Você está na página 1de 11

This is an open access article published under an ACS AuthorChoice License, which permits

copying and redistribution of the article or any adaptations for non-commercial purposes.

Article
pubs.acs.org/IECR

Modeling and Dynamic Optimization of Microalgae Cultivation


in Outdoor Open Ponds
Abdulla Malek, Luca C. Zullo, and Prodromos Daoutidis*,

Department of Chemical Engineering & Materials Science, University of Minnesota, Minneapolis, Minnesota 55455, United States
VerdeNero LLC, 955 Highway 169 North, Plymouth, Minnesota 55441, United States

ABSTRACT: This paper presents a model-based dynamic optimization study of the operation of an outdoor open pond for
microalgae cultivation. A nonlinear mathematical model based on rst-principles for predicting the growth of microalgae in open
ponds is developed and validated against literature data. To account for the impact of weather vagaries on the cultivation of
microalgae, data for local climatic conditions is incorporated into the model. The supply of dissolved CO2 to the algal culture
from a CO2 rich gas is modeled as well. Optimal monthly operating proles for the dilution rate, CO2 gas ow rate, and makeup
water ow rate are determined based on minimization of the cultivation cost. The case study included in the analysis is for
cultivating Nannochloropsis Salina over an annual production cycle in California, United States of America. The dynamic
optimization identied a set of operation proles that reduced the cultivation cost by at least 15% as opposed to relying on
heuristic approaches for improving operation.
and optimization based on computational uid dynamics
(CFD) modeling have been carried out as well.10
The economics of algae cultivation are highly sensitive to the
productivity of microalgae and maintaining steady operation in
outdoor cultivation at commercial scale is challenging, because
varying weather conditions greatly aect the growth rate.7 The
identication of optimal operating conditions mitigating the
impact of such environmental factors can maximize the productivity of microalgae over a production cycle. Most of the
studies on optimization of microalgae cultivation are based on
trial-and-error and/or design heuristics.1114 Some model-based
optimization studies have appeared focusing on laboratory algae
growth systems without accounting for weather variations.15,16
The objective of this study is to develop a dynamic
optimization formulation for determining optimal operating
conditions for improving outdoor algae production over a
production period. Specically, a rst-principles model is
developed for algae cultivation in outdoor open ponds
considering the eect of daily varying local climatic conditions
on algae growth. The model accounts for the eect of medium
temperature, irradiance level, and nutrient availability on the
growth of microalgae as well as the transfer of CO2 from a CO2
rich gas to the growth culture. Model validation against
experimental results from the literature is then conducted.
Finally, a dynamic optimization problem is formulated to
determine the optimal dilution rate, makeup water ow rate,
and CO2 gas ow rate monthly proles that minimize the cost
of producing microalgae in a representative location (Imperial
County in California, USA) over the course of a year.

1. INTRODUCTION
1

In 2013, global CO2 emissions reached 35.3 billion tonnes.


Cutting greenhouse gas emissions toward achieving the targets
pledged under the United Nations Framework Convention on
Climate Change (UNFCCC) is a universal challenge.2 This is
stimulating a global interest in nding renewable alternatives to
fossil fuels, thus ensuring environmental and energy sustainability. A potential replacement for petroleum is biofuels, which
can be integrated with the current fuel infrastructure. Satisfying
50% of the transportation fuel demand in the United States of
America (USA) with biofuels derived from palm oil would
require devoting approximately 24% of USA existing cropping
land.3 Thus, relying on energy crops as feedstock for biofuels
production might disrupt the food supply chain. On the other
hand, utilizing algal biofuels would reduce that cultivation
land requirement to 2.5%, making algal biomass a promising
feedstock for biofuels production.3 Moreover, carbon constitutes up to 50% of dry weight algal biomass (DW); hence,
microalgae cultivation would assist in mitigating CO2 as approximately 1.83 kg of CO2 is required to produce each kilogram
of DW.4 A well-to-pump net CO2 emissions of 20.9 kg GJ1
of energy generated has been reported in a life cycle analysis
study for an algal biodiesel process.5
Microalgae are widely cultivated for a variety of products
including food additives, pigments, antibiotics, and nutraceuticals.6 However, the production cost remains too high for highvolume, low-value markets; for example, a conservative estimate
for the cost of producing algae-based green diesel is around
$10 gal1.7 In parallel with advances in algal biotechnology,
process systems engineering can play an important role in
optimizing the economics of algae-based commodity products
such as biofuels. Several optimization studies have appeared in
the literature focusing on the design and synthesis of
downstream processes for biofuels production and CO2
mitigation using microalgae.8,9 Studies on bioreactor design
2015 American Chemical Society

Special Issue: Sustainable Manufacturing


Received:
Revised:
Accepted:
Published:
3327

September 1, 2015
October 26, 2015
October 27, 2015
October 27, 2015
DOI: 10.1021/acs.iecr.5b03209
Ind. Eng. Chem. Res. 2016, 55, 33273337

Article

Industrial & Engineering Chemistry Research

Figure 1. Approximation of the open pond model as a cascade of compartments.

2. PROCESS DESCRIPTION AND MODELING


The economics of mass cultivation of microalgae are more
favorable for the open pond system than the photobioreactor
congurations.7,17 Several open pond modeling studies in the
literature18,19 have adopted the concept proposed and validated
in refs 20 and 21 where the hydrodynamics of an open pond
are modeled by a cascade of continuous stirred-tank reactors
(CSTRs). In this approach, an individual open pond is approximated as a series of n compartments where the content leaving
the last compartment is recirculated to the rst compartment as
demonstrated in Figure 1. Assuming each compartment to be
well-mixed, it can be modeled as a CSTR and the corresponding mass and energy balances can be constructed for
the species considered in this model: microalgae, water, and
CO2. Models for the growth of microalgae in open ponds can
be found in refs 2123; whereas refs 19 and 24 also incorporate
modeling of the CO2 transfer. The microalgae growth kinetics
formulation presented herein is based on the work in ref 23 and
the CO2 transfer modeling is adopted from ref 24.
2.1. Water Material and Energy Balances. In an
insulated open pond, water enters through the pond feed
(Ffeed), or from precipitation, and leaves via evaporation
(Fevap), harvest (Fharvest), or consumption by microalgae in
photosynthesis. Typically algae are grown in geographic areas
lacking rains, hence the amount of water added by
precipitation is hereby neglected. Water consumed during
photosynthesis is also negligible. For any CSTR in the pond,
water ows in (Fin) from the preceding CSTR and ows out
(Fout) to the succeeding CSTR in the direction of ow.
Therefore, the depth of water (H) in an arbitrary CSTR in a
segmented open pond can be found from
dH
LW i = Fin, i Fout, i Fevap, i
(1)
dt
where L is the length and W is the width of each CSTR, which
are assumed constant and identical for all of the CSTRs. The
following relations govern the ow within the pond between
the ith CSTR and the preceding CSTR
Fin, i = Fout, i 1 = WHi 1v
(2)

To compensate for evaporation losses, after harvesting the


algal biomass the water is combined with makeup water
(Fmakeup) and recycled back to the pond as Ffeed as shown in
Figure 2. Regardless of the microalgae type, i.e. freshwater vs

Figure 2. Modeling the open pond as a cascade of continuous stirredtank reactors (CSTRs) with an internal and an external recycle stream.

marine, only freshwater is considered for Fmakeup, because even


for a seawater pond to maintain the salinity at the desired level
freshwater has to be added. Water blowdown for preserving
culture quality is neglected. The amount of water evaporating
from each CSTR is calculated accordingly
Fevap, i =

(4)

Ev, i = (19.0 + 0.95U 2)(Psat, i Pair)

(5)

Le, i = 597.3 0.57Tw, i

(6)

where U is the wind speed above the pond water and Tw is the
water temperature.25 The saturation vapor pressure at the water
temperature (Psat) and the vapor pressure in the overlaying air
(Pair) are computed using Antoines equation.
The temperature of the water greatly aects the growth rate
of algae. Hence, an energy balance for water is derived to track
the water temperature in each compartment

n
i=1

Le, i

where is the water density. The latent heat ux (Ev),


according to Daltons Law, and the latent heat of evaporation
(Le) are given by

where v is the water velocity, and i [1, n]. The pond feed and
harvest are taken into account by adding Ffeed and subtracting
Fharvest from eq 1 for the rst and nth CSTRs, respectively. The
dilution rate (D) dictates the rate of algal biomass harvest
Fharvest = DLW Hi

LWEv, i

LWc p
(3)

d
(HiTw, i) = c p(Fin, iTw, i 1 Fout, iTw, i) + LWEi
dt
(7)

3328

DOI: 10.1021/acs.iecr.5b03209
Ind. Eng. Chem. Res. 2016, 55, 33273337

Article

Industrial & Engineering Chemistry Research


where cp is the specic heat capacity of water. The term E
represents the heat exchanged through the water surface of
each compartment. It accounts for the heat addition from the
absorbed solar irradiance (Es) and atmospheric long-wave (Ea),
and the heat loss due to evaporation (Ev), water long-wave
(Ew), and conduction to atmosphere (Ec). These terms are
calculated based on the model developed in ref 25
4

Ea = (Tair + 273) (1 + 0.031 Pair )(1 5)

(8)

Ew, i = (Tw, i + 273)4

(9)

Ec, i = 2(19 + 0.95U 2)(Tw, i Tair)

(10)

Es = 3Ia

(11)

of algae considered in the model are the temperature of the


culture, the irradiance level, and the nutrient concentrations.
The maximum algae growth rate (max) is achieved when those
factors are at their optimal levels, otherwise the actual growth
rate depends on the deviations from these levels as
i = max fN, i fT, i fI, i average

where f N, f T, and f I, are the attenuation factors for nutrient,


temperature, and light limitations, respectively.23 Most of these
factors are species-specic, and so their detailed forms depend
on the particular algae strain in study.
The marine alga Nannochloropsis Salina (N. Salina) chosen
for this study is a promising strain for algal biofuels production
because of its high lipid content and robustness in outdoor
cultivation.29 Since there are limited studies on outdoor cultivation of N. Salina, the mass and energy balances in the model
are validated using data from the literature for the protein rich
freshwater Spirulina which is widely cultivated outdoors at
commercial scale for food-grade products.30 Microalgae
proliferate within a certain range around an optimal temperature and growth stops completely away from such range of
temperature. The term f T can be estimated using the following
exponential relation

where Ia is the daily average solar irradiance at the pond surface


and Tair is the air temperature. The other parameters are
dened in Table 1 including their assigned values as well as
values for the mass balance parameters.
Table 1. Parameter Denitions and Values for the Water
Mass and Energy Balances
parameter

description

value

cp

1
2
3
5

water velocity in the open pond (cm s1)


density of water (g cm3)
specic heat of water (cal K1 g1)
StefanBoltzmann constant (cal cm2 d1 K4)
water emissivity
atmospheric attenuation coecient
Bowens coecient (mmHg C1)
radiation absorption factor
reection coecient

2526
1.0
1.0
11.7 108
0.9725
0.625
0.47
0.927
0.0325

fT, i

fI, i average =

d
(HiCalgae, i) = Fin, iCalgae, i 1 Fout, iCalgae, i
dt
+ (i Bi )LWHiCalgae, i

feed

= (1 Heff )Calgae, n

Fharvest
Ffeed

(12)

(13)

where Heff is the harvest eciency which is around 95% for


harvest of algal biomass.28 The areal productivity of the system
(Pr) is often used as an evaluation criterion and could be
dened as
D
Pr = Heff
n

i=1

Hi

tp

t p

fI, i dt dz

(18)

(19)

where I is the photosynthetically active radiation (PAR)


experienced by the microalgae and Imax is the maximum
radiance above which algae growth does not increase any
further. However, for some microalgae strains, including
Spirulina, exceeding Imax actually damages the photosynthetic
reaction, known as the photoinhibition eect. This phenomenon is captured in the relation proposed by Steele35

(14)

The basal metabolism rate increases exponentially with the


water temperature
Bi = Bm e kB(Tw,i TB)

1
Hi

fI, i = 1 eIi / Imax

HiCalgae,i

(17)

where is the fractional length of a day having daylight


(photoperiod), tp is the length of a day, and z is the vertical distance from the water surface. Since autotrophic microalgae do
not grow at night, this time integral is formulated to generate
an average over the daylight hours; using a daily average instead
would lead to overestimation of the growth rate. Various
formulas have been developed in the literature32 for modeling
f I, and the model proposed in ref 33 ts the experimental data
presented in ref 34 for N. Salina

After harvesting the algal biomass, the water is recycled back to


the pond and the concentration of algae in the feed is given by
Calgae

k T,1(Tw,i Topt)2
, Tw, i > Topt
e
=
ek T,2(Topt Tw,i)2 , T T
w, i
opt

where Topt is the optimal temperature for algae growth and kT,1
and kT,2 are tting constants.23 The optimal temperature for
Spirulina is within the range of 2442 C depending on the
particular strain of Spirulina, and for N. Salina it is around
28 C.28,31 Autotrophic microalgae drive photosynthesis by the
energy from light photons, and due to the shading eect, the
microalgae closer to the pond surface are exposed to higher
light intensities; hence, the mean value of f I is considered

2.2. Microalgae Mass Balance and Growth Kinetics. In


the absence of algae grazers, the concentration of algae in the
pond (Calgae) would depend on the growth rate (), the harvest
rate, and the rate of algal biomass deterioration due to
respiration and other basal metabolism processes (B)
LW

(16)

(15)

fI, i =

where Bm is the metabolic rate at a reference temperature (TB),


and kB is a tting constant.23 Factors aecting the growth rate
3329

Ii
Imax

e1 Ii / Imax

(20)
DOI: 10.1021/acs.iecr.5b03209
Ind. Eng. Chem. Res. 2016, 55, 33273337

Article

Industrial & Engineering Chemistry Research


The BeerLambert law is commonly used to model I as
corrected for the attenuating eect of the growth culture

derived in ref 24
yout, i =

I
Ii = 34 a eke,iz

(21)

eKLai(1 g,i)WHiWs/ Q g,iHe

where 4 is a factor for converting from total radiation to PAR.


The extinction coecient (ke) is related to the water/background turbidity (kw) and the concentration of algae in the
pond
ke, i = k w + .0088Calgae, iR Chl + .054(Calgae, iR Chl)2/3

(22)

ai = db2

N0, i =

(23)

where KC and Ks are the half saturation and inhibition constants


for CO2, respectively.24 Autotrophic microalgae acquire carbon
by consuming dissolved CO2 supplied articially and/or from
diusion of atmospheric CO2, hence the balance for the CO2 in
the pond water (CCO2) is

g, i =

PC =

R gTg

(yin yout, i )

6 (1 g, i)

(28)

6Q g, i
6Q g, i + vWW
s

(29)

cost CO2, t + cost water, t + costenergy, t


harvest t

t=1

(24)

(30)

where t is an integer denoting the number of the day starting


from the beginning of the production cycle, and Tc is the length
of the time horizon. The daily amount of algal biomass
harvested, cost of CO2, cost of water, and cost of energy are
calculated by

where RCO2 is the CO2 requirement and G is the rate of CO2


supplementation by bubbling CO2 rich gas into the pond water.
The mass transfer coecient Katm regulates the diusion of
CO2 to/from the atmosphere as driven by the dierence
between CCO2 and the equilibrium concentration of atmospheric CO2 in water (CCO2,atm).
2.3. CO2 Sump Stations. Sump stations located at the
pond middle are used for bubbling CO2 gas into the growth
culture in a concurrent or countercurrent arrangement. Modeling
the CO2 transfer from the gas bubbles to the pond water in a
sump station allows analyzing the eect of bubbling rate on
algae growth which is useful for the optimization problem. For
a given CO2 mole fraction in the inlet gas (yin), the term G can
be calculated from determining the CO2 mole fraction in the
outgassing bubbles (yout)
Gi =

Q g, i
db3

Tc

(i Bi )HiR CO2Calgae, i

PgQ g, i

(27)

where 6 is a correction factor for the compression of gas under


water.40
2.4. Process Economics. The algal biomass production
cost (PC), a criterion to evaluate dierent scenarios in this
work, is dened as

(FinCCO2, i 1 FoutCCO2, i + Gi)


d
(HiCCO2, i) =
dt
LW
K atm(CCO2, i CCO2,atm)

WW
s (v b v )

The gas hold up is determined from the volume ratio of gas in


the sump

2
CCO
2, i

Ks

No, i

where db is the bubble diameter, and vb is the gas bubble


terminal velocity.24 According to ref 39, the number of bubbles
of gas formed at the sump bottom (N0) is

CCO2, i
K C + CCO2, i +

(26)

where He is the dimensionless Henrys constant, KL is the mass


transfer coecient for the CO2 transfer from the gas phase to
the liquid phase, Ws is the width of the sump station, and g is
the gas hold up. For countercurrent ow, the gas bubbles total
interfacial area (a) can be approximated from

where RChl is the chlorophyll content of the algal biomass.36


There are at least 30 chemical elements required for growing
microalgae of which the most important ones, i.e. usually the
growth limiting ones, are nitrogen, phosphorus, and carbon.37
However, for the purpose of determining f N in this analysis,
only carbon is considered, because it constitutes a big fraction
of the algal biomass.38 Assuming that all nutrients are abundant except for carbon, a Monod type equation modied for
the inhibitory eect of oversupply of CO2 can be used to
estimate f N
fN, i =

y Pg

HeCCO2, i + in
CCO2, i
Pg

HeR gTg

R gTg

harvest t = Heff Calgae, n , tFharvest, t

(31)

cost CO2, t = xCO2CO

Q g,i ,t

(32)

i=1

cost water, t = x wFmakeup, t

(33)

costenergy, t = xeEmixing, t

(34)

where xCO2, xw, and xe are the prices of CO2, freshwater, and
electricity, respectively. The CO2 gas density (CO2) is
calculated using the ideal gas law. The daily energy requirement
for mixing the open pond water using a paddle wheel (Emixing) is
given by

(25)

where Rg is the universal gas constant, and the gas ow rate


(Qg), temperature (Tg), and pressure (Pg) are assumed
constant throughout the entire water column. Assuming plug
ow for the gas phase, yout is estimated based on the model

Emixing =
3330

89(hfriction + hbend)vW i = 1 Hi
nMeff

(35)
DOI: 10.1021/acs.iecr.5b03209
Ind. Eng. Chem. Res. 2016, 55, 33273337

Article

Industrial & Engineering Chemistry Research


where Meff is the eciency of the mixing system, 8 is a unit
conversion factor, and 9 is the number of hours the paddle
wheel is running daily.26 From the Gauckler-Manning formula
for open channel ow one can calculate the head losses from
friction of the pond bottom (hfriction) and ow around each
bend (hbend)
hfriction

hbend

nL
= v 2n02 4/3
r

37v 2
=
g

Table 2. Parameter Denitions and Values for the


Microalgae Growth Kinetics Model
parameter
Bm

description

value

metabolic rate at reference


temperature (d1)
TB
reference temperature for metabolic
rate (C)
kB
metabolic rate exponential tting
constant (C1)

fractional length of a day having


daylight
length of a day (h)
tp
kw
water (background) turbidity (m1)
4
coecient for photosynthetically active
radiation (PAR)
Species-Specic Parameters
max
maximum algae growth rate (d1)
Topt
optimal temperature for algae growth
(C)
kT,1
temperature limitation tting constant
(C2)
kT,2
temperature limitation tting constant
(C2)
Imax
maximum irradiance for algae growth
(E m2 s1)
RChl
algal biomass chlorophyll content
(g Chl kg1 DW)
Initial Conditions
H
depth of water in a CSTR in the open
pond (m)
Tw
temperature of the water in the CSTR
(C)
Calgae
concentration of microalgae in the
pond (g DW m3)
molar concentration of CO2 in the
CCO2
pond (mol m3)

(36)

(37)

where n0 is a roughness factor also known as the GaucklerManning coecient, r is the channel hydraulic radius, 7 is the
kinetic loss coecient, and g is the acceleration of gravity.41
The mathematical model described above was coded in
gPROMS ModelBuilder v4.0 installed in a 64-bit Windows 7
CPU equipped with an i7 processor at 3.4 GHz and 16 GB of
RAM.42

3. MODEL VALIDATION
3.1. Experimental Setup and Model Assumptions.
Spirulina was cultivated in a 450 m2 outdoor open pond for 10
months in Malaga, Spain.43,44 The experiment resembles a
scenario for commercial scale algal biomass production,
especially given the prolonged cultivation period and varying
environmental conditions. Therefore, the water temperature,
biomass areal concentration, and productivity proles reported
in refs 43 and 44 were used to validate the ones predicted by
the developed model.
In the experiment, the depth of water was maintained at
30 cm, and the growth culture was prepared with modied
Zarrouks Medium providing adequate nutrient levels.43,45
Moreover, the pond was inoculated with an areal concentration
of 15 g DW m2 and harvest was started after 13 days and only
interrupted during February due to heavy rains.44 The
corresponding model assumptions and parameters are (1) the
open pond was discretized into n = 18 compartments each
having a length and width of L = W = 5 m; (2) the makeup
water ow rate was set to match the evaporation losses to x
the depth at H = 0.3 m; (3) the nutrient concentrations were
at their optimal values; therefore, eq 23 was set to f N,i = 1 and
eqs 2429 were excluded; (4) after the inoculation period
(13 days), the dilution rate was set to an estimate D = 0.10 d1
based on ref 46 except for 17 days during February where there
was no harvest.
The initial conditions and values assigned to the kinetic
parameters for Spirulina growth are shown in Table 2. Figure 3
shows the solar radiation and air temperature data for the
period of the experiment at Malaga, which were obtained from
the European Database of daylight and Solar Radiation47 and
the Tutiempo Network,48 respectively. These data, including
data for humidity, wind speed, and photoperiod, were used in
the simulation of the model consisting of eqs 118 and 2023.
The time integration of the ordinary dierential equations was
performed using the DASOLV solver which is based on variable
time step Backward Dierentiation Formulas (BDFs).49
This solver embeds the MA48 subsolver which uses a direct
LU-factorization algorithm to solve a set of linear algebraic
equations. To increase numerical stability the default value of
the PivotStabilityFactor was changed to 0.9, which is one of
the setting parameters of the MA48 subsolver.

0.0423
2023
0.06923
0.5
24
0.325
2.0550
Spirulina
1.451
27.523

N. Salina
1.334
2752

0.00523

0.0152

0.00423

0.0352

20053

5834

6.643

1754

Spirulina
0.3

N. Salina
0.3

22.1a

12.4a

15

15
0.02a

Assumed in equilibrium with the ambient environment.

Figure 3. Daily global solar radiation and average air temperature at


Malaga from September 23, 1997, to July 31, 1998.47,48

3.2. Simulation Results. The predicted water temperature


prole shown in Figure 4 is slightly lower than the one
determined experimentally which is due to the fact that the air
temperature data used herein are marginally lower than the
measurements at the experiment.43,44 As a result, the growth
was inhibited in the winter but advanced in the summer, and
consequently, the areal concentration prole was slightly
altered. However, as shown in Figure 5 the predicted productivity is in good agreement with the experimentally determined productivity with a mean percent error of 16.3%, which
demonstrates the adequacy of the proposed model.
3331

DOI: 10.1021/acs.iecr.5b03209
Ind. Eng. Chem. Res. 2016, 55, 33273337

Article

Industrial & Engineering Chemistry Research

Table 3. Values Assigned to the Parameters of the CO2 Mass


Balance and Transfer Model
parameter

description

value

KC
Ks
CCO2,atm

half saturation constant for CO2 (mol CO2 m3)


inhibition constant for CO2 (mol CO2 m3)
equilibrium concentration of atmospheric CO2 in
water (mol m3)
CO2 requirement per unit of algae
(mol CO2 g1 DW)
CO2 gas bubble terminal velocity (cm s1)
temperature of CO2 gas (C)
pressure of CO2 gas (atm)
mole fraction of CO2 in the bubbling gas at inlet
universal gas constant (atm m3 mol1 K1)
dimensionless Henrys constant
width of the CO2 gas sump station (m)
diameter of CO2 gas bubble (mm)
coecient for gas compression under water depth
of 30 cm
mass transfer coecient for diusion of CO2 to/
from atmosphere (m d1)
mass transfer coecient for CO2 transfer from gas
bubbles to water (m d1)

9 104 24
18024
0.02

RCO2
vb
Tg
Pg
yin
Rg
He
Ws
db
6

Figure 4. Simulation results for Spirulina cultivation in Malaga, Spain,


from September 1997 to July 1998: algal biomass areal concentration
and temperature of the pond water.

Katm
KL

0.04256
30 assumed
31626
1.226
1
8.2 105
0.8317
0.326
257,58
0.9640
2.45
9.5959

Table 4. Parameter Denitions and Values for the Process


Economics Model
parameter

Figure 5. Comparison between monthly averaged values of the


predicted productivity and the productivity determined experimentally
in ref 44.

description

value

xw
xe
xCO2

price of agricultural water ($ m3)


price of electricity ($ kW1 h1)
price of pure CO2 gas ($ tonne1)

0.01660
0.0461
4062

Meff
n0
r
7

eciency of the paddle wheel mixing system


Gauckler-Manning coecient
channel hydraulic radius
kinetic loss coecient

40% assumed
2.08 107 26
0.2926
226

concentration of microalgae in the pond also resulting in a


lower productivity. Moreover, in an extremely sunny day this
could be damaging to the growth rate for a microalga that
experiences photoinhibition which would lead to an even lower
productivity. Therefore, local climatic conditions have to be
considered when searching for the optimal D. Another
important operating parameter is the rate of bubbling of the
CO2 gas (Qg) into the pond culture. Increasing Qg provides
more dissolved CO2 to the growth culture, but it increases the
loss of CO2 through degassing as well, and CO2 is a costly
feedstock. The operating depth of water in the pond, controlled
by Fmakeup, inuences the system productivity and economics
through aecting the light penetration, CO2 absorption and
mixing energy requirement.
For the base case scenario, the operation was set based on
best practice as reported in the literature for commercial scale
cultivation. The optimal D was determined by employing the
empirical harvest scheme proposed in reference.46 It starts by
holding o harvest and allowing the algae to grow until the
stationary phase of growth is reached followed by ramping
up D in steps of 0.05 d1 as demonstrated in Figure 6. Each
time the concentration of algae in the pond is stabilized for
several days the dilution rate is increased and the optimum is
the one yielding the maximum algae productivity. The next step
is to determine the algal biomass areal concentration, which
when reached after inoculation, harvest at the identied D
should commence. Typically the CO2 is added to the culture
according to the demand for carbon as determined by the

4. DYNAMIC OPTIMIZATION
4.1. Problem Denition. Imperial County in California is
one of the suitable places for cultivating N. Salina in USA owing
to the warmer weather and availability of resources including
land, water and CO2.26 This site was selected for the
optimization case study and the daily weather conditions in a
typical meteorological year for this site were obtained from the
National Solar Radiation Data Base.55 The 4 ha open pond
proposed in ref 26 for the production of algal biofuels at
commercial scale was adopted herein. Therefore, the pond was
discretized into 44 compartments with L = W = 30 m and the
CO2 gas was set to be introduced at the bottom of the 12th and
34th compartments. The parameter values selected for the
growth kinetics of N. Salina are shown in Table 2. The values
assigned to the parameters of modeling the CO2 transfer and
process economics are shown in Tables 3 and 4, respectively.
For a production cycle of 1 y, the program, comprising
eqs 119 and 2137, contains 1043 variables and 1876
parameters.
4.2. Base Case. The dilution rate (D) is one of the main
operating parameters aecting the productivity of an algal
system. A low D creates a dense culture where light availability
becomes limited due to the shading eect. This reduces the
growth rate of microalgae and consequently leads to a lower
productivity. On the other hand, a high D would reduce the
3332

DOI: 10.1021/acs.iecr.5b03209
Ind. Eng. Chem. Res. 2016, 55, 33273337

Article

Industrial & Engineering Chemistry Research

Figure 8. Sensitivity analysis demonstrating the eect of the dilution


rate (D) on the production cost (PC). The minimum PC is at D of
0.31 d1.

Figure 6. Demonstration of the heuristic approach for nding the


optimal dilution rate. The dilution rate is increased in steps of 0.05 d1
each time the microalgae reach a stationary phase.

conditions, therefore, the dilution rate prole needs to be


determined using dynamic optimization. Furthermore, in the
previous scenarios around 22% of the CO2 fed to the pond was
lost to the atmosphere due to degassing, and Chart 1 shows

pH of the culture, hence Qg was set to the level just enough


to prevent any limitation on growth due to CO2 deciency.63
The depth of culture is usually maintained at a certain level;
therefore, Fmakeup was set to the level that compensates for
evaporation losses.4
The productivity prole shown in Figure 6 is consistent with
the experimental results reported in ref 46; note that the
highest productivity was achieved when D is 0.10 d1. From
Figure 7, it was observed that the growth rate is mostly
inhibited by the limited availability of sunlight and this
limitation is higher for sunnier days. Considering that N.
Salina does not experience photoinhibition, this could only
mean that the shading eect of the high algal biomass concentration is the main factor responsible for limiting the
productivity of the system. Comparing Figures 6 and 7, it was
noticed that the growth rate peaks when the areal concentration
is around 35 g DW m2. Therefore, the base case was simulated
starting with no harvest followed by D = 0.10 d1 once the
concentration exceeded 50 g DW m2 and PC was found to be
$240 tonne1. Sensitivity analysis can provide a better insight
into the eect of the dilution rate on PC as demonstrated in
Figure 8, suggesting an optimal D of 0.31 d1 which improves
the base case PC to $196 tonne1.
An alternative to nding a single dilution rate is to aim for
the optimal areal concentration and adjust the dilution rate accordingly.64 However, in outdoor cultivation the optimal areal
concentration is a function of the uncontrolled environmental

Chart 1. Breakdown of the Production Cost (PC) of


$196 tonne1 for the Improved Base Case Scenario Using
a Dilution Rate of 0.31 d1

that the cost of the CO2 gas constitutes 59% of the improved
base case PC; hence, Qg needs to be optimized as well. Also,
although the cost of makeup water is insignicant as Chart 1
shows, optimizing Fmakeup would aect the CO2 absorption.
4.3. Optimization Problem. The optimization problem
considered determines the proles of the dilution rate, CO2 gas
ow rate, and makeup water ow rate that

minimize PC eq 30

D , Q g , Fmakeup

Figure 7. Eect of the attenuation factors on the growth rate during the search for the optimal dilution rate and biomass algal concentration in the
pond using the heuristic approach.
3333

DOI: 10.1021/acs.iecr.5b03209
Ind. Eng. Chem. Res. 2016, 55, 33273337

Article

Industrial & Engineering Chemistry Research

Figure 9. Comparing results from the improved base case and the optimization study: (a) algal biomass areal concentration; (b) dilution rate;
(c) CO2 gas ow rate; (d) makeup water ow rate.

Figure 10. Eect of attenuation factors on the growth rate for the optimal case scenario.

subject to the following constraints:


(1)
(2)
(3)
(4)

Q g 0,

mass and energy balances eqs 114


growth kinetics eqs 1519 and 2123
CO2 transfer eqs 2429
process economics eqs 3037

Fmakeup 0,

0 D 0.5,

0.2 H 0.4

assuming that the CO2 gas and makeup water availability are
unlimited. The bounds on the dilution rate and depth were set
based on typical operation in commercial algae facilities.4 The
improved base case values for the decision variables were used
as an initial guess and the dynamic optimization was performed
using the CVP_SS solver which implements a control vector
parametrization algorithm based on the single-shooting
method.65 It took the optimizer 1443 s to nd the optimal
operating proles shown in Figures 9bd corresponding to a
PC of $167 tonne1 which is 15% less than in the improved

Note that this formulation contains nonlinear growth kinetics


and transient balances making it a nonlinear dynamic optimization problem. Assuming that the optimal areal concentration varies only on a monthly basis, the time horizon was
divided into 12 control intervals creating 36 optimization decision variables which are subject to the following
bounds:
3334

DOI: 10.1021/acs.iecr.5b03209
Ind. Eng. Chem. Res. 2016, 55, 33273337

Article

Industrial & Engineering Chemistry Research

9 = daily hours the paddle wheel is operating (24 h d1)


B = rate of basal metabolism processes (d1)
Bm = metabolic rate at reference temperature (d1)
= fractional length of a day with daylight
Calgae = mass concentration of microalgae in the pond
(g DW m3)
Calgaefeed = concentration of algae in the feed (g DW m3)
CCO2 = molar concentration of CO2 in the pond (mol m3)
CCO2,atm = equilibrium concentration of atmospheric CO2 in
water at 20 C (mol m3)
costCO2 = cost of CO2 supplied to the open pond ($ d1)
costwater = cost of freshwater supplied to the pond ($ d1)
costenergy = cost of energy supplied to the pond ($ d1)
cp = specic heat capacity of water (cal K1 g1)
D = dilution rate (d1)
db = diameter of CO2 gas bubble (mm)
E = heat exchanged through the water surface of the CSTR
(cal cm2 d1)
Ea = heat added to the CSTR water from the atmospheric
long-wave (cal cm2 d1)
Ec = heat lost from the CSTR water by conduction to
atmosphere (cal cm2 d1)
Emixing = energy requirement for mixing the open pond water
(kW h d1)
Es = heat added to the CSTR water from the absorbed solar
irradiance (cal cm2 d1)
Ev = latent heat ux (cal cm2 d1)
Ew = heat lost from the CSTR water through water longwave (cal cm2 d1)
= water emissivity
g = gas hold up
Fevap = water evaporating from a CSTR in the open pond
(m3 d1)
Ffeed = ow rate of the open pond feed stream (m3 d1)
Fharvest = ow rate of the open pond harvest stream (m3 d1)
Fin = water entering a CSTR from the previous CSTR within
the open pond (m3 d1)
Fmakeup = makeup water ow rate (m3 d1)
Fout = water leaving a CSTR to the following CSTR within
the open pond (m3 d1)
f I,f N,f T = attenuation factors for light, nutrient, and
temperature limitations
g = acceleration of gravity (m s2)
G = rate of CO2 bubbling into the pond water (mol d1)
H = depth of water in a CSTR in the open pond (m)
harvest = algal biomass harvested daily (tonne d1)
He = dimensionless Henrys constant
Heff = eciency of algal biomass harvest (%)
hfriction/bend = head loss from friction and from ow around
the bends and sumps (m)
i = index for the location of the CSTR relative to the other
CSTRs in the open pond
I = visible irradiance absorbed at the pond surface over
daylight hours (E m2 s1)
Ia = daily average solar irradiance at the pond surface (W m2)
Imax = maximum irradiance above which algae growth does
not increase (E m2 s1)
Katm = mass transfer coecient for diusion of CO2 to/from
the atmosphere (m d1)
kB = exponential tting constant for metabolic rate (C1)
KC = half saturation constant for CO2 (mol CO2 m3)
ke = extinction coecient related to water turbidity and algae
concentration (m1)

base case. Interestingly, Figure 9a shows that the algal biomass


areal concentration in the improved base case and the optimal
case are almost identical after around 3 months of cultivation;
hence, it seems that the optimal areal concentration is indeed
close to 30 g DW m2. As demonstrated in Figures 9a and b,
the optimal dilution rate prole follows the algal biomass areal
concentration during the year, but at the end of the production
cycle D goes up so that the remaining algae in the pond is
completely harvested. Note from Figure 9c that the optimal
CO2 gas ow rate is on average less than in the improved base
case scenario, which reduces the CO2 loss from the pond to the
atmosphere from an annual average of 1.10.58 kg CO2 kg
DW1 produced. Although, this causes growth limitation due to
CO2 deciency as shown in Figure 10, apparently the gain on
algae growth does not justify the increased loss of CO2 to the
atmosphere from an economic point of view.

5. CONCLUSIONS
A mathematical model for estimating the growth of microalgae
in an outdoor open pond based on local climatic conditions was
developed in this work. The model was validated against
literature data for the production of Spirulina in an outdoor
open pond in Malaga, Spain. The simulated algal biomass
productivity agreed with experimental data, with a mean
percent error of approximately 16%. A dynamic optimization
problem was formulated for determining the location-specic
optimal monthly operating proles for the dilution rate, CO2
gas ow rate, and makeup water ow rate. A case study was
conducted for the cultivation of N. Salina in California, USA.
The operating proles generated by the optimization lowered
the cultivation cost by at least 15% when compared with other
case scenarios where best practice operation and sensitivity
analysis were employed.
Based on the analysis presented, algae facilities adopting
outdoor open ponds and having access to local meteorological
data can use the formulation presented in this work to lower
the cultivation cost by identifying optimal operating conditions.
Furthermore, the proposed formulation can be useful for
determining facility locations for algal biomass production by
comparing candidate sites based on optimized operations.

AUTHOR INFORMATION

Corresponding Author

*E-mail: daout001@umn.edu.
Notes

The authors declare no competing nancial interest.

ACKNOWLEDGMENTS
The authors acknowledge the nancial support from Abu Dhabi
National Oil Company (ADNOC).
NOMENCLATURE
a = total interfacial area of CO2 gas bubbles (m1)
1 = atmospheric attenuation coecient
2 = Bowens coecient (mmHg C1)
3 = radiation absorption factor
4 = fraction accounting for the visible portion of solar
irradiance
5 = reection coecient
6 = correction factor for the compression of gas under water
7 = kinetic loss coecient
8 = conversion factor (9.8 W s kg1 m1)
3335

DOI: 10.1021/acs.iecr.5b03209
Ind. Eng. Chem. Res. 2016, 55, 33273337

Article

Industrial & Engineering Chemistry Research

(6) Foley, P. M.; Beach, E. S.; Zimmerman, J. B. Algae as a source of


renewable chemicals: opportunities and challenges. Green Chem. 2011,
13, 1399.
(7) Davis, R.; Aden, A.; Pienkos, P. T. Techno-economic analysis of
autotrophic microalgae for fuel production. Appl. Energy 2011, 88,
35243531.
(8) Gong, J.; You, F. Optimal Design and Synthesis of Algal
Biorefinery Processes for Biological Carbon Sequestration and
Utilization with Zero Direct Greenhouse Gas Emissions: MINLP
Model and Global Optimization Algorithm. Ind. Eng. Chem. Res. 2014,
53, 15631579.
(9) Martn, M.; Grossmann, I. E. Simultaneous Optimization and
Heat Integration for Biodiesel Production from Cooking Oil and
Algae. Ind. Eng. Chem. Res. 2012, 51, 79988014.
(10) Smith, J. D.; Neto, A. A.; Cremaschi, S.; Crunkleton, D. W.
CFD-Based Optimization of a Flooded Bed Algae Bioreactor. Ind. Eng.
Chem. Res. 2013, 52, 71817188.
(11) Gross, M. Development and optimization of algal cultivation
systems. M.S. Thesis, Iowa State University, Ames, USA, 2013.
(12) Park, K.-H.; Lee, C.-G. Optimization of Algal Photobioreactors
Using Flashing Lights. Biotechnol. Bioprocess Eng. 2000, 5, 186190.
(13) McKnight, K. Optimizing Growth of Microalgae for Use as a
Potential Biofuel Feedstock. Ph.D. Dissertation, Lawrence Berkeley
National Laboratory, Berkeley, USA, 2013.
(14) Richmond, A.; Cheng-Wu, Z. Optimization of a flat plate glass
reactor for mass production of Nannochloropsis sp. outdoors. J.
Biotechnol. 2001, 85, 259269.
(15) Thornton, A.; Weinhart, T.; Bokhove, O.; Zhang, B.; Sar, D. M.
v. d.; Kumar, K.; Pisarenco, M.; Rudnaya, M.; Savcenco, V.;
Rademacher, J.; Zijlstra, J.; Szabelska, A.; Zyprych, J.; Schans, M. v.
d.; Timperio, V.; Veerman, F. Modeling and optimization of algae
growth, CASA-Report 1059; Eindhoven University of Technology,
The Netherlands, 2010.
(16) He, L.; Subramanian, V. R.; Tang, Y. J. Experimental analysis
and model-based optimization of microalgae growth in photobioreactors using flue gas. Biomass Bioenergy 2012, 41, 131138.
(17) Richardson, J. W.; Johnson, M. D.; Outlaw, J. L. Economic
comparison of open pond raceways to photo bio-reactors for profitable
production of algae for transportation fuels in the Southwest. Algal Res.
2012, 1, 93100.
(18) Jupsin, H.; Praet, E.; Vasel, J. L. Dynamic mathematical model
of high rate algal ponds (HRAP). Water Sci. Technol. 2003, 48, 197
204.
(19) Yang, A. Modeling and Evaluation of CO2 Supply and
Utilization in Algal Ponds. Ind. Eng. Chem. Res. 2011, 50, 11181
11192.
(20) Miller, S. B.; Buhr, H. O. Mixing characteristics of a high-rate
algae pond. Water SA 1981, 7, 815.
(21) Buhr, H. O.; Miller, S. B. A dynamic model of the high-rate
algal-bacterial wastewater treatment pond. Water Res. 1983, 17, 2937.
(22) James, S. C.; Boriah, V. Modeling Algae Growth in an OpenChannel Raceway. J. Comput. Biol. 2010, 17, 895906.
(23) Cerco, C. F.; Cole, T. Users Guide to the CE-QUAL-ICM ThreeDimensional Eutrophication Model; US Army Corps of Engineers:
Vicksburg, 1995.
(24) Ketheesan, B.; Nirmalakhandan, N. Modeling microalgal growth
in an airlift-driven raceway reactor. Bioresour. Technol. 2013, 136, 689
696.
(25) Chapra, S. C. Surface Water-Quality Modeling; The McGRAWHill Companies, INC.: Boulder, CO, 1997; pp 560576, 603621.
(26) Lundquist, T. J.; Woertz, I. C.; Quinn, N. W. T.; Benemann, J.
R. A Realistic Technology and Engineering Assessment of Algae Biofuel
Production; Energy Biosciences Institute: Berkeley, CA, 2010.
(27) Weyer, K. M.; Bush, D. R.; Darzins, A.; Willson, B. D.
Theoretical Maximum Algal Oil Production. BioEnergy Res. 2010, 3,
204213.
(28) Vonshak, A.; Tomaselli, L. Arthrospira (Spirulina): systematics
and ecophysiology. In Ecology of Cyanobacteria; Whitton, B. A., Potts,

KL = mass transfer coecient for CO2 transfer from gas


phase to liquid phase (m d1)
Ks = inhibition constant for CO2 (mol CO2 m3)
kw = water (background) turbidity (m1)
kT,1,kT,2 = tting constants of attenuation factor for
temperature limitation (C2)
L = length of the CSTR in the open pond (m)
Le = latent heat of evaporation (cal g1)
Meff = eciency of the paddle wheel mixing system
= growth rate of microalgae (d1)
max = maximum algae growth rate (d1)
n = number of compartments the open pond is segmented to
n0 = Gauckler-Manning coecient; a roughness factor (d
m1/3)
N0 = number of bubbles of gas formed at the sump bottom
(s1)
Pair = vapor pressure in the overlaying air (mmHg)
PC = algal biomass production cost ($ tonne1)
Pg = pressure of CO2 gas (atm)
Pr = productivity of algal biomass in the open pond (g DW
m2 d1)
Psat = saturation vapor pressure for a given water temperature
(mmHg)
Qg = CO2 gas ow rate (m3 d1)
r = hydraulic radius of pond channel (m)
Rg = universal gas constant (atm m3 mol1 K1)
RCO2 = carbon dioxide requirement (mol CO2 g1 DW)
RChl = algal biomass chlorophyll content (g Chl kg1 DW)
= density of water (g cm3)
CO2 = density of CO2 gas (g cm3)
= StefanBoltzmann constant (cal cm2 d1 K4)
tp = length of a day in hours (24 h)
Tair = daily average local air temperature (C)
TB = reference temperature for metabolic rate (C)
Tc = time horizon of the production cycle
Tg = temperature of CO2 gas (C)
Topt = optimal temperature for algae growth (C)
Tw = temperature of the water in the CSTR (C)
U = wind speed above the open pond water (m s1)
v = water velocity in the open pond (cm s1)
vb = gas terminal velocity (cm s1)
W = width of the CSTR in the open pond (m)
Ws = width of the CO2 gas sump station (m)
xCO2 = price of CO2 ($ tonne1)
xe = price of electricity ($ kWh1)
xw = price of agricultural water ($ m3)
yin/out = CO2 mole in the inlet/outgassing bubbles
z = vertical distance from the water surface in the pond (m)

REFERENCES

(1) Olivier, J. G. J.; Janssens-Maenhout, G.; Muntean, M.; Peters,


J. A. H. W. Trends in global CO2 emissions: 2014 Report; PBL
Netherlands Environmental Assessment Agency: The Hague, 2014.
(2) Davenport, C. A Climate Accord Based on Global Peer Pressure.
http://www.nytimes.com/2014/12/15/world/americas/lima-climatedeal.html (accessed June 1, 2015).
(3) Chisti, Y. Biodiesel from microalgae. Biotechnol. Adv. 2007, 25,
294306.
(4) Weissman, J. C.; Goebel, R. P.; Benemann, J. R. Photobioreactor
Design: Mixing, Carbon Utilization, and Oxygen Accumulation.
Biotechnol. Bioeng. 1988, 31, 336344.
(5) Sander, K.; Murthy, G. S. Life Cycle Analysis of Algae Biodiesel.
Int. J. Life Cycle Assess. 2010, 15, 704714.
3336

DOI: 10.1021/acs.iecr.5b03209
Ind. Eng. Chem. Res. 2016, 55, 33273337

Article

Industrial & Engineering Chemistry Research

(49) Model Developer Guide, version 3.5; Process Systems Enterprise


Limited: London, U.K., 2012.
(50) Stefan, H. G.; Cardoni, J. J.; et al. Model of Light Penetration in
a Turbid Lake. Water Resour. Res. 1983, 19, 109120.
(51) Bhattacharya, S.; Shivaprakash, M. Evaluation of three Spirulina
species grownunder similar conditions for their growthand biochemicals. J. Sci. Food Agric. 2005, 85, 333336.
(52) National Alliance For Advanced Biofuels And Bioproducts
Synopsis (NAABB): Full Final Report Section II. http://www.energy.
gov/eere/bioenergy/downloads/national-alliance-advanced-biofuelsand-bioproducts-synopsis-naabb-nal (accessed Aug 18, 2015).
(53) Vonshak, A. Spirulina: growth, physiology and Biochemistry. In
Spirulina platensis (Arthrospira): physiology, cell-biology, and biotechnology; Vonshak, A., Ed.; Taylor & Francis: London, U.K., 1997; pp 43
65.
(54) Volkman, J. K.; Brown, M. R.; Dunstan, G. A.; Jeffrey, S. W. The
Biochemical Composition of Marine Microalgae from the Class
Eustigmatophyceae. J. Phycol. 1993, 29, 6978.
(55) National Solar Radiation Data Base. 19912005 Update:
Typical Meteorological Year 3. http://rredc.nrel.gov/solar/old_data/
nsrdb/1991-2005/tmy3/ (accessed June 16, 2015).
(56) Becker, E. W. Microalgae: Biotechnology and Microbiology;
Cambridge University Press: New York, NY, 1994.
(57) Hughmark, G. A. Holdup and Mass Transfer in Bubble
Columns. Ind. Eng. Chem. Process Des. Dev. 1967, 6, 218220.
(58) Talbot, P.; Gortares, M. P.; Lencki, R. W.; Noiie, J. d. l.
Absorption of CO2 in Algal Mass Culture Systems: A Different
Characterization Approach. Biotechnol. Bioeng. 1991, 37, 834842.
(59) Carvalho, A. P.; Malcata, F. X. Transfer of Carbon Dioxide
within Cultures of Microalgae: Plain Bubbling versus Hollow-Fiber
Modules. Biotechnol. Prog. 2001, 17, 265272.
(60) About IID Water. http://www.iid.com/water/about-iid-water
(accessed June 29, 2015).
(61) Wholesale Electricity and Natural Gas Market Data. http://
www.eia.gov/electricity/wholesale/index.cfm (accessed Aug 27, 2015).
(62) Leading options for the capture of CO2 emissions at power stations,
Technical Report PH3/14; IEA greenhouse gas R&D programme:
Cheltenham, U.K., 2000.
(63) Zeiler, K. G.; Heacox, D. A.; Toon, S. T.; Kadam, K. L.; Brown,
L. M. The use of microalgae for assimilation and utilization of carbon
dioxide from fossil fuel-fired power plant flue gas. Energy Convers.
Manage. 1995, 36, 707712.
(64) Richmond, A. Outdoor Mass Cultures of Microalgae. In
Handbook of Microalgal Mass Culture; Richmond, A., Ed.; CRC Press,
Inc.: Boca Raton, FL, 1986; pp 285329.
(65) Optimisation Guide, version 3.5; Process Systems Enterprise
Limited: London, U.K., 2012.

M., Eds.; Kluwer Academic Publishing: The Netherlands, 2000; pp


505523.
(29) Griffiths, M. J.; Harrison, S. L. Lipid productivity as a key
characteristic for choosing algal species for biodiesel production. J.
Appl. Phycol. 2009, 21, 493507.
(30) Belay, A.; Ota, Y.; Miyakawa, K.; Shimamatsu, H. Current
knowledge on potential health benefits of Spirulina. J. Appl. Phycol.
1993, 5, 235241.
(31) Boussiba, S.; Vonshak, A.; Cohen, Z.; Avissar, Y.; Richmond, A.
Lipid and Biomass Production by the Halotolerant Microalga
Nannochloropsis salina. Biomass 1987, 12, 3747.
(32) Grima, E. M.; Fernandez, F. G. A.; Camacho, F. G.; Chisti, Y.
Photobioreactors: light regime, mass transfer, and scaleup. J. Biotechnol.
1999, 70, 231247.
(33) Van Oorschot, J. L. P. Conversion of Light Energy in Algal
Cultures; Med. van. Lund. Wang.; Wageningen Veenman & Zonen,
1955; Vol. 55, pp 225277.
(34) Huesemann, M. H.; Van Wagenen, J. V.; Miller, T.; Chavis, A.;
Hobbs, S.; Crowe, B. A Screening Model to Predict Microalgae
Biomass Growth in Photobioreactors and Raceway Ponds. Biotechnol.
Bioeng. 2013, 110, 15831594.
(35) Steele, J. H. Microbial Kinetics and Dynamics in Chemical
Reactor Theory. In Chemical reactor theory: a review; Lapidus, L.,
Amundson, N. R., Eds.; Prentice-Hall: Englewood Clis, NJ, 1977; pp
405483.
(36) Riley, G. A. II. Physical Oceanography. Oceanography of Long
Island Sound, 19521954; Bulletin of the Bingham Oceanographic
Collection; Yale University, New Haven, CT, 1956; Vol. 15, pp 15
46.
(37) Grobbelaar, J. U. Algal Nutrition. In Handbook of Microalgal
Culture Biotechnology and Applied Phycology; Richmond, A., Ed.;
Blackwell: Oxford, U.K., 2006; pp 97115.
(38) Kaplan, D.; Richmond, A. E.; Dubinsky, Z.; Aaronson, S. Algal
Nutrition. In Handbook of Microalgal Mass Culture; Richmond, A., Ed.;
CRC Press: Boca Raton, FL, 1986; pp 147198.
(39) McGinnis, D. F.; Little, J. C. Predicting diffused-bubble oxygen
transfer rate using the discrete-bubble model. Water Res. 2002, 36,
46274635.
(40) Weissman, J. C.; Goebe, R. P. Design and Analysis of Microalgal
Open Pond Systems for the Purpose of Producing Fuels: a Subcontract
Report, SERI/STR-231-2840; Solar Energy Research Institute: Golden,
CO, 1987.
(41) Hudson, N. W. Field measurement of soil erosion and runo; Food
and Agriculture Organization of the United Nations: Rome, Italy,
1993.
(42) gPROMS, version 4.0; Process Systems Enterprise: London,
U.K., 2015; www.psenterprise.com/gproms.
(43) Jimenez, C.; Cosso, B. R.; Labella, D.; Niell, F. X. The
Feasibility of industrial production of Spirulina (Arthrospira) in
Southern Spain. Aquaculture 2003, 217, 179190.
(44) Jimenez, C.; Cosso, B. R.; Niell, F. X. Relationship between
physicochemical variables and productivity in open ponds for the
production of Spirulina: a predictive model of algal yield. Aquaculture
2003, 221, 331345.
(45) Zarrouk, C. Contribution a letude dune cyanophycee: inuence
de divers facteurs physiques et chimiques sur la croissance et la
photosynthese de Spirulina maxima (Setch et Gardner) Geitler. Ph.D.
Thesis, University of Paris, Paris, France, 1966.
(46) Cheng-Wu, Z.; Zmora, O.; Kopel, R.; Richmond, A. An
industrial-size flat plate glass reactor for mass production of
Nannochloropsis sp. (Eustigmatophyceae). Aquaculture 2001, 195,
3549.
(47) Satel-Light, The European Database of Daylight and Solar
Radiation. http://www.satel-light.com/indexs.htm (accessed June 16,
2015).
(48) Tutiempo Network, S.L., Climate Malaga/Aeropuerto - Climate
data (84820). http://en.tutiempo.net/climate/07-1997/ws-84820.
html (accessed June 16, 2015).
3337

DOI: 10.1021/acs.iecr.5b03209
Ind. Eng. Chem. Res. 2016, 55, 33273337

Você também pode gostar