, J-H. and Raa, C. (2012) Magnesium ammonium phosphate
formation, recovery and its application as valuable resources: a review. Journal of
Chemical Technology and Biotechnology, 88, 181189.

Mahapatra, K., Kumar, M.S. Vaidya, A.N. and Chakrabarti, T. (2007) Production and
recovery process of polyhydroxybutyrate (PHB) from waste activated sludge.
Journal of Environmental Science and Engineering, 49, 164-169.

Manara, P. and Zabaniotou, A. (2012) Towards sewage sludge based biofuels via
thermochemical conversion A review. Renewable and Sustainable Energy
Reviews, 16, 2566-2582.

McAdam, E.J., Lffler, D., Martin-Garcia, N., Eusebi, A.L., Lester, J.N., Jefferson, B. and
Cartmell, E. (2011) Integrating anaerobic processes into wastewater treatment.
Water Science and Technology, 63, 1459-1466.

McCarty, P.L., Bae, J. and Kim, J. (2011) Domestic wastewater treatment as a net
energy producer can this be achieved? Environmental Science and Technology,
45, 7100-7106.
Mengmeng, C., Hong, C., Qingliang, Z., Shirley, S.N. and Jie, R. (2009) Optimal
production of polyhydroxyalkanoates (PHA) in activated sludge fed by volatile fatty
acids(VFAs) generated from alkaline excess sludge fermentation. Bioresource
Technology, 100, 1399-1405.

Miller, M. and OConnor, G.A. (2009) The longer-term phytoavailability of biosolids-


phosphorus. Agronomy Journal, 101, 889-896.

Mininni, G., Di Bartolo Zuccarello, R., Lotito, V., Spinosa, L. and Di Pinto, A.C. (1997) A
design model of sewage sludge incineration plants with energy recovery. Water
Science and Technology 36, 211-218.

Mo, W. and Zhang, Q. (2013) Energy-nutrients-water nexus: Integrated resource


recovery in municipal wasterwater treatment plants. Journal of Environmental
Management, 127, 255-267.

Morris, R, Smith, S.R, Bellett-Travers, D.M. and Bell, J.N.B. (2003) Reproducibility of the
nitrogen response and residual fertiliser value of conventional and enhanced-
treated biosolids. Workshop focussed on research at Imperial College London on
Recycling Nutrient Management Issues. In, Horan, N.J. (Ed) Proceeding of the Joint
CIWEM Aqua Enviro Technology Transfer 8th European Biosolids and Organic
Residuals Conference, 24 26 November, Wakefield.

Morse, G.K., Lester, J.N. and Perry, R. (1993) The Economic and Environmental
Impact of Phosphorus Removal from Wastewater in the European Community.
Selper Publications, Chiswick.

Mudhoo, A., Forster-Carneiro, T. and Snchez, A. (2011) Biohydrogen production


and bioprocess enhancement: A review. Critical Review in Biotechnology, 31, 250-
263.

Nabarlatz, D., Vondrysova, J., Jenicek, P., Stber, F., Font, J., Fortuny, A., Fabregat, A.
and Bengoa, C. (2008) Extraction of enzymes from activated sludge. Waste
Management and the Environment IV, 109, 249-257.

NEBRA; North East Biosolids and Residuals Association (2007) A National Biosolids
Regulation, Quality, End Use & Disposal Survey. Nebra, Tamworth, New Hampshire,
US.

ORiordan, E.G., Dodd, V.A., Tunney, H. and Fleming, G.A. (1987a) The fertiliser
nutrient value of activated sludge under grassland filed conditions. Irish Journal of
Agricultural Research, 26, 213-229.

ORiordan, E.G., Dodd, V.A., Tunney, H. and Flemming, G.A. (1987b) Fertiliser value of
sewage sludge: Phosphorus. Irish Journal of Agricultural Research, 26, 53-61.

degaard, H. (2004) Sludge minimization technologies an overview. Water


Science and Technology, 49, 31-40.
degaard, H., Paulsrud, B. and Karlsson, I. (2002) Wastewater sludge as a resource:
disposal strategies and corresponding treatment technologies aimed at sustainable
handing of wastewater sludge. Water Science and Technology, 46, 295-303.

Oladeji, O.O., OConnor, G.A. and Sartain, J.B. (2008) Relative phosphorus
phytoavailability of different phosphorus sources. Communications in Soil Science
and Plant Analysis, 39, 2398-2410.

Oloffson, U., Bignert, A. and Haglund, P. (2012) Time-trends of metals and organic
contaminants in sewage sludge. Water Research, 46, 4841-4851.

Pagnanelli, F., Mainelli, S., Bornoroni, L., Dionisi, D. and Toro, L. (2009) Mechanisms of
heavy-metal removal by activated sludge. Chemosphere, 75, 1028-1034.

Paredes, D., Kuschk, P., Mbwette, T.S.A., Stange, F., Mller, R.A. and Kser, H. (2007)
New aspects of microbial nitrogen transformations in the context of wastewater
treatment A review. Engineering in Life Sciences, 7, 13-25.

Petzet, S. and Cornel, P. (2011) Towards a complete recycling of phosphorus in


wastewater treatment options in Germany. Water Science and Technology, 64, 29-
35.

Platts, McGraw Hill Financial (2014) Platts Global Low-Density Polyethylene (LDPE)
Price Index [on-line]. http://www.platts.com/news-feature/2012/pgpi/ldpe
[accessed: 12 Feb 2014]

Porter, A.W. and Hay, A.G. (2007) Identification of opdA, a gene involved in
biodegradation of the endocrine disruptor octylphenol. Applied and Environmnetal
Microbiology, 73, 7373-7379.

POST; Parliamentary Office of Science and Technology (2007) Energy and Sewage.
Number 282. http://www.parliament.uk/business/publications/research/briefing-
papers/POST-PN-282/energy-and-sewage-april-2007 [accessed: 29 Jan 2014]

Rhu, D.H., Lee, W.H., Kim, J.Y. and Choi, E. (2003) Polyhydroxyalkanoate (PHA)
production from waste. Water Science and Technology, 48, 221228

Ruiken, C.J., Breuer, G., Klaversma, E., Santiago, T. and van Loosdrecht, M.C.M.
(2013) Sieving wastewater Cellulose recovery, economic and energy evaluation.
Water Research, 47, 43-48.

Rulkens, W. (2008) Sewage sludge as a biomass resource for the production of


energy: overview and assessment of the various options. Energy & Fuels, 22, 9-15.

Rulkens, W.H. (2004) Sustainable sludge management what are the challenges for
the future? Water Science and Technology, 49, 11-19.
Rulkens, W.H. and Bien, J.D. (2004) Recovery of energy from sludge comparison of
the various options. Water Science and Technology, 50, 213-221.

Sachdeva, V., Tyagi, R.D. and Valro, J.R. (2000) Production of biopesticides as a
novel method of wastewater sludge utilization/disposal. Water Science and
Technology, 42, 211-216.

Sano, D., Matsuo, T. and Omura, T. (2004) Virus-binding proteins recovered from
bacterial culture derived from activated sludge by affinity chromatography assay
using a viral capsid peptide. Applied and Environmental Microbiology, 70, 3434
3442.

Satoh, H., Mino, T. and Matsuo, T. (1999) PHA production from activated sludge.
International Journal of Biological Macromolecules, 25, 105-109.

Schowanek, D., David, H., Francaviglia, R., Hall, J., Kirchmann, H., Krogh, P.H.,
Schraepen, N. and Smith, S.R. (2007) Probabilistic risk assessment for linear
alkylbenzene sulfonate (LAS) in sewage sludge used on agricultural soil. Regulatory
Toxicology and Pharmacology, 49, 245-259.

Serafim, L.S., Lemos, P.C., Albuquerque, M.G.E. and Reis, M.A.M. (2008) Strategies for
PHA production by mixed cultures and renewable waste materials. Applied
Microbiology and Biotechnology, 81, 615-628.

Serna-Maza, A., Heavens, S. and Banks, C.J. (2014) Ammonia removal in food waste
anaerobic digestion using a side-stream stripping process. Bioresource Technology,
152, 307-315.

Shi, C.Y. (2011) Mass Flow and Energy Efficiency of Municipal Wastewater Treatment
Plants. IWA Publishing, London.

Shieh, T.R. (1988) Bacillus thuringiensis biological insecticide and biotechnology. ACS
Symposium Series, 362, 207-216.

Siddiquee, M.N. and Rohani, S. (2011) Lipid extraction and biodiesel production from
municipal sewage sludges: A review. Renewable and Sustainable Energy Reviews,
15, 1067-1072.

Smith, K.M., Fowler, G.D., Pullket, S. and Graham, N.J.D. (2009) Sewage sludge-based
adsorbents: A review of their production, properties and use in water treatment
applications. Water Research, 43, 2569-2594.

Smith, S.R. (2009) Organic contaminants in sewage sludge (biosolids) and their
significance for agricultural recycling. Philosophical Transactions of the Royal Society
A 367, 3871-3872.
Smith, S.R. and Hadley, P. (1988) A comparison of the effects of organic and
inorganic nitrogen fertilisers on the growth response of summer cabbage (Brassica
oleracea var. capitata cv Hispi F1). Journal of Horticultural Science, 63, 615-620.

Smith, S.R. and Hadley, P. (1989) A comparison of organic and inorganic nitrogen
fertilisers: Their nitrate-N and ammonium-N release characteristics and effects on the
growth response of lettuce (Lactuca sativa L. cv. Fortune). Plant and Soil, 115, 135-
144.

Smith, S.R. and Hadley, P. (1990) Carbon and nitrogen mineralisation characteristics
of organic nitrogen fertilisers in a soil-less incubation system. Fertiliser Research, 23,
97-103.

Smith, S.R., Morris, R., Bellett-Travers, D.M. and Bell, J.N.B. (2003) Managing the
phosphate supply from sewage sludge amended agricultural soil. Workshop
focussed on research at Imperial College London on Recycling Nutrient
Management Issues. In Horan, N.J. (Ed) Proceeding of the Joint CIWEM Aqua Enviro
Technology Transfer 8th European Biosolids and Organic Residuals Conference, 24
26 November, Wakefield.

Stark, K. (2004) Phosphorus recovery - experiences from European countries,


Integration and optimisation of urban sanitation systems. In Plaza, E., Levlin, E.,
Hultman, B. (eds) Proceedings of Polish-Swedish seminars, 6 8 June, Stockholm.
http://www2.lwr.kth.se/forskningsprojekt/Polishproject/rep12/StarkSthlm19.pdf
[accessed: 8 Feb 2014]

Takabatake, H., Satoh, H., Mino, T. and Matsuo, T. (2002) PHA


(polyhydroxyalkanoate) production potential of activated sludge treating
wastewater. Water Science and Technology, 45, 119-126.

Takiguchi, N., Kishino, M., Kuroda, A., Kato, J. and Ohtake, H. (2004) A laboratory-
scale test of anaerobic digestion and methane production after phosphorus
recovery from waste activated sludge. Journal of Bioscience and Bioengineering,
97, 365368.

Tauseef, S.M., Abbasi, T. and Abbasi, S.A. (2013) Energy recovery from wastewaters
with high-rate anaerobic digesters. Renewable and Sustainable Energy Reviews, 19,
704-741.

Tchobanoglous, G., Burton, F.L. and Stensel, H.D. (2003) Wastewater Engineering
Treatment and Reuse. Fourth Edition. McGraw-Hill, New York.

Thierbach, R.D. and Hanssen, H. (2002) Utilisation of energy from digester gas and
sludge incineration at Hamburgs Khlbrandhft WWTP. Water Science and
Technology 46, 397-403.
Tiehm, A., Nickel, K., Zellhorn, M. and Neis, U. (2001) Ultrasonic waste activated
sludge disintegration for improving anaerobic stabilization. Water Research, 35,
2003-2009.

Ting, C.H., Lin, K.R., Lee, D.J. and Tay, J.H. (2004) Production of hydrogen and
methane from wastewater sludge using anaerobic fermentation. Water Science
and Technology, 50, 223-228.

Tyagi, V.K. and Lo, S-L. (2013) Sludge: A waste or renewable source for energy and
resources recovery? Renewable and Sustainable Energy Reviews, 25, 708-728.

UNEP/IFIA; United Nations Environment Programme/International FertilizerIndustry


Association (2001) Environmental Aspects of Phosphate and Potash Mining. UNEP,
Paris. http://www.elaw.org/system/files/PotashMining.pdf [accessed: 5 Feb 2014]

Verachtert, H., Ramasamy, K., Meyers, M. and Bever, J. (1982) Investigations on


cellulose degradation in activated sludge plants. Journal of Applied Bacteriology,
52, 185-190.

Verstraete, W., Van de Caveye, P. and Diamantis, V. (2009) Maximum use of


resources present in domestic used water. Bioresource Technology, 100, 5537-
5545.

Vidyarthi, A.S., Tyagi, R.D., Valero, J.R. and Surampalli, R.Y. (2002) Studies on the
production of B. thurningiensis based biopesticides using wastewater sludge as a
raw material. Water Research, 36, 4850-4860.

Walker, M., Iyer, K., Heaven, S. and Banks, C.J. (2011) Ammonia removal in
anaerobic digestion by gas stripping: An evaluation of process alternatives using a
first order rate model based on experimental findings. Chemical Engineering
Journal, 178, 138-145.

Wang, H., Fang, M., Fang, Z. and Bu, H. (2010) Effects of sludge pretreatments and
organic acids on hydrogen production by anaerobic fermentation. Bioresource
Technology, 101, 8731-8735.

Wang, Y-Q, Zhang, J-S., Zhou, J-T. and Zhang, Z-P. (2009) Biodegradation of 4-
aminobenzenesulfonate by a novel Pannonibacter sp. W1 isolated from activated
sludge. Journal of Hazardous Materials, 169, 1163-1167.

Water UK (2009) How the Water Industry is Managing its Contribution to Climate
Change [on-line]. http://www.water.org.uk/home/policy/climate-
change/mitigation [accessed: 25 Feb 2014]

Webber, M.D. (1991) Resource recovery through unconventional uses of sludge. In


Hall, J.E. (ed) Alternative Uses of Sewage Sludge, pp. 343-358. Pergamon Press,
Oxford.
Webber, M.D., Duvoort-van Engers, L.E. and Berglund, S. (1986) Future developments
in sludge disposal strategies. In Davis R.D. (ed) Factors Influencing Sludge Utilisation
Practices in Europe, pp. 103-116. D. Reidel, Dordrecht.

Winter, P., Gonzalez, C., and Smith, S. (2012) The influence of mesophilic
temperatures over 35oC on anaerobic digester performance. In Horan, N.J. (ed)
17th European and Organic Resources Conference, Seminar and Exhibition. 19-21
November, The Royal Armouries, Leeds.

Winter, P., Rus Perez, E. and Smith, S.R. (2013) Optimum anaerobic digestion loading
rates. In Horan, N.J. (ed) 18th European Biosolids & Organic Resources Conference
Proceedings. 18-20 November, Manchester.

Yenign, O. and Demirel, B. (2013) Ammonia inhibition in anaerobic digestion: A


review. Process Biochemistry, 48, 901-911.

Yuvakkumara, R., Elangoa, V., Rajendrana, V. and Kannanb, N. (2014) High-purity


nano silica powder from rice husk using a simple chemical method Journal of
Experimental Nanoscience, 201, 272-281.

Zhang, H.L., Lin, Y.M. and Wang, L. (2009) Biosorption of copper by calcium alginate
from excess activated sludge. Environmental Technology, 30, 1461-1467.

Zou, J., Dai, Y., Pan, K., Jiang, B., Tian, C., Tian, G., Zhou, W., Wang, L., Wang, X. and
Fu, H. (2013) Recovery of silicon from sewage sludge for production of high-purity
nano-SiO2. Chemosphere, 90, 23322339.

Zou, J., Dai, Y., Tian, C.,Pan, K., Jiang, B., Wang, L., Zhou, W., Tian, G.,Wang, X., Xing,
Z. and Fu, H. (2012) Structure and properties of noncrystalline nano-Al(OH)3
reclaimed from carbonized residual wastewater treatment sludge. Environmental
Science and Technology, 46, 45604566.

You might also like