Microalgae Recovery from Water for Biofuel ... - ACS Publications

Jun 17, 2016 - ABSTRACT: There is a pressing need to develop efficient and sustainable approaches to harvesting microalgae for biofuel production and ...
0 downloads 0 Views 700KB Size
Subscriber access provided by UNIV OF CAMBRIDGE

Article

Microalgae Recovery from Water for Biofuel Production Using CO2-Switchable Crystalline Nanocellulose Shijian Ge, Pascale Champagne, Hai-Dong Wang, Philip G. Jessop, and Michael F. Cunningham Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.6b00732 • Publication Date (Web): 17 Jun 2016 Downloaded from http://pubs.acs.org on June 17, 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 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14

Environmental Science & Technology

Microalgae Recovery from Water for Biofuel Production Using CO2-Switchable Crystalline Nanocellulose Shijian Ge a, Pascale Champagne a,b,*, Haidong Wang b, Philip G. Jessop,c Michael F. Cunningham b a

Department of Civil Engineering, Queen’s University, 58 University Avenue, Kingston, Ontario K7L 3N6, Canada b Department of Chemical Engineering, Queen’s University, 19 Division Street, Kingston, Ontario K7L 3N6, Canada c Department of Chemistry, Queen’s University, 90 Bader Lane, Kingston, Ontario K7L 3N6, Canada *Corresponding author: E-mail: [email protected]; Phone: (613)533-3053; Fax: (613)533-2128.

1

ACS Paragon Plus Environment

Environmental Science & Technology

15

Abstract: There is a pressing need to develop efficient and sustainable approaches to harvesting

16

microalgae for biofuel production and water treatment. CO2-switchable crystalline nanocellulose

17

(CNC) modified with 1-(3-aminopropyl)imidazole (APIm) is proposed as a reversible coagulant

18

for harvesting microalgae. Compared to native CNC, the positively charged APIm-modified

19

CNC, which dispersed well in carbonated water, showed appreciable electrostatic interaction

20

with negatively charged Chlorella vulgaris upon CO2-treatment. The gelation between the

21

modified CNC, triggered by subsequent air sparging, can also enmesh adjacent microalgae

22

and/or microalgae-modified CNC aggregates, thereby further enhancing harvesting efficiencies.

23

Moreover, the surface charges and dispersion/gelation of APIm-modified CNC could be

24

reversibly adjusted by alternatively sparging CO2/air. This CO2-switchability would make the

25

reusability of re-dispersed CNC for further harvesting possible. After harvesting, the supernatant

26

following sedimentation can be reused for microalgal cultivation without detrimental effects on

27

cell growth. The use of this approach for harvesting microalgae presents an advantage to other

28

current methods available because all materials involved, including the cellulose, CO2 and air,

29

are natural and biocompatible without adverse effects on the downstream processing for biofuel

30

production.

31 32

Keywords: crystalline nanocellulose, microalgae, coagulation, carbon dioxide, biofuel,

33

Chlorella vulgaris

34

2

ACS Paragon Plus Environment

Page 2 of 28

Page 3 of 28

Environmental Science & Technology

35

Introduction

36

Microalgae are a promising alternative third generation feedstock for the biofuel production

37

industry due to their higher photosynthetic efficiency and lipid contents (15-77 % of cell mass) 1-

38

5

39

problematic in water systems due to their rapid increase or accumulation during algal blooms.7-9

40

Thus, efficient microalgal harvesting or removal from water is not only critical for biofuel

41

production, but also important in the mitigation of aquatic systems.

42

than common feedstocks such as crops.6 Additionally, microalgae have been shown to be

Microalgal harvesting or separation from water still represents a major technological and

43

economic barrier for both the microalgae-based biofuel and water treatment industries.

44

Conventional microalgal separation methods include gravity, precipitation, centrifugation,

45

microstraining, flotation and filtration.10 These methods are often energy- and/or time-consuming

46

and thus undesirable for both the low-cost production of microalgal biofuels and full-scale water

47

treatment.11 More recently, the application of chemical flocculants such as cationic inorganics or

48

polymers,12 magnetophoretic separation using native or cationic polymer coated magnetic

49

nanoparticles (NPs) 13, 14 and bio-flocculation induced by microbes (e.g. bacteria, fungi) 10, 15

50

have been reported. However, these methods involve the use of additives, which have the

51

potential to contaminate and have adverse effects on both microalgal cells and culture media

52

when using microalgae as either inocula and/or recycling supernatant after harvesting. As such,

53

approaches that are sustainable and feasible for large-scale applications are still being

54

investigated, and the challenge will lie in the development of technologies that can facilitate the

55

separation of microalgae in a technically, environmentally and economically viable manner.

3

ACS Paragon Plus Environment

Environmental Science & Technology

56

Cellulose is the most abundant natural and renewable organic polymer on Earth,16 and it is

57

regarded as an almost infinite source of raw material.17 Moreover, crystalline nanocellulose

58

(CNC), derived from the acid hydrolysis of cellulose fibers, has attracted significant interest

59

from both researchers and engineers due to its environmentally benign nature (biodegradability)

60

and physicochemical properties such as nanoscale dimensions, high specific surface area and

61

unique optical properties.18 In general, CNC properties can be manipulated for a variety of

62

purposes through modification of the hydroxyl groups on the CNC surface.19 Kan et al 20

63

proposed a pH-responsive P4VP-g-CNCs grafted using the surface-initiated polymerization of 4-

64

vinylpyridine (P4VP) with a ceric(IV) ammonium nitrate initiator, which showed reversible

65

flocculation and sedimentation properties with changes in pH. Recently, Vandamme et al.21, 22

66

demonstrated the applicability of CNC functionalized with cationic pyridinium and imidazole

67

groups for microalgal flocculation, and also noted that in contrast to conventional polymer

68

flocculants, the flocculation efficiency of cationic CNC was relatively unaffected by algal

69

organic matter. In their studies, however, the recovery and reuse of the CNC and the culture

70

medium were not investigated. Such recycling could be an important consideration in the

71

development of a sustainable and economic technology. Moreover, further optimization of CNC

72

dosage requirement is essential to minimize the cost of the CNC.

73

To allow for the recovery and reuse of flocculants or coagulants, detachment of the

74

flocculants or coagulants from the microalgae must be achieved, for example by inducing

75

changes in their surface properties, such as surface charge and wetting properties.13, 23-25 Such

76

changes may be triggered by the presence of switchable or stimuli-responsive groups on the

77

surface of the flocculants or coagulants. We recently reported a new CNC which is a CO2-

78

switchable nanomaterial prepared by surface modification with 1-(3-aminopropyl)imidazole 4

ACS Paragon Plus Environment

Page 4 of 28

Page 5 of 28

Environmental Science & Technology

79

(APIm).26 Such APIm-modified CNC was positively charged in the presence of CO2 resulting

80

from the protonation of the APIm groups by the carbonated water (Equation 1), which can

81

probably promote the coagulation or attachment of microalgae cells carrying negative charges.21

82

Additionally, the chemically bonded imidazole groups on the CNC surface can respond to the

83

CO2 stimulus in an effective and repeatable manner. Specifically, the APIm-modified CNC

84

disperses well in water in the presence of CO2, while subsequent removal of CO2 through

85

sparging of the dispersion with N2 gives rise to the formation of aggregates. This

86

dispersion/aggregation cycle can be performed repeatedly by alternating treatments with CO2

87

and N2, which could allow for the recovery of the CNC, potentially resulting in a decrease in the

88

economic and ecological cost of the microalgal harvesting process.

(1)

89

90

In this study, the APIm-modifed CNC was synthesized as previously reported 26. The

91

following questions were then addressed: (1) does surface modification of the CNC with APIm

92

create a CO2-switchable surface charge and reversible size changes in contrast to native CNC; (2)

93

can the APIm-modified CNC be used to harvest a model biofuel-producing microalgal species,

94

Chlorella vulgaris (C. vulgaris) with a reduced dose demand and harvesting time compared to

95

the native CNC; (3) can the process be modified to improve the harvesting efficiency, for

96

example by changing the method or flow rate of inert gas introduction; (4) are the colloidal

97

interactions between CNC particles and microalgae consistent with the

98

Derjaguin−Landau−Verwey−Overbeek (DLVO) theory; and (5) can the culture medium, as well 5

ACS Paragon Plus Environment

Environmental Science & Technology

99

as the harvested and concentrated microalgae-CNC aggregates, be recycled to decrease

100

operational costs and increase the sustainability of this CNC-based microalgal separation process?

101

Materials and methods

102

Microalgal culture. A 25 L glass carboy was used to grow C. vulgaris in modified Bold's

103

Basal Medium (MBBM) at room temperature (23.0 ± 0.5°C).10 The MBBM contained the major

104

ions of Na+, K+, Mg2+, Ca2+, Fe2+, Zn2+, Mn2+, Cu2+, Co2+, H+, NO3- , H2PO4- , HPO42-, BO33-,

105

SO42-, Cl-, OH-, MoO42-, and EDTA2- with a total ionic strength of 10.4 mM. The molarity of

106

each ion is listed in Table S1. The initial solution pH, dissolved oxygen, and oxidation reduction

107

potential of the culture medium were 6.8 ± 1.0, 12 ± 2 mg·L-1, and 170 ± 31 mV, respectively.

108

The system was aerated with filtered ambient air (0.039 % CO2) at a rate of 200 mL·min-1.

109

Continuous irradiation (27.4 µmoles·m-2·s-1) was applied. During cultivation, samples were

110

collected to measure optical density at 680 nm (OD680) using a spectrophotometer (Hach Method

111

8171).27 Biomass concentration (g·L-1) was gravimetrically quantified by dry cell weights

112

(DCW), which was performed by drying 0.45 µm membrane-filtered microalgae in an oven at

113

105 oC to a constant weight.10

114

Preparation and characterization of APIm-modified CNC. The native CNC was

115

provided by FPInnovations, Canada. The APIm-modified CNC was synthesized as previously

116

reported by our group, which used 1,10-carbonyldiimidazole (CDI, reagent grade, Sigma-Aldrich)

117

and 1-(3-aminopropyl)imidazole (APIm, ≥ 97 %, Sigma-Aldrich).26 Zeta potentials, average

118

hydrodynamic diameters, and particle size distributions (PSD) of both native and APIm-modified

119

CNC suspended in water, under the various experimental conditions mentioned below, were

120

characterized at 25 oC by dynamic light scattering (DLS) on a Zetasizer Nano ZS instrument 6

ACS Paragon Plus Environment

Page 6 of 28

Page 7 of 28

Environmental Science & Technology

121

(Malvern Instruments, UK) using DTS 160 disposable folded capillary cells. Refractive indices

122

of 1.500 and 1.347 were used, respectively, for APIm-modified CNC and C. vulgaris cells for

123

calculating the scattering wave vector.

124

Microalgae separation with APIm-modified CNC. Harvesting experiments began once C.

125

vulgaris had reached their exponential growth stage. The APIm-modified CNC stock dispersion

126

in water (~10.5 g·L-1, 35 mL) was vortexed (3 times at 3000 rpm for 2 min each, with 30 s

127

intervals), sparged with CO2 (99.995%, MEGS) for 10 min, and briefly centrifuged (15,000 ×g,

128

20oC, 1 min) to remove floating particles. Then the supernatant containing well-dispersed APIm-

129

modified CNC was mixed into the 20 mL suspension of C. vulgaris in a 50 mL glass specimen

130

bottle or falcon tube at room temperature with an initial microalgal concentration of 0.2-0.4 g·L-1.

131

Afterwards, CO2 was sparged into the suspension for 1 min. After sparging with air (lab air

132

system) or N2 (99.9999%, MEGS) gas for a specified time (5-10 min), microalgae-CNC

133

aggregates were allowed to flocculate and settle for 10 min. The gas sparging was conducted

134

under 1 atm of CO2, N2 or air. Finally, liquid samples were taken from 1.0 cm below the surface

135

of the microalgal suspension for an optical absorbance measurement at 680 nm using a Hach

136

Method 8171 spectrophotometer. Three indicators including harvesting efficiency (HE) for the

137

evaluation of the efficiency of proposed harvesting technique, recovery efficiency (RE) for

138

efficiency of the coagulant (native or modified CNC), and recovery capacity (RC) for the

139

harvesting performance as attributed to microalgal quantities (gram algae) per gram CNC, were

140

used to evaluate microalgal separation performance:28, 29

141

HE (%)=[1-(C t /C0 )]×100%

(2)

7

ACS Paragon Plus Environment

Environmental Science & Technology

Page 8 of 28

143

C t C0 )]×100% C't C'0 ( C − Ct )V RC (g-algae ⋅ g-CNC-1 ) = 0 m

144

where C0 and Ct are the microalgal concentrations in the supernatant before and after separation

145

(g·L-1), C0’ and Ct’ are the microalgal concentrations without addition of native or APIm-

146

modified CNC in the control group (g·L-1), V is the volume (20 mL) of microalgal suspension

147

and m is the mass of APIm-modified CNC added (g). APIm-modified CNC concentrations were

148

gravimetrically quantified by dry cell weights (g·L-1). Microalgal separation was studied by

149

varying the APIm-modified CNC dosage, non-acidic gas (pure N2 or air) addition, and air

150

sparging time as described below. All experiments below were performed in duplicate or

151

triplicate.

152

Effect of CNC surface modification. Different quantities of native or APIm-modified CNC were

153

added to the C. vulgaris suspensions at 0.2~0.4 g-DCW·L-1 to achieve different mass ratios of

154

coagulant to microalgal biomass (0.01, 0.02, 0.05, 0.1, 0.2, 0.3, 0.4, and 0.5 g-CNC·g-algae-1).

155

All doses were calculated based on the dry weights of both microalgae and CNC. The mixtures

156

were then sparged with CO2 for 1 min followed by air alone at a gas flow rate of 140 mL·min-1

157

for 10 min.

158

Effect of inert gas and flow rates. The sparging of nitrogen-containing gas (either pure N2 gas or

159

air) was used to flush CO2 from the aqueous phase. These two nitrogen-containing gases (N2 or

160

air) were compared to evaluate their effectiveness for microalgal separation. In addition, three

161

flow rates (25, 80 and 140 mL·min-1) were applied. The separation experiments were performed

162

with an optimized dose (0.05 g-CNC·g-algae-1) of APIm-modified CNC determined above.

142

(3)

RE (%)=[1-(

(4)

8

ACS Paragon Plus Environment

Page 9 of 28

Environmental Science & Technology

163

Effect of air sparging time. After 1 min of CO2 sparging, different air sparging times (0, 1, 3, 5,

164

7, 10, 13, 20 min) were investigated to optimize the sparging time required for coagulation.

165

These were performed at two APIm-modified CNC doses (0.05 and 0.49 g-CNC·g-algae-1) for

166

comparison.

167

Effect of pH adjustment. Different pH cycles (4.9/7.9, 4.6/7.8 and 4.2/7.2) were artificially

168

generated through the addition of 1 M HCl and 1 M NaOH to mimic the CO2/air-treated samples

169

for three APIm-modified CNC doses (0.05, 0.29 and 0.49 g-modified CNC.g-1-algae). In another

170

pH adjustment experiment, the use of HCl/air treatment was compared. A control experiment

171

having only a pH adjustment of 4.6/7.8 (with 1 M HCl and 1 M NaOH), but without any APIm-

172

modified CNC was also tested.

173

Assessment of supernatant and APIm-modified CNC reuse. After harvesting the

174

microalgae with APIm-modified CNC, concentrations of nitrate and phosphorus in the used

175

medium were adjusted to the levels in the MBBM. The medium was then reused to cultivate C.

176

vulgaris. For comparison, another two microalgal culture supernatants were used, which were

177

obtained after harvesting by either centrifugation (as a control) or coagulation using alum (20

178

mg·L-1). Both recycled supernatants were then reused for microalgal cultivation. In all three

179

cases, the biomass growth in the recycled medium was monitored for 7 days in 250 mL flasks.

180

Cultivation conditions were previously described.30

181

To determine the recyclability of APIm-modified CNC (0.49 g-modified CNC.g-1-algae), the

182

collected microalgae-CNC aggregates (~2.0-2.5 g-algae·L-1 in ~3 mL) were reused to harvest

183

new microalgae batches following the same harvesting procedures as noted above. The process

184

was repeated for five cycles to further assess the recyclability. 9

ACS Paragon Plus Environment

Environmental Science & Technology

Page 10 of 28

185

Comparison of colloidal interaction. Quantitative information on the nonspecific interactive

186

forces between CNC particles and microalgal cells can be directly obtained with Ohshima’s soft-

187

particle DLVO theory assuming that Lifshitz-van der Waals and electrostatic forces are the

188

dominant forces.31, 32 The computation methods for the van der Waals and electrostatic forces

189

vary with the geometry of the interacting entities. The diameter of spherical microalgal cells is

190

approximately 2-5 µm, 33 and the CNCs are usually 100–300 nm in length and 10-20 nm in

191

width. In this study, the CNC was assumed to be spherical to simplify the analysis to sphere-to-

192

sphere geometry, as was employed in other studies and to allow for comparison of the interaction

193

between microalgae and flocculants such as metal oxide NPs, or bacteria. 10, 29, 34, 35 The retarded

194

Lifshitz-van der Waals and the electrostatic interaction energy (the linearized version of the

195

Poisson−Boltzmann expression) for sphere-to-sphere geometry were calculated as per Equations

196

(5) and (6) when h95 % RE was achieved with

449

each cycle, indicating that following CO2 sparging, the re-dispersed APIm-modified CNC

450

particles have sufficient adsorption sites to coagulate additional microalgal cells. The pKa of the

451

imidazole functionalities have been reported to be in the range of 6.0-6.5.40, 41 In our previous

452

study,26 it was calculated that up to 94 % of imidazole rings could be protonated upon exposure

453

to CO2, while sparging N2 reduced this value to 26%. These results imply that over the pH range

454

obtained with CO2/air sparging, the imidazole groups on the APIm-modified CNC can switch

455

from being almost fully protonated to only partially protonated, which directly influences the

456

attachment or detachment of CNC and microalgae, and the ability to reuse the APIm-modified

457

CNC.

458

Such CO2-switchable APIm-modified CNC has been demonstrated to be greener than the

459

recyclable polyampholytic flocculants which require the addition of acid and base,25 is likely to

460

be more energy-efficient than the magnetic Fe3O4-ZnO nanocomposites which require UV

461

irradiation,23 and more beneficial for the microalgae harvesting process than single-use

462

commercial flocculants.42 As such, this recyclable CO2-switchable APIm-modified CNC has the

463

potential to provide a sustainable solution to microalgal harvesting and cultivation. Moreover, 23

ACS Paragon Plus Environment

Environmental Science & Technology

464

the saturated microalgae-modified CNC aggregates have the potential for use in biofuel

465

production through subsequent anaerobic digestion, hydrothermal liquefaction or other

466

conversion technologies. To some extent, such recycling and biofuel conversion could

467

compensate for the use of more expensive CNC materials as coagulants, requiring considerably

468

higher initial capital investment, which would otherwise hamper their application in industry. A

469

life cycle assessment (LCA) of such an integrated process would also be highly valuable in

470

assessing its water, energy and environmental footprint, as well as its techno-economic viability.

471

Supporting Information

472

Additional information includes Table S1-S8, and Figures S1−S5. This material is available free

473

of charge via the Internet at http://pubs.acs.org.

474

Acknowledgements

475

The authors thank the Ontario Ministry of Research Innovation – Ontario Research Fund, the

476

National Science and Engineering Research Council (NSERC), the Canada Research Chairs

477

program (PC, PGJ) and the Ontario Research Chairs program (MFC).

24

ACS Paragon Plus Environment

Page 24 of 28

Page 25 of 28

Environmental Science & Technology

478

References

479 480 481 482 483 484 485 486 487 488 489 490 491 492 493 494 495 496 497 498 499 500 501 502 503 504 505 506 507 508 509 510 511 512 513 514 515 516 517 518 519 520 521

1. Milledge, J.; Heaven, S., A review of the harvesting of micro-algae for biofuel production. Rev Environ Sci Biotechnol 2013, 12, 165-178. 2. Suali, E.; Sarbatly, R., Conversion of microalgae to biofuel. Renew Sust Energ Rev 2012, 16, 43164342. 3. Chisti, Y., Biodiesel from microalgae. Biotechnol Adv 2007, 25, 294-306. 4. Del Río, E.; Armendáriz, A.; García-Gómez, E.; García-González, M.; Guerrero, M. G., Continuous culture methodology for the screening of microalgae for oil. J Biotechnol 2015, 195, 103-107. 5. Allen, J. W.; DiRusso, C. C.; Black, P. N., Triacylglycerol synthesis during nitrogen stress involves the prokaryotic lipid synthesis pathway and acyl chain remodeling in the microalgae Coccomyxa subellipsoidea. Algal Res 2015, 10, 110-120. 6. Sharma, Y. C.; Singh, B.; Korstad, J., A critical review on recent methods used for economically viable and eco-friendly development of microalgae as a potential feedstock for synthesis of biodiesel. Green Chem 2011, 13, 2993-3006. 7. Lee, S. O.; Kim, S.; Kim, M.; Lim, K. J.; Jung, Y., The Effect of Hydraulic Characteristics on Algal Bloom in an Artificial Seawater Canal: A Case Study in Songdo City, South Korea. Water 2014, 6, 399-413. 8. Lou, X.; Hu, C., Diurnal changes of a harmful algal bloom in the East China Sea: Observations from GOCI. Remote Sens Environ 2014, 140, 562-572. 9. Jiang, Q.; Jie, Y.; Han, Y.; Gao, C.; Zhu, H.; Willander, M.; Zhang, X.; Cao, X., Self-powered electrochemical water treatment system for sterilization and algae removal using water wave energy. Nano Energy 2015, 18, 81-88. 10. Agbakpe, M.; Ge, S.; Zhang, W.; Zhang, X.; Kobylarz, P., Algae harvesting for biofuel production: Influences of UV irradiation and polyethylenimine (PEI) coating on bacterial biocoagulation. Bioresour Technol 2014, 166, 266–272. 11. Ahmad, A. L.; Yasin, N. H. M.; Derek, C. J. C.; Lim, J. K., Comparison of harvesting methods for microalgae Chlorella sp. and its potential use as a biodiesel feedstock. Environ Technol 2014, 35, 2244– 2253. 12. Rashid, N.; Rehman, S. U.; Han, J.-I., Rapid harvesting of freshwater microalgae using chitosan. Process Biochem 2013, 48, 1107-1110. 13. Ge, S.; Agbakpe, M.; Zhang, W.; Kuang, L., Heteroaggregation between PEI-Coated Magnetic Nanoparticles and Algae: Effect of Particle Size on Algal Harvesting Efficiency. ACS Appl. Mater. Interfaces 2015, 7, 6102-6108. 14. Chiang, Y.-D.; Dutta, S.; Chen, C.-T.; Huang, Y.-T.; Lin, K.-S.; Wu, J. C. S.; Suzuki, N.; Yamauchi, Y.; Wu, K. C. W., Functionalized Fe3O4@Silica Core–Shell Nanoparticles as Microalgae Harvester and Catalyst for Biodiesel Production. ChemSusChem 2015, 8, 789-794. 15. Taylor, R. L.; Rand, J. D.; Caldwell, G. S., Treatment with algae extracts promotes flocculation, and enhances growth and neutral lipid content in Nannochloropsis oculata—a candidate for biofuel production. Mar Biotechnol 2012, 14, 774-781. 16. Klemm, D.; Heublein, B.; Fink, H. P.; Bohn, A., Cellulose: fascinating biopolymer and sustainable raw material. Angew Chem Int Ed Engl 2005, 44, 3358-3393. 17. Brinchi, L.; Cotana, F.; Fortunati, E.; Kenny, J. M., Production of nanocrystalline cellulose from lignocellulosic biomass: Technology and applications. Carbohyd Polym 2013, 94, 154-169. 18. Peng, B.; Dhar, N.; Liu, H.; Tam, K., Chemistry and applications of nanocrystalline cellulose and its derivatives: a nanotechnology perspective. Can J Chem Eng 2011, 89, 1191-1206. 25

ACS Paragon Plus Environment

Environmental Science & Technology

522 523 524 525 526 527 528 529 530 531 532 533 534 535 536 537 538 539 540 541 542 543 544 545 546 547 548 549 550 551 552 553 554 555 556 557 558 559 560 561 562 563 564 565 566 567

Page 26 of 28

19. Habibi, Y.; Lucia, L. A.; Rojas, O. J., Cellulose nanocrystals: chemistry, self-assembly, and applications. Chem Rev 2010, 110, 3479-3500. 20. Kan, K. H. M.; Li, J.; Wijesekera, K.; Cranston, E. D., Polymer-Grafted Cellulose Nanocrystals as pH-Responsive Reversible Flocculants. Biomacromolecules 2013, 14, 3130-3139. 21. Vandamme, D.; Eyley, S.; Van den Mooter, G.; Muylaert, K.; Thielemans, W., Highly charged cellulose-based nanocrystals as flocculants for harvesting Chlorella vulgaris. Bioresour Technol 2015, 194, 270-275. 22. Eyley, S.; Vandamme, D.; Lama, S.; Van den Mooter, G.; Muylaert, K.; Thielemans, W., CO 2 controlled flocculation of microalgae using pH responsive cellulose nanocrystals. Nanoscale 2015, 7, 14413-14421. 23. Ge, S.; Agbakpe, M.; Zhang, W.; Kuang, L.; Wu, Z.; Wang, X., Recovering Magnetic Fe3O4-ZnO Nanocomposites from Algal Biomass Based on Hydrophobicity Shift under UV Irradiation. ACS applied materials & interfaces 2015, 7, 11677-11682. 24. Prochazkova, G.; Podolova, N.; Safarik, I.; Zachleder, V.; Branyik, T., Physicochemical approach to freshwater microalgae harvesting with magnetic particles. Colloid Surface B 2013, 112, 213-218. 25. Morrissey, K. L.; Keirn, M. I.; Inaba, Y.; Denham, A. J.; Henry, G. J.; Vogler, B. W.; Posewitz, M. C.; Stoykovich, M. P., Recyclable polyampholyte flocculants for the cost-effective dewatering of microalgae and cyanobacteria. Algal Res 2015, 11, 304-312. 26. Wang, H. D.; Jessop, P.; Bouchard, J.; Champagne, P.; Cunningham, M., Cellulose nanocrystals with CO2-switchable aggregation and redispersion properties. Cellulose 2015, 22, 3105-3116. 27. Arbib, Z.; Ruiz, J.; Álvarez-Díaz, P.; Garrido-Pérez, C.; Perales, J. A., Capability of different microalgae species for phytoremediation processes: Wastewater tertiary treatment, CO2 bio-fixation and low cost biofuels production. Water Res 2014, 49, 465-474. 28. Hu, Y.-R.; Wang, F.; Wang, S.-K.; Liu, C.-Z.; Guo, C., Efficient Harvesting of Marine Microalgae Nannochloropsis maritima using Magnetic Nanoparticles. Bioresour. Technol. 2013, 138, 387-390. 29. Ge, S.; Agbakpe, M.; Wu, Z.; Kuang, L.; Zhang, W.; Wang, X., Influences of Surface Coating, UV Irradiation and Magnetic Field on the Algae Removal Using Magnetite Nanoparticles. Environ Sci Technol 2015, 49, 1190-1196. 30. Ge, S.; Champagne, P., Nutrient removal, microalgal biomass growth, harvesting and lipid yield in response to centrate wastewater loadings. Water Res 2016, 88, 604-612. 31. Ohshima, H., Electrophoresis of soft particles. Advances in colloid and interface science 1995, 62, 189-235. 32. Hayashi, H.; Tsuneda, S.; Hirata, A.; Sasaki, H., Soft particle analysis of bacterial cells and its interpretation of cell adhesion behaviors in terms of DLVO theory. Colloids and Surfaces B: Biointerfaces 2001, 22, 149-157. 33. de-Bashan, L. E.; Bashan, Y.; Moreno, M.; Lebsky, V. K.; Bustillos, J. J., Increased pigment and lipid content, lipid variety, and cell and population size of the microalgae Chlorella spp. when coimmobilized in alginate beads with the microalgae-growth-promoting bacterium Azospirillum brasilense. Can J Microbiol 2002, 48, 514-521. 34. Toh, P. Y.; Ng, B. W.; Ahmad, A. L.; Chieh, D. C. J.; Lim, J., The role of particle-to-cell interactions in dictating nanoparticle aided magnetophoretic separation of microalgal cells. Nanoscale 2014, 7, 12838-12848. 35. Ma, S.; Zhou, K.; Yang, K.; Lin, D., Heteroagglomeration of Oxide Nanoparticles with Algal Cells: Effects of Particle Type, Ionic Strength and pH. Environ Sci Technol 2015, 49, 932-939. 36. Schenkel, J.; Kitchener, J., A test of the Derjaguin-Verwey-Overbeek theory with a colloidal suspension. Transactions of the Faraday Society 1960, 56, 161-173. 26

ACS Paragon Plus Environment

Page 27 of 28

568 569 570 571 572 573 574 575 576 577 578 579 580 581

Environmental Science & Technology

37. Hadjoudja, S.; Deluchat, V.; Baudu, M., Cell surface characterisation of Microcystis aeruginosa and Chlorella vulgaris. J Colloid Interf Sci 2010, 342, 293-299. 38. Wang, N.; Hsu, C.; Zhu, L.; Tseng, S.; Hsu, J.-P., Influence of metal oxide nanoparticles concentration on their zeta potential. J Colloid Interf Sci 2013, 407, 22-28. 39. Lee, K.; Lee, S. Y.; Na, J.-G.; Jeon, S. G.; Praveenkumar, R.; Kim, D.-M.; Chang, W.-S.; Oh, Y.-K., Magnetophoretic harvesting of oleaginous Chlorella sp. by using biocompatible chitosan/magnetic nanoparticle composites. Bioresour Technol 2013, 149, 575-578. 40. Kim, T.-i.; Rothmund, T.; Kissel, T.; Kim, S. W., Bioreducible polymers with cell penetrating and endosome buffering functionality for gene delivery systems. J Control Release 2011, 152, 110-119. 41. Lin, W.; Kim, D., pH-Sensitive micelles with cross-linked cores formed from polyaspartamide derivatives for drug delivery. Langmuir 2011, 27, 12090-12097. 42. Ummalyma, S. B.; Mathew, A. K.; Pandey, A.; Sukumaran, R. K., Harvesting of microalgal biomass: Efficient method for flocculation through pH modulation. Bioresour Technol 2016, http://dx.doi.org/10.1016/j.biortech.2016.03.114.

582

27

ACS Paragon Plus Environment

Environmental Science & Technology

Graphic abstract

ACS Paragon Plus Environment

Page 28 of 28