Vous êtes sur la page 1sur 13

Algal Research 4 (2014) 76–88

Contents lists available at ScienceDirect

Algal Research
journal homepage: www.elsevier.com/locate/algal

A critical analysis of paddlewheel-driven raceway ponds for algal biofuel


production at commercial scales
Jonathan N. Rogers a, Julian N. Rosenberg a,b, Bernardo J. Guzman a, Victor H. Oh a, Luz Elena Mimbela c,
Abbas Ghassemi c, Michael J. Betenbaugh a, George A. Oyler a,b,d, Marc D. Donohue a,b,e,⁎
a
Department of Chemical & Biomolecular Engineering, Johns Hopkins University, 3400 N. Charles St, Baltimore, MD 21218, United States
b
Synaptic Research, LLC, 1448 South Rolling Road, Baltimore, MD 21227, United States
c
Institute for Energy & the Environment, New Mexico State University, Las Cruces, NM 88003, United States
d
Department of Biochemistry, University of Nebraska–Lincoln, 1901 Vine Street, Lincoln, NE 68588, United States
e
Applied Physics Laboratory, Johns Hopkins University, 11100 Johns Hopkins Road, Laurel, MD 20723, United States

a r t i c l e i n f o a b s t r a c t

Article history: Microalgae have been promoted as the next frontier of green biotechnology and gained widespread attention as
Received 11 June 2013 desirable feedstocks for biofuels. Using conservative assumptions for microalgal growth rates (15 g m−2 d−1)
Received in revised form 13 September 2013 and total lipid content (25%), the entire “pond-to-pump” lifecycle of algal biofuels for 1000 bbl d−1 of crude
Accepted 6 November 2013
algae oil production is modeled with approximately 4875 ha of raceway ponds for solar collection and cultivation
Available online 19 December 2013
and 1463 MLD (385 MGD) of water handling capacity in the current analysis. Technoeconomic analysis based on
Keywords:
an array of 6000 modular 0.8 ha (2 acre) paddlewheel-driven ponds in New Mexico identified several cost bar-
Algae riers and resources challenges (i.e., nutrient and water resources). For 10- and 20-year capital return scenarios,
Biofuel the cost of algal oil production – $4.10 L−1 ($15.52 gal−1) and $3.21 L−1 ($12.14 gal−1), respectively – requires
Raceway ponds substantial capital and facility maintenance investments with principal cost sensitivities attributed to extraction
Sustainability efficiency and lipid content. Baseline conditions result in an energy return on investment (EROI) of 2.73. Uncer-
Technoeconomic analysis tainty in energy requirements for paddlewheels as well as water supply and circulation significantly affect the
EROI and operating costs. Alternative strategies to address the major cost barriers are needed for algal biofuels
to realize their full potential.
© 2013 The Authors. Published by Elsevier B.V. Open access under CC BY-NC-ND license.

1. Introduction States has a goal of 17% reduction in CO2 emissions by 2020. Current
projections show energy-related CO2 emissions in 2020 to be only 9%
The social, environmental, and economic pressures of human activi- below their 2005 level [2]. Therefore, an intensified expansion of carbon
ty require ever increasing energy resources. The rise of developing na- neutral renewables will be necessary to meet the milestone within the
tions coupled with a predicted expansion of the world population to coming decade.
at least 9 billion by 2050 [1] correlates to a global increase in energy Microalgae are perhaps the most prolific source of photosynthetic
use from 533 quadrillion (1015) kJ in 2008 to 812 quadrillion kJ by biomass on the planet. The controlled cultivation of microalgae on
2035 [2]. Photosynthetic biomass grown as a bioenergy crop has the po- large-scale farms offers an avenue to enhance domestic energy produc-
tential to contribute a significant amount of renewable fuel while simul- tion while minimizing land resources requirements. In 2010, liquid
taneously absorbing point sources of carbon dioxide (CO2). The United biofuels represented only 1% of the total U.S. fuel portfolio and are

Abbreviations: AD, Anaerobic Digestion; ANL, Argonne National Laboratory; APD, Algae Process Description Tool; ASP, Aquatic Species Program; BD, Biodiesel; BGY, Billion Gallons per
Year; BLY, Billion Liters per Year; CO2, Carbon Dioxide; COP, Cost of Production; DAF, Dissolved Air Flotation; DARPA, Defense Advanced Research Projects Agency; DAP, Diammonium
Phosphate; DOE, U.S. Department of Energy; DW, Dry Weight; EE, Extraction Efficiency; EERE, Energy Efficiency & Renewable Energy; EIA, Energy Information Administration; EISA,
Energy Independence and Security Act of 2007; EPA, Environmental Protection Agency; EROI, Energy Return on Investment; FWC, Flue-gas and Wastewater Co-utilization; g L− 1,
Grams per Liter; GHG, Green House Gas; ha, Hectare; HDPE, High-Density Polyethylene; HRP, High Rate Ponds; kW, Kilowatts; kWh, Kilowatt Hours; LANL, Los Alamos National
Laboratory; LCA, Life-Cycle Analysis; LEA, Lipid Extracted Algae; LLE, Liquid–Liquid Extraction; MGY, Million Gallons per Year; MGD, Million Gallons per Day; MLD, Million Liters per
Day; NMSU, New Mexico State University; NREL, National Renewable Energy Laboratory; NWIS, National Water Information System; PA, Polyacrylamide; PBR, Photobioreactor; PE,
Photosynthetic Efficiency; RA, Resource Assessment; RD, Renewable Diesel; RFS, Renewable Fuel Standard; TEA, Techno-Economic Analysis; USGS, U. S. Geological Survey; WWT,
Wastewater Treatment; WWTP, Wastewater Treatment Plant.
⁎ Corresponding author at: Department of Chemical & Biomolecular Engineering, Johns Hopkins University, 3400 N. Charles St, Baltimore, MD 21218, United States.
E-mail address: mdd@jhu.edu (M.D. Donohue).

http://dx.doi.org/10.1016/j.algal.2013.11.007
2211-9264/© 2013 The Authors. Published by Elsevier B.V. Open access under CC BY-NC-ND license.
J.N. Rogers et al. / Algal Research 4 (2014) 76–88 77

expected to reach 4% by 2035 with nearly all of their consumption at- equivalent to 18.9 and 37.8 billion liters per year (BLY), were explored
tributed to the transportation sector [2]. Algal biofuels may ease this in- for the geography and climate of New Mexico based on technical and
evitable energy transition by fulfilling a significant role in our portfolio economic analyses. In particular, this algae oil process model focuses
of alternatives to fossil fuels. on upstream cultivation using conventional technologies (i.e., raceway
Microalgae have been promoted as one of the more promising third- ponds) and locally sourced water and energy inputs. The impact of
generation biofuels for their ability to accumulate substantial amounts well-established harvesting, dewatering, and separating of methodolo-
of lipids, divide rapidly, grow in low quality water, absorb CO2, and gies on the sustainability of downstream processing was also assessed.
grow on non-arable land [3]. There is a wealth of literature that docu- The results of this technoeconomic study identify the major energy de-
ments the commercial scale growth of various microalgal species for mands, water requirements, and capital investments associated with
natural products as well as the progression of both basic and applied bi- traditional algal biofuel production. The ultimate conclusions regarding
ological research, improvements to photobioreactor (PBR) and pond de- cost of production (COP) and sustainability are held in comparison with
sign, and lifecycle analyses of algal biofuels [4–6]. The mass cultivation empirical data from the New Mexico State University (NMSU)
of microalgae was pioneered in the early 1950s with Chlorella [7] and microalgae cultivation testbed.
quickly transitioned into a modular production process using Oswald's
raceway design termed “high rate ponds” (HRPs) for large-scale
recirculating algae cultivation [8]. Although there is some debate sur- 2. Methods
rounding the “carbon neutrality” of biofuels in general, microalgae
offer significant advantages over other alternatives. Yet, the lack of uni- 2.1. Production assumptions and scalability
formity in these technologies and microbial crops still makes assess-
ment of the costs and energy requirements complex. Some of the The model in the present study was developed in order to estimate
major factors affecting algal biomass productivity include inherent pho- the scale of a facility to produce 1000 barrels of crude algae oil per day
tosynthetic constraints as well as the bioprocessing challenges related (bbl d− 1). Nutrient demand was determined from the Redfield ratio
to large-scale cultivation of microbes in water. elemental composition of C106H181O45N16P, resulting in nutrient re-
When cultivating algae in an artificial environment (e.g., outdoor quirements of 525.1 mg C g− 1-algae DW, 91.9 mg N g− 1-algae DW,
pond), it is essential that growth factors are plentiful in order to maxi- and 12.7 mg P g−1-algae DW. All references made to algal biomass in
mize growth rates [9]. While CO2 can be acquired from the atmosphere, this model pertain to the biomass dry weight (DW). As a simplifying as-
it is commonly fed into algae media to improve production [10–12]. In sumption, a year averaged microalgal growth rate of 15 g m−2 d−1 was
addition to CO2, nitrogen and phosphorus are the major nutrients chosen. Estimates for growth culture density (0.5 g L−1), harvesting
required for algae growth. It has become an accepted measure that ma- rate (10%), extraction efficiency (80%), and lipid content (25%) were
rine plankton have relatively constrained elemental ratios of 106:16:1 used as the baseline condition of the model. Due to the uncertainty of
(C:N:P) [13–15]. Although some algae species, primarily cyanobacteria, lipid fractions and the effects of specific growth conditions, all lipids
can fix nitrogen from the air [16,17], most microalgae require a soluble were assumed to be useable as a precursor to renewable diesel (RD)
form, such as urea or ammonia [18]. conversion. Based on the assumptions, a growth surface area of
In raceway ponds, paddlewheels are used to maintain constant 4875 ha will produce 730,000 ± 2000 kg d−1 algal biomass, yielding
mixing of the algae. A single paddlewheel has been shown to provide 1000 bbl d−1 algae oil. So as not to exceed the paddlewheel capabilities
sufficient mixing for algae biomass cultivation for arrays of connect- individual ponds were sized at 0.81 ha each, requiring 6000 ponds for
ed ponds covering areas as large as 5 ha [19,20]. While this shows production. The growth surface area has a depth of 30 cm resulting in a
promise for paddlewheels in HRPs, larger ponds require scale-up of total volume of 14.627 billion liters. By maintaining the culture density
the number of paddlewheels to maintain sufficient mixing. With a at 0.5 g L−1 the ponds can contain 7.3 M kg-algae, of which only 10% of
turbulent mixing velocity, the culture can maintain a uniform densi- the total raceway volume will be harvested daily. This allows the ponds
ty. The addition of eddies generated from the paddlewheel helps to to maintain production and operate in a continuous steady state by
reduce the residence time of the algae in the dark regime. At a low balancing the harvesting rate with the photosynthetic growth rate.
velocity, laminar flow will decrease the productivity of the pond; Wages were estimated to contribute 12% of the total operating costs.
yet, as the velocity increases, the power required to generate the This estimate was chosen based on similar analyses of the industry [6].
new velocity increases cubically. This presents problems for the Maintenance costs were estimated as 3% of the process equipment and
mixing velocity at rates greater than 30 cm s− 1, specifically mani- raceway pond costs [27]. All equipment was assumed to operate for
fested in increased energy costs [21]. 24 h d−1 365 d yr−1. As a simplifying assumption, an average cost of
Algae based biofuels received recent support as qualified feed- electricity of $0.11 kWh−1 was used to account for seasonal variations.
stock, making algae-derived biofuels eligible for $1.01 tax credit One of the main goals of the current study was to evaluate the poten-
per gal (Section 40 of United States Code). Despite the general public tial scalability of algae facilities. The U.S. Energy Independence Security
awareness of the various biological sources of liquid fuels and their Act of 2007 (EISA) has mandated 36 billion gallons (136B L) of renew-
potentially significant contribution to greenhouse gas (GHG) reduc- able fuel by 2022 that will in turn be ramped up to 60 billion gallons
tion [22], there remains little consensus on the lifecycle analyses (227B L) of biofuel by 2030 with provisions to emphasize the develop-
(LCA) and TEA of these biofuels [23,24]. Some of the most recent in- ment of advanced, non-corn ethanol biofuels [28]. To evaluate the
formation on algal biofuel technoeconomics shows that algae may scale up feasibility of algae production, we assumed that algae oil will
have a favorable energy return on investment (EROI) compared to contribute a moderate portion of these mandates by producing 5 BGY
fossil fuels, first-, and even second generation biofuels [25]. Howev- (18.9 BLY) of algae oil by 2022 with an increase to 10 BGY (37.8 BLY)
er, the existence of multiple biofuel production pathways, different by 2030. The current model was designed with New Mexico as a possi-
productivity assumptions and limited commercial-scale production ble site for algal production due to its climate (low seasonal variation
makes it difficult to establish theoretical mass and energy balance and high solar irradiance) and geography (flat topography). The current
equations [26]. study evaluated the capital and operating requirements associated with
In the present study, the production costs and EROI of a hypothetical algae-oil production of 1000 bbl d−1. Results from a single model pro-
algae farm and biocrude oil refinery of commercially relevant produc- duction plant were used to assess the scalability and key limitations of
tion capacity (1000 bbl d−1) were evaluated for a variety of operating algal biofuels from both economic and sustainability perspectives [25].
scenarios. The overall feasibility of this facility and its scalability to The process flow of baseline estimates for the current model is shown
meet 5 and 10 billion gallon per year (BGY) production goals, in Fig. 2.1a.
78 J.N. Rogers et al. / Algal Research 4 (2014) 76–88

(1,438 MLD)

(21 MLD)
Recycle
Recycle
Total water Flocculant
recycle DAP Urea
(33,200 kg d-1)
(1,460 MLD) (16,300 kg d-1) (37,100 kg d-1)

New water Primary


(28 MLD) (386.5 MGD) Flocculant
CO2 absorber raceway ponds dewatering
add.
CO2 (static mixers) (paddlewheels) (lamella
Algae (static mixers)
(1.3M kg d-1) clarifiers)
0.5 g L-1
Evaporation Algae
Nutrient (25 MLD)
(24 MLD) 30 g L-1
recycle
Secondary
dewatering
Feed LEA
(centrifugation)
supplement (585,000 kg d-1)

(4 MLD) Algae
200 g L-1
Phase
Biomass- Cell
Green crude separation Lipid extraction
solvent disruption
(1000 bbl d-1) (Distillation (static mixer)
separator (sonication)
column)

New hexanes
Hexane recycle (17.5 kg hr-1)
(350,000 kg hr-1)

Fig. 2.1a. Single plant baseline operating flow diagram. Dashed lines indicate possible pathways for LEA. Additional processing such as anaerobic digestion would be necessary for nutrient
breakdown and the production of biogas. An alternative pathway would be to sell LEA as a nutritional supplement for livestock or aquaculture.

2.2. Economic, energy, and resource estimates the clarifiers was assumed to be 30 g L−1 [30]. Centrifuges were used
for additional dewatering of the algae water stream to 200 g L−1 [6].
The plant capital and operating costs were estimated using industri- The current model requires 12 centrifuges, totaling $6.6 M in capital in-
al estimates and reports from similar processes. Three static mixers vestment, with each operating at 55 kW continuously and costing
would facilitate CO2 and nutrient addition to the ponds totaling about $0.64 M annually (GEA Westfalia). The current model requires 13
$0.17 M capital investment (Westfall Manufacturing). Paddlewheel es- sonicators for cell disruption, totaling $3.25 M. The sonicators operate
timates were based on the requirements for a 0.81 ha raceway with a at 16 kW continuously and process up to 12 m3 h−1, costing $0.20 M
channel 12.2 m across and 30 cm deep. The 6000 ponds in the current annually (Hielsher Ultrasonics). In the baseline scenario, extraction is
model require a total of 24,000 paddlewheel units, and span over performed on the 20%-solid algae slurry using hexane at a cost of
4875 ha. Capital cost estimates for paddlewheels totaled a $120 M cap- $3000 per ton. A small static mixer is used to combine the solvent and
ital investment (Waterwheel Factory Inc.). The impacts of three energy algae during lipid extraction costing $4500 (Westfall Manufacturing). A
conditions were evaluated in the model. In the baseline scenario, the solvent to algae-DW mass ratio of 10:1 was assumed with a 1 h recycle
paddlewheels were calculated to require 0.22 W m−2 (Appendix B.1) time, totaling a 350 t requirement for extraction. Hexane was estimated
and contribute $10.1 M of annual operating costs. Paddlewheel energy to contribute $1.05 M of initial capital investment. It was assumed that
scenario 2 requirements are 0.73 W m−2 based on power usage specs there would be a daily solvent loss of 0.005%, contributing to an addition-
from an industrial supplier (Waterwheel Factory Inc.), resulting in al $0.4 M of annual operating costs. In order to separate the LEA from the
annual operating costs of $34.5 M. In paddlewheel energy scenario 3, solvent-crude mix the current model uses an oil-water separator with
assumptions from the current model are applied to pilot plant 3000 GPM capacity and residence time of 1 h (Hydro-Flo Technologies),
paddlewheel energy measurements from the NMSU testbed to assess totaling a $0.25 M capital investment. The separator operates at 6 kW
scale up feasibility. Two pilot ponds, each with a growth surface area continuously and contributes about $6000 of annual operating costs.
of 32 m2, were operated by paddlewheels with individual motors. The present model uses a distillation column, $0.5 M capital investment
Paddlewheel energy measurements were taken using a fluke meter [31], with a 595 GPM (2252 L min−1) capacity for solvent recovery and
while circulating water in the ponds continuously 24 h d−1 to deter- purification of the crude oil stream. Energy calculations (Appendix B.2)
mine energy consumption under load. The total energy consumed by assume continuous operation of the column to contribute an annual en-
the two paddle wheel motors was measured to be 8.16 W m−2. Scaling ergy cost of $2.7 M. In an alternative scenario the economic and energy
up to the production capacity of that model would not be feasible for impact of drying the algae slurry to 10%-water content is evaluated
commercial scale and would result in $383.1 M of annual operating (Appendix B.2).
costs. Mixing energy inputs exceeded the theoretical growth mixing re- The proposed liner is a high-density polyethylene (HDPE) liner [32].
quirements at the test scale and represented a significant discrepancy A 40-mil HDPE liner was modeled to line 6000 ponds at a price of
between experimental ponds and commodity-scale ponds used for $0.77 m−2, totaling $250 M [6]. The landscaping cost was estimated at
algal cultivation. Further optimization is needed to determine minimum $0.16 m−2, contributing a total of $57 M [33]. Covering the raceways
mixing energy requirements and will assist in bridging the gap between cost an estimated $0.98 m−2, totaling $315 M [34], which included a
pilot- and commercial-scale operations. transparent PE cover and greenhouse structure. Additional infrastruc-
The current model assumes a polymer flocculant cost of $100 per ton ture costs needed for the plant include railways, roads, buildings,
of algae processed and contributes $26.7 M annually to the process (SNF pipes, and construction. The current model assumes that railways and
Polydyne). Two static mixers are required for flocculant addition totaling roads will be necessary for the transport of raw materials and access
$0.12 M (Westfall manufacturing). Estimates for two, 76.2 m diameter to the facility. The study estimates that a $1.85 M investment will pro-
lamella clarifiers, added a capital investment of $2.5 M based on a project vide 8.0 km of rail and 4.8 km of road for the facility [35,36]. Buildings
for the City of Detroit WWT, from Monroe Environmental [29]. The are estimated to contribute $0.45 M for a chemical storage warehouse,
5.96 kW rakes were assumed to run continuously for the operation of a central operations facility, process buildings, and additional storage
each clarifier, costing $11,500 annually. The algae water stream leaving [37]. Construction costs are estimated to contribute $2.5 M for the
J.N. Rogers et al. / Algal Research 4 (2014) 76–88 79

facility, based on a similar study done on the profitability of soybean 12.7 mg-P g−1-algae DW at an uptake efficiency of 50% [6], contribut-
biodiesel from Iowa State [37]. Sizing estimates for piping were made ing $7.36 M of annual operating costs. In addition to DAP, the model
using the volumetric flow rates throughout the process to maintain an uses urea (46% N, $379 ton−1 [47]) as a nitrogen source at an uptake ef-
average flow velocity of ~2.5 m s−1. Estimates were simplified by as- ficiency of 76% to meet the 91.9 mg-N g−1-algae DW [6], contributing
suming 188,000 ft of piping requirements at an average cost of $24.43 M of annual operating costs (Table 2.2a).
$11.43 m−1, totaling approximately $7 M [38,39].
Land was assumed to cost $3,707 ha−1 for property in New Mexico, 3. Results
totaling approximately $18 M [40], and did not consider CO2 (flue gas)
or railway accessibility. Based on a resource assessment performed by 3.1. Results overview
The Solar Energy Research Institute, the Tularosa Basin and Crow Flats
Basin are comprised of approximately 39,000 ha and a regional saline- The assumptions made for algal growth, dewatering, and extraction
water availability of 750–1630 billion gallons (2839–6170 billion L) suit- efficiency allowed us to estimate the costs and energy usage for each
able for microalgal cultivation [41]. Although the current study operates step of the process. The production goal of 1000 bbl-algae oil per day
under the assumption of using fresh-water algae, it assessed the feasibility in New Mexico was found to require 730,000 kg ± 2000 kg-algae d−1
of New Mexico for a theoretical strain of marine-algae with otherwise (depending on inconsistencies in bioproduction) via a growth surface
identical characteristics to that used in the present model. area of 4875 ha. The plant also produces 585 tons of lipid extracted
Several simplifying assumptions were applied to the availability of algae (LEA) daily that can potentially be sold as a feed supplement or
water, which assumed to be accessible and that pumping was the only used for anaerobic digestion and nutrient recycle. The estimated cost
associated cost. No permit fees, taxes, or transportation considerations of production that covers operating costs and a 10- or 20-year return
were applied to the baseline estimate for the cost of water. An average of capital investment resulted in a baseline cost of $4.10 L− 1
of ground water level measurements in New Mexico taken by the ($15.52 gal−1) or $3.21 L−1 ($12.14 gal−1) of crude oil, respectively.
National Water Information System (NWIS) of 114 m (Table A.1) was The operating cost of production was estimated to be $134 M annually.
used to calculate the pumping energy requirements to fill and maintain Five operating factors – paddlewheels, nutrients, water, maintenance,
the raceway ponds. Using groundwater from this depth, filling the pond and harvesting – were identified for their major impact on either cost
to the total operating volume of 14.63 BL (3.864 BG) was found to cost or sustainability to the process. For the purpose of this study, the plant
$2.9 M. Two scenarios of 6 m (baseline) and 12 m head were estimated was classified as sustainable if it could annually achieve its nameplate
for the on-site pumping circulation requirements of the 1463 MLD capacity without any significant detriment to the surrounding
(386.5 MGD) processed, contributing annual operating costs of $5.7 M
and $11.5 M for the scenarios respectively (Appendix B.3). Industrial Table 2.2a
turbine pumps were used as the basis for the capital cost estimates of Key assumptions and model details.

the pumps throughout the plant, totaling $1.27 M of capital costs Algae strain Chlorella vulgaris
[42,43]. The current model estimates that approximately 1457 MLD Elemental composition of algae biomass C106H181O45N16P
(385 MGD) of water are recycled daily throughout the dewatering Average annual areal productivitya 15 g m−2 d−1
Biomass lipid content 25 wt.%
and back end processes. Despite the high percentage of process water
Daily oil production 1000 bbl d−1 (159,000 l d−1)
recovered, a daily loss of about 0.03% of the total plant water volume Density 920 kg m−3 (146,268 kg-oil d−1)
equates to 3.67 MLD (0.97 MGD) for the 4875 ha plant. Far more critical Extraction efficiency (base) 80%
however is the amount of water lost due to evaporation from the ponds. Required daily biomass (DW) 730,000 ± 1,500 kg d−1
Evaporation loss for the plant is 243.8 MLD (64.4 MGD, 0.5 cm d− 1) Dilution rate 10%
Total biomass in raceways 7.31 M kg
and 24.2 MLD (6.4 MGD, 0.05 cm d−1) for the two evaporation scenar- Required growth surface area for daily 4875 ha
ios. Groundwater pumping for replacing 247.6 MLD (0.5 cm d− 1 production
evaporation, 3.67 MLD process loss) and 28.0 MLD (0.05 cm d−1 evap- Growth surface area per pond 0.81 ha
oration, 3.67 MLD process loss) was found to have an annual cost of Number of ponds in plant 6000
Maximum culture density 0.5 g L−1
$18.3 M and $2.1 M respectively. The total annual water consumption
Raceway depth 30 cm
per plant, producing 57.9 million liters (15.3 million gallons) of Raceway volume 14.62 billion liters
crude-algae oil/yr, is estimated at 90.5 billion and 10.2 billion liters of CO2 recovery to culture 50%
water for the uncovered and covered raceway pond designs, respective- Nitrogen recovery to culture 76%
ly. The two scenarios were found to require from 176 L-water to P recovery to culture 50%
Net N demand 91.9 mg g−1-algae DW
1560 L-water per L-algae oil. Our wastewater treatment costs were es- Net P demand 12.7 mg g−1-algae DW
timated based on a facility of similar processing capacity totaling a Pond mixing baseb 0.22 W m−2
$245 M capital investment [44]. Only relevant WWT steps were consid- Pond mixing 2c 0.73 W m−2
ered due to lower process requirements. Pond mixing 3d 8.16 W m−2
Lamella separators 3.91 × 10−4 kWh kg−1-algae DW
Nutrient consumption was estimated using the Aspen Plus v7.3
Centrifuge power 2.17 × 10−2 kWh kg−1-algae DW
(Aspen Technology, Inc.) chemical process modeling software suite. Cell disruption 6.83 × 10−3 kWh kg−1-algae DW
The uptake was divided into photosynthesis and biosynthesis in order Biomass-separator 1.97 × 10−4 kWh kg−1-algae DW
to adhere to the program's capabilities. Photosynthesis was modeled Hexane recovery 4.68 × 10−1 kWh kg−1-oil
by combining CO2, water, and light to produce sugars and oxygen. Con- Groundwater depth 114.3 m (375 ft)
Pumping water from off-site 2.12 × 10−3 kWh L−1
version efficiency considered poor light penetration in ponds, inherent Evaporation (base) 0.05 cm d−1
photosynthetic loss, and cellular energy use; therefore, we chose a 4% On-site pumping circulation base 9.84 × 10−5 kWh L−1
overall conversion of solar energy [45] and a 50% conversion of CO2 to (6 m head)
meet the demand of 525.1 mg C g− 1-algae DW in order to simplify On-site pumping circulation 2 1.97 × 10−4 kWh L−1
(12 m head)
mass balances. Biosynthesis modeled the conversion of sugar, a nitrogen
source, and a phosphorus source into algal biomass (lipid, starch, a
The annual average areal productivity of 15 g m−2 d−1 in New Mexico s accounted
protein), water, and oxygen. The current model found an annual re- for by assuming 19 g m−2 d−1 peak algal biomass production in the spring and summer
months with 8 g m−2 d−1 during the fall and winter [46,48].
quirement of 0.280 million tons of CO2 at a cost $40 per ton [46], con- b
Calculated in Appendix B.1.
tributing $11.2 M of annual operating costs. Diammonium phosphate c
Based on power usage specs from an industrial supplier (Waterwheel Inc.).
(DAP, 18% N, 46% P, $499 ton−1 [47]) is used to meet the demand of d
Pilot plant paddlewheel energy measurements from NMSU testbed.
80 J.N. Rogers et al. / Algal Research 4 (2014) 76–88

$350.00
environment. Additionally, capital investment was estimated to be ap-
proximately $1035 M. Five capital expenses; paddlewheels, wastewater $300.00
treatment equipment, land/landscaping, pond liners, and pond covers $250.00
were found to contribute over 95% of the total capital investment for

Millions
$200.00
the plant. A comparison of the operating and capital costs is shown in
Figs. 3.1a and 3.1b respectively. Complete tables of the results (present- $150.00
ed in both $ L−1 and $ gal−1) can be found in the supplementary mate-
$100.00
rial (Tables A.2, A.3).
$50.00

3.2. Economic sensitivity analysis $0.00

Multiple scenarios were modeled for adjustments to the major oper-


ating and capital factors. The sensitivity analysis was performed for the
10- and 20-year return of capital investment scenarios and can be seen
in Figs. 3.2a and 3.2b, respectively. Tables A.4 and A.5 display the COP
sensitivity analysis results in terms of $ L−1 and $ gal−1 respectively.
The sensitivity analysis shows that the cost of production is the most
heavily influenced by the useable lipid content of the algae and the ex- Fig. 3.1b. Baseline capital cost comparison. Capital investment was estimated to be approx-
traction efficiency (EE) of lipids in the process. Changes to the algal imately $1035 M. Paddlewheels, wastewater treatment equipment, land/landscaping,
pond liners, and pond covers were found to contribute over 95% of the total capital invest-
lipid content and EE directly impact productivity and do not affect the
ment for the plant.
scale of the growth infrastructure and only slightly impact the processing
requirements of the model. Therefore, as a simplifying assumption only
productivity was varied and all operating and capital costs were kept algae crude. Increasing EE in the 10- and 20-yr capital return scenarios
constant for the sensitivity analysis of lipid content and EE. By increasing lowers COP by $0.82 L−1 and $0.64 L−1 respectively. Lowering EE has
the algal lipid content to 35% productivity would increase from the base- a more pronounced impact on the cost L−1; increasing the COP by
line scenario (25% lipid content) of 57.9 MLY to 81.4 MLY and lower COP $1.37 L−1 (10-yr) and $1.07 L−1 (20-yr). The significant investment re-
by $1.17 L−1 (10-yr) and $0.92 L−1 (20-yr). Contrarily, decreasing the quired for pond liners and covers result in a strong dependence to the
algal lipid content to 15% lowers the production to 34.8 MLY and sub- capital return scenarios. It can be seen that the impact on cost L−1 is re-
stantially increases COP by $2.73 L−1 (10-yr) and $2.14 L−1 (20-yr). laxed significantly in the 20-yr capital return case. Whether or not the
This range of oil contents (±10%) extended from our baseline of 25% pond uses a liner impacts the COP by ±$0.43 L−1 and ±$0.22 L−1 in
total lipids represents a single biological variable with the ability to sur- the 10-yr and 20-yr scenarios respectively. Whether or not to cover the
pass DOE's 2014 target of reducing the “modeled mature plant cost of ponds also shows a reduced impact on COP in the 20-yr return scenario
algal open pond oil” [49]. Therefore, one priority of future research will at $0.27 L−1 as opposed to $0.54 L−1 in the 10-yr return case. The paddle
be to evaluate different approaches to increasing oil content, whether wheel energy requirements are independent of capital expense and are
by genetic intervention or controlled nutrient regimens, such as nitrogen therefore unchanged for the two capital return scenarios. Raising the
deprivation. Additionally, future research can increase productivity by paddlewheel energy requirements from 0.22 W m−2 to 0.73 W m−2 re-
improving the average areal growth rate of the algal biomass used. A de- sulted in a $0.42 L−1 increase. The decrease in net harvesting efficiency
crease in the growth rate to 10 g m−2 d−1 from the baseline scenario uses the simplifying assumption that all operating and capital costs re-
(15 g m−2 d−1) would reduce the productivity to 38.6 MLY and sub- main the same and final production is decreased from 57.9 MLY
stantially increase COP by $2.05 L−1 (10-yr) and $1.60 L−1 (20-yr). (100%) to 49.3 MLY (85%) of algae oil. The lower harvesting efficiency re-
Increasing the growth rate to 20 g m−2 d−1 would increase the produc- sults in a $0.72 L−1 (10-yr return) and $0.57 L−1 (20-yr return) increase.
tivity to 77.2 MLY and lower COP by $1.02 L−1 (10-yr) and $0.80 L−1 It can be seen clearly that flocculant costs constitute nearly all of the net
(20-yr). The three EE scenarios would result in either an increase to harvesting costs. The flocculant costs contribute ±$0.23 L−1 (both re-
72.7 MLY (100% EE) or decrease to 43.5 MLY (60% EE) production of turn scenarios) of the total impact, ±$0.24 L−1 (both return scenarios)
when adjusting total harvesting costs ±50%. The economic impact of N
and P recycling operated under a simplifying assumption that eliminated
$30.00 the cost of DAP and urea addition, resulting in a $0.55 L−1 decrease. The
extraction costs only considered the solvent costs and the energy costs
$25.00
required to run the extraction equipment. The extraction costs in the cur-
$20.00 rent model assumed that the use of hexane at a 10:1 solvent to algae DW
ratio would achieve 80% extraction efficiency from 20%-solids algae [6]. It
Millions

$15.00 can be seen that decreasing the baseline extraction costs by 50% resulted
in a COP decrease of $0.03 L−1 in both scenarios. There is not much infor-
$10.00
mation available about extraction techniques being used at the commer-
$5.00
cial scale. Conventional solvent extractions used in agricultural processes
require the biomass to have less than 10% water content [4]. Drying the
$0.00 algae paste leaving the centrifuges (185.4 tons, 20%-solids) to 10%
water content requires 122 M kWh of energy annually, and causes COP
to increase by $0.23 L−1 (10- and 20-yr scenarios). The costs associated
with evaporation rate are based on the calculations performed for the
energy requirements of pumping replacement groundwater from a
depth of 114 m. Using the assumption that water levels and pumping
costs would remain constant, increasing the evaporation rate from
Fig. 3.1a. Baseline annual operating cost comparison. Operating cost of production was
0.05 cm d−1 to 0.5 cm−1 had a $0.28 L−1 increase on COP. Increasing
estimated to be $134 M annually. Paddlewheels, nutrients, water, maintenance, and
harvesting were identified for their major impact on either cost or sustainability to the the on-site pumping capacity for 1463 MLD (386.5 MGD) from a 6 m
process. to 12 m head resulted in a $0.10 L−1 increase on COP.
J.N. Rogers et al. / Algal Research 4 (2014) 76–88 81

Lipid content (35% : 25% : 15 %)

Extraction efficiency (100% : 80% : 60%)

Areal productivity g/m2/d (20 : 15: 10)

Pond liner (no liner : liner : replace liner once)

Cover (no cover : cover : replace cover once)

Paddle wheels (0.22 W per m2 : 0.73 W per m2)

Net harvesting efficiency (100% : 85%)

Net Harvesting costs (50% : base : 150%)

Flocculant ($50 : $100 : $150 per ton-algae)

$ per ton CO2 ($20 : $40 : $60)

N+P recycle (100% : 0%)

Extraction costs (50% : base : Dry algae to 90% DW)

Evaporation rate (0.05 : 0.5 cm/day)

On-site pumping capacity (20 ft : 40 ft head)

-$1.50 -$1.00 -$0.50 $4..10 +$0.50 +1.00 +$1.50 +$2.00 +$2.50 +$3.00

Fig. 3.2a. Sensitivity analysis on COP (10-yr return of capital investment). Estimates are based on the baseline operating costs of $134 MM and annual capital payments of $104 MM to
cover a 10 year return of the $1.04 billion capital investment. The baseline COP was found to be $4.10 L−1 ($15.52 gal−1).

3.3. Energy sensitivity analysis aspects of the plant (listed from high to low) are due to the
paddlewheels, on-site pumping circulation, solvent recovery, pumping
Maintaining a positive net energy balance is a major issue for biodie- to fill the ponds, and pumping replacement water. The baseline scenario
sel production. The amount of energy generated must be greater than results in an EROI of 2.73. The alternative energy graph shows the im-
the amount of energy that is input to make it. The current model pro- pact for increases in the requirements for replacement water pumping
duces 53.4 M kg-algae oil yr− 1. Algal biodiesel has been reported to (147.2 M kWh annual energy increase), on-site pumping circulation
have a higher heating value of 41 MJ kg−1 [50]. Under this assumption (52.3 M kWh annual energy increase), paddlewheels (221.6 M kWh
we can estimate that the annual oil production as approximately annual energy increase), and algae dewatering to 90%-solids
608.2 M kWh of energy. The energy return on investment (EROI) (122.3 M kWh annual energy increase). The current model found that
must be above 1, otherwise the plant will consume more energy than in the worst-case scenario, in which the energy increases from all of
it generates and is inherently unsustainable. Alternative energy scenar- the alternative scenarios were combined, the EROI dropped to 0.80.
ios were evaluated for the amount of replacement water required due to This is an unsustainable scenario in which the plant consumes more en-
evaporation, on-site pumping capacity, and paddlewheels. The major ergy than it produces and highlights the importance of minimizing en-
energy requirements of the plant for both the baseline and alternative ergy inputs for commercial-scale algal cultivation.
scenarios can be seen in Fig. 3.3a.
The annual energy requirements show the energy dependence of 4. Discussion
different processes throughout the model. The combined energy of all
plant processes must be less than the energy generated through oil pro- Scalable production capability is a critical element in analyzing the
duction. The baseline scenario shows that the most energy intensive viability of algae oil as a renewable fuel source. Algae may play a crucial

Lipid content (35% : 25% : 15 %)

Extraction efficiency (100% : 80% : 60%)

Areal productivity g/m2/d (20 : 15: 10)

Pond liner (no liner : liner : replace liner once)

Cover (no cover : cover : replace cover once)

Paddle wheels (0.22 W per m2 : 0.73 W per m2)

Net harvesting efficiency (100% : 85%)

Net Harvesting costs (50% : base : 150%)

Flocculant ($50 : $100 : $150 per ton-algae)

$ per ton CO2 ($20 : $40 : $60)

N+P recycle (100% : 0%)

Extraction costs (50% : base : Dry algae to 90% DW)

Evaporation rate (0.05 : 0.5 cm/day)

On-site pumping capacity (20 ft : 40 ft head)

-$1.50 -$1.00 -$0.50 $3..21 +$0.50 +1.00 +$1.50 +$2.00 +$2.50 +$3.00

Fig. 3.2b. Sensitivity analysis on COP (20-yr return of capital investment). Estimates are based on the baseline operating costs of $134 MM and annual capital payments of $52 MM to cover
a 20 year return of the $1.04 billion capital investment. The baseline COP was found to be $3.21 L−1 ($12.14 gal−1).
82 J.N. Rogers et al. / Algal Research 4 (2014) 76–88

35000
operating costs and result in an annual difference of 221.6 M kWh. In
30000 the third scenario, the paddlewheel energy measurements for the
Baseline
Alternative NMSU pilot pond totaled 8.16 W m−2 and require further optimization
25000
to determine the minimum mixing energy requirements for experi-
Million kWh

20000 mental growth. The uncertainty when scaling up production is clearly


shown by the difference between the theoretical and measured
15000
paddlewheel energy requirements and indicates the need for the devel-
10000 opment of alternative options. Other mixing devices, such as air lift
pumps, mixing boards, Archimedes screws, and mechanical pumps
5000
have all been found to suffer from inflexibility in operation and high
000 capital costs despite relatively good efficiency [19]. Promising alterna-
tives may eliminate mechanical mixing by utilizing large-scale open-
ponds with sloped or corrugated designs [57,58].

4.2. Pond liners

The liner is one of the most expensive items for raceway ponds.
Fig. 3.3a. Annual plant energy comparison. Alternative energy scenarios for the amount of
Liners enable a raceway pond to be built in otherwise unsuitable terrain.
replacement water required due to evaporation (28.0 MLD to 247.6 MLD), on-site
pumping capacity (20 ft to 40 ft head), paddlewheels (0.22 W m−2 to 0.73 W m−2), Liners can be made from different materials such as clay, concrete, as-
and the additional dewatering of the algae to 90%-DW for extraction are compared with phalt, fiberglass, and HDPE [59]. Although ponds can be located in
the baseline energy values. areas where clay is abundant and inexpensive, it can crack when the
ponds are dry and cause the ponds to lose essential nutrients and
role in contributing to the EISA biofuels mandate of 36 billion gallons water. Clay ponds cannot be cleaned like those utilizing a HDPE liner
(136B L) of renewable fuel by 2022 and 60 billion gallons (227B L) by and can lead to an increase in potential contamination [6,56]. Despite
2030. In order for algae oil to make a significant contribution to the re- being costly, liners may be necessary to help mitigate these problems.
newable fuel industry, we proposed 5 BGY (18.9B LY) of algae oil by In the current model, liners constitute a massive portion ($250 M,
2022 with an increase to 10 BGY (37.8B LY) by 2030. New Mexico was 24%) of the capital costs for the facility. Liners represent a crucial cost
evaluated as a possible site for algal production to meet these demands. prohibitive element in the use of raceway ponds for algal biomass pro-
The current model evaluated the COP impact of both 10-yr and 20-yr duction and a recent demonstration plant construction by Sapphire
capital investment return scenarios. The baseline algae oil COP was Energy in New Mexico without liners has indicated the viability of pro-
found to be $4.10 L− 1 ($15.52 gal−1) (10-yr) and $3.21 L− 1 ceeding successfully in the cultivation of algae without.
($12.14 gal−1) (20-yr). The COP results of this model fall in line with
the priorities set by the DOE Office of Energy Efficiency & Renewable En- 4.3. Flocculant options
ergy (EERE) for Bioenergy Technologies to reduce the cost of open pond
algal oil to $14.31 per gallon ($3.78 L−1) of gasoline equivalent by 2014 Harvesting single cells from liquid suspensions is energy intensive
[49]. Although this model does not take conversion costs of the algae and difficult when applied to continuous operation of large-scale
crude to useable fuel, it identifies key areas that have the potential to re- bioprocessing facilities [60]. Membrane filtration by size exclusion is a
duce COP and help realize the EERE goals. simple and cost effective separation technique; however, even ad-
vanced filtration methods are subject to caking and erosion based on
the throughput of biomass needed for biofuel production, which can
4.1. Paddlewheels lead to significant reductions in efficiency over the lifetime of the mem-
brane [61]. Low-speed centrifugation may be a more feasible alternative
Paddlewheels are needed for the cultivation of algae, but their main for algae if a flocculant agent is used to increase the average particle size
drawbacks come from their operating costs. The present model assumes of microalgae slurries from 10 μm to 10 mm by forming large masses of
that the paddlewheels will maintain a mixing velocity of 30 cm s−1 for many small particles to facilitate solid separation from a liquid media
the full 365 days with constant operation and an average productivity of [62]. The use of cationic polymers is a well-established method of floc-
15 g d−1. Raceway ponds are inherently subject to varying intensities of culating particulate matter from large volumes of liquid as in wastewa-
natural light throughout the course of the day. Absence of light during ter treatment facilities or paper processing plants [63,64] Dissolved air
the night is inevitable and could result in biomass losses as high as flotation and settling clarifiers have also been successfully implemented
25% [51]. Sufficient light penetration is variable in any algae cultivating to harvest flocculated algal biomass at demonstration scales. The algal
system, due to the density of the algae culture. Thus, there is an inherent biomass harvesting process modeled in the present study relies on a
light attenuation regime in raceway ponds (i.e., algae cells near the flocculation step followed by lamella clarifiers to bring the ~ 2 g L−1
bottom of the ponds may be in the dark regime) [52]. Therefore, algae suspension to a 30 g L− 1 slurry, followed by centrifugation of
ample mixing should be provided to minimize residence time in this slurry to a 200 g L−1 paste.
the dark regime [53–55]. Alternatively, one method of reducing the op- Polyacrylamide (PA) flocculants are most widely used and can be ef-
erating costs of the paddlewheels is to decrease the mixing velocity dur- fective with microalgal cells; however, if the biomass residues are to be
ing the night. Techniques such as operating the paddlewheels on a used as LEA for other co-products, PA may foul the biomass from a
variable schedule (e.g. run at 30 cm s−1 for 14 h, and then reduce it to health perspective. This process was designed using a flocculant dose
25 cm s−1 for 108 h) can be used to increase paddlewheel productivity estimated at $100 per ton of DW algae harvested based on the operat-
and efficiency [56]. Another method to reduce paddlewheel-operating ing parameters. While flocculant costs in excess of $100 per ton of
costs would be to cease operations during the winter. During the winter, DW algae are prohibitive for biofuel production, a target of less than
algae productivity decreases due to the low temperatures. This method $40 per ton may be an achievable production target with alternative
is an extreme case, when the operating costs far exceed the expected re- flocculant materials. Metal ion based flocculants have been deemed in-
turn [56]. compatible with end biofuel standards and pH-induced changes are
In both operating cases (0.22 W m−2 and 0.73 W m− 2) the unfeasible for algal biofuel production at commercial scales; thus, bio-
paddlewheels represent a significant portion of the total annual logical approaches to flocculation have gained significant attention in
J.N. Rogers et al. / Algal Research 4 (2014) 76–88 83

recent years. These processes can occur either as a natural “auto- feed is still being researched, leading to an uncertainty in selling price
flocculation” mechanism in mature or nutrient deprived algae cul- estimated to be from $250–$1000 ton−1 [82]. Each 1000 bbl d−1
tures or instigated by a biological agent, such as the introduction facility could therefore potentially generate between $53.3 and
of another microorganism (algae, bacteria, cyanobacteria, predators $213.5 million from the sale of LEA as a protein supplement.
or grazers) [65–68] or a biomolecule with the bridging characteris-
tics similar to a polymer or adhesive qualities of a coagulant [69]. By 4.6. Nutrient consumption
employing these methods, some estimates project that operating
costs can be reduced by $200–400 per million gallons of process Nutrients represent an essential element of algae growth. The costs
water treated by transitioning from chemical flocculants to and sustainability of which depend heavily on the nutrient source and
bioflocculation [70]. Furthermore, the production of biological floc- recyclability. As algae-oil technology continues to develop, it will be cru-
culants is highly scalable and can be incorporated into the algae bio- cial to supplement CO2, nitrogen, and phosphorus requirements by
fuel production plant. recycling nutrients and co-locating algae farms with power plants and
wastewater [4]. Increased demand of limited resources and the exhaus-
4.4. Extraction efficiency, lipid content, and areal growth rate tion of optimal growth locations are some of the difficulties associated
with scaling up production.
The extraction efficiency, algae lipid content, and areal growth rate Carbon is the largest nutrient requirement and is fed into the autotro-
directly correlate to the overall productivity of the operation. The sensi- phic process in the form of CO2, in which 768 tons are required daily for
tivity of these three factors was shown to have the highest economic production in the plant. The production process examined in this study
implications and represents a valuable opportunity for future research requires roughly 0.28 million tons of CO2 annually. At the price of
to improve. It is crucial for the extraction process to be as efficient as $40 per ton, CO2 annually contributes $11.2 M of the annual operating
possible in order to minimize costs and maximize productivity. Some costs. The source of CO2 can have a significant impact on price and the
dewatering is important to reduce processing volumes, however exten- cost of production. CO2 sourced from an ethanol refinery may be provid-
sive drying must be avoided and techniques for efficient extraction from ed for as low as $20 whereas CO2 sourced from flue gas or other industri-
aqueous biomass (less than 20%-solids) must be optimized to avoid al sources may range from $40–$60 [39]. Therefore, operating expenses
costs and energy requirements. Because such a substantial portion of can be minimized co-locating commercial-scale algal biofuel develop-
COP is in the capital investment required for the growth infrastructure, ments with suitable CO2 sources. Scaling up to achieve the contribution
and increases to the algae lipid content from 25% to 35% or microbial goals of 5 BGY and 10 BGY to the EISA biofuel mandates would require
growth rate from 15 g m−2 d−1 to 20 g m−2 d−1 improve production 92 million and 183 million tons of CO2, respectively. According to the
without a need for the development of additional growth area. As such, EPA, coal-fired power plants in the United States emitted 1.6B metric -
there is a strong incentive to genetically engineering microalgae for aug- tons of CO2 in 2009 [83]. The 5 BGY and 10 BGY scale up contributions
mented areal growth and oil productivity [71], which can significantly would require 5.7% and 11.4% of all coal-fired power plant emissions in
impact the cost of production, as evidenced by our sensitivity analysis. the United States, respectively. Algae have fantastic potential to seques-
There is an equally strong motivation to control the unintended release ter the emitted CO2 to be recycled into biofuel [84], however many of
of such genetically modified (GM) organisms into the natural environ- the coal-fired plants are located in areas not suitable for algae-growth.
ment; such scenarios for GM algae have been investigated recently Additional work needs to be done at existing coal-fired plants to assess
[72–75]. Additionally, the areal growth rate of microalgae has been the potential construction of algae-farms at their location. Efforts should
shown to be dependent on seasonal variations [48] averaged on an annual also be focused on building future coal-fired plants in areas that will also
basis for simplicity in this model. These seasonal variations create a differ- be suitable for algae-oil production.
ence between production and processing capacity, resulting in excess algal The present model relies on urea as a nitrogen source, which con-
production in peak months and excess in processing capacity in winter, tributes $24.43 million of the annual operating costs. Ammonia is an-
which must be considered when sizing future commercial scale develop- other optional nitrogen source that is similar in price and availability
ments. The current model assumes that all extracted material is useable to urea [6]. A single 1000 bbl d− 1 plant that produces 57.9 MLY
for conversion to fuel; however, high triglyceride content would be the (15.3 MGY) of oil is estimated to require 64,459 tons of urea annually
most suitable for fuel conversion via transesterification. Alternative tech- for operation. Scaling up in attempts to contribute to the EISA biofuels
niques such as hydrothermal processing or pyrolysis may allow for higher mandates would require 21.1 million and 42.2 million tons of urea an-
yields via the conversion of the entire algal biomass into fuel, however no nually for the production of 5 BGY and 10 BGY of algae-oil, respectively.
such systems are currently being used at the industrial scale [76,77]. Natural gas is the main input used to produce ammonia and with the
major increase in supply of natural gas over the past few years' nitrogen
4.5. Lipid extracted algae supply is not currently an area of concern [85] and may lower the asso-
ciated operating costs. In this study, the phosphorus requirement is met
A potential benefit of the use of algae for biofuel production is the with 14,746 tons of DAP annually. The demand was found to be
potential value of the spent biomass as animal feed. The use of spent 12.7 mg-P/g-algae DW at an uptake efficiency of 50%. 40.4 tons d−1 of
biomass as a side product has significant contributions to the value of DAP are used to meet the gross requirement 18.6 tons d−1 of phospho-
algae biomass. Nutritional and toxicological evaluations have demon- rus. In the current model, DAP annually contributes $7.36 million of the
strated that algae biomass is suitable as a feed supplement or substitute annual operating costs. Scale up for the 5 and 10 BGY production sce-
for conventional animal feed sources [78,79]. For LEA to be considered narios would increase the demand to 4.82 and 9.64 million tons of
for use as a feedstock the algae strain must have a composition suited DAP, respectively. Phosphate rock is a non-renewable resource that
for both nutritional and fuel production purposes [80]. For example de- has taken 10–15 Ma to form. World population growth ensures the
spite the high nutritional value of Spirulina, it contains very low lipid need for phosphate fertilizer to grow crops for food and biofuels [86].
content and is not considered a viable feedstock for biofuel production The largest deposits are found in northern Africa, China, the Middle
[81]. Although lipid extraction will remove beneficial fatty acids (such East, and the United States. Large resources have been identified on
as omega-3 PUFAs) and lower the overall nutritional value of the algae the continental shelves and seamounts in the Atlantic Ocean and the Pa-
biomass, the LEA may still be used as a high protein supplement. cific Ocean however they currently cannot be recovered economically
The theoretical commercial scale algae facility in this study is esti- with the current technology. At the current rate of production, the phos-
mated to produce 585 tons d−1 of lipid-extracted algae. As mentioned phate reserves in the United States will be depleted in about 50 yr [87].
the quality of LEA as a protein supplement in animal and aquaculture There are no substitutes for phosphorus in agriculture. This dependence
84 J.N. Rogers et al. / Algal Research 4 (2014) 76–88

will force the United States to rely on imported fertilizers to grow crops. as it requires 168% and 704% of all of the freshwater consumed in the
The growth of biofuels will only serve to expedite the depletion of state for the covered and uncovered raceway scenarios, respectively.
existing phosphorus reserves. From 2006 to 2012 the average value of The impact on the total freshwater consumed in the United States is
phosphate rock increased from $27.89 to $96.90 per ton [86,88]. also evident as algae production is increased to 10 BGY, foreshadowing
While chemical processes can produce synthetic nitrogen fertilizers, potential water versus fuel debate by reaching levels equivalent to 14%
this process does not provide a sustainable outlook for algal biofuels of all current U.S. freshwater consumption. The scale up effects of this
due to the fossil fuel demand of the Haber–Bosch process. It is critical algae facility pertaining to total freshwater consumption in New
to algae biofuels to continue to consider growth conditions that utilize Mexico [97] and the United States [98] can be seen in Table A.6.
recycled nitrogen and phosphorus sources. In 1990, of the estimated 2.467 × 1013 m3 of groundwater reserves
Efforts must continue to identify methods that will lower the costs and in New Mexico, 1.85 × 1013 m3 were characterized as moderately saline,
improve the sustainability associated with nutrient consumption. Inex- very saline, or brine. Only Tularosa Basin and Crows Flats were deemed
pensive sources of CO2 such as from a flue gas provide an appealing option suitable for large-scale microalgal cultivation based on the selection
for culture enrichment. Nutrient-rich wastewater can supplement or criteria established by the Solar Energy Research Institute (SERI)
eliminate the need for additional N and P and the recovery of these vital (water quality, land slope, and climatic conditions) [41]. The selected re-
elements through anaerobic digestion of algae biomass offers another im- gions were found to contain about 2.84–6.17 × 109 m3 of available sa-
portant option to reclaim the fertilizer components used in algal biofuel line water with just over 39,000 ha of suitable land [41,99]. Studies
production [89,90]. This flue-gas and wastewater co-utilization (FWC) have been performed to determine the salinity tolerance of algae for bio-
strategy will not only lower costs of raw materials but also offset waste fuel production as utilizing saline water can help to curb egregious fresh-
products and contribute to the sustainability of microalgal biofuels [91]. water requirements [41,99]. Maintaining all previous assumptions for a
Despite the benefits of FWC, the maximum potential of algae production marine algal species, the land available would allow for the
from FWC may still only make a moderate impact on United States petro- construction of eight 1000 bbl d−1 facilities at 4875 ha each, requiring
leum dependence for transportation fuels. A GIS based national assess- from 229–870 billion liters of water annually for the covered and uncov-
ment considering the relative abundance of the input resources and ered raceway designs, respectively. The facilities in the current model
proximity for the economic and biofuel production potential of FWC would therefore be able to produce an estimated 463.3 million liters
found that less than 200 MGY (757 MLY) could be produced annually (122.4 million gallons) of oil annually based on the 1990 SERI resource as-
[92]. Additional strategies for increasing biofuel production must include sessment. More current resource assessments have indicated that addition-
nutrient recycling and the utilization of nutrients from livestock waste al saline or “brackish” water resources may be available to facilitate
[93]. additional development algal biofuels in New Mexico, however favor loca-
tions situated around the Gulf Coast, the eastern seaboard, and areas adja-
4.7. Land and water requirements: covered vs. uncovered raceways cent to the Great Lakes to meet water requirements [6].
The low seasonal variation, high solar irradiance, and flat topography
Algal production is heavily dependent on large volumes of water for in New Mexico make it a valuable location for the development of algal
production. The costs and energy associated with replenishing lost water biofuels, however groundwater depletion in New Mexico is already a se-
to the process were found to be highly dependent on evaporative losses rious issue and increased withdrawals could further aggravate freshwa-
and groundwater levels. This section assesses the sustainability and pro- ter scarcity in the state. Therefore, algal biofuel development efforts in
ductivity of scaling-up an algal-facility as it pertains to water availability. the state must use algal strains with the ability to utilize saline or “brack-
Despite the reductions in COP for uncovered raceways, the water ish” water. Scaling algal biofuel production in New Mexico must focus on
requirements pose serious sustainability issues when scaling up produc- reducing water losses and maximizing available resources. From an envi-
tion. For equivalent water availability of a plant using covered raceways ronmental engineering perspective, the water reduction and sustainabil-
can support 4.2 times the production capacity as a plant operating with ity gained by using covered raceways most likely outweighs the increase
the uncovered design. While evaporative loss of culture volume can seri- in cost of production. Covering the raceway ponds will also help to mit-
ously affect production due to water availability in arid regions, evapora- igate contamination and environmental damage however it must be de-
tive cooling plays an important and necessary role in maintaining signed to facilitate cooling and allow for optimal growth.
appropriate pond temperature [6]. Water loss in the range of 0.3–
0.6 cm d−1 [6] is not uncommon and accounts for a significant hurdle 5. Conclusions
to sustainable algal biofuel production [25]. Thus, a balance must be
struck between the natural discharge of heat by evaporation and the rec- The current study evaluates sustainability and economic require-
lamation of this valuable water that is lost in the process [94,95]. Addi- ments of a 1000 bbl d−1 algal biofuel facility based in New Mexico.
tional research must account for seasonal variations to optimize growth Several concerns arose during the development of this model; however,
conditions without jeopardizing plant sustainability due to evaporative alternatives and solutions exist and must continue to be developed in
losses [96]. order to realize the full potential of algal biofuels. Current technology
As a preliminary metric, it was desired to evaluate freshwater re- and production techniques limit the scalability of algal biofuel produc-
quirements of the current model and scale up compared to the total tion. Energy, water, CO2, and nutrient requirements in particular
freshwater consumption in both New Mexico and the United States as represent significant obstacles to the sustainability of algal biodiesel
a whole. The 2005 New Mexico Water Report found that 4.88 trillion - at the industrial scale. A critical analysis of the available resources
liters (1.29 trillion gallons) of freshwater are consumed every year in New Mexico has shown that these challenges of water, CO2,
[97]. The state's freshwater is extremely limited, of which 77.86% is and nutrient sourcing at large-scale production must be carefully
used agriculturally. Due to limited supply, both fresh and slightly saline considered. Development must focus on reducing water losses
waters are now used for domestic, industrial, and agricultural purposes. and maximizing available saline and “brackish” water resources.
Water rights holders are also unlikely to support the demands of an Multiple pathways are available for producing fuels from algae and addi-
increase in algal production [41]. In 2005, the United States freshwater tional research needs to be done to develop the most economic and sus-
consumption totaled 482.3 trillion liters (127.4 trillion gallons) annual- tainable solutions. Both brackish water tolerant freshwater and marine
ly [98]. Assuming sufficient land for production, it is evident algae strains must be cultivated using various water sources to identify
that the freshwater required to build facilities in New Mexico the strains ideally suited for production under a variety of conditions.
rapidly matches and surpasses all other freshwater consumption in the Several options are available to mitigate the sustainability challenges
state. The 5 BGY goal is completely unattainable in New Mexico alone that were identified by the present model. Flue-gas and wastewater
J.N. Rogers et al. / Algal Research 4 (2014) 76–88 85

utilization will help to supplement some of the nutrient demands Appendix A. Tables
of algal growth. Integration with livestock production will provide
nutrients to the system that can be used for algal production; however,
Table A.1
locating these facilities in close proximity also depends on the available New Mexico groundwater depth measurements.
land and resources. The systems that are the most ideal for integration
New Mexico groundwater depth measurements [104]
with algal production (flue gas sources, WWT, etc.) already exist
for the large part. Therefore, they must be evaluated for the feasibility National water information system
of incorporating algae growth into their facilities [91]. Future develop- Site number Depth ft
312323108293302 265
ment of algal biofuels will be best served if new sources of unconven-
312503108332801 400
tional nutrients are developed in conjunction with algal growth in an 312705108304801 520
integrated biorefinery model [20,93]. Additionally, despite the potential 312722108350401 250
economic impact of utilizing LEA as an animal feed supplement, nutri- 312804108332301 300
312951108295101 392
ent sustainability concerns may necessitate the use of LEA for nutrient
320104103120301 500
recycle. Average 375 ft = 114 m
Freshwater availability represents a serious issue to algal biofuel
production without robust water reclamation methods and evaporative
loss minimization. Utilizing saline or brackish water will be necessary to
sustainably grow algae in New Mexico [95]. Depleting ground water Table A.2
depth levels will increase the pumping energy requirements to fill and Baseline operating cost results for a single plant. Comparison of annual costs and
contribution to COP in $ gal−1 and $ L−1. Annual production: 58.03 MLY (15.33 MGY).
maintain the ponds. Efforts must focus on reducing evaporative losses,
recovering water during processing, and maximizing available saline Baseline operating costs
and “brackish” water resources. $MM yr−1 Cost $ gal−1 Cost $ L−1
The disparity between the theoretical and measured paddlewheel
Paddlewheels $10.07 $0.66 $0.17
energy use reflects the uncertainty in scale-up and demonstrates the Lamella separator $0.01 $0.00 $0.00
importance of minimizing energy requirements. Paddlewheel-driven Centrifuge $0.64 $0.04 $0.01
raceway ponds have historically been the method of choice for Cell disruption $0.20 $0.01 $0.00
algae cultivation due to their presumed low-cost compared to more Biomass-separator $0.01 $0.00 $0.00
Hexanes recovery $2.75 $0.18 $0.05
sophisticated photobioreactor designs. However, raceway ponds are On-site pumping circulation $5.76 $0.38 $0.10
severely limited by their operating efficiencies and physical aspect ra- DAP $7.36 $0.48 $0.13
tios to promote maximal algal biomass productivity. While these Urea $24.43 $1.59 $0.42
ponds can accomplish the objective of tertiary wastewater treatment Pumping replacement water $2.07 $0.13 $0.04
Replacement hexane $0.40 $0.03 $0.01
with microalgae at small scales, the capital costs, operating costs,
Absorber CO2 $11.21 $0.73 $0.19
and energy requirements of paddlewheels at large scales indicate the Flocculant $26.70 $1.74 $0.46
need for alternative bioreactor designs for algal biofuel production Wastewater treatment $1.10 $0.07 $0.02
[100–102]. Maintenance $28.55 $1.86 $0.49
Algae growth and harvesting costs dominate the economic invest- Wages $13.05 $0.85 $0.22
Total operating $134.30 $8.76 $2.31
ment in facility scale-up. The amount of growth area required for pro-
duction correlates directly to the largest capital expenditures. A pond
liner is a necessity for the majority of geographic locations suitable for
algae growth and a partial cover will reduce evaporative loss while Table A.3
allowing necessary cooling to occur. Cultivating strains with a more Baseline capital cost results for a single plant. Total investment and cost contribution to
rapid growth rates and augmented total lipid contents have the highest COP in $ gal−1 and $ L−1 for 10-yr and 20-yr capital return scenarios. Annual production:
potential to decrease COP as these biological parameters increase pro- 58.03 MLY (15.33 MGY).
ductivity without imposing additional capital investment to the growth Baseline capital costs
infrastructure [103].
$MM 10 yr 10 yr 20 yr 20 yr
Collectively, the modeled data pertaining to energy, nutrient, and $ gal−1 $ L−1 $ gal−1 $ L−1
water requirements for expansive deployment of microalgal biofuel
CO2 absorber (static mixer) $0.17 $0.001 $0.000 $0.001 $0.000
production in New Mexico pose certain challenges but are likely to be Paddlewheels $120.00 $0.783 $0.207 $0.391 $0.103
addressable with technological advances in algae cultivation methods. Flocculant addition (static mixer) $0.11 $0.001 $0.000 $0.000 $0.000
The continued development of mitigation strategies for each of these Lamella separators $2.50 $0.016 $0.004 $0.008 $0.002
concerns will build a foundation to withstand the inevitable and neces- Centrifuges $6.60 $0.043 $0.011 $0.022 $0.006
Cell disruption (sonicators) $3.25 $0.021 $0.006 $0.011 $0.003
sary scrutiny for emerging energy technologies, such as algal biofuels.
Lipid extraction (static mixer) $0.00 $0.000 $0.000 $0.000 $0.000
Ongoing research in this active area of renewables will need to refine Biomass-solvent separator $0.25 $0.002 $0.000 $0.001 $0.000
algae production systems in order to gain both public acceptance and Distillation column $0.50 $0.003 $0.001 $0.002 $0.000
secure the economic viability of biofuels. Pumps $1.27 $0.008 $0.002 $0.004 $0.001
Operating water $2.96 $0.019 $0.005 $0.010 $0.003
Hexane $1.05 $0.007 $0.002 $0.003 $0.001
Wastewater treatment equipment $245.00 $1.598 $0.422 $0.799 $0.211
Acknowledgements Land $18.00 $0.117 $0.031 $0.059 $0.016
Liners $250.00 $1.631 $0.431 $0.815 $0.215
This work was supported, in part, by funds from the U.S. Department Landscaping $57.00 $0.372 $0.098 $0.186 $0.049
Cover $315.00 $2.055 $0.543 $1.027 $0.271
of Energy’s National Energy Technologies Laboratory through grant
Railways $1.25 $0.008 $0.002 $0.004 $0.001
number DE-OE0000098 to New Mexico State University. Partial support Roads $0.60 $0.004 $0.001 $0.002 $0.001
was also provided by a fellowship to JNR from the Johns Hopkins Envi- Operating buildings $0.45 $0.003 $0.001 $0.001 $0.000
ronment, Energy, Sustainability & Health Institute (E²SHI). We appreci- Construction costs $2.50 $0.016 $0.004 $0.008 $0.002
ate the constructive input from Dr. Scott Williams and his generous help Pipes $7.08 $0.046 $0.012 $0.023 $0.006
Total capital $1035.55 $6.76 $1.78 $3.38 $0.89
reviewing this manuscript.
86 J.N. Rogers et al. / Algal Research 4 (2014) 76–88

Table A.4 Appendix B. Calculations


Cost of production sensitivity analysis ($ per liter). Estimates are based on the baseline
operating costs of $134 MM and annual capital payments of $52 MM to cover a 20 year
return of the $1.04 billion capital investment.
B.1. Energy requirements for raceway ponds [56]

Cost of production sensitivity analysis ($ per liter) The head loss in bends is calculated by,
10 yr baseline COP 20 yr baseline COP
$4.10 L−1 $3.21 L−1
 
2
K v
COP COP COP COP hb ¼ ðB:1:1Þ
decrease increase decrease increase 2g

Lipid content (35%: 25%: 15%) −$1.17 $2.73 −$0.92 $2.14


Extraction efficiency (100%: 80%: 60%) −$0.82 $1.37 −$0.64 $1.07 in which K is the kinetic loss coefficient for 180° bends
Areal productivity g/m2/d (20: 15: 10) −$1.02 $2.05 −$0.80 $1.60 (theoretically = 2), v is the velocity of the raceway (0.3 m s−1), and
Pond liner −$0.43 $0.43 −$0.22 $0.22 g is the acceleration due to gravity (9.8 m s−2). Resulting in
(no liner: liner: replace liner once)
hb = 0.01834.
Cover −$0.54 $0.54 −$0.27 $0.27
(no cover: cover: replace cover once)
The friction loss across the length of the raceway is calculated using
Paddle wheels – $0.42 – $0.42 Manning's Equation,
(0.22 W m-2: 0.73 W m-2)
Net harvesting efficiency (100%: 85%) – $0.72 – $0.57  
2 2 L
Net harvesting costs (50%: base: 150%) −$0.24 $0.24 −$0.24 $0.24 hc ¼ v  n  ðB:1:2Þ
R =3
4
Flocculant ($50: $100: $150 ton-algae−1) −$0.23 $0.23 −$0.23 $0.23
$ ton−1 pure CO2 ($20: $40: $60) −$0.10 $0.29 −$0.10 $0.29
N + P recycle (100%: 0%) −$0.55 – −$0.55 – in which n is the roughness factor (0.015 for polyethylene), R is the
Extraction costs −$0.03 $0.23 −$0.03 $0.23
channel hydraulic radius (2112), and L is the channel length (643.9).
(50%: base: dry algae to 90% DW)
Evaporation rate (base 0.05: 0.5 cm d−1) – $0.28 – $0.28 Resulting in hc = 0.46635.
On-site pumping capacity – $0.10 – $0.10 The energy requirement per pond is calculated by,
(base 20 ft: 40 ft head)
 
Q wh
W ¼ 9:8  ðB:1:3Þ
Table A.5
e
Cost of production sensitivity analysis ($ per gallon). Estimates are based on the baseline
operating costs of $134 M and annual payments of $104 M to cover a 10 year return of in which Q is the volumetric flow rate (1.1 m3 s−1), w is the unit mass of
the $1.04 billion capital investment.
water (998 kg m3), h is the total head loss (htotal = 0.64984), e is the
Cost of production sensitivity analysis ($ per gallon) paddle wheel and drive system efficiency (40% assumed), and 9.8 is
10 yr baseline COP 20 yr baseline COP the conversion factor in W-s kg-m−1. The energy calculations resulted
$15.52 gal−1 $12.14 gal−1 in 1741.5 W pond− 1 (0.22 W m− 2) or 10,448.8 kW for the whole
COP COP COP COP
plant. When operated continuously the total paddlewheel energy use
decrease increase decrease increase totals 91.531 M kWh annually.
Lipid content (35%: 25%: 15%) −$4.43 $10.34 −$3.47 $8.09
Extraction efficiency (100%: 80%: 60%) −$3.10 $5.17 −$2.43 $4.05 B.2. Energy required for solvent recovery and drying algae to 10% water
Areal productivity g/m2/d (20: 15: 10) −$3.88 $7.76 −$3.03 $6.07 content
Pond liner (no liner: liner: replace liner once) −$1.63 $1.63 −$0.82 $0.82
Cover −$2.05 $2.05 −$1.03 $1.03
(no cover: cover: replace cover once) cp  w  ðT b −20Þ
Paddle wheels – $1.59 – $1.59
E¼ ðB:2:1Þ
3600
(0.22 W m-2: 0.73 W m-2)
Net harvesting efficiency (100%: 85%) – $2.74 – $2.14
Net harvesting costs (50%: base: 150%) −$0.92 $0.92 −$0.91 $0.91 For solvent recovery: E is the energy in kWh, cp is the specific heat
Flocculant ($50: $100: $150 ton-algae−1) −$0.87 $0.87 −$0.87 $0.87
of hexane (2.26 KJ kg − 1 °C − 1 [105]), w is the number of kg of
$ ton−1 pure CO2 ($20: $40: $60) −$0.37 $1.10 −$0.37 $1.10
N + P recycle (100%: 0%) −$2.07 – −$2.07 – hexane entering the column every hour (92,696 kg hr− 1), Tb is the
Extraction costs −$0.11 $0.88 −$0.11 $0.88 boiling point of hexane (69.0 °C [106]), 20 is the starting tempera-
(50%: base: dry algae to 90% DW) ture of the hexane, and 3600 is a conversion factor from KJ to kWh.
Evaporation rate (0.05: 0.5 cm d−1) – $1.06 – $1.06 Under continuous operation the column will require 25.0 M kWh
On-site pumping capacity (20 ft: 40 ft head) – $0.38 – $0.38
annually.
For drying algae to 10% water content: E is the energy in kWh, cp is
Table A.6 the specific heat of water (4.18 KJ kg− 1 °C− 1), w is the number of kg
Scale up assessment of algae freshwater requirements. Compares the water requirements of water entering the dryer every hour (150,290 kg hr− 1), Tb is the
when scaling the model for two evaporation conditions to the total freshwater boiling point of water (100.0 °C), 20 is the starting temperature
consumption in New Mexico and the United States.
of the water, and 3600 is a conversion factor from KJ to kWh.
Scale up assessment of algae freshwater requirements Under continuous operation the column will require 122.3 M kWh
New Mexico annual freshwater consumption (BGY): 1287.2 annually.
United States annual freshwater consumption (BGY): 127,385.0
1 Plant (15.3 MGY production) Water consumption (BGY) % of NM % of US B.3. Energy required for pumping groundwater and on-site circulation
Covered (0.05 cm d−1 evaporation) 6.57 0.5% 0.005%
[107]
Uncovered (0.5 cm d−1 evaporation) 27.73 2.2% 0.022%
327 Plants (5 BGY production)
Covered (0.05 cm d−1 evaporation) 2160 167.7% 1.7%
E ¼φW h ðB:3:1Þ
Uncovered (0.5 cm d−1 evaporation) 9060 703.7% 7.1%
654 Plants (10 BGY production)
Covered (0.05 cm d−1 evaporation) 4320 335.3% 3.4% in which E represents the total energy in Watt hours, W is the pumping
Uncovered (0.5 cm d−1 evaporation) 18,120 1407.4% 14.2%
out W cubic meters of groundwater (or on-site circulation), h is the
J.N. Rogers et al. / Algal Research 4 (2014) 76–88 87

average groundwater depth (or m of head), and φ is a coefficient de- [23] B.J.L.B. Williams, L.M.L. Laurens, Microalgae as biodiesel & biomass feedstocks: re-
view & analysis of the biochemistry, energetics & economics, Energy Environ. Sci. 3
fined by: (2010) 554–590.
[24] M.E. Huntley, D.G. Redalje, CO2 mitigation and renewable oil from photosynthetic
microbes: a new appraisal, Mitig. Adapt. Strateg. Glob. Chang. 12 (4) (2007)
γρg 573–608.
φ¼ ðB:3:2Þ
1000 [25] NationalResearch Council, Sustainable Development of Algal Biofuels, National
Academies Press, 2012, ISBN 978-0-309-26032-9. (Available at: http://www.nap.
edu/catalog.php?record_id=13437).
in which γ is the pumping efficiency (0.5), ρ is the density of water [26] A. Kendall, J. Yuan, Comparing life cycle assessments of different biofuel options,
Curr. Opin. Chem. Biol. (2013).
(1000 kg m− 3), and g is the acceleration due to gravity (9.8 m s−2). [27] Maintenance Estimate, Retrieved December 13, 2012 from http://www.
The current model assumes a groundwater depth of 114 m and two reliableplant.com/Read/7050/cost-replacement-value.
conditions for replacement water of 28,000 m3 (0.05 cm d−1 evapora- [28] EISA, Mandates Renewable Fuel Standard, Retrieved April 10th, 2013 from http://
www.epa.gov/otag/fuels/renewablefuels/.
tion) and 247,450 m3 (0.5 cm d−1 evaporation). The two replacement [29] Monroe Environmental, Retrieved December 13, 2012 from http://www.
conditions require 18.8 M kWh yr−1 and 165.9 M kWh yr−1 respec- monroeenvironmental.com/PDF/ME_Circular_Clarifiers.pdf.
tively. On-site pumping circulates 1.46 M m3 d−1 for the two scenarios [30] WEF, Clarifier Design, WEF Manual of Practice No. FD-8, 2nd ed., McGraw-Hill,
2005.
of 6 m and 12 m head. The two pumping capacities are calculated to re-
[31] Max S. Peters, et al., Plant Design and Economics for Chemical Engineers, 5th ed.
quire 52.3 M kWh yr−1 and 104.6 M kWh yr−1 respectively. McGraw-Hill, 2002. 793.
[32] Henri P. Sangam, R. Kerry Rowe, Effects of exposure conditions on the depletion of
antioxidants from high-density polyethylene (HDPE) geomembranes, Can.
Appendix C. Supplementary data
Geotech. J. 39 (6) (2002) 1221–1230.
[33] EPA, East Helena Capital Costs for Lands to be Developed, Retrieved December 13,
Supplementary data to this article can be found online at http://dx. 2012 from http://www.epa.gov/region8/superfund/mt/easthelena/03_CostTables.
doi.org/10.1016/j.algal.2013.11.007. pdf.
[34] J.N. Rosenberg, A. Mathias, K. Korth, M.J. Betenbaugh, G.A. Oyler, Microalgal bio-
mass production and CO2 sequestration from an integrated ethanol biorefinery
References in Iowa: a technical appraisal and economic feasibility evaluation, Biomass
Bioenergy 35 (2011) 3865–3876.
[1] J. Chamie, et al., World Population to 2300, Department of Economic and Social [35] Mike Good, Railroad Construction Cost Estimates, Henry Country Library, 2008.
Affairs, Population Division, ST/ESA/SER.A/236, United Nations, New York, NY, (Retrieved December 13, 2012 from, http://tacnet.missouri.org/history/railroads/
2004. rrcosts.html).
[2] Annual Energy Outlook, U.S. Energy Information Administration, DOE/EIA- [36] Generic Cost Per Mile Models, Retrieved December 13, 2012 from http://
0383(2012) June 25, 20122012. (www.eia.gov/forecasts/aeo). citizensforsouthernillinois.org/?page_id=722012.
[3] A. Richmond, Handbook of Microalgal Culture: Biotechnology and Applied Phycol- [37] Iowa State University Extension and Outreach. “Biodiesel Profitability”. Unpub-
ogy, Wiley-Blackwell, 2008. lished result retrieved December 13, 2012 from, http://www.extension.iastate.
[4] J. Sheehan, et al., A Look Back at the US Department of Energy's Aquatic Species edu/agdm/energy/xls/d1-15biodieselprofitability.xlsx.
Program: Biodiesel from Algae. TP-580-24190, National Renewable Energy Labora- [38] Harrison Machine & Plastic Co., Retrieved December 13, 2012 from http://www.
tory, Golden, CO, 1998. harrisonplastic.com/aboutpvc.html.
[5] E.D. Frank, et al., “Life-Cycle Analysis of Algal Lipid Fuels with the GREET Model.” [39] Mueller Industries, Wholesale Pipe Price Lists, Retrieved December 13, 2013 from
ANL/ESD/11-5, Center for Transportation Research, Energy Systems Division, Argonne http://www.muellerindustries.com/catalogs/wholesale-price-lists.
National Laboratory, Oak Ridge, 2011. [40] Lands of New Mexico, Retrieved May 5, 2013 from http://www.landsofnewmexico.
[6] R. Davis, D. Fishman, E.D. Frank, M.S. Wigmosta, et al., Renewable Diesel from com/new_mexico/index.cfm2013.
Algal Lipids: An Integrated Baseline for Cost, Emissions, and Resource Potential [41] R. Lansford, J. Hernandez, P.J. Enis, D. Truby, C. Mapel, Evaluation of available saline
from a Harmonized Model, Argonne National Laboratory, Technical Report: water resources in New Mexico for the production of microalgae, Report, Solar
ANL/ESD/12-4, NREL/TP-5100-55431, PNNL-214372012. (www.nrel.gov/docs/ Energy Research Institute, Golden, Colorado, SERI/TP 232-3597, 1990, (83 pp.).
fy12osti/55431.pdf). [42] All Cost Data, Turbine Pumps, Cast Iron, Retrieved December 13, 2013 from http://
[7] John S. Burlew, Algal culture, From Laboratory to Pilot Plant, Carnegie Inst. www.allcostdata.info/browse.html/152901000.
WashingtonPubl 600, 1953. 1. [43] Pentair Co., Propeller Pumps, Retrieved December 13, 2013 from http://www.
[8] W.J. Oswald, C.G. Golueke, Biological transformation of solar energy, Adv. Appl. fairbanksnijhuis.com/EngineeredProduct_N_Propeller.aspx.
Microbiol. 11 (1960) 223–242. [44] Oakland County Water and Wastewater Masterplan, Alternative Analysis, vol. 4,
[9] R.E. Lee, Phycology, Cambridge University Press, New York, 1980. Sept 2007, (URS Project No. 13647263).
[10] Dragoljub Bilanovic, et al., Freshwater and marine microalgae sequestering of [45] Robert E. Blankenship, et al., Comparing photosynthetic and photovoltaic ef-
CO2 at different C and N concentrations—Response surface methodology analysis, ficiencies and recognizing the potential for improvement, Science (2011)
Energy Convers. Manag. 50 (2) (2009) 262–267. 805–809(332.6031).
[11] Sheng-Yi Chiu, et al., Lipid accumulation and CO2 utilization of Nanochloropsis [46] A. Sun, R. Davis, M. Starbuck, A. Ben-Amotz, R. Pate, P.T. Pienkos, Comparative cost
oculata in response to CO2 aeration, Bioresour. Technol. 100 (2) (2009) 833–838. analysis of algal oil production for biofuels, Energy 36 (8) (2011) 5169–5179.
[12] Bei Wang, et al., CO2 bio-mitigation using microalgae, Appl. Microbiol. Biotechnol. [47] Index Mundi, DAP Fertilizer vs Urea, Retrieved December 13, 2012 from http://
79 (5) (2008) 707–718. www.indexmundi.com/commodities/?commodity=dap-
[13] A.C. Redfield, On the proportions of organic derivatives in sea water and their rela- fertilizer&commodity=urea.
tion to the composition of plankton, in: R.J. Daniel (Ed.), James Johnstone Memorial [48] Craig Behnke, Sapphire Energy, Inc., Development of Value-Added Products
Volume, University Press of Liverpool, 1934, pp. 177–192. from Residual Algae Biomass, DOE EERE Biomass Peer Review (May 20-23, 2013,
[14] A.C. Redfield, The biological control of chemical factors in the environment, Am. Alexandria, VA), Slide 7, May 20–23, 2013, Accessed September 11, 2013. Available
Sci. 46 (1958) 205–221. from: https://www2.eere.energy.gov/biomass/peer_review2013/Portal/.
[15] Adam C. Martiny, et al., Strong latitudinal patterns in the elemental ratios of ma- [49] Office of EERE, Bioenergy Technologies At-a-Glance, Retrieved May 23, 2013 from
rine plankton and organic matter, Nat. Geosci. 6 (4) (2013) 279–283. http://www1.eere.energy.gov/office_eere/bo_budget_fy14.html.
[16] David T. Welsh, et al., Denitrification, nitrogen fixation, community primary pro- [50] Pascal Schlagermann, et al., Composition of Algal Oil and Its Potential as Biofuel,
ductivity and inorganic-N and oxygen fluxes in an intertidal Zostera noltii meadow, J. Combust. (2012)(Article ID 285185) (http://www.hindawi.com/journals/jc/
Mar. Ecol. Prog. Ser. 208 (2000) 65–77. 2012/285185/abs/).
[17] José Moreno, et al., Outdoor cultivation of a nitrogen-fixing marine cyanobacteri- [51] Yusuf Chisti, Biodiesel from microalgae, Biotechnol. Adv. 25 (3) (2007)
um, Anabaena sp. ATCC 33047, Biomol. Eng. 20 (4) (2003) 191–197. 294–306.
[18] Tom Berman, Sara Chava, Algal growth on organic compounds as nitrogen sources, [52] Maria J. Barbosa, et al., Optimization of biomass, vitamins, and carotenoid yield on
J. Plankton Res. 21 (8) (1999) 1423–1437. light energy in a flat-panel reactor using the A-stat technique, Biotechnol. Bioeng.
[19] John R. Benemann, William J. Oswald, Systems and economic analysis of 89 (2) (2005) 233–242.
microalgae ponds for conversion of CO2 to biomass, Final Report. No. DOE/ [53] Otto Pulz, Karl Scheibenbogen, Photobioreactors: design and performance with re-
PC/93204–T5, California Univ, Berkeley, CA (United States), 1996, (Dept. of Civil spect to light energy input, Bioprocess and Algae Reactor Technology, Apoptosis,
Engineering). Springer, Berlin Heidelberg, 1998. 123–152.
[20] Rupert Craggs, Donna Sutherland, Helena Campbell, Hectare-scale demonstration [54] Jeffrey M. Gordon, Juergen E.W. Polle, Ultrahigh bioproductivity from algae, Appl.
of high rate algal ponds for enhanced wastewater treatment and biofuel produc- Microbiol. Biotechnol. 76 (5) (2007) 969–975.
tion, J. Appl. Phycol. 24 (3) (2012) 329–337. [55] M. Janssen, P. Slenders, J. Tramper, L.R. Mur, R.H. Wijffels, Photosynthetic efficiency
[21] Venkatesh Vasudevan, et al., Environmental performance of algal biofuel technol- of Dunaliella tertiolecta under short light/dark cycles, Enzyme Microb. Technol. 29
ogy options, Environ. Sci. Technol. 46 (4) (2012) 2451–2459. (4–5) (2001) 298–305.
[22] M. Jensen, A.H. Andersen, Biofuels: a contested response to climate change, [56] T.J. Lundquist, et al., A Realistic Technology and Engineering Assessment of Algae
Sustainability Sci. Pract. Policy 9 (1) (2013) 42–56. Biofuel Production, Energy Biosciences Institute, 2010. 1.
88 J.N. Rogers et al. / Algal Research 4 (2014) 76–88

[57] DARPA, Defense Sciences Office. Biofuels, Retrieved April 30, 2013 from http:// [83] EPA, CO2 Emissions Estimates, Retrieved April 26th, 2013 from http://www.epa.
www.darpa.mil/Our_Work/DSO/Programs/Biofuels.aspx. gov/greenpower/pubs/calcmeth.htm.
[58] Rupert J. Craggs, Paul J. McAuley, Valerie J. Smith, Wastewater nutrient removal by [84] K.L. Kadam, Power plant flue gas as a source of CO2 for Microalgae cultivation: eco-
marine microalgae grown on a corrugated raceway, Water Res. 31 (7) (1997) nomic impact of different process options, Energy Convers. Manag. 38 (1997)
1701–1707. 505–510(Suppl.).
[59] Ami Ben-Amotz, Large scale open algae ponds, Technical report, The National Insti- [85] W. Huang, Impact of Rising Natural Gas Prices on U.S. Ammonia Supply, USDA
tute of Oceanograpy, Israel, 2008. WRS-0702August 2007.
[60] S. Hellwig, J. Drossard, R.M. Twyman, R. Fischer, Plant cell cultures for the produc- [86] U.S. Geological Survey, Mineral Commodity Summaries, January 2013.
tion of recombinant proteins, Nat. Biotechnol. 22 (2004) 1415–1422. [87] S.M. Jasinski, Phosphate Rock, Statistics and Information, US Geological Survey,
[61] N.R. Mann, R.S. Herbst, J.R. Hess, V. Kochergin, Enhanced flux, erosion resistant 2006.
membranes for biomass separations—Part II, Membr. Technol. 2004 (2004) 5–9. [88] U.S. Geological Survey, Mineral Commodity Summaries, January 2006.
[62] A. Schlesinger, D. Eisenstadt, A. Bar-Gil, H. Carmely, S. Einbinder, J. Gressel, Inex- [89] P. Bohutskyi, E. Bouwer, Biogas production from algae and cyanobacteria through
pensive non-toxic flocculation of microalgae contradicts theories; overcoming a anaerobic digestion: a review, analysis, and research needs, Adv. Biofuels Bioprod.
major hurdle to bulk algal production, Biotechnol. Adv. 30 (2012) 1023–1030. (2013) 873–975.
[63] B. Ghernaouta, D. Ghernaoutb, A. Saibabpages, Algae and cyanotoxins removal by [90] S. Aslan, I.K. Kapdan, Batch kinetics of nitrogen and phosphorus removal from syn-
coagulation/flocculation: a review, Desalin. Water Treat. 20 (2010) 133–143. thetic wastewater by algae, Ecol. Eng. 28 (1) (2006) 64–70.
[64] N.B. Wyatt, L.M. Gloe, P.V. Brady, J.C. Hewson, A.M. Grillet, M.G. Hankins, P.I. Pohl, [91] Shaikh A. Razzak, et al., Integrated CO2 capture, wastewater treatment and biofuel
Critical conditions for ferric chloride-induced flocculation of freshwater algae, production by microalgae culturing—a review, Renew. Sust. Energy Rev. 27 (2013)
Biotechnol. Bioeng. 109 (2) (2012) 493–501. 622–653.
[65] G. Gutzeit, D. Lorch, A. Weber, M. Engels, U. Neis, Bioflocculent algal-bacterial biomass [92] Nolan Orfield, Greg Keoleian, Nancy Love, A GIS based national assessment of algal
improves low-cost wastewater treatment, Water Sci. Technol. 52 (12) (2005) 9–18. biofuel production potential through flue-gas and wastewater coutilization, 6th In-
[66] S. Salim, R. Bosma, M.H. Vermuë, R.H. Wijffels, Harvesting of microalgae by ternational Conference of the International Society for Industrial Ecology (ISIE)
bio-flocculation, J. Appl. Phycol. 23 (5) (2010) 849–855. Proceedings. Berkeley, CA, June 7–10 2011, Abstract #591, 2011.
[67] S. Salim, M.H. Vermuë, R.H. Wijffels, Ratio between autoflocculating and target [93] Keri B. Cantrell, et al., Livestock waste-to-bioenergy generation opportunities,
microalgae affects the energy-efficient harvesting by bio-flocculation, Bioresour. Bioresour. Technol. 99 (17) (2008) 7941–7953.
Technol. 118 (2012) 49–55. [94] E.R. Venteris, R. Skaggs, A.M. Coleman, M.S. Wigmosta, A GIS cost model to as-
[68] Method and System for Efficient Harvesting of Microalgae and Cyanobacteria, sess the availability of freshwater, seawater, and saline groundwater for algal
WIPO Patent No. 2011040955, 8 Apr. 2011. biofuel production in the United States, Environ. Sci. Technol. 47 (9) (2013)
[69] G.A. Oyler, J.N. Rosenberg, D.P. Weeks, Single Chain Antibodies for Photosynthetic 4840–4849.
Microorganisms and Method of Use, US 2012/032662, 2012. [95] Yuan-Kun Lee, Microalgal mass culture systems and methods: their limitation and
[70] T.J. Lundquist, D. Hammond, F.B. Green, W.J. Oswald, Tertiary Pilot Study Report, potential, J. Appl. Phycol. 13 (4) (2001) 307–315.
Recycled Wastewater Facilities Plan, Oswald Engineering Associates, Inc., City of [96] Peter Waller, et al., The algae raceway integrated design for optimal temperature
St. Helena, California, 2004. 70. management, Biomass Bioenergy 46 (2012) 702–709.
[71] J.N. Rosenberg, G.A. Oyler, L. Wilkinson, M.J. Betenbaugh, A green light for [97] John Longworth, New Mexico water use by categories 2005, New Mexico Office of
engineered algae: redirecting metabolism to fuel a biotechnology revolution, the state Engineer Technical Report 52, 2005, (http://www.ose.state.nm.us/PDF/
Curr. Opin. Biotechnol. 19 (2008) 430–436. Publications/Library/TechnicalReports/TechReport-052.pdf).
[72] K.J. Flynn, A. Mitra, H.C. Greenwell, J. Sui, Monster potential meets potential monster: [98] J.F. Kenny, N.L. Barber, S.S. Hutson, K.S. Linsey, J.K. Lovelace, M.A. Maupin, Estimat-
pros and cons of deploying genetically modified microalgae for biofuels production, J. ed Use of Water in the United States in 2005: U.S. Geological Survey Circular, 2009.
R. Soc. Interface 3.1 (2013)(http://www.ncbi.nlm.nih.gov/pmc/issues/222341/). 1344 (52 pp.).
[73] K.J. Flynn, H.C. Greenwell, R.W. Lovitt, R.J. Shields, Selection for fitness at the indi- [99] R. Lansford, J. Hernandez, P.J. Enis, Evaluation of available saline water resources in
vidual or population levels: modeling effects of genetic modifications in New Mexico for the production of microalgae, FY 1986 Aquatic Species Program
microalgae on productivity and environmental safety, J. Theor. Biol. 263 (3) Annual Report, Solar Energy Research Institute, Golden, Colorado, 1987,
(2010) 269–280. pp. 227–248, (SERI/SP-231–3071).
[74] J. Gressel, C.J.B. van der Vlugt, H.E.N. Bergmans, Environmental risks of large [100] B. Ketheesan, N. Nirmalakhandan, Feasibility of microalgal cultivation in a
scale cultivation of microalgae: mitigation of spills, Algal Res. 2 (2013) pilot-scale airlift-driven raceway reactor, Bioresour. Technol. 108 (2012) 196–202.
286–298. [101] L.B. Christenson, R.C. Sims, Rotating algal biofilm reactor and spool harvester for
[75] W.J. Henley, R.W. Litaker, L. Novoveská, C.S. Duke, H.D. Quemada, R.T. Sayre, Initial wastewater treatment with biofuels by-products, Biotechnol. Bioeng. 109 (2012)
risk assessment of genetically modified (GM) microalgae for commodity-scale bio- 1674–1684.
fuel cultivation, Algal Res. 2 (1) (2013) 66–77. [102] A. Ozkana, K. Kinneya, L. Katza, H. Berberoglub, Reduction of water and energy re-
[76] P.E. Savage, R.B. Levine, C.M. Huessman, Hydrothermal processing of biomass, Ther- quirement of algae cultivation using an algae biofilm photobioreactor, Bioresour.
mochemical conversion of biomass to liquid fuels and chemicals2010. 192–221. Technol. 114 (2012) 542–548.
[77] Xiaoling Miao, Wu. Qingyu, Changyan Yang, Fast pyrolysis of microalgae to pro- [103] H.C. Greenwell, L.M.L. Laurens, R.J. Shields, R.W. Lovitt, K.J. Flynn, Placing
duce renewable fuels, J. Anal. Appl. Pyrol. 71 (2) (2004) 855–863. microalgae on the biofuels priority list: a review of the technological challenges,
[78] V.K. Dhargalkar, X.N. Verlecar, Southern Ocean seaweeds: a resource for explora- J. R. Soc. Interface 7 (2010) 703–726.
tion in food and drugs, Aquaculture 287 (2009) 229–242. [104] USGS, NWIS Groundwater Levels for New Mexico, Retrieved April 23, 2013 from
[79] Susan M. Renaud, et al., Effect of temperature on growth, chemical composition http://nwis.waterdata.usgs.gov/nm/nwis/gwlevels.
and fatty acid composition of tropical Australian microalgae grown in batch cul- [105] The Engineering Toolbox, Retrieved May 10, 2013 from http://www.
tures, Aquaculture 211 (1) (2002) 195–214. engineeringtoolbox.com/specific-heat-fluids-d_151.html2013.
[80] M.R. Brown, et al., The vitamin content of microalgae used in aquaculture, J. Appl. [106] OSHA, Hexane Boiling Point, Retrieved May 10, 2013 from https://www.osha.gov/
Phycol. 11 (3) (1999) 247–255. SLTC/healthguidelines/n-hexane/recognition.html2013.
[81] Jasvinder Singh, Gu. Sai, Commercialization potential of microalgae for biofuels [107] Tingju Zhu, Claudia Ringler, Ximing Cai, Energy Price and Groundwater Extraction
production, Renew. Sust. Energ. Rev. 14 (9) (2010) 2596–2610. for Agriculture: Exploring the Energy–Water–Food Nexus at the Global and Basin
[82] Jose A. Olivares, Challenges and opportunities in algal biodiesel development, Levels, International Conference of Linkages Between Energy and Water Manage-
No. LA-UR-11-00834; LA-UR-11-834, Los Alamos National Laboratory (LANL), 2011. ment for Agriculture in Developing Countries, Hyderabad, India, 2007.

Vous aimerez peut-être aussi