Downstream Processing Technology Review(1)

49
1 BioMara Project Processing Technology Review Simon Murray, R. Elaine Groom, Julie-Anne Hanna & Conall Watson QUESTOR Centre, School of Chemistry and Chemical Engineering, Queen’s University Belfast

description

Chemical Engineering Downstream Processing Technology

Transcript of Downstream Processing Technology Review(1)

1

BioMara Project

Processing Technology Review

Simon Murray, R. Elaine Groom, Julie-Anne Hanna & Conall Watson

QUESTOR Centre, School of Chemistry and Chemical Engineering,

Queen’s University Belfast

2

Table of Contents 1 Introduction ................................................................................................................. 4

2 Cultivation................................................................................................................... 5

2.1 Requirements for algal growth ............................................................................. 5

2.1.1 Light .............................................................................................................. 5

2.1.2 Carbon dioxide .............................................................................................. 5

2.2 Microalgae ............................................................................................................ 6

2.2.1 Species selection ........................................................................................... 6

2.2.2 UV utilisation ................................................................................................ 6

2.2.3 Oxygen poisoning ......................................................................................... 7

2.2.4 Nutrient uptake.............................................................................................. 7

2.3 Growth methods ................................................................................................... 7

2.3.1 Raceway ponds ............................................................................................. 8

2.3.2 Photobioreactors ........................................................................................... 9

2.4 Macroalgae ......................................................................................................... 11

3 Harvesting ................................................................................................................. 13

3.1 Macroalgae ......................................................................................................... 13

3.2 Microalgae .......................................................................................................... 14

3.2.1 Sedimentation ............................................................................................. 15

3.2.2 Flocculation................................................................................................. 15

3.2.3 Flotation ...................................................................................................... 16

3.2.4 Membrane filtration .................................................................................... 16

3.2.5 Centrifugation ............................................................................................. 17

4 Dewatering ................................................................................................................ 17

5 Lipids extraction ....................................................................................................... 18

5.1 Mechanical methods ........................................................................................... 18

5.2 Solvent extraction ............................................................................................... 19

5.2.1 Switchable solvents ..................................................................................... 20

5.3 Supercritical CO2................................................................................................ 20

5.4 Ultrasonic extraction .......................................................................................... 20

3

5.5 Flash depressurisation ........................................................................................ 21

6 Fuel Production ......................................................................................................... 22

6.1 Biodiesel ............................................................................................................. 22

6.2 Bioethanol .......................................................................................................... 24

6.3 Biogas ................................................................................................................. 24

6.4 Biobutanol .......................................................................................................... 25

6.5 Other uses of ‘waste’ products ........................................................................... 26

6.6 Fuel distribution ................................................................................................. 26

7 Conclusion ................................................................................................................ 27

8 Literature review ....................................................................................................... 28

9 References ................................................................................................................. 42

4

1 Introduction As global fossil fuel supplies dwindle and atmospheric carbon dioxide concentrations

rise, there is increasing pressure to find viable biofuel alternatives to petroleum products.

The European Parliament has set a target of 10% of road transport fuel to come from

renewable sources by 2020. Usage in the U.K. is currently estimated at less than 3%,

making this an urgent challenge.

In response to this challenge, amongst a climate of rising fuel prices and concerns over

energy security, there has been an increased focus on the search for alternatives to

petroleum derived products to act as transport fuel. The production of biofuels from

biomass such as micro- and macro-alga has become very attractive to both academic

researchers and energy companies; biodiesel (Ahmad et al., 2011), bioethanol (Harun et

al., 2010), biohydrogen (Park et al., 2009), biogas (Salerno et al., 2009) and biobutanol

(Maron, 2009) are all being investigated. Biodiesel and bioethanol are of special interest

as these fuels are similar to existing petroleum diesel and gasoline and are therefore

compatible with existing infrastructure (Chisti, 2008, Mata et al., 2010).

The production of biofuels and value-added chemicals from algal biomass involves

several common unit operations, regardless of end product (Figure 1).

Figure 1: Unit operations involved in algal fuel production

Whilst this report focusses on the different options available for the downstream

processing of algal biomass into transport fuel, it also included details of the cultivation

step, as the choices during the cultivation stage have a significant impact for the other

processes in the production line.

Previously, review of the production process have identified a number of challenges

which must be overcome in order to make algal biofuels to viable (Brennan & Owende,

2010b):

1. Species selection must balance requirements for biofuel production and extraction

of valuable co-products

2. Attaining higher photosynthetic efficiency

3. Development of single species cultivation, evaporation reduction and CO2

diffusion losses

5

4. Potential for negative energy balance and accounting for requirements in water

pumping, CO2 transfer, harvesting and extraction.

5. Few plants in operation, therefore lack of information for scale-up (desire to be

innovative

6. Use of flue gases due to poisonous nature of NOx and SOx.

2 Cultivation

2.1 Requirements for algal growth

The potential of using ‘waste’ streams in algal cultivation is attractive; for example flue

gas from power plants can be utilised as a carbon dioxide source (Stephenson et al.,

2010), whilst nutrients can be obtained from industrial wastewater (Chinnasamy et al.,

2010), or anaerobic digestate (Wang et al., 2010).

2.1.1 Light

The theoretical maximum efficiency of solar energy conversion (total solar energy to

primary photosynthetic products) is calculated as being approximately 10% (Hulatt &

Thomas, 2011), but metabolic activities within the cell, such as lipid and protein

production, mean this figure is not obtained and measured values are typically in the

range 1 – 3%.

2.1.2 Carbon dioxide

The growth of most photosynthetic algae is autotrophic, carbon is taken up in the

inorganic form i.e. carbon dioxide, carbonate or hydrogencarbonate. Although this can be

sourced from the atmosphere in open growth systems, the driving force for the

replacement of CO2 in the growth media is low and the CO2 concentration can become

the limiting factor in growth (Williams & Laurens, 2010). Therefore, to maintain high

biomass productionCO2 needs to be continually replaced by addition of liquid or gaseous

CO2.

A low cost source is to use flue gases. CO2 makes up between 3% - 15% of flue gases

from fossil-fuel power plants, depending on the fuel source and plant design (Packer,

2009). This is also a form of carbon capture which adds to the attractiveness of this

option.

CO2 requirements can be calculated from growth rates and algal cell compositions

(Stephenson et al., 2010), using an assumed value for efficiency of transfer to the growth

media, and utilisation of, CO2 (Carvalho et al., 2006).

The supply and uptake of CO2 is a key processing parameter in the production of algal

biomass for fuel production. Despite this, the technology for supply of CO2 to algal

6

biomass is as yet poorly developed. This is in stark contrast to the large body of research

on genetic improvement of algal species for specific applications (Carvalho et al., 2006).

2.2 Microalgae

A two-step process to maximise the production of fuel by microalgae is proposed by

Stephenson et al. (2010). Microalgae would first be grown to a high concentration under

nutrient sufficient conditions. The supply of nutrient is then discontinued, causing the

microalgae to accumulate molecules such as triacylglycerides within their cells.

The mechanism by which microalgae stop producing new cells, and instead begin to

accumulate fuel molecules in their cells, is triggered by stress conditions such as nutrient

limitations (Chen et al., 2011).

2.2.1 Species selection

Evidence from the NREL Aquatic Species Programme (ASP) suggests that allowing the

system to self-select the organism is the best approach for cultivation. The key

parameters – temperature, light intensity, pH etc. – can then be optimized for the

prevalent species following small scale modelling work (Sheehan et al., 1998).

Recently, commercial algae-fuel operations have taken one of two approaches: either

choosing to first optimise the production methods (for example an oil-company), or to

first optimise the algal species, through use of genetic modification (Kanellos, 2008).

The ease of downstreaming processing is also highly dependent on the algal species. For

example, Park et al. (2011) found harvesting of algal cells with a settling process most

effective when Pediastrum sp. dominated a mixed algal culture.

2.2.2 UV utilisation

It has been estimated that only about 10% of total solar energy can be utilised by

microalgae (Sheehan et al., 1998). Above this level, light is simply wasted, and may in

fact damage the photosynthetic apparatus (photoinhibition). This presents a major

challenge in microalgae culture – how can high conversion of incident light be achieved?

This is a major consideration in photobioreactors.

Chojnacka & Noworyta (2004) compared growth rates of algae in autotrophic and

mixotrophic modes. The researchers found that, in both modes, no increase in the specific

growth rate was observed above a light intensity of 30 W/m2, with autotrophic growth

being photoinhibited at light intensities above 50 W/m2.

It has been shown experimentally that microalgae cultures exposed to short

(milliseconds) of intense light followed by longer periods of darkness, achieved the same

light conversion as cultures constantly exposed to the same average photon flux. Another

7

approach, using diffusion of light throughout the culture using optical fibres, was also

attempted in Japan, but the high cost of optical fibres is likely to be prohibitive.

2.2.3 Oxygen poisoning

Algal cells take up carbon dioxide and use solar radiation to convert this to oxygen,

which is then expelled from the algal cells. The level of dissolved oxygen present in the

culture can become inhibitory in high concentrations, and become toxic to the algae at

concentrations in excess of 35 mg/l (Carvalho et al., 2006). Chisti (2008) recommended

that oxygen levels are maintained below 400% of air saturation levels to prevent

inhibition occurring.

Although not a problem in open systems (e.g. raceway ponds), where oxygen levels will

be in equilibrium with atmospheric conditions, in closed systems air bubbling zones may

be required to strip excess oxygen from solution (Zebib, 2008).

2.2.4 Nutrient uptake

In a lab scale study, Wang et al. (2010) used diluted dairy slurry digestate as a nutrient

source for cultivation of Chlorella sp. Using a 25 times dilution of the digestate, 100%

ammoniacal nitrogen, 76-83% total nitrogen and 63-75% total phosphorous removal was

achieved over a 21 day period. Up to 38% reductions in COD concentrations were also

observed, with no significant difference in removals obtained at different dilutions.

A study utilising textile effluent for nutrient removal found 99%, 100% and 75%

removals of nitrate-nitrogen, ammoniacal-nitrogen and phosphate-phosphorous

respectively within 24 hours (Chinnasamy et al., 2010). Removal increased to 99.7%,

100% and 98.8% within 72 hours when bubbled with ambient air. Lower phosphate

removal was obtained in cultures bubbled with CO2 enriched (6%) air.

Use of wastewater to provide nutrient is beneficial as, in addition to the obvious savings

in nutrient costs, 100 kg of nitrogen (as N) and 10 kg of phosphorous (as P) can

potentially be removed from wastewater per tonne of algal biomass produced

(Chinnasamy et al., 2010).

2.3 Growth methods

Brennan & Owende (2010a) summarised the advantages and disadvantages of the various

cultivation options for microalgae production (Figure 2).

8

Figure 2: Advantages and limitations of open ponds and photobioreactors (Brennan & Owende,

2010b)

In general, open systems are characterised by low operational and capital costs and low

productivities, whilst closed systems are characterised by high operational and capital

costs and high productivities. Despite their higher productivities and other advantages,

uptake of closed systems for existing commercial applications was not widespread, with

the simple operation of open systems being considered preferred (Carvalho et al., 2006).

2.3.1 Raceway ponds

Raceway ponds are currently used for most commercial microalgae cultivation operations

(Stephenson et al., 2010). Capable of producing algal suspensions with concentrations

up to 600 mg/l (Shelef et al., 1984), the operation is simple, with a paddle wheel

circulating the biomass and CO2 added by sparging the gas into the bottom of a sump

placed into the raceway (Stephenson et al., 2010).

Raceway ponds, however, have several disadvantages (Stephenson et al., 2010):

Potential for high water losses by evaporation;

Potential for contamination with alternative microalgae species.

The requirement for water for cultivation in raceway ponds has been estimated to be as

high as 11-13 million L/ha.year. In a life cycle analysis study, Stephenson et al. (2010)

report evaporative losses ranging from 10x10-3

m3/m

2day for tropical climates to 3x10

-3

m3/m

2day for the climate experienced in the U.K.

The NREL review of the US Department of Energy’s ASP states that invasion by native

species is inevitable, and that therefore the production system should self-select the

organisms. The conditions and genetics of these species should then be optimised for fuel

production (Sheehan et al., 1998).

9

Bennemann and Eisenberg (Sheehan et al., 1998) carried out investigations using a 0.1 ha

raceway pond. Mixing velocities of 15 cm/s were achievable with a power input of 1

kWh/day, but this increased significantly to 10 kWh/day for a mixing velocity of 30 cm/s

(as head loss is related to the square of velocity).

Power requirements decrease with increasing pond depth (Weissman et al., 1988), but

obviously lead to reduced productivity due to reduced incident sunlight per unit volume.

2.3.2 Photobioreactors

Most photobioreactors are one of three types:

(i) vertical airlift/bubble columns;

(ii) horizontal tubular reactors;

(iii) Continuous Stirred-Tank Reactor (CSTR).

The cost of biomass production in photobioreactors (PBRs) may be one order of

magnitude higher than in ponds (Mata et al., 2010), but PBRs offer advantages in terms

of better control of contaminants, temperature and substrate concentrations. Additionally,

photobioreactors offer better volume to land area ratios and reduced harvesting costs

(Carvalho et al., 2006). The main limitations include: overheating, bio-fouling, oxygen

accumulation, difficulty in scale-up, high capital cost and damage of cells by shear stress

(Mata et al., 2010).

In column reactors mixing is provided by the turbulence caused by the rising gas bubbles,

whilst in horizontal reactors a circulation pump is required for mixing purposes. For this

reason, horizontal reactors have higher energy consumption, although this is balanced by

a higher biomass productivity due to the angle toward sunlight facilitating efficient light

harvesting (Carvalho et al., 2006).

Figure 3: Column reactors at Daithi O’Mhurchu

Marine Research Station, Bantry, Cork, Ireland

Figure 4: Horizontal bioreactor at SARDI,

Adelaide, Australia (Rone-Clarke, 2010)

10

Large diameter tubes have a higher specific volume, but the energy required for mixing is

larger and they have limited exposed surface area, meaning algal growth is limited by

availability of UV radiation. The optimum size varies with solar conditions and algae

species, but Zebib (2008), estimates the optimum diameter should not exceed 0.1 m.

The lengths of tubular bioreactors are limited by other factors including the need to

remove excess oxygen and control CO2 concentration with associated pH effects (Ugwu

et al., 2008).

A CSTR consists of a single tank which is continuously agitated by a motor driven

paddle. The mixing rate is such that concentrations of substrate and products are the same

in all areas of the reactor; allowing easier control of CO2 and O2 concentration and pH.

CSTRs are commonly applied in the pharmaceutical industry for biosynthesis reactions

and have been used by U.S. company OriginOil for the production of microalgae for the

production of fuel and other valuable chemicals. The Helix BioReactor system uses an

array of light emitting diodes (LEDs) as part of the agitator (Figure 5), which can be

tuned to specific frequencies for optimisation of photosynthesis (Eckelberry, 2010).

Figure 5: Origin Oil Helix BioReactor system (Eckelberry, 2010)

Recently, the exterior of several buildings has been clad in algal photobioreactors,

including a 15 unit apartment building in Hamburg, Germany (Figure 6). The bioreactors

are provided with nutrients from the greywater produced by the building, and turn to face

the sun to maximise production. The biomass is periodically harvested and converted to

biogas which provides a portion of the energy required to heat the building (Yirka, 2013).

11

Figure 6: The Bio Intelligent Quotient House, Hamburg, Germany (Yirka, 2013)

2.4 Macroalgae

The majority of macroalgae used for commercial operations is cultivated. Various

approaches are used including the use of land-based tanks, coastal lagoons, inshore

waters and more open waters.

Regardless of the location of cultivation, the first stage usually takes place on land, where

seaweed is seeded onto string or nets prior to ongrowing in coastal waters (Figure 7).

After a growing period of 6 – 8 months, the seaweed is ready for harvest for biofuel

production or other uses (Figure 8).

Figure 7: Kelp seeded onto string prior to ongrowing (Schmid et

al., 2003b)

Figure 8: Harvested kelp

(Source: SAMS)

12

Large near shore seaweed cultivation industries are well developed in countries such as

China and Chile, but these facilities are very labour intensive and the economic viability

of such operations relies on the availability of relatively cheap labour. For a seaweed

production operations to succeed in the UK, especially for production of a relatively low

value commodity such as energy, optimisation of all stages of the production and

processing steps is required (Schmid et al., 2003b).

A combination of mechanical failures and issues with nutrient availability rendered early

attempts to cultivate seaweed in offshore locations unsuccessful. Since these failures,

however, there has been significant practical development, and prototypes for kelp

production in the North Sea have been successfully tested (Schmid et al., 2003a).

A recent development in seaweed cultivation is Integrated Multi-Trophic Aquaculture

(IMTA), an aquaculture technique in which the by-products of one process, including

waste, are provided as inputs for another. This typically involves the combination of fed

aquaculture (e.g. finfish, prawns) with organic extractive aquaculture (e.g. shellfish) and

inorganic extractive aquaculture (i.e. seaweed) (Wanner et al., 1994).

In IMTA, finfish are grown in net cages and fed as in standard fish farming. In the same

vicinity shell fish such as mussels are grown supported on ropes as shown in Figure 10,

and seaweeds such as kelp are grown supported on strings similar to that illustrated in

Figure 7 and Figure 8. All three cultivated organisms are valuable commodities, so that

the overall productivity of the operation is greater than if the individual products were

grown in monocultures.

Figure 9: : Conceptual diagram of IMTA (Source:

www.oceansfortomorrow.ca)

Figure 10: Mussels growing on ropes

(Source: http://www.dfo-

mpo.gc.ca/science/publications/article/2009/11-23-

09-eng.html)

IMTA is seen as more sustainable than single species aquaculture operations (Debus &

Wanner, 1992). In fish farming, waste material such as uneaten food and faeces, fall to

the seabed below the cages. Decomposition of this material leads to altered nutrient levels

13

with associated changes to the ecosystem (Lynch, 1978). In IMTA, each element of the

IMTA system is arranged is such a way so that the extractive shellfish and seaweed

occupy the nutrient plume, where they uptake these nutrients and lessen the effect on the

environment (Wanner et al., 1994). In this way, nutrient availability issues which

hindered early commercial seaweed cultivation projects are avoided.

The majority of IMTA deployments to date have been for academic research purposes.

Systems have been operated in various parts of Asia and Europe, South and North

America. The greatest concentration of research has occurred in the Bay of Fundy, New

Brunswick/Nova Scotia, Canada. In this area, industry, academia and government are

collaborating to expand production to commercial scale, with a working industrial scale

IMTA unit expected by 2014 (Modin et al., 2007/6).

In the U.K., the Scottish Marine Institute has carried out a number of IMTA projects. The

REDWEED project (2003-2006) examined bioremediation by seaweed growing adjacent

to sea based fish farms and the MERMAIDS project studied salmon/oysters/seaweed

IMTA systems in the Western Isles (Schnobrich et al., 2007). The projects quantified the

bioremediation potential of IMTA systems and made recommendations of best practice,

but no assessment of the economic aspects of the use of such systems in the U.K. has yet

been published.

3 Harvesting

3.1 Macroalgae

In Northern Ireland, The Environment and Heritage Service’s position is that mechanical

harvesting will not be supported, unless it can be demonstrated that it will not

significantly adversely affect the environment (EHS (N.I.), 2007).

Seaweed that washes up on beaches is regularly removed, especially in tourist areas

where it, along with other deposits and litter, is considered an eyesore. Collection of

beach cast material varies from manual cleaning with rakes to mechanical operations

where the top layer of sand is removed, sieved to remove seaweed and litter, and returned

to the beach (Zemke-White et al., 2005).

Beach cast seaweed begins to decompose within 2-3 days of being washed up on the high

tide line. The decay of seaweed provides nutrients to an area which is likely to otherwise

receive low levels of nutrient influx (McLaughlin et al., 2006). The removal of these

deposits can have a very significant adverse impact on biodiversity (EHS (N.I.), 2007).

However, in areas with high levels of eutrophication, removal of beach cast seaweed can

have positive effects such as increasing water clarity (McLaughlin et al., 2006).

The Blue Flag criteria states that beach cast seaweed should only be removed where it is

considered a nuisance, and that, when removed, it should be disposed in an

“environmentally friendly manner” e.g. composting or use as fertiliser (EHS (N.I.),

2007).

14

Issues surrounding the use of beach cast seaweed concern sand content, which can give

mechanical handling issues, and salt concentration, meaning the seaweed requires

washing in order to prevent interference with fuel production processes.

3.2 Microalgae

The cultivation of microalgae typically results in suspensions of cells with concentrations

of 200 – 600 mg/l. Harvesting refers to the concentration step into slurries or pastes

containing 5000 – 25000 mg/l (Shelef et al., 1984). Due to the cost of concentrating the

relatively dilute solutions produced by the cultivation step, up to 50% of the final product

costs can be due to downstream processing (Greenwell et al., 2010), with 20 – 30%

attributable to the harvesting step alone (Brennan & Owende, 2010b).

As such, the harvesting of algal cultures has been identified by researchers as a major

bottleneck in the production of algal based biofuels (Uduman et al., 2010).

Harvesting can either take place a one-step process, where the chosen technology is

capable of delivering a sufficiently concentration slurry for oil extraction/fuel production;

or by a cascade approach utilising multiple technologies (Figure 11), similar to that

suggested by Bilad et al. (2012) and discussed in Section 3.2.4.

Figure 11: Schematic diagram of microalgal production and processing (Uduman et al., 2010)

The efficient separation, dewatering and drying (if required) of microalgae is probably

the most significant factor in determining the economic feasibility of any microalgae

production system (Shelef et al., 1984).

Algae are difficult to concentrate, as they typically have negatively charged surfaces due

to ionization of functional groups on cell walls. The intensity of these charges is a

function of algal species, ionic strength of the medium, pH and temperature (Shelef et al.,

1984).

15

3.2.1 Sedimentation

Sedimentation (or settling) is a technique by which a feed suspension is separated into a

concentrated slurry and clarified liquid. Used extensively in the wastewater treatment

industry, it is a low-cost separation technology. Suspensions are provided with quiescent

conditions so that the denser solid particles sink to the bottom of the tank and are

removed as an underflow, whilst the clarified liquid layer leaves as a top product (Figure

12).

Figure 12: Settling tank at Antrim WwTW

It is not widely practised with algal cultures, as settling rate is highly influenced by

density difference, and algal cells have a density close to that of water. Without the

addition of flocculants, the reliability of the process is poor (Uduman et al., 2010). With

the addition of flocculants, slurries of up to 1.5% TSS can be obtained; without

flocculation additional, reliability is poor (Uduman et al., 2010).

3.2.2 Flocculation

The addition of chemicals to induce algae flocculation is a common procedure where

sedimentation, floatation, filtration and centrifugation processes are utilised (Shelef et al.,

1984). The chemicals used fall into two broad groups:

a) Inorganic polyvalent metal ions

b) Polymeric organics

Harith et al. (2009) carried out a comparison study of the flocculation efficiency of two

inorganic anionic flocculants (Magnafloc® LT 25 & LT 27) and organic chitosan. Over

90% flocculation efficiency was achieved with each of the flocculants at optimal

conditions, but the authors noted the higher dosage requirement for chitosan (~20 mg/L)

compared to the Magnafloc® (~1 mg/L) to achieve this efficiency.

However, there has been limited investigation on the effect the addition of these

flocculation has to the fuel production, and the ability of the produced fuels to meet the

relevant quality standards. Flocculants are often expensive and used in high

concentrations, which can cause problems during further processing steps and in the

reuse of waste materials.

16

3.2.3 Flotation

Flotation is a unit process by which bubbles of gas are used to drag suspended particles to

the surface of a liquid, from where they can be removed from suspension. Two types of

flotation are suitable for algal harvesting:

Dissolved air flotation (DAF)

Suspended air flotation (SAF)

DAF units use a compressor to supersaturate flotation water with air in a saturator. The

flotation water is then released at atmospheric pressure, causing air to precipitate as small

bubbles. SAF is similar to DAF, but bubbles are created by conventional sparging.

Surfactants are used to reduce surface tension and prevent agglomeration of bubbles and

therefore bubble size. This eliminates the need for a compressor and saturator and

associated costs (Wiley et al., 2009).

The efficiency of flotation processes can be improved via the addition of flocculants. In

jar tests involving alum, cationic, anionic & neutral polyacrylamide polymers and

chitosan, Sim et al. (1988) found alum to be most effective for clarification of water.

Solids concentrations of up to 5% were achieved. Chitosan was also effective at a dose of

50 mg/l, but none of the polyacrylamide polymers was effective at economically low

doses.

3.2.4 Membrane filtration

Membrane filtration can be used to concentrate algal suspensions and remove protozoans

and viruses allowing reuse of residual nutrients. Whilst the majority of studies

investigating membrane filtration in relation to algal cells have explored methods of

preventing fouling in water filtration applications, there have been some reports of use of

membranes for algal cell harvesting.

Using hollow fibre PVC ultrafiltration membranes with a molecular weight cut-off

(MWCO) of 50 kDa, Zhang et al. (2010) concentrated algal suspensions from 1 g/L by a

factor of 155. Using a cross-flow configuration, backwashing and air scouring techniques

were developed that allowed maintenance of consistently high flux levels in continuous

operation.

Whilst the results from the Zhang et al. study are promising, cross-flow membrane

filtration is associated with high energy usage. Using a submerged microfiltration

membrane, Bilad et al. (2012) were able to achieve a 15 times concentration with a

specific energy consumption of just 0.27 kWh/m3

for a culture of Chlorella vulgaris.

Although this is still a relatively low concentration, the authors suggest it would be

economic to use centrifugation to bring the suspension up to the concentration required

for fuel production.

17

3.2.5 Centrifugation

Centrifugation is a technique which utilises centrifugal force to accelerate sedimentation

of suspensions. Suspensions are rotated around a fixed axis, with denser particles forced

to the outside. Centrifuges are used in many applications including blood separation,

separation of cream from milk and thickening of wastewater treatment sludge.

Centrifuges have been used for algal cell harvesting and dewatering in many

investigations. Sim et al. (1988) carried out an investigation using a centrifuge for

harvesting of sewage grown algae. Slurry with dry solids content in excess of 15% was

obtained, with an energy consumption estimated at 1.3 kWh/m3 of pond water.

Despite the widespread use by academic studies, there has been no recent energy analysis

of the use of centrifugation for harvesting of algal biomass.

4 Dewatering The products of harvesting processes oftentimes have significant water content which

must be reduced before further processing is possible. By dehydrating the harvested algal

slurry to a moisture content of 12-15%, the product is stable and storable and suitable for

all forms of lipid extraction processes (Shelef et al., 1984).

In lab scale experiments, lyophilisation or freeze drying is commonly employed (e.g.

Wahlen et al., 2011), but due to the energy involved in maintaining a vacuum, this

commands significant energy consumption and its industrial application is therefore

limited (Huang et al., 2009). Whilst economic for high value pharmaceutical and food

products, it is unlikely to be viable for algal fuel products.

For dewatering of low value products, such as wastewater sludges etc., thermal drying

techniques such as rotary, spray and flash dryer are used (Shelef et al., 1984). Sun drying

is also employed where climate permits, as this can realise an order of magnitude saving

in energy use when compared to thermal techniques (Ciurzynska & Lenart, 2011).

For fuel production processes, unit operations which are capable of oil extraction from

algal slurries, such as Origin Oil’s Single-Step Extraction™ (Eckelberry, 2010) are of

great interest. This process uses an applied electromagnetic field to force algal cells to

release their cells contents which are then easily separated by density difference (light

lipids rise to the top, biomass sinks to bottom (Figure 13)). This, it is claimed, is capable

of processing algal slurries of 1 g/L (dry basis) for as little of $0.20/kg – approximately

one-sixth the cost of conventional techniques.

18

Figure 13: Origin Oil Single-Step Extraction process (Eckelberry, 2010)

5 Lipids extraction Most microalgae strains are relatively small and have a thick cell wall, meaning harsh

conditions are required to rupture the cells for lipid extraction (Wijffels & Barbosa,

2010).

Lipids for fuel production can be extracted from algal biomass using two groups of

methods: mechanical methods where pressure is used to rupture the cells and squeeze the

oil from the cells and chemical methods where solvents are used to extract the lipids. A

combination of both can also be used, which has been shown to most effective in some

cases (Samori et al., 2010).

5.1 Mechanical methods

Mechanical methods involve placing dried algae cells under pressure, forcing the cells

walls to rupture and release the contained lipids.

The most frequently used mechanical method is the screw press, which works in a similar

way to an extruder for polymer processing (Figure 14). Dry algae are conveyed by a

rotating screw along a horizontal annulus whose cross section decreases along the length.

The ruptured cells and oil emerge through a nozzle and are easily separated.

Approximately 75% oil recovery is possible (Farag, 2010).

19

Figure 14: Oil Screw Press (Farag, 2010)

The screw press is commonly used for dewatering operations in other industries such as

paper production, sewage sludge and digestate separations, and for various food

processes including extraction of lipids from olives and seeds.

Although the simplest technology, mechanical methods are very energy intensive, both in

the requirement for dry input material and the energy involved in creating sufficient

pressure for cell rupture.

5.2 Solvent extraction

Several solvents have been used successfully including hexane, ethanol and hexane-

ethanol mixtures, obtaining up to 98% extraction of lipids. Although ethanol is a very

good solvent, it does not selectively extract lipids, extracting also sugars, amino acids,

salts and pigments. This is not desirable for biodiesel production, where the purpose of

extraction is just the lipids and fatty acids (Mata et al., 2010), and the use of ethanol as a

solvent reduces the amount of energy that can be recovered from residues via anaerobic

digestion or other processes.

In a study aimed at optimising the analytical determination of lipid content in lyophilized

microalgae, Ryckebosch et al. (2012) compared the effectiveness of a number of organic

solvents and organic solvent blends. The study found a 1:1 blend of chloroform and

methanol to be most effective for lipid extraction from Chlorella vulgaris, using a

technique involving vortex mixing and separation of solvent and aqueous layers by

centrifugation.

20

5.2.1 Switchable solvents

The potential of switchable solvents for extraction of lipids from algal biomass has also

been investigated (Samori et al., 2010). Switchable solvents are capable of turning from a

non-ionic form to an ionic liquid by bubbling CO2 through the solution and recovered by

bubbling N2. Despite being described by the authors as having potential, the processing

steps used appear very complex, with long extraction and centrifugation steps, and

characterised by hydrocarbon recoveries of less than 20%.

5.3 Supercritical CO2

Supercritical fluids are fluids which are held at or above a critical temperature and

pressure. At these conditions, no distinct liquid and gas phases exist. Supercritical fluids

typically are able to effuse through solids like a gas and dissolve materials like a liquid.

Supercritical fluid extraction is more effective than traditional solvent separation

methods, with recoveries of 100% being possible (Demirbas, 2009).

Capital costs involved in supercritical extraction methods are higher than traditional

solvent extractions (Pellerin, 2003), however operating costs can be considerably lower

as separation of solvent and product is simpler. Solvent can be easily recycled and

residues are suitable for uses such as animal feed. Additionally, the process is relatively

rapid, due to the inherent low viscosity and high diffusivities of supercritical fluids.

Supercritical carbon dioxide is often the choice for supercritical extractions due to its

near ambient critical temperature (304 K) and similar solvation properties to hexane.

5.4 Ultrasonic extraction

Ultrasonic extraction utilises ultrasonic waves (20 kHz - 2 MHz) to create bubbles in a

solvent material. When a bubble collapses near a cell wall, it creates a shock wave which

causes the cell wall to rupture, releasing the cell contents into the solvent.

Cravotto et al. (2008) employed ultrasound and microwaves in order to improve solvent

extraction of oil from biomass sources including cultivated microalgae. Ultrasound was

found to be the most effective for microalgae, improving the extraction yield from 4.8%

to 25.9% whilst reducing the extraction time from 4 hours to 20 minutes.

No details of large cell algal harvesting operations deploying this technology are

currently published, but the technology is being tested in anaerobic digestion

applications. Fitted onto a recycle loop, the Ultrawaves reactor (Figure 15) is designed to

cause anaerobic sludge cells to collapse, allowing their cell contents to become available

for biogas production and boost gas yields (Ultrawaves Reactors Ltd, 2011).

21

Figure 15: Ultrawaves reactor

5.5 Flash depressurisation

Cellruptor, a technology developed by Eco-Solids International for boosting of biogas

yields from wastewater sludge digestion, uses more moderate conditions than

supercritical CO2. Whereas supercritical CO2 methods typically use temperatures and

pressures in the ranges 313 – 333 K (40 – 60ºC) and 20 – 30 MPa (Couto et al., 2010),

the Eco-Solids method uses pressures up to 10 barg (1MPa).

22

Figure 16: Pretreated lignocellulosic biomass

Figure 17: Lab scale Cellruptor

system

Under these pressures, a gas (recycled biogas or exhaust gases) are forced to dissolve in

the aqueous phase (including inside the cells) before it is rapidly depressurized. This

sudden change in osmotic pressure causes the dissolved gases to form bubbles, the force

of which causes cells to rupture (Gill, 2008). This method has been applied to rupture of

wastewater sludge cells, microalgae cells and as a pre-treatment for ligno-cellulosic

biomass for biofuels production (Figure 16).

The energy consumption of this process compares very favourably with other

technologies such as supercriticial CO2 and ultrasonic methods. The energy consumption

of the Cellruptor system is estimated to 3 orders of magnitude lower than that of

ultrasonic methods (Gill, 2008). However, this technology is not yet mature and no data

is available on the effectiveness with algal biomass.

6 Fuel Production

6.1 Biodiesel

The term Biodiesel refers to any fuel made from renewable feedstocks with

characteristics similar to those of petroleum derived diesel (Ahmad et al., 2011).

Biodiesel was first produced in the mid-19th

century, approximately 40 years before

Rudolf Diesel developed the first diesel engine.

Biodiesel is formed by the transesterifcation reaction of fats or oils with alcohols to form

Fatty-Acid Methyl Esters (FAME) (Karmakar et al., 2010). As alcohols are barely

soluble in fats or oils, the reaction proceeds very slowly without the use of catalysts to

increase alcohol solubility, Catalysts can be either homogenous acidic or alkaline,

23

biocatalysts or heterogenous. The fastest reaction rates are achieved with alkali catalysts,

and as such they are the most widely used for commercial applications, despite the

formation of soaps as side products, with associated separation issues (Karmakar et al.,

2010).

The basic process by which biodiesel is produced is the same regardless of feedstock or

catalysis choice. The fats or oils are heated and blended with methanol containing

catalysis. The reaction takes place at the interface between the two phases, with the

reaction reaching completion after 1-2 hours (Lee & Saka, 2010). A simplified process

flow diagram for the production of biodiesel from waste greases and vegetable oils is

shown in Figure 18.

Figure 18: Schematic diagram of biodiesel production

Wahlen et al. (2011) achieved single vessel lipid extraction and transesterification using

methanol acidified with sulphuric acid. At a catalyst concentration of 2.0% (v/v), 32.9 mg

of FAME was produced from 100 mg of lycophilised algal biomass. The reaction went to

completion with 2.5 hours, with the majority of reaction taking place in the first 100

minutes.

A recently developed alternative to the use of catalysts to improve organic solubility is to

employ supercritical conditions, at which distinct phases no longer exist (Madras et al.,

2004). With the reaction taking place in one phase, esterification time is greatly reduced.

Saka & Kusdiana (2001) achieved catalyst free conversion of rapeseed oil and methanol

to biodiesel at conditions of 45-65 MPa and 350℃ with a reaction time of 240s.

Approximately 95% conversion was obtained. Separation was significantly easier as no

soap layer was formed in the absence of alkali catalyst.

24

Although the use of high temperatures favours a short reaction time and high conversion,

it can have a negative effect on the biodiesel properties. Xin et al. (2008) found thermal

decomposition took place when production took place at temperatures greater than 300℃,

which adversely impacted the cold flow properties of the fuel.

Algal oils, unlike other plant oils, contain nitrogen, phosphorus and sulphur which are

problematic with regard to engine performance. However, these impurities are likely to

end up in the soluble fraction following transesterification, and be virtually absent from

algal biodiesel (Williams & Laurens, 2010).

6.2 Bioethanol

Recent work by Lee et al. (2011), employed inverse metabolic engineering of

Saccharomyces cerevisiae to enhance the fermentation of galactose to ethanol, without

any decline in glucose fermentation. Although these studies were carried out in the

laboratory with model sugar solutions, galactose is one the most abundant sugars

contained in red seaweeds and therefore this work is of relevance in increasing the

viability of bioethanol production from marine macroalgae.

Bioethanol has the potential to replace gasoline in today’s cars with little or no

modification to engines (Mata et al., 2010), and has already been established as a fuel for

transportation. In countries such as Brazil, where 15% of total energy is derived from

sugarcane, bioethanol use is widespread (Sivakumar et al., 2010).

Unlike petroleum gasoline, however, ethanol is corrosive and miscible in water. As water

infiltration is evitable in submerged pipeline systems, ethanol cannot be distributed using

the existing pipeline infrastructure. Instead, it must be transported by rail or truck,

considerably increasing the net energy cost (Sivakumar et al., 2010).

Estimates of fuel production efficiency indicate that potentially 2,000 gallons of ethanol

per acre per year are possible using macroalgae (Bio Architecture Lab, 2010).

6.3 Biogas

Algal biomass is suitable for biogas production via anaerobic digestion of both whole

cells and of the residue resulting from lipids extraction (Sialve et al., 2009). Anaerobic

digestion is especially suitable for wastes with high moisture (80 – 90%) content

(Brennan & Owende, 2010b).

Anaerobic digestion of algal biomass residue following lipid extraction can increase the

viability of the process, as additional energy as biogas from algal biomass is recovered

(Sialve et al., 2009).

Where oil content of the algal cells remains below 40%, there is little benefit in

employing through an extraction step to recover oil followed by digestion of residues

over anaerobic digestion of whole algal cells. Therefore, the better strategy is probably

25

the direct AD of harvested biomass (following a suitable concentration step) (Sialve et

al., 2009).

The performance of an anaerobic digester may be compromised by the low C/N ratio of

algal biomass (Brennan & Owende, 2010b). This problem can be resolved by co-

digestion with a high C/N ratio waste product (e.g. waste paper). This also reduces the

potential for inhibition by excess ammonium (Sialve et al., 2009).

The specific methane yield, expressed in litres of CH4 per gram of volatile solid, can be

estimated from the composition of the structure of chemical compounds which have the

general formula CaHbOcNd by use of Equation 1 (Buswells’ formula), where Vm is the

normal molar of methane (Sialve et al., 2009).

mVdcba

dcbaB *

141612

3240 Equation 1

Two factors have been found to have significantly inhibitory effects on the anaerobic

digestion of algal biomass; ammonium and sodium toxicity (Sialve et al., 2009).

Acetoclastic methanogen bacteria are the most sensitive to ammonia, with inhibition

being observed in the range 1.7 – 14 g/L but following acclimatisation of the bacterial

consortium the inhibitory effects were lessened.

Toxicity has been reported with sodium ions above concentrations of 0.5M, although it

has been feasible to produce adapted consortia which did not show any inhibition in high

salinity conditions such as those experienced in marine water.

In general, toxicity effects of both ammonia and sodium had a greater effect in

thermophilic rather than mesophilic conditions (Sialve et al., 2009).

6.4 Biobutanol

Butanol is an alcohol with a four carbon chain backbone. This longer chain makes

butanol more polar compared to two carbon ethanol, resulting in properties which are

more similar to those of petroleum gasoline. Butanol can be used as vehicle fuel in an

undiluted form meaning, unlike ethanol, it has the potential to completely replace

petroleum gasoline.

Butanol is immiscible in water, leading to a lessening of the corrosion and water

absorption problem anticipated with ethanol in fuel distribution systems. However, there

are technical obstacles that need to be overcome (mainly related to biobutanol’s relatively

high viscosity) before it is acknowledged as a suitable fuel by engine manufacturers.

Biobutanol is produced commericially by the anaerobic conversion of carbohydrate into

acetone, butanol and ethanol by Clostridium Acetobutylicum. This process is known as

26

Acetone-Butanol-Ethanol fermentation (ABE fermentation), with a product ratio of 3:6:1

(Nigam & Singh, 2011).

Despite the interest of large industrial players, there are limited publications on the

production of biobutanol from algal biomass. DuPont, citing commercial competitive

reasons, have not disclosed how much butanol they expect to produce from microalgae

(Maron, 2009), but have entered into R&D projects with several energy providers.

6.5 Other uses of ‘waste’ products

Unless major advances can be made in the productivity of algal biomass and efficiency of

processing techniques, it is likely that production of transport fuel from algal biomass

will only become economically viable following increases in the price of crude oil with

associated increases in gasoline-derived transport fuels.

The traditional transesterification reaction by which biodiesel is produced from

triacylgerides leads to the production of glycerol as a waste product. Although glycerol is

used in soap manufacture, there is a large quantity already on the market due to increased

production of biodiesel from terrestrial crops (Packer, 2009). As such, alternative uses of

glycerol need to be developed to match the increased production of biodiesel.

Work by Siles et al. (2010) used glycerol containing waste and wastewater from biodiesel

production from used-cooking oil for biogas production. Using a mixture containing 15%

glycerol and 85% wastewater, close to 100% biodegradation was obtained, giving a

methane yield of 310 ml CH4/gCOD removed at 1 atm and 25°C.

This use of glycerol is synergistic with the approach taken by Sialve et al. (2009), which

used anaerobic digestion to recover maximum energy from algal biomass and identified

the need to co-digest with a nitrogen poor substrate.

6.6 Fuel distribution

The most economically viable option for transportation of biofuels involves the use of the

existing distribution and storage infrastructure. There may be potential issues in this

regard dealing with pipeline ownership (Sivakumar et al., 2010).

Where a pipeline and/or storage tank is located underground, it is inevitable that

infiltration of water will occur. With traditional gasoline, this is removed through a series

of ‘traps’ where water and gasoline are allowed to separate due to differences in density.

Unlike traditional gasoline, ethanol is hydroscopic and potentially corrosive meaning that

it cannot be transported in existing pipelines (Sivakumar et al., 2010). Alternative

transport methods, such as rail or truck, significantly add to the net energy cost.

27

7 Conclusion Commercially viable algal biofuels are not yet a reality. Research has delivered a

significant body of work since the oil crises of the 1970s in response to the threat of

dwindling reserves and high crude oil prices, demonstrating the potential of micro and

macro algae resources. The majority of this work has tended to focus on single operations

along the supply chain, often failing to address how these technologies will function in

real life.

A more holistic view of the production processes involved is required for the

development of commercially viable algal biofuels. This is necessary to ensure use is

made of existing infrastructure and promote adoption of new fuels by the public.

28

8 Literature review

Author Papers main focus Growth medium Algal

population

Method of determining pop.

growth

Growth Technologies

used Extraction method

Conversion method

Study conclusions

(Chinnasamy et al., 2010)

Industrial effluent as a medium.

Treated carpet mill industrial effluent.

Individual algae/

Consortium of isolates from wastewater.

Chlorophyll a content.

Flasks & Raceway ponds

N/A Lipid extraction and biodiesel conversion.

Consortium performed best treated waters

(expected 29.3 tons of biomass & 4060L/

ha/year)

(Wang et al., 2010)

Cow manure as a medium.

Undigested and digested dairy

manure. Chlorella sp.

Optical density of the inoculated

manure as algal density indicator.

250ml Erlenmeyer

Flasks

Centrifugation and dried for lipid

content.

Transesterification to biodiesel

N/A

(Wahlen et al., 2011)

Extraction, In situ (one step)

tranesterification, biodiesel

Artificial sea water media for marine strains/ Modified Bristol's media for

fresh water strains/ BG-11

7 separate strains of algae

N/A

500ml Baffled Flasks/ 5L Cell-

Stir Flasks/ 220L Raceway

Centrifugation/ In situ (one step)

tranesterification of FAME's and

biodiesel

In situ (one step) tranesterification

of FAME's and biodiesel using methanol and

chloroform

Eliminating the need for drying of biomass

reducing energy requirements.

(Patil et al., 2011)

Single-step supercritical process

for extraction and transesterification,

biodiesel

N/A

Algal paste Nannochloropsis sp. Obtained from outside

source.

N/A Outdoor open

raceways

Single-step supercritical process

for extraction and transesterification

for biodiesel.

Single-step supercritical process for

extraction and transesterification

for biodiesel.

Eliminating the need for drying of biomass

reducing energy requirements.

29

(Lardon et al., 2009)

LCA and review on biodiesel.

N/A N/A N/A N/A N/A N/A N/A

(Stephenson et al., 2011)

Theoretical Review on Algal Cultivation, exploting possibility

of improving photosynthesis

through genomics.

N/A N/A N/A N/A N/A N/A N/A

(Kumar et al., 2010)

Determine best total ammonia

nitrogen level for optimal culture

growth.

Anaerobically digested piggery

effluent/ Pure mineral media.

Chlorella vulgaris

Using haemocytometer,

Algal Density (cells/ml)

12L Plastic bags N/A N/A

Highest algal growth rate at 20 mg/L TAN.

Piggery efflluent good feed for algal growth. Supplementation of

specific nutrients may increase GR.

(Chojnacka & Noworyta, 2004)

Algae cultivation, effects of light and

glucose on autotrophic,

heterotrophic and mixotrophic algae.

Zarrouk liquid medium

Spirulina sp. Spectrophotomete

r

2 L Photobioreactor/ working vol. 1

L

N/A N/A

Autotrophic culture optimum light

intenisty 30-50 W m -2, inhibition above

50Wm-2 Mixotrophic culture light >30 W m-2 and

>0.5g/L glucose Heterotrophic culture

>0.5g/L glucose

30

(Valderrama et al., 2003)

Chemical (Astaxantin,

Phycocyanin) extraction using supercritical CO2

N/A

Dried H. Pluviali, & Spirulina

maximasourced externally

N/A Artificial ponds.

Using an extraction unit with natural substances, pure supercritical CO2/

ethanol.

N/A

Solvent/feed ratio and the use of

cosolvents have great effects on the total substance extacted.

(Cravotto et al., 2008)

Simultaneous microwave and

ultrasound assisted extraction of methyl

esters.

N/A Milled seaweed

externally sourced

N/A N/A Microwave and

ultrasound assisted extraction

N/A

Extraction times increases 10 fold and yields increased 50-500% in comparison

to conventional methods of extraction.

(Couto et al., 2010)

Evaluate microalgae growth and

docosahexaenoic acid production. To

compare the supercritical CO2

method with convential lipid

extraction.

f2+NPM medium supplemented with glucose.

C.cohnii N/A Cylinder-conical 100L fermenter.

Centrifuged, freeze dried. SC-CO2 method used

Transesterification and extraction with hexane to obtain methyl

esters.

N/A

(Catchpole et al., 2009)

Extraction and fractionation of

lipids using near-critical fluids.

REVIEW

N/A N/A N/A N/A N/A N/A N/A

31

(Harun et al., 2010)

Using microalgal biomass as a

substrate for yeast fermentation and

bioethanol production

Yeast (Luria broth and Terrific broth)) Chlorococum

sp.

Yeast Growth measured using optical density.

100L bag photobioreactor

s outdoors

Lipid extraction using supercritical

CO2.

Ethanol production from fermentation

of biomass in 500mL Erlenmeyer

flasks.

1. Highest yeast conc. produced highest conc. of ethanol.

2. Algal cells produce more ethanol in

fermentation after the lipid extraction process as cell walls are ruptured making more simple sugars available for ethanol

production. 3. 30°C best

temperature for ethanol production.

4. Algal concentration and medium choice

have great effects on ethanol production.

(Lee et al., 2011)

Improving the efficiency of

galactose fermentation in bioethanol using

genetically modified yeast cells.

Galactose is an abundant sugar in

marine plant biomass such as red

seaweed.

Yeast (YP medium, YSC medium)

N/A Yeast Growth

measured using optical density.

N/A N/A N/A

1. The overexpression of a number of genes

greatly increased both galactose

consumption rate and ethanol productivity. 2. Certain genes may partially alleviate the

effects of glucose repression.

32

(Karmakar et al., 2010)

Review focusing mainly on different biodiesel feedstock

properties, and briefly on the

manufacturing process.

N/A N/A N/A N/A N/A N/A N/A

(Chisti, 2008)

Review of the prospects of algae derived biodiesels.

Production of microalgal biomass.

Comparison with bioethanol and

biodiesel economics.

N/A N/A N/A N/A N/A N/A N/A

(Demirbas, 2009)(Madras et

al., 2004)

Extraction, fractionation and

transesterifcation of two species of algae

and microalgae.

N/A

Cladophora fracta, Chlorella protothecoides obtained from

external supplier.

N/A N/A

Milling and the use of hexane in a

Soxhlet extractor for 18 hrs.

Transesterification using supercritical

methanol

Lipid fraction and heating value in

Chlorella protothecoidesis is much higher than

inCladophora fracta.

33

{{597 Madras,Giridhar

2004}}

Transesterification of vegetable oil

using supercritical methanol and

ethanol. Effects of enzyme loading, oil:alcohol ratio,

reaction time and temp.

Transesterification of vegetable oil

using supercritical CO2 and an enzyme.

N/A N/A N/A N/A N/A

Transesterification using supercritical methanol, ethanol

and CO2 (with lipase).

High rates of conversion with both

supercritical methanol and ethanol

at 80-100% conversion. Lower

conversion with supercritical CO2 and enzyme at 27-30%.

(Hansen et al., 2005)

REVIEW The blending of ethanol

and diesel and properties of

different blends. Includes, engine

performance, durability, emissions

information

N/A N/A N/A N/A N/A N/A N/A

34

(Lin et al., 2011)

REVIEW Development of

biodiesel, different feedstocks and

conversion technologies.

Environmental and social impacts of biodiesel, food security, land

change, water use.

N/A N/A N/A N/A N/A N/A N/A

(Lee & Saka, 2010)

REVIEW biodiesel The properties of

homogeneous and heterogeneous

catalysts. Effects of temp. Pressure,

molar ratio, water, free fatty acids and a number of new

supercritical techniques.

N/A N/A N/A N/A N/A N/A N/A

(Gheshlaghi et al., 2009)

REVIEW Metabolic pathways of butanol producing bacteria clostridia. Enzymes

involved.

N/A N/A N/A N/A N/A N/A N/A

35

(Harvey & Meylemans,

2011)

REVIEW Properties of biobutanol as a

alternative fuel source and in blends

with diesel. Commercialization

of biobutanol.

N/A N/A N/A N/A N/A N/A N/A

(Sierra et al., 2008)

Characterization of a flat panel

photobioreactor. Mixing , heat transfer, solar

radiation.

N/A N/A N/A N/A N/A N/A

Major aspects determing

productivity is the location, deployment and orientation of the

reactor. Flat bioreactors require less power supply than their tubular

counterparts.

(Tan et al.)

REVIEW Current level of

development of China's biomass

resource utilization. Technological level.

Resource distribution.

N/A N/A N/A N/A N/A N/A

Huge potential growth area within

China. Large resource base but relatively low exploitation.

36

(Stripp & Happe, 2009)

REVIEW into the history of the

discovery of a class of bacterial proteins

which can evolve hydrogen gas. Bound to the

photosynthetic electron transport

chain through ferredoxin.

N/A N/A N/A N/A N/A N/A N/A

(Yan et al., 2010)

Coupling of biohydrogen and

polyhydroxyalkanoates production

through anaerobic digestion of blue

algae. Examination of blue algae

pretreatments with bases, heat and

microwaves.

Mixture of macro and micro-

minerals with added sugars,

Blue Algae obtained from a euthrophic lake

in China and Bascillus cereus isolated from active sludge.

N/A 5 L anaerobic digester, 3 L fermenter.

N/A Anaerobic digestion.

Basification of blue algae before digestion

showed higher efficiencies than

other pretreatments.

(Biagini et al., 2010)

Study into different 4 hydrogen

generation systems. N/A N/A N/A N/A N/A N/A

Hydrogen production maximized for the

gasification/separation process.

37

(Greenwell et al., 2010)

REVIEW Very extensive focusing

on microalgal cultivation, growth

technologies, harvesting,

conversion of lipids to fuels.

N/A N/A N/A N/A N/A N/A N/A

(Sivakumar et al., 2010)

REVIEW Looks at a number of different

feedstocks for bioethanol and

biodiesel production. Starch

based, lignocellulose

based, plant based, microbial based and

algal based.

N/A N/A N/A N/A N/A N/A N/A

(Festel, 2008)

REVIEW Economic aspects of biofuel

production including first and second generation

biofuels.

N/A N/A N/A N/A N/A N/A N/A

38

(Larkum, 2010)

REVIEW Compares the potential of solar energy in

water organisms vs. land plants. Solar

footprint and carbon footprint.

N/A N/A N/A N/A N/A N/A N/A

(Amin, 2009)

REVIEW Describes a number of different

microalgal lipid conversion methods

including gasification, liquefaction,

pyrolysis, hydrogenation,

fermentation and transesterification.

N/A N/A N/A N/A N/A N/A N/A

(FitzPatrick et al., 2010)

REVIEW focusing on lignocellulosic

feedstocks for the production of

valuable chemical compounds and

biofuels. Biorefinery.

Fractionation.

N/A N/A N/A N/A N/A N/A

Good prospects to make better use of

cheap and abundant forestry waste.

39

(da et al., 2009)

REVIEW Pretreatment of lignocellulosic

wastes. Physical, solvent, chemical (acidification and

basification), heat, oxidative

pretreatments.

N/A N/A N/A N/A N/A N/A

Pretreatments play a central role and

inefficient pretreatment

strategies are an important cost

deterent for the use of lignocellosic

wastes.

(Mata et al., 2010)

REVIEW Comprehensive

review of the biodiesel production process. Selection of

microalgal strains based on lipid

content, harvesting, extraction,

conversion. Growth technologies. Discussion of

numerous secondary

applications for microalgal

production.

N/A N/A N/A N/A N/A N/A

Considerable investment and development of expertise is still

needed before the utilization of

microalgae for biodiesel production

is economically viable.

40

(Collet et al., 2011)

Life cycle analysis of microalgae culture

couple to biogas production.

Compares with algal biodiesel processes.

N/A Chlorella vulgaris

N/A Raceway pond N/A

Anaerobic digestion and

purified by bubbling through

pressurized water.

From both an environmental and an

economic stand-point, a combination

of both lipid extraction, biodiesel

production and anaerobic digestion may be preferable.

(Siles Lopez et al., 2009)

Anaerobic digestion of glycerol by-product of the

transesterification of triacylglycerol

during the production of

biodiesel.

N/A N/A N/A N/A N/A Anaerobic digestion

Glycerol-containing waste produces

substantial quantities of methane.

(Fernandez et al., 2010)

Kinetics of mesophilic

anaerobic digestion using different

initial amounts of municipal solid

waste.

N/A N/A N/A N/A N/A Anaerobic digestion

Higher active biomass and higher coefficient for the production of

methan at lower loadings.

41

(Doukova et al., 2010)

Anaerobic digestion for methane production,

ammonium fertilizer production,

microalgal by-products harvesting

within closed system.

Mixture of macro and micro-minerals.

Chlorella Vulgaris

Optical density. Photobioreactor N/A N/A N/A

(Eckelberry, 2010)

Short booklet showing a number of algal harvesting,

dewatering and extraction

technologies.

N/A N/A N/A N/A

Polymer flocculation,

Decanters/Centrifuges, Hydroclones,

Indirect/Direct heat, Fluid bed,

microwave, expellers/presses,

solvents, supercritical CO2,

enzymes, ultrasonification, live

extraction.

N/A N/A

42

9 References

Ahmad, A.L., Yasin, N.H.M., Derek, C.J.C. & Lim, J.K. (2011), "Microalgae as a

sustainable energy source for biodiesel production: A review", Renewable and

Sustainable Energy Reviews, 15, 1, p. 584-593.

Amin, S. (2009), "Review on biofuel oil and gas production processes from microalgae",

Energy Conversion and Management, 50, 7, p. 1834-40.

Biagini, E., Masoni, L. & Tognotti, L. (2010), "Comparative study of thermochemical

processes for hydrogen production from biomass fuels", Bioresource technology,

101, 16, p. 6381-6388.

Bilad, M.R., Vandamme, D., Foubert, I., Muylaert, K. & Vankelecom, I.F.J. (2012),

"Harvesting microalgal biomass using submerged microfiltration membranes",

Bioresource technology, 111, 0, p. 343-352.

Bio Architecture Lab 2010, , Berkeley California Synthetic Biotech and Biofuels.

Available online: http://www.ba-lab.com/ [2010, 11/22] .

Brennan, L. & Owende, P. (2010a), "Biofuels from microalgae:a review of technologies

for production, processing, and extractions of biofuels and co-products", Renewable

and Sustainable Energy Reviews, 14, 2, p. 557-77.

Brennan, L. & Owende, P. (2010b), "Biofuels from microalgae—A review of

technologies for production, processing, and extractions of biofuels and co-

products", Renewable and Sustainable Energy Reviews, 14, 2, p. 557-577.

Carvalho, A.P., Meireles, L.A. & Malcata, F.X. (2006), "Microalgal reactors: A review of

enclosed system designs and performances", Biotechnology progress, 22, 6, p. 1490-

1506.

Catchpole, O.J., Tallon, S.J., Eltringham, W.E., Grey, J.B., Fenton, K.A., Vagi, E.M.,

Vyssotski, M.V., MacKenzie, A.N., Ryan, J. & Zhu, Y. (2009), "The extraction and

fractionation of specialty lipids using near critical fluids", The Journal of

Supercritical Fluids, 47, 3, p. 591-597.

Chen, M., Tang, H., Ma, H., Holland, T.C., Ng, K.Y.S. & Salley, S.O. (2011), "Effect of

nutrients on growth and lipid accumulation in the green algae Dunaliella tertiolecta",

Bioresource technology, 102, 2, p. 1649-1655.

Chinnasamy, S., Bhatnagar, A., Hunt, R.W. & Das, K.C. (2010), "Microalgae cultivation

in a wastewater dominated by carpet mill effluents for biofuel applications",

Bioresource technology, 101, 9, p. 3097-3105.

43

Chisti, Y. (2008), "Biodiesel from microalgae beats bioethanol", Trends in

biotechnology, 26, 3, p. 126-131.

Chojnacka, K. & Noworyta, A. (2004), "Evaluation of Spirulina sp. growth in

photoautotrophic, heterotrophic and mixotrophic cultures", Enzyme and microbial

technology, 34, 5, p. 461-465.

Ciurzynska, A. & Lenart, A. (2011), "Freeze-Drying - Application in Food Processing

and Biotechnology - A Review", Polish Journal of Food and Nutrition Sciemces, 61,

3, p. 165-171.

Collet, P., Hélias, A., Lardon, L., Ras, M., Goy, R. & Steyer, J. (2011), "Life-cycle

assessment of microalgae culture coupled to biogas production", Bioresource

technology, 102, 1, p. 207-214.

Couto, R.M., Simoes, P.C., Reis, A., Da Silva, T.L., Martins, V.H. & Sanchez-Vicente,

Y. (2010), "Supercritical fluid extraction of lipids from the heterotrophic microalga

Crypthecodinium cohnii", Engineering in Life Sciences, 10, 2, p. 158-164.

Cravotto, G., Boffa, L., Mantegna, S., Perego, P., Avogadro, M. & Cintas, P. (2008),

"Improved extraction of vegetable oils under high-intensity ultrasound and/or

microwaves", Ultrasonics sonochemistry, 15, 5, p. 898-902.

da, C.S., Chundawat, S.P.S., Balan, V. & Dale, B.E. (2009), "'Cradle-to-grave'

assessment of existing lignocellulose pretreatment technologies", Current opinion in

biotechnology, 20, 3, p. 339-347.

Debus, O. & Wanner, O. (1992), "Degradation of xylene by a biofilm growing on a gas-

permeable membrane", Water Science and Technology, 26, 3, p. 607-616.

Demirbas, A. (2009), "Production of biodiesel from algae oils", Energy Sources, Part A:

Recovery, Utilization and Environmental Effects, 31, 2, p. 163-168.

Doukova, I., Katanek, F., Maleterova, Y., Katanek, P., Doucha, J. & Zachleder, V.

(2010), "Utilization of distillery stillage for energy generation and concurrent

production of valuable microalgal biomass in the sequence: Biogas-cogeneration-

microalgae-products", Energy Conversion and Management, 51, 3, p. 606-611.

Eckelberry, R. (2010), "Algae Harvesting, Dewatering and Extraction", World Biofuels

Markets, .

EHS (N.I.) (2007), "Environmentally Sustainable Seaweed Harvesting in Northern

Ireland", .

Farag, I.H. (2010), "Microalgae Production of Biodiesel", .

44

Fernandez, J., Perez, M. & Romero, L.I. (2010), "Kinetics of mesophilic anaerobic

digestion of the organic fraction of municipal solid waste: Influence of initial total

solid concentration", Bioresource technology, 101, 16, p. 6322-6328.

Festel, G.W. (2008), "Biofuels - Economic Aspects", Chemical Engineering &

Technology, 31, 5, p. 715-720.

FitzPatrick, M., Champagne, P., Cunningham, M.F. & Whitney, R.A. (2010), "A

biorefinery processing perspective: Treatment of lignocellulosic materials for the

production of value-added products", Bioresource technology, 101, 23, p. 8915-

8922.

Gheshlaghi, R., Scharer, J.M., Moo-Young, M. & Chou, C.P. (2009), "Metabolic

pathways of clostridia for producing butanol", Biotechnology Advances, 27, 6, p.

764-781.

Gill, N. (2008), "Cellruptor a highly efficient biomass processing & biofuels processing

platform technology", .

Greenwell, H.C., Laurens, L.M.L., Shields, R.J., Lovitt, R.W. & Flynn, K.J. (2010),

"Placing microalgae on the biofuels priority list: A review of the technological

challenges", Journal of the Royal Society Interface, 7, 46, p. 703-726.

Hansen, A.C., Zhang, Q. & Lyne, P.W.L. (2005), "Ethanol–diesel fuel blends –– a

review", Bioresource technology, 96, 3, p. 277-285.

Harith, Z.T., Yusoff, F.M., Mohamed, M.S., Din, M.S.M. & Ariff, A.B. (2009), "Effect

of different flocculants on the flocculation performance of microalgae, Chaetoceros

calcitrans, cells", African Journal of Biotechnology, 8, 21, p. 5971-5978.

Harun, R., Danquah, M.K. & Forde, G.M. (2010), "Microalgal biomass as a fermentation

feedstock for bioethanol production", Journal of Chemical Technology and

Biotechnology, 85, 2, p. 199-203.

Harvey, B.G. & Meylemans, H.A. (2011), "The role of butanol in the development of

sustainable fuel technologies", Journal of Chemical Technology and Biotechnology,

86, 1, p. 2-9.

Huang, L., Zhang, M., Mujumdar, A.S., Sun, D., Tan, G. & Tang, S. (2009), "Studies on

decreasing energy consumption for a freeze-drying process of apple slices", Drying

Technology, 27, 9, p. 938-946.

Hulatt, C.J. & Thomas, D.N. (2011), "Energy efficiency of an outdoor microalgal

photobioreactor sited at mid-temperate latitude", Bioresource technology, 102, 12, p.

6687-6695.

45

Kanellos, M. 2008, November 21, 2008-last update, Can Solix Cut the Cost of Making

Algae by 90%? [Homepage of GreenTech Media], [Online]. Available online:

http://www.greentechmedia.com/articles/read/can-solix-cut-the-cost-of-making-

algae-by-90-5247/ [2011, February 17th] .

Karmakar, A., Karmakar, S. & Mukherjee, S. (2010), "Properties of various plants and

animals feedstocks for biodiesel production", Bioresource technology, 101, 19, p.

7201-7210.

Kumar, M.S., Miao, Z.H. & Wyatt, S.K. (2010), "Influence of nutrient loads, feeding

frequency and inoculum source on growth of Chlorella vulgaris in digested piggery

effluent culture medium", Bioresource technology, 101, 15, p. 6012-6018.

Lardon, L., Helias, A., Sialve, B., Steyer, J. & Bernard, O. (2009), "Life-cycle assessment

of biodiesel production from microalgae", Environmental Science and Technology,

43, 17, p. 6475-6481.

Larkum, A.W.D. (2010), "Limitations and prospects of natural photosynthesis for

bioenergy production", Current opinion in biotechnology, 21, 3, p. 271-276.

Lee, J. & Saka, S. (2010), "Biodiesel production by heterogeneous catalysts and

supercritical technologies", Bioresource technology, 101, 19, p. 7191-7200.

Lee, K., Hong, M., Jung, S., Ha, S., Yu, B.J., Koo, H.M., Park, S.M., Seo, J., Kweon, D.,

Park, J.C. & Jin, Y. (2011), "Improved galactose fermentation of Saccharomyces

cerevisiae through inverse metabolic engineering", Biotechnology and

bioengineering, 108, 3, p. 621-631.

Lin, L., Cunshan, Z., Vittayapadung, S., Xiangqian, S. & Mingdong, D. (2011),

"Opportunities and challenges for biodiesel fuel", Applied Energy, 88, 4, p. 1020-

1031.

Lynch, W. 1978, Handbook of Silicone Rubber Fabrication, Van Nostrand Reinhold

Company, New York.

Madras, G., Kolluru, C. & Kumar, R. (2004), "Synthesis of biodiesel in supercritical

fluids", Fuel, 83, 14-15, p. 2029-2033.

Maron, D.F. (2009), DuPont's 'Unique' Seaweed Ventre Nets DOE Cash.

Mata, T.M., Martins, A.A. & Caetano, N.S. (2010), "Microalgae for biodiesel production

and other applications: A review", Renewable and Sustainable Energy Reviews, 14,

1, p. 217-232.

46

McLaughlin, E., Kelly, J., Birkett, D., Maggs, C. & Dring, M. (2006), "Assessment of the

Effects of Commercial Seaweed Harvesting on Intertidal and Subtidal Ecology in

Northern Ireland", .

Modin, O., Fukushi, K. & Yamamoto, K. (2007/6), "Denitrification with methane as

external carbon source", Water Research, 41, 12, p. 2726-2738.

Nigam, P.S. & Singh, A. (2011), "Production of liquid biofuels from renewable

resources", Progress in Energy and Combustion Science, 37, 1, p. 52-68.

Packer, M. (2009), "Algal capture of carbon dioxide; biomass generation as a tool for

greenhouse gas mitigation with reference to New Zealand energy strategy and

policy", Energy Policy, 37, 9, p. 3428-3437.

Park, J.B.K., Craggs, R.J. & Shilton, A.N. (2011), "Recycling algae to improve species

control and harvest efficiency from a high rate algal pond", Water research, 45, 20,

p. 6637-6649.

Park, J., Lee, J., Sim, S.J. & Lee, J. (2009), "Production of hydrogen from marine macro-

algae biomass using anaerobic sewage sludge microflora", Biotechnology and

Bioprocess Engineering, 14, 3, p. 307-315.

Patil, P.D., Gude, V.G., Mannarswamy, A., Deng, S., Cooke, P., Munson-McGee, S.,

Rhodes, I., Lammers, P. & Nirmalakhandan, N. (2011), "Optimization of direct

conversion of wet algae to biodiesel under supercritical methanol conditions",

Bioresource technology, 102, 1, p. 118-122.

Pellerin, P. (2003), "Comparing Extraction by Traditional Solvants with Supercritical

Extraction from an Economic and Environmental Standpoint", , p. 111.

Rone-Clarke, B. 2010, , NCRIs Algae and Biofuels Facility, South Australian Research

and Development Institute (SARDI) [Homepage of Ausbiotech], [Online]. Available

online: http://www.ncrisbiofuels.org/facilities [2013, 16/01/2013] .

Ryckebosch, E., Muylaert, K. & Foubert, I. (2012), "Optimization of an analytical

procedure for extraction of lipids from microalgae", JAOCS, Journal of the

American Oil Chemists' Society, 89, 2, p. 189-198.

Saka, S. & Kusdiana, D. (2001), "Biodiesel fuel from rapeseed oil as prepared in

supercritical methanol", Fuel, 80, 2, p. 225-231.

Salerno, M., Nurdogan, Y. & Lundquist, T.J. (2009), "Biogas production from algae

biomass harvested at wastewater treatment ponds", , p. 204-208.

47

Samori, C., Torri, C., Samori, G., Fabbri, D., Galletti, P., Guerrini, F., Pistocchi, R. &

Tagliavini, E. (2010), "Extraction of hydrocarbons from microalga Botryococcus

braunii with switchable solvents", Bioresource technology, 101, 9, p. 3274-3279.

Schmid, T., Helmbrecht, C., Panne, U., Haisch, C. & Niessner, R. (2003a), "Process

analysis of biofilms by photoacoustic spectroscopy", Analytical and Bioanalytical

Chemistry, 375, 8, p. 1124-1129.

Schmid, T., Panne, U., Haisch, C. & Niessner, R. (2003b), "Photoacoustic absorption

spectra of biofilms", Review of Scientific Instruments, 74, 1, p. 754-6.

Schnobrich, M.R., Chaplin, B.P., Semmens, M.J. & Novak, P.J. (2007), "Stimulating

hydrogenotrophic denitrification in simulated groundwater containing high dissolved

oxygen and nitrate concentrations", Water research, 41, 9, p. 1869-1876.

Sheehan, J., Dunahay, T., Benemann, J. & Roessler, P. (1998), "A Look Back at the U.S.

Department of Energy's Aquatic Species Program - Biodiesel from Algae", .

Shelef, G., Sukenik, A. & Green, M. (1984), "Microalgae Harvesting and Processing: A

Literature Review", .

Sialve, B., Bernet, N. & Bernard, O. (2009), "Anaerobic digestion of microalgae as a

necessary step to make microalgal biodiesel sustainable", Biotechnology Advances,

27, 4, p. 409-416.

Sierra, E., Acien, F.G., Fernandez, J.M., Garcia, J.L., Gonzalez, C. & Molina, E. (2008),

"Characterization of a flat plate photobioreactor for the production of microalgae",

Chemical Engineering Journal, 138, 1-3, p. 136-147.

Siles Lopez, J.A., Martin Santos, Maria de,los Angeles, Chica Perez, A.F. & Martin

Martin, A. (2009), "Anaerobic digestion of glycerol derived from biodiesel

manufacturing", Bioresource technology, 100, 23, p. 5609-5615.

Siles, J.A., Martín, M.A., Chica, A.F. & Martín, A. (2010), "Anaerobic co-digestion of

glycerol and wastewater derived from biodiesel manufacturing", Bioresource

technology, 101, 16, p. 6315-6321.

Sim, T.,-S., Goh, A. & Becker, E.W. (1988), "Comparison of Centrifugation, Dissolved

Air Flotation and Drum Filtration Techniques for Harvesting Sewage-grown Algae",

Biomass, 16, p. 51-62.

Sivakumar, G., Vail, D.R., Xu, J., Burner, D.M., Lay Jr., J.O., Ge, X. & Weathers, P.J.

(2010), "Bioethanol and biodiesel: Alternative liquid fuels for future generations",

Engineering in Life Sciences, 10, 1, p. 8-18.

48

Stephenson, A.L., Kazamia, E., Dennis, J.S., Howe, C.J., Scott, S.A. & Smith, A.G.

(2010), "Life-Cycle Assessment of Potential Algal Biodiesel Production in the

United Kingdom: A Comparison of Raceways and Air-Lift Tubular Bioreactors",

Energy & Fuels, 24, 7, p. 4062-4077.

Stephenson, P.G., Moore, C.M., Terry, M.J., Zubkov, M.V. & Bibby, T.S. (2011),

"Improving photosynthesis for algal biofuels: toward a green revolution", .

Stripp, S.T. & Happe, T. (2009), "How algae produce hydrogen - News from the

photosynthetic hydrogenase", Dalton Transactions, 45, p. 9960-9969.

Tan, T., Shang, F. & Zhang, X. "Current development of biorefinery in China",

Biotechnology Advances, In Press, Corrected Proof,.

Uduman, N., Qi, Y., Danquah, M.K., Forde, G.M. & Hoadley, A. (2010), "Dewatering of

microalgal cultures: A major bottleneck to algae-based fuels", Journal of Renewable

and Sustainable Energy, 2, 1.

Ugwu, C.U., Aoyagi, H. & Uchiyama, H. (2008), "Photobioreactors for mass cultivation

of algae", Bioresource technology, 99, 10, p. 4021-4028.

Ultrawaves Reactors Ltd (2011), Ultrasound reactor for sewage sludge disintegration.

Valderrama, J.O., Perrut, M. & Majewski, W. (2003), "Extraction of Astaxantine and

phycocyanine from microalgae with supercritical carbon dioxide", 48, 4, p. 827-830.

Wahlen, B.D., Willis, R.M. & Seefeldt, L.C. (2011), "Biodiesel production by

simultaneous extraction and conversion of total lipids from microalgae,

cyanobacteria, and wild mixed-cultures", Bioresource technology, 102, 3, p. 2724-

2730.

Wang, L., Li, Y., Chen, P., Min, M., Chen, Y., Zhu, J. & Ruan, R.R. (2010), "Anaerobic

digested dairy manure as a nutrient supplement for cultivation of oil-rich green

microalgae Chlorella sp", Bioresource technology, 101, 8, p. 2623-2628.

Wanner, O., Debus, O. & Reichert, P. (1994), "Modelling the spatial distribution and

dynamics of a xylene-degrading microbial population in a membrane-bound

biofilm", Water Science and Technology, 29, 10, p. 243-251.

Weissman, J.C., Goebel, R.P. & Benemann, J.R. (1988), "PHOTOBIOREACTOR

DESIGN: MIXING, CARBON UTILIZATION, AND OXYGEN

ACCUMULATION", Biotechnology and bioengineering, 31, 4, p. 336-344.

Wijffels, R.H. & Barbosa, M.J. (2010), "An outlook on microalgal biofuels", Science,

329, 5993, p. 796-9.

49

Wiley, P.E., Brenneman, K.J. & Jacobson, A.E. (2009), "Improved algal harvesting using

suspended air flotation", Water Environment Research, 81, 7, p. 702-708.

Williams, P.J.L.B. & Laurens, L.M.L. (2010), "Microalgae as biodiesel biomass

feedstocks: Review analysis of the biochemistry, energetics economics", Energy and

Environmental Science, 3, 5, p. 554-590.

Xin, J., Imahara, H. & Saka, S. (2008), "Oxidation stability of biodiesel fuel as prepared

by supercritical methanol", Fuel, 87, 10–11, p. 1807-1813.

Yan, Q., Zhao, M., Miao, H., Ruan, W. & Song, R. (2010), "Coupling of the hydrogen

and polyhydroxyalkanoates (PHA) production through anaerobic digestion from

Taihu blue algae", Bioresource technology, 101, 12, p. 4508-4512.

Yirka, B. 2013, 12/04/2013-last update, First algae powered building goes up in

Hamburg [Homepage of Phys.org], [Online]. Available online:

http://phys.org/news/2013-04-algae-powered-hamburg.html#jCp [2013, 13/04/2013]

.

Zebib, T. (2008), Microalgae Grown in Photobioreactors for Mass Production of

Biofuel, unknown edn, Rutgers University, Department of Bioenvironmental

Engineering.

Zemke-White, W.L., Speed, S.R. & McClary, D.J. (2005), "Beach-cast seaweed: a

review".

Zhang, X., Hu, Q., Sommerfeld, M., Puruhito, E. & Chen, Y. (2010), "Harvesting algal

biomass for biofuels using ultrafiltration membranes", Bioresource technology, 101,

14, p. 5297-5304.