You are on page 1of 56

Contents

1. 2. Executive Summary .............................................................................................................................. 6 Introduction .......................................................................................................................................... 6 3.3. Software Used in the Analysis.......................................................................................................... 11 3.3.1 CA GREET ................................................................................................................................ 12 3.3.2 Simapro ...................................................................................................................................... 12 3.3.3 ASPEN ....................................................................................................................................... 13 3.4.2. Policy relevant metrics .............................................................................................................. 13 4. Cultivation and Harvesting .................................................................................................................. 15 4.1 Microalgae cultivation .................................................................................................................. 15 4.1.1 Strain selection ........................................................................................................................... 15 4.1.2. Open-pond systems ................................................................................................................... 16 4.1.3 Mixing ........................................................................................................................................ 16 4.1.4. Pumping .................................................................................................................................... 18 4.1.5 CO2 ............................................................................................................................................. 18 4.1.6 Fertilizers ................................................................................................................................... 19 4.2. Microalgae harvesting .................................................................................................................. 21 4.2.1. Dewatering for transesterification pathway ........................................................................... 21 4.2.1.1. Flocculant ............................................................................................................................... 22 4.2.1.2. Centrifuge .............................................................................................................................. 22 4.2.1.3. Drying .................................................................................................................................... 22 4.2.1.4. Summary of dewatering mass and energy inputs for transesterification pathway ................. 23 4.2.2. Dewatering for AD pathway ..................................................................................................... 24 4.2.2.1. Natural settling ....................................................................................................................... 24 4.2.2.2. EC/EF..................................................................................................................................... 24 5. Transesterification and AD of LERs ................................................................................................ 26 5.1. Solvent oil extraction ................................................................................................................... 26 5.1.1. Homogenization ........................................................................................................................ 27 5.1.2. Solvent Extraction ..................................................................................................................... 27 5.1.3. Alkali Refining.......................................................................................................................... 27 5.2. Base catalyzed transesterification ................................................................................................ 28

5.2.1. Summary of transesterification inputs ...................................................................................... 28 5.3. Biodiesel yield ............................................................................................................................. 29 Table 5-3 Biodiesel yield .................................................................................................................... 30 6. Anaerobic Digestion (AD) .............................................................................................................. 31 6.1. Digester operating conditions ...................................................................................................... 31 6.2. Biogas upgrading ....................................................................................................................... 32 6.3. Biogas Yield ................................................................................................................................ 33 7. Hydrogasification ............................................................................................................................ 35 8. Transportation of fuel production inputs and outputs ....................................................................... 36 8.1. Transportation of fuel production inputs...................................................................................... 36 8.2. Transportation and distribution of fuel products.......................................................................... 36 8.2.1 Transportation and distribution of biodiesel .............................................................................. 36 8.2.2. Transportation and distribution of CNG ................................................................................... 36 8.2.3 Transportation and distribution of syngas .................................................................................. 36 9. Coproducts .......................................................................................................................................... 37 9.1. Energy allocation ......................................................................................................................... 37 9.1.1. Energy allocation to algal oil .................................................................................................... 37 9.1.2. Energy Allocation to Biodiesel ................................................................................................. 37 9.2. Aquaculture feed .......................................................................................................................... 38 9.3.1. Biogas from lipid extracted algae residuals .............................................................................. 39 9.3.2. Compost and fertilizers from solid digestates ........................................................................... 39 9.4. WWT offsets ................................................................................................................................ 39 10. Capital (fuel production infrastructure) ............................................................................................ 41 10.1. Cultivation Capital ..................................................................................................................... 41 10.2. Harvesting Capital...................................................................................................................... 41 10.3. Biodiesel plant capital ................................................................................................................ 42 10.4. Biogas plant capital .................................................................................................................... 42 10.5. Hydrogasification plant capital .................................................................................................. 42 11. Waste and dismantling (end of infrastructure life cycle) ................................................................. 43 11.1. Cultivation waste and dismantling ............................................................................................. 43 11.2. Harvesting waste and dismantling ............................................................................................. 43 11.3. BD plant waste and dismantling ................................................................................................ 44

11.4. Biogas plant waste and dismantling ........................................................................................... 44 12.0. Results and discussion ....................................................................................................................... 46 12.1. Transesterification and AD pathway .............................................................................................. 46 12.2. Transesterification and hydrogasification pathway....................................................................... 49 12.3. AD pathway .................................................................................................................................... 49 12.4. Hydrogasification pathway ............................................................................................................ 52 13. Conclusion ............................................................................................................................................ 53 14. References ............................................................................................................................................. 54

Abbreviations AD anaerobic digestion BD biodiesel BOD biochemical oxygen demand CA California CAS Conventional Activated Sludge CARB California Air Resources Board CEC California Energy Commission CH3ONa Sodium methoxide CH4 methane CO2 carbon dioxide CO2-e carbon dioxide equivalent COD chemical oxygen demand CNG Compressed Natural Gas CSTR Continuously stirred tank reactor EC/EF Electro Flocculation/Electro Flotation EOR Enhanced Oil Recovery EPA U.S. Environmental Protection Agency FAME Fatty Acid Methyl Ester GHG greenhouse gas GREET Greenhouse Gases, Regulated Emissions, and Energy Use in Transportation HCl Hydrochloric acid HRAP High Rate Algal Ponds HRT Hydraulic Retention Time IPCC Intergovernmental Panel on Climate Change kgCO2-e kilogram(s) of carbon dioxide equivalent kWh Kilowatt-hour LCA life-cycle analysis LCFS Low Carbon Fuel Standard LER lipid extracted residual(s) LHV lower heating value MJ Megajoule MW Megawatts N nitrogen NaOH Sodium hydroxide NER Net Energy Ratio (ratio of energy inputs to energy outputs) NEV Net Energy Value (difference of net energy MJ/L fuel outputs and net energy MJ/L inputs) NH3 ammonia N2O nitrous oxide OLR Organic Loading Rate P phosphorus PBR photobioreactor

PTW pump to wheels (considers the usage by a vehicle portion of the life cycle) PVC Polyvinyl chloride RFS Renewable Fuels Standard TS total solids TSS total suspended solids VOC volatile organic compound VS volatile solids WTP well to pump WTW well to wheels (i.e., cradle to grave; considers entire fuel life cycle) WWT wastewater treatment

1. Executive Summary 2. Introduction Transportation accounts for about 50 percent of energy consumption and for over 38 percent of total greenhouse gas (GHG) emissions in the state of California (CA) (CEC, 2011) and the demand for transportation fuels is increasing. The number of light-duty vehicles is projected to grow from 25.6 million on-road vehicles in 2003 to 35.6 million by 2025. The increasing demand for fuel and the depletion of in-state and Alaskan petroleum will mean CA will become increasingly dependent on foreign imports of crude oil. Improving vehicle efficiency has been identified as the most effective way to reduce petroleum dependence. However, the California Energy Commission (CEC) and the California Air Resources Board (CARB) have concluded that improving vehicle efficiency alone will not be enough. In addition to improved vehicle efficiency, biofuels have been proposed as a part of the solution to the problems of dependence on foreign oil, and GHG emissions (CEC, 2011). Here, biofuel is defined as any fuel that is produced from any plant- or animal-based feedstock (often referred to as biomass). Biofuels offer a way to produce transportation fuels from renewable sources or waste materials and may reduce net carbon dioxide (CO2) emissions because the CO2 emitted during combustion of the fuel is captured during the growth of the feedstock. Policies such as the Renewable Fuels Standard (RFS) and the Low Carbon Fuel Standard (LCFS) lay the foundation for reaching the state of Californias goals of reducing imported petroleum, achieving significant reductions of GHG emissions from the use of renewable fuels, encouraging the development and expansion of the renewable fuels sector. The RFS program was created under the Energy Policy Act (EPAct) of 2005, and established the first renewable fuel volume mandate in the United States. As required under EPAct, the original RFS program (RFS1) required 7.5 billion gallons of renewable fuel to be blended into gasoline by 2012. The Energy Independence and Security Act (EISA) of 2007 updates the RFS to include diesel and increases the volume of renewable fuel required to be blended into transportation fuel from 9 billion gallons in 2008 to 36 billion gallons by 2022. For a biofuel to comply with the RFS requirement, it must demonstrate that greenhouse gas (GHG) emissions over the lifecycle of the fuel from production of the feedstock to processing of the fuel to combustion achieve a 50 percent reduction in emissions relative to petroleum-based fuel (Schremp et al., 2011). With a similar goal but different approach, the CA LCFS is focused on a performance standard for the entire transportation fuel pool rather than a volumetric requirement for biofuels only. The LCFS requires a reduction in the carbon intensity of transportation fuels that are sold, supplied, or offered for sale in the state by a minimum of 10% by 2020. Carbon intensity

reductions are based on reformulated gasoline mixed with 10% corn-derived ethanol and lowsulfur diesel fuel (Schremp et al., 2011). The renewable fuel mandate is based on first generation feedstocks such as corn, and soybeans. Current commercial processes convert simple sugars, starches, or oils to produce biofuels. The fermentation of cornstarch (from the corn kernel), sugar beets, or sugarcane produces ethanol, and the transesterification of oils (e.g., soybean or palm oil) produces biodiesel. The drawback of this technology is that competition with food crops for the use of arable land may lead to global increases in prices for food grains and vegetable oils (Pimentel et al., 2008). In addition, large scale production of corn ethanol may lead to higher emissions of GHGs than gasoline due to impacts associated with indirect land use changes (carbon emissions) and increased fertilizer use (nitrous oxide emissions) (Searchinger et al., 2008; Crutzen et al., 2008). Questions raised about the long term sustainability of first generation biofuels suggest it would be prudent to critically evaluate the next generation of biofuels before any large scale production takes place. Owing to its high vegetable oil content (2-75%) and biomass production rates, microalgae have the potential to displace feed stocks used for food (Wiley et al., 2010; Mata et al., 2010). Microalgae could have significant environmental benefits because they do not compete for arable land with food crops (Chisti, 2007), treat waste water to the tertiary level (Craggs et al., 1996), and may be able to utilize waste CO2 emissions from industrial or municipal processes to support pond growth (Kadam, 1997; Oswald and Golueke, 1968; Sheehan, et al., 1998). Algal biomass can be transformed into a variety of fuels including biodiesel, renewable diesel, renewable gasoline, renewable jet fuel, ethanol, and biomethane. However, it should be pointed out that large scale commercial algal biofuel plants do not exist. Algal biofuels are in the research and development phase and are considered next generation or emerging technologies. As such there are substantial uncertainties associated with potential of algal biofuel as an alternative transportation fuel and technology as defined by the RFS. For these reasons, algae biofuels have been the subject of several recently published life-cycle analyses (LCAs). In LCA, estimates are made for the materials and energy resources required for, and the environmental emissions resulting from, the production of a product or service. A common procedure for developing an LCA is to define the goal and scope, boundary for analysis, list the inputs, outputs, and externalities of interest, and define a functional unit for which the units can be evaluated. Quantitative metrics developed in LCA can help decision makers evaluate the viability of various potential systems for producing biofuels. Greenhouse gas (GHG) emissions are directly linked to energy use and as such, much emphasis is placed on energy analysis. A common approach is the net energy analysis which may be based on the computation of the ratio of total energy inputs from a system, in the form of fuels or electricity, to the total outputs. When calculating the GHG emission reductions from the use of biofuels, it is important to examine the full life-cycle of emissions from the fuel. Potential greenhouse gas emission reductions vary widely, depending on choices made at each step, from feedstock selection and production through conversion of the feedstock into a fuel, and then to final fuel use. Fossil fuels are often used in growing and processing feedstocks, which can increase the life-cycle emissions

for the biofuel. Changes in land use and land management practices to grow biofuel feedstocks also affect the GHG profile of a fuel. This study investigates the contribution algal biofuels may have in displacing fossil fuels in the future algae biofuels industry. The LCA consists of detailed models of each of the feedstock processing stages (cultivation, dewatering, extraction, and fuel conversion) combined with a standard set of metrics and boundary conditions. The four fundamental pathways are illustrated below in Figure 1. Pathways one and two include lipid extraction with conversion of algae cake through thermochemical or anaerobic digestion. Pathways three and four include direct conversion of whole algae through thermochemical or anaerobic digestion. This LCA seeks to answer the question of which pathway is most energetically favorable and has the greatest reductions in GHGs, and which pathway if any, satisfy the requirements for renewable fuels as defined by the RFS. As a benchmark for comparison, the net energy and GHGs for conventional diesel and soybean biodiesel (BD) are also evaluated.

Figure 2-1 Fundamental energy pathways

The materials and energy input associated with the cultivation of algae was assumed to be the same regardless of energy pathway referenced in Figure 2-1. Similarly, the processes are for anaerobic digestion (AD) and biogas upgrading are the same whether the feedstock is whole algal biomass as in pathway IV or lipid extracted residuals (LERs) in pathway II. However, the harvesting method was optimized for upstream processing of each energy pathway. In the case of transesterification, referenced in Figure 2-1, pathways I and II, the optimal harvesting method was determined chemical flocculant followed by dewatering with the centrifuge. For pathway IV, AD, the combination of natural settling and electro flocculation/electro flotation (EC/EF)

was assumed to be the optimal harvesting method. Three cases were considered for this analysis: (1) the baseline scenario where it was assumed nutrients were provided with commercially procured fertilizers and CO2, (2) a scenario where waste water (WW) provides the fertilizers for the pond but the CO2 was commercially procured, and (3) a scenario where fluegas was assumed to provide the CO2 but fertilizers were commercially procured. These scenarios are hereafter referred case 1, case 2, and case 3. Figures 2-2 and 2-3 are process overviews for fuel pathways 2 and 4 referenced in Figure 2-1.

Figure 2-2 Fundamental energy pathway II, Transesterification and AD of lipid extracted residuals

Figure 2-3 Fundamental energy pathway IV, AD of whole algal biomass

Because some processes apply to more than one fundamental energy pathway, as in the case of cultivation and harvesting, the report is organized topically by process. Section 3 discusses the methods used to conduct the analysis. The remaining section topics follow in order: 4. Cultivation and harvesting; 5. Transesterification and AD of LERs; 6. AD of whole algal biomass; 7. Steam hydrogasification; 8. Transportation of fuel production material inputs and distribution of finished fuel products; 9. Coproduct accounting; 10. Capital and infrastructure; 11. Waste and dismantling; 12. Results and discussion; and 13. Conclusions.

10

3. Methodology The material and energy flows associated with the cultivation, harvesting, and fuel production of microalgae were taken from literature or estimated using first principles. For upstream net energy and GHG emissions, the study followed an attributional approach for the fuel product and the product by material, energy flows or services that follow supply chain logic. The Greenhouse Gases, Regulated Emissions, and Energy Use in Transportation (GREET) framework was adopted for the analysis. In GREET energy is defined at the process level and emissions are defined at the technology level (Elgowainy, 2011). The relationship between energy and emissions is through technologies. A process employs technologies, and technologies employ fuels which in turn produce emissions. 3.1. Net energy calculation methodology The process related parameters include the input/output relation such as the efficiency, yield and energy intensity, and coproduct amount. The energy accounting throughout a pathway was calculated using the following (Elgowainy, 2011): where x is the energy in process fuels or input materials and upxi is upstream energy needed to produce xi amount of fuel or material i. 3.2. GHG emission methodology The emissions were calculated using the following equation (Elgowainy, 2011): [ ( ) () ( )

where Ei is the total process emissions of pollutant i, EFi(j, k) is the emissions factor of pollutant i when fuel j is used in technology k, and Share (j,k) is the share of fuel j used in technology k. Several emission impacts were not included due to a lack of data or high degree of uncertainty and include land use change, fugitive emissions from the anaerobic digester and biogas scrubber, and nitrous oxide emissions from the pond and WW used as nutrient source. Each of these impacts should be considered in future LCAs.

3.3. Software Used in the Analysis A description of the net energy and GHG impacts of microalgae transportation fuels requires a consistent model of the material inputs, energy use, and products for the process using various simulation tools. The modeling tools used in this analysis include, California (CA) GREET model, and Simapro and ASPEN software programs. Life cycle impacts (LCIs) were obtained from the CA GREET model and Simapro software databases. ASPEN software was used to simulate the hydrogasification fuel pathway.

11

3.3.1 CA GREET The framework and structure for the lifecycle inventory and system boundary for this study are provided in CA GREET developed by Argonne National Laboratories. The platform for the CA GREET model is a Microsoft Excel workbook and evaluates fossil fuel use (natural gas, coal, and petroleum) as well as the emissions of carbon dioxide, volatile organic compounds, carbon monoxide, nitrogen oxides, particulates, and sulfur oxides from all lifecycle stages for a given fuel production pathway. A summary of the life cycle impact factors are provided in Table 3-1. Table 3-1 Life cycle impact factors obtained in CA GREET Item Nitrogen Phosphorus Transportation of chemicals Electricity Natural Gas Hexane Methanol Sodium hydroxide BD transportation and distribution 3.3.2 Simapro Simapro is a LCA software program designed to model products and systems from a life cycle perspective. The software includes several LCI databases, accounts for various parameters, and has an option to address uncertainty with Monte Carlo analysis. In this study, Simapro was used in this study to obtain impacts not included in the CA GREET model. Table 3-2 is a summary of the impact factors obtained in Simapro. Table 3-2 Life cycle impact factors obtained in Simapro Item Concrete PVC pond liner PVC pipes Steel Cold rolled sheet steel Fiber glass Disposal concrete Waste water Carbon dioxide Synthetic flocculent Lime Chloroform Unit Energy Use (MJ) GHG (kg CO2e) 1 kg concrete 0.60 0.12 1 kg PVC 53.00 2.00 1 kg PVC pipe 59.10 3.18 1 kg steel 9.95 4.85 1 kg sheet steel 0.06 0.01 1 kg fiber glass 69.90 0.91 1 kg concrete 0.19 0.01 1 cubic meter wastewater 0.59 0.30 1 kg CO2 4.26 0.40 1 kg flocculent 60.30 1.88 1 kg lime 3.84 0.76 1 kg chloroform 36.60 4.11 Unit 1 kg N as urea 1 kg P as superphosphate 1 kg transported unit 1 MJ electricity 1 MJ natural gas 1 kg hexane 1 kg methanol 1 kg sodium hydroxide 1 kg transported unit Energy Use (MJ) 53.20 14.00 0.14 2.26 1.07 53.96 31.45 23.56 0.06 GHG (kg CO2e) 2.93 1.03 0.01 0.13 0.08 0.86 0.50 1.65 0.0002

12

Sulfuric Acid 3.3.3 ASPEN

1 kg sulfuric acid

1.72

0.12

Aspen Plus is a process modeling tool for conceptual design, optimization, and performance monitoring for the chemical, polymer, specialty chemical, metals and minerals, and coal power industries. The software includes the worlds largest database of pure component and phase equilibrium data for conventional chemicals, electrolytes, solids, and polymers. The database is regularly updated data from the U.S. National Institute of Standards and Technology (NIST). 3.4. Scope of analysis and policy relevant metrics 3.4.1. System Boundary The system boundary used in this analysis includes the production of algae and biofuels, BD (via a transesterification reaction), biogas (with anaerobic digestion), and syngas (hydrogasification), from well-to-tank and is represented by the dashed-line box in Figure 3-1. The terminology, well-to-pump (WTP), and framework for the analysis is based on the CA GREET model. The boundary includes all upstream mass and energy flows that are required to make the chemical and energy resources required for processing. Emissions are accounted for during all stages of fuel production and coproducts and include GHGs associated with direct energy (e.g. grid electricity and natural gas consumption) as well as upstream energy use for materials (e.g. production and transportation of consumable chemicals and structural materials to build the algae-to-fuel refinery, etc). While the inputs associated with the materials used to build the capital (pond and fuel processing facilities) are included in the analysis, algae farm labor is not.
Cultivation: CO2 and fertilizers (production and transport), electricity, pond capital Harvesting: Chemicals (production and transport), electricity, natural gas (NG), and capital pond capital Fuel: Chemicals (production and transport), electricity, NG, coproducts, and pond capital Vehicle: Vehicle use

Well-to-Pump (WTP)

Pump-to-Wheel (PTW)

Wells-to-Wheels (WTW)

Figure 3-1 LCA system boundaries enclosed in dashed lines; adapted from CA GREET (Frank et al., 2011)

3.4.2. Policy relevant metrics It is important to choose the appropriate metric for the goal of the assessment, and the analysis framework support the selected metric. Commonly used metrics used to determine the energy and GHG performance of a biofuel system are quantities of fuel produced, yield per unit land

13

area, or vehicle distance driven. The results reported in this study are net energy MJ/MJ fuel and the policy-relevant climate metric of net kg CO2-equivalent emissions / MJ fuel.

14

4. Cultivation and Harvesting The primary components of the LCA include: (1) microalgae cultivation in outdoor high rate algal ponds (HRAP), (2) microalgae harvesting using either a combination of chemical flocculants and/or natural settling, and mechanical processes such as centrifugation, and EC/EF, and (3) simulation of fuel conversion processes including transesterification for BD, hydrogasification for syngas, and AD for biogas. For the analysis of scenarios considered in this study, regardless of the fuel conversions process, it is assumed the simulation of the microalgae cultivation will be the same. 4.1 Microalgae cultivation Some key considerations for modeling the net energy and GHGs of the cultivation process include (1) strain selection, (2) HRAP and photobioreactor (PBR) design considerations, (3) mixing, (4) water availability and pumping, (5) methods for CO2 supply and delivery, and (6) nutrient supply. 4.1.1 Strain selection The growth rate, biomass fractions (protein, carbohydrates, and lipids), and heating value of the of the algae biomass are constrained by the strain selection. Several authors report high lipid productivity of algae cells (ranging from 2 to 75% by mass) resulting from nitrogen starvation (Wiley, et al., 2010). However, intentionally cultivating algae in nutrient stressed conditions inhibits cell division, which can lead to decreased overall biomass productivity (Wiley, et al., 2010). Lipid content is an important consideration for the production of BD but less so for the production of methane-rich biogas by AD or syngas by hydrogasification (Wiley et al., 2010; Sialve et al., 2009). The genus and species selected for this study is Chlorella vulgaris because of the abundance of data in literature for the growth and biomass fractions, and subsequent fuel conversion processes. As noted in the literature, algae biomass production rates can vary highly between and even within species and depend on the level of inputs (Wiley et al., 2010). A commonly reported growth rate in the literature is 25 g m-2 day-1 (Lardon et al., 2009; Davis et al., 2010; Frank et al., 2011; Clarens et al., 2011) for open pond systems. Due to the consensus in the literature, the base-line scenario in this study assumed a growth rate of 25 g m-2 and biomass components reported in Lardon et al., (2009) (Table 3): 282 gkg C.vulgaris -1 protein, 175 gkg C.vulgaris -1 lipid, and 495 gkg C.vulgaris -1 carbohydrates. This results in net biomass fractions of approximately 18% lipids, 51% carbohydrates, and 29% protein with a calorific value of 17.5 MJ. Table 3-3 Biomass fractions (Lardon et al., 2009) Fraction protein carbohydrate Formula C4.43H7O1.44N1.16 C6H12O6 Molar mass (gmol-1) 100.1 180.0 Net calorific value (MJkg-1) 15.5 13.0

15

lipid

C40H74O5

634.0

38.3

4.1.2. Open-pond systems HRAP systems are the most commonly reported growth systems in the literature and all very large commercial systems used today are of this type (Dragone et al, 2010). Systems such as raceway ponds are simple to construct, less expensive to build and operate, and have a larger production capacity than most closed systems. Furthermore, they can utilize sunlight and the nutrients can be provided by channeling the water from sewage/water treatment plants making this the least expensive method of large-scale algal biomass production (Wiley et al., 2010). HRAPs are typically made of a closed loop, oval shaped recirculation channels, generally between 0.2 and 0.5 m deep, with mixing and circulation required to stabilize algae growth and productivity. In a continuous production cycle, algae broth and nutrients are introduced in front of the paddlewheel and circulated through the loop to the harvest extraction point. The paddlewheel is in continuous operation to prevent sedimentation. Although these systems are the most widely used at the industrial level, HRAPs still present significant technical challenges. HRAPs are susceptible to weather conditions, evaporation, and lack controls for water temperature and lighting, which make these systems dependent on the prevailing regional climate conditions (daily and annual temperature range, annual rainfall and rainfall pattern, number of sunny days, and degree of cloud cover) (Dragone et al., 2010). Furthermore, contamination by predators and other fast growing heterotrophs have restricted the commercial production of algae in open culture systems to fast growing, naturally occurring species (Dragone et al, 2010). Consequently, the species that can successfully grow in open systems is limited to Dunaliella (tolerant of high salinity), Spirulina (tolerant of high alkalinity) and Scenedemus and Chlorella (tolerant of nutrient-rich media) (Dragone et al., 2010). While PBRs allow for the regulation and control of nearly all the biotechnologically important parameters and generate higher biomass concentrations compared with HRAPs, the high operating and construction cost of PBR systems may limit their application for large-scale production of low value products such as biofuel (Lardon et al., 2009; Clarens et al., 2010; Wiley et al., 2010). Due to the consensus in the literature, the HRAP was the pond design choice for this study. It was assumed the HRAP would be constructed of concrete and lined with PVC with a Manning friction factor of 0.010 and a depth of 0.30 m. The pond infrastructure lifetime was assumed to be 30 years. At the end of the 30-year lifespan concrete is sent to the landfill for disposal whereas steel and PVC are recycled.

4.1.3 Mixing Mixing contributes to the growth of microalgae (Dragone et al., 2010) and prevents cells from settling and thermal stratification, distributes nutrients, breaks down diffusion gradients at the cell surface (Oswald, 1988). Paddlewheels are typically used to mix open pond systems and the power consumption for these systems depends not only on mixing speed but also on pond depth,

16

bottom roughness, the mechanical efficiency of the paddle wheel, and operating time (Oswald, 1988). For this study, it was assumed the pond is mixed at a moderate velocity of 20 cm/sec and a depth of 30 cm, with a train efficiency of 40% (Benemann et al., 1996). The mixing energy required per hectare (kWhha-1) as taken from Oswald (1988): Power requirements can be determined on the basis of change in depth of the pond, d, and the quantities of water in motion according to the relation: ; Where P is power (kW); Q the quantity of water in motion (m3s-1); W is the specific weight of water, 9.804 (kNm-3 at 10C); d is the change in depth (m), and e is the efficiency of the paddle wheel. For a channel of width w and depth d the quantity of flow Q (m3s-1) is given by:

in which w and d are expressed in meters and V is in ms-1. For this study we assumed paddle wheel channel dimensions and resulting power consumption as follows: a 6 m wide channel, 30 cm deep, with liquid flowing at a velocity of 20 cms-1, the power required, assuming a paddle wheel efficiency of 0.40 is:

For 24h, the power consumption would be

For a 1 hectare pond in continuous operation (365 days per year) in units MJha-1:

or 0.34 MJ/kg algae.

17

4.1.4. Pumping Pumping is required to bring piped water and/or WW and recycled effluent to the ponds and to remove the pond medium for biomass extraction. From a design perspective, many assumptions have to be made that will influence the overall power requirements which is likely oversimplified in the literature. Some of the questions to consider include: (1) water source (on-site bore well, municipal supply, or off-site) which means power requirements are site specific, (2) headloss which is influenced by pipeline roughness, the length of pipe, the pipe diameter, and the velocity of the water, (3) volumetric flow rate, and (4) the pump efficiency. The energy for pumping will depend on the power, and the assumed operating time. For this study, several design choices and assumptions were made with respect to the pumping energy calculations. Based on a steady state system and an algal density of 0.5 g/L, the flow rate was calculated to be 41 m3h-1. PVC pipe extends 500 meters with a diameter of 0.20 meters. Head losses were calculated with the Colebrook White equation (Munson et al., 2009). The static lift is assumed to be 10 meters. The power for pumping was calculated using the equation ) (Munson et al., 2009): ( where Q is the flowrate (m3s-1), g is gravity (9.8 ms-2), is the density of water (999 kgm-3), h is the headloss (m), and e is the efficiency of the pump. This results in approximately 51,614.93 MJ of pumping power per hectare per year or 0.63 MJ/kg algae. 4.1.5 CO2 Providing CO2 to algal cultures increases growth rates and maintains optimal pH. Flue gas from coal-fired powered plants has been suggested as a source of CO2 for algal cultivation operations (Kadam, 2002). While this may decrease the net energy associated with delivering CO2 to offsite locations, there is uncertainty regarding how certain flue gas constituents (e.g., SOX, NOX, mercury) will affect algae growth rates (Clarens et al., 2010). Furthermore, collocation of algal cultivation systems with power plants may limit access to WW effluent, thereby increasing fertilizer requirements for media preparation (Wiley et al., 2010). Waste CO2 generated during AD processes at wastewater treatment (WWT) facilities has been suggested as a source of carbon that may have fewer impurities compared with flue gas while also providing access to WW effluent (Wiley et al., 2010). Most commercial CO2 is recovered as a by-product of other processes, such as the production of ethanol by fermentation, the manufacture of ammonia, petroleum refineries petrochemical plants, hydrogen production plants, and other combustion and industrial process sources (EPA, 2010). Currently, the majority of CO2 transported for commercial use is transported by pipeline. In the western US, carbon dioxide pipelines extend over 2500 km where they carry 50 MtCO2 yr-1 from natural sources to enhanced oil recovery projects (EOR) in the west Texas and elsewhere (IPCC, 2005). The CO2 being used in EOR is primarily produced from naturally-occurring underground CO2 reservoirs, but is also captured from natural gas processing, coal syngas, and ammonia plants. In this study, two scenarios were considered for delivering CO2 to the algal ponds: (1) pure CO2 was delivered by truck, and (2) CO2 was supplied via piped flue gas consisting of 15% CO2.

18

In the first scenario it was assumed CO2 was bubbled into the pond via an automated control system where energy consumption was related to upstream processes associated with the manufacture of CO2 and transportation via truck (Clarens et al., 2010). Here we assume 100% utilization efficiency, and based on the biomass fractions given in Table 1, the algae require 1.8 kg CO2/kg algae. However, because the biogas scrubbing water is not regenerated, the CO2 and other impurities in the scrubbed from the biogas are recycled to the pond reducing the overall demand for energy and GHG emissions associated with the procurement of commercially purchased CO2. In the second scenario, it was assumed direct energy consumption would be required for delivery of the gas via pipeline and for pressurization and injection in the ponds. In a study conducted by Clarens et al., (2011) it is argued that both the power plant and the algae system benefit from the capture of CO2 and thus the energy burden is allocated between the power plant and the algae farm on an energy basis. This was done by expanding system boundaries for algae cultivation to also include the power plant from which CO2 is ultimately derived. The amount of CO2 burden shared by the power plant and the algae system is X Mg. The amount of algae that can be derived from X Mg CO2 is Y Mg, which in turn, gives rise to A MJ of algae-derived energy in the form of BD and/or bioelectricity. X Mg of shared CO2 also corresponds to some quantity of fossil fueled electricity production, f, as computed using the factor 1 Mg CO2/MWh for a conventional power plant [7]. The fraction of CO2 procurement burden allocated to algae farming, fALGAE- CO2, is given by the following equation (Clarens et al., 2011):

In the case of flue-gas derived CO2, it was assumed the efficiency of the delivery system is 75% (Lundquist et al., 2010) and based on the biomass fractions given in Table 1 results in a demand of 2.4 kg CO2/kg algae. However, as in the case of commercially purchased CO2, the demand for flu-gas derived CO2 is reduced by the recycle of biogas upgrading scrubbing water. It should be emphasized that for this study we made simplified estimates of the energy required for CO2 pressurization and distribution, and that a more detailed analysis would require data that is site specific. For instance, it is likely that some of the land within the served area will either be unsuitable for algal ponds or already in use. In a study conducted by Benemann and Oswald (1996) it is estimated that transport of CO2 was limited to 2.5 to 5.0 km due to economic constraints, with 2.5 km more likely. 4.1.6 Fertilizers Algae growth is limited by several nutrients, such as nitrogen (N), phosphorous (P), and carbon. Each nutrient is an important component of the algal growth cycle. On the basis of Table 3, it is possible to estimate N requirements and according to algae composition. The necessary amounts of N and P are assumed to be proportional to the protein content, and then indirectly to the N fraction of the biomass (Lardon et al., 2009).

19

For the baseline scenario, with a 29% protein biomass fraction, the fertilizer demand was assumed to be 46 g N/kg algae and 9.9 g P/kg algae (Lardon et al., 2009). Two scenarios were considered for nutrients: (1) commercially purchased fertilizers, and (2) secondary WW from conventional activated sludge (CAS). In both scenarios, nutrients were assumed to be recovered from the anaerobic digester to offset the procurement of off-site fertilizers required to support pond growth. Here, the baseline scenario considers a recovery rate of 80% of the N and P retained as ammonium/NH3 and P in the supernatant that is returned to the pond. The 20% fraction of N remaining in the solids is treated as organic N with a bioavailability of 40% that is exported as a fertilizer replacement co-product that displaces the average emissions for N and P production (Frank et al., 2011). In the case of commercially purchased fertilizers, the fossil energy demand and equivalent CO2 emissions associated with production include transportation and distribution of the chemical fertilizers. In CA GREET, the N and P are delivered to the farm in a medium duty truck that travels a one-way distance of 30 miles. Due to the recycling of N and P in the AD effluent, the resulting demand of off-site fertilizer was assumed to be 9.2 g N/kg algae and 1.98 g P/kg algae. In the second scenario, N and P are provided by secondary WW from CAS. There are many sources of eutrophic or mixed waters (such as animal litter, secondary wastewater, and agricultural or industrial effluents) rich in N, P, and other nutrients and minerals. The use of WWT effluents have offsets associated with reduced need for fertilizer production and WWT and the associated material and energy inputs. Additionally, the use of WW offsets freshwater usage. The N and P content of secondary WW from CAS is reported to be 15-35 mg/L for N and 4-10 mg/L for P (Clarens, 2010). For this study, the baseline scenario takes the midpoint values, with 25 mg/L for N and 7 mg/L for P. Here, it was assumed WW provides the necessary nutrients such that additional fertilizer would not be required. In this case, there are coproducts associated with the avoidance of WWT, and nutrients recycled from the digester. 4.1.8 Summary of cultivation inputs. Table 4-1 summarizes the material and energy inputs and upstream energy associated with the cultivation of microalgae. The numbers 1, 2, and 3 represent scenarios commercially purchased fertilizers and CO2, fertilizers from CAS WW and commercially purchased CO2, and commercially purchased fertilizers and flue gas CO2, respectively. The letters L, B, and H represent low (worst case scenario), base-line, and high (best case scenario), respectively. To address uncertainty in the analysis, the electricity consumption associated with cultivation was adjusted by 20%, an increase for low and a decrease for high scenarios. In addition, the mineralization rate of the liquid AD digestates was adjusted to 0.70 and 0.90 from the baseline for the low and high cases, respectively. In case 2, where it was assumed CAS WW provides the nutrients for the pond, nitrogen was adjusted to 15 and 35 mg/L and phosphorous to 4 and 10 mg/L for the low and high cases, respectively. Lastly, the amount of CO2 in the biogas dictates the amount ultimately recycled to the pond with the high case corresponding to higher biogas yields and lower CO2 demand, and vice versa for the low case.

20

Table 4-1 Cultivation mass and energy inputs


Cultivation Phase Direct Energy Mixing electricity (MJ/kg algae) Pumping electricity (MJ/kg algae) CO2 pressurization (MJ/kg algae) CO2 Application rate (kg CO2/kg algae) Subtotal direct energy (MJ/kg algae) Indirect Energy Nitrogen Impact Factor (MJ N/kg N) N Application rate (kg N/kg algae) Nitrogen (MJ N/kg algae) Phosphorus Impact Factor (MJ/kg P2O5) P2O5 application (kg P2O5/kg algae) Phosphorous (MJ P2O5/kg algae) Subtotal fertilizers (MJ fertilizers/kg algae) Electricity Impact Factor (MJ/MJ electricity) Electricity (MJ/kg algae) CO2 (MJ CO2/kg CO2) CO2 Application rate (kg CO2/kg algae) CO2 (MJ CO2/kg algae) Capital (Pond and equipment) (MJ/kg algae) Waste and dismantling (MJ/kg algae) Subtotal Indirect Energy (MJ/kg algae) 1L 0.40 0.75 1B 0.34 0.63 1H 0.27 0.50 2L 0.40 0.75 2B 0.34 0.63 2H 0.27 0.50 3L 0.40 0.75 0.39 2.17 1.54 53.20 0.014 0.73 14.00 0.003 0.04 0.78 2.26 1.94 3B 0.34 0.63 0.41 2.00 1.37 53.20 0.009 0.49 14.00 0.002 0.03 0.52 2.26 1.72 3H 0.27 0.50 0.41 1.82 1.18 53.20 0.005 0.24 14.00 0.001 0.01 0.26 2.26 1.48

1.16 53.20 0.014 0.73 14.00 0.003 0.04 0.78 2.26 1.46 4.26 1.39 5.92 1.34 0.45 9.94 11.10

0.96 53.20 0.009 0.49 14.00 0.002 0.03 0.52 2.26 1.21 4.26 1.39 5.92 1.34 0.45 9.44 10.40

0.77 53.20 0.005 0.24 14.00 0.001 0.01 0.26 2.26 0.97 4.26 1.39 5.92 1.34 0.45 8.94 9.71

1.16 53.20 0.005 0.26 14.00 0.001 0.01 0.27 2.26 1.46 4.26 1.58 6.73 1.34 0.45 10.24 11.40

0.96 53.20 0.000 0.00 14.00 0.000 0.00 0.00 2.26 1.21 4.26 1.39 5.92 1.34 0.45 8.92 9.89

0.77 53.20 0.000 0.00 14.00 0.000 0.00 0.00 2.26 0.97 4.26 1.22 5.20 1.34 0.45 7.96 8.73

Total Cultivation Phase (MJ/kg algae)

1.34 0.45 4.51 6.05

1.34 0.45 4.03 5.40

1.34 0.45 3.53 4.71

4.2. Microalgae harvesting Separating the algae from water remains a major hurdle to industrial scale processing partly because of the small size of the algal cells, with unicellular eukaryotic algae typically 330 m (Molina Grima et al., 2003), and cyanobacteria as small as 0.22 m (Chorus and Bartram, 1999). In addition, the algae are suspended in relatively broths of 200600 mg/l (Uduman et al., 2010), and require large volumes of water be processed. The initial harvesting step also affects later processes downstream. Lowering the energy inputs in harvesting algae and harvesting in a way that allows for the utilization of coproducts remains a challenge. Given the high energy inputs associated with harvesting, the dewatering process should be optimized for the fuel production pathway. For instance, it is typically necessary to dry the algae cake to 90% for oil extraction and transesterification for the production of BD (Lardon et al., 2009) whereas for biogas produced in the process of AD, no drying is required as typically the organic loading rate is 5% total solids (Collett et al., 2011). 4.2.1. Dewatering for transesterification pathway In this study, for the baseline scenario, it was assumed the algae would be dried to 90% dry matter prior to transesterification. This would involve several sequential steps: (1) concentration of the dilute algal broth from 0.5 g/L to 20 g/L with the use of chemical conditioners, (2) dewatering of the algal slurry with a centrifuge resulting in a 10 fold concentration of 200 g/L,

21

and (3) algal cake is dried to 900 g/L with an industrial air dryer. The assumptions for each dewatering step are discussed in more detail in the sections that follow. 4.2.1.1. Flocculant Chemical conditioning with coagulating agents is frequently employed in WWT operations to improve the efficiency of the subsequent dewatering. In the literature, it is commonly assumed that the chemical conditioning techniques for sewage sludge apply to algae (Lardon et al., 2009; Clarens et al., 2010; Clarens et al., 2011; Frank et al., 2011). For the baseline scenario in this study, it was assumed that the addition of 0.5 g m-3 of a synthetic flocculant and the addition of lime up to a pH of 11 (i.e., 300 g m-3) would flocculate 90% of the algal biomass (Lardon et al., 2009). This results in flocs characterized by a settling speed of 2 mh-1 and a concentration of 20 kg m-3 (Lardon et al., 2009). 4.2.1.2. Centrifuge A centrifuge is a separation process that uses centrifugal force to separate suspended solids from liquids (Wiley et al., 2011). In the literature, recovery of algal slurries of 10-22% total solids content with a feed concentration of 0.1 to 2% have been reported using a decanter bowl centrifuge (Wiley et al., 2011). Energy consumption and dewatering performance vary considerably depending on the type of centrifuge employed. Mass retention efficiencies of 85 97% are reported for self-cleaning disc stack centrifuges and of 8595% for a decanter bowl (Frank et al., 2011). Power requirements for a centrifuge are within the range of 0.9-8 kWh/m3 (Wiley et al., 2010). For this study, it was assumed the Evodos spiral plate centrifuge would dewater the algae to a dry weight concentration of 200 kg m-3. The Evodos spiral plate centrifuge is reported to dewater algae with a retention efficiency of 95%, a dry solids concentration of 31.5% with a dry solids feed concentration of 0.3% resulting in a concentration factor of 100 (Evodos, 2011). According to the manufacturers specifications, 1.1 kWh/m3 is required to operate the centrifuge, and 0.25 kWh/m3 to operate the pump (Evodos, 2011). We assume the manufacturers power specifications, 95% retention efficiency, and a concentration factor of 10. 4.2.1.3. Drying In the literature, it is generally assumed that the algal biomass must be dried to 90% dry matter prior to lipid extraction based on industrial scale processes for canola and soybean oil extraction (Lardon et al., 2009; Sheehan et al., 1998). In this study, two industrial scale drying scenarios were considered (1) the first, and the baseline scenario, it was assumed a gas powered industrial air dryer would provide the drying heat and (2) the other scenario assumed algal biomass would be dried from the waste heat provided from a coal power plant. In the first scenario, it was assumed that algae leaving the centrifuge would require additional dewatering requiring 13.8 MJ heat energy per kg of dry algae and 400 Wh/kg algae to operate the conveyer belt required to move the algae to the drying unit (Lardon et al., 2009, Hassebrauck and Ermel, 2006).

22

The second scenario, it was assumed the energy for drying the algae biomass would be provided from power plant flue gas waste heat. Waste heat is routinely recovered from exhaust gases for use in preheating the combustion air so that higher overall thermal efficiencies may be obtained. The power required for operation of the conveyer dryer is 0.0094 MJ/kg algae biomass (Clarens et al., 2011). Based on the estimate that a 500 MW power plant can provide up to 4.4x109 MJ/year of waste heat (Clarens et al., 2011), it seems reasonable to assume that sufficient waste heat would be available for drying the algae biomass. There are approximately 1,008 power plants in CA with a total capacity of 69,709 MW. As noted in section 4.1.5., it is likely that some of the land within the power plant area will either be unsuitable for algal ponds or already in use. In this study we made simplified estimates of the energy required for biomass drying. A more detailed analysis would require site specific information and include spatial analysis of available land suitable for algae production near power plants. 4.2.1.4. Summary of dewatering mass and energy inputs for transesterification pathway

Table 4-2 summarizes the material and energy inputs and upstream energy associated with the harvesting of microalgae. The numbers 1, 2, and 3 represent scenarios commercially purchased fertilizers and CO2, fertilizers from CAS WW and commercially purchased CO2, and commercially purchased fertilizers and flue gas CO2, respectively. The letters L, B, and H represent low (worst case scenario), base-line, and high (best case scenario), respectively. To address uncertainty in the analysis, the electricity consumption and natural gas heat associated with harvesting was adjusted by 20%, an increase for low and a decrease for high scenarios. Table 4-2 Harvesting mass and energy inputs for transesterification and AD pathway
Harvesting Phase Direct Energy Dewatering Electricity (MJ/kg algae) Drying heat (MJ/kg algae) Electricity for conveyer belt (MJ/kg algae) Subtotal direct energy (MJ/kg algae) Indirect Energy Flocculant Impact factor (MJ/kg flocculant) Flocculant (kg synthetic flocculant/kg algae) Flocculant (MJ/kg algae) Lime Impact factor(MJ/kg lime) Lime (kg lime/kg algae) Lime (MJ/kg algae) Electricity impact factor(MJ/MJ Electricity) Electricity (MJ/kg algae) Natural Gas impact factor (MJ/MJ NG) Natural gas (MJ/kg algae) Transportation impact factor (MJ/kg chemical) 1,2 L 0.29 16.6 1.73 18.6 60.3 0.001 0.06 3.84 0.30 1.15 2.26 2.54 1.07 1.16 0.14 1,2 B 0.24 13.8 1.44 15.5 60.3 0.001 0.06 3.84 0.30 1.15 2.26 2.12 1.07 0.97 0.14 1,2 H 0.19 11.0 1.15 12.4 60.3 0.001 0.06 3.84 0.30 1.15 2.26 1.70 1.07 0.77 0.14 3L 0.29 0.00 1.73 2.0 60.3 0.001 0.06 3.84 0.30 1.15 2.26 2.54 1.07 0.00 0.14 3B 0.24 0.00 1.44 1.7 60.3 0.001 0.06 3.84 0.30 1.15 2.26 2.12 1.07 0.00 0.14 3H 0.19 0.00 1.15 1.3 60.3 0.001 0.06 3.84 0.30 1.15 2.26 1.70 1.07 0.00 0.14

23

Transportation of chemicals (MJ/kg algae) Capital (Equipment) (MJ/kg algae) Waste and dismantling (MJ/kg algae) Indirect energy harvesting phase (MJ/kg algae) Total Harvesting Phase (MJ/kg algae)

0.04 0.99 0.24 6.2 24.8

0.04 0.99 0.24 5.6 21.1

0.04 0.99 0.24 5.0 17.3

0.04 0.99 0.24 5.0 7.0

0.04 0.99 0.24 4.6 6.3

0.04 0.99 0.24 4.2 5.5

4.2.2. Dewatering for AD pathway For the baseline scenario in this study, the optimal harvesting method was assumed to be natural settling and EC/EF. The process of natural settling was desirable for AD because this process eliminates chemical conditioners which may have an adverse effect on the microbes in the digester. This harvesting method was selected for the baseline for this fuel pathway because it is low in energy demand and it provides the optimal total solids concentration for anaerobic digesters, 5%. AD does not require subsequent drying of the biomass for fuel production. 4.2.2.1. Natural settling Natural settling is inexpensive, low in energy intensity, and does not require chemical conditioners. This procedure involves the separation of flocculated algae cells from liquid using gravitation forces (Wiley et al., 2010). For this study it was assumed C. vulgaris settles with a sedimentation velocity of 3.575 m d-1, allowing the collection of 65% of the algal biomass with a concentration 20 times higher than in the culture stream after one hour (Collet et al., 2011). This results in a stream leaving the natural settler with an algal concentration of 10 kg/m3. 4.2.2.2. EC/EF EC/EF is a flotation technology that generates coagulating species for destabilizing algal suspensions in situ, through the electrochemical oxidation of consumable metal electrodes (Wiley et al, 2010). Electrodes, made of aluminum or iron, release charged ions into solution when direct current is applied to the system (Wiley et al., 2010). The ions hydrolyze to coagulating agents, polymeric iron or aluminum hydroxide (Wiley et al, 2010). Hydrogen and oxygen bubbles are produced during this process at the anode and cathode, respectively, which float the sample to the surface of the flotation cell, where it is removed with a skimming mechanism (Wiley et al., 2010). One of the key advantages of EC/EF is that it generates coagulating species in situ, thereby eliminating costs associated with purchasing, transporting, storing, and handling chemicals (Wiley et al, 2010). Another advantage is that it has few moving parts, only requires electricity and periodic replacement of the sacrificial electrodes (Wiley et al., 2010). In the literature, the power to operate the EC/EF is found to be in the range of 0.3-2 kWh/m3. For this study, the midpoint of the range, 1.15 kWh/m3, was the assumed the energy requirement for EC/EF device.

24

4.2.2.3. Summary of dewatering mass and energy inputs for AD pathway Table 4-3 summarizes the material and energy inputs and upstream energy associated with the harvesting of microalgae. The numbers 1, 2, and 3 represent scenarios commercially purchased fertilizers and CO2, fertilizers from CAS WW and commercially purchased CO2, and commercially purchased fertilizers and flue gas CO2, respectively. The letters L, B, and H represent low (worst case scenario), base-line, and high (best case scenario), respectively. To address uncertainty in the analysis, the electricity consumption associated with EC/EF was adjusted from the baseline 1.15 kWh/m3 to 0.3 and 2 kWh/m3, for the high and low scenarios, respectively. Table 4-3 Harvesting mass and energy inputs for AD pathway
Harvesting Phase Direct Energy EC/EF (MJ/kg algae) Subtotal direct energy (MJ/kg algae) Indirect Energy Electricity impact factor(MJ/MJ Electricity) Indirect energy dewatering electricity (MJ/kg algae) Capital (Equipment) (MJ/kg algae) Waste and dismantling (MJ/kg algae) Indirect energy harvesting phase (MJ/kg algae) Total Harvesting Phase (MJ/kg algae) 1,2,3 L 0.720 0.720 2.26 0.91 0.990 0.239 2.136 2.856 1,2,3 B 0.414 0.414 2.26 0.52 0.990 0.239 1.750 2.164 1,2,3 H 0.108 0.108 2.26 0.14 0.990 0.239 1.365 1.473

4.3.1. Dewatering for steam hydrogasification pathway

25

5. Transesterification and AD of LERs BD is a renewable fuel that can be produced from biological oils derived from plants, animals, or microbes (Wahlen et al., 2010). The oils (triglycerides) are converted by transesterification using alcohols (e.g., methanol) and catalyst (base or acid) to yield glycerol and the fatty acid alkyl ester or fatty acid methyl ester (FAME) if methanol is the alcohol as illustrated in Figure 3. Biodiesel can be produced through direct transesterification of algal biomass or by a two-step process by which lipids are extracted, collected, and transesterified (Wiley et al., 2010). Either process requires lipid extraction using combinations of solvents and alcohols, such as chloroform/methanol, hexane/ispropanol, or petroleum ethers and methanol (Wiley et al., 2011). The advantage of the direct method is that it combines lipid extraction and transesterification into one process, making it less time consuming than two-step transesterificiation processes (Wiley et al., 2010). Another scenario proposed in LCAs is a wet extraction of the algal oil. Lardon et al. (2009) proposes wet extraction results in lower drying demands but increased solvent use and lower overall yields. However, there is very little experimental data for the transesterification process of algal biomass with the two-step method and even less for the direct and wet extraction methods. Therefore, for this study, the analysis will be limited to the two-step process requiring drying of the algal biomass. The two-step transesterification is a sequential process that involves the process of homogenization to break the algae cell walls, lipid extraction with hexane solvent, alkali refining, and transesterification with methanol. Section 5.1 provides an overview of the solvent extraction, oil recovery process, and assumptions made in this study. Section 5.2 provides an overview of the transesterification process and assumptions made in this study. 5.1. Solvent oil extraction

Recovery of algal oil involves homogenization of the algal slurry to increase lipid yield, solvent extraction with hexane, and alkali refining of the extracted oil. Each step of the oil recovery and the mass and energy associated with each process is discussed in the following sections.

26

5.1.1. Homogenization It was assumed that homogenization would disrupt the algal cell walls requiring 67 MJ of electricity per cubic meter of algal slurry (Stephenson et al., 2010). Furthermore, it was also assumed two passes would be required to achieve 96% efficiency (Stephenson et al., 2010). Therefore, this results in a demand of 0.670 MJ/kg algae for homogenization. 5.1.2. Solvent Extraction The standard solvent extraction process uses n-hexane produced from petroleum. Most of the nhexane used in oil extraction is recovered and recycled, with a small fraction of loss. However, it should be emphasized there is very little experimental data available on algal lipid extraction with solvents and the process has not been demonstrated at scale. Most LCA studies have extrapolated data from bench scale or have based the extraction on soy or canola oil mills (Lardon et al., 2009; Stephenson et al., 2010; Lundquist et al., 2010). In calculating emissions and energy use, it was assumed that steam was generated from natural gas. N-hexane is a straight-chain hydrocarbon manufactured by distillation of straight-run gasoline produced from crude oil or natural gas liquids (Huo et al., 2008). In CA GREET, hexane is assumed to be produced from crude oil, and its upstream production energy use and emissions are adopted from energy use and emissions calculated for production of LPG from crude oil (Huo et al., 2008). Because hexane is volatile, the amount of hexane lost during soy oil extraction is assumed to be in the form of VOC emissions to the atmosphere (Huo et al., 2008). For this study it was assumed 1.03 MJ heat/kg algae, 0.328 MJ electricity/kg algae would be required for hexane solvent extraction of the algal biomass (Frank et al., 2011). Furthermore, 2 g of hexane are lost for each dry kg of algae (Lardon et al., 2009).

5.1.3. Alkali Refining Before the transesterification process, the free fatty acids must be removed from the algal lipids. It was assumed that this would be achieved by alkali refining whereby the addition of sodium hydroxide (NaOH) reacts with the fatty acids to form soaps that can be scrubbed from the lipid mixture. This results in a demand for 23 g NaOH/kg algal oil. 5.1.4. Summary of algal oil extraction Table 5-1 summarizes the material and energy inputs and upstream energy associated with the oil extraction process. The numbers 1, 2, and 3 represent scenarios commercially purchased fertilizers and CO2, fertilizers from CAS WW and commercially purchased CO2, and commercially purchased fertilizers and flue gas CO2, respectively. The letters L, B, and H represent low (worst case scenario), base-line, and high (best case scenario), respectively. To address uncertainty in the analysis, the electricity consumption and natural gas heat associated with oil extraction was adjusted by 20%, an increase for low and a decrease for high scenario.

27

Table 5-1 Oil extraction mass and energy inputs


Oil Extraction Phase Direct Energy Homogenization electricity (MJ/kg algae) Solvent recovery electricity (MJ/kg algae) Heat for solvent recovery (MJ/kg algae) Subtotal direct energy (MJ/kg algae) Indirect Energy Hexane Impact factor (MJ/kg hexane) Hexane (kg hexane/kg algae) Indirect energy hexane (MJ/kg algae) Sodium hydroxide impact factor (MJ/kg NaOH) Sodium hydroxide (kg NaOH/kg algae) Sodium hydroxide (MJ/kg algae) Impact factor electricity (MJ/MJ electricity) Indirect energy (MJ/kg algae) Impact factor NG (MJ/MJ NG) Indirect energy NG (MJ NG/kg algae) Transportation impact factor (MJ/kg chemical) Transportation of chemicals (MJ/kg algae) Subtotal indirect energy Total Oil Extraction Phase 1,2,3 L 0.804 0.394 1.24 2.43 54.0 0.0020 0.108 23.6 0.0039 0.0914 2.26 1.51 1.07 0.087 0.14 0.0008 1.7956 4.23 1,2,3 B 0.670 0.328 1.03 2.03 54.0 0.002 0.108 23.6 0.004 0.091 2.26 1.26 1.07 0.072 0.14 0.0008 1.53 3.56 1,2,3 H 0.536 0.262 0.824 1.62 54.0 0.002 0.108 23.6 0.0039 0.0914 2.26 1.01 1.07 0.058 0.14 0.0008 1.2638 2.89

5.2. Base catalyzed transesterification It was assumed base-catalyzed transesterification coverts algae oil into BD. In this case, the base sodium methoxide (CH3ONa) is the catalyst. Hydrochloric acid (HCl) is required to neutralize the sodium methoxide. Methanol is the alcohol for the transesterification reaction. Heat and electricity are required to heat wash water to remove impurities in the BD and recover residual methanol and glycerin. The demands for sodium methoxide, methanol, electricity, heat, and HCl were based on GREET soybean biodiesel production. The following chemicals were assumed to be required for the transesterification reaction to produce 1 kg of BD: 0.0125 kg sodium methoxide, 0.100 kg methanol, and 0.007 kg HCl. The natural gas heat and electricity required to produce 1 kg BD were assumed to be 0.784 MJ natural gas heat and 0.119 MJ electricity. 5.2.1. Summary of transesterification inputs

Table 5-2 summarizes the material and energy inputs and upstream energy associated with the transesterification process. The numbers 1, 2, and 3 represent scenarios commercially purchased fertilizers and CO2, fertilizers from CAS WW and commercially purchased CO2, and commercially purchased fertilizers and flue gas CO2, respectively. The letters L, B, and H represent low (worst case scenario), base-line, and high (best case scenario), respectively. To address uncertainty in the analysis, the electricity consumption and natural gas heat associated

28

with transesterification was adjusted by 20%, an increase for low and a decrease for high scenarios. Table 5-2 Transesterification mass and energy inputs
Transesterification Phase Direct Energy Heat for methanol recovery (MJ/kg algae) Heat for glycerin recovery (MJ/kg algae) Subtotal direct energy (MJ/kg algae) Indirect Energy Methanol (MJ methanol/kg methanol) Methanol (kg methanol/kg algae) Methanol (MJ methanol/kg algae) Sodium methoxide (MJ/kg CH3ONa) Sodium methoxide (kg CH3ONa/kg algae) Sodium methoxide (MJ/kg algae) Hydrochloric Acid Impact factor (MJ/kg HCl) Hydrochloric Acid (kg HCl/kg algae) Indirect energy Hydrochloric Acid (MJ HCl/kg algae) Natural Gas impact factor (MJ/MJ NG) Indirect energy natural gas (MJ/kg algae) Capital (plant and equipment) (MJ/kg algae) Waste and dismantling (MJ/kg algae) Transportation impact factor (MJ/kg chemical) Transportation of chemicals to plant (MJ/kg algae) Algae BD transportation and distribution (MJ/kg algae) Subtotal indirect energy (MJ/kg algae) Total Transesterification Phase 1,2,3 L 0.159 0.024 0.183 31.5 0.013 0.405 53.0 0.002 0.112 1.97 0.009 0.018 1.07 0.013 0.020 0.001 0.140 0.003 0.062 0.573 0.756 1,2,3 B 0.132 0.020 0.152 31.5 0.013 0.405 53.0 0.002 0.112 1.97 0.009 0.018 1.07 0.011 0.020 0.001 0.140 0.003 0.062 0.571 0.723 1,2,3 H 0.106 0.016 0.122 31.5 0.013 0.405 53.0 0.002 0.112 1.97 0.009 0.018 1.07 0.009 0.020 0.001 0.140 0.003 0.062 0.569 0.690

5.3. Biodiesel yield In this study it was assumed 5.93 kg algae would produce 1 kg of algal BD (Lardon et al., 2009). Based on a pond yield of 91,250 kg algae/ha, LHV of BD of 37.8 MJ/kg BD (Lardon et al., 2009), density of 0.87 kg/L (Sander and Murthy, 2010), this results in a biorefinery yield of 0.194 L BD/ha, and an energetic fuel yield of 525,885 MJ/ha. Table 5-3 summarizes the BD yields.

29

Table 5-3 Biodiesel yield


Biodiesel Yield Pond yield (kg algae/ha) Algae harvesting efficiency (kg algae out/kg algae in) Algae input (kg algae/kg oil) Biorefinery yield (L/kg algae) Volumetric fuel yield (L/ha) Energetic fuel yield (MJ/ha)

91250 0.90 5.34 0.194 17687 581661

5.4. AD of LERs The material and energy inputs associated with operating the AD are the same whether the substrate is LERs or whole algae. Details regarding the assumptions related to the operation of the AD and biogas upgrading are given in Section 6. In this case, biogas was treated as a coproduct, and the net energy yield is discussed in Section 9. 5.5. Feedstock inputs Following GREET methodology, the conversion of the algal biomass into algal oil, and the oil feedstock into algal BD is accounted for by adding the difference between the LHVs of the algal biomass and algal oil, and the difference between the LHVs of the algal oil and the BD. The calculation is as follows: ( ) ( )

30

6. Anaerobic Digestion (AD)

Recent LCAs demonstrate harvesting and oil extraction from algae represent a high energy debt and suggest is worth investigating an alternative fuel pathway to algal BD (Lardon et al,, 2009, Clarens, et al, 2010, Clarens et al., 2011; Collet et al., 2011). In this study the fuel pathway under consideration is AD of the whole algal biomass to produce biogas consisting of 96% CH4 for use in compressed natural gas (CNG) vehicles. AD is a well known technology widely used for the treatment of WW and more recently in livestock effluents. The idea of anaerobically digesting whole algae biomass was recently suggested by Sialve et al., 2009 and has been considered in recent LCAs (Collet, et al., 2011; Clarens et al., 2011). Because AD reduces the amount of energy for dewatering, input stream of 5% total solids versus 20% for BD, and eliminates the oil extraction and transesterification steps, it could reduce the total energy requirements. In addition, by re-circulating the liquid fraction of the digestates toward the algal ponds, a significant part of the fertilizers could be recycled. 6.1. Digester operating conditions In this study, AD of the whole biomass as well as the LERs was considered. The energy inputs associated with operating the digester are the same whether the digester substrate was whole algal biomass or LERs. AD of organic matter is a complex process and the yield and energy requirements are highly dependent on the substrate, operating environment, digester type, organic loading rate (OLR), and hydraulic retention time (HRT) to mention just a few. For this study it was assumed the continuously stirred reactor (CSTR) was operated at a minimum HRT of 46 days with an OLR of 1.4 g Chemical Oxygen Demand (COD)/L-d (Collet et al., 2011). This heat necessary for operating the anaerobic digestion plant, 2.45 MJ/kg algae, was provided by burning part of the biogas produced in a boiler. The energy to operate the digester was assumed to require 0.389 MJ electricity/kg total solids (TS) to mix the CSTR and 0.091 MJ electricity/kg TS to operate the centrifuge to dewater the solids from the liquid digestates. 6.1.1 Summary of AD mass and energy inputs Table 6-1 summarizes the material and energy inputs and upstream energy associated with the AD process. The numbers 1, 2, and 3 represent scenarios commercially purchased fertilizers and CO2, fertilizers from CAS WW and commercially purchased CO2, and commercially purchased fertilizers and flue gas CO2, respectively. The letters L, B, and H represent low (worst case scenario), base-line, and high (best case scenario), respectively. To address uncertainty in the analysis, the electricity consumption associated with AD was adjusted by 20%, an increase for low and a decrease for high scenarios. Table 6-1 AD mass and energy inputs
Anaerobic Digestion Direct Energy Electricity mixing of digesters (MJ/kg TS) Electricity centrifuge digestates (MJ/kg TS) 1,2,3 L 0.467 0.1089 1,2,3 B 0.389 0.091 1,2,3 H 0.311 0.0726

31

Subtotal direct energy (MJ/kg algae) Indirect Energy Indirect Electricity Impact Factor (MJ/MJ electricity) Indirect energy electricity (MJ/kg TS) Capital (plant and equipment) (MJ/kg TS) Waste and dismantling (MJ/kg TS) Transportation of chemicals (MJ/kg chemical) Solid digestates (kg organics/kg TS) Transportation of solid digestates (MJ/kg TS) Subtotal indirect energy (MJ/kg TS) Total Anaerobic Digestion Phase

0.575 2.26 0.725 0.00008 0.140 0.063 0.009 0.734 1.309

0.480 2.260 0.604 0.00008 0.140 0.063 0.009 0.613 1.093

0.384 2.26 0.483 0.00008 0.140 0.063 0.009 0.492 0.876

6.2. Biogas upgrading To achieve a biogas composition of 96% CH4 or greater, the impurities in the biogas such as CO2 and hydrogen sulphide (H2S) and must be removed. There are a number of technologies for biogas upgrading including pressure swing adsorption, membrane separation, cryogenic distillation, and absorption with water scrubbing. Water scrubbing is lower in energy intensity than the other methods mentioned, does not require chemicals, and if the water is not regenerated, the CO2 removed from the biogas can be re-circulated to the algal pond reducing the demand CO2 from off-site sources. Water scrubbing is an absorption process that works because H2S and CO2 are more soluble in water than CH4. The absorption process is purely physical. In this case, the biogas is pressurized and fed to the bottom of a packed column where water is fed on the top and so the absorption process is operated counter-currently. The resulting energy demand for the water scrubbing absorption process was assumed to be 0.299 MJ/kg TS. 6.2.1 Summary of biogas upgrading mass and energy inputs Table 6-2 summarizes the material and energy inputs and upstream energy associated with the biogas upgrading process. The numbers 1, 2, and 3 represent scenarios commercially purchased fertilizers and CO2, fertilizers from CAS WW and commercially purchased CO2, and commercially purchased fertilizers and flue gas CO2, respectively. The letters L, B, and H represent low (worst case scenario), base-line, and high (best case scenario), respectively. To address uncertainty in the analysis, the electricity and natural gas associated with biogas upgrading was adjusted by 20%, an increase for low and a decrease for high scenarios.

Table 6-2 Biogas upgrading mass and energy inputs


Gas upgrading, transportation and compression at station Direct Energy Electricity pressurized water srubbing (MJ/kg TS) Electricity transportation via pipeline (MJ/kg TS) Electricity compression at station (MJ/kg TS) Natural gas transportation via pipeline (MJ/kg TS) Natural gas compression at station (MJ/kg TS) Subtotal direct energy (MJ/kg TS) 1,2,3 L 0.359 0.0002 0.253 0.004 0.663 1.279 1,2,3 B 0.299 0.0002 0.211 0.003 0.553 1.066 1,2,3 H 0.239 0.0002 0.169 0.002 0.442 0.852

32

Indirect Energy Electricity impact factor (MJ/MJ electricity) Indirect energy electricity (MJ/kg TS) Natural gas impact factor (MJ/MJ NG) Indirect energy NG (MJ/kg TS) Capital (plant and equipment) (MJ/kg TS) Waste and dismantling (MJ/kg TS) Subtotal indirect energy (MJ/kg TS) Total upgrading phase

2.260 0.771 1.070 0.047 0.000 0.818 2.096

2.260 0.642 1.070 0.039 0.000 0.681 1.747

2.260 0.514 1.070 0.031 0.000 0.545 1.398

6.3. Biogas Yield Decomposition of organic matter in anaerobic processes is extremely complex, involving the interaction of anaerobic and facultative bacteria. The process can be characterized by three distinct and sequential stages comprising (1) hydrolysis, (2) organic acid formation, and (3) methane fermentation. In the first stage, insoluble, complex organics such as lipids, cellulose, and protein are hydrolyzed by extracellular enzymes to smaller molecules accessible by microbes. In the second stage, the smaller organic molecules are subsequently decomposed by smaller fast-growing, acedogenic bacteria to alcohols and short-chain fatty acids such as acetic, propionic, and butyric, acids. And, in the third stage, most of the organic acids and all of the H2 are metabolized by methanogenic bacteria, with the end result being production of a mixture of biogas with approximately 55% to 70% CH4 and 30% to 40% CO2. Typically the yield of methane is reported as liters (L) of methane per g of volatile solids (VS). When the composition of the organic matter is known, it is possible to evaluate the theoretical methane yields that can be expected from the anaerobic digestion (Sialve et al., 2009). The yield can be estimated in Eq. 1 (Sialve et al., 2009): ( ) ( ) ( ) ( )

In this equation, the organic matter is stoichiometrically converted to methane, carbon dioxide and ammonia. The specific methane yield expressed in liters of CH4 per gram of volatile solids (VS) can thus be calculated as (Sialve et al., 2009): ( ) where Vm is the molar volume of methane. The ratio rG of methane to carbon dioxide can be computed from (Sialve et al., 2009) Eq. 3 and 4: ( )

33

where n is the average carbon oxidation state in the substrate (Sialve et al., 2009) ( ) In this study, the yield was calculated using Equations 1 and 2 and the biomass fractions of Table 1. The resulting estimated theoretical yield was found to be 0.520 L CH4/g VS. However, there is little experimental data for the AD of algae and as a result there is a high degree of uncertainty in the expected yield of CH4, the amount of volatile solids in the total solids (algae), and the resulting ratio of CH4 to CO2. However, in experiments that have anaerobically digested C. vulgaris, the yields range between 0.15 0.35 L CH4/g VS, 0.85-0.95 g VS/g TS, and CH4% between 62 -76 (Ras et al., 2011; Collet et al., 2011; Sialve et al., 2009). In Figure 5, experimental and theoretical biomethane yields of lipid extracted and whole algae biomass are plotted against hydraulic retention time (HRT). The yield reported by Ehimen et al. (2011) includes the codigestion of glycerin in the reported yields.

Figure 6-1 Plot of methane yield of C. vulgaris versus digestion time reported in the literature. The symbols a, b, and * denote experimental result, theoretical result, and lipid extracted biomass, respectively. To reflect the uncertainty in this analysis the yield was assumed be an average of the maximum theoretical yield and half of the theoretical yield, resulting in a value of 0.390 L CH4/kg VS. Similarly, the ratio VS/TS and %CH4, values of 0.80 and 70% were assumed, respectively. This results in a yield of 244 L/TS. The percent CH4 was assumed 70%. The resulting CH4 yield for anaerobically digested whole algae is 312 L CH4/kg algae. After deducting the biogas required to heat the digester results in a balance of 244 L CH4/TS. The LHV of biomethane (96% CH4) is

34

0.037 MJ/L. The volumetric fuel yield is 22,237842 L/ha. The energetic fuel yield is 825,150 MJ/ha. To conduct sensitivity analysis, high and low values were also calculated, with 400 L/kg TS and 114 L/kg TS, 0.9 and 0.7, and 75% and 65%, digester yield, VS/TS, % CH4, respectively. The biogas yield is summarized in Table 6-3. Table 6-3 Summary of whole algae biogas yields
Whole Algae Biomass Biogas Yields Pond yield (kg algae/ha) Digester yield (L/kg algae) Volumetric fuel yield (L/ha) Energetic fuel yield (MJ/ha) Low 91250 114 10,375,342 384,984 Base 91250 244 22,237,842 825,150 High 91250 400 36,472,842 1,353,349

7. Hydrogasification

35

8. Transportation of fuel production inputs and outputs

8.1. Transportation of fuel production inputs

A medium duty diesel truck was assumed to provide transportation of consumables such as chemicals for fuel transformation requiring 0.140 MJ/kg unit based on a round trip distance of 100 miles. 8.2. Transportation and distribution of fuel products

The transportation and distribution of BD is assumed to be a combination of barge, rail, and heavy duty trucks requiring diesel fuel. The transportation and distribution of CNG is assumed to be via pipeline requiring natural gas and electricity. 8.2.1 Transportation and distribution of biodiesel It was assumed algal biodiesel is transported and distributed in the same manner as soybean biodiesel, by barge, rail, and heavy duty truck. All totaled, the transportation and distribution energy is 0.062 MJ/kg algae. 8.2.2. Transportation and distribution of CNG

It was assumed transportation and distribution of CNG by pipeline requires 0.427 MJ/ton-mile. Based on a transportation distance of 50 miles, the resulting demand was found to be 0.021 MJ/kg biogas. However, 6% of this energy is provided by electricity and the other 94% is provided by natural gas. In addition, energy is required to compress the biomethane. In this study, it was assumed 1.33 MJ electricity/kg biogas and 3.49 MJ natural gas/kg biogas is required for compression at the station. 8.2.3 Transportation and distribution of syngas

36

9. Coproducts

Coproduct credits in this analysis are treated with a hybrid of displacement and energy allocation for BD and displacement for the case of biogas from whole algal biomass. There are two coproduct energy allocations in the case of BD, oil extraction and biodiesel. Net energy credit is only given to coproducts not used in the fuel production process. For example, when the process uses biogas produced on-site, recovered nutrients or CO2 the net consumption of these nutrients and fuels by the algae pathway system is reduced accordingly. However, the biogas combusted in a boiler to heat the digester incurred a GHG penalty. 9.1. Energy allocation

In this study, biogas from LERs was allocated to algal oil feedstock and glycerin was allocated to BD. The methods for calculating and applying these allocations are discussed in the following sections. The energy allocation method is the ratio of coproduct in all products by energy. The main fuel product and coproduct carry energy and emissions burden based on their ratios in the total products. 9.1.1. Energy allocation to algal oil

For the baseline scenario, the energy allocated to algal oil is calculated as the ratio of the LHV of the algal oil to the sum of the LHV of the algal oil and the LHV of the biogas per kilogram of algal oil:

The algal oil energy allocation factor was applied to all net energy and GHGs associated with the processes leading up to and including oil extraction, such as the net inputs for cultivation, harvesting, oil extraction, anaerobic digestion, biogas upgrading, and coproducts (biogas, fertilizer, and compost of solid digestates). 9.1.2. Energy Allocation to Biodiesel Because of the relatively low market value of glycerin, in this study, the energy allocation method is used to apply energy credit. For every kg of BD produced, a 0.213 kg of glycerin is produced (Wang et al., 2008). The LHV of glycerin was assumed to be 18.56 MJ/kg glycerin (Wang et al., 2008). Thus, the net coproduct energy applied is 3.13 MJ/kg BD. The energy allocation is calculated as the ratio of the LHV of biodiesel to the sum of the LHV of biodiesel and the LHV of glycerin:

37

The energy allocation factor was applied to all net energy and GHGs associated with the transesterification process, coproduct glycerin, and the LHV of BD. 9.2. Aquaculture feed The displacement method was based on the displacement of microalgae used as fish and rotifer feed in aquaculture by the microalgal extract produced in the microalgae-to-fuel process. An averaged value for the energy dedicated to cultivation of microalgae for fish feed in aquaculture of 7.6 MJ kg-1 of dry algae (Batan et al., 2010). Aquaculture feed was not a coproduct considered in this analysis but is included for comparison purposes. 9.3. Biogas from lipid extracted algal residuals An overview of the biogas production plant for the lipid extracted algae residuals is provided in Figure 6. The coproducts associated with AD of lipid extracted algae residuals are compost (comprised of the carbon and portion of the N and P unavailable to plants), recycled N and P, and biogas.

Figure 9-1 Overview of algae biogas plant for lipid extracted residuals

38

9.3.1. Biogas from lipid extracted algae residuals

The process of AD of lipid extracted algae residuals was assumed to be the same as was reported in 3.4.2. Thus the energy required for operating and mixing the digester and for upgrading the biogas is the same as that reported in section 3.4.2. Similar to whole algae biomass, the theoretical CH4 yield was calculated using Equations 1 and 2 and Table 1. However, the assumed yield of biogas and biogas constituents are different because the biomass fractions are assumed to be 62% carbohydrates and 35% proteins instead of 18% lipids, 29% protein, and 51% carbohydrates. The resulting estimated theoretical yield was found to be 0.415 L CH4/g VS. To reflect the uncertainty in this analysis the yield was assumed be an average of the maximum theoretical yield and half of the theoretical yield, resulting in a value of 0.311 L CH4/g VS. The ratio VS/TS and %CH4, values of 0.90 and 51% were assumed, respectively. The resulting CH4 yield for an anaerobically digested lipid extracted algae is 280 L CH4/kg algae. After deducting the biogas required to heat the digester results in a balance of 212 L CH4/kg algae. This results in a coproduct credit of 26.1 MJ/kg BD for lipid extracted algae residuals. 9.3.2. Compost and fertilizers from solid digestates In this analysis compost and fertilizers were given net energy and GHG emissions credit with the displacement method. The energy and GHGs associated with the production of compost, and fertilizers N and P were displaced by the production of compost and fertilizers from the solid digester effluent. For this study, it was assumed that all carbon not in the CH4 or CO2 was a solid digestate, and that 20% of the fertilizers end up in the solid digestates. 20% of the N and P used to fertilize the pond do not mineralize and end up in the solid digestate, of which 60% is unavailable to plants. This portion unavailable to plants was given credit as compost. The remaining 40% of the solid N and P was assumed to be plant available and thus was given credit as N and P fertilizer. Therefore, the resulting amount of displaced compost and fertilizers was assumed to be 58 g compost/kg algae, 3.68 g N/kg algae, and 0.8 g P/kg algae.

9.4. WWT offsets

The use of WWT effluents offsets demand for fertilizer production and the associated material and energy inputs; and freshwater usage since WWT effluent can be utilized as the algae growth medium. In this study, we assume CAS WW completely supplants the need for freshwater for the algae cultivation ponds. The N content was assumed to be 25 mg/L and the P content 7 mg/L. N removal was assumed to proceed via nitrification and subsequent de-nitrification with addition of methanol as external carbon source. A ratio of 3.4 kg methanol per 1 kg N eliminated was used to compute the mass of methanol required to reduce the WW initial nitrogen concentration

39

down to 3.0 mg/L (Clarens, et al., 2010). Electricity consumption for aeration during nitrification was also assessed, using a value of 10 MJ per 1 kg N eliminated. P removal proceeds via chemical precipitation with ferrous sulfate. A ratio of 1.8 kg Fe per 1 kg P removed was used to compute the amount of ferrous sulfate required to reduce the wastewater P concentration down to 0.1 mg/L (Clarens, et al., 2010). Energy consumption for transportation of the resulting precipitant sludge was estimated to be 2 MJ per kg P eliminated. As a general rule of thumb it was assumed one person generates 100 gallons of WW per day (Neandross, 2010) and based on the population of California, 37,253,956 in 2010, there are approximately 3.7 billion gallons of WW that could be repurposed for algae cultivation. (U.S. Census Bureau, 2011).

40

10. Capital (fuel production infrastructure) In this analysis, the infrastructure and machines were assumed to have 30 and 10 year life-spans, respectively. The inventory of materials for the infrastructure and machines was found in the literature (Lardon et al., 2009). The capital for BD and biogas plants were found to be very small compared to the other categories in the analysis. The pond area was assumed to be 1000 square meters. The annual growth rate was assumed to be 9.125 kg algae m2-y. The pond materials and machinery are allocated to 1 kg of algae by Equation 5:
(
( )

) )

(5)

10.1. Cultivation Capital An inventory of the materials for the pond infrastructure and machines is provided in Table 10-1. The PVC liner was assumed to have a density of 1.3 kg/m3 and thickness of 1 mm. The capital allocated to 1kg of algae was calculated using Equation 5 and the capital inventory in Table 101. Table 10-1 Cultivation capital and infrastructure allocated to 1 kg algae Capital Inventory Foundation, walls 450 Pond liner 1200 Pipes 687 Paddlewheel, steel 50 Paddlewheel, glass fiber 256 Pump, steel 20 Capital allocated to 1kg algae 1.64384 kg concrete/kg algae 0.00002 kg PVC liner/kg algae 0.00251 kg PVC pipes/kg algae 0.00022 kg steel/kg algae 0.00281 kg glass fiber/kg algae 0.00022 kg steel/kg algae

t concrete m2 PVC kg PVC kg steel kg glass fiber kg steel

10.2. Harvesting Capital An inventory of the materials for the pond infrastructure and machines is provided in Table 10-2. The capital allocated to 1kg of algae was calculated using Equation 5 and the capital inventory in Table 10-2. Table 10-2 Harvesting capital and infrastructure allocated to 1 kg algae Capital Inventory Foundation, walls for settling ponds Centrifuge Dryer Capital allocated to 1kg algae 1.256621 kg concrete/kg algae 0.023014 kg steel/kg algae 0.043836 kg steel/kg algae

344 t concrete 2100 kg steel 4000 kg steel

41

10.3. Biodiesel plant capital The biodiesel plant was assumed to be similar to a soybean biodiesel plant. The LCIs for the biodiesel plant were obtained from the Simapro data base. The oil extraction capital is based on the inventory for a soy oil mill, where the infrastructure includes the processing of the oil and the cake. The transesterification capital is based on the inventory for a vegetable oil transesterification plant. The resulting net fossil energy and GHG impact s were found to be 0.020 MJ/kg algae and 10.4. Biogas plant capital The plant digester is assumed to be constructed of 113,400 kg of steel. With the assumption that cold rolled steel was used in the construction results in 7.65 x 10-5 MJ of fossil energy. The energy demand associated with the biogas upgrading plant was even smaller. Therefore, for the purposes of this analysis, the energy and GHG emissions associated with the biogas plant capital were assumed to be negligible. 10.5. Hydrogasification plant capital

42

11. Waste and dismantling (end of infrastructure life cycle) In this analysis, the infrastructure and machines were assumed to have 30 and 10 year life-spans, respectively. At the end of the useful life of the infrastructure or machine it was assumed energy would be required to dismantle and dispose of the waste. The inventory of materials for disposal was found in the literature (Lardon et al., 2009). The pond area was assumed to be 1000 square meters. The annual growth rate was assumed to be 9.125 kg algae m2-y. The pond materials and machinery are allocated to 1 kg of algae by Equation 5:
(
( )

) )

(5)

11.1. Cultivation waste and dismantling

The disposal of capital allocated to 1kg of algae was calculated using Equation 5 and the pond waste and dismantling inventory in Table 11-1.

Table 11-1 Cultivation waste and dismantling allocated to 1 kg algae Inventory Item 450 t concrete 1200 m2 PVC 687 kg PVC 50 kg steel 256 kg glass fibre 20 kg steel 2100 m3 y-1 Mass allocated to 1 kg algae 1.643836 kg concrete/kg algae recycle - none recycle - none recycle - none recycle - none recycle - none 0.230137 m3 wastewater/kg algae

Foundation, walls Pond liner Pipes Paddlewheel Fiber glass Pump Water flush

11.2. Harvesting waste and dismantling

The disposal of capital allocated to 1kg of algae was calculated using Equation 5 and the harvesting waste and dismantling inventory in Table 11-2. Table 11-2 Harvesting waste and dismantling allocated to 1 kg algae Inventory Item Foundation, walls 344 t concrete Steel, rotary press 1200 m2 PVC Mass allocated to 1 kg algae 1.256621 kg concrete/kg algae recycle no impact

43

Steel, dryer

687 kg PVC

recycle no impact

11.3. BD plant waste and dismantling The waste and dismantling of the BD plant capital was considered negligible. However, the treatment of WW was considered and based on values in Simapro for WW disposal associated with vegetable oil transesterification. The value was found to be 0.00539 MJ/kg BD. 11.4. Biogas plant waste and dismantling The materials used in the construction of the biogas plant were assumed to be made of steel and recycled. Thus no net energy was associated with the waste and dismantling of the biogas plant. 11.5. Hydrogasification plant waste and dismantling

44

45

12.0. Results and discussion

Three cases were considered for this analysis: (1) the baseline scenario where it was assumed nutrients were provided with commercially procured fertilizers and CO2, (2) a scenario where WW provides the fertilizers for the pond but the CO2 was commercially procured, and (3) a scenario where flue-gas was assumed to provide the CO2 but fertilizers were commercially procured. The net energy ratio (NER), the ratio of the input energy to the output energy, and greenhouse gas emissions (kg CO2e/MJ fuel) was calculated for each scenario described above. In addition, a low value, base-line, and high value were considered where the low (L) value represents the high energy and low biogas yield, and the high (H) value represents the low energy and high biogas yield. The results of the transesterification and AD pathway are discussed in section 12.1 and the AD of whole algae biomass pathway is discussed in section 12.2.

12.1. Transesterification and AD pathway The NER was calculated for case 1 and found to be 0.22, 0.32, and 0.43 for low, base, and high scenarios, respectively. Similarly, for case 2, the NER was found to be 0.26, 0.45, and 0.63 for low, base and high scenarios, respectively. And, case 3 the NER was found to be 0.33, 0.46, and 0.59 for low, base, and high scenarios, respectively. The results for each case are summarized in Table 12-1. For comparison, the net energy ratios of soybean BD and conventional diesel were also calculated and were found to be 0.74 and 5.56, respectively. Table 12-1 Summary of energy allocated to 1 L of fuel for transesterification pathway

Totals Total cultivation phase (MJ/L) Total harvesting phase (MJ/L) Total oil extraction phase (MJ/L) Feedstock Input (MJ/L) Total fuel production phase (MJ/L) Total anaerobic digestion phase (MJ/L) Total gas upgrading phase (MJ/L)

1L 36.5 78.4 13.9 72.3

1B 26.2 51.1 9.0 63.8

1H 19.8 34.1 5.9 58.5

2L 37.5 78.4 13.9 72.3

2B 24.9 51.1 9.0 63.8

2H 17.8 34.1 5.9 58.5

3L 19.9 22.3 14.3 72.3

3B 13.6 15.2 9.0 63.8

3H 9.6 10.9 5.9 58.5

Soybean BD

Conventional Diesel

10.7

15.3 50.0

1.8

3.7

3.5

3.3

3.7

3.5

3.3

3.4

3.5

3.1

6.0

4.7

4.3

2.8

1.8

4.3

2.8

1.8

4.3

2.8

1.8

6.9

4.4

2.9

6.9

4.4

2.9

6.9

4.4

2.9

46

Input energy (MJ/L) Reported HV of fuel (MJ/L) Coproduct credits (MJ/L) Coproducts as % of total energy Output energy (MJ/L) Net energy value, NEV (MJ/L) Net Energy Ratio

216.0 -32.9 -18.2 8.4 47.9 168.1 0.22

160.7 -32.9 -22.6 14.0 52.2 108.5 0.32

126.4 -32.9 -25.3 20.0 55.0 -71.4 0.43

217.0 -32.9 -27.5 12.7 57.2 159.8 0.26

159.4 -32.9 -41.5 26.0 71.2 -88.2 0.45

124.3 -32.9 -48.7 39.2 78.4 -45.9 0.63

143.5 -32.9 -18.2 12.7 47.9 -95.6 0.33

112.3 -32.9 -22.5 20.1 52.2 -60.1 0.46

92.6 32.9 25.3 27.3 54.9 37.7 0.59

82.0 -33.3 -27.7

6.4 -35.8

-61.0

-35.8

0.74

5.6

350.0 300.0 250.0 200.0 MJ/L 150.0 100.0 50.0 0.0 -50.0 -100.0 Reported HV of fuel (MJ/L) Total gas upgrading phase (MJ/L) Total anaerobic digestion phase (MJ/L) Total fuel production phase (MJ/L) Feedstock Input (MJ/L) Coproduct credits (MJ/L)

Soybean BD

Conventional Diesel

Total oil extraction phase (MJ/L) Total harvesting phase (MJ/L) Total cultivation phase (MJ/L)

1B

2B

1H

2H

3B

Figure 12-1 Net energy summary for transesterification + anaerobic digestion The GHGs were also calculated for each case. For case 1, the GHGs were found to be 0.18, 0.11, and 0.07 kg CO2e/MJ for cases low, base-line, and high scenarios, respectively. For case 2, the GHGs were found to be 0.17, 0.09, and 0.05 kg CO2e/MJ for cases low, base-line, and high scenarios, respectively. For case 3, the GHGs were found to be 0.01, 0.00, and -.01 kg CO2e/MJ for cases low, base-line, and high scenarios, respectively. The results for each case are summarized in Table 12-2. For comparison, the GHGs of soybean BD and conventional diesel and were found to be -0.072, and 0.017 kg CO2e/MJ, respectively. Case three, CO2 provided by flue gas and commercially procured fertilizers, meets the requirement of the EPA RFS with

3H

2L

3L

47

reductions greater than 50% compared to conventional diesel. Graphs of the net energy summaries are presented in Figure 12-1 and summaries of GHG emissions in Figure 12-2.

Table 12-2 Summary of GHGs allocated to 1 MJ of fuel for transesterification pathway


Totals Total cultivation phase (kg CO2e/L) Total harvesting phase (kg CO2e/L) Total oil extraction phase (kg CO2e/L) 0.8 Total transesterification phase (kg CO2e/L) Total AD phase (kg CO2e/L BD) Total biogas upgrade (kg CO2e/L BD) Biomass uptake (kg CO2e/L) 5.9 Coproduct credits (kg CO2e/L) 0.1 Net GHG emissions (kg CO2e/L) HV of Fuel (MJ/L) 32.9 kg CO2e/ MJ Fuel 0.18 0.11 0.07 0.17 0.09 0.05 0.01 0.00 (0.01) (0.07) 0.02 32.9 32.9 32.9 32.9 32.9 32.9 32.9 32.9 0.1 0.1 0.3 0.6 0.5 0.2 0.1 0.0 4.5 3.7 5.9 4.5 3.7 5.9 4.5 3.7 0.5 0.3 0.8 0.5 0.3 0.8 0.5 0.3 1L
1B 1H 2L 2B 2H 3L 3B 3H Soybean Biodiesel Conventional Diesel

3.4

2.5

2.0

3.6

2.5

1.8

1.8

1.3

0.9

5.7

3.8

2.6

5.7

3.8

2.6

2.2

1.6

1.2

0.1 1.2

0.1 0.9

0.1 0.7

0.1 1.2

0.1 0.9

0.1 0.7

0.2 1.2

0.2 0.9

0.2 0.7

0.4

0.3

0.2

0.4

0.3

0.2

0.4

0.3

0.2

5.6

3.5

2.2

5.6

2.9

1.6

0.4

0.1

(0.2)

48

0.20

0.15

0.10 kg CO2e/MJ

0.05

(0.05)

(0.10) Soybean Biodiesel 1L 2L 3L Conventional Diesel 1B 2B 1H 2H 3B 3H

Figure 12-2 Net GHGs for transesterification + anaerobic digestion

12.2. Transesterification and hydrogasification pathway 12.3. AD pathway

The NER was calculated for case 1 and found to be 0.13, 0.33, and 0.63 for low, base, and high scenarios, respectively. Similarly, for case 2, the NER was found to be 0.21, 0.59, and 1.26 for low, base and high scenarios, respectively. And, case 3 the NER was found to be 0.16, 0.43, and 1.05 for low, base, and high scenarios, respectively. The results for each case are summarized in Table 12-3. The NERs for the AD pathway perform better than transesterification and AD of LERs with a value greater than 1 for the high scenario for cases 1 and 2.

49

Table 12-3 Summary of energy allocated to 1 L of fuel for AD pathway

Totals Total cultivation phase (MJ/L) Total harvesting phase (MJ/L) Biomass input (MJ/L) Total digester phase (MJ/L) Total gas purification, transportation & compression phase (MJ/L) Input energy LHV of biogas (MJ/L) Coproduct credits (MJ/L) Coproducts as % of total energy Output energy (MJ/L) Net energy value, NEV (MJ/L) Net Energy Ratio 0.35

1L 0.14 0.03 0.13 0.01 0.02 0.33 0.04 0.00 2.07 0.04 0.28 0.13

1B 0.06 0.01 0.04 0.00 0.00 0.12 0.04 0.00 1.71 0.04 0.08 0.33

1H 0.04 0.00 0.01 0.00 0.00 0.06 0.04 0.00 0.90 0.04 0.02 0.66

2L 0.13 0.03 0.13 0.01 0.02 0.31 -0.04 -0.03 16.10 0.07 -0.25 0.21

2B 0.05 0.01 0.04 0.00 0.01 0.12 -0.04 -0.03 42.86 0.07 -0.05 0.59

2H 0.03 0.00 0.01 0.00 0.00 0.05 -0.04 -0.03 70.07 0.07 0.01 1.26

3L 0.06 0.03 0.13 0.01 0.02 0.25 0.04 0.00 3.42 0.04 0.21 0.16

3B 0.03 0.01 0.04 0.00 0.01 0.09 0.04 0.00 2.69 0.04 0.05 0.43

3H 0.01 0.00 0.01 0.00 0.00 0.04 0.04 0.00 1.69 0.04 0.00 1.05

Biomass input (MJ/L) 0.30 Coproduct credits (MJ/L) 0.25 0.20 0.15 Total gas purification, transportation & compression phase (MJ/L) 0.10 Total digester phase (MJ/L) 0.05 0.00 -0.05 -0.10 1L 1B 1H 2L 2B 2H 3L 3B 3H Total harvesting phase (MJ/L) LHV of biogas (MJ/L)

MJ/L fuel

Total cultivation phase (MJ/L)

Figure 12-3 Net energy summary for anaerobic digestion of whole algae biomass

50

The GHGs were also calculated for each case. For case 1, the GHGs were found to be 0.008, 0.018, and -0.004 kg CO2e/MJ for cases low, base-line, and high scenarios, respectively. For case 2, the GHGs were found to be -0.027, -0.031, and -0.036 kg CO2e/MJ for cases low, baseline, and high scenarios, respectively. For case 3, the GHGs were found to be -0.153, -0.078, and -.052 kg CO2e/MJ for cases low, base-line, and high scenarios, respectively. The results for each case are summarized in Table 12-4. Case three, CO2 provided by flue gas and commercially procured fertilizers, meets the requirement of the EPA RFS. Graphs of the net energy summaries are presented in Figure 12-4. In all of the scenarios, low, baseline and high for each case, the GHGs were lower than that of conventional diesel. Table 12-4 Summary of GHGs allocated to 1 MJ of fuel for AD pathway
Totals Total cultivation phase (kg CO2e/L) Total harvesting phase (kg CO2e/L) Total AD phase (kg CO2e/L BD) Total biogas upgrade (kg CO2e/L BD) Biomass uptake (kg CO2e/L) Coproduct credits (kg CO2e/L) Net GHG emissions (kg CO2e/L) HV of Fuel (MJ/L) kg CO2e/ MJ Fuel 0.008 (0.018) (0.004) (0.027) (0.031) (0.036) (0.153) (0.078) (0.052) 1L 0.011 0.001 0.002 0.001 0.003 0.000 0.001 0.007 0.016 0.000 0.001 0.000 0.037 (0.001) 0.037 (0.000) 0.037 (0.001) 0.037 (0.001) 0.037 (0.001) 0.037 (0.006) 0.037 (0.003) 0.037 0.000 0.001 0.001 0.001 0.001 0.000 0.000 (0.002) 0.037 0.005 0.016 0.007 0.005 0.016 0.007 0.005 0.000 0.001 0.000 0.000 0.001 0.000 0.000 0.001 0.003 0.001 0.001 0.003 0.001 0.001 0.000 0.002 0.001 0.000 0.002 0.001 0.000
1B 1C 2L 2B 2H 3L 3B 3C

0.005

0.003

0.010

0.004

0.003

0.005

0.002 0.001

0.0500 kg CO2e/MJ -

(0.0500) (0.1000) (0.1500) (0.2000) 1L 1B 1C 2L 2B 2H 3L 3B 3C

Figure 12-4 Net GHGs for anaerobic digestion of whole algae biomass

51

12.4. Hydrogasification pathway

52

13. Conclusion

53

14. References Batan, L., Quinn, J., Willson, B., and Bradley, T. (2010) Environ Sci Technol 44: 7495-7980. Benemann, J.; Oswald, W. Systems and Economic Analysis of Microalgae Ponds for Conversion of CO2 to Biomass Final Report; Department of Energy: Pittsburgh, PA, 1996; p 201. CEC. Consumer Energy Center. Why should care about the vehicles we buy and drive. Web accessed November 7, 2011. http://www.consumerenergycenter.org/transportation/why.html. Schremp, G., Weng-Gutierrez, M., Eggers, R., Bahreinian, A., Gage, J., van der Werf, Y., Zipay, G., McBride, B., Lawson, L., and Yowell, G. (2011). Transportation Energy Forecasts and Analyses for the 2011 Integrated Energy Policy Report. CEC Draft Staff Report No. CEC6002011007SD. Web accessed November 7, 2011. http://www.energy.ca.gov/2011publications/CEC-600-2011-007/CEC-600-2011-007-SD.pdf Chaumont D. Biotechnology of algal biomass production: a review of systems for outdoor mass culture. Journal of Applied Phycology. 1993; 5:593-604. Clarens, A., Resurreccion E., Shite M., Colosi L., (2010). Environmental life cycle comparison of algae to other bioenergy feedstocks. Environ Sci Technol 44:1813-1819. Clarens, A., Nassau, H., Resurreccion E., White, M., Colosi L. (2011). Environmental Impacts of Algae-Derived Biodiesel and Bioelectricity for Transportation. Environ Sci Technol DOI: 10.1021/es200760n. Craggs, R.J.; Adey, W.H.; Jenson, K.R.; St. John, M.S.; Green, F.B. and Oswald, W.J. (1996). Phosphorus removal from wastewater using an algal turf scrubber. Water Sci. Technol. 33 (7), 191198. 90 A.C. Wilkie, W.W. Mulbry / Bioresource Technology 84 (2002) 8191 Dragone G, Fernandes B, Vicente A, Teixeira JA. Third generation biofuels from microalgae. In: Vilas AM, editor. Current research, technology and education topics in applied microbiology and microbial biotechnology. Badajoz: Formatex Research Center; 2010. p. 135566. Ehimen EA, Sun ZF, Carrington CG, Birch EJ, Eaton-Rye JJ. (2011). Anaerobic digestion of microalgae residues resulting from the biodiesel production process. Appl Energy 88(10):345463. Elgowainy, A. GREET 1: A Fuel-Cycle Model for Alternative Fuels and Light-Duty Vehicles. GREET Training WorkshopArgonne National Laboratory December 7-8, 2011. Frank et al., (2011). Life-Cycle Assessment of Algal Lipid Fuels with the GREET model. Argonne National Laboratory Report No. ANL/ESD/11-5.

54

Huo, et al. (2008). Life-Cycle Assessment of Energy and Greenhouse Gas Effects of SoybeanDerived Biodiesel and Renewable Fuels. Argonne National Laboratory Report No. ANL/ESD/08-2. IPCC, Special Report on Carbon Dioxide Capture and Storage (2005), available at: http://www.ipcc.ch/pdf/special-reports/srccs/srccs_summaryforpolicymakers.pdf. Janssen M, Tramper J, Mur LR, Wijffels RH. (2003). Enclosed outdoor photobioreactors: Light regime, photosynthetic efficiency, scaleup, and future prospects. Biotechnology and Bioengineering. 81:193-210. Kadam, K.L. (1997). Power plant flue gas as a source of CO2 for microalgae cultivation: economic impact of different process options. Energy Conv Manag 38: S505-S510.-510-29417. Kadam, K. (2001) Microalgae production from power plant flue gas: environmental implications on a life cycle bases. Tech. Rep. NREL/TP Lardon L., Helias, A., Sialve, B., Steyer J., Bernard, O. (2009) Life cycle assessment of biodiesel production from microalgae. Environ Sci Technol 43:6475-6481. Mata, T.; Martins, A.; Caetano, N. (2010) Microalgae for Biodiesel Production and Other Applications: A Review. Renewable Sustainable Energy Rev., 14 (1), 217232. Munson, B.R., Young, D.F., Okiishi, T.H., Huebsch, W.W. (2009). Fundamentals of Fluid Mechanics. 6th edition; John Wiley & sons, Inc. Oswald W.J. and C.G. Golueke. (1968). Large scale production of algae. In: Matele s RI, Tanebaum SR, editors. Single Cell Protein. Cambridge: MIT Press. p 271-305. Oswald W.J. (1988). Large-scale algal culture systems. In: Borowtizka, M, and Borowitzka, L. editors. Micro-algal Biotechnology. Cambridge University Press. p. 357-394. Pulz O. (2001).Photobioreactors: production systems for phototrophic microorganisms. Applied Microbiology and Biotechnology. 57:287-293. Ras, M., et al., (2011). Experimental Study on a Coupled Process of Production and Anaerobic Digestion of Chlorella vulgaris. Bioresource Technology 102(1):200206. Sander, K. and G. Murthy. (2010). Life cycle analysis of algae biodiesel. Int J Life Cycle Assess 15: 704-714. Sialve, B., et al., (2009).Anaerobic Digestion of Microalgae as a Necessary Step to Make Microalgal Biodiesel Sustainable. Biotechnology Advances 27(4):409416. Sheehan, J., T. Dunahay, J. Benemann and P. Roessler. (1998). A look back at the U.S. Department of Energys Aquatic Species Program Biodiesel from algae. National Renewable

55

Energy Laboratory, Golden, CO, NREL/TP-580-24190. 291 p. Stephenson, A. L., et al., (2010). Life-Cycle Assessment of Potential Algal Biodiesel Production in the United Kingdom: A Comparison of Raceways and Air-Lift Tubular Bioreactors. Energy and Fuels 24(7):40624077. Tredici MR. Mass production of microalgae: photobioreactors. In: Richmond A, eds. Handbook of Microalgal Culture: Biotechnology and Applied Phycology. Oxford: Blackwell Science; 2004: 178-214. Wiley, P. E., Campbell, J. E, and McKuin, B. (2011). Production of Biodiesel and Biogas from Algae: A Review of Process Train Options. Water Environment Research. 83, 4: 326-338(13)

56

You might also like