Thermal Reactor Model for Large-Scale Algae ... - ACS Publications

Mar 7, 2016 - Thermal Reactor Model for Large-Scale Algae Cultivation in Vertical. Flat Panel ... productivity of microalgae strongly depends on the c...
0 downloads 0 Views 1MB Size
Subscriber access provided by the Henry Madden Library | California State University, Fresno

Article

A Thermal Reactor Model for Large-Scale Algae Cultivation in Vertical Flat Panel Photobioreactors Christian Hermann Endres, Arne Roth, and Thomas Brück Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.5b05414 • Publication Date (Web): 07 Mar 2016 Downloaded from http://pubs.acs.org on March 13, 2016

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Environmental Science & Technology is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 26

Environmental Science & Technology

1

TITLE

2

A Thermal Reactor Model for Large-Scale Algae Cultivation in Vertical Flat Panel

3

Photobioreactors

4 5 6

AUTHORSHIP

7

Christian H. Endresa,b,*, Arne Rotha, Thomas B. Brückb

8 9 10

AFFILIATIONS

11

a

12

Bauhaus Luftfahrt, Willy-Messerschmitt-Str. 1, 85521 Ottobrunn, Germany, Tel. +49-

13

89307484948, E-mail: [email protected]

14 15

b

16

Division of Industrial Biocatalysis, Department of Chemistry, Technische Universität

17

München, Lichtenbergstraße 4, 85748 Garching, Germany

18 19 20

ABSTRACT

21

Microalgae grow significantly faster than terrestrial plants and are a promising feedstock for

22

sustainable value added products encompassing pharmaceuticals, pigments, proteins and most

23

prominently biofuels. As the biomass productivity of microalgae strongly depends on the

24

cultivation temperature detailed information of the reactor temperature as a function of time

25

and geographical location is essential to evaluate the true potential of microalgae as an

26

industrial feedstock. In the present study a temperature model for an array of vertical flat plate 1 ACS Paragon Plus Environment

Environmental Science & Technology

Page 2 of 26

27

photobioreactors is presented. It was demonstrated that mutual shading of reactor panels has a

28

decisive effect on the reactor temperature. By optimizing distance and thickness of the panels

29

the occurrence of extreme temperatures and the amplitude of daily temperature fluctuations in

30

the culture medium can be drastically reduced, while maintaining a high level of irradiation

31

on the panels. The presented model was developed and applied to analyze the suitability of

32

various climate zones for algae production in flat panel photobioreactors. Our results

33

demonstrate that in particular Mediterranean and tropical climates represent favorable

34

locations. Lastly, the thermal energy demand required for the case of active temperature

35

control is determined for several locations.

36 37 38

TOC ABSTRACT

39

40 41

2 ACS Paragon Plus Environment

Page 3 of 26

Environmental Science & Technology

42

INTRODUCTION

43

Microalgae grow significantly faster than terrestrial plants and can accumulate large quantities

44

of lipids.1–5 Additionally agricultural land or freshwater is not necessarily required for algae

45

cultivation.4–8 Therefore, microalgae are considered a promising sustainable feedstock for a

46

broad range of value added products encompassing pharmaceuticals, pigments, proteins and,

47

most prominently, biofuels.1,9–12 Various concepts for cultivating microalgae are currently

48

under development, generally classified as either closed photobioreactors or open ponds.

49

While open pond cultivation is normally characterized by a simple system configuration,

50

closed photobioreactors offer high controllability of growth conditions, albeit at the cost of

51

increased system complexity.2,13,14

52

The crucial aspect of algae cultivation from an economic perspective is the biomass yield in

53

relation to the required effort and investments. One of the most important parameters

54

effecting the yield is the temperature of the cultivation medium (in addition to solar

55

irradiation and nutrient availability).15,16 A typical range for algae cultivation is 10 °C to

56

40 °C. Temperatures below 10 °C usually result in very low growth rates. At subzero

57

temperatures ice formation in the reactor and at instrumentations poses an additional problem.

58

Temperatures above the 40 °C-threshold are only tolerated by few thermophilic algae and

59

may lead to cell death in case of less adapted species. It is therefore mandatory to keep algae

60

within a favorable temperature regime, preferably close to the optimum temperature of the

61

respective strain, to guarantee high biomass production rates.

62

In a laboratory environment temperature can easily be controlled. This cannot be applied to

63

the same extent in an industrial-sized plant, as temperature regulation would require the

64

installation of heat exchangers, pumps and pipes, thus substantially adding to capital and

65

energy costs. However, without active temperature control, closed photobioreactor systems

66

may overheat during hot days with reactor temperatures reaching values up to 55 °C.17

67

Consequently, it is crucial to evaluate the time-dependent reactor temperature profile already 3 ACS Paragon Plus Environment

Environmental Science & Technology

Page 4 of 26

68

in the planning phase of a commercial microalgae cultivation plant. In this respect an accurate

69

process simulation allows the assessment of the economic potential with reference to a given

70

geographical location and plant design.

71

Recently, several studies reported on the development of temperature models and

72

complementary research involving various types of microalgae cultivation systems.18–23 Those

73

studies differ widely in terms of scope (e.g. type of reactor system) and levels of accuracy.

74

One of the most advanced temperature models for a closed photobioreactor system has been

75

reported by Béchet et al.18,19 Their model computes the reactor’s heat balance as a function of

76

various static and dynamic parameters for a bubble column reactor concept, a system mostly

77

used in laboratory-scale experiments. One current limitation of this important study is that

78

only a single, stand-alone reactor was considered. The reported approach thus neglects

79

shading effects that naturally occur in larger applications. To date, shading effects on reactor

80

temperature have only been taken into account by a single study.23 However, in the respective

81

publication shading was only examined as a static parameter, neglecting the effects of the

82

sun’s position with respect to the orientation and geometry of the panels.

83

In our study, we now introduce a dynamic temperature model for a complete array of flat

84

panel photobioreactors providing a new framework for the economic assessment of algae

85

processes under industrial relevant conditions. We have selected flat panel reactors because

86

they represent one of the most commonly used closed photobioreactor systems for academic

87

research13,21,24–26 and commercial activities alike.27,28 In our model the heat balance is

88

calculated as a function of a broad range of parameters such as reactor geometry (e.g. panel

89

thickness), inter-panel distance, orientation of reactor panels (north-south vs. east-west) as

90

well as local climatic conditions and irradiation. All first order reflections are taken into

91

account and a detailed multilayer ground model is introduced allowing for an accurate

92

calculation of its thermal radiation. Both features represent significant advances with respect

93

to the current literature. Most importantly, mutual shading of the reactor panels, calculated as 4 ACS Paragon Plus Environment

Page 5 of 26

Environmental Science & Technology

94

a dynamic function of the sun’s position and reactor dimensions is taken into account for the

95

first time. The functionality of the proposed model is demonstrated in an extensive parameter

96

study for various locations in the U.S. Trade-offs between parameters are identified and

97

discussed. It is demonstrated that shading has a massive impact on both, reactor temperature

98

and the amount of available light for photosynthesis.

99

The temperature model introduced in this paper therefore represents a substantial step forward

100

relative to the work conducted in this field so far and provides a valuable basis for the site-

101

specific estimation of microalgae productivities, and thus for the assessment of the economic

102

and environmental performance.

103 104

METHODOLOGY

105

Photobioreactor Characteristics. The cultivation system selected for this work is an array of

106

vertical flat plate photobioreactors. The single reactors are arranged in long parallel lines. An

107

illustration of the concept can be found in the supplementary information (Fig. S1). Reactors

108

at the borders of the array are neglected as they receive higher levels of irradiation not being

109

representative for the majority of reactors in the field. Each single reactor is defined by its

110

dimensions (panel thickness, height and width) and the distance to the opposing panels. The

111

reactor volume is given by the respective reactor dimensions. Within this study the width and

112

height of the reactor are kept constant at 2 m and 1 m, respectively. Even though a fixed panel

113

height is used throughout the publication, the presented results can be easily transferred to

114

other reactor heights, provided that the ratio between reactor height and panel distance is kept

115

constant (Fig. S8, supplementary information). Thus the thermal behavior of a reactor that is 1

116

m high and 0.5 m apart from the next panel is basically identical to a reactor of 2 m in height

117

with a panel distance of 1 m.

118

For heat exchange only the back and front surfaces of the reactor panel are considered and the

119

small areas at the edges of the reactor are neglected. The culture medium is continuously 5 ACS Paragon Plus Environment

Environmental Science & Technology

120

homogenized (pneumatic agitation), therefore temperature is assumed to be constant over the

121

reactor volume at a certain point in time. As the reactor wall is thin, the temperature of the

122

wall is considered to be identical to the temperature of the culture medium. Site-specific

123

meteorological and irradiation data were adopted from the National Solar Radiation Data

124

Base.29 The provided data sets describe a typical meteorological year and are specifically

125

intended for computer simulations of solar energy conversion systems.

Page 6 of 26

126 127

Temperature Modeling Approach. The calculation of the reactor temperature is based on a

128

balance of all relevant heat fluxes, which can be expressed by the following equation

129 130

Equation 1 R R R 2 − 1 =    ∙ ∆, 

131 132

with the reactor volume VR, the density ρR (997 kg m-3) 30 and the specific heat capacity of the

133

culture medium cpR (4181 J kg-1 K-1) 30T1 and T2 are the reactor temperatures at the beginning

134

and the end of the considered time interval, ∆t, and Q̇i represents heat fluxes affecting the

135

reactor.

136

Solving the equation for T2, the temperature at the end of each interval can be calculated from

137

the temperature of the previous time step, provided that all heat fluxes are known. Using

138

MATLAB® (The MathWorks®, Inc., Natick, MA) as software environment for the simulation,

139

the reactor temperature and heat fluxes are updated every minute, resulting in 525 601 data

140

points for a complete year. The time to generate a single temperature profile amounts to

141

approximately 12 min (Intel® Core™i5 2.53 GHz, 4 GB RAM). The starting temperature for

142

the culture medium was set to 20 °C and the following heat fluxes are considered in the

143

model: 6 ACS Paragon Plus Environment

Page 7 of 26

Environmental Science & Technology

144 145

- Direct sunlight

146

- Diffuse sunlight

147

- Atmospheric long-wave irradiation

148

- Heat radiation from the reactor panels

149

- Heat radiation from the ground

150

- Reflections of direct, diffuse and thermal radiation at the reactor panels

151

- Reflections of direct, diffuse and thermal radiation at the ground

152

- Heat transfer through natural air convection

153

- Heat transfer related to the aeration of the reactor panels

154 155

Calculation Details of Heat Fluxes. In the following fundamental assumptions and details

156

regarding the calculation of the heat fluxes are explained. For a complete mathematical

157

description please refer to the supporting information.

158

The amount of direct sunlight contributing to increase the reactor temperature depends on the

159

angle the sun beams make with the reactor surface. This angle of incidence is calculated for

160

every data point.31 Part of the incident sunlight is lost for the reactors due to reflections at the

161

reactor wall. The reflection losses are calculated with the Fresnel equation, assuming the

162

reflectivity index of air, the reactor wall (glass, plastic) and the culture medium (water) being

163

1, 1.5 and 1.33, respectively. In contrast to the case of an isolated reactor, an array of flat plate

164

photobioreactors causes shading, and thus reduces the amount of light reaching the panel

165

surface. The equations used in this study to determine this heat flux are based on published

166

literature32.

167

Not all sunlight reaching the surface of the reactor is converted into heat, but part of it is

168

scattered/reflected back by the algae cells. Béchet et al.18 approximated the absorptivity of

169

algae biomass based on Kirchhoff’s law of thermal radiation that states that for a given 7 ACS Paragon Plus Environment

Environmental Science & Technology

Page 8 of 26

170

wavelength, emissivity and absorptivity of a material are identical. However, the emissivity is

171

often measured for wavelengths in the far infrared and is therefore not suited to determine the

172

absorptivity in the visible spectrum of light. This becomes obvious when looking at the

173

emissivity of water, which has a value of around 0.9 for a temperature of 273 K.33 Assuming

174

that the emissivity determined for infrared radiation equals the absorptivity in the visible

175

spectrum of light, water would absorb 90 % of the incoming light. As most of the light is

176

absorbed, water would appear as very dark body. In the present study the emissivity is only

177

used for radiation in the far infrared, while the albedo is utilized as a measure for the

178

reflectivity of the algae cells in the visible spectrum of light. For dense algae cultures an

179

albedo of 0.3 is used. This value is in accordance with typical values for thick, plant leaves.34.

180

As we assume that light is either absorbed or reflected by the opaque algae culture,

181

transmission through the panels is zero and not considered in the temperature model. The

182

influence of the albedo on central outcomes of the temperature simulation is further discussed

183

in the supporting information (Sec. S2.2).

184

Atmospheric longwave radiation is thermal radiation from the air layers above the cultivation

185

plant. Therefore, this heat flux can be expressed by the Stefan-Boltzmann-law. The emissivity

186

of the atmosphere depends on factors such as the humidity of the air and cloud coverage. In

187

our model the Brutsaert-Equation35 in combination with the cloud cover model by Crawford

188

and Duchon36 is used to calculate this emissivity.

189

Analogous to atmospheric radiation, heat radiation of the ground is calculated with the Stefan-

190

Boltzmann equation. The temperature of the top soil layer is determined with a multilayer

191

ground temperature model. The heat exchange between different soil layers is taken into

192

account as well as the property of the ground to store thermal energy. The first, top layer is

193

exposed to the environment and thus affected by irradiation and convection. Additionally, all

194

ground layers exchange thermal energy via heat conduction. In total the model consists of 13

195

layers of varying thickness and reaches down 16.4 m. The first layer is 2 mm thick and 8 ACS Paragon Plus Environment

Page 9 of 26

Environmental Science & Technology

196

thickness doubles for every further layer. Very fine layers near the top soil are required to

197

display daily temperature fluctuations. With increasing depth these fluctuations are less

198

pronounced and therefore thicker layers are considered in order to reduce computing time. At

199

a depth of 16.4 m the temperature is considered constant during the year, which is in

200

accordance with actual measurements of the soil temperature.37,38

201

As in particular deep soil layers are capable to store thermal energy for a long time, it is

202

essential to choose realistic ground temperatures as starting condition for the simulation. In

203

order to determine these temperatures, an iterative procedure is used, where the ground

204

temperatures of the last time step in December form the starting condition of the next

205

iteration. A detailed description of the ground layer model can be found in the supporting

206

information.

207

First order reflections between the panels and between panel and ground are considered in the

208

model (see supporting information). The heat exchange between the air and a vertical flat

209

plate is calculated based on equations for natural convection.33 For simplification it is

210

assumed that convection at a panel surface is not influenced by convection at the ground or

211

other panels. Forced convection caused by wind is neglected for two reasons. Firstly, wind

212

speed is generally measured 10 m above the ground, thus wind speed near the ground would

213

be substantially lower. Secondly, the array of vertical flat plate photobioreactors provides

214

protection from wind.

215

The last heat flux considered in the model is heat exchange caused by aeration of the

216

photobioreactors. Various elements contribute to this heat flux. First, heat is exchanged due to

217

the different temperatures of the instreaming gas bubbles and the culture medium. Second,

218

during their way through the reactor, gas bubbles are saturated with water from the medium.

219

Third, prior to injection, the gas is compressed in order to overcome the hydraulic pressure in

220

the reactor. As the bubbles rise through the reactor they expand releasing the compression

221

energy. All these elements are taken account of to calculate the heat flux. 9 ACS Paragon Plus Environment

Environmental Science & Technology

Page 10 of 26

222 223

RESULTS AND DISCUSSION

224

Temperature Profiles as Function of Geographic Location. In a primary step, the general

225

temperature profiles are analyzed and characterized (Fig. 1). For this purpose, we selected

226

three US locations within three different climate zones. Of the chosen locations, Boston in

227

Massachusetts corresponds to a cold climate according to the Köpper-Geiger climate

228

classification.39 By contrast, Sacramento in California and Hilo in Hawaii are situated more

229

south and therefore have a temperate climate, with dry summers (Mediterranean) and a

230

tropical climate respectively. Temperature profiles of additional locations, varying panel

231

distances and east-west panel orientation are described in the supporting information (Fig. S9

232

– S11). For the interpretation of the presented temperature profiles it is important to keep in

233

mind that the selected panel distance of 0.5 m already results in significant shading of the

234

panels.

235

10 ACS Paragon Plus Environment

Page 11 of 26

Environmental Science & Technology

236 237 238

Figure 1. Temporal temperature profile for photobioreactors located in (A) Boston, MA, (B)

239

Sacromento, CA and (C) Hilo, HI. Black and grey lines correspond to the reactor and air

240

temperature respectively. (panel distance: 0.5 m, panel thickness: 0.05 m, orientation: north-

241

south)

242 243

In all instances the reactor temperature (black line) follows the daily and diurnal fluctuations

244

of the air temperature (grey line). However, the average temperature of the reactors is slightly

245

warmer and day-night-cycles are more pronounced. In summer the reactor temperature may

246

exceed 40 °C even for cold climates such as Boston, posing a potential threat for cell

247

cultures.15,16 Additionally, reactor and air temperature in Boston drops to 0 °C and below for 11 ACS Paragon Plus Environment

Environmental Science & Technology

248

nearly one quarter of the year. Under these circumstances the cultivation time is rather

249

limited. Both Sacramento and Hilo show less extreme reactor temperature profiles. In

250

Sacramento, the reactor temperature drops below 0 °C only on few days of the year while it

251

occasionally exceeds 40 °C. In Hilo on Hawaii, the reactor and air temperature stays

252

constantly at an elevated level, which is typical for a tropical climate. However, the reactor

253

temperature also exceeds 40 °C at some days for the chosen reactor geometry.

254

As demonstrated above, Sacramento with its Mediterranean climate is well suited for

255

cultivating algae. Moreover, the profile shows a seasonal temperature cycle, typical for all

256

non-tropical regions. Thus, Sacramento is selected as exemplary site for further parameter

257

studies (Fig. 2 to 4).

Page 12 of 26

258 259

Analysis of Affecting Heat Fluxes. To optimize thermal management, it is essential to

260

understand the impact of the various heat fluxes on the total heat balance of the reactor. We

261

thus analyzed the time-related profiles of the most important heat fluxes during three

262

exemplary days in late spring (Fig. 2A).

263

12 ACS Paragon Plus Environment

Page 13 of 26

Environmental Science & Technology

264 265 266

Figure 2. (A) Temporal course of the most important heat fluxes affecting the reactor

267

temperature. (B) Yearly average of the positive value of the heat fluxes. (for both subfigures:

268

location: Sacramento, panel distance: 0.5 m, panel thickness 0.05 m, orientation: north-south)

269 270

As expected most heat fluxes follow the typical day-night-cycle. Additionally, heat fluxes

271

such as direct (“DNI”) and diffuse irradiation (“DHI”) strongly depend on daily weather

272

conditions. For days with essentially no clouds, heat resulting from direct sunlight is very

273

high (day 114). In contrast, for a 100 %-cloud coverage (day 116), direct sunlight is

274

significantly reduced, while at the same time diffuse radiation increases, as more light is

275

scattered by the water molecules in the atmosphere. Diffuse light is partially compensating for 13 ACS Paragon Plus Environment

Environmental Science & Technology

Page 14 of 26

276

the losses of direct irradiation, but the overall heat flux still decreases with increasing cloud

277

coverage. When clouds occasionally pass the sun (day 115), the temporal profiles of direct

278

radiation displays several peaks in contrast to the smooth profile of day 114. The individual

279

peaks represent phases of partially blue sky, when direct light is momentarily not blocked by

280

clouds.

281

To quantify the total impact of the heat fluxes the yearly average of the positive value is

282

calculated and compared (Fig. 2B). On a yearly basis, the reactor heat radiation (“Reac. IR”)

283

is the most important of all heat fluxes, followed by the ground heat radiation (“Ground IR”)

284

and the atmospheric heat radiation (“Atmos. IR”). Heat from the visible spectrum of light,

285

DNI and DHI, contributes only to around 6 % to the total heat flux. However, they represent

286

the only light fraction that can be converted via photosynthesis into biomass. All further heat

287

fluxes, mostly reflections of the ground and the reactor panels, contribute around 5 % to the

288

total heat flux and are not displayed separately but treated as combined heat flux in the figure

289

(“Other”). In this context, it is important to note, that even though individual reflection-related

290

heat fluxes only have minor impact on the heat balance, the sum of all these heat fluxes is

291

comparable to the irradiation of sunlight. Thus, reflections have to be taken into account for

292

simulations of the temperature in closed photobioreactors.

293 294

Captured Sunlight as Function of Reactor Geometry. A photobioreactor is a biomass

295

production system. Consequently, it is important to evaluate its ability to produce biomass in

296

the context of temperature management. As biomass production is driven by radiation in the

297

visible spectrum of light, the energy of sunlight captured by the reactor panels is used as a

298

figure of merit to analyze the performance of photobioreactor systems. (Fig. 3)

299

Two performance indicators are examined. First, captured sunlight is related to the ground

300

area occupied by the reactor panels, being a measure for the areal productivity of a given

301

panel array (lines with empty markers, left y-axis). Second, captured sunlight is related to the 14 ACS Paragon Plus Environment

Page 15 of 26

Environmental Science & Technology

302

reactor volume (lines with filled markers, right y-axis). Within certain limits, more light per

303

volume translates into higher biomass productivity and therefore into lower operational and

304

capital costs. Captured sunlight with respect to the reactor volume thus can be described as a

305

measure for the efficiency of a photobioreactor system.

306

307 308 309

Figure 3. Quantity of captured sunlight with respect to the ground area and the reactor

310

volume displayed as function of the panel distance and panel thickness. Lines with empty

311

markers and filled markers correspond to the left and right y-axis, respectively. Square, circle,

312

diamond and triangle markers indicate panel thicknesses of 0.025 m, 0.05 m, 0.1 m and

313

0.15 m. (location: Sacramento, CA, orientation: north-south)

314 315

According to Figure 3, the captured sunlight per ground area increases with decreasing panel

316

distance until it peaks at a distance between of 0.2 m and 0.4 m. This peak is a result of a

317

simplification made for the calculation in the model. Typically the top surface can be

318

considered small in comparison to the sides of the reactor. Additionally the top may be

319

covered by a frame holding the panel or instrumentations preventing the light from entering

320

the reactor. Therefore, the top area is neglected in the current model for heat exchange 15 ACS Paragon Plus Environment

Environmental Science & Technology

Page 16 of 26

321

including irradiation. However, for the case of a transparent top surface, i.e. light can enter

322

the reactor through the top, captured sunlight related to the ground area would constantly

323

increase even for small panel distances and be basically independent of the panel thickness

324

(not shown here).

325

The theoretical maximum value for the areal light capture is represented by the global

326

horizontal irradiation (GHI). For the given location, Sacramento, the average GHI is

327

6491 MJ m-2 a-1. Though, even for small panel distances this value cannot be reached, as a

328

fraction of the incident light always is reflected back to the atmosphere. For the given

329

configurations around two thirds of the GHI can be captured by the reactor panels.

330

When captured sunlight is not related to the ground area but to the reactor volume small panel

331

distances are not very favorable, as only little light is captured by the individual panel. By

332

increasing the panel distance, more and more light is captured per culture volume up to a

333

distance of around 2 m. From this point onwards the captured sunlight per reactor volume

334

stays more or less constant.

335

As small panel distances lead to good areal productivities and large panel distances result in

336

high photobioreactor efficiencies trade-offs are necessary to optimize the overall performance

337

of an algae cultivation plant. As already mentioned, panel distances above 2 m are no

338

reasonable option as the light captured by the reactor increases only very moderately while

339

the areal productivity decreases at the same time. Conversely, very small panel distances

340

below 0.3 m would require a great number of reactors to capture the sunlight resulting in high

341

investment costs. Thus, panel distances between 0.3 m and 2 m appear the most interesting

342

range in terms of biomass productivity. Most notably, small panel thicknesses exert a positive

343

effect on captured sunlight related to the reactor volume, while having neutral to positive

344

effect (depending whether the top surface area is neglected or not) when related to the ground

345

area. Thus, from a performance point of view, thin panels should be preferred in production

346

plant design. 16 ACS Paragon Plus Environment

Page 17 of 26

Environmental Science & Technology

347 348

Limitations in Cultivation Time as Result of Extreme Reactor Temperatures. The

349

thermal behavior of the reactor as a function of the panel thickness and panel distance is

350

examined in Figure 4. For this purpose the temperature range for technically and

351

economically reasonable algae cultivation is set to 0 °C – 40 °C. The temperature model is

352

applied to compute the number of days in a year when reactor temperature exceeds (Fig. 4A)

353

or drops below (Fig. 4B) those limits. The remaining days correspond to the total cultivation

354

time.

355

356 357 358

Figure 4. (A) Number of days too hot (Treactor > 40 °C) and (B) days too cold (Treactor < 0 °C)

359

for algae cultivation displayed as function of the panel distance and panel thickness. Square,

360

circle, diamond and triangle markers correspond to panel thicknesses of 0.025 m, 0.05 m,

361

0.1 m and 0.15 m, respectively. (location: Sacramento, CA, orientation: north-south)

362 17 ACS Paragon Plus Environment

Environmental Science & Technology

Page 18 of 26

363

According to our results, both panel distance and thickness of the panels strongly influence

364

the thermal behavior of the reactor. For small distances and thicknesses the number of

365

extreme temperature events in both directions of the temperature scale is very small. Thus

366

nearly year-round cultivation is possible at the exemplary site of Sacramento. However, with

367

increasing panel distances and decreasing panel thicknesses the number of days with

368

temperatures outside the tolerable range significantly increases. Interestingly, this not only

369

true for hot periods, as during cold phases tighter packed panels ‘loose’ less heat to the

370

atmosphere than further distanced panels.

371

Very thin panels are especially prone to extreme reactor temperatures. Therefore, cultivation

372

time may be reduced to less than half of the year under certain conditions. In order to still

373

make use of thin panels in outdoor cultivation, it is therefore advisable to choose a small

374

panel distance in order to reduce the number of days with extreme temperatures. Thicker

375

panels are less sensitive to extreme temperature events as they contain a larger liquid volume

376

and react slower on the influence of the various heat fluxes. Hence, from an exclusive

377

temperature management point of view, thick reactor panels should be preferred to thin

378

panels.

379

However, the reactor performance, expressed by the amount of captured sunlight, improves

380

with decreasing panel thickness (Fig. 3). Considering both, temperature management and

381

reactor performance, it is thus advisable to choose a medium panel thickness of 0.05 m to

382

0.1 m. In order to optimize areal productivity and reduce extreme temperature events a panel

383

distance of 0.5 m to 1 m appears reasonable.

384 385

Suitability of Various Geographic Locations for Algae Cultivation. Among other criteria,

386

the suitability of a location for algae cultivation depends on the quantity of sunlight provided

387

at the respective site. Moreover, temporal reactor temperatures determine the total cultivation

388

time. Taking both aspects into account for various US-locations, we calculated the quantity of 18 ACS Paragon Plus Environment

Page 19 of 26

Environmental Science & Technology

389

captured sunlight at days of tolerable reactor temperatures (0 °C – 40 °C) (Fig. 5A, light grey

390

bar). The considered spots, from north to south are: Forks (WA), Boston (MA), Sacramento

391

(CA), Phoenix (AZ), New Orleans (LA) and Hilo (HI). Further, for comparison, the GHI

392

(black bar) and the captured sunlight without temperature-related limitation in cultivation

393

time are determined (medium grey bar). Based on our previous results a panel distance of

394

0.5 m and a panel thickness of 0.05 m are chosen for the analysis.

395

396 397 398

Figure 5. (A) Captured sunlight received at days of tolerable temperatures (0 °C ≤ Treactor ≤

399

40 °C) in relation to captured sunlight for all reactor temperatures and in relation to the global

400

horizontal irradiation. (B) Heating and cooling demand for various temperature regimes and

401

locations. (for both subfigures: panel distance 0.5 m, panel thickness 0.05 m, orientation:

402

north-south)

403 19 ACS Paragon Plus Environment

Environmental Science & Technology

Page 20 of 26

404

As expected, GHI increases from the most northern location Forks over Boston and

405

Sacramento to Phoenix. As a consequence of typically high cloud coverage in subtropical and

406

tropical regions, the GHIs in New Orleans and Hilo (yearly average around 6000 MJ m-2 a-1)

407

do not reach the value of Phoenix (7500 MJ m-2 a-1).

408

According to the GHI, Phoenix in Arizona has the highest potential for algae production.

409

However, the ‘true’ production potential is much lower (light grey bar). As the temperature of

410

the culture medium in Phoenix often exceeds 40 °C, the cultivation period is limited to a

411

relatively small fraction of the year. This indicates that deserts are not a favorable site for

412

algae production in photobioreactors. According to our study the best places for algae

413

cultivation can be found in Mediterranean (Sacramento) and tropical climates (Hilo).

414 415

Energy Requirements for Active Temperature Control. Apart from passively controlling

416

the cultivation temperature by thermally optimizing the reactor design, temperature can be

417

actively controlled through a cooling and heating system. We therefore calculated the thermal

418

energy that has to be removed (negative values) or applied (positive value) in order to keep

419

the reactor temperature within certain intervals (Fig. 5B). For interpretation of the results, this

420

thermal energy demand is related to the energy content stored in form of algae biomass:

421

Assuming an areal productivity of 20 g m-2 d-1 and a heating value of 22 MJ kg-1 for algae

422

biomass40, the amount of chemical energy bound in algae biomass is 161 MJ m-2 a-1. For cold

423

climates this value is often in the same order of magnitude as the thermal energy required to

424

heat the reactors, e.g. the thermal energy required to keep the temperature above 0 °C in

425

Boston is around 400 MJ m-2 a-1. Thus heating is only a reasonable option if algae are not

426

cultivated for energetic use or if waste heat from other sectors could be adopted.

427

In order to remove thermal energy from the reactors, a cooling medium is required. Assuming

428

that 500 MJ m-2 a-1 of thermal energy need to be removed, around 35 t m-2 a-1 of water are

429

required when the temperature difference of incoming to outgoing cooling water is 15 °C. 20 ACS Paragon Plus Environment

Page 21 of 26

Environmental Science & Technology

430

Further assuming that the water source is situated 10 m below plant level, around

431

3.5 MJ m-2 a-1 of mechanical energy are required for pumping. When comparing this value to

432

the 161 MJ m-2 a-1 provided by the algae biomass cooling is in principle an acceptable option.

433

However, it has to be kept in mind that the energy to provide cooling water is usually higher

434

as friction losses in the pipes and the efficiency of the pump have to be considered. Moreover,

435

the height difference between the water source and the algae plant can be larger than 10 m,

436

resulting in a higher energy demand.

437

Both, heating and cooling of the algae reactors requires the installation of heat exchangers,

438

pumps, pipes, etc. Thus, even if cooling and heating are acceptable solutions from an

439

energetic perspective, active temperature control may still not be economically viable due to

440

high investment costs.

441 442

Implications for Commercial Algae Cultivation in Photobiobreactors. It has been

443

demonstrated that temperature simulation is important to adequately assess the potential of

444

algae cultivation at specific geographical locations. Additionally, temperature simulation can

445

be used to optimize reactor dimensioning with respect to both, light input and cultivation

446

temperature. Therefore, temperature simulation represents a significant asset to the process of

447

an integrated reactor and plant design. In this context, mutual shading of the panels has a

448

decisive effect on reactor temperature and should not be neglected.

449

Algae cultivated in outdoor photobioreactors are exposed to strong temperature variations.

450

According to our simulations, at most of the examined locations reactor temperature exceeded

451

40 °C at least once during summer. In winter, however, very cold reactor temperatures around

452

0 °C are common for nearly all climates with the exception of tropical regions. Consequently,

453

it is essential for commercial cultivation that algae strains can withstand such unfavorable

454

temperatures, in particular the occurrence of temperature peaks exceeding the 40 °C

455

threshold. High productivity over a broad temperature range might prove economically 21 ACS Paragon Plus Environment

Environmental Science & Technology

Page 22 of 26

456

beneficial and should therefore be taken into account for strain selection. As an alternative to

457

a single algae strain for year-round cultivation two strains, one adapted to cold and one to

458

warm cultivation temperatures, can be utilized in different seasons.13,41,42

459

Lastly, our study indicates that deserts represent very challenging areas for algae cultivation

460

in closed photobioreactors. Although deserts usually provide high levels of irradiation and

461

large areas of available non-arable land and are therefore suggested as suitable for algae

462

cultivation, reactor temperatures in these regions may climb to 50 °C and above. As water for

463

cooling reactors is usually rare in these areas, other technical modifications would be

464

necessary to reduce cultivation temperature. One option is the usage of nets spanned over

465

panel rows that provide shade during the hottest times of the year. The impact of such an

466

approach on reactor temperatures however is not clear and has to be addressed by future

467

research. Furthermore, technical measures for cooling are generally associated with costs.

468

Very few algae, mostly cyanobacteria, are capable of resisting cultivation temperatures of

469

50 °C.43 By using such thermophile organisms algae cultivation in uncooled photobioreactors

470

is theoretically possible even in hot environments like deserts. However, currently little

471

practical experience exists for large-scale cultivation of thermophilic algae.

472

In the context of the demonstrated importance of temperature simulation additional research

473

should be focused on extending the simulation tool to cover promising reactor types other

474

than flat panel reactors. Most important, however, is the implementation of a temperature-

475

sensitive growth model for microalgae to compute realistic productivity values for outdoor

476

algae cultivation plants and site-specific growth conditions. Such an implementation is subject

477

to our ongoing work.

478 479

ACKNOWLEDGEMENTS

480

We gratefully acknowledge the financial support by the German Federal Ministry of

481

Education and Research (Project: Advanced Biomass Value, 03SF0446C) and the support 22 ACS Paragon Plus Environment

Page 23 of 26

Environmental Science & Technology

482

granted by the TUM Graduate School. We further thank Christoph Falter, Valentin Batteiger,

483

Andreas Sizmann and Oliver Boegler for many fruitful discussions.

484 485

ASSOCIATED CONTENT

486

S1 Description of the temperature model, S2 Supplementary temperature simulations and

487

related content. This information is available free of charge via the Internet at

488

http://pubs.acs.org/.

489 490

REFERENCES

491

(1)

Chisti, Y. Biodiesel from microalgae. Biotechnol. Adv. 2007, 25, 294–306.

492 493

(2)

Richmond (Ed.), A. Handbook of microalgal culture: biotechnology and applied phycology; Blackwell Science Ltd: Oxford, Ames, Carlton, 2004.

494 495 496

(3)

Mutanda, T.; Ramesh, D.; Karthikeyan, S.; Kumari, S.; Anandraj, a.; Bux, F. Bioprospecting for hyper-lipid producing microalgal strains for sustainable biofuel production. Bioresour. Technol. 2011, 102, 57–70.

497 498 499

(4)

Scott, S. a; Davey, M. P.; Dennis, J. S.; Horst, I.; Howe, C. J.; Lea-Smith, D. J.; Smith, A. G. Biodiesel from algae: challenges and prospects. Curr. Opin. Biotechnol. 2010, 21, 277–286.

500 501

(5)

Wijffels, R. H.; Barbosa, M. J. An outlook on microalgal biofuels. Science 2010, 329, 796–799.

502 503

(6)

Pienkos, P. T.; Darzins, A. The promise and challenges of microalgal-derived biofuels. Biofuels, Bioprod. Biorefining 2009, 3, 431–440.

504 505 506

(7)

Schenk, P. M.; Thomas-Hall, S. R.; Stephens, E.; Marx, U. C.; Mussgnug, J. H.; Posten, C.; Kruse, O.; Hankamer, B. Second generation biofuels: high-efficiency microalgae for biodiesel production. BioEnergy Res. 2008, 1, 20–43.

507 508

(8)

Pittman, J. K.; Dean, A. P.; Osundeko, O. The potential of sustainable algal biofuel production using wastewater resources. Bioresour. Technol. 2011, 102, 17–25.

509 510

(9)

Guedes, A. C.; Amaro, H. M.; Malcata, F. X. Microalgae as sources of high addedvalue compounds - a brief review of recent work. Biotechnol. Prog. 2011, 27, 597–613.

511 512 513

(10)

Harun, R.; Singh, M.; Forde, G. M.; Danquah, M. K. Bioprocess engineering of microalgae to produce a variety of consumer products. Renew. Sustain. Energy Rev. 2010, 14, 1037–1047. 23 ACS Paragon Plus Environment

Environmental Science & Technology

Page 24 of 26

514 515

(11)

Norsker, N.; Barbosa, M. J.; Vermuë, M. H.; Wijffels, R. H. Microalgal production — A close look at the economics. Biotechnol. Adv. 2011, 29, 24–27.

516 517

(12)

Clark, J. H.; Luque, R.; Matharu, A. S. Green chemistry, biofuels, and biorefinery. Annu. Rev. Chem. Biomol. Eng. 2012, 3, 183–207.

518 519

(13)

Borowitzka, M. A.; Moheimani, N. R. Algae for biofuels and energy; Borowitzka, M. A.; Moheimani, N. R., Eds.; Springer Netherlands: Dordrecht, 2013.

520 521 522

(14)

Olivieri, G.; Salatino, P.; Marzocchella, A. Advances in photobioreactors for intensive microalgal production: configurations, operating strategies and applications. J. Chem. Technol. Biotechnol. 2014, 89, 178–195.

523 524

(15)

Bernard, O.; Rémond, B. Validation of a simple model accounting for light and temperature effect on microalgal growth. Bioresour. Technol. 2012, 123, 520–527.

525 526

(16)

Goldman, J. C. Outdoor algal mass cultures - II. Photosynthetic yield limitations. Water Res. 1979, 13, 119–136.

527 528

(17)

Mata, T. M.; Martins, A. a.; Caetano, N. S. Microalgae for biodiesel production and other applications: a review. Renew. Sustain. Energy Rev. 2010, 14, 217–232.

529 530 531

(18)

Béchet, Q.; Shilton, A.; Fringer, O. B.; Muñoz, R.; Guieysse, B. Mechanistic modeling of broth temperature in outdoor photobioreactors. Environ. Sci. Technol. 2010, 44, 2197–2203.

532 533

(19)

Béchet, Q.; Shilton, A.; Guieysse, B. Full-scale validation of a model of algal productivity. Environ. Sci. Technol. 2014, 48, 13826–13833.

534 535 536

(20)

Béchet, Q.; Shilton, A.; Park, J. B. K.; Craggs, R. J.; Guieysse, B. Universal temperature model for shallow algal ponds provides improved accuracy. Environ. Sci. Technol. 2011, 45, 3702–3709.

537 538 539 540

(21)

Zhang, K.; Kurano, N.; Miyachi, S. Outdoor culture of a cyanobacterium with a vertical flat-plate photobioreactor: effects on productivity of the reactor orientation, distance setting between the plates, and culture temperature. Appl. Microbiol. Biotechnol. 1999, 52, 781–786.

541 542

(22)

Goetz, V.; Le Borgne, F.; Pruvost, J.; Plantard, G.; Legrand, J. A generic temperature model for solar photobioreactors. Chem. Eng. J. 2011, 175, 443–449.

543 544 545

(23)

Pereira, D. a; Rodrigues, V. O.; Gómez, S. V; Sales, E. a; Jorquera, O. Parametric sensitivity analysis for temperature control in outdoor photobioreactors. Bioresour. Technol. 2013, 144, 548–553.

546 547 548

(24)

Sierra, E.; Acién, F. G.; Fernández, J. M.; García, J. L.; González, C.; Molina, E. Characterization of a flat plate photobioreactor for the production of microalgae. Chem. Eng. J. 2008, 138, 136–147.

24 ACS Paragon Plus Environment

Page 25 of 26

Environmental Science & Technology

549 550 551

(25)

Rodolfi, L.; Chini Zittelli, G.; Bassi, N.; Padovani, G.; Biondi, N.; Bonini, G.; Tredici, M. R. Microalgae for oil: strain selection, induction of lipid synthesis and outdoor mass cultivation in a low-cost photobioreactor. Biotechnol. Bioeng. 2009, 102, 100–112.

552 553

(26)

Richmond, A.; Cheng-Wu, Z. Optimization of a flat plate glass reactor for mass production of Nannochloropsis sp. outdoors. J. Biotechnol. 2001, 85, 259–269.

554

(27)

Algenol LLC http://www.algenol.com/ (accessed Jul 29, 2015).

555

(28)

Subitec GmbH http://subitec.com/de (accessed Jul 29, 2015).

556 557 558

(29)

National Renewable Energy Laboratory (NREL). National solar radiation data base (1991 -2005 update: typical meteorological year 3) http://rredc.nrel.gov/solar/old_data/nsrdb/ (accessed Mar 26, 2015).

559 560

(30)

NIST Standard Reference Database Number 69. In NIST Chemistry WebBook; Linstrom, P. J.; Mallard, W. G., Eds.; National Institute of Standards and Technology.

561 562

(31)

Stine, W. B.; Geyer, M. Power from the sun; published online at http://www.powerfromthesun.net/book.html, 2001.

563 564

(32)

Slegers, P. M.; Wijffels, R. H.; van Straten, G.; van Boxtel, A. J. B. Design scenarios for flat panel photobioreactors. Appl. Energy 2011, 88, 3342–3353.

565 566

(33)

VDI-Wärmeatlas; Kabelac, S.; Kind, M.; Martin, H.; Mewes, D.; Schaber, K.; Stephan, P., Eds.; 11th ed.; Springer Berlin Heidelberg: Berlin, Heidelberg, 2013.

567 568

(34)

Gates, D. M.; Keegan, H. J.; Schleter, J. C.; Weidner, V. R. Spectral properties of plants. Appl. Opt. 1965, 4, 11.

569 570

(35)

Brutsaert, W. On a derivable formula for long-wave radiation from clear skies. Water Resour. Res. 1975, 11, 742–744.

571 572 573

(36)

Crawford, T. M.; Duchon, C. E. An improved parameterization for estimating effective atmospheric emissivity for use in calculating daytime downwelling longwave radiation. J. Appl. Meteorol. 1999, 38, 474–480.

574 575

(37)

Schmidt, W. L.; Gosnold, W. D.; Enz, J. W. A decade of air – ground temperature exchange from Fargo, North Dakota. Glob. Planet. Change 2001, 29, 311–325.

576 577 578 579

(38)

Henning, A.; Limberg, A. Veränderung des oberflächennahen Temperaturfeldes von Berlin durch Klimawandel und Urbanisierung - Variation of the subsoil temperature field in Berlin as a result of climate change and urbanization. Brand. Geowissenschaftliche Beiträge 2012, 19, 81–92.

580 581

(39)

Peel, M. C.; Finlayson, B. L.; McMahon, T. A. Updated world map of the KöppenGeiger climate classification. Hydrol. Earth Syst. Sci. 2007, 11, 1633–1644.

582 583

(40)

Weyer, K. M.; Bush, D. R.; Darzins, A.; Willson, B. D. Theoretical maximum algal oil production. BioEnergy Res. 2009, 3, 204–213. 25 ACS Paragon Plus Environment

Environmental Science & Technology

Page 26 of 26

584 585 586

(41)

Apel, A. C.; Weuster-Botz, D. Engineering solutions for open microalgae mass cultivation and realistic indoor simulation of outdoor environments. Bioprocess Biosyst. Eng. 2015, 38, 995–1008.

587

(42)

Tamiya, H. Mass culture of algae. Annu. Rev. Plant Physiol. 1957, 8, 309–334.

588 589 590

(43)

Manjre, S. D.; Deodhar, M. a. Screening of thermotolerant microalgal species isolated from western ghats of Maharashtra, India for CO2 sequestration. J. Sustain. Energy Environ. 2013, 4, 61–67.

591

26 ACS Paragon Plus Environment