12
Simple method for measuring the spectral absorption cross-section of microalgae Razmig Kandilian a,b , Antoine Soulies b , Jeremy Pruvost b , Benoit Rousseau c , Jack Legrand b , Laurent Pilon a,n a Mechanical and Aerospace Engineering Department, Henry Samueli School of Engineering and Applied Science, University of California, 420 Westwood Plaza, Eng. IV 37-132, Los Angeles, CA 90095, USA b Université de Nantes, GEPEA CNRS, UMR 6144, Bd de l'Université, CRTT - BP 406, 44602 Saint-Nazaire Cedex, France c Laboratoire de Thermocinétique de Nantes, Université de Nantes, UMR CNRS 6607, Nantes, France HIGHLIGHTS A simple procedure for measuring microalgae spectral absorption cross-section is proposed. It is based on normalhemispherical transmittance and reectance measurements. It was validated against direct measurements of the radiation characteristics of C. vulgaris. It can be used for other suspensions of absorbing and scattering particles. It can be used to predict and control light transfer and biomass productivity in PBRs. article info Article history: Received 15 October 2015 Received in revised form 16 February 2016 Accepted 23 February 2016 Available online 3 March 2016 Keywords: Light transfer Photobioreactors Light absorption Microalgae Growth modeling abstract The spectral absorption cross-section of microalgae is an essential parameter in modeling the microalgae metabolism and growth kinetics as well as in estimating the productivity and efciency of photo- bioreactors. This paper presents a simple experimental procedure for retrieving the average spectral absorption cross-section of concentrated microalgal suspensions. The method combines experimental measurements of the normalhemispherical transmittance and reectance of the suspensions in con- ventional cuvettes with an inverse method and analytical expressions obtained from the modied two- ux approximation accounting for absorption and multiple scattering. The method was validated with direct measurements of the scattering phase function and of the absorption and scattering cross-sections of freshwater microalgae Chlorella vulgaris. It was able to retrieve the absorption cross-section with acceptable accuracy. The latter was then used to estimate successfully the uence rate using a simplied light transfer model, the local rate of photon absorption, and the biomass productivity in at plate PBRs. & 2016 Elsevier Ltd. All rights reserved. 1. Introduction Cultivation of photosynthetic microorganisms has received sig- nicant interest in recent years as a way to produce a broad range of chemical compounds including food supplements, fertilizers, hydro- gen gas, and lipids for biodiesel production (Richmond, 2004). A wide variety of microalgae and cyanobacteria species can be cultivated in fresh or brackish waters as well as in seawater. Cultivation typically takes place in outdoor raceway ponds or photobioreactors. Light transfer in such systems is arguably one of the most important parameters affecting both the growth rate and metabolism of the photosynthetic microorganisms (Cornet et al., 1998; Pilon et al., 2011). Illumination conditions within the culture can vary strongly with time depending on the season, on the local weather, and on the photo- adaptation of the microorganisms which reduces or increases their pigment concentrations under excessive illumination or light limiting conditions, respectively. Physical modeling coupling light transfer to growth kinetics and metabolic activities can facilitate the optimum design, operation, and control of photobioreactors (Cornet et al., 1992, 1992, 1995; Cornet and Dussap, 2009; Murphy and Berberoğlu, 2011; Pilon et al., 2011; Takache et al., 2010, 2012; Pruvost et al., 2012; Wheaton and Krishnamoorthy, 2012; Kandilian et al., 2014; Kong and Vigil, 2014). To do so, knowing the radiation characteristics of microalgae and the amount of light they absorb to drive the photo- synthesis is essential. In fact, the radiation characteristics of micro- organisms vary signicantly among species (Berberoğlu and Pilon, Contents lists available at ScienceDirect journal homepage: www.elsevier.com/locate/ces Chemical Engineering Science http://dx.doi.org/10.1016/j.ces.2016.02.039 0009-2509/& 2016 Elsevier Ltd. All rights reserved. n Corresponding author. Tel.: þ1 310 206 5598; fax: þ1 310 206 2302. E-mail address: [email protected] (L. Pilon). Chemical Engineering Science 146 (2016) 357368

Simple method for measuring the spectral absorption cross ...pilon/Publications/CES2016...Simple method for measuring the spectral absorption cross-section of microalgae Razmig Kandiliana,b,

  • Upload
    others

  • View
    3

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Simple method for measuring the spectral absorption cross ...pilon/Publications/CES2016...Simple method for measuring the spectral absorption cross-section of microalgae Razmig Kandiliana,b,

Chemical Engineering Science 146 (2016) 357–368

Contents lists available at ScienceDirect

Chemical Engineering Science

http://d0009-25

n CorrE-m

journal homepage: www.elsevier.com/locate/ces

Simple method for measuring the spectral absorption cross-sectionof microalgae

Razmig Kandilian a,b, Antoine Soulies b, Jeremy Pruvost b, Benoit Rousseau c,Jack Legrand b, Laurent Pilon a,n

a Mechanical and Aerospace Engineering Department, Henry Samueli School of Engineering and Applied Science, University of California, 420 WestwoodPlaza, Eng. IV 37-132, Los Angeles, CA 90095, USAb Université de Nantes, GEPEA CNRS, UMR 6144, Bd de l'Université, CRTT - BP 406, 44602 Saint-Nazaire Cedex, Francec Laboratoire de Thermocinétique de Nantes, Université de Nantes, UMR CNRS 6607, Nantes, France

H I G H L I G H T S

� A simple procedure for measuring microalgae spectral absorption cross-section is proposed.

� It is based on normal–hemispherical transmittance and reflectance measurements.� It was validated against direct measurements of the radiation characteristics of C. vulgaris.� It can be used for other suspensions of absorbing and scattering particles.� It can be used to predict and control light transfer and biomass productivity in PBRs.

a r t i c l e i n f o

Article history:Received 15 October 2015Received in revised form16 February 2016Accepted 23 February 2016Available online 3 March 2016

Keywords:Light transferPhotobioreactorsLight absorptionMicroalgaeGrowth modeling

x.doi.org/10.1016/j.ces.2016.02.03909/& 2016 Elsevier Ltd. All rights reserved.

esponding author. Tel.: þ1 310 206 5598; faxail address: [email protected] (L. Pilon).

a b s t r a c t

The spectral absorption cross-section of microalgae is an essential parameter in modeling the microalgaemetabolism and growth kinetics as well as in estimating the productivity and efficiency of photo-bioreactors. This paper presents a simple experimental procedure for retrieving the average spectralabsorption cross-section of concentrated microalgal suspensions. The method combines experimentalmeasurements of the normal–hemispherical transmittance and reflectance of the suspensions in con-ventional cuvettes with an inverse method and analytical expressions obtained from the modified two-flux approximation accounting for absorption and multiple scattering. The method was validated withdirect measurements of the scattering phase function and of the absorption and scattering cross-sectionsof freshwater microalgae Chlorella vulgaris. It was able to retrieve the absorption cross-section withacceptable accuracy. The latter was then used to estimate successfully the fluence rate using a simplifiedlight transfer model, the local rate of photon absorption, and the biomass productivity in flat plate PBRs.

& 2016 Elsevier Ltd. All rights reserved.

1. Introduction

Cultivation of photosynthetic microorganisms has received sig-nificant interest in recent years as a way to produce a broad range ofchemical compounds including food supplements, fertilizers, hydro-gen gas, and lipids for biodiesel production (Richmond, 2004). A widevariety of microalgae and cyanobacteria species can be cultivated infresh or brackish waters as well as in seawater. Cultivation typicallytakes place in outdoor raceway ponds or photobioreactors. Lighttransfer in such systems is arguably one of the most importantparameters affecting both the growth rate and metabolism of the

: þ1 310 206 2302.

photosynthetic microorganisms (Cornet et al., 1998; Pilon et al., 2011).Illumination conditions within the culture can vary strongly with timedepending on the season, on the local weather, and on the photo-adaptation of the microorganisms which reduces or increases theirpigment concentrations under excessive illumination or light limitingconditions, respectively. Physical modeling coupling light transfer togrowth kinetics and metabolic activities can facilitate the optimumdesign, operation, and control of photobioreactors (Cornet et al., 1992,1992, 1995; Cornet and Dussap, 2009; Murphy and Berberoğlu, 2011;Pilon et al., 2011; Takache et al., 2010, 2012; Pruvost et al., 2012;Wheaton and Krishnamoorthy, 2012; Kandilian et al., 2014; Kong andVigil, 2014). To do so, knowing the radiation characteristics ofmicroalgae and the amount of light they absorb to drive the photo-synthesis is essential. In fact, the radiation characteristics of micro-organisms vary significantly among species (Berberoğlu and Pilon,

Page 2: Simple method for measuring the spectral absorption cross ...pilon/Publications/CES2016...Simple method for measuring the spectral absorption cross-section of microalgae Razmig Kandiliana,b,

R. Kandilian et al. / Chemical Engineering Science 146 (2016) 357–368358

2007; Berberoğlu et al., 2008; Heng et al., 2014) or during the culti-vation process. This is particularly true during nitrogen starvationused to trigger lipid accumulation (Kandilian et al., 2013; Heng et al.,2014; Heng and Pilon, 2014; Kandilian et al., 2014).

The goal of the present study is to develop a simple yet accu-rate experimental method for determining the spectral absorptioncross-section of microalgae over the photosynthetically activeradiation (PAR) region, corresponding to wavelength between 400and 700 nm. This method does not require the knowledge of thescattering phase function of the microorganisms or the use of anephelometer, which can be costly and difficult to operate.Moreover, it only requires an integrating sphere spectro-photometer. It also aims to demonstrate that the use of theabsorption cross-section alone with a simplified light transfermodel is sufficient to predict light transfer, growth kinetics, andbiomass productivity of PBRs. The unicellular microalgae Chlorellavulgaris was used to illustrate the new method because of theinterest in the wide range of compounds it can produce.

2. Background

2.1. Chlorella vulgaris

Chlorella vulgaris is a unicellular microalgae with sphericalshape and 1–6 μm in diameter (Safi et al., 2014). This particularspecies of green microalgae has garnered special attention for itslarge areal biomass productivity and its relatively short cell dou-bling time (Griffiths and Harrison, 2009). In addition, it has beenidentified as a potential source of protein for human and animalfeed (Tokuşoglu and Ünal, 2003). It is also an important source forthe commercial production of carotenoids and various bio-pharmaceuticals (Mendes et al., 2003). It can also produce lipids tobe converted into biodiesel (Brennan and Owende, 2003). Finally,this freshwater species can be cultivated in wastewater forsimultaneous wastewater treatment and production of value-added products (Feng et al., 2011).

2.2. One-dimensional radiative transfer equation

Fig. 1 shows a schematic of a microalgae suspension of thick-ness ℓ illuminated with collimated, unpolarized, and normallyincident light along with the associated coordinate system. Lighttransfer through such absorbing, anisotropically scattering, andnon-emitting suspension can be treated as one-dimensional (1D)along the z-direction taken as the direction of the incident light. Itis governed by the 1D radiation transfer equation (RTE) expressedin terms of the local spectral light intensity Iλðμ; zÞ (in W/m2 sr μm) at location z, in direction θ, and at wavelength λ as(Modest, 2013),

μ∂Iλ∂z

¼ �ðκλþσs;λÞIλþσs;λ

2

Z 1

�1pλðμ0;μÞIλðμ0; zÞdμ0 ð1Þ

where μ¼ cos θ is the direction cosine, κλ and σs;λ are respec-tively the absorption and scattering coefficients while pλðμ0;μÞ isthe azimuthally symmetric scattering phase function. The latterrepresents the probability that photons coming from direction μ0

¼ cos θ0 be scattered in direction μ¼ cos θ. It is normalized suchthat

12

Z 1

�1pλðμ0;μÞ dμ0 ¼ 1: ð2Þ

For the azimuthally symmetric problem, pλðμ0;μÞ depends only onμ0 ¼ cos θ0 where θ0 is the angle between the radiation incidenton the scatterer along direction μ0 ¼ cos θ0 and the intensityscattered in direction μ¼ cos θ, i.e., pλðμ0;μÞ ¼ pλðμ0Þ. From

geometrical consideration, μ0 can be expressed as μ0 ¼ μμ0 þð1�μ2Þ1=2ð1�μ02Þ1=2 (Modest, 2013).

In cases when scatterers are much larger than the radiationwavelength, scattering is mainly forward and the scattering phasefunction can be estimated based on the transport approximationas pλðμ0Þ ¼ 1�gλþ2gλδð1�μ0Þ (McKellar and Box, 1981). Here,δðμÞ is the Dirac delta function and gλ is the asymmetry factor atwavelength λ defined as

gλ ¼12

Z 1

�1pλðμ0Þμ0 dμ0: ð3Þ

Then, the RTE can be expressed as

μ∂Iλ∂z

¼ �ðκλþσs;tr;λÞIλþσs;tr;λ

2

Z 1

�1Iλðμ0; zÞdμ0 ð4Þ

where σs;tr;λ ¼ ð1�gλÞσs;λ is the transport scattering coefficient.In the context of light transfer in photobioreactors, the

absorption and scattering coefficients are linearly proportional tothe number density NT of the microalgae cells (in #/m3) andexpressed as

κλ ¼ Cabs;λNT and σs;λ ¼ Csca;λNT ð5Þ

where Cabs;λ and Csca;λ (in m2) are the average spectral absorptionand scattering cross-sections of a cell, respectively. Alternatively,the biomass concentration of the microalgal suspension X,expressed in mass of dry weight per unit volume (g/L or kg/m3), isoften measured instead of NT. Then, one can express κλ and σs;λ as

κλ ¼ Aabs;λX and σs;λ ¼ Ssca;λX ð6Þ

where Aabs;λ and Ssca;λ are the average spectral mass absorptionand scattering cross-sections expressed in m2/kg dry weight.Typically, the biomass concentration X in conventional PBR varyfrom 0.1 to 2.0 kg/m3 (Takache et al., 2010).

2.3. Fluence rate

The local spectral fluence rate GλðzÞ in a PBR is defined as theirradiance incident at location z from all directions. It is expressedin W/m2 μm or μmolhν=m2 s μm and defined, in 1D radiationtransfer, as

GλðzÞ ¼ 2πZ 1

�1Iλðz;μÞ dμ: ð7Þ

Cornet and co-workers solved the above 1D RTE using theSchuster–Schwarzschild two-flux approximation to predict thelocal fluence rate GλðzÞ in flat plate PBR illuminated on one sidewith a black (Cornet et al., 1992) or diffusely reflecting (Pottier etal., 2005) wall on the other. Their solution accounted for multipleand anisotropic scattering. For example, the local fluence rate GλðzÞin a flat plate PBR with a transparent front window and a diffuselyreflecting back wall of reflectance ρλ exposed to collimated spec-tral irradiance q″in;λ normally incident onto the photobioreactor as(Pottier et al., 2005),

GλðzÞ ¼ 2q″in;λ½ρλð1þαλÞe�δλL�ð1�αλÞe�δλL�eδλzþ½ð1þαλÞeδλL�ρλð1�αλÞeδλL�e�δλz

ð1þαλÞ2eδλL�ð1�αλÞ2e�δλL�ρλð1�α2λÞeδλLþρλð1�α2

λÞe�δλL

ð8Þ

where αλ and δλ are expressed as (Pottier et al., 2005)

αλ ¼

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiAabs;λ

Aabs;λþ2bλSsca;λ

vuut and δλ ¼ XffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiAabs;λðAabs;λþ2bλSsca;λÞ

q:

ð9Þ

Page 3: Simple method for measuring the spectral absorption cross ...pilon/Publications/CES2016...Simple method for measuring the spectral absorption cross-section of microalgae Razmig Kandiliana,b,

Fig. 1. Schematic of one-dimensional light transfer in a refracting, absorbing, andscattering microalgae suspension illuminated with collimated and normallyincident light.

Table 1Kinetic model parameters used for predicting biomass productivity of C. vulgaris(Soulies et al., 2016).

Parameter Value Units

Mx 0.024 kg/molJNADH2 1:80� 10�3 molNADH2

=kg s

νO2 �X 1.13 molO2 =molXρM 0.8ϕ0O2 1:10� 10�7 molO2 =μmolhν

νNADH2 �O22 molNADH2 =molO2

K 4:0� 104 μmolhν=kg s

Kr 500 μmolhν=kg s

R. Kandilian et al. / Chemical Engineering Science 146 (2016) 357–368 359

Here, bλ is the backward scattering fraction defined as

bλ ¼12

Z π

π=2pλð0;θÞ sin θ dθ: ð10Þ

Microalgae suspensions in PBRs scatter visible light strongly inthe forward direction. Then, bλ tends to zero, αλ approaches unity,and δλ ¼ Aabs;λX ¼ κλ. Thus, the above expression for GλðzÞ can besimplified to (Lee et al., 2014)

GλðzÞ ¼ q″in;λe�Aabs;λXzþρλq

″in;λe

�Aabs;λXð2L� zÞ: ð11Þ

The first term represents the remainder of the incident irradianceafter it has traveled a distance z in the microalgae suspensionwhile the second term accounts for the incident irradiance thathas traveled once through the suspension, got reflected by theback wall, and continued traveling a distance ðL�zÞ in the sus-pension. Note that reflectance ρλ¼0 for PBRs with a transparentback wall. This simplified expression has been validated againstpredictions by Eq. (8) and by the numerical solutions of the 3D RTEusing the discontinuous Galerkin method combined with thediscrete-ordinate method for outdoor open ponds and flat platePBRs (Lee et al., 2014). Note that Eq. (11) differs from Beer-Lam-bert's law in that (i) it depends only on the absorption coefficientκλ ¼ Aabs;λX and not on the extinction coefficient βλ ¼ ðAabs;λþSsca;λÞX and (ii) it accounts for reflection within the PBR. More impor-tantly, the simplified expression for the fluence rate GλðzÞ given byEq. (11) depends only on the average spectral mass absorptioncross-section Aabs;λ of the suspension.

2.4. Growth kinetics and productivity

The time rate of change of the microorganism mass con-centration X(t) in continuous cultures satisfies the mass balanceequation (Pires, 2015)

dXdt

¼ ⟨μ⟩X�DX ð12Þ

where D (in h�1) is the dilution rate of the culture and ⟨μ⟩ (in h�1)corresponds to the volume-average specific growth rate of themicroorganisms. In addition, the volumetric biomass productivity⟨rX ⟩ is defined as ⟨rX ⟩¼ ⟨μ⟩X. For PBRs operated in continuousmode, steady-state conditions are achieved when dX

dt ¼ 0. Thiscondition is satisfied when ⟨μ⟩¼D (Pires, 2015).

Several models have been developed to predict the averagespecific growth rate ⟨μ⟩ as a function of the illumination conditionsin the PBR over the PAR region. The most general approach is todefine the average specific growth rate ⟨μ⟩ over the entire flat-plate PBR depth L from the local growth rate μðzÞ at location zaccording to (Fouchard et al., 2009)

⟨μ⟩¼ 1L

Z L

0μðzÞ dz: ð13Þ

The local growth rate μðzÞ is related to the local oxygen productionrate JO2

ðzÞ (in molO2=kgX � s) by Takache et al. (2012)

μðzÞ ¼ JO2ðzÞMx

νO2 �X: ð14Þ

Here, Mx (in kg/mol) is the molar mass of the biomass and νO2 �X

the stoichiometric coefficient of the oxygen production. Then, thelocal oxygen production rate JO2

ðzÞ can be expressed as a functionof the local rate of photon absorption (LRPA) AðzÞ according to(Takache et al., 2012)

JO2ðzÞ ¼ ρM

KKþAðzÞϕ

0O2AðzÞ� JNADH2

νNADH2 �O2

Kr

KrþAðzÞ ð15Þ

where ρM is the maximum energy yield for photon conversion, ϕ0O2

is the quantum yield of O2 (in molO2=μmolhν) for the Z-scheme of

photosynthesis, and K and Kr (in μmolhν=kgX � s) are, respectively,the half-saturation constant for photosynthesis and the saturationconstant describing the inhibition of respiration in light. Note thatthe first term on the right hand side of Eq. (15) represents oxygenproduction due to photosynthesis while the second term repre-sents oxygen consumption due to respiration. Similarly, JNADH2

andνNADH2 �O2

are the specific rate and stoichiometric coefficient ofcofactor regeneration on the respiratory chain. The kinetic modelpresented by Eqs. (14) and (15) was developed and experimentallyvalidated for the green microalgae C. reinhardtii by Takache et al.(2012). Recently, Pruvost and co-workers (Pruvost et al., 2016;Soulies et al., 2016) adjusted the model parameter values to adaptthe model to C. vulgaris by fitting the model to experimentallymeasured biomass productivity at various culture dilution ratesand incident irradiances. Detailed description of the kinetic modelparameters as well as how they were estimated can be found inCornet and Dussap (2009); Takache et al. (2012); Soulies et al.(2016). Table 1 summarizes the kinetic model parameters andtheir corresponding values for C. vulgaris (Soulies et al., 2016).

Finally, the LRPA AðzÞ can be expressed as (Cornet et al., 1992;Cornet and Dussap, 2009)

AðzÞ ¼Z 700

400Aabs;λGλðzÞ dλ: ð16Þ

Similarly, the mean rate of photon absorption (MRPA) ⟨A⟩ can beobtained by integrating the LRPA over the volume of the flat-platePBR as

⟨A⟩¼ 1L

Z L

0AðzÞ dz: ð17Þ

Here, it is interesting to note that, for strongly forward scatteringmedia, the MRPA ⟨A⟩ depends only on Gλ, Aabs;λ, and L where GλðzÞ is

Page 4: Simple method for measuring the spectral absorption cross ...pilon/Publications/CES2016...Simple method for measuring the spectral absorption cross-section of microalgae Razmig Kandiliana,b,

R. Kandilian et al. / Chemical Engineering Science 146 (2016) 357–368360

given by Eq. (11). This emphasizes the importance of determining thespectral mass absorption cross-section Aabs;λ over the PAR region.

2.5. Microalgae radiation characteristics

The radiation characteristics of microalgae can be directlymeasured experimentally (Berberoğlu and Pilon, 2007; Berberoğluet al., 2008; Pilon et al., 2011) or predicted theoretically (Pottieret al., 2005; Berberoğlu et al., 2007; Dauchet et al., 2015). Directexperimental measurements can be performed on microorganismsof any size and shape. However, they require expensive equipmentand can be time consuming, and need to be repeated for differentgrowth conditions, such as different pH, temperature, and/orphoton flux density change, as these parameters can alter theradiation characteristics of the microalgae (Berberoğlu et al., 2008;Pilon et al., 2011). Recently, Bellini et al. (2014) developed anexperimental method to determine the absorption Aabs;λ and thetransport scattering ð1�gλÞSsca;λ cross-sections of microalgae. Theauthors measured the total reflectance (diffuse plus specularreflectance) and the normal–hemispherical transmittance ofmicroalgae suspensions using a double integrating sphere spec-trophotometer and retrieved Aabs;λ and ð1�gÞSsca;λ using theinverse-adding-doubling (IAD) method (Prahl et al., 1993). Themicroalgae suspensions featured a concentration between1.1�1013 and 2.4�1013 cells/m3. This method was not able toestimate the absorption and the transport scattering cross-sectionaccurately especially for samples with low absorption coefficient.Indeed, the authors stated that “in samples with low absorptionand high scattering coefficients, the unwanted light losses may beinterpreted by the IAD algorithm as absorption phenomena insidethe sample rather than losses in the outside world, leading to anoverestimation of the absorption coefficient” (Bellini et al., 2014).They recommended for this method to be only used for qualitativeanalysis of microalgae cultures. Alternatively, the radiation char-acteristics of microalgae can be estimated theoretically usingelectromagnetic wave theory and solutions of Maxwell's equationssuch as the Lorenz–Mie theory. However, such theoretical pre-dictions rely on several simplifying assumptions such as treatingthe cells as homogeneous and spherical with some effectivespectral complex index of refraction predicted based on simplifiedmodels (Pottier et al., 2005; Berberoğlu et al., 2007; Dauchet et al.,2015). For example, fractal colonies and aggregates of microalgae(Kandilian et al., 2015) and filamentous cyanobacteria such asNostoc pondiforme and Anabaena iyengari (Lee and Pilon, 2013), orthe dumbbell shaped cyanobacteria Synechocystis sp. (Heng et al.,2015) cannot be treated as spherical. By contrast, a simpleexperimental method that is able to determine the absorptioncross-section of microorganisms with any arbitrary shapes andoptical properties could be valuable.

This paper presents a simple method for measuring the spectralabsorption cross-section Cabs;λ or Aabs;λ from simple and rapid mea-surements of normal–hemispherical transmittance and reflectance ofrelatively concentrated suspensions in conventional cuvettes. In fact,we have argued that the absorption cross-section is sufficient topredict the local fluence rate GλðzÞ, the MRPA ⟨Α⟩, and the averagegrowth rate ⟨μ⟩.

3. Materials and methods

3.1. Microalgae cultivation and sample preparation

The microalgae C. vulgaris CCAP 211/19 were obtained from theculture collection of algae and protozoa (CCAP, Scotland, UK).The microorganisms were cultivated in a 4 cm thick, 1.4 L torusPBR constructed with transparent PMMA described in detail in

Pottier et al. (2005). The PBR was illuminated with 200 μmolhν=m2

s white LEDs incident on one side. It was operated in continuousmode by injection of fresh medium at the dilution rate D of0.06 h�1 using a peristaltic pump. The microalgae were grown inmodified Sueoka mediumwith the following composition (in gramsper liter of MilliQ water): NH4Cl 1.45, MgSO4–7H2O 0.28, CaCl2 0.05,KH2PO4 0.61, NaHCO3 1.68, ZnSO4–7H2O 0.022, H3BO3 0.0114,MnCl2–4H2O 0.00506, CoCl2–6H2O 0.00161, CuSO4–5H2O 0.00157,FeSO4–7H2O 0.00499, (NH4)6Mo7O24–4H2O 0.0011, KOH 0.05,Na2EDTA 0.02. The culture temperature was maintained at 22 °Cthanks to a liquid cooled heat exchanger attached to the stainlesssteel PBR back wall. The culture medium pH was continuouslymonitored using a pH sensor (Mettler Toledo SG 3253) and wasmaintained at 7.5 by automatically injecting CO2 gas when itexceeded 7.5. Agitation of the culture was achieved by mechanicallymixing the culture using a marine propeller attached to an electricmotor. The culture under steady-state continuous operation had drybiomass concentration X of 0.33 g/L. To obtain samples with largerbiomass concentrations, a small volume of culture was harvestedand centrifuged at 10,000 g (ThermoScientific Sorvall RC 6 Plus,Massachusetts, USA) for 5 min at 4 °C and suspended in 3 ml ofphosphate buffer saline (PBS) solution. The volume of culturesampled was chosen based on the biomass concentration desiredfor optical measurements and ranged from 0.035 to 111 g/L. Notethat the biomass concentration of only the most concentratedsample was measured, i.e., 11175 g/L and the biomass concentra-tions of the remaining samples, namely, 55.672.5, 11.170.5,5.5670.25, 1.1170.05, 0.55670.025, and 0.11170.0125 g/L, wereestimated based on the dilution performed.

3.2. Culture characterization

The cell size distribution of C. vulgaris was measured using 165microscope images captured using a digital camera (AxioCam MRc)mounted on an optical microscope (Zeiss Axio Scope A1). The dia-meters of 2873 cells were measured manually using an image pro-cessing software (Axio Vision Routine) and their volume equivalentsphere radius was reported for the sake of completeness. Theseresults were not used in either experimental approaches investigated.

The culture dry biomass concentration X (in g/L) was determinedgravimetrically by filtering 5 mL of culture through a pre-dried andpre-weighed 0.45 μm pore size glass-fiber filter (Whatman GF/F). Thefilters were dried overnight in an oven at 105 °C and weighed, in ananalytical balance (Mettler Toledo XA204, Columbus, OH) with asensitivity of 0.1 mg, after being cooled in a desiccator for 10 min. Thesamples were analyzed in triplicates and the reported biomass con-centration corresponded to the mean value. In addition, the celldensity NT (in #/m3) was counted using a 200 μm deep Malassezcounting chamber.

The volume fraction of water in the cells xw was estimatedusing the relation derived by Pottier et al. (2005) and given by

xw ¼ 1� XNT

1V32ρdm

ð18Þ

where ρdm corresponds to the density of dry biomass and wastaken as 1350 kg/m3 (Dauchet, 2012) while V32 is the cell volumecorresponding to the Sauter mean diameter of the microorgan-isms. The latter is defined as the diameter of a sphere that has thesame volume to surface area ratio as the original cell.

The concentration of photosynthetic pigments in C. vulgarisgrown under the above mentioned conditions were estimatedspectrometrically in terms of mass of pigments per unit volume ofsuspension (mg/m3). A volume of 0.5 mL of culture was first cen-trifuged at 13,400 rpm (12,100 g) for 10 min. The medium was dis-carded and the cells were resuspended in 1.25 mL pure methanoland sonicated for 10 s. Pigments were extracted for 1 h at 45oC and

Page 5: Simple method for measuring the spectral absorption cross ...pilon/Publications/CES2016...Simple method for measuring the spectral absorption cross-section of microalgae Razmig Kandiliana,b,

R. Kandilian et al. / Chemical Engineering Science 146 (2016) 357–368 361

the extract was centrifuged. The optical density ODλ of the super-natant was measured at wavelengths 750, 665, 652, and 480 nmusing a UV-VIS-NIR spectrophotometer (Agilent Cary 5000, SantaClara, CA). All extractions were performed in triplicates. Chlorophyll aand b concentrations, respectively denoted by CChla and CChlb (in mg/L), were estimated according to the correlations (Ritchie, 2006)

CChla ¼ �8:0962ðOD652�OD750Þþ16:5169ðOD665½�OD750Þ�V2ℓ

�1V �11 ð19Þ

and CChlb ¼ 27:4405ðOD652�OD750Þ½�12:1688ðOD665�OD750Þ�V2ℓ

�1V �11 ð20Þ

where V1 and V2 are the volumes of the culture and of the solvent(methanol), respectively while ℓ is the thickness of the cuvette usedto measure the optical density. Similarly, the total photoprotective(PPC) and photosynthetic (PSC) carotenoid concentration CPPCþPSC (inmg/L) was estimated according to (Strickland and Parsons, 1968)

CPPCþPSC ¼ 4ðOD480�OD750ÞV2ℓ�1V �1

1 : ð21ÞThe corresponding mass fraction of pigment “i” per dry weight ofbiomass can be estimated as wi ¼ Ci=X.

3.3. Direct radiation characteristics measurements

3.3.1. ExperimentsFirst, the total scattering phase function pðμ0Þ of the microalgae

suspension was measured at 632.8 nm using a polar nephelometerequipped with He–Ne laser. The experimental setup and dataanalysis were previously reported in detail by Berberoğlu et al.(2008) and need not be repeated. Due to probe interference withthe incident laser beam, it was only possible to collect measure-ments for scattering angles θ0 up to 160°. Then, the asymmetryfactor g633 and the back-scattered fraction b633 at 632.8 nm wereestimated according to Eqs. (3) and (10), respectively.

Second, the normal–normal transmittance of dilute suspen-sions of C. vulgaris in a cuvette, denoted by Tnn;λ, was measuredusing a UV-VIS-NIR spectrophotometer equipped with a detectorwith a 8° half acceptance angleΘa (Agilent Cary 5000, Santa Clara,CA). The normal–hemispherical transmittances Tnh;λ of the samesuspensions were measured using an integrating sphere attach-ment (Agilent Cary DRA-2500, Santa Clara, CA) to the spectro-photometer. Due to masking by the cuvette holder, the cuvettearea exposed to the normally incident beam was 20� 8 mm andthe cuvette holder was positioned directly in front of the 19�17 mm opening of the integrating sphere. The size of the incidentbeam was 9� 5 mm. These measurements were performed with1 cm pathlength quartz cuvettes (11010-40 Hellma Analytics,Mülheim, Germany) in the wavelength range from 350 to 750 nmwith 1 nm spectral resolution. To avoid absorption and scatteringby the growth medium, the microalgae were centrifuged at13,400 rpm for 10 min and washed twice with PBS solution andsuspended in PBS. Two biomass concentrations were considered,namely 0.035 and 0.049 dry mass g/L. Such dilute suspensionsensure that single scattering prevails so that measurements ofTnn;λ and Tnh;λ can be used in the direct measurements of theradiation cross-sections of C. vulgaris. Note that the samples weremanually shaken prior to each of the transmission measurementsso as to prevent sedimentation. The sample normal–normal Tnn;λand normal–hemispherical Tnh;λ transmittances were measuredthree times and the results were averaged for each concentration.

3.3.2. Data analysisThe extinction coefficient of the microalgae suspension βλ was

obtained from normal–normal transmittance measurements. In orderto correct for the effects of reflection and refraction by the cuvette, themeasurements were calibrated using the transmittance of the

reference medium (e.g., PBS) denoted by Tnn;λ;ref . Then, the apparentextinction coefficient βn

λ was defined as (Davies-Colley et al., 1986)

βn

λ ¼ �1ℓln

Tnn;λ

Tnn;λ;ref

!: ð22Þ

In addition, microalgae are relatively large compared with thewavelength of light and thus scatter strongly in the forward direction.A fraction ϵn of the light that is scattered in the forward directionreaches the detector due to its finite acceptance angle. This fractioncan be expressed as a function of the scattering phase function pλðθÞpreviously measured (Agrawal and Mengüç, 1991; Privoznik et al.,1978)

ϵn ¼12

Z Θa

0pλðθÞ sin θ dθ ð23Þ

where Θa is the half acceptance angle of the detector. Then, theapparent extinction coefficient can be related to the actual absorptionand scattering coefficients by (Davies-Colley et al., 1986),

βn

λ ¼ κλþσs;λ�ϵnσs;λ so that βλ ¼βn

λ�ϵnκλ1�ϵn

: ð24Þ

The absorption coefficient κλ was obtained from normal–hemispherical transmittance Tnh;λ measurements. Then, theapparent absorption coefficient κn

λ can be expressed as

κn

λ ¼ �1ℓln

Tnh;λ

Tnh;λ;ref

!: ð25Þ

Due to imperfect reflections at the inner surface of the integratingsphere and to the geometry of the experimental setup, the detectorwas unable to capture all the scattered light. To account for thesephenomena, the apparent absorption coefficient κn

λ can be related tothe absorption and scattering coefficients by (Pilon et al., 2011)

κn

λ ¼ κλþð1�ϵhÞσs;λ ð26Þwhere ϵh represents the fraction of scattered light not collected bythe integrating sphere or not measured by the detector. It accountsfor backward scattered light that does not enter the integratingsphere and for forward scattered light that does not reach thedetector because of imperfect reflection, for example. This correctionfactor is assumed to be constant over the PAR region. Davies-Colleyet al. (1986) assumed that the microalgae did not absorb radiation atwavelength λo. For example, λo¼750 nm for Chlamydomonas rein-hardtii (Berberoğlu et al., 2008) and Chlorella sp. (Berberoğlu et al.,2009). At this wavelength, the absorption coefficient vanishes, i.e.,κλ ¼ 0 m�1, and ϵh can be estimated from Eq. (26) as (Pilon et al.,2011)

ϵh ¼ 1�κn

λoσs;λo

ð27Þ

Combining Eqs. (24)–(27) yields the expression for the actualabsorption coefficient κλ given, in terms of measurable quantities, by(Pilon et al., 2011)

κλ ¼ κn

λ�κn

λo

βn

λ�κn

λ

βn

λo �κn

λo

!ð28Þ

Finally, the actual extinction coefficient βλ can be obtained by sub-stituting κλ into Eq. (24). Then, the scattering coefficient can becomputed using the definition σs;λ ¼ βλ�κλ.

The absorption and scattering coefficients κλ and σs;λ weredivided by the sample dry biomass concentration X to obtain theaverage spectral mass absorption and scattering cross-sectionsAabs;λ and Ssca;λ according to Eq. (6). The spectral cross-sectionsAabs;λ and Ssca;λ for different cell densities/concentrations shouldcollapse onto a single line if single and independent scattering

Page 6: Simple method for measuring the spectral absorption cross ...pilon/Publications/CES2016...Simple method for measuring the spectral absorption cross-section of microalgae Razmig Kandiliana,b,

Fig. 2. Block diagram of the procedure used to retrieve the average mass absorption and transport scattering cross-sections Aabs;λ and ð1�gλÞSsca;λ of concentratedsuspensions at a given wavelength λ from spectral normal–hemispherical transmittance and reflectance measurements. We used P¼120 individuals per generation for amaximum of 50 generations.

R. Kandilian et al. / Chemical Engineering Science 146 (2016) 357–368362

prevailed (Van De Hulst, 1957). This provided further validation ofthe experimental procedure and data analysis.

Finally, the direct measurement method with the previouslydescribed apparatus and data analysis were successfully validatedby comparing the measurements with predictions by Lorenz–Mietheory for monodisperse spherical latex particles 1.9870.01 or4.51870.158 μm in diameter (Polysciences Inc., Warrington, PA).

Fig. 3. Experimentally measured size distribution of C. vulgaris with mean radiusr ¼ 1:98 μm.

4. Inverse method

4.1. Experiments

The normal–hemispherical transmittance Tnh;λ and reflectanceRnh;λ of C. vulgaris suspensions, with mass concentration rangingfrom 0.111 g/L to 111 g/L, were systematically measured using theabove described spectrophotometer/integrating sphere assembly.Measurements for the different concentrations were performedwithin less than 1 h in order to minimize changes in cell biomass,composition, and pigment concentrations and thus their radiationcharacteristics during the successive measurements. For the rela-tively large concentrations considered, multiple scattering pre-vailed and the direct measurement method, previously describedin Section 3.3 and valid for single scattering, could not be used.Instead, experimentally measured values of Tnh;λ and Rnh;λ wereused as input parameters in the inverse method to retrieve theaverage spectral mass absorption Aabs;λ and scattering Ssca;λ cross-sections of C. vulgaris.

4.2. Forward problem formulation

Let us consider one-dimensional radiation transfer in microalgalsuspension exposed to normally incident, collimated, monochro-matic, and randomly polarized light as illustrated in Fig. 1. The sus-pension was assumed to be well-mixed and the microorganismswere randomly oriented and azimuthally symmetric. Under theseconditions and due to strong forward scattering by microalgae sus-pensions and the relatively thin cuvettes, light transfer can be

assumed to be one-dimensional (Pilon et al., 2011; Pottier et al., 2005;Lee et al., 2014; Dauchet et al., 2015; Stramski and Piskozub, 2003).We further assumed that the scattering phase function was given bythe transport approximation. Then, the spectral normal–hemi-spherical transmittance and reflectance of the suspension, afteraccounting for internal reflection, can be expressed as (Dombrovskyet al., 2005)

Tnh;λ;pred ¼ T0nh;λþ

2½ð1þρ1;λÞ exp ð�τtr;λ;ℓÞþAλ=ζλ� ð29Þ

and Rnh;λ;pred ¼ R0nh;λþ

2ð1þBλ=ζλþCtr;λÞ ð30Þ

Page 7: Simple method for measuring the spectral absorption cross ...pilon/Publications/CES2016...Simple method for measuring the spectral absorption cross-section of microalgae Razmig Kandiliana,b,

Fig. 4. Experimentally measured phase function of C. vulgaris at 632.8 nm alongwith the associated Henyey–Greenstein phase function for asymmetry factorg633¼0.974.

λλλ

Fig. 5. Directly measured average spectral mass absorption Aabs;λ and scatteringSsca;λ cross-sections of C. vulgaris suspensions between 350 and 750 nm for biomassconcentrations X of 0.035 and 0.049 g/L.

R. Kandilian et al. / Chemical Engineering Science 146 (2016) 357–368 363

where τtr;λ;ℓ is the transport optical thickness defined as τtr;λ;ℓ ¼βtr;λℓ¼ ½Aabs;λþð1�gλÞSsca;λ�Xℓ. Here, R0

nh;λ and T0nh;λ are the spectral

normal–hemispherical reflectance and transmittance of the suspen-sion ignoring multiple scattering expressed as (Dombrovsky et al.,2005; Modest, 2013)

R0nh;λ ¼

ρ1;λþð1�ρ1;λÞ2Ctr;λ

1�ρ1;λCtr;λand T0

nh;λ ¼ð1�ρ1;λÞ21�ρ1;λCtr;λ

exp ð�τtr;λ;ℓÞ

ð31Þwhere ρ1;λ is the normal–normal reflectivity of the quartz/air inter-face given by Fresnel's equations. For an optically smooth surfacewith normally incident radiation and assuming that the absorptionindex kλ of the quartz was negligible compared with its refractiveindex nλ so that

ρ1;λ ¼ðnλ�1Þ2ðnλþ1Þ2

ð32Þ

The parameters Aλ, Bλ, Ctr;λ, and Dλ are defined as (Dombrovsky et al.,2005)

Aλ ¼ðγ1;λ�γ2;λρ1;λÞðφλsλþcλÞ exp ð�τtr;λ;LÞ�ðγ2;λ�γ1;λCtr;λÞ

ð1þφ2λÞsλþ2φλcλ

ð33Þ

Bλ ¼ðγ1;λ�γ2;λρ1;λÞ exp ð�τtr;LÞ�ðγ2;λ�γ1;λCtr;λÞðφλsλþcλÞ

ð1þφ2λÞsλþ2φλcλ

ð34Þ

Ctr;λ ¼ ρ1;λexpð�2τtr;λ;ℓÞ ð35Þ

Dλ ¼γλð1�μ2

c;λÞχλζ2λ

ζ2λ�1ð36Þ

Here, the parameters γλ, μc;λ, γλ, γ1;λ, γ2;λ, φλ, ζλ, sλ, and cλ are givenby Dombrovsky et al. (2005)

γλ ¼1�ρ1;λ

1þρ1;λ; μc;λ ¼

ðn2λ�1Þ1=2nλ

;

γλ ¼γλ

1þμc;λ; γ1;λ ¼ 1�2γλ; γ2;λ ¼ 1þ2γλ;

χλ ¼ωtr;λ

1�ωtr;λ

1�ρ1;λ

1�ρ1;λCtr;λ; φλ ¼ 2γλ=ζλ; ζ

2λ ¼

4ð1þμc;λÞ2

1�ωtr;λ

1�ωtr;λμc;λ;

sλ ¼ sinh ðζλτtr;λ;ℓÞ; and cλ ¼ cosh ðζλτtr;λ;ℓÞ ð37Þ

where μc;λ is the director cosine of the critical angle θc beyond whichtotal internal reflection occurs while the transport single scatteringalbedo is defined as ωtr;λ ¼ σtr;s;λ=βtr;λ. In order to predict the nor-mal–hemispherical transmittance and reflectance Tnh;λ and Rnh;λ

using Eqs. (29)–(37), one needs to know (i) the thickness ℓ of thesuspension, (ii) the index of refraction of the quartz cuvette nλ, (iii)the absorption coefficient of the suspension κλ, and (iv) the transportscattering coefficient σs;tr;λ ¼ σs;λð1�gλÞ. Here, the thickness ℓ of thecuvette was 10 mm. The spectral refraction index nλ of quartz wasgiven by the dispersion relation (Heraeus, 2015)

n2λ−1¼ B1λ

2

λ2−C1þ B2λ

2

λ2−C2þ B3λ

2

λ2−C3ð38Þ

Page 8: Simple method for measuring the spectral absorption cross ...pilon/Publications/CES2016...Simple method for measuring the spectral absorption cross-section of microalgae Razmig Kandiliana,b,

a

b

0

5

10

15

20

25

30

35

350 450 550 650 750

Nor

mal

-hem

isphe

rica

l ref

lect

ance

, Rnh

,(%

)

Wavelength, (nm)

X = 0.1113 g/LX = 0.5565 g/LX =1.113 g/LX = 5.565 g/LX = 11.13 g/LX = 55.65 g/LX = 111.3 g/L

X=0.111 g/LX=0.556 g/LX=1.11 g/LX=5.56 g/LX=11.1 g/LX=55.6g/LX=111 g/L

R. Kandilian et al. / Chemical Engineering Science 146 (2016) 357–368364

where the wavelength λ is expressed in μm and ranges from 0.2 to2.5 μm. The constants B1, B2, B3, C1, C2, and C3 are respectivelyequal to 4.73115591�10�1, 6.31038719�10�1, 9.06404498�10�1,1.29957170�10�2 µm2, 4.12809220 �10�3 µm2, and 9.87685322�101 µm2 (Heraeus, 2015). On the other hand, in the visible part of thespectrum, quartz can be treated as non-absorbing, i.e., kλ � 0 (Her-aeus, 2015). This corresponded to μc around 0.73 or to a critical angleθc, for total internal reflection, of about 43° while the reflectivity ρ1;λwas about 3.5%.

4.3. Inverse method optimization

Fig. 2 shows a block diagram of the procedure used to simul-taneously retrieve the spectral mass absorption Aabs;λ and trans-port scattering ð1�gλÞSsca;λ cross-sections from the normal–hemispherical transmittance Tnh;λ and reflectance Rnh;λ measure-ments over the PAR region. The objective is to find the values ofAabs;λ and ð1�gλÞSsca;λ that minimize the difference between thepredicted and experimentally measured normal–hemisphericaltransmittance and reflectance of the microalgal suspension in theleast-square sense. Various optimization algorithms can be used tominimize, for each wavelength, the objective function δλ definedas,

δλ ¼Tnh;λ;pred�Tnh;λ

Tnh;λ

� �2

þ Rnh;λ;pred�Rnh;λ

Rnh;λ

� �2

: ð39Þ

Here, the inverse method to retrieve Aabs;λ and ð1�gλÞSsca;λ wasimplemented in Microsoft Excel using the built in non-linear optimi-zation solver based on the generalized reduced gradient (GRG) algo-rithm. A simple Excel file with instruction is available in digital formonline (Pilon, 2015) or directly from the corresponding author uponrequest. For the sake of comparison, the inverse method was alsoimplemented based on genetic algorithm using the general purposefunction optimization code PIKAIA (Charbonneau and Knapp, 1995;Charbonneau, 2002, 2002). Here, the spectral mass absorption andtransport scattering cross-sections Aabs;λ and ð1�gλÞSsca;λ wereretrieved and assumed to range from 0 to 1000 m2/kg based on pastexperience and on the literature (Pilon et al., 2011). The geneticalgorithm used a maximum of 50 generations with a populationconsisting of P¼120 individuals. The convergence criteria was set asδλo10�4. Note that both retrieval methods gave identical results.

0

10

20

30

40

50

60

70

80

90

100

350 450 550 650 750

Nor

mal

-hem

isphe

rica

l tra

nsm

ittan

ce, T

nh,(%

)

Wavelength, (nm)

Fig. 6. Experimentally measured normal–hemispherical (a) reflectance Rnh;λ and(b) transmittance Tnh;λ between 350 and 750 nm for C. vulgaris suspensions inquartz cuvettes with a path length of 1 cm with different values of biomass con-centrations X ranging from 0.111 to 111 g/L.

5. Results and discussion

5.1. Chlorella vulgaris characterization

Fig. 3 shows the measured size distribution of C. vulgarisobtained with 2873 cells. The cells featured a mean radius rs of1.98 μmwith a standard deviation of 0.41 μm. Moreover, the meanratio of minimum and maximum axis of each cell fitted to anellipse was 0.81. The cell number density NT was related to the drybiomass concentration X by NT ¼ 1:17� 1014X. Thus, the massfraction of water in the microorganisms was estimated as xw¼82wt.% based on Eq. (18). The pigment concentrations weremeasured spectrophotometrically according to the proceduredescribed earlier. The concentrations of Chl a, Chl b, and totalcarotenoids were such that CChla ¼ 25:9570:82 mg=L;CChlb ¼ 4:8170:36 mg=L;CPPCþPSC ¼ 6:14 70:29 mg=L. These results corre-sponded to the weight percentage wChla ¼ 4:4170:14 wt:%;wChlb

¼ 0:8270:05 wt:%, and wPPCþPSC ¼ 1:0470:06 wt:%.

5.2. Direct measurements

Fig. 4 shows the azimuthally symmetric scattering phasefunction of C. vulgaris measured experimentally at 632.8 nm. It is

evident that C. vulgaris scattered visible light strongly in the for-ward direction, as expected for such large scatterers with meansize parameter χ ¼ 2πrs=λ ranging from 16 to 36 over the PAR. Infact, the asymmetry factor g633 and back-scattered fraction b633were 0.974 and 0.0042, respectively. Fig. 4 also plots the corre-sponding Henyey–Greenstein approximate phase function. Inaddition, the fraction ϵn of light scattered in the forward directionreaching the detector was estimated to be ϵn ¼ 0:71. This valuewas used in the direct measurements of the extinction cross-section [Eq. (24)].

Fig. 5a and b respectively plot the average mass absorptionAabs;λ and scattering Ssca;λ cross-sections (in m2/kg) as functions ofwavelength between 350 and 750 nm measured directly for smallbiomass concentrations of 0.035 and 0.049 g/L. First, the cross-sections measured for both concentrations collapse on the samecurve confirming that single scattering conditions prevailed

Page 9: Simple method for measuring the spectral absorption cross ...pilon/Publications/CES2016...Simple method for measuring the spectral absorption cross-section of microalgae Razmig Kandiliana,b,

R. Kandilian et al. / Chemical Engineering Science 146 (2016) 357–368 365

(Van De Hulst, 1957). Fig. 5 also indicates that the scattering cross-section Ssca;λ was much larger than the absorption cross-sectionAabs;λ. Note, however, that the transport scattering cross-sectionSsca;λð1�gλÞ was smaller than the absorption cross-section since gλapproached unity. The absorption peaks of Chl a were apparent at435, 630, and 676 nm, based on the absorption peaks of in vivophotosynthetic pigments reported by Bidigare et al. (1990). Theshoulder from 455 to 485 nm can be attributed to superposition ofthe absorption peaks of Chl b at 475 nm and of PPC around 462and 490 nm (Bidigare et al., 1990). Another absorption peak of Chlb can be observed around 650 nm. These absorption and scatteringcross-sections obtained from direct measurements will be treatedas references in evaluating the performance of the inverse method.

5.3. Results from inverse method

5.3.1. Normal–hemispherical transmittance and reflectanceFig. 6a and b respectively show the normal–hemispherical

reflectance Rnh;λ and transmittance Tnh;λ as a function of wavelengthbetween 350 and 750 nm for C. vulgaris concentration ranging from0.111 to 111 g/L. It indicates that the transmittance decreased sharply

a

c

0

100

200

300

400

500

600

700

800

400 450 500 550 600 650 700 750

Tran

spor

t ext

inct

ion

cros

s-se

ctio

n, E

tr,ex

t,(m

2 /kg)

Wavelength, (nm)

0

100

200

300

400

500

600

700

400 450 500 550 600 650 700 750

Abs

orpt

ion

cros

s-se

ctio

n, A

abs,

(m2 /k

g)

Wavelength, (nm)

Fig. 7. Retrieved average spectral mass (a) absorption Aabs;λ , (b) transport scattering ð1�between 400 and 750 nm for biomass concentrations X of 0.111, 0.556, and 1.11 g/L.

as the mass concentration increased. In fact, for concentration largeror equal to 5.56 g/L, the transmittance, over the PAR region, fell belowthe detection limit of the spectrophotometer. On the other hand, thereflectance remained relatively constant around 4%. Interestingly, itincreased with increasing concentration beyond 700 nm due to thefact that the suspension absorption coefficient vanished and scatter-ing, including back-scattering, increased. Overall, only normal–hemi-spherical transmittance and reflectance measurements for cell con-centrations 0.111, 0.556, and 1.11 g/L had a large enough signal tonoise ratio to be used in the inverse method. Data for larger con-centrations led to very noisy and unrealistic retrieved values of bothAabs;λ and Ssca;λð1�gλÞ.

5.3.2. Model validationThe measured transmittance Tnh;λ and reflectance Rnh;λ of C. vul-

garis suspensions of biomass concentration X equal to 0.111, 0.556,and 1.11 g/L over the PAR region were in reasonably good agreementwith predictions by the modified two-flux approximation (Eqs. (29)–(37)) using the average absorption Aabs;λ and scattering Ssca;λ cross-sections and the asymmetry factor g measured directly with dilutesuspensions (see Supplementary Material). This establishes that

b

Direct measurementInverse method (X=0.111 g/L)Inverse method (X=0.556 g/L)Inverse method (X=1.11 g/L)

0

10

20

30

40

50

60

400 450 500 550 600 650 700 750

Tran

spor

t sca

tter

ing

cros

s-se

ctio

n, (1

-g)S

sca,

(m2 /k

g)

Wavelength, (nm)

gλÞSsca;λ , and (c) transport extinction Etr;ext;λ cross-sections of C. vulgaris suspensions

Page 10: Simple method for measuring the spectral absorption cross ...pilon/Publications/CES2016...Simple method for measuring the spectral absorption cross-section of microalgae Razmig Kandiliana,b,

R. Kandilian et al. / Chemical Engineering Science 146 (2016) 357–368366

(1) the radiation transfer in cuvettes containing microalgae culturescan be treated as one-dimensional and (2) the modified two-fluxapproximation is an adequate model for predicting the associatednormal–hemispherical transmittance and reflectance. Note that thetransmittance was underpredicted and reflectance was overpredictedat wavelengths between 500 nm and 600 nm. This may be due to thefact that the asymmetry factor gλ was assumed to be constant overthe PAR region. In fact, the spectral asymmetry factor gλ of C. vulgariscan vary by as much as 80% over the PAR region relative to its value at633 nm (Dauchet, 2012). Moreover, the use of transport approxima-tion may have also contributed to the inaccuracy in the predictedtransmittance and reflectance, particularly at higher cell concentra-tions. Indeed, Dombrovsky et al. (2005) state that “the error of thetransport approximation increases with optical thickness but it isinsignificant in the range of weak extinction”.

Fig. 8. Comparison of (a) fluence rate ratio GPARðzÞ=q″in and (b) local rate of photonabsorption (LRPA) AðzÞ as a function of depth in a 1 cm thick flat plate PBR pre-dicted by Eqs. (8) and (16), respectively. The predictions used the radiation char-acteristics obtained from either direct measurements or retrieved by the inversemethod for concentrations 0.111, 0.556, and 1.11 g/L.

5.3.3. Retrieved cross-sectionsFig. 7 compares the averagemass (a) absorption Aabs;λ, (b) transport

scattering ð1�gλÞSsca;λ, and (c) transport extinction Etr;ext;λ ¼ Aabs;λþð1�gλÞSsca;λ cross-sections as functions of wavelength between 400and 750 nm either measured directly or retrieved by the inversemethod from measurements of Tnh;λ and Rnh;λ for three differentconcentrations. First, the retrieved spectral average mass absorptioncross-section Aabs;λ was similar for each biomass concentration, asexpected from its definition (Eq. (6)). In addition, the values of Aabs;λretrieved by the inverse method were in relatively good agreementwith those obtained from direct measurements except above 700 nmwhere it should vanish. Note that the inverse method retrieved Aabs;λfor each wavelength independently. Yet, the retrieved spectralabsorption cross-section Aabs;λ was a continuous and smooth functionof wavelength. On the other hand, the retrieved transport scattering

ν

ν

ν

ν

Fig. 9. Comparison of biomass productivity ⟨rX ⟩ of C. vulgaris grown in a 1 cm thickflat-plate PBRs exposed to 200 or 500 μmolhν=m2 s incident white LED light, pre-dicted using Eqs. (12)–(16) as a function of either (a) the steady-state biomassconcentration X or (b) the dilution rate D. The predictions used the radiationcharacteristics obtained from either direct measurements or retrieved by theinverse method.

Page 11: Simple method for measuring the spectral absorption cross ...pilon/Publications/CES2016...Simple method for measuring the spectral absorption cross-section of microalgae Razmig Kandiliana,b,

R. Kandilian et al. / Chemical Engineering Science 146 (2016) 357–368 367

cross-section ð1�gλÞSsca;λ depended on the concentration consideredand featured non-physical fluctuations as a function of wavelength. Itwas also significantly smaller than the values obtained from directmeasurements. Finally, excellent agreement was found between allthree retrieved transport extinction cross-sections Etr;ext;λ and theexperimentally measured one. Note that the inverse method over-estimated the value of Aabs;λ and compensated for it by under-predicting the transport scattering cross-section ð1�gλÞSsca;λ of themicroorganisms. This was due to the fact that normal–hemisphericaltransmittance Tnh;λ and reflectance Rnh;λ measurements of microalgaewere not very sensitive to the cells' transport scattering cross-sectionas microalgae scattered light mainly in the forward direction. Inaddition, the forward scattered light was collected by the integratingsphere used in the measurements of Tnh;λ. In fact, the absorptioncross-section dominated over the transport scattering cross-section(i.e., Aabs;λ⪢ð1�gλÞSsca;λ) so that small errors in Aabs;λ had strongeffects on the shape and magnitude of the transport scattering cross-section ð1�gλÞSsca;λ. This was confirmed by the fact that goodagreement was observed between the direct and inverse methods forthe transport extinction cross-section Etr;ext;λ.

Moreover, the same inverse procedure and algorithm was usedto retrieve the three parameters Aabs;λ, Ssca;λ, and gλ independently.This resulted in strongly oscillating values of Ssca;λ and gλ. How-ever, the same values of Aabs;λ and ð1�gλÞSsca;λ as those shown inFig. 7 were obtained. This can be attributed to the facts that (i) thesolution of the ill-posed inverse problem was not unique and (ii)the modified two-flux approximation for the expressions ofTnh;λ;pred and Rnh;λ;pred depended on the transport scattering coef-ficient σs;tr;λ and not on σs;λ and gλ separately.

As argued earlier, the average mass absorption cross-sectionAabs;λ is the most important radiation characteristic in analyzing,predicting, and controlling coupled light transfer and growthkinetics in PBRs. In fact, Fig. 8a shows the fluence rate ratioGPARðzÞ=q″in in a 1 cm thick flat plate PBR illuminated normally onone side with a front transparent window and a back reflectivityρλ ¼ 0:02. The fluence rate ratio was predicted by using either(i) the two-flux approximation given by Eq. (8) based on thedirectly measured radiation characteristics Aabs;λ and Ssca;λ with bλ¼ 0:0042 or (ii) the simplified two-flux approximation given by Eq.(11) based on the retrieved values of Aabs;λ for concentrations 0.111,0.556, and 1.11 g/L.

Fig. 8b compares the local rate of photon absorption (LRPA) AðzÞ as a function of flat plate PBR thickness predicted by Eq. (16)using the fluence rate shown in Fig. 8a and the associatedabsorption and scattering cross-sections. Here, the light sourcewas assumed to be collimated white LEDs delivering a total inci-dent flux of 200 μmolhν=m2 s as previously used experimentally(Soulies et al., 2016). Fig. 8 indicates that the absorption cross-section retrieved by the inverse method combined with the sim-plified two-flux approximation (Eq. (11)) led to acceptable pre-dictions of the fluence rate GPAR(z) and of the LRPA AðzÞ. This wasachieved while requiring fewer experimental measurements andonly an integrating sphere spectrophotometer.

Finally, Fig. 9a and b shows the volume averaged steady-statebiomass productivity ⟨rX ⟩ of C. vulgaris as a function of steady-statebiomass concentration X and dilution rate D, respectively. Productivitywas predicted using Eqs. (12)–(16) combined with either (i) the two-flux approximation given by Eq. (8) and the directly measuredradiation characteristics Aabs;λ, Ssca;λ with bλ ¼ 0:0042 or (ii) thesimplified two-flux approximation given by Eq. (11) and the retrievedvalues of Aabs;λ. A torus type 4 cm thick flat-plate PBR in continuousmode was simulated exposed to 200 or 500 μmolhν=m

2 s of colli-mated white LEDs and the dilution rate D varied from 0.02 to 0.1 h�1

as previously used experimentally (Soulies et al., 2016). Fig. 9 illus-trates that the productivity predicted using the fluence rate estimatedwith the simplified two-flux approximation and the absorption cross-

section retrieved by the inverse method was very similar to pro-ductivity predictions using the fluence rate estimated by the two-fluxapproximation and the directly measured radiation characteristics.Indeed, the relative difference between these two methods was lessthan 10% for all dilution rates D investigated and for both realisticincident irradiances considered.

6. Conclusion

This paper presented a method to easily measure the spectralabsorption cross-section of randomly oriented microalgae of anyshape from conventional normal–hemispherical transmittanceand reflectance measurements of well-mixed suspensions inconventional cuvettes. It was successfully demonstrated withsuspension of C. vulgaris with relatively large biomass concentra-tions such that multiple scattering prevailed. Even though themethod was unable to properly retrieve the transport scatteringcross-section, the retrieval of the absorption cross-section com-bined with the simplified two-flux approximation was sufficient topredict accurately the local fluence rate, the local rate of photonabsorption, and the biomass productivity of flat plate PBRs. Thisnew method can also be applied to other absorbing and scatteringparticles of various shapes with large enough size parameter toensure strong forward scattering and therefore a small transportscattering cross-section compared with the absorption cross-section.

Acknowledgement

This work was supported by the French National ResearchAgency project DIESALG (ANR-12-BIME-0001-02) for biodieselproduction based on solar production of microalgae, and is part ofthe French “BIOSOLIS” research program on developing photo-bioreactor technologies for mass-scale solar production (http://www.biosolis.org/). Additional support was provided by the CNRSprogram Cellule énergie. R.K. is grateful to the Embassy of Francein the United States, Office for Science and Technology for theChateaubriand Fellowship. L.P. thanks the Région Pays de la Loirefor the Research Chair for International Junior Scientists.

Appendix A. Supplementary data

Supplementary data associated with this paper can be found inthe online version at http://dx.doi.org/10.1016/j.ces.2016.02.039.

References

Agrawal, B.M., Mengüç, M.P., 1991. Forward and inverse analysis of single andmultiple scattering of collimated radiation in an axisymmetric system. Int. J.Heat Mass Transf. 34, 633–647.

Bellini, S., Bendoula, R., Latrille, E., Roger, J.-M., 2014. Potential of a spectroscopicmeasurement method using adding-doubling to retrieve the bulk opticalproperties of dense microalgal media. Appl. Spectrosc. 68 (October (10)),1154–1167.

Berberoğlu, H., Pilon, L., 2007. Experimental measurement of the radiation char-acteristics of Anabaena variabilis ATCC 29413-U and Rhodobacter sphaeroidesATCC 49419. Int. J. Hydrog. Energy 32, 4772–4785.

Berberoğlu, H., Yin, J., Pilon, L., 2007. Light transfer in bubble sparged photo-bioreactors for H2 production and CO2 mitigation. Int. J. Hydrog. Energy 32,2273–2285.

Berberoğlu, H., Melis, A., Pilon, L., 2008. Radiation characteristics of Chlamydomonasreinhardtii CC125 and its truncated chlorophyll antenna transformants tla1,tlaX, and tla1-CWþ . Int. J. Hydrog. Energy 33, 6467–6483.

Berberoğlu, H., Gomez, P.S., Pilon, L., 2009. Radiation characteristics of Botryococcusbraunii, Chlorococcum littorale, and Chlorella sp used for CO2 fixation and biofuelproduction. J. Quant. Spectrosc. Radiat. Transf. 110, 1879–1893.

Page 12: Simple method for measuring the spectral absorption cross ...pilon/Publications/CES2016...Simple method for measuring the spectral absorption cross-section of microalgae Razmig Kandiliana,b,

R. Kandilian et al. / Chemical Engineering Science 146 (2016) 357–368368

Bidigare, R., Ondrusek, M., Morrow, J., Kiefer, D., 1990. In vivo absorption propertiesof algal pigments. Ocean Opt. X 1302, 290–301.

Brennan, L., Owende, P., 2013. Biofuels frommicroalgae: towards meeting advancedfuel standards. In: Lee, James W. (Ed.), Advanced Biofuels and Bioproducts.Springer, New York, pp. 553–599.

Charbonneau, P., Knapp, B., 1995. A User's Guide to PIKAIA 1.0. Technical Report,NCAR Technical Note 418þ IA, National Center for Atmospheric Research.

Charbonneau, P., 2002. An Introduction to Genetic Algorithms for NumericalOptimization. Technical Report, NCAR Technical Note 450þ IA, National Centerfor Atmospheric Research.

Charbonneau, P., 2002. Release Notes for PIKAIA 1.2. Technical Report, NCARTechnical Note 451þSTR, National Center for Atmospheric Research.

Cornet, J.F., Dussap, C.G., 2009. A simple and reliable formula for assessment ofmaximum volumetric productivities in photobioreactors. Biotechnol. Prog. 25,424–435.

Cornet, J.-F., Dussap, C.G., Dubertret, G., 1992. A structured model for simulation ofcultures of the cyanobacterium Spirulina platensis in photobioreactors: I. Cou-pling between light transfer and growth kinetics. Biotechnol. Bioeng. 40,817–825.

Cornet, J.-F., Dussap, C.G., Cluzel, P., Dubertret, G., 1992. A structured model forsimulation of cultures of the cyanobacterium Spirulina platensis in photo-bioreactors: II. Identification of kinetic parameters under light and minerallimitations. Biotechnol. Bioeng. 40, 826–834.

Cornet, J.-F., Dussap, C.G., Gross, J.B., Binois, C., Lasseur, C., 1995. A simplifiedmonodimensional approach for modeling coupling between radiant lighttransfer and growth kinetics in photobioreactors. Chem. Eng. Sci. 50,1489–1500.

Cornet, J.-F., Dussap, C.G., Gros, J.-B., 1998. Kinetics and energetics of photosyntheticmicro-organisms in photobioreactors. In: Bioprocess and Algae Reactor Tech-nology, Apoptosis, Advances in Biochemical Engineering Biotechnology, vol. 59.Springer, Berlin, Heidelberg, pp. 153–224.

Dauchet, J., Blanco, S., Cornet, J.-F., Fournier, R., 2015. Calculation of the radiativeproperties of photosynthetic microorganisms. J. Quant. Spectrosc. Radiat.Transf. 161, 60–84.

Dauchet, J., 2012. Analyse Radiative des Photobioréacteurs (Ph.D. thesis). UniversitéBlaise Pascal, Clermont Ferrand II, France.

Davies-Colley, R.J., Pridmore, R.D., Hewitt, J.E., 1986. Optical properties of somefreshwater phytoplanktonic algae. Hydrobiologia 133, 165–178.

Dombrovsky, L., Randrianalisoa, H., Baillis, D., Pilon, L., 2005. Use of Mie theory toanalyze experimental data to identify infrared properties of fused quartz con-taining bubbles. Appl. Opt. 44 (33), 7021–7031.

Feng, Y., Li, C., Zhang, D., 2011. Lipid production of Chlorella vulgaris cultured inartificial wastewater medium. Bioresour. Technol. 102 (1), 101–105.

Fouchard, S., Pruvost, J., Degrenne, B., Titica, M., Legrand, J., 2009. Kinetic modelingof light limitation and sulfur deprivation effects in the induction of hydrogenproduction with Chlamydomonas reinhardtii: part I. Model development andparameter identification. Biotechnol. Bioeng. 102, 232–245.

Griffiths, M.J., Harrison, S.T.L., 2009. Lipid productivity as a key characteristic forchoosing algal species for biodiesel production. J. Appl. Phycol. 21 (5), 493–507.

Heng, R.-L., Pilon, L., 2014. Time-dependent radiation characteristics of Nanno-chloropsis oculata during batch culture. J. Quant. Spectrosc. Radiat. Transf. 144,154–163.

Heng, R.-L., Lee, E., Pilon, L., 2014. Radiation characteristics and optical properties offilamentous cyanobacteria Anabaena cylindrica. J. Opt. Soc. Am. A 31 (4),836–845.

Heng, R.-L., Sy, K.C., Pilon, L., 2015. Absorption and scattering by bispheres, quad-spheres, and circular rings of spheres and their equivalent coated spheres. J.Opt. Soc. Am. A 32 (1), 46–60.

Heraeus, 2015. Quartz glass for optics: data and properties, ⟨http://www.heraeus-quarzglas.com/media/webmedia_local/downloads/broschren_mo/DataandProperties_Optics_fusedsilica.pdf⟩.

Kandilian, R., Lee, E., Pilon, L., 2013. Radiation and optical properties of Nanno-chloropsis oculata grown under different irradiances and spectra. Bioresour.Technol. 137, 63–73.

Kandilian, R., Pruvost, J., Legrand, J., Pilon, L., 2014. Influence of light absorption rateby Nannochloropsis oculata on triglyceride production during nitrogen starva-tion. Bioresour. Technol. 163, 308–319.

Kandilian, R., Tsao, T.-C., Pilon, L., 2014. Control of incident irradiance on a batchoperated flat-plate photobioreactor. Chem. Eng. Sci. 119, 99–108.

Kandilian, R., Heng, R.-L., Pilon, L., 2015. Absorption and scattering by fractalaggregates and by their equivalent coated spheres. J. Quant. Spectrosc. Radiat.Transf. 151, 310–326.

Kong, B., Vigil, R.D., 2014. Simulation of photosynthetically active radiation dis-tribution in algal photobioreactors using a multidimensional spectral radiationmodel. Bioresour. Technol. 158, 141–148.

Lee, E., Pilon, L., 2013. Absorption and scattering by long and randomly orientedlinear chains of spheres. J. Opt. Soc. Am. A 30 (9), 1892–1900.

Lee, E., Pruvost, J., He, X., Munipalli, R., Pilon, L., 2014. Design tool and guidelines foroutdoor photobioreactors. Chem. Eng. Sci. 106, 18–29.

McKellar, B.H.J., Box, M.A., 1981. The scaling group of the radiative transfer equa-tion. J. Atmos. Sci. 38 (5), 1063–1068.

Mendes, R.L., Nobre, B.P., Cardoso, M.T., Pereira, A.P., Palavra, A.F., 2003. Super-critical carbon dioxide extraction of compounds with pharmaceutical impor-tance from microalgae. Inorg. Chim. Acta 356, 328–334.

Modest, M.F., 2013. Radiative Heat Transfer, third ed. Elsevier, Oxford, UK.Murphy, T.E., Berberoğlu, H., 2011. Effect of algae pigmentation on photobioreactor

productivity and scale-up: a light transfer perspective. J. Quant. Spectrosc.Radiat. Transf. 112, 2826–2834.

Pilon, L., Berberoğlu, H., Kandilian, R., 2011. Radiation transfer in photobiologicalcarbon dioxide fixation and fuel production by microalgae. J. Quant. Spectrosc.Radiat. Transf. 112, 2639–2660.

Pilon, L., 2015. Excel spreadsheet for retrieving the spectral absorption cross-section from normal–hemispherical reflectance and transmittance, ⟨http://repositories.cdlib.org/escholarship/⟩ or ⟨http://www.seas.ucla.edu/pilon/downloads.htm⟩ (accessed 15.06.15).

Pires, J.C.M., 2015. Chapter 5 – mass production of microalgae. In: Kim, Se-Kwon(Ed.), Handbook of Marine Microalgae. Academic Press, Boston, pp. 55–68.

Pottier, L., Pruvost, J., Deremetz, J., Cornet, J.-F., Legrand, J., Dussap, C.G., 2005. Afully predictive model for one-dimensional light attenuation by Chlamydomo-nas reinhardtii in a torous photobioreactor. Biotechnol. Bioeng. 91, 569–582.

Prahl, S.A., Van Gemert, M.J.C., Welch, A.J., 1993. Determining the optical propertiesof turbid media by using the adding-doubling method. Appl. Opt. 32 (4),559–568.

Privoznik, K.G., Daniel, K.J., Incropera, F.P., 1978. Absorption,extinction, and phasefunction measurements for algal suspensions of Chlorella pyrenoidosa. J. Quant.Spectrosc. Radiat. Transf. 20, 345–352.

Pruvost, J., Cornet, J.F., Goetz, V., Legrand, J., 2012. Theoretical investigation ofbiomass productivities achievable in solar rectangular photobioreactors for thecyanobacterium Arthrospira platensis. Biotechnol. Prog. 28, 699–714.

Pruvost, J., Le Gouic, B., Lepine, O., Legrand, J., Le Borgne, F., 2016. Microalgae culturein building-integrated photobioreactors: biomass production modelling andenergetic analysis. Chem. Eng. J. 284, 850–861.

Richmond, A., 2004. Handbook of Microalgal Culture. Blackwell Science Ltd, Oxford,UK.

Ritchie, R., 2006. Consistent sets of spectrophotometric chlorophyll equations foracetone, methanol and ethanol solvents. Photosynth. Res. 89, 27–41.

Safi, C., Zebib, B., Merah, O., Pontalier, P.-Y., Vaca-Garcia, C., 2014. Morphology,composition, production, processing and applications of Chlorella vulgaris: areview. Renew. Sustain. Energy Rev. 35, 265–278.

Soulies, A., Pruvost, J., Castelain, C., Burghelea, T.I., Legrand, J., 2016. Investigationand modeling of the effects of light spectrum and incident angle on Chlorellavulgaris in photobioreactors. Biotechnol. Prog. (in press)

Stramski, D., Piskozub, J., 2003. Estimation of scattering error in spectrophotometricmeasurements of light absorption by aquatic particles from three-dimensionalradiative transfer simulations. Appl. Opt. 42 (18), 3634–3646.

Strickland, J.D., Parsons, T.R., 1968. A Practical Handbook of Seawater Analysis.Fisheries Research Board of Canada, Ottawa, Canada.

Takache, H., Christophe, G., Cornet, J.-F., Pruvost, J., 2010. Experimental and theo-retical assessment of maximum productivities for the microalgae Chlamydo-monas reinhardtii in two different geometries of photobioreactors. Biotechnol.Prog. 26 (2), 431–440.

Takache, H., Pruvost, J., Cornet, J.-F., 2012. Kinetic modeling of the photosyntheticgrowth of Chlamydomonas reinhardtii in a photobioreactor. Biotechnol. Prog. 28(3), 681–692.

Tokuşoglu, Ö, Ünal, M.K., 2003. Biomass nutrient profiles of three microalgae:Spirulina platensis, Chlorella vulgaris, and Isochrisis galbana. J. Food Sci. 68 (4),1144–1148.

Van De Hulst, H.C., 1957. Light Scattering by Small Particles. Wiley, New York, NY.Wheaton, Z.C., Krishnamoorthy, G., 2012. Modeling radiative transfer in photo-

bioreactors for algal growth. Comput. Electron. Agric. 87, 64–73.