You are on page 1of 15

Biochemical Engineering Journal 14 (2003) 5165

Online estimation of stirred-tank microalgal photobioreactor cultures


based on dissolved oxygen measurement
Jian Li, Ning Shou Xu, Wei Wen Su
Department of Molecular Biosciences and Bioengineering, University of Hawaii at Manoa, Manoa, HI 96822, USA

Accepted after revision 5 August 2002

Abstract
Photobioreactor sensing presents unique challenges not met in conventional fermentors. First and foremost, photobioreactor processes
are governed by photosynthesis, and hence many parameters important to photobioreactors are not considered in conventional fermentors.
Furthermore, photobioreactor processes are typically associated with stringent cost constraints, and thus the use of complex sensing hard-
ware is precluded. This calls for innovative approaches to address sensing problems in photobioreactors. Here, we report the development
of an effective model-based estimator that is capable of tracking key culture states in stirred-tank microalgal photobioreactor systems. A
marine micro-alga Dunaliella salina was used as a model organism in this study. Extended Kalman filter (EKF) was applied here to provide
optimal estimates of photobioreactor states, based on a dynamic process model in conjunction with online dissolved oxygen measurement.
The process model consists of a growth model and a light transport model. The former associates growth to average light intensity in the
reactor, while taking into account both photoinhibition and oxygen inhibition. The latter is based on a modified radial model to estimate
the average light intensity. The estimator is capable of estimating biomass density, specific growth rate, dissolved oxygen concentration,
photosynthetic efficiency, and average light intensity in the photobioreactor illuminated either with constant incident light at different in-
tensity levels or with time-varying incident lights. For the latter, an auxiliary internal model EKF was used to accurately track the variation
rate of the incident lights. This paper also presents a detailed analysis on the tuning of EKF for optimal estimation. This state estimation
system offers a cost-effective means for monitoring the process dynamics of microalgal photobioreactor cultures online, through which
the productivity of such a process could be optimized.
2002 Elsevier Science B.V. All rights reserved.
Keywords: Biokinetics; Bioprocess monitoring; Bioreactors; Extended Kalman filter; Microalgae; State estimation

1. Introduction from power plant stack gas). In addition, bioreactor-based


photosynthetic microalgal cultures are being consid-
Photobioreactors are used for culturing photosynthetic mi- ered as a part of the Closed Ecological Life Support
croorganisms such as microalgae, cyanobacteria, plant cells, System [1].
and photosynthetic bacteria, for various biotechnological ap- Process sensing is imperative for optimizing photobiore-
plications. Large-scale production of cyanobacteria biomass actor productivity. With the ability to accurately monitor
as food supplement has been in commercial operation for the process dynamics, timely control actions can be taken
many years. There also exist several microalgal-based pro- to assure an optimal environment in the photobioreactor.
cesses for producing nutraceutical-type products, such as The aim of this work is to develop an effective sensing
polyunsaturated fatty acids (e.g. eicosapentaenoic acid and system for estimating key culture parameters in microal-
docosahexaenoic acid) and carotenoids (e.g. astaxanthin), gal photobioreactors. Process sensing in photobioreactors
as well as specialty chemicals (such as radio-labeled com- presents unique challenges unparalleled in conventional fer-
pounds). In aquaculture, microalgal cultures are used as mentors. First and foremost, photobioreactor processes are
feed. Large-scale microalgal cultures also find applications governed by photosynthesis, and hence many parameters
in energy (e.g. biohydrogen) production and environmental important to photobioreactors are not considered in conven-
remediation (e.g. wastewater treatment and removal of CO2 tional fermentors. Furthermore, photobioreactor processes
are typically associated with stringent cost constraints, and
Corresponding author. Tel.: +1-808-956-3531; fax: +1-808-956-3542. thus the use of complex sensing hardware is precluded. This
E-mail address: wsu@hawaii.edu (W.W. Su). calls for innovative approaches to address sensing problems

1369-703X/02/$ see front matter 2002 Elsevier Science B.V. All rights reserved.
PII: S 1 3 6 9 - 7 0 3 X ( 0 2 ) 0 0 1 3 5 - 3
52 J. Li et al. / Biochemical Engineering Journal 14 (2003) 5165

Nomenclature in photobioreactors. In microalgal photobioreactors, pH,


dissolved oxygen, temperature, and light intensity are com-
A partial derivative matrix, defined in Eq. (6) monly monitored. Measurement of these parameters alone,
C partial derivative matrix for calculating the while useful, does not provide much in-depth information
filtering gain, defined in Eq. (6) on culture physiological states. Meanwhile, electronic sen-
E0 parameter in the light model (Eq. (12)) sors for measuring a variety of other culture and process
E1 parameter in the light model (Eq. (12)) parameters (such as culture turbidity, sugar concentration,
h[ ] measurement function and metabolic activities) are available, however, they are
Iav average light intensity in the culture quite costly. In order to monitor multiple parameters, a man-
Im average light intensity at which cell ifold of sensors will need to be installed, which makes it
growth is ceased impractical.
IM amplitude in Eq. (23) In this study, we took a state estimation approach for
I0 incident light intensity on the reactor online detection of key state variables in microalgal photo-
external surface bioreactors. State estimation entails the use of process mod-
Ir changing rate of incident light intensity els together with limited process measurements to provide
kl a volumetric oxygen mass transfer coefficient noise-filtered measurements, estimates of system states that
K filtering gain matrix are not readily measurable, and identification of uncertain
Ka biomass light absorption coefficient system dynamics [2]. Specifically, extended Kalman filter
KI saturation constant (EKF) was used here to provide optimal estimates of relevant
n parameter in the light model (Eq. (12)) culture states in the photobioreactor, based on a dynamic
O dissolved oxygen concentration process model in conjunction with readily available online
O saturated oxygen concentration in the media process measurements. This sensing approach requires only
Om dissolved oxygen concentration at which minimal instrumentation and cost. In addition, it is versatile
cell growth is ceased enough that several culture states can be estimated in paral-
P covariance matrix of state estimation error lel without needing complex sensor hardware. In this study,
Q covariance matrix of system noise an EKF estimation scheme was developed that is capable of
R covariance matrix of measurement noise monitoring biomass density, specific growth rate, dissolved
R reactor radius (a constant) oxygen level, photosynthetic efficiency (in terms of pho-
Ro net oxygen production rate tosynthetic oxygen evolution), and average light intensity
Roev specific photosynthetic oxygen evolution rate in the reactor, based on incident light information and on-
Rore specific respiration rate line dissolved oxygen measurement. A marine micro-alga
Romax max. oxygen generation rate Dunaliella salina was used as the model organism.
Romin equivalent to the specific respiration rate
v measurement noise vector
Vo measurement noise of dissolved oxygen 2. Materials and methods
concentration
VI measurement noise of I0 2.1. Microorganism and medium
w dynamic noise vector
Wi noise The green alga D. salina (Teod.) UTEX 1644 was cul-
X cell density (cell dry weight) tured in the chemically defined hypersaline liquid medium
Y measurement vector containing 1.5 M NaCl, 15 mM KNO3 , 1 mM K2 HPO4 ,
Y measurement of dissolved oxygen 1 mM CaCl2 , 5 mM MgSO4 , 25 mM NaHCO3 and 40 mM
concentration TrisHCl at pH 7.4, supplemented with a mixture of
YI measurement of I0 micronutrients containing, in final concentration, 50 M
Y0 yield coefficient H3 BO3 , 50 M EDTA, 10 M MnCl2 , 5 M FeCl3 , 2 M
Na2 MnO4 , 1.5 M NaVO3 , 0.8 M ZnSO4 , 0.4 M CuSO4 ,
Greek letters 0.2 M CoCl2 , and CO2 as carbon source. This medium
[ ] dynamic state function was modified from that used by Pick et al. [3], and the
specific growth rate (h1 ) modification was made based on biomass composition of
max maximum specific growth rate major nutrient elements and verified via growth studies
instantaneous angular frequency of I0 variation (Joanne Radway, personal communication). The medium
phase angle in Eq. (23) was autoclaved at 121 for 20 min before use. The culture
state vector was maintained in Erlenmeyer flasks, illuminated using
state estimate dim cool-white fluorescent lights (<100 E m2 s1 ) under
I internal model state vector a 12 h/12 h light/dark cycle at room temperature without
shaking, and sub-cultured once every two weeks.
J. Li et al. / Biochemical Engineering Journal 14 (2003) 5165 53

system consists of a pH sensor (405-DPAS-SC-K8S/200,


Metter Toledo, Urdorf, Switzerland), a pH meter (model
430, Corning, Corning, New York), a CO2 mass flow con-
troller (FMA-A2404, Omega Engineering, Stamford, CT),
and a supervisory computer which receives pH signals from
the pH meter and activates the mass flow controller through
a data acquisition board (AT-MIO-16DE-10, National In-
struments, Austin, TX). The PID algorithm was coded in
LabVIEW (National Instruments, Austin, TX). The flow
rate of the aeration stream was controlled using a similar
mass flow controller via LabVIEW. The photobioreactor was
also equipped with a dissolved oxygen sensor (InPro6000,
Ingold-Metter-Toledo, Wilmington, MA) connected to an
oxygen meter (model 01971-00, Cole Parmer, Vernon Hills,
IL), and the signals were logged into the supervisory com-
puter through the data acquisition board. The volumetric
oxygen transfer coefficient, kl a, in the photobioreactor was
measured using the gassing-in method [4].
All the experiments were performed at a temperature of
29 C which was controlled by the BioFloIII thermostat
unit. The cell concentration of D. salina culture was deter-
mined by measuring the culture turbidity at 750 nm using a
spectrophotometer (UV60, Shimadzu, Columbia, MD). The
turbidity data were converted to cell dry weight based on a
calibration curve. Cell dry weight was determined directly
in some of the culture samples (those with higher cell con-
centrations) to validate the dry weight data derived from
turbidity measurement. The cell dry weight concentration of
D. salina culture was determined by filtering the culture sus-
Fig. 1. Schematic representation of the photobioreactor system.
pensions through pre-dried and pre-weighed 0.45 m mem-
brane filters. The cells on the filter were then rinsed with a
0.3 M NaCl solution to remove excess salts from the culture
2.2. Bioreactor culture conditions and measurements medium, and dried in an oven at 90 C till constant weight.
We have confirmed experimentally that the growth of D.
Bioreactor cultures of D. salina were conducted in an salina was not limited by the macro nutrients (phosphate,
instrumented bench-top photobioreactor. This reactor was iron, and nitrate) under the culture conditions employed
modified from a 3 l stirred-tank fermenter (BiofloIII, New in this study, indicated by the residual nutrients detected
Brunswick Scientific, Edison, NJ). The original BioFloIII in the medium after each culture experiments (data not
culture vessel contains a dished stainless steel bottom cham- shown) and the increased biomass production correspond-
ber which also serves as a heat exchanger. This chamber ing to the increases in illumination intensities. Furthermore,
takes up about 1/5 of the reactor working volume, prevent- by considering the equilibrium concentrations of dissolved
ing uniform illumination throughout the reactor, and hence carbon dioxide species, molecular carbon dioxide should
it was replaced with a flat stainless steel base plate (Fig. 1). also be in excess, under the condition that the culture pH
In addition, a jacketed and lengthened glass cylinder (with was controlled at 7.4 by bubbling CO2 gas [4]. The incident
an inner diameter and height of 13.5 and 25.0 cm, respec- light intensity on the reactor surface was measured using a
tively) was used in place of the original BioFlo glass jar. flat-surface quantum sensor (LI-190SA, LI-COR, Lincoln,
The photobioreactor was illuminated by an external light NE) connected to a micrologger (21X, Campbell Scientific,
source. The light source was constructed using six compact Logan, UT). Because the average light intensity in the reac-
fluorescent light bulbs (BiaxTM D/E, F26DBX/840/4P, GE tor cannot be measured directly, we measured the local light
Lighting, Cleveland, OH) connected to two dimmable bal- intensity at 30 distinct locations inside the photobioreactor
lasts (Mark X REZ-3S32, Advance Transformer Co., Rose- using a micro quantum sensor (US-SQS/LI, Heinz Walz
mont, IL), and the light intensity could be adjusted manually GmbH, Effeltrich , Germany) linked to a light meter (LI-250,
via a dimmer (NF-10WH, Lutron Electronics, Coopersburg, LI-COR, Lincoln, NE). This fiber-optic micro light sensor
PA). Culture pH was precisely maintained at pH 7.4 0.05 has a spherical sensor tip (with a diameter of 3 mm) that can
by controlled supplementation of the air feed stream with sense light from all directions, as opposed to the flat-surface
CO2 gas based on a PID control system. The pH control quantum sensors which detect unidirectional light. The light
54 J. Li et al. / Biochemical Engineering Journal 14 (2003) 5165

measurements were taken at five horizontal planes uniformly modeling of the radiant field in photoreactors has been the
located along the vertical axis of the reactor, and with each subject of numerous studies [1014]. Basically, the light
plane containing six evenly distributed measurement points models developed thus far differ in how light propagates and
along the radial axis. The average light intensity was calcu- attenuates inside the reactor. For cylindrical reactors with
lated by taking the weighted average of all measurements. uniform external illumination, light is assumed to propagate
inside the photobioreactor with various degrees of light dif-
2.3. Extended Kalman filter fusion [15]. In the absence of any light diffusion (i.e. light ray
travels in one direction), it results in the so-called radial
The dynamics of a non-linear biochemical process can be model [16]. In cases where light rays are diffused (i.e. light
expressed in the following general form: travels in all directions), a diffuse model has been pro-
posed [15]. The local light intensity is calculated by consid-
(t) = [(t)] + w(t), (t)|t=0 = 0 (1) ering light rays from all directions. When light diffusion is
Y (t) = h[(t)] + v(t) (2) considered to occur only in certain region of the reactor, such
as the center, this leads to the partially diffuse model. Notice
where (t) is the state vector with an initial value of 0 , Y (t) that when the diffuse region is restricted only to the center
the measurement vector, w(t) the system noise, representing of the reactor, the diffuse model becomes the radial model
modeling error and unknown disturbances, and v(t) is the [14]. In terms of light attenuation through cell culture, most
measurement noise. Both system and measurement noises published formulas were based upon the Lambert-Beers
are assumed to be independent random white noises with law. Detailed mechanistic modeling approaches have been
zero mean, with corresponding covariance matrices Q and taken to reflect the actual light absorption and scattering
R. When the measurements are taken continuously in time, phenomena occurred in the heterogeneous photobioreactor
the extended Kalman filter (EKF) algorithm for calculating containing cell particles [12,14,17]. To this end, mechanis-
the optimal state estimate (t) based on the available mea- tic models considering either isotropic [18] or nonisotropic
surement up to current time t is given by [5]: scattering [11] have been reported. These mechanistic mod-
els however are difficult to use in an online state estimation
(t) = [ (t)] + K(t){Y (t) h[ (t)]}, (t)|t=0 = 0 (3) system due to the introduced computational complexity.
Average light intensity models for a cylindrical reactor
with
with uniform external illumination can be derived by first
K(t) = P (t)C(t)R 1 (4) calculating the local light intensity I (r) (at a radial position
r) based on a particular mode of light propagation and the
where K(t) is the filtering gain matrix, and P (t) the covari- formula of light attenuation. Using the assumption that light
ance matrix of filtering error satisfying the following matrix distribution inside the bioreactor vessel is vertically uniform,
Riccati equation: the average light intensity, Iav , can then be calculated as an
integral of I(r) across the horizontal plane of the cylindrical
P (t) = A(t)P (t) + P (t)AT (t) K(t) R K T (t) + Q,
reactor, divided by the reactor cross-sectional area:
P (t)|t=0 = P 0 (5)  R
1
where Iav = 2 rI(r) dr (7)
  R2 0
[(t)]  h[(t)] 
A(t) = , C(t) = (6) In the case of radial light propagation, the following for-
T (t)  (t) T (t)  (t) mula of I(r) was deduced by adopting LambertBeers law
and energy shell balance [15]:
I0 R  Ka X(Rr) 
3. Results and discussions I (r) = e + eKa X(R+r) (8)
r
3.1. Light model development where I0 is the incident light intensity on the reactor external
surface, X the biomass concentration, and Ka is the light
Light availability plays a key role in photobioreactor op- extinction coefficient. Based on the LambertBeers law, Ka
eration. In modeling photobioreactors it is a common prac- was estimated from the light absorption data of the same
tice to link cell growth with the average light intensity in the culture used in the experiment for measuring average light
reactor [68]. Quantification of light distribution in photo- intensity.Substituting Eq. (8) into Eq. (7) and integrating the
bioreactors is a prerequisite for estimating the average light equation, an analytical solution is reached as follows.
intensity. However, modeling of light propagation and atten- 2I0
uation inside a bioreactor is complicated by several factors Iav (I0 , X, Ka ) = (1 e2Ka XR ) (9)
Ka XR
such as spectral and physical properties of the light source,
the geometrical properties of the reactor, optical properties Another formula of I(r) was derived for the case of diffused
of the culture, and physiology of the cells [9]. Mathematical light propagation with the assumption of LambertBeers
J. Li et al. / Biochemical Engineering Journal 14 (2003) 5165 55

Fig. 2. Comparison of three light models.

law [9]. As the biomass concentration X approaches zero, Iav


 approaches I0 E0 E1 . This semi-empirical three-parameter
I0
I (r) = eKa X[r cos + R 2 r 2 sin2 ]
d (10) model can fit the experimental data well, within a large
0 range of cell concentrations (Fig. 2). Compared with en-
Combination of Eq. (10) with Eq. (7) gives another average tirely empirical second-order polynomial models, which
light intensity model also contain three adjustable parameters, Eq. (12) provides
a better fit to the experimental results (data not shown). This
Iav (I0 , X, Ka ) new model implicitly compensates for some of the deficien-
  cies encountered in the radial model, such as light diffusion
2I0 R Ka X[r cos + R 2 r 2 sin2 ]
= r e d dr (11) and deviation from the LambertBeers law. Eq. (12) is
R2 0 0
simple to use, which makes it well suited for incorporation
The previous two Iav models (Eqs. (9) and (11)) were into the state estimator since it reduces the computational
tested against the measured data of average light intensity. load of EKF. Moreover, when used as a part of the state
The radial and diffuse models were simulated, with the estimator, the light model can be tuned dynamically to cope
resulting Iav versus X curves presented in Fig. 2, which with uncertainty of the model structure and/or parameter
indicated that the former overestimated the average light values, via model parameter evolution (i.e. by including
intensity while the latter did the opposite. A number of the model parameters in the state vector with other state
factors may contribute to this discrepancy: (1) the behavior variables). This will be discussed further in Section 3.3.2.
of the actual light source is likely to be between purely ra-
dial and purely diffused; (2) due to varying degrees of light 3.2. Dynamic process model development and parameter
scattering, light attenuation in the culture may not be fully estimation
described by the Lambert-Beers law especially in dense
cultures and long light path conditions; and (3) light inten- A central element of a state estimator is a state model that
sity distribution in the bioreactor is not entirely equal in the adequately depicts the process kinetic behavior. In develop-
vertical direction as assumed in the calculation. In addition, ing such state models, it is customary to keep the model
Ka might not be a constant during the cultivation process structure as simple as possible while retaining the models
due to changes in cellular pigment content [6]. Considering capability to reflect the overall dynamics of the process [2].
the complexity of the system, a semi-empirical approach By doing so, it should reduce the computational load of the
was taken here to derive a model based upon the structure of state estimator, which is particularly significant when the di-
the radial model (mainly because of its mathematical sim- mension of the state vector is large. With this in mind, we
plicity) but included three new parameters (E0 , E1 and n): developed an unstructured model to simulate the dynamics
I0 E0 n of cell density, dissolved oxygen concentration, and average
Iav = n
(1 eE1 X ) (12) light intensity in the photobioreactor for culturing D. salina
X
56 J. Li et al. / Biochemical Engineering Journal 14 (2003) 5165

under light-limited growth conditions. Light and oxygen are describe the process dynamics in terms of the kinetic behav-
two of the most important factors that affect the productiv- iors of biomass and dissolved oxygen concentrations:
ity of microalgal photobioreactor processes. The majority
X = X (13)
of photobioreactor processes are, for the most part, oper-
ated under light-limited conditions. However, photobioreac- O = Ro kl a(O O ) (14)
tor cultures may sometimes be exposed to excessive light
intensity. Under such condition, protein D1 in photosystem where is the specific growth rate, O the dissolved oxygen
II can be damaged leading to photoinhibition and reduced level, O the dissolved oxygen level in the media that is in
growth rate due to reduction in the number of active pho- equilibrium with the bulk gas phase in the sparged air stream,
ton traps [19]. In addition, accumulation of high levels of kl a the volumetric oxygen transfer coefficient, and Ro is
dissolved oxygen resulting from photosynthesis, especially the net oxygen production rate which reflects the combined
in poorly mixed high-density cultures, may lead to growth effect of photosynthetic oxygen evolution and respiration.
inhibition [20]. In the growth experiments reported here, In modeling the specific growth rate, we have taken into
substrate limitation was precluded, indicated by the residual account light limitation, light inhibition, and oxygen inhibi-
nutrients detected in the medium after each culture experi- tion, and the model was based on the average light intensity.
ments (data not shown) and the increased biomass produc- Hyperbolic equation was widely used to model light-limited
tion corresponding to the increases in I0 (refer to the data in algal growth [10,21]. On the basis of the hyperbolic model,
Fig. 3). Other culture variables, such as temperature, media a light inhibitory term and an oxygen inhibitory term were
pH, and oxygen mass transfer, were precisely controlled in included in our specific growth rate model:
our photobioreactor system at constant levels.    
Iav Iav O
The process model reported here was formulated based on = max 1 1 (15)
KI + Iav Im Om
mass balances and saturation-type kinetics, with assumption
of unsegregated, balanced growth [4]. Among the three state where max is the maximum specific growth rate, KI the
variables included in the model (i.e. X, O and Iav ), the dis- light saturation constant, Im represents the light intensity
solved oxygen level (O) could be continuously measured on- level at which the cell growth is totally inhibited, and Om is
line. Biomass and average light intensity are closely linked to the oxygen level at which the cells cease to grow. Similar
photosynthesis (and thus photosynthetic oxygen evolution), models with various modifications have been used success-
and hence both of these state variables should be observable fully in modeling other microalgal cultures [7,22]. Some of
via dissolved oxygen measurement. For a well-mixed batch the reported models accounted for the maintenance require-
photobioreactor, the following mass-balances were used to ment but left out the oxygen inhibitory effect. Considering

Fig. 3. Modeling and filtering results for short-term culture experiments operated under constant I0 . From top to bottom, the values of I0 are 400, 350,
280, and 200 E m2 s1 , respectively. Legend: experimental data for X (), experimental data for O ( ), model (. . . ), filter result (---).
J. Li et al. / Biochemical Engineering Journal 14 (2003) 5165 57

that all cultures in this study were conducted under continu- Table 1
ous illumination and underwent active cell growth, the main- Model parameter values
tenance energy was assumed negligible in our model. The Parameter Value
net oxygen production rate, Ro , was modeled as follows: max 0.119 h1
    KI 202.1 E m2 s1
Iav Iav Im 2005 E m2 s1
Ro = (Roev Rore )X = Romax 1
KI + Iav Im Om 47.9 mg l1
   Ro max 324.2 mg g1 h1
O
1 Romin X (16) Ro min 22.2 mg g1 h1
Om kl a for short-term experiments 13.5 h1
kl a for long-term experiments 6.9 h1
Here, the specific photosynthetic oxygen evolution rate, E0 0.26 g0.87 l0.87
Roev , was assumed to be linearly correlated with the specific E1 10.0 l0.87 g0.87
growth rate, with a ratio of Romax /max between the two n 0.87
rates. Such linear correlation has been reported in the liter-
ature, for example, by Rebolloso Fuentes et al. [23] in their [27] and Neidhardt et al. [28], Romax and Romin were esti-
study of Porphyridium cruentum, and it is justified by the mated from the data of Cao et al. [26] and Loeblich [29], and
fact that cell growth is directly coupled with photosynthe- Om from unpublished data (Joanne Radway, personal com-
sis. The specific respiration rate of many algal species was munication). All these data were obtained using D. salina
observed to be influenced by the average light intensity [24] culture. The multivariable regression was carried out as fol-
as well as by the history of the illumination that the cells ex- lows. First, second-order polynomials were used to fit the
perienced [25]. However, the respiration rate was typically original experimental data (to obtain smoothed profiles of X
less than 10% of the oxygen evolution rate. Therefore, Rore and O together with their derivatives, X and O). The origi-
was considered as a constant here, represented by Roin . nal differential model equations were then transformed into
The state model (i.e. the growth model plus the light an algebraic form (by rearranging the terms in the equa-
model) was fitted to data from culture experiments con- tions). Subsequently, the equations were subject to the re-
ducted under various light conditions, utilizing only one set gression analysis by using the Marquardt non-linear least
of model parameters. Among the model parameters, the oxy- squares method [30]. Values of the model parameters ob-
gen mass transfer coefficient, kl a, was determined experi- tained from the regression analysis are presented in Table 1.
mentally. For the rest of the parameters, the estimation was These values are generally in good agreement with the liter-
done based on a combination of literature data and non-linear ature values. For instance, the value of Romax from regres-
least square fitting of the state model to the data derived sion was 324.2 mg g1 h1 , while 370 and 468 mg g1 h1
from the short-term experiments conducted at four different were reported by Loeblich [29] and Cao et al. [26], respec-
incident light intensity conditions. To initiate the non-linear tively. Using parameters in Table 1, the model simulation
model regression, initial values of the parameters were es- results are presented in Figs. 3 and 4 for short-term (<15 h)
timated from the literature data, and in one case from un- and long-term (>100 h) experiments, respectively. As indi-
published results. The initial values for max was estimated cated in Figs. 3 and 4, the model could capture the overall
from the batch growth data published by Cao et al. [26], KI trends of the key culture states, namely X and O, in both
and Im were estimated from the data of Baroli and Melis short- and long-term cultures, although some discrepancy

Fig. 4. Modeling and filtering results for long-term culture experiments operated under constant I0 (upper panel: I0 = 60 E m2 s1 ; lower panel:
I0 = 100 E m2 s1 ). Legend: experimental data for X (), experimental data for O ( ), model (. . . ), filter result (---).
58 J. Li et al. / Biochemical Engineering Journal 14 (2003) 5165

Fig. 5. Model parameter sensitivity analysis. Legend: ( ) for X, and ( ) for O.

between the model and the experiment was apparent, espe- Iav could not be measured directly and the determination
cially as the cultivation progressed. As to be seen later, the of its value could be affected by several factors such as
model discrepancy can be reduced by using EKF. modeling error associated with the Iav versus X correlation
A parameter sensitivity analysis was conducted to quan- (Eq. (12)), random disturbance on I0 due to ambient lights,
titatively determine the influence of each parameter on the and potential variation in the biomass light absorption effi-
model. The sums of squared differences of the simulation ciency resulted from different extents of pigmentation dur-
results (on X and O) upon a 5% increase or a 5% decrease ing the cultivation process [6]. Therefore, Iav was included
from the base value (as indicated in Table 1) of each param- as one of the estimated states. Combining Eqs. (12)(16)
eter were examined. The results were normalized and the gives the following state model:
relative sensitivity of all parameters presented in Fig. 5. Base
X
on this analysis, the model was found to be most sensitive to  Romax 

max for biomass estimation, and Romax for dissolved oxy- X Romin X + kl a(O O)
O = max
gen estimation. The estimated values for these parameters

were in accordance with the literature data. Moreover, we Iav I0 Iav
have tested an adaptive EKF that included max and Romax I0 I (X) +
I0
in the state vector, to compensate any error that may exist
WX
in the values assigned for these model parameters. This will
be discussed further in Section 3.3.2. + WO (17)
WI
3.3. State estimation where

E1 Xn E0 n
3.3.1. Basic EKF I (X) = n E0 E1 e n (1 eE1 X )
X
The extended Kalman filter algorithm was used to esti- 
n Iav
mate cell density (X), dissolved oxygen concentration (O), = n E0 E1 eE1 X (18)
and average light intensity (Iav ), based on the dynamic I0
growth and light transfer models described earlier, and to Since dissolved oxygen (i.e. O) was used as the only mea-
eliminate noise in the measured state (in this case, O). Dis- sured state, the measurement equation becomes:
solved oxygen was chosen as the measured state because
Y (t) = O(t) + VO (t) = C(t) + VO (t) (19)
it is readily measurable in a continuous mode. In addition,  
dissolved oxygen dynamics is closely related to photosyn- where C = 0 1 0 . In this case, Q = diag[VarWX ,
thesis and cell growth. As such, bioreactor states could be VarWO , VarWI ] and R = VarVO , and the detailed expres-
updated continuously. The photobioreactor states are also sion of A used in Eq. (5) for calculating P (t) is given by
greatly influenced by the average light intensity. However, Eq. (A.1) in the Appendix A.
J. Li et al. / Biochemical Engineering Journal 14 (2003) 5165 59

The covariance matrix R can be determined on the basis When I0 is kept at a constant level, I0 = 0. The state esti-
of the standard deviations corresponding to the noises of mation results under this condition are given in Figs. 3 and 4
the instruments that were used to measure the states. In our for short-term (<15 h) and long-term (>100 h) experiments,
system, as dissolved oxygen was the only measured state respectively. Figs. 3 and 4 show that while the model alone
and the measurement was assumed to be characterized by can generally fit the experimental results for X under dif-
a Gaussian noise with a standard deviation of ca. 3%, we ferent I0 values and different culture durations, propagation
set R = [VarVO ] = [0.001]. The system noise covariance of modeling error was augmented as the cultivation period
matrix Q can be set based on the variance of the model was extended. Note that after applying EKF, the modeling
uncertainty of each state. For the photobioreactor system, errors could be rectified. To further illustrate this point,
Q was chosen as: diag[105 , 103 , 4]. Here, the covariance Fig. 6 shows that when a significant initial estimation error
matrix Q was chosen in a diagonal form according to the exists in X, the model output retains a large error through-
usual assumption that the individual components in w (i.e. out the entire culture period, while the EKF can correct the
system noise vector) are uncorrelated. For the covariance initial estimation error immediately, provided that the value
matrix of initial estimation error, P 0 , a diagonal form was for Px 0 is set correctly (as stated before, Px 0 should be set
assumed at first. The initial value of the states was esti- at a value close to the variance of the initial state estimation
mated based upon off-line measurements. The values of error). This distinctive advantage of EKF is mainly due to
the diagonal elements in the P 0 matrix should be set close the fact that the EKF is able to make full use of the feed-
to the variance of the initial state estimation errors. Based back information picked up from the online measurement
on this rule, P 0 was chosen as: diag[PX0 , PO0 , PIav0 ] = of O, even if the measurement is somewhat noisy.
diag[0.0001, 0.05, 6]. The choice of covariance matrices Upon a closer examination of the estimation results for
is important for state estimation. If fast tracking is required, cases where the initial estimation of X is inaccurate, several
more weight should be placed on the error term in Eq. (3), interesting observations were noted. If estimation of X is
which means small R and large P should be chosen. How- the primary concern, then increasing Px 0 alone (up to 0.1),
ever, weighing too much on the error term will cause the while keeping PIavo unchanged, may well cope with very
system unstable. As the system is non-linear and time large initial estimation errors in X (up to 0.3 g l1 ; note that
variant, the stability of system at different time will not be 0.1 (0.3)2 ), although in this case the estimation of Iav
the same. becomes distorted (Fig. 6A). In the photobioreactor state

Fig. 6. Effect of P 0 on state estimation: (A) effect of Px 0 (while setting Cov(X0 , Iav0 ) = 0), and (B) effect of Cov(X0 ,Iav0 ). Legend: for (A), Px0 = 104
(), Px0 = 103 (. . . ), Px0 > 102 ( ); for (B), Cov(X0 , Iav0 ) = 0 (---); Cov(X0 , Iav0 ) = 0.0245 (). For both (A) and (B): model (---),
experimental data for X (), and experimental data for O ( ).
60 J. Li et al. / Biochemical Engineering Journal 14 (2003) 5165

Fig. 7. Effect of Q on state estimation. Legend: experimental data for X (), experimental data for O ( ), model (---), VarWX = 107 (---),
VarWX = 103 (. . . ), VarWX = 102 (), VarWX = 101 ( ).

model presented here, two of the state variables, X and Iav , in Fig. 7. When Var WX was set too high (>0.01), it re-
are actually negatively correlated (cf. Eq. (12) and Fig. 2), sulted in a significant overestimation of X. Conversely, when
therefore, the assumption of a diagonal P0 matrix may not the value of Var WX was set too low (<107 ), the estima-
be entirely correct. With such an assumption, the initial tion of X almost reduced to the model output. Also noted
estimation errors are considered uncorrelated. As seen in in Fig. 7, when Var WX was set too low, the estimation of
Figs. 3 and 4, however, this assumption did not lead to Iav was distorted. The estimation result was not so sensitive
erroneous estimation since the initial estimation of X was to Var WO and Var WI , and hence the results are not pre-
reasonably accurate in those simulations. If a significant sented here. Fig. 8 illustrates the effect of R on the state
error exists in the initial estimate of X, as shown in Fig. 6A, estimation. When the value for R was set either too high
the assumption of a diagonal P0 matrix may lead to erro- (>0.1) or too low (<0.00005), notable underestimation of X
neous estimation of Iav in the starting stage (during the first was observed, with substantial distortion in Iav and even O.
25 h). To overcome this problem,
a non-zero covariance The simulations shown in Figs. 68 depicted the sensitivity
term Cov(X0 , Iav0 ) = PX0 PI av0 was incorporated of the estimator to the filter settings and could be used as
into the matrix P 0 . This way, the estimation of Iav was guidelines for tuning the extended Kalman filter.
improved compared with the case where diagonal P 0 was We observed in numerical simulations that, if the initial
used, as seen from Fig. 6B. In Fig. 6B, Px0 = 104 , and state estimation error variance P 0 , the dynamical noise
PI av0 = 6. It should be pointed out, however, when the variance Q, and the measurement noise variance R were
initial estimation errors in X exceeded ca. 0.15 g l1 , the all adjusted proportionally, the state estimation results re-
estimation of Iav remained distorted even if the additional mained unchanged. We show here this phenomenon can be
covariance term was incorporated into the matrix P 0 . explained by analyzing Eqs. (3)(5). Eq. (3) implies that for
Next, the effect of Q on the state estimation was exam- a given process the state estimation result depends mainly
ined. Recall that Q = diag[VarWX , VarWO , VarWI ], the on the gain matrix K(t), which is proportional to P (t) and
estimation results under various Var WX values are presented inversely proportional to R as shown by Eq. (4). Inserting

Fig. 8. Effect of R on state estimation. Legend: experimental data for X (), experimental data for O ( ), model (---), filter results with
R = 2 105 ( ), R = 1 (---), R = 101 (. . . ), R = 102 to 104 ().
J. Li et al. / Biochemical Engineering Journal 14 (2003) 5165 61

Eq. (4) into Eq. (5) gives If we intentionally set the initial values of max and Romax
incorrectly (in this case, we set max = 0.099 h1 and
P (t) = A(t)P (t) + P (t)AT (t) Romax = 300 mg g1 h1 ), Fig. 9 shows that the adap-
P (t)C(t) R 1 C T (t)P (t) + Q, tive EKF is capable of adjusting these parameters towards
their correct values (i.e. max = 0.119 h1 and Romax =
P (t)|t=0 = P 0 (20)
324.2 mg g1 h1 ) at a satisfactory rate. Similarly, the adap-
Then, with P 0 = P  0 , Q(t) = Q (t), and R(t) = tive EKF was also able to rectify deviations in the initial
R  (t) (i.e. P 0 , Q and R are adjusted proportionally), settings of E0 and E1 with respect to their correct values
Eq. (20) becomes (E0 = 0.26 g0.87 l0.87 and E1 = 10 l0.87 g0.87 ) as shown
in Fig. 9, while setting PE0 0 = 0.1, VarWE0 = 0.0000002 for
P (t) = A(t)P (t) + P (t)AT (t) E0 , and PE1 0 = 1; VarWE0 = 0.31 for E1 . The estimation
P (t)C(t) (R  )1 C T (t)P (t) + Q , curve for O in these cases is similar to that in Fig. 4 and
thus not presented here again.
P (t)|t=0 = P 0 (21.a)
or 3.3.3. Auxiliary internal model EKF
In outdoor applications, the light intensity on the pho-
P (t) = A(t) P (t) + P (t)A (t)
1 1 1 T tobioreactor surface, I0 , is time-varying. Under such a
condition, one needs accurate information for both I0 and
1 P (t)C(t) (R  )1 C T (t) 1 P (t) + Q ,
its derivative Ir , in order to achieve precise state estimation
P (t) 
|t=0 = P 0 (21.b) (recall Eq. (17) which represents the state model for the
photobioreactor process). This is complicated by the fact
Therefore, the new filtering error variance obtained under that, under outdoor conditions, I0 measurement is typically
the same A(t) and C(t) should be P  (t) = (1/)P (t). This associated with substantial noises, and hence calculation
leads to the following conclusion: of its derivative can be prone to errors. To cope with the
K  (t)=P  (t)C(t)(R  )1 = 1 P (t)C(t)( 1 R)1 =K(t) (22) filtering problem associated with I0 and Ir , we took the fol-
lowing internal model approach. Successful implementation
and hence the state estimates (t) remain unchanged. Ac- of such an approach in state estimation has been reported
cording to Eq. (6), this in turn guarantees that A(t) and C(t) in discrete-time systems [31]. With the internal model ap-
will not change. The previous analysis reveals that, when proach, we assume that I0 satisfies a sinusoidal function
choosing P 0 , Q and R, it is important to pay attention to with an unknown but almost constant angular frequency
their relative values. within a small time interval around current time instant t, i.e.
I0 = IM sin(t + ) (23)
3.3.2. Adaptive EKF
In practice, uncertainty may exist in model structure and where IM is the amplitude, and is the phase angle. Then,
measurement precision, and certain model parameters may
possess slowly time-varying property. To cope with these Ir = I0 = IM cos(t + ) (24)
factors, two versions of adaptive EKF were considered here and
to allow parameter evolution, by including in the state vec-
tor, max and Romax in one case, and E0 and E1 in the other, Ir = I0 = 2 IM sin(t + ) = 2 I0 (25)
in addition to the common state variables X, O and Iav (i.e. Thus by setting a state vector as I = [I0 , Ir , ]T to
= [X, O, Iav , max , Romax ]T or = [X, O, Iav , E0 , E1 ]T ). include the unknown but approximately constant frequency
The parameters, max and Romax were selected based on , an internal model state equation for I0 can be derived as
results from the sensitivity analysis, while E0 and E1 were follows
chosen due to their potential time-varying property. Here we
consider max and Romax or E0 and E1 to be relatively sta- I0 Ir WI0

tionary throughout the cultivation, and thus the dynamics of Ir = 2 I0 + WIr (26)
the parameter change can be modeled as i = 0 + Wi where 0 W
i and Wi represent the adaptive model parameters and the
noises, respectively. Here the corresponding measurement and the corresponding measurement equation is
equation remains the same as Eq. (19) in form, except that YI = I0 + VI = C I I + VI (27)
C = [ 0 1 0 0 0 ]. The detailed expressions of A used
in Eq. (5) for calculating P (t) in the two versions of adaptive where C I = [ 1, 0, 0 ].
EKF are given by Eqs. (A2)(A4) in the Appendix A. Notice that here the amplitude IM and the phase angle
The parameter evolution process for max and Romax do not appear in the internal model state Eq. (26). We need
is illustrated in Fig. 9 while setting P0 = 0.1 VarW = only to consider the angular frequency . This is an advan-
0.000031 for max , and PR0 = 3, VarWR = 3125 for Romax . tage of the internal model approach. Since this state equation
62 J. Li et al. / Biochemical Engineering Journal 14 (2003) 5165

Fig. 9. Adaptive EKF with parameter evolution. Legend: EKF with adaptive max and Romax ( ), EKF with adaptive E0 and E1 (---).

Fig. 10. State estimation under varying incident light intensity. Legend: experimental data for X (), experimental data for O ( ), measured data
for I0 ( ), model (. . . ), filtering result ().
J. Li et al. / Biochemical Engineering Journal 14 (2003) 5165 63

is non-linear, an auxiliary EKF is also needed for estimat- major states of D. salina culture grown in a stirred-tank
ing I . The filter parameters for such an auxiliary EKF are: photobioreactor. The estimator was scrutinized under a wide
QI = diag[VarWI0 , VarWIr , VarW ], RI = VarVI , and range of operating conditions including short- (less than
15 h) and long- term (over 100 h) cultures under different
0 1 0 levels of constant incident light intensities, as well as under
AI = 2 0 2I0 (28) the condition of time-varying incident light intensities. This
0 0 0 state estimation system offers a cost-effective means for
Once I0 and its derivative Ir are calculated from the monitoring the process dynamics of microalgal photobiore-
internal model, the values are plugged into Eq. (17) for actor cultures online, and could substantially improve the
bioreactor state estimation. Fig. 10 shows the state estima- productivity of such a process.
tion results under stepwise varying I0 , which approximately
simulates the daily sunlight variation. The parameters for
Acknowledgements
the auxiliary EKF were selected as follows. Considering
that the time-varying I0 function contains a first angular
frequency component 1 2 /18 0.35 rad h1 with The D. Salina strain was obtained from Dr. Anastasios
a period of about 18 h (corresponding to the total culture Melis at the University of California, Berkeley. This work
period), and a second angular frequency component 2 was supported by the NSF ERC Program. Contract grant
2/1 6.28 rad h1 with a period of 1 h (corresponding number: EEC-9731725.
to the hourly stepwise change in I0 ), an intermediate value
1.4 rad h1 was chosen as the initial estimation of the angu-
lar frequency . Since I0 by definition is exactly equal to Appendix A. Appendix: Detailed expressions of A for
Ir , and hence VarWI0 = 0. RI = VarVI = 40 was chosen calculating P (t)
based on the precision of the light meter. The remaining
parameters do not affect the estimation result significantly In the basic EKF (Section 3.3.1), A is expressed as:



X X
O Iav

Ro max Ro max Ro max
Ro min X kl a X
A= = A (A.1)
max max O max Iav


I (X) Ir
I0 I0 I (X) I0 I (X) +
X O Iav I0
In the two versions of adaptive EKF (Section 3.3.2):

and were chosen empirically as VarWIr = 10, VarW = A B
A= (A.2)
0.007, PI0 0 = 100, PIr 0 = 100, P0 = 0.007. It can be 023 022
seen from Fig. 10 that, by incorporating such an auxiliary
internal model EKF whose outputs can well track the I0 for = [X, O, Iav , max , Romax ]T , B is expressed as:

and its variation rate Ir , the EKF offers satisfactory state
estimations under time-varying I0 condition, while the model max X 0

alone fails to give proper outputs.

B= 0 X (A.3)
max

4. Conclusion
I0 I (X) 0
max
A simple unstructured process model coupled with a
semi-empirical light model was shown here to give a reason- and for = [X, O, Iav , E0 , E1 ]T , B is expressed as:

able representation of the D. salinaphotobioreactor process
dynamics, although with the model alone it was not pos- X X
E0 E1
sible to accurately track all culture states, especially under

long-term cultivations and time-varying illuminating con- B = Ro max Iav Ro max Iav (A.4)
X X
ditions. By incorporating EKF with the model, and based max Iav E0
max Iav E1

only on online measured dissolved oxygen concentration,
a state estimator was successfully developed to track the 0 1
64 J. Li et al. / Biochemical Engineering Journal 14 (2003) 5165

where [5] A.H. Jazwinski, Stochastic Processes and Filtering Theory, Academic
Press, New York, 1970.
I (X) [6] E. Molina Grima, F. Garca Camacho, J.A. Snchez Prez, J.M.
Fernndez Sevilla, F.G. Acin Fernndez, A. Contreras Gmez, A
X    mathematical model of microalgal growth in light-limited chemostat
E1 1 E1 Xn 1
= nE0 + nE21 X n1 e culture, J. Chem. Tech. Biotechnol. 61 (1994) 167173.
X X n+1 X n+1 [7] M. Taya, M. Miya-Oka, Y. Toyo-Oka, M. Kino-Oka, S. Tone, K.
Ono, Growth characteristics of liverwort cells, in a photoautotrophic
(A.5) suspension culture, J. Ferment. Bioeng. 80 (1995) 580585.
  [8] J.F. Cornet, J. Albiol, Modeling photoheterotrophic growth kinetics of
max Iav Iav Rhodospirillum rubrum in rectangular photobioreactors, Biotechnol.
= 1 (A.6)
O Om (KI + Iav ) Im Prog. 16 (2000) 199207.
[9] J.C. Ogbonna, H. Yada, H. Tanaka, Light supply coefficient: a
new engineering parameter for photobioreactor design, J. Ferment.
  Bioeng. 80 (1995) 369376.
O max
= 1 [10] E.G. Evers, A model for light-limited continuous cultures: growth,
Iav Om (KI + Iav ) shading, and maintenance, Biotechnol. Bioeng. 38 (1991) 254259.
    [11] H. Kurata, S. Furusaki, Nonisotropic scattering model for estimation
KI Iav Iav of light absorption rates in a suspension culture of Coffea arabica
1 (A.7) cells, Biotechnol. Prog. 9 (1993) 8692.
KI + Iav Im Im [12] J.F. Cornet, C.G. Dussap, J.B. Gros, Conversion of radiant light
   energy in photobioreactors, Am. Inst. Chem. Eng. J. 40 (1994) 1055
Iav Iav O 1066.
= 1 1 (A.8) [13] A. Brucato, L. Rizzuti, Simplified modeling of radiant fields in
max KI + Iav Im Om heterologous photoreactors. 1. Case of zero reflectance, Ind. Eng.
Chem. Res. 36 (1997) 47404747.
Iav [14] T. Katsuda, T. Arimoto, K. Igarashi, M. Azuma, J. Kato, S.
= , i = 0, 1 (A.9)
Iav Ei
Takakuwa, H. Ooshima, Light intensity distribution in the external
Ei illuminated cylindrical photo-bioreactor and its application to
hydrogen production by Rhodobacter capsulatus, Biochem. Eng. J.
  5 (2000) 157164.
I (X) Ir Iav
i = I0 I (X) + + , i = 0, 1 [15] T. Matsuura, J.M. Smith, Light distribution in cylindrical
Ei Ei I0 Ei photoreactors, Am. Inst. Chem. Eng. J. 16 (1970) 321324.
[16] J.E. Huff, C.A. Walker, The photochlorination of chlorform in
(A.10) continuous flow systems, Am. Inst. Chem. Eng. J. 8 (1962) 193200.
[17] Y.S. Yun, J.M. Park, Attenuation of monochromatic and
Iav I0 n Iav
(1 eE1 X ) =
polychromatic lights in Chlorella vulgaris suspensions, Appl.
= n
(A.11)
E0 X E0 Microbiol. Biotechnol. 55 (2001) 765770.
[18] J. Koizumi, S. Aiba, Significance of the estimation of light absorption
Iav
rate in the analysis of growth of Rhodopseudomonas spheroids, Eur.
n
= I0 E0 eE1 X (A.12) J. Appl. Microbiol. Biotechnol. 10 (1980) 113123.
E1 [19] G.H. Krause, Photoinhibition of photosynthesis: an evaluation of
damaging and protective mechanism, Physiol. Plant 74 (1988) 566
 
I (X) n 1 Iav 574.
= n nE1 eE1 X (A.13) [20] F.J. Marquez, K. Sasaki, N. Nishio, S. Nagai, Inhibitory effect
E0 I0 E0 of oxygen accumulation on the growth of Spirulina platensis,
 Biotechnol. Lett. 17 (1995) 225228.
I (X) n 1 Iav
= n E0 (1 E1 X n )eE1 X
[21] J.F. Cornet, C.G. Dussap, G. Dubertret, Structured model for
(A.14)
E1 I0 E1 simulation of cultures of the cyanobacterium Sprirulina platensis,
in photobioreactors. I. Coupling between light intensity transfer and
growth kinetics, Biotechnol. Bioeng. 40 (1992) 817825.
[22] M.P. Gentile, H.W. Blanch, Physiology and xanthophyll cycle activity
References
of Nannochloropsis gaditanna, Biotechnol. Bioeng. 75 (2001) 112.
[23] M.M. Rebolloso Fuentes, J.L. Garca Snchez, J.M. Fernndez
[1] J.F. Cornet, C.G. Dussap, J.J. Leclercq, Simulation, design and model Sevilla, F.G. Acin Fernndez, J.A. Snchez Prez, E. Molina Grima,
based predictive control of photobioreactors, in: M. Hofman, P. Outdoor continuous culture of Porphyridium cruentum in a tubular
Thonart (Eds.), Engineering and Manufacturing for Biotechnology, photobioreactor: quantitative analysis of the daily cyclic variation of
Kluwer Academic Publishers, Dordrecht, 2001, pp. 227238. culture parameters, J. Biotechnol. 70 (1999) 271288.
[2] G. Stephanopoulos, S. Park, Bioreactor state estimation, in: H.J. [24] J.U. Grobbelaar, B.M.A. Kroon, T. Burger-Wiersma, L.R. Mur,
Rehm, G. Reed, A. Phler, P. Stadler, K. Schgerl (Eds.), Influence of medium frequency light/dark cycles of equal duration
Biotechnology: A Multi-Volume Comprehensive Treatise, Measuring, on the photosynthesis and respiration of Chlorella pyrenoidosa,
Modeling and Control, vol. 4, 2nd ed., VCH, New York, 1991, Hydrobiologia 238 (1992) 5362.
pp. 225249. [25] X. Wu, J.C. Merchuk, A model intergrating fluid dynamics in
[3] U. Pick, L. Karni, M. Avron, Determination of ion content and ion photosynthesis and photoinhibition processes, Chem. Eng. Sci. 56
fluxes in the halotolerant algae Dunaliella salina, Plant Physiol. 81 (2001) 35273538.
(1986) 9296. [26] H. Cao, L. Zhang, A. Melis, Bioenergetic and metabolic processes
[4] J.E. Bailey, D.F. Ollis, Biochemical Engineering Fundamentals, 2nd for the survival of sulfur-deprived Dunaliella salina (chlorophyta),
ed., McGraw-Hill, Singapore, 1986. J. App. Phycol. 13 (2001) 2534.
J. Li et al. / Biochemical Engineering Journal 14 (2003) 5165 65

[27] I. Baroli, A. Melis, Photoinhibition and repair in Dunaliella salina [29] L.A. Loeblich, Photosynthesis and pigments influenced by light
acclimated to different growth irradiances, Planta 198 (1986) 640 intensity in the halophile Dunaliella salina (Chlorophyta), J. Mar.
646. Biol. Asso. UK 62 (1982) 493508.
[28] J. Neidhardt, J.R. Benemann, L. Zhang, A. Melis, Photosystem-II [30] D. Marquardt, An algorithm for least squares estimation of nonlinear
repair and chloroplast recovery from irradiance stress: relationship parameters, SIAM J. Appl. Math. 11 (1963) 431441.
between chronic photoinhibition, light-harvesting chlorophyll [31] N.S. Xu, Random Signal Estimation and System Control, Beijing
antenna size and photosynthetic productivity in Dunaliella salina Polytechnic University Press, Beijing, 2000.
(green algae), Photosynth. Res. 56 (1998) 175184.

You might also like