1-Hydroxyanthraquinone:

Activity in denitrificans and potential application for biomass reduction

in wastewater treatment facilities

A Thesis Submitted to the Faculty

of

Drexel University

By

Lynn Ann Arlauskas-Dekleva

In Partial Fulfillment of the Requirement

For the Degree of

Doctor of Philosophy

February 2007

ii

Dedication

This work is dedicated to the memory of my mother, Florence Arlauskas. She always felt that an education was the key to achieving your full potential and strongly supported my pursuits in both my undergraduate and graduate education. My mother and friend- this one is for you!

iii

Acknowledgments

This work is the result of a great many people who dedicated their time and effort to support my goal. Over the years, I have relied heavily on the advice and encouragement of my advisor, Dr. Michael Gealt. His patience, guidance and support enabled me to persevere and finally complete this body of work. I also wish to thank my other committee members; Dr. Charles Haas, Dr. Susan Kilham, Dr. Irene Legiec and Dr.

Raj Mutharasan for their time and valuable suggestions that helped direct and enrich this work. I appreciate the many hours you dedicated to reviewing and discussing the issues and wish to thank you for all of your suggestions.

This research is an extension of the pioneering work by Dr. Martin Odom who provided valuable guidance on the use of this compound in denitrifying systems. Martin provided priceless assistance and free access to both his laboratory and personnel. I wish to thank Martin‘s laboratory associates, Eva Nagel, JoAnn Payne, and Jolanta Wilczek who assisted in the setup and analysis of the initial laboratory investigations and welcomed me into their laboratory.

The DuPont company wastewater expert, Robert Reich, provided valuable information on industrial wastewater treatment facilities and insight into the potential and limitations associated with the implementation of the uncoupler technology in an industrial setting. His vast knowledge and hands-on experience were critical to the success of this program and I truly enjoyed the numerous and sometimes lengthy interactions. I wish to thank the head of the wastewater laboratory, Marge Ventura, and iv

her laboratory associates. They provided a wealth of information and support especially in the setup and maintenance of the wastewater reactors and gracefully looked the other way during the numerous midnight acquisitions.

I wish to thank Dr. Lisa Laffend for all her help and support. Lisa served as a member of my committee for my Candidacy exam and supported this work through its completion. Lisa‘s support as a colleague and friend was invaluable. She was always available for discussions on recent laboratory results, provided valuable guidance on the continuous reactor experiments and encouraged me to continue when things go tough.

Thanks to Lisa‘s laboratory associates, Keith Burlew and Keith Cantera, for their guidance and assistance with the setup and running of the continuous reactor.

I wish to acknowledge Dr. John Gannon, Dr. Ning Wang and Patrick Folsom for allowing free and unrestricted access to the respirometer, known affectionately as

―Ralph‖, and the laboratory facilities at the DuPont Glasgow site. The aerobic studies in this work could not have been possible without their support.

I wish to acknowledge the assistance of many individuals who were called on numerous times for their input, guidance and support. Steve Constable who provided information on biosolids production and disposal costs; Dr. Barabara Larsen and Beth

Fenner for analytical information, analysis and support; Dr. David Short and Sid Bledsoe for their help with the continuous reactor studies; Reggie Corbitt and Stan Miller for providing information on the operation of the Little Rock Wastewater Treatment facility;

Jim Hobsetter for access to information on municipal waste treatment facilities; my management, Dr. David Griffith, Dr. David Short, Dr. Julie Rakestraw, Dr. John Gannon v

and Dr. Michael Hennessey for their support of my graduate education; my first DuPont supervisor, Dr. Dale Peterson, who saw and encouraged my questioning ; and Dr.

Henn Kilkson who identified the engineer deep inside my soul. Henn‘s encouragement of my graduate education was unwavering although it initially began in the Biological

Sciences.

Last but not least, I wish to thank and acknowledge the role of my family. My husband, Mark Dekleva and children Tyler and Garrison have always supported and encouraged my educational pursuits. They stood by during long sessions at the laboratory and laptop knowing how much this work meant to me. This degree is as much mine as theirs; I truly couldn‘t have done it without their love and support. Thanks boys!!! vi

Table of Contents

List of Tables ...... ix List of Figures ...... xi Abstract ...... xiii Chapter 1 Problem Description ...... 1 References ...... 3 Chapter 2 Effect of Chemical Uncouplers on Microbial Respiration and Biomass Production: Implications for their use in Wastewater Treatment Facilities ...... 8 Introduction ...... 8 Respiration ...... 10 ...... 16 The Process of ATP Synthesis and Its Uncoupling ...... 20 Application of Chemical Uncouplers to Activated Sludge Processes ...... 29 References ...... 36 Chapter 3 The activity of 1-hydroxyanthraquinone on in Paracoccus denitrificans...... 49 Abstract ...... 49 Introduction ...... 49 Materials and Methods ...... 57 Organism and Growth Conditions ...... 57 Chemicals and Stock Solutions ...... 58 Experimental Procedures ...... 59 Results and Discussion ...... 61 Comparison of 1-Hydroxyanthraquinone with respiratory inhibitors and uncouplers of oxidative phosphorylation on denitrifying cultures of Paracoccus denitrificans .. 61 vii

Effect of 1-Hydroxyanthraquinone, 9,10-Anthraquinone and 2,4-Dinitrophenol on anaerobic cultures of Paracoccus denitrificans using nitrate or nitrite as the terminal electron acceptor ...... 67 References ...... 79 Chapter 4 Effect of 1-hydroxyanthraquinone on biomass accumulation and utilization in Paracoccus denitrificans and in mixed wastewater cultures ...... 91 Abstract ...... 91 Introduction ...... 91 Materials and Methods ...... 99 Organisms and Growth Conditions ...... 99 Paracoccus...... 99 Activated Sludge Cultures ...... 99 Chemicals and Stock Solutions ...... 100 Experimental Procedures ...... 101 Respirometry Studies ...... 101 Paracoccus Continuous Studies ...... 103 Biomass Determinations ...... 105 Results ...... 106 Effect of OHAQ on Paracoccus ...... 106 Paracoccus Substitution Studies ...... 109 Paracoccus Continuous Culture Studies ...... 114 Mixed Culture Batch Studies...... 116 Discussion ...... 118 References ...... 127 Chapter 5 The economic implications of metabolic uncoupling in wastewater treatment facilities ...... 137 References ...... 148 Chapter 6 Summary and Conclusions ...... 152 Chapter 7 Barriers to Implementation...... 155 viii

Lack of Pilot Scale Demonstrations ...... 155 Acclimation ...... 156 Effluent Discharges to Environment ...... 158 References ...... 160

ix

List of Tables

Chapter 2 Effect of Chemical Uncouplers on Microbial Respiration and Biomass Production: Implications for their use in Wastewater Treatment Facilities

Table 1: Classic Uncouplers of Respiration...... 28

Chapter 3 The activity of 1-hydroxyanthraquinone on anaerobic respiration in Paracoccus denitrificans

Table 1: Respiratory Inhibitors and Uncouplers of Oxidative Phosphorylation Studied in Pa. denitrificans under Denitrifying Conditions with Nitrate as the Electron Acceptor ...... 62

Table 2: Nitrogen production (mean mMoles N2/bottle) of Pa. denitrificans with 10 mg/L OHAQ, electron transport inhibitors (rotenone, NBD-Cl, and antimycin), ATPase inhibitors (DCC, Tinopal) and uncouplers of oxidative phosphorylation (DNP, FCCP and CCCP)...... 63

Table 3: The effect of 1-hydroxyanthraquinone (OHAQ) and 2,4-Dinitrophenol (DNP) on biomass production and nitrogen respiration rate in anaerobic cultures of Pa. denitrificans with nitrate as the Terminal Electron Acceptor...... 67

Table 4: The effect of 1-hydroxyanthraquinone (OHAQ), 9,10-Anthraquinone (AQ) and 2,4-Dinitrophenol (DNP) at a Test Concentration of 10 mg/L on biomass production and nitrogen respiration rate in anaerobic cultures of Pa. denitrificans with nitrate as the electron acceptor...... 69

Table 5: The effect of 1-hydroxyanthraquinone (OHAQ), 9,10-Anthraquinone (AQ) and 2,4-Dinitrophenol (DNP) at a Test Concentration of 10 mg/L on biomass production and nitrogen respiration rate in anaerobic cultures of Pa. denitrificans with nitrite as the electron acceptor...... 72

Chapter 4 Effect of 1-hydroxyanthraquinone on biomass accumulation and oxygen utilization in Paracoccus denitrificans and in mixed wastewater cultures

Table 1: The effect of 1, 5 and 10 mg/L OHAQ on biomass production and oxygen utilization rate in batch aerobic cultures of Pa. denitrificans...... 107

Table 2: Biomass Reductions, Log P, pKa and Water Solubilities of Test Compounds evaluated in Batch Aerobic Cultures of Pa. denitrificans ...... 112 x

Table 3: Summary of the Continuous Pa. denitrificans Reactor Results for the Control and OHAQ Studies. Reactor was Fed BTZ-3 Medium containing Acetone (for the control) or 10 mg/L OHAQ at a rate of 1.5 ml/min...... 115

Table 4: The effect of 1-hydroxyanthraquinone (OHAQ) at a test concentration of 10 mg/L on Oxygen Consumption rate and biomass production in aerobic mixed cultures derived from Municipal Wastewater Treatment facility...... 117

Chapter 5 The economic implications of metabolic uncoupling in wastewater treatment facilities

Table 1: Operating Conditions for Industrial and Municipal Wastewater Treatment Facilities used in Economic Analysis...... 139

Table 2: Economic analysis for Industrial Wastewater Treatment Facility ...... 143

Table 3: Economic analysis for a Municipal Wastewater Treatment Facility with a Hydraulic Flow Rate of 6,319 GPM...... 144

Table 4: Economic analysis for a Municipal Wastewater Treatment Facility with a Hydraulic Flow Rate of 817 GPM...... 145

xi

List of Figures

Chapter 2 Effect of Chemical Uncouplers on Microbial Respiration and Biomass Production: Implications for their use in Wastewater Treatment Facilities Figure 1: Simplified relationship between catabolism and anabolism ...... 9 Figure 2: Aerobic Electron Transport Pathway in Paracoccus denitrificans ...... 11 Figure 3: Electron Transport Pathway in Mitochondria ...... 11 Figure 4: The Q-Cycle in Mitochondria ...... 15 Figure 5: Anaerobic electron transport chain in Paracoccus denitrificans ...... 17

Figure 6: Schematic Model for the Arrangement of the F0F1 subunits of ATPase ...... 21 Figure 7: The molecular mechanism of H+ translocation mediated by a protonophoric uncoupler ...... 26

Chapter 3 The activity of 1-hydroxyanthraquinone on anaerobic respiration in Paracoccus denitrificans Figure 1: Simplified relationship between catabolism and anabolism ...... 51 Figure 2: Anaerobic electron transport chain in Paracoccus denitrificans ...... 55 Figure 3: The effect of OHAQ and Uncouplers on nitrogen production in anaerobic cultures of Pa. denitrificans with nitrate as the Terminal Electron Acceptor...... 66 Figure 4: The effect of 1-hydroxyanthraquinone (OHAQ), 9,10-Anthraquinone (AQ) and 2,4-Dinitrophenol (DNP) at a Test Concentration of 10 mg/L on nitrogen Production in anaerobic cultures of Pa. denitrificans with nitrate as the electron acceptor...... 70 Figure 5: The effect of 1-hydroxyanthraquinone (OHAQ), 9,10-Anthraquinone (AQ) and 2,4-Dinitrophenol (DNP) at a Test concentration of 10 mg/L on nitrogen respiration in anaerobic cultures of Pa. denitrificans with nitrite as the electron acceptor...... 73 Figure 6: Distribution of Electrons in Substituted Anthraquinone Compounds with Electron Donating Groups ...... 76

Chapter 4 Effect of 1-hydroxyanthraquinone on biomass accumulation and oxygen utilization in Paracoccus denitrificans and in mixed wastewater cultures Figure 1: Simplified relationship between catabolism and anabolism ...... 93 Figure 2: Aerobic Electron Transport Pathway in Paracoccus denitrificans ...... 97 xii

Figure 3: The effect of OHAQ on Oxygen Utilization in Aerobic Batch Cultures of Pa. denitrificans ...... 108 Figure4: Anthraquinone test compounds ...... 110 Figure 5: The Effect of OHAQ to Biomass Ratio on Biomass Reductions in Batch and Continuous Paracoccus and Batch Mixed Culture Studies...... 120 Figure 6: Distribution of Electrons in Substituted Anthraquinone Compounds with Electron Donating Groups ...... 122 Figure 7: The molecular mechanism of H+ translocation mediated by a protonophoric uncoupler ...... 123

Chapter 5 The economic implications of metabolic uncoupling in wastewater treatment facilities Figure 1: The Yearly Reagent Costs for Reagent Doses for Reagent Doses of 1 to 10 mg/L and Reagent Costs of $2.5/KG to $20.5/KG in a Moderate Flow Municipal Wastewater Treatment Facility...... 147

xiii

Abstract Activity in Paracoccus denitrificans and potential application for biomass reduction in wastewater treatment facilities Lynn Ann Arlauskas-Dekleva Michael A. Gealt, Ph.D.

The effect of 1-hydroxyanthraquinone (OHAQ) on respiration rate and biomass production was evaluated in Paracoccus denitrificans and in mixed cultures derived from a municipal wastewater treatment facility. OHAQ addition resulted in classic

―uncoupling‖ activity characterized by an increase in the respiration rate and decrease in the biomass production. In anaerobic batch cultures, OHAQ increased the nitrogen production rate 64.5 % and decreased biomass production 30.7% in Paracoccus. In aerobic batch cultures, OHAQ increased the oxygen utilization rate 38.5% and decreased the biomass production 39.3%. OHAQ addition to a continuous Paracoccus culture resulted in an increase in the respiratory activity compared to the steady state control values. There was a 10.4% in the carbon evolution rate (CER) and a 10.8 % increase in the specific oxygen utilization rate (SOUR). The growth of the organism was not adversely affected by the presence of OHAQ suggesting that the compound could increase the capacity of wastewater treatment facilities without modification to existing equipment. The addition OHAQ to mixed cultures derived from a municipal wastewater treatment facility resulted in an increase in the initial oxygen utilization rate and a xiv

reduction in the overall biomass production. The overall oxygen utilization was not adversely affected by OHAQ. In test vessels containing 300 mg/L MLSS, 10 mg/L

OHAQ increased the initial oxygen consumption rate 16.9% and decreased the biomass production 10.8 %.

The activity of OHAQ derivatives was related to the electronic properties and position of the substitution on the anthraquinone ring system. There was no correlation to the Log P, pKa or water solubility of the compound.

These studies demonstrate that the activity of OHAQ is the result of the

―uncoupling‖ of aerobic and anaerobic respiration. The results indicate that OHAQ has the potential to reduce both the capital and operating costs of biological wastewater treatment systems, especially in the industrial setting, by simultaneously improving the metabolic capacity and decreasing the biomass production rates. The uncoupling activity of anthraquinones combined with the low toxicity of compounds suggests that these compounds may be used for reducing biomass production in commercial bioprocesses in addition to traditional wastewater treatment systems.

1

Chapter 1 Problem Description

The primary purpose of biological waste treatment processes is the removal of organic material 51, thereby decreasing its environmental burden before its discharge.

The activated sludge process is the most common biological treatment process for domestic and industrial wastewater 55 and has proven to be both a reliable and efficient treatment technology capable of producing high-quality effluents 173. Wastewaters typically contain a complex mixture of components which are degraded by a mixed microbial population in biochemical reactions 99 that convert the organic material into new cell mass (biomass) and respiration products (carbon dioxide and water) 94, 98, 99.

The excess biomass produced in the activated sludge process can be significant and requires additional processing and disposal 173. The costs associated with the handling and disposal of biomass from biological treatment processes typically represent from

40 to 60% of a wastewater treatment plant operating costs 27, 99, 190 and may require significant capital expenditures.

The disposal of excess biomass may include landfill, incineration and use on agricultural land 104. In 2000, approximately 50% of the 7.5 millon dry metric tons of biosolids produced were land applied 188 151, 14% were disposed of in a landfill and

22% were incinerated 38. Projections of an increase in biosolids production 38 combined with increasing political, economic and regulatory pressure 54 emphasizes the need for technologies that minimize biomass production in wastewater treatment facilities. 2

There are many technologies that can reduce biomass production rates but one technology in particular, chemical uncoupling, has shown a lot of promise. Chemical uncouplers disrupt the link between the catabolic (energy yielding oxidation) step and the formation of ATP and the production of new cells in anabolism 134 therefore limiting the cells ability to capture energy from substrate oxidation 98, 203. Laboratory studies have suggested that the application of the uncoupler technology to activated sludge units can result in a significant reductions in the biomass production rate 93, 99,

101, 199, 201-203. The most active uncoupler compounds are usually the most toxic and a serious problem with the uncoupler technology in wastewater treatment is the potential presence of the compound in the effluent where it might be toxic to the environment.

A number of uncoupler compounds, (e.g. DNP, chlorinated phenols) are designated as toxic pollutants subject to effluent limitations 63, 64; impeding their utility at a full-scale wastewater treatment facility.

Anthraquinones, are found widely dispersed in the environment 204, occurring as the result of both natural and commercial processes 7, 18, 50, 204. They exhibit a wide variety of biological effects 7, 9, 10, 16, 18, 29, 34, 67, 70, 80-82, 131, 132, 148, 149, 189, but the focus of this investigation is their potential uncoupling activity in denitrifying and aerobic cultures. Uncoupling activity has been demonstrated with mitochondria 10, 16, 82 and sulfate respiring microorganisms 29 but not with aerobic or anaerobic microorganisms

29, 189.

The objective of this investigation is to determine if 1-hydroxyanthraquinone

(OHAQ) can reduce biomass production through the process of uncoupling without 3

adversely affecting the respiration rate of the organism. The specific research goals of

the program are to:

Develop anaerobic and aerobic systems to evaluate the activity of OHAQ.

Determine if 1-hydroxyanthraquinone possesses uncoupling activity under

aerobic and denitrifying conditions.

Evaluate if the activity can be expanded to mixed cultures derived from a

wastewater treatment facility.

References

1. Anke, H., et al., Metabolic Products of Microorganisms. 185*. The Anthraquinones of the Aspergillus glaucus Group. I. Occurrence, Isolation, Identification and Antimicrobial Activity. Archives of Microbiology, 1980. 126: p. 223-230.

2. Bakola-Christianopoulou, M.N., L.B. Ecateriniadou, and K.J. Sarrris, Evaluation of the Antimicrobial Activity of a New Series of Hydroxy-Quinone Chelates of Some Transition Metals. European Journal of Medicinal Chemistry - Chim. Ther, 1986. 21(5): p. 385-390.

3. Betina, V. and S. Kuzela, Uncoupling effect of fungal hydroxyanthraquinones on mitochondrial oxidative phosphorylation. Chem.-Biol. Interact., 1987. 62(2): p. 179-89.

4. Boos, K.S. and E. Schlimme, Anthraquione Dyes: A New Class of Potent Inhibitors of Mitochrondrial Adenine Nucleotide Translocation and Oxidative Phosphorylation. FEBS Letters, 1981. 127(1): p. 40-44. 4

5. Brown, J.P., A Review of the Genetic Effects of Naturally Occurring Flavonoids, Anthraquinones and Related Compounds. Mutation Research, 1980. 75: p. 243- 277.

6. Constable, S.W., Solids disposal costs. 1995, Personal Communication.

7. Cooling III, F.B., et al., Inhibition of Sulfate Respiration by 1,8- Dihydroxyanthraquinone and Other Anthraquinone Derivatives. Applied and Environmental Microbiology, 1996. 62(8): p. 2999-30004.

8. Cudlin, J., et al., Biological activity of hydroxyanthraquinones and their glucosides toward microorganisms. Folia Microbiol., 1976. 21(1): p. 54-7.

9. EPA, Biosolids Generation, Use and Disposal in The United States. 1999, United States Environmental Protection Agency, Municipal and Industrials Solid Waste Division, Office of Solid Waste: Washington, DC. p. 74.

10. Garcia-Lopez, P.M., L. Kung Jr, and J.M. Odom, In Vitro Inhibition of Microbial Methane Production by 9,10-Anthraquinone. Journal of Animal Science, 1996. 74: p. 2276-2284.

11. Gaudy Jr, A.F. and E.T. Gaudy, Mixed Microbial Populations. Advances in Biochemical Engineering, 1972: p. 97-143.

12. Goldstein, N., The State of Biosolids in America. Biocycle 2000. 41(1212): p. 50- 56.

13. Grady Jr, C.P.L. and H.C. Lim, Biological Wastewater Treatment: Theory and Applications. Pollution Engineering and Technology, ed. P.N. Cheremisinoff. Vol. 12. 1980, New York: Marcel Dekker, Inc. 963.

14. Health, N.I.o., Hazardous Substances Databank Trichlorophenol. 2000, National Library of Medicine's TOXNET System.

15. Health, N.I.o., Hazardous Substances Databank Dinitrophenol. 2000, National Library of Medicine's TOXNET System. 5

16. Hilgert, I., et al., Antitumor and immunosuppressive activity of hydroxyanthraquinones and their glucosides. Folia Biol., 1977. 23(2): p. 99-109.

17. House, R.V. and J.H. Dean, Structural requirements for suppression of cytotoxic T lymphocyte induction by polycyclic compounds. In Vitro Toxicol., 1987. 1(3): p. 149-62.

18. Karukstis, K.K., et al., Alternative Measures of Photosystem II Electron Transfer Inhibition in Anthraquinone-Treated Chloroplasts. Photochemistry and Photobiology, 1992. 55(1): p. 125-132.

19. Kavanagh, F., Activities of Twenty-Two Antibacterial Substances Against Nine Species of . Journal of Bacteriology, 1947. 54: p. 761-766.

20. Kawai, K., et al., The inhibition of mitochondrial respiration by anthraquinone mycotoxins. The relationship between the position of hydroxyl group on the anthraquinone nucleus and the uncoupling effect on oxidative phosphorylation. Proc. Jpn. Assoc. Mycotoxicol., 1987. 26: p. 43-6.

21. Liu, Y., Effect of Chemical Uncoupler on the Observed Growth Yield in Batch Culture of Activated Sludge. Water Research, 2000. 34(7): p. 2025-2030.

22. Liu, Y. and J.-H. Tay, Strategy for Minimization of Excess Sludge Production from the Activated Sludge Process. Biotechnology Advances, 2001. 19: p. 97- 107.

23. Low, E.W. and H.A. Chase, The Use of Chemical Uncouplers for Reducing Biomass Production During Biodegradation. Water Science and Technology, 1998. 37(4/5): p. 399-402.

24. Low, E.W. and H.A. Chase, Reducing Production of Excess Biomass During Wastewater Treatment. Water Research, 1999. 33(5): p. 1119-1132.

25. Low, E.W., et al., Uncoupling of Metabolism to Reduce Biomass Production in the Activated Sludge Process. Water Research, 2000. 34(12): p. 3204-3212. 6

26. Mayhew, M. and T. Stephenson, Low Biomass Yield Activated Sludge: A Review. Environmental Technology, 1997. 18: p. 883-892.

27. Odom, J.M., Anthraquinone inhibition of methane production in methanogenic bacteria, in 42 pp. 1994, Bio-Technical Resources LP, USA: PCT Int. Appl.

28. Oettmeier, W., K. Masson, and A. Donner, Anthraquinone inhibitors of photosystem II electron transport. FEBS Letters, 1988. 231(1): p. 259-62.

29. Okey, R.W. and H.D. Stensel, Uncouplers of Oxidative Phosphorylation. Modeling and Predicting Their Impact on Wastewater Treatment and the Environment, in ACS Symposium Series Emerging Technologies in Hazardous Waste Management V, D.W. Tedder and F.G. Pohland, Editors. 1995, American Chemical Society: Washington. p. 284-309.

30. Powell, A.K., Growth-Inhibiting Action of Some Pure Substances. Nature, 1944. 153: p. 345.

31. Prestera, T., H.J. Prochaska, and Talalay, Inhibition of NAD(P)H:(Quinone- Acceptor) Oxidoreductase by Cibacron Blue and Related Anthraquinone Dyes: A Structure-Activity Study. Biochemistry, 1992. 31: p. 284-293.

32. Renner, R., Scientist debate fertilizing soils with sewage sludge. Environmental Science & Technology, 2000. 34(11): p. 242A-243A.

33. Strand, S.E., G.N. Harem, and H.D. Stensel, Activated-Sludge Yield Reduction Using Chemical Uncouplers. Water Environment Research, 1999. 71(4): p. 454- 458.

34. WEF, Biosolids Fact Sheets. 1998, Water Environment Federation.

35. Weimer, P.J., et al., Anthraquinones as inhibitors of sulfide production by sulfate- reducing bacteria in sewage, oil wells, process tanks, or biomass fermentation, in 22 pp. Cont.-in-part of U.S. Ser. No. 652,406, abandoned. 1993, du Pont de Nemours, E. I., and Co., USA: U.S. 7

36. Wheatley, A.D., Effluents, in Biodeterioration 7, D.R. Houghton, R.N. Smith, and H.O.W. Eggins, Editors. 1988, Cranfield Institute of Technology: Bedford. p. 193-206.

37. Yang, X.-F., M.-L. Xie, and Y. Liu, Metabolic Uncouplers Reduce Excess Sludge Production in an Activated Sludge Process. Process Biochemistry, 2003. 38: p. 1373-1377.

38. Ye, F.X., D.S. Shen, and Y. Li, Reduction in Excess Sludge Production by Addition of Chemical Uncouplers in Activated Sludge Batch Cultures. Journal of Applied Microbiology, 2003. 95: p. 781-786.

39. Ye, F.X. and Y. Li, Reduction of Excess Sludge Production by 3,3',4',5- Tetrachlorosalicylanilide in an Activated Sludge Process. Applied Microbiology and Biotechnology, 2005. 67: p. 269-274.

40. Ye, F.X. and Y. Li, Uncoupled Metabolism Stimulated by Chemical Uncoupler and Oxic-Settling-Anaerobic Combined Process to Reduce Excess Sludge Production. Applied Biochemistry and Biotechnology, 2005. 127(3): p. 187-199.

41. Yoshimi, N., et al., The mRNA overexpression of inflammatory enzymes, phospholipase A2 and cyclooxygenase, in the large bowel mucosa and neoplasms of F344 rats treated with naturally occurring carcinogen, 1- hydroxyanthraquinone. Cancer Lett., 1995. 97(1): p. 75-82.

8

Chapter 2 Effect of Chemical Uncouplers on Microbial Respiration and Biomass Production: Implications for their use in Wastewater Treatment Facilities

Lynn Ann Arlauskas-Dekleva and Michael A. Gealt

Introduction

The primary purpose of biological treatment processes is the removal of organic

material 51, thereby decreasing its environmental burden before its discharge. The

activated sludge process is the most common biological treatment process for domestic

and industrial wastewater 97, 154 and has proven to be a reliable and efficient treatment

technology capable of producing high-quality effluents 173. Wastewaters typically

contain a complex mixture of components which are degraded by a mixed microbial

population in biochemical reactions that convert the organic material into new cell

mass (biomass) and respiration products (carbon dioxide and water) 94, 98, 99. The

excess biomass produced in the activated sludge process can be significant and requires

additional processing and disposal 173. The costs associated with the handling and

disposal of biomass from biological treatment processes can represent from 40 to 60%

of a wastewater treatment plant operating costs 27, 99, 190 and may require significant

capital expenditures.

The metabolism of organic wastes, in catabolism, provides the energy required

for cell synthesis (anabolism) in the form of an energy rich molecule, ATP 24, 97, 112. 9

The relationship between ATP formation by catabolism and its use in biomass synthesis 24, shown in Figure 1, illustrates the relationship between the two processes.

FIGURE 1: SIMPLIFIED RELATIONSHIP BETWEEN CATABOLISM 24 AND ANABOLISM (ADAPTED FROM REFERENCE )

Biomass

Substrate ATP

Catabolism Anabolism

Substrate

CO + H O 2 2 ADP + Pi

Energy

ATP, synthesized in a process known as respiration 127, 167, provides the energy required for cell synthesis 112 and serves as a link between substrate oxidation and biomass synthesis 24. Chemical compounds known as uncouplers limit the cells ability to capture energy from substrate oxidation 98, 202 and disrupt the energetic link between catabolism (oxidation of organic matter) and anabolism (ATP generation) 105.

10

Respiration

In most organisms, the majority of the ATP is synthesized in a process known

as respiration 127, 167 that utilizes a sequence of oxidation-reduction reactions to transfer

electrons from the substrate to oxygen 60 or, in the absence of oxygen, to inorganic

acceptors, e.g. nitrate 19. The electrons travel through a series of membrane-associated

electron carrier proteins and lipids 191, known as the electron transport chain or

respiratory network 181. The components of the electron transport chain are organized

into dehydrogenase and oxidase complexes connected by quinones 191 and catalyze the

downhill transfer electrons from NADH and couple to the O2 through a redox potential

span of 1.1V 127.

The electron transport chain is localized in a membrane essentially

impermeable to most ions, including OH- and H+ 61. In prokaryotes, the membrane is

associated with the outer cellular membrane. In eukaryotic systems, the membrane is

the inner membrane of the . The flow of electrons is coupled to the

translocation of protons across the membrane that generates an electrical or membrane

potential ( ) and proton gradient ( pH) on the opposite sides of the membrane 40, 57,

61, 68, 114, 125.

In bacteria there is considerable variation in the components of the electron

transport chain, depending on the species and the culture growth conditions, 119. The

composition of the aerobic electron transport chain of Paracoccus denitrificans, a gram

negative 171, 181 non-fermenting facultative anaerobe 15, 57, 77, 184, shown in Figure 2 57, 75, 11

119, 181, 191, closely resembles the four enzymatic complexes (I-IV) of a mitochondrial electron transport chain, Figure 3 21, 57, 74, 75, 83, 119, 127, 161, 181, 198.

FIGURE 2: AEROBIC ELECTRON TRANSPORT PATHWAY IN 57, 75, 119, 181, PARACOCCUS DENITRIFICANS (MODIFIED FROM REFERENCES 191)

cyt o O2

dehydrogenase UQ bc1 c cyt aa3 O2

dehydrogenase

FIGURE 3: ELECTRON TRANSPORT PATHWAY IN MITOCHONDRIA 119, 191

Complex III

NADH dehydrogenase UQ Cytochrome bc1 Cytochrome c Cytochrome c oxidase

Complex I Complex IV Succinate dehydrogenase Complex II

12

In the mitochondrial and Paracoccus electron transport chains, the electrons from NADH flow to oxygen via the NADH dehydrogenase enzyme complex, quinones, cytochrome bc1 complex, cytochrome c and a terminal cytochrome c oxidase

181. Most bacteria, unlike the mitochondrial respiratory chain, contain a number of alternative pathways in which electrons are transferred from specific dehydrogenases to a variety of terminal oxidases. This respiratory network enables bacteria to respond to a variety of environmental stimuli 8, 181.

The Paracoccus aerobic pathway, Figure 2, branches at ubiquinone (UQ) leading either through the bc1 complex to cytochrome aa3 (cyt aa3), or bypassing the

191 bc1 complex to cytochrome o (cyt o) . The extent to which the o-type cytochrome is active under aerobic conditions is thought to vary with different Paracoccus strains and growth conditions 160, 179, 184. For example, the cytochrome o pathway may function as a ―proton-motive release valve‖ under conditions of diminished ATP consumption.

Electron flow through the two coupling sites (bc1 and cytochrome aa3) generates a proton-motive force, P, which drives ATP production. As the ATP consumption decreases, the proton-motive builds to a point of retarding electron flow through the bc1 complex. The diversion of electron through the cytochrome o pathway, a non- coupling site, maintains the oxidation rates of NADH and ubiquinol. In a similar manner, the pathway may provide a survival mechanism in the presence of bc1 complex inhibitors or toxins 179, 191. 13

Complex I (or NADH-Coenzyme Q Reductase) of the mitochondrial electron transport chain contains at least 46 different subunits 198 and is composed of NADH dehydrogenase, a non-covalently bound FMN, Coenzyme Q (CoQ) and eight FeS

(iron-) clusters 2, 74, 127, 161, 198. The enzyme complex is an L-shaped structure that spans the inner mitochondrial membrane and protrudes into the mitochondrial matrix 2,

127, 161, 198. The NADH dehydrogenase complex (NDH-1) from Paracoccus, composed of 10-14 dissimilar subunits 197, 198, non-covalently bound FMN and FeS groups 196, 197, is similar to the mitochondrial Complex I in terms of electron carriers, inhibitor specificity and amino acid sequence of subunits but is much simpler in structure 85, 198.

The NADH-Quinone Oxidoreductase and NDH-1 complexes generate a transmembrane electrochemical potential through redox-driven proton translocation 181 by means of coupling the transfer of electrons from NADH to the translocation of four protons across the membrane 181, 191, 196, 198.

Mitochondrial Complex II (or Succinate-Ubiquinone Oxidoreductase) is composed of succinate dehydrogenase, flavin adenine dinucleotide (FAD) and three

FeS clusters 2, 127. The four-subunit enzyme in mammalian systems is composed of two integral membrane protein subunits and two hydrophilic proteins that lie in the matrix space 74, 161. The direction of electron flow through this reversible enzyme complex is dictated by the relative concentrations of reactants and products and the enzyme participates in both the respiratory chain and the tricarboxylic acid (TCA) cycles; 161. Complex II is not a coupling site in the respiratory pathway 127 and 14

transfers electrons from succinate to ubiquinone, which is located within the inner membrane 74.

Complex III, Cytochrome C Reductase (or Coenzyme Q -Cytochrome C

191 Reductase, also called the bc1 complex) , is a multi-subunit enzyme present in the respiratory chain of mitochondria and prokaryotes, possessing Class I terminal oxidases 8. The enzyme is composed of cytochrome c1, b-type cytochromes and an

8, 127, 169 iron-sulfur (FeS) protein . The cytochrome bc1 complex spans the inner membrane, extending into both the cytoplasm and matrix of the mitochondria 161. The

FeS protein, or Rieske protein, first described and isolated by Rieske et al. in 1964, contains a Fe2S2 cluster in which one iron atom (Fe1) is coordinated to the protein by

43, 91 two cysteine residues and the other iron (Fe2) by two histidine residues . The two inorganic sulfide ions bridge the two iron ions forming a flat, rhombic cluster and hydrogen bond with the main-chain nitrogens. The iron ligated to the cysteines remains in the ferric state, regardless of the reduction state of the cluster, while the histidine-ligated iron goes from a ferric state to a ferrous state when reduced 43, 185.

The oxidation of ubiquinone (QH2) and reduction of quinone (Q) occur during the reaction cycle 161 and is associated with translocation of four protons 74, 127. The reaction cycle, Figure 4, begins with the oxidation of ubiquinol to the semi-quinone anion, UQ¯, and the release of two protons on the cytoplasmic side of the membrane at the cytoplasmic (P-phase) of the enzyme 127, 161. The electron is passed to the high-

161 potential FeS complex, which subsequently transfers the electron to cytochrome c1 .

The semi-quinone anion is oxidized to UQ and then is transferred to heme bl , then 15

heme bh , where it is used to reduce a molecule of ubiquinone at the matrix to its semi- quinone form. A second QH2 is oxidized at the cytoplasmic site and the semi-quinone is reduced to QH2, picking up to protons from the matrix side of the membrane. The full reaction cycle consists of the oxidation of two molecules of ubiquinol with the associated release of four protons to the cytoplasmic side of the membrane, the reduction of one molecule of ubiquinone coupled to the uptake of two protons from the matrix space, and the reduction of two molecules of cytochrome c 161.

161 FIGURE 4: THE Q-CYCLE IN MITOCHONDRIA

Fe2S2 c 1 c Cytoplasm (P-phase)

e- - QH2 Q Q e-

b L

b H

- QH2 Q Q

Matrix (N-phase)

Complex IV, also known as Cytochrome C Oxidase, catalyzes the transfer of four electrons from ferrocytochrome c to dioxygen, in the terminal step of respiration,

8, 127, 161. Cytochrome c oxidase is composed of 13-subunits in mitochondria; bacterial 16

cytochrome c oxidase contains three or four subunits161. The enzyme is an integral

membrane protein 122 that contains two spectrally distinct, though chemically identical

127 8, heme a species, heme a and heme aa3 , and two copper complexes CuA and CuB

127 . Electrons from cytochrome c are transferred to a copper center, CuA, located on

the electrically positive (P) cytoplasmic side of the membrane and transferred via heme

122 a to a heme-copper binuclear center (heme a3 and CuB) . Molecular oxygen binds to

the reduced heme a3 and is reduced by four electrons and four protons taken from the

negative (N) matrix side of the membrane, to form water 122, 161. Complex IV, through

a series of conformational changes, couples this reduction reaction with the

unidirectional transport of up to four additional matrix protons (vectorial protons) 161.

Denitrification

A reduction in the molecular oxygen concentration induces the anaerobic

respiratory network in Paracoccus denitrificans. This organism can utilize nitrate as a

terminal electron acceptor, and to do so synthesizes four enzyme complexes that

catalyze the reduction of nitrate, nitrite, , and in a process

known as denitrification 4, 19, 36, 84, 140, 181, 191, 192. The conversion to denitrifying

conditions influences the composition of the electron transport chain of Pa.

11, 21, 157, 160 denitrificans; there is a sharp decrease in synthesis of cytochromes a + a3

and an increase in the c-type cytochromes 11, 157, 160, including cytochrome o 57, 191. The 17

anaerobic electron transport pathway includes several branches to the individual reductases 139, 191 and the denitrifying enzymes are ‗plugged‘ into the respiratory chain, shown in Figure 5, at the level of either ubiquinone or cytochrome c 42, 171.

FIGURE 5: ANAEROBIC ELECTRON TRANSPORT CHAIN IN 191 PARACOCCUS DENITRIFICANS MODIFIED FROM

- NO3 nitrate reductase

- NO2

dehydrogenase UQ bc1 c nitrite reductase NO nitric oxide reductase

1/2 N O nitrous oxide 2 reductase

1/2 N2

Ubiquinol transfers electrons directly to nitrate reductase, bypassing the bc1 complex and the c cytochromes 32, 76, 171, 191. Electron flow from ubiquinol to nitrite, nitrous, and nitric reductases proceeds via the cytochrome bc1 complex to cytochrome c 11, 19, 42, 77, 139, 171, 180, 181. There are two proton-translocating sites in the anaerobic electron transport chain of Pa. denitrificans. The first site is the NADH dehydrogenase

181, 191 and the second site is the bc1 complex . 18

The first step in denitrification is catalyzed by a membrane bound nitrate reductase, which converts nitrate to nitrite by the reaction 19, 76, 87, 171:

- + - - NO3 + 2H + 2e  NO2 + H2O

The reduction of nitrate is achieved at the active site of the -subunit located on the cytoplasmic (N) side of the membrane 181 and the conversion of nitrate is coupled to the translocation of protons 69 derived from the cytoplasm 78. The enzyme complex consists of three different subunits; one transmembrane subunit, , and two soluble subunits, - and associated with the cytoplasmic membrane 181. Electrons from ubiquinol, on the periplasmic side of the membrane, are passed through the membrane via the b-type hemes in the -subunit and iron-sulfur centers in the -subunit to the molybdenum cofactor in the -subunit 181 and two electrons are released to the periplasm 42.

The subsequent reaction, mediated by nitrite reductase, a soluble periplasmic enzyme 19, 30, 42, 69, 84, 90, 171, converts nitrite into gaseous nitric oxide by the reaction 181:

- + - NO2 + 2H + e  NO + H2O

The enzyme contains both c- and d1-hemes and can receive electrons from

42 cytochrome bc1 via cytochrome c550 on the periplasmic face of the membrane . The active site of the enzyme is located at the periplasmic side of the cytoplasmic 19

3, 14 membrane where the transfer of two electrons through cytochrome bc1, as in the aerobic and mitochondrial electron transport chain, is coupled to the translocation of four protons to the periplasmic side of the membrane 12, 14, 138, 171.

Nitric oxide reductase, an integral membrane protein 127, 191, catalyzes the reduction of nitric oxide to nitrous oxide by the reaction 181:

+ - 2NO + 2H + 2e  N2O + H2O

The enzyme consists of two subunits and contains both b- and c-type hemes and a non-heme iron 42, 181. The nitric oxide is reduced at the periplasmic surface of the membrane by electrons from the cytochrome bc1 complex and net outward charge transfer 42

The final step in denitrification, the reduction of nitrous oxide to dinitrogen is

181 carried out by the periplasmic N2O reductase :

+ - N2O + 2H +2e  N2 + H2O

The enzyme is a soluble periplasmic enzyme that contains two copper centers,

181 center A and center Z . The complex is linked to cytochrome bc1 via cytochrome

127 c550 and the protons required for reduction of nitrous oxide are taken from the 20

periplasmic side of the plasma membrane 11, 31. The reduction of nitrous oxide is

accompanied by proton translocation 11, 106.

The Process of ATP Synthesis and Its Uncoupling

The net movement of charges across the membrane in the electron transport

chain generates an electrical or membrane potential ( ) and proton gradient ( pH) on

the opposite sides of the membrane 57, 61, 68, 114. The energy contained in the pH and

known collectively as the electrochemical gradient or proton-motive force ( P) 57,

61, 68, 114-116, is recovered by the cell as the protons return across the membrane 103. The

cell exploits this movement of protons at the to synthesize ATP or to

drive nutrient transport and other endergonic reactions 57, 61, 114-116.

ATP is synthesized from ADP and inorganic phosphate (Pi) in a reversible

reaction mediated and facilitated by the transmembrane enzyme complex, ATPase

(ATP synthase) 35, 45, 53, 187, 200. The synthesis of ATP from ADP and inorganic

phosphate (Pi) is dependent on the driving force of the charge and proton gradient

components of P, and on the physical import of protons across the membrane 35, 142.

The ATPase complex from the membranes of mitochondria, chloroplasts and bacteria

are similar in structure and mechanism 37, 47, 48, 53, 206 and consist of a membrane

37, 44, 48, 53, 117, 118, 187, 200, 206 extrinsic F1 and transmembrane F0 components , shown

schematically in Figure 6 56, 187. 21

FIGURE 6: SCHEMATIC MODEL FOR THE ARRANGEMENT OF THE 56, 187 F0F1 SUBUNITS OF ATPASE

The F1 component, is located at the interior surface of the cell membrane and possesses catalytic and proton translocating properties 117, 136. In mitochondria and

22, bacteria, F1 is composed of five different subunits with the stoichiometry 3 3 1 1 1 22

136 which form structure consisting of a catalytic hexagon ring of alternating and subunits surrounding a central cavity containing the -subunit 53, 156, 187, 205.

The lipophillic F0 component, an integral membrane component, forms the

118, 124 channel through which protons are imported to F1 . The F0 component contains

22, 136 three essential subunits with a stochiometry a1b2c9-12 . Subunits a and c are hydrophobic proteins that span the membrane 56 and play an essential role in proton

45, 56, 185 translocation through F0 . Subunit b has a hydrophilic domain exposed to the cytoplasm and is anchored to the membrane by a hydrophobic N-terminal segment 45,

56. Subunit b does not directly participate in proton translocation 56 but is required for

45 F1 binding . The F0 subunits are proposed to be arranged in a rotary fashion, with a ring of c subunits surrounding the transmembrane helices of subunits a and b with the polar domain of subunit b extending from the center 45.

All eight subunits are required for E. coli ATP synthase 187. Chloroplast and cyanobacterial ATP synthase contain nine different subunits and fifteen subunits have been identified in the animal mitochondrial enzyme complex 141, 187. ATP synthase has been shown to exist in a dimeric state in yeast and mammalian mitochondria 158, 159 while monomeric ATP synthase has been found in Pa. denitrificans, Acetobacter woodii and spinach chloroplasts 159.

The F0 and F1 components of the ATPase enzyme complex are connected through two stalks: a central stalk formed by the connection between F1 subunits and

and the c9-12 ring of F0 and a peripheral one, that links the / hexagon of F1 to the F0 23

subunit via the hydrophilic b subunits 136, 205. The and subunits are proposed to reside in the stalk structure, subunit is required for F1 binding and cross-links to the

22, 45 , and subunits of F1 and the c subunit ring of F0 .

The F1 subunit has six alternating nucleotide-binding sites, three catalytic and three non-catalytic, on the hexagon formed by three / dimers 53, 136. The catalytic sites, designated DP, TP and E 1, have different catalytic site conformations that switch their affinities at a fixed step in the catalytic sequence 45, 48, 187. The three subunits exhibit cooperativity, positive in catalysis and negative in nucleotide binding, and demonstrate differing interactions with the central -subunit helices 48, 53.

Experimental results suggest that all three catalytic sites are occupied under physiological conditions (ATP> 1mM) 48, 187 and that the E catalytic site is where Pi must initially bind for ATP synthesis 1. Only one catalytic site is active at a time 53, 187 but it is necessary for the other catalytic sites to have bound substrates for product release from the high affinity site 187. The major energy requiring step is not the synthesis of ATP at catalytic sites, but rather the simultaneous and highly cooperative binding of substrates to, and the release of products from, these sites 37, 48, 187.

ATP synthesis or hydrolysis by F0F1 is a combination of catalysis, mechanical work ( -subunit rotation and torque generation) and proton transport 48, 135. ATP synthesis or hydrolysis is coupled to proton transport by the mechanical rotation of the

53, 56 33, 117, 125 - -c assembly extending through F1F0 and takes place at the F1 module .

187 Proton binding to a critical carboxyl group of subunit c in F0 that drives the rotation 24

of the - -c assembly rotation triggering conformational and affinity changes in the catalytic sites on the subunits that results in the release of the ATP product 22, 37, 48, 135,

142.

In coupled conditions, the ATPase enzyme complex can work reversibly as a synthase or hydrolase of ATP, consuming or pumping protons respectively 44, 121, 136.

Overall, the complete membrane-bound enzyme complex catalyzes the reaction 33, 114:

+ + ATP + H20 + 2H in ADP + Pi + 2H out

The stoichiometry of proton translocation through the H+-ATPase was initially determined to be equal to 2-3 but is now considered to be equal to 4 in plant chloroplasts 45, 59. The shift in stoichiometry may be due to the diversity of test material (chloroplasts, mitochondria, bacteria), techniques (steady-state vs. pH or salt jumps) and experimental approaches. The H+-ATP stoichiometry was reported to be 3 in rat-liver mitochondria 59, 120, 165, 1 in the plasma membrane of the fungus Neurospora

143 143 crassa , and 2 in the F0F1 ATPase of E. coli .

ATP provides the energy required for cell synthesis 112 and serves as a link between substrate oxidation and biomass synthesis 24. Anabolic pathways involve the use of free energy to build molecules required by the cell 105 for growth or other energy linked reactions 24. Energy transfer between the two pathways is in the form of ATP

105. Uncouplers of oxidative phosphorylation are thought to act as proton conductors, 25

dissipating the proton gradient across the membrane and decreasing the magnitude of

P 144 that drives the production of ATP in the cell membranes 66, 203 without appreciably affecting the electron transfer mediators 175. Oxidation of the substrate still occurs but the phosphorylation of ADP to ATP is reduced and less energy is available for the formation of biomass 202.

The effectiveness of various uncouplers is correlated with their capacity to act as proton ionophores 68 126 170 and catalyze electrogenic transfer of protons across the membrane 114 203. As a result of the proton and membrane potential components of the

P, the interior of a coupling membrane has a higher pH and a more negative electric potential than the exterior 113. The electrogenic transfer of protons introduces positive electrical charges into the interior of the cell forming a potential (locally positive) opposite that of the membrane potential 98, 191. This results in the destruction of both the pH gradient and membrane potentials which are203 required for complete metabolic uncoupling 33.

The mechanism of H+ translocation mediated by a weakly acidic uncoupler is illustrated in Figure 7 107, 175, 176.

26

FIGURE 7: THE MOLECULAR MECHANISM OF H+ TRANSLOCATION MEDIATED BY A PROTONOPHORIC UNCOUPLER 107, 175, 176 (ADAPTED FROM )

A- A-

H+ H+

HA HA + — Membrane

The original Mitchell model, proposed in 1966, assumes that the cell membrane is impermeable to most ions and the weak acid uncouplers are soluble in the membrane in both the neutral and anionic forms 107, 175, 176. The anions (A-) move from right to left through the interior of the membrane driven by the higher negative interior potential. This increase in the concentration of anions at the left-hand interface drives the heterogenous reaction:

- + A if + H HAif

- An anion at the interface (A if) combines with a proton from the aqueous phase

+ (H ) to form the undissociated acid (HAif ). The neutral form of the uncoupler diffuses through the membrane to the interior surface, driven by the concentration gradient of the compound across the membrane. As the acid form encounters the more alkaline 27

environment at the interior interface, the acid deprotonates and releases a proton into the cytoplasm. The anion then returns electrophoretically to the positively charged outside from the negatively charged inside of the energized membrane. The net result of the uncoupler cycle is the transport of H+ into the cell and dissipation of both the H+ gradient and membrane potential resulting in uncoupling 107, 175, 176.

Operationally, this results in a dramatic reduction in the P/O ratio; i.e. the moles of inorganic phosphate taken up per atom of oxygen consumed, while respiration, and therefore oxygen consumption, continues in the presence of substrate 66, 144, 176.

During uncoupling, a larger fraction of the substrate consumed in respiration is in support of proton translocation to maintain the proton gradient rather than the production of ATP. It is important to note that the synthesis of biomass is reduced or completely inhibited and energy-linked reactions such as active transport of ions are suppressed 175.

A wide variety of compounds are known to be uncouplers of oxidative phosphorylation, a listing of the most common uncouplers is provided in Table 1.

Uncouplers are, in general, moderately weak acids with an acid dissociation constant

176 (pKa) in the range or 5-7 and at physiological pH the compounds will exist in both the neutral (HA) and the anionic (A-) forms 61. Uncouplers generally contain an acidic hydrogen group, an electron-withdrawing moiety (-NO2, -CF3, -CN) and bulky hydrophobic group(s) 66, 176 which impart lipid solubility to the compounds, thought to be of primary importance for induction of uncoupling 175, 176.

28

TABLE 1: CLASSIC UNCOUPLERS OF RESPIRATION (ADAPTED 144 FROM )

2,4- Dinitrophenol CCCP (carbonyl cyanide m-chlorophenyl hydrazone) FCCP (carbonyl cyanide-p-trifluromethoxyphenyl hydrazone) p-Nitrophenol Dicoumarol (3,3'-methylene bis-4-hydroxycoumarin) Salicylanilides TCS (Tetrachlorosalicylanilide) S 13 (2',5-dichloro-3-tert-butyl 4'-nitrosalicylanilide) Benzimidazoles TTFB (4,5,6,7-tetrachloro-2'-trifluoromethyl-benzimidazole)

The most frequently studied uncoupler has been 2,4-dinitrophenol (DNP) 134, which was reported to have an effect on oxidative phosphorylation long before its mode of action was understood 68. DNP was shown to inhibit ATP synthesis in rabbit kidney preparations by Loomis and Lipmann, in 1948 96, stimulate the removal of the organic fraction of synthetic sewage by Rich and Yates in 1955 152, temporarily inhibit metabolic activity but not uncouple growth of Psuedomonas putida in batch cultures by

Low and Chase in 1996 97 and increase the rate of glucose consumption and decrease biomass production by 70% in continuous cultures of Saccharomyces cerevisia in 1975 by Shah et al 164. Mitchell, in 1961, after observing that the addition of dinitrophenol equilibrated the pH gradient across bacterial and mitochondrial membranes 68, 114 proposed that dinitrophenol and other ―uncoupler‖ compounds shuttle protons across the cell membranes 114 short-circuiting the energy producing mechanisms. 29

Uncoupler studies in denitrification systems have focused primarily on the

inhibition rather than the stimulation of respiration. Several weak acid-type uncouplers

and their analogues have been reported to uncouple aerobic but inhibit nitrogen oxide

respirations in microbial cultures. Carbonyl cyanide m-chlorophenylhydrazone

(CCCP) resulted in the uncoupling of aerobic respiration in Pseudomonas denitrificans

and Pseudomonas aeruginosa while CCCP, carbonyl cyanide p-

trifluoromethoxyphenylhydrazone (FCCP) and DNP inhibited the reductions of nitrate,

nitrite and nitrous oxide in Ps. denitrificans, Ps. aeruginosa 186 and Pa. denitrificans 87.

The rank order of the uncouplers as inhibitors of dentrification was found to be similar

to that for oxidative phosphorylation inhibition, release of respiratory control

(uncoupling), or increase in proton permeability of membranes, i.e. CCCP FCCP >

DNP 186. The addition of CCCP to denitrifying cultures of Pa.denitirificans was found

- to cause a lag phase in NO3 reduction but not in membrane vesicles with nitrate

reductases facing the outer bulk medium 13 and it was suggested that uncouplers

- eliminated the P-dependent transport of NO3 across the membrane.

Application of Chemical Uncouplers to Activated Sludge Processes

Biological treatment by the activated sludge process has become the major

technology for treating both municipal and industrial wastewaters worldwide 203.

Previously, the volume of wastewater treated, maximum substrate removal, and

effluent quality were the main criteria for the activated sludge process. However, 30

recently the rising cost for the treatment and disposal of excess sludge produced and restrictive legislation have emphasized the need to minimize biomass generated from activated sludge operations 203. The concept of uncouplers for biomass reduction in activated sludge systems is not new, and while there is a large amount of uncoupler literature, reports on their application in activated sludge systems are limited.

In 1955, Rich and Yates 152, in their pioneering work with uncouplers in activated sludge, found the addition of 5 and 15mg/L DNP increased the substrate uptake rate and inhibited growth in activated sludge units while the addition of higher concentrations of DNP resulted in a reversible decrease in the rate of substrate uptake.

Shah, et al 164 employing yeast to metabolize sugary wastes in a continuous waste treatment system found an 85% increase in degradation rate and 70% decrease in cell growth when the aqueous concentration of DNP was 5 x 10-6M (0.92 mg/L). A material balance on carbon utilization confirmed that most of the glucose consumed was degraded aerobically to carbon dioxide and water with negligible amounts going for cell growth or production of acids and alcohols.

Recently, the bulk of the literature has focused on the use of chemical uncouplers such as DNP, para-nitrophenol (p-NP), pentachlorophenol (PCP), chlorophenols (CP) and 3,3‘,4‘,5-tetrachlorosalicylanilide (TCS) in activated sludge cultures. Mayhew and Stephenson 105 investigated the effect of eight chemical inhibitors on the oxygen uptake rate of activated sludge cultures obtained from a municipal wastewater treatment facility. There was no difference between the control and experimental oxygen uptake rate between 0 and 25 hrs, after which the oxygen 31

uptake rate increased in the DNP and rotenone samples. Supplementation of lab-scale activated sludge simulations with 35 mg/L DNP resulted in decreased biomass production rates without an adverse effect on process efficiency. The average yield coefficient was 0.42 and 0.30 for the control and DNP simulation, respectively, while the biochemical oxygen demand (BOD) removal was unaffected by the addition of

DNP.

Okey and Stensel 133, in 1993, reported that the addition of halogenated phenols, such as 2,4-dichlorphenol (DCP), to unacclimmated cultures derived from an activated sludge and soil inoculum, resulted in a 25 to 45% increase in oxygen consumption when tested at ratios of 0.02 to 0.08 mg DCP/mg MLSS. In Warburg studies studies, where the MLVSS concentration ranged from 500 to 3000mg/L, the ratio of uncoupler to MLVSS concentration was found to correlate better with the extent of uncoupler activity than the concentration of uncoupler alone 133.

In batch studies with Psueudomonas putida, Low and Chase 97 found that the addition of p-nitrophenol (p-NP) inhibited biomass production without adversely affecting the removal of organic material. The observed biomass yields were found to decrease as the influent pH decreased in continuous studies while the specific substrate uptake rate increased in the presence of p-NP. The biomass reduction increased from

62% to 77% as the pH was decreased from 7.0 to 6.2 in feeds supplemented with 100 mg/L p-NP. In a bench scale activated sludge unit, the biomass production rate was decreased 49% in the presence of 100mg/L p-NP, but the total substrate removal rate was reduced by 25%. This decline in overall performance was attributed to the design 32

of the activated sludge unit resulting in solids removal in the final effluent and a decrease in the reactor biomass concentration 98.

Yang 199 reported that substituted phenols (p-chlorophenol, m-chlorophenol, m- nitrophenol, and o-nitrophenol) were effective in reducing biomass production in batch activated sludge studies. The acidity constant (pKa) of the metabolic uncoupler highly influenced its effectiveness in reducing biomass production and that lower pKas resulted in higher decreases in the biomass production. m-Chlorophenol was the most effective in reducing sludge production and had the least effect on process performance than the other compounds in the study. The addition of 20 mg/L of m-chlorophenol resulted in an 86.9% sludge reduction while the COD removal efficiency was only decreased by 13.2% as compared to the control test.

Strand and Stensel 172 compared eleven chemical uncouplers for their effectiveness in batch studies using activated sludge inoculums. The addition of

20mg/l CCCP and FCCP resulted in a reduction in biomass production of 41% and

33%, respectively while DNP did not have an effect on biomass production. 2,4,5-

Trichlorophenol (TCP) and 4-chloro-2-nitrophenol (CNP) were the most effective compounds tested and resulted in a 50% reduction in biomass when tested at concentrations of 3 and 4 mg/L, respectively. In completely mixed activated sludge

(CMAS) units, TCP functioned as an uncoupler when added to the influent at 10 mg/L.

When tested at lower concentrations (2.5 mg/L), TCP was degraded to levels below that required for uncoupling and long-term (80 days) exposure led to the development of a microbial population that were resistant to the uncoupling effect of TCP. 33

Uncoupling by TCP was more effective in a plug-flow activated sludge unit when tested at 2mg/L. Yields decreased from 0.4 g volatile suspended solids (VSS)/g chemical oxygen demand (COD) to 0.3 g VSS/g COD while the specific oxygen uptake rate (SOUR) increased from 8 to 16 mg O2/gVSS/hr.

In batch activated sludge cultures, Harem 60 found that DNP reduced the relative yields at concentrations of 40 and 80 mg/l but had little or no effect at the lower test concentrations of 10 and 20 mg/l 173. The most effective uncoupler found during these screening studies were CNP and TCP, which reduced the biomass production 68% and 73% relative to the controls when tested at a concentration of

20mg/L. In subsequent studies, TCP was tested in CMAS systems treating simulated municipal wastewater. Initially, TCP reduced the average yield by 50% and SOUR increased from 8 O2/gVSS/hr for the control reactor to approximately 20 O2/gVSS/hr.

After long-term exposure to the uncoupler, the biomass yield in the TCP reactor increased and effluent TCP levels decreased indicating the biomass became resistant to the uncoupler compound.

Chen et. al. 23 studied the feasibility of using TCS as a metabolic uncoupler to reduce biomass production in batch and continuous activated sludge cultures. TCS was found to be an effective chemical agent limiting biomass production when tested at concentrations > 0.4 mg/L. TCS reduced biomass production by 40% when added to continuous cultures daily to achieve a test concentration of 0.8 mg/L. Substrate removal efficiency and effluent nitrogen concentration were not adversely affected by the presence of TCS. 34

Ye et al. 201 studied the effect of o-chlorophenol (oCP), DCP, TCS, pNP and

DNP on biomass production in activated sludge cultures. The uncouplers were tested at a concentration of 20 mg/L in batch screening studies and resulted in reductions in biomass productions of 42.8 to 69.1% relative to the control culture. The most effective uncouplers were DNP, p-NP and TCS which resulted in biomass reductions of 69.1%, 63.7% and 56.1%. TCS reduced biomass production in a laboratory-scale

CMAS by 30% when tested at a concentration of 40mg/day 202. The substrate removal capability was not adversely affected by the presence of TCS, but effluent nitrogen concentration increased during the 60-day continuous operation. The SOUR of the culture was increased in the presence of TCS but the sludge settability from the test reactor was comparable to that of the control. Subsequent studies by Ye and Li 203 investigated the effects of three uncoupled metabolic systems: conventional activated sludge with the addition of TCS, an oxic-settling-anaerobic (OSA) process and the

TCS/OSA combined process on the biomass production rate. TCS was an effective chemical uncoupler reducing the biomass production by 34% compared to the control but had an adverse effect on the effluent nitrogen concentration and sludge settability when tested at 50mg/L for the first 3 months and 100 mg/L in the last 3 months. The

OSA process decreased biomass production by 26% and did not have an adverse effect on substrate removal or sludge settability. The most effective system was the combined process, which resulted in a 46.9% decrease in biomass production.

While there is an abundance of literature on the topic of the activity of uncouplers, there is just a handful of investigations into the use or activity of 35

uncouplers in biological waste treatment systems. Although there have been reports on the ―evidence‖ or the ―likelihood‖ of uncoupling in full-scale treatment systems, most of the information is speculative and lacks a solid scientific investigation. Okey and

Stensel 133, 134 observed behavior indicative of uncoupling at two full-scale activated sludge systems but could not detect any organic uncouplers at a high enough concentration to produce these effects. The lack of information on the use of uncouplers at the pilot or full scale is at best discouraging. Chemical uncouplers may result in a reduction in both the capital and operating costs of biological waste treatment systems by significantly improving the capacity of biological waste treatment systems and simultaneously decreasing the biomass production rates. The need for efficient, cost-effective and environmentally sound compound for waste treatment in both the municipal and industrial markets warrants the further investigation of the use of chemical uncouplers for biomass reduction.

36

References

1. Ahmad, Z. and A.E. Senior, Identification of Phosphate Binding Residues of Escherichia coli ATP Synthase. Journal of Bioenergetics and Biomembranes, 2005. 37(6): p. 437-440.

2. Alberts, B., et al., Molecular Biology of the Cell. Second Edition ed. 1989, New York: Garland Publishing, Inc. 1219.

3. Alefounder, P.R. and S.J. Ferguson, The Location of Dissimilatory Nitrite Reductase and the Control of Dissimilatory Nitrate Reductase by Oxygen in Paracoccus denitrificans. Biochem. J., 1980. 192: p. 231-240.

4. Alefounder, P.R., et al., Selection and Organisation of Denitrifying Electron- Transfer Pathways in Paracoccus denitrificans. Biochimica et Biophysica Acta, 1983. 724: p. 20-39.

5. Anraku, Y., Bacterial Electron Transport Chains. Annual Review of Biochemistry, 1988. 57: p. 101-32.

6. Boogerd, F.C., H.W. van Verseveld, and A.H. Stouthamer, Electron Transport to Nitrous Oxide in Paracoccus denitrificans. FEBS Letters, 1980. 113(2): p. 279- 284.

7. Boogerd, F.C., H.W. Van Verseveld, and A.H. Stouthamer, Respiration-Driven Proton Translocation with Nitrite and Nitrous Oxide in Paracoccus denitrificans. Biochimica et Biophysica Acta, 1981. 638(2): p. 181-191.

8. Boogerd, F.C., K.J. Appeldoorn, and A.H. Stouthamer, Effects of electron transport inhibitors and uncouplers on denitrification in Paracoccus denitrificans. FEMS Microbiol. Lett., 1983. 20(3): p. 455-60. 37

9. Boogerd, F.C., H.W. Van Verseveld, and A.H. Stouthamer, Dissimilatory nitrate uptake in Paracoccus denitrificans via a .DELTA.~.mu.H+-dependent system and a nitrate-nitrite antiport system. Biochim. Biophys. Acta, 1983. 723(3): p. 415- 27.

10. Boogerd, F.C., et al., Reconsideration of the efficiency of energy transduction in Paracoccus denitrificans during growth under a variety of culture conditions. A new approach for the calculation of P/2e- ratios. Arch. Microbiol., 1984. 139(4): p. 344-50.

11. Bryan, B.A., Physiology and Biochemistry of Denitrification, in Denitrification, Nitrification, and Atmospheric Nitrous Oxide, C.C. Delwiche, Editor. 1981, John Wiley & Sons: New York. p. 67-84.

12. Calder, K., K.A. Burke, and J. Lascelles, Induction of Nitrate Reductase and Membrane Cytochromes in Wild Type and Chlorate-Resistant Paracoccus denitrificans. Archives of Microbiology, 1980. 126: p. 149-153.

13. Capaldi, R.A. and B. Schulenberg, The subunit of bacterial and chloroplast F1F0 ATPases Structure, arrangement, and role of the subunit in energy coupling withing the complex. Biochimica et Biophysica Acta, 2000. 1458: p. 263-269.

14. Chen, Y.X., F.X. Ye, and X.S. Feng, The use of 3,3',4',5-tetrachlorosalicylanilide as a chemical uncoupler to reduce activated sludge yield. Journal of Chemical Technology and Biotechnology, 2004. 79: p. 111-116.

15. Chudoba, P., J. Chudoba, and B. Capdeville, The Aspect of Energetic Uncoupling of Microbial Growth in the Activated Sludge Process-OSA System. Water Science & Technology, 1992. 26(9-11): p. 2477-2480.

16. Constable, S.W., Solids disposal costs. 1995, Personal Communication.

17. Cox, C.D. and W.J. Payne, Separation of Soluble Denitrifying Enzymes and Cytochromes from Pseudomonas perfectomarinus. Canadian Journal of Microbiology, 1973. 19: p. 861-872. 38

18. Cox, R.B. and J.R. Quayle, The Autotrophic Growth of Micrococcus denitrificans on Methanol. Biochemical Journal, 1975. 150: p. 569-571.

19. Craske, A. and S.J. Ferguson, The Respiratory Nitrate Reductase from Paracoccus denitrificans. Journal of Biochemistry, 1986. 158: p. 429-436.

20. Cross, R.L., The Mechanism and Regulation of ATP Synthesis by F1-ATPases. Biochemistry, 1981. 50: p. 681-714.

21. Das, A. and L.G. Ljungdahl, Composition and Primary Structure of the F1F0 ATP Synthase from the Obligately Anaerobic Bacterium Clostridium thermoaceticum. Bacteriology, 1997. 179(11): p. 3746-3755.

22. Delwiche, C.C., Denitrification, in Inorganic Nitrogen Metabolism: Function of Metallo-Flavoprotein, W.D. McElroy and B. Glass, Editors. 1956, John Hopkins Press: Baltimore. p. 233-256.

23. Duncan, T.M., Rotation of subunits during catalysis by Escherichia coli F1- ATPase. Proceeding of the National Academy of Sciences, 1995. 92(10964- 10968).

24. Ferguson, S. The Redox Reactions of the Nitrogen and Sulphur Cycles. in Nitrogen and Sulphur Cycles Symposium 42. 1988: Society for General Microbiology.

25. Ferguson, S.J., Denitrification: a question of the control and organization of electron and ion transport. Trends Biochem. Sci., 1987. 12(9): p. 354-7.

26. Ferraro, D.J., L. Gakhar, and S. Ramaswamy, Rieske business: Structure-function of Rieske non-heme oxygenases. Biochemical and Biophysical Research Communications, 2005. 338: p. 175-190.

27. Fillingame, R.H., Biochemistry and Genetics of Bacterial H+ - Translocating ATPases. Bioenergetics, 1981. II: p. 35-106. 39

28. Fillingame, R.H., Membrane Sectors of F- and V-type H+-transporting ATPases. Current Opinion in Structural Biology, 1996. 6: p. 491-498.

29. Futai, M., Reconstitution of ATPase Activity From the Isolated , , and Subunits of the Coupling Factor, F1, of Escherichia coli. Biochemical and Biophysical Research Communications, 1977. 79(4): p. 1231.

30. Futai, M., et al., Synthase (H+ ATPase): coupling between catalysis, mechanical work and proton translocation. Biochimica et Biophysica Acta, 2000. 1458: p. 276-288.

31. Gaudy Jr, A.F. and E.T. Gaudy, Mixed Microbial Populations. Advances in Biochemical Engineering, 1972: p. 97-143.

32. Gogol, E.P., et al., Ligand-dependent structural variations in Escherichia coli F1 ATPase revealed by cryoelectron microscopy. Proceeding of the National Academy of Sciences, 1990. 87: p. 9585-9589.

33. Groth, G., Molecular Models of the Structural Arrangement of Subunits and the Mechanism of Proton Translocation in the Membrane Domain of F1F0 ATP synthase. Biochimica et Biophysica Acta, 2000. 1458: p. 417-427.

34. Haddock, B.A. and C.W. Jones, Bacterial Respiration. Bacteriological Reviews, 1977. 41(1): p. 47-99.

35. Haraux, F. and Y. De Kouchkovsky, Energy Coupling and ATP Synthase. Photosynthesis Research, 1998. 57(3): p. 231-251.

36. Harem, G.N., Investigation of Chemical Uncouplers to Reduce Activated Sludge Yield, in Civil Engineering. 1995, University of Wahington.

37. Harold, F.M., Conservation and Transformation of Energy by Bacterial Membranes. Bacteriological Reviews, 1972. 36(2): p. 172-230.

38. Heytler, P.G., Uncouplers of Oxidative Phosphorylation. Methods in Enzymology, 1979. LV: p. 462-473. 40

39. Hinkle, P.C. and R.E. McCarty, How Cells Make ATP. Scientific American, 1978. 238: p. 104-123.

40. Hochstein, L.I. and G.A. Tomlinson, The Enzymes Associated with Denitrification. Annual Review of Microbiology, 1988. 42: p. 231-261.

41. Jassal, B., Electron Transport Chain. 2005 http://www.reactome.org/cgi- bin/eventbrowser?DB=gk_current&ID=163200&ZOOM=2.

42. John, P. and F.R. Whatley, Paracoccus denitrficans and the Evolutionary Origin of the Mitochondrion. Nature, 1975. 254: p. 495-498.

43. John, P., Aerobic and Anaerobic Bacterial Respiration Monitored by Electrodes. Journal of General Microbiology, 1977. 98: p. 231-238.

44. John, P. and F.R. Whatley, The Bioenergetics of Paracoccus denitrificans. Biochemica et Biophysica Acta, 1977. 463: p. 129-153.

45. Jones, R.W., A. Lamont, and P.B. Garland, The Mechanism of Proton Translocation Driven by the Respiratory Nitrate Reductase Complex of Escherichia coli. Biochem. J., 1980. 190: p. 79-94.

46. King, M.W., Oxidative Phosphorylation. 2006, Indiana State University School of Medicine.

47. Knowles, R., Denitrification. Microbiological Reviews, 1982. 46(1): p. 43-70.

48. Kotlyar, A.B., Albracht, Simon P.J., and R.J.M. van Spanning, Comparison of Energizatin of Complex I in Membrane Particles from Paracoccus denitrificans and Bovine Heart Mitochondria. Biochimica et Biophysica Acta, 1998. 1365: p. 53-59.

49. Kristjansson, J.K., B. Walter, and T.C. Hollocher, Respiration-Dependent Proton Translocation and the Transport of Nitrate and Nitrite in Paracoccus denitrificans and Other Denitrifying Bacteria. Biochemistry, 1978. 17(23): p. 5014-5019. 41

50. Lam, Y. and D.J.D. Nicholas, Nitrite reductase with cytochrome oxidase activity from Micrococcus denitrificans. Biochim. Biophys. Acta, 1969. 180(3): p. 459- 72.

51. Link, T.A., The role of the 'Rieske' iron sulfur protein in the hydroquinone oxidation (QP) site of the cytochrome bc1 complex. FEBS Letters, 1997. 412: p. 257-264.

52. Liu, Y. and J.-H. Tay, Strategy for Minimization of Excess Sludge Production from the Activated Sludge Process. Biotechnology Advances, 2001. 19: p. 97-107.

53. Loomis, W.F. and F. Lipmann, Reversible Inhibition of the Coupling Between Phosphorylation and Oxidation. Biological Chemistry, 1948. 173: p. 807-808.

54. Low, E.W. and H.A. Chase. The Use of Uncouplers to Inhibit Biomass Production in Wastewater Treatment. in Icheme Research Event. 1996. Europe: Institute of Chemical Engineers.

55. Low, E.W. and H.A. Chase, The Use of Chemical Uncouplers for Reducing Biomass Production During Biodegradation. Water Science and Technology, 1998. 37(4/5): p. 399-402.

56. Low, E.W. and H.A. Chase, Reducing Production of Excess Biomass During Wastewater Treatment. Water Research, 1999. 33(5): p. 1119-1132.

57. Maloney, P.C., Energy Coupling to ATP Synthesis by the Proton-Translocating ATPase. Journal of Membrane Biology, 1982. 67: p. 1-12.

58. Mayhew, M. and T. Stephenson, Biomass Yield Reduction: Is Biochemical Manipulation Possible Without Affecting Activated Sludge Process Efficiency? Water Science & Technology, 1998. 38(8-9): p. 137-144.

59. McEwan, A.G., et al., Nitrous Oxide Reduction by Members of the Family Rhodospirillaceae and the Nitrous Reductase of Rhodopseudomonas capsulata. Journal of Bacteriology, 1985. 164(2): p. 823-830. 42

60. McLaughlin, S.G. and J.P. Dilger, Transport of Protons Across Membranes by Weak Acids. Physiological Reviews, 1980. 60(3): p. 825-863.

61. Millis, N.F., Effect of Uncoupling Oxidative Phosphorylation on the Economics of the Activated Sludge Process. Chem. Eng. Common., 1986. 45: p. 135-144.

62. Mitchell, P. and J. Moyle, Stoichiometry of Proton Translocation Through the Respiratory Chain and Adenosine Triphosphatase Systems of Rat Liver Mitochondria. Nature, 1965. 208: p. 147-151.

63. Mitchell, P., Chemiosmotic Coupling in Oxidative and Photosynthetic Phosphorylation. Biological Reviews, 1966. 41: p. 445-502.

64. Mitchell, P. and J. Moyle, Proton Translocation Coupled to ATP Hydrolysis in Rat Liver Mitochondria. European Journal of Biochemistry, 1968. 4(4): p. 530- 539.

65. Mitchell, P. and J. Moyle, Estimation of Membrane Potential and pH Difference across the Cristae Membrane of Rat Liver Mitochondria. European Journal of Biochemistry, 1969. 7(4): p. 471-484.

66. Mitchell, P., Cation-Translocating Adenosine Triphosphate Models: How Direct is the Participation of Adenosine Triphosphate and its Hydrolysis Products in Cation Translocation? FEBS Letters, 1973. 33(3): p. 267-274.

67. Mitchell, P., A Chemiosmotic Molecular Mechanism for Proton-Translocating Adenosine Triphosphatases. FEBS Letters, 1974. 43(2): p. 189-194.

68. Moat, A.G. and J.W. Foster, Microbial Physiology. Second Edition ed. 1988, New York: John Wiley & Sons. 597.

69. Moreau, F. and J.D. De Virville, Stoichiometry of Proton Translocation Coupled to Substrate Oxidation in Plant Mitochondria. Plant Physiology, 1985. 77: p. 118- 123. 43

70. Muneyuki, E., et al., F0F1-ATP synthase: general structural features of "ATP- engine" and a problem on free energy transduction. Biochimica et Biophysica Acta, 2000. 1458: p. 467-481.

71. Namslauer, A., et al., Redox-Coupled Proton Translocation in Biological Systems: Proton Shuttling Cytochrome c Oxidase. Proceeding of the National Academy of Sciences, 2003. 100(26): p. 15543-15547.

72. Negrin, R.S., D.L. Foster, and R.H. Fillingame, Energy-transducing H+-ATPase of Escherichia coli. The Journal of Biological Chemistry, 1980. 255(12): p. 5643- 5648.

73. Neidhardt, F.C., J.L. Ingraham, and M. Schaechter, Physiology of the Bacterial Cell. 1990, Sunderland, Massachusetts: Sinauer Associates, Inc. Publishers.

74. Neijssel, O.M., The Effect of 2,4-Dinitrophenol on the Growth of Klebsiella aerogenes NCTC 418 in Aerobic Chemostat Cultures. FEMS Letters, 1977. 1: p. 47-50.

75. Nicholls, D.G. and S.J. Ferguson, Bioenergetics 2. 1992, New York: Academic Press. 255.

76. Okey, R.W. and H.D. Stensel, Uncouplers and Activated Sludge - The Impact on Synthesis and Respiration. Toxicological and Environmental Chemistry, 1993. 40: p. 235-254.

77. Okey, R.W. and H.D. Stensel, Uncouplers of Oxidative Phosphorylation. Modeling and Predicting Their Impact on Wastewater Treatment and the Environment, in ACS Symposium Series Emerging Technologies in Hazardous Waste Management V, D.W. Tedder and F.G. Pohland, Editors. 1995, American Chemical Society: Washington. p. 284-309.

78. Omote, H., et al., The -subunit rotation and torque generation in F1-ATPase from wild-type or uncoupled mutant Escherchia coli. Proceeding of the National Academy of Sciences, 1999. 96: p. 7780-7784. 44

79. Pacheco-Moises, F., et al., Sulfite Inhibits the F1F0-ATP Synthase and Activates the F1F0-ATP of Paracoccus dentrificans. Journal of Bioenergetics and Biomembranes, 2002. 34(4): p. 269-278.

80. Parsonage, D., A.J. Greenfield, and S.J. Ferguson, The high affinity of Paracoccus denitrificans cells for nitrate as an electron acceptor. Analysis of possible mechanisms of nitrate and nitrite movement across the plasma membrane and the basis for inhibition by added nitrite of oxidase activity in permeabilized cells. Biochim. Biophys. Acta, 1985. 807(1): p. 81-95.

81. Payne, W.J., Reduction of Nitrogenous Oxides by Microorganisms. Bacteriological Reviews, 1973. 37(4): p. 409-452.

82. Payne, W.J., Denitrification. TIBS, 1976: p. 220-222.

83. Pedersen, P.L., Transport ATPases: Structure, Motors, Mechanism and Medicine: A Brief Overview. Journal of Bioenergetics and Biomembranes, 2005. 37(6): p. 445-449.

84. Perez, J.A. and S.J. Ferguson, Kinetics of Oxidative Phosphorylation in Paracoccus denitrificans. 1. Mechanism of ATP Synthesis at the Active Sites(s) of F0F1-ATPase. Biochemistry, 1990. 29: p. 10503-10518.

85. Perlin, D.S., et al., H+/ATP Stoichiometry of Proton Pumps from Neurospora crassa and Escherichia coli. Archives of Biochemistry and Biophysics, 1986. 248(1): p. 53-61.

86. Phillips-Jones, M.K. and M.E. Rhodes-Roberts, Studies of Inhibitors of Respiratory Electron Transport and Oxidative Phosphorylation. Bacteriology, 1991. 27: p. 203-224.

87. Rich, L.G. and J. Oscar W. Yates, The Effect of 2,4-Dinitrophenol on Activated Sludge. Applied Microbiology, 1955. 3: p. 95-98.

88. Rocher, M., et al., Towards a Reduction in Excess Sludge Production in Activated Sludge Processes: Biomass Physicochemical Treatment and Biodegradation. Applied Microbiology and Biotechnology, 1999. 51: p. 883-890. 45

89. Sabbert, D., S. Engelbrecht, and W. Junge, Functional and idling rotatory motion with F1-ATPase. Proceeding of the National Academy of Sciences, 1997. 94: p. 4401-4405.

90. Sapshead, L.M. and J.W.T. Wimpenny, The Influence of Oxygen and Nitrate on the Formation of the Cytochrome Pigments of the Aerobic and Anaerobic Respiratory Chain of Micrococcus denitrificans. Biochimica et Bioophysica Acta, 1972. 267: p. 388-397.

91. Schagger, H., Respiratory Chain Supercomplexes. IUBMB Life, 2001. 52: p. 119- 128.

92. Schagger, H., Respiratory chain supercomplexes of mitochondria and bacteria. Biochimica et Biophysica Acta, 2002. 1555: p. 154-159.

93. Scholes, P.B. and L. Smith, Composition and Properties of the Membrane-Bound Respiratory Chain System of Micrococcus dentrificans. Biochimica et Biophysica ACTA, 1968. 153: p. 363-375.

94. Schultz, B.E. and S.I. Chan, Structures and Proton-Pumping Strategies of Mitochondrial Respiratory Enzymes. Annual Review of Biophysics and Biomolecular Structure, 2001. 30: p. 23-65.

95. Shah, P.M., et al., Enhancement of Specific Waste Water Treatment by the Uncoupler 2,4-dinitrophenol. Food Science, 1975. 40(2): p. 302-305.

96. Sholtz, K., I. Gorskaya, and A. Kotelnikova, The Stoichiometry of Proton Translocation through H+-ATPase of Rat-Liver Mirochondria. European Journal of Biochemistry, 1983. 136(1): p. 129-134.

97. Soboll, S., Regulation of Energy Metabolism in Liver. Bioenergetics and Biomembranes, 1995. 27(6): p. 571-582.

98. Steinrucke, P., et al., The Cyctochrome c Reductse/Oxidase Respiratory Pathway of Paracoccus denitrificans: Genetic and Functional Studies. Journal of Bioenergetics and Biomembranes, 1991. 23(2): p. 227-239. 46

99. Stockdale, M. and J. Sewyn, Effects of ring substitutions on the activity of phenols as inhibitors and uncouplers of mitochondrial respiration. European Journal of Biochemistry, 1971. 21: p. 565-574.

100. Stouthamer, A.H., Metabolic Regulation Including Anaerobic Metabolism in Paracoccus denitrificans. Journal of Bioenergetics and Biomembranes, 1991. 23(2): p. 163-185.

101. Strand, S.E. and H.D. Stensel, Sludge Yield Reduction Using Chemical Uncouplers. 1994, Electric Power Research: Seattle.

102. Strand, S.E., G.N. Harem, and H.D. Stensel, Activated-Sludge Yield Reduction Using Chemical Uncouplers. Water Environment Research, 1999. 71(4): p. 454- 458.

103. Terada, H., The Interaction of Highly Active Uncouplers with Mitochondria. Biochimica et Biophysica Acta, 1981. 639: p. 225-242.

104. Terada, H., Uncouplers of Oxidative Phosphorylation. Environmental Health Perspectives, 1990. 87: p. 213-218.

105. Trumpower, B.L., The Three-Subunit Cytochrome bc1 Complex of Paracoccus denitrificans. Its Physiological Function, Structure, and Mechanism of Electron Transfer and Energy Transduction. Journal of Bioenergetics and Biomembranes, 1991. 23(2): p. 241-255.

106. Urata, K. and T. Satoh, Mechanism of Nitrite Reduction to Nitrous oxide in a Photodenitrifier, Rhodopseudomonas sphaeroides forma sp. denitrificans. Biochimica et Biophysica Acta, 1985. 841: p. 201-207.

107. Van Spanning, R.J.M., et al., Regulation of oxidative phosphorylation: the flexible respiratory network of Paracoccus denitrificans. J. Bioenerg. Biomembr., 1995. 27(5): p. 499-512.

108. Vignais, P.M., et al., The electron transport system and hydrogenase of Paracoccus denitrificans. Curr. Top. Bioenerg, 1981. 12: p. 115-96. 47

109. Vik, S.B. and R.R. Ishmukhametov, Structure and Function of Subunit a of the ATP Synthase of Escherichia coli. Journal of Bioenergetics and Biomembranes, 2005. 37(6): p. 445-449.

110. Walter, B., et al., Inhibition of Denitrification by Uncouplers of Oxidative Phosphorylation. Biochemistry, 1978. 17(15): p. 3039-3045.

111. Weber, J. and A.E. Senior, Catalytic mechanism of F1-ATPase. Biochimica et Biophysica Acta, 1997. 1319: p. 19-58.

112. Wheatley, A.D., Effluents, in Biodeterioration 7, D.R. Houghton, R.N. Smith, and H.O.W. Eggins, Editors. 1988, Cranfield Institute of Technology: Bedford. p. 193-206.

113. White, D., The Physiology and Biochemistry of Prokaryotes. 1995, New York: Oxford University Press. 378.

114. Wiesmann, U., Biological Nitrogen Removal from Wastewater. Advances in Biochemical Engineering Biotechnology, 1994. 51: p. 113-154.

115. Yagi, T., Bacterial NADH-Quinone Oxidoreductases. Journal of Bioenergetics and Biomembranes, 1991. 23(2): p. 211-225.

116. Yagi, T., et al., NADH Dehydrogenases: From Basic Science to Biomedicine. Journal of Bioenergetics and Biomembranes, 2001. 33(3): p. 233-242.

117. Yagi, T. and A. Yagi-Matsuno, The Proton-Translocating NADH-Quinone Oxidoreductase in the Respiratory Chain: The Secret Unlocked. Biochemistry, 2003. 42: p. 2266-2274.

118. Yang, X.-F., M.-L. Xie, and Y. Liu, Metabolic Uncouplers Reduce Excess Sludge Production in an Activated Sludge Process. Process Biochemistry, 2003. 38: p. 1373-1377. 48

119. Yasuda, R., et al., Rotation of the Subunit in F1-ATPase; Evidence That ATP Synthase Is a Rotary Motor Enzyme*. Bioenergetics and Biomembranes, 1997. 29(3): p. 207-209.

120. Ye, F.X., D.S. Shen, and Y. Li, Reduction in Excess Sludge Production by Addition of Chemical Uncouplers in Activated Sludge Batch Cultures. Journal of Applied Microbiology, 2003. 95: p. 781-786.

121. Ye, F.X. and Y. Li, Reduction of Excess Sludge Production by 3,3',4',5- Tetrachlorosalicylanilide in an Activated Sludge Process. Applied Microbiology and Biotechnology, 2005. 67: p. 269-274.

122. Ye, F.X. and Y. Li, Uncoupled Metabolism Stimulated by Chemical Uncoupler and Oxic-Settling-Anaerobic Combined Process to Reduce Excess Sludge Production. Applied Biochemistry and Biotechnology, 2005. 127(3): p. 187-199.

123. Zharova, T.V. and A. Vinogradov, Energy-dependent Transformation of F0F1- ATPase in Paracoccus denitrificans Plasma Membranes. The Journal of Biological Chemistry, 2004. 279(13): p. 12319-12324.

124. Zhou, Y., T.M. Duncan, and R.L. Cross, Subunit rotation in Escherchia coli F0F1- ATP synthase during oxidative phosphorylation. Proceeding of the National Academy of Sciences, 1997. 94: p. 10583.

49

Chapter 3 The activity of 1-hydroxyanthraquinone on anaerobic respiration in Paracoccus denitrificans

Lynn Ann Arlauskas-Dekleva, J. Martin Odom, Jolanta Wilchek, JoAnn Payne, Eva Nagel, and Michael A. Gealt

Abstract

The activity of 1-hydroxyanthraquinone (OHAQ) on nitrogen respiration and biomass production was evaluated in anaerobic batch cultures of Paracoccus denitrificans using nitrate or nitrite as the terminal electron acceptor. The addition of 10 mg/L of OHAQ resulted in a significant increase in the nitrogen production rate and decrease in biomass production in batch denitrification studies. The activity of OHAQ was different than the other electron transport and ATPase inhibitors and uncouplers of oxidative phosphorylation evaluated in this test system. The hydroxyl group at position 1 of the anthraquinone ring system was required for uncoupling activity in the denitrifying system used in these studies. The uncoupling activity combined with the low toxicity of OHAQ suggests that this compound may be used to reduce both the capital and operating costs of commercial bioprocesses in addition to traditional biological wastewater treatment systems.

Introduction

The primary purpose of biological treatment processes is the removal of organic

material 51, thereby decreasing its environmental burden before its discharge. The 50

activated sludge process is the most common biological treatment process for domestic and industrial wastewater 97, 154 and has proven to be a reliable and efficient treatment technology capable of producing high-quality effluents 173. Wastewaters typically contain a complex mixture of components which are degraded by a mixed microbial population in biochemical reactions 99 that convert the organic material into new cell mass (biomass) and respiration products (carbon dioxide and water) 94, 98, 99. The excess biomass produced in the activated sludge process can be significant and requires additional processing and disposal 173. The costs associated with the handling and disposal of biomass from biological treatment processes can represent from 40 to 60% of a wastewater treatment plant operating costs 27, 99, 190 and may require significant capital expenditures.

The metabolism of organic wastes, in catabolism, provides the energy required for cell synthesis (anabolism) in the form of an energy rich molecule, ATP 24, 97, 112.

The relationship between ATP formation by catabolism and its use in biomass synthesis 24, shown in Figure 1, illustrates the relationship between the two processes.

51

FIGURE 1: SIMPLIFIED RELATIONSHIP BETWEEN CATABOLISM 24 AND ANABOLISM (ADAPTED FROM REFERENCE )

Biomass

Substrate ATP

Catabolism Anabolism

Substrate

CO + H O 2 2 ADP + Pi

Energy

In most organisms, the majority of the ATP is synthesized in a process known as respiration 127, 167 that utilizes a sequence of oxidation-reduction reactions to transfer electrons from the substrate to oxygen 60 or in the absence of oxygen to inorganic acceptors, e.g. nitrate 19. The electrons travel through a series of membrane-associated electron carrier proteins and lipids 191, known as the electron transport chain or respiratory network 181. The electron transport chain is localized in a membrane essentially impermeable to most ions, including OH(-) and H(+) 61. In prokaryotes, the membrane is associated with the outer cellular membrane. In eukaryotic systems the membrane is the inner membrane of the mitochondrion. The flow of electrons is coupled to the translocation of protons across the membrane that generates an electrical 52

or membrane potential ( ) and proton gradient ( pH) on the opposite sides of the membrane 40, 57, 61, 68, 114, 125.

The energy contained in the pH and known collectively as the electrochemical gradient or proton-motive force ( P) 57, 61, 68, 114-116, is recovered by the cell as the protons return across the membrane 103. The cell exploits this movement of

57, 61, 114-116 protons at the cell membrane to synthesize ATP from ADP and Pi in a reversible reaction 35, 45, 53, 187, 200. ATP provides the energy required for cell synthesis

112 and serves as a link between substrate oxidation and biomass synthesis 24.

One approach for decreasing biomass production involves the use of chemical compounds, known as uncouplers, which disrupt the direct energetic link between catabolism (oxidation of organic matter) and anabolism (ATP generation) 105, therefore limiting the cells ability to capture energy from substrate oxidation 98, 203. Uncouplers transport protons across biological membranes, dissipating the proton gradient across the membrane and decreasing the magnitude of P 49, 144. The rate of ATP synthesis and biomass production are reduced 66, 175, 203 while respiration, and therefore oxygen consumption, continues in the presence of substrate 66, 98, 144, 176.

Chemical uncouplers are a structurally diverse group of molecules 95 144. In general, they are moderately weak lipophilic acids with an acid dissociation constant

176 (pKa) in the range of 5-7 . Uncouplers generally contain an acidic hydrogen, an

66, 176 electron-withdrawing moiety (-NO2, -CF3, -CN) and bulky hydrophobic group(s) 53

and at physiological pH the compounds will exist in both the neutral (HA) and the anionic (A-) forms 61.

Laboratory studies have shown that the application of the uncoupler technology to activated sludge units can result in a significant reduction in the biomass production rate 93, 99, 101, 199, 201-203. The most active uncoupler compounds are usually the most toxic and a serious problem with the uncoupler technology in wastewater treatment is the potential presence of the compound in the effluent where it might be toxic to the environment. A number of uncoupler compounds, (e.g. DNP, chlorinated phenols) are designated as toxic pollutants subject to effluent limitations 63, 64; impeding their utility at a full-scale wastewater treatment facility.

Anthraquinones, are found widely dispersed in the environment 204 occurring as the result of both natural and commercial processes 7, 18, 50, 204. Anthraquinones, are an important class of compounds found as raw materials 25 in the textile 58, plastics, cosmetic, 204 and pharmaceutical 67, 70, 128, 174 industries. They exhibit a wide variety of biological effects 7, 9, 10, 16, 18, 29, 34, 67, 70, 80-82, 131, 132, 148, 149, 189 including inhibition of electron transport in chloroplasts 80, 132, uncoupling of mitochondrial 10, 16, 82 and sulfate respiration29 and inhibition of bacterial growth 7, 9, 34, 81 and respiration 29, 131, 189.

The uncoupling activity of anthraquinones in sulfate-reducing bacteria 29 and mitochondria 10 combined with their low systemic toxicity 62 suggest that these compounds may have some utility for biomass reduction in wastewater treatment systems. Laboratory studies have focused primarily on the application of anthraquinones for growth inhibition of aerobic and anaerobic microorganism 29, 50, 131, 54

189 rather than its potential use as an uncoupler. Several weak acid-type uncouplers and their analogues have been reported to uncouple aerobically but inhibit nitrogen oxide respiration in microbial cultures. The purpose of this study is to evaluate the effect of

1-hydroxyanthraquinone (OHAQ) on anaerobic batch cultures of Paracoccus denitrificans using nitrate as the terminal electron acceptor. The activity of OHAQ on nitrogen and biomass production is compared to respiratory inhibitors and uncouplers of oxidative phosphorylation.

Paracoccus was chosen as the model system because of its ability to grow aerobically or under anaerobic conditions with nitrate, nitrite, or nitrous oxide as the terminal electron acceptor 21, 77, 171, 191]. Formerly known as Micrococcus denitrificans

184, Pa. denitrificans is a gram negative 171, 181, non-fermenting facultative anaerobic bacteriuum 15, 57, 77, 184, commonly found in soil and sludge 181. The organism, in response to a reduction in the molecular oxygen concentration, synthesizes four enzyme complexes that catalyze the sequential reduction of nitrate, nitrite, nitric oxide, and nitrous oxide to nitrogen gas in a process known as denitrification 4, 19, 36, 84, 140, 181,

191, 192.

The overall conversion of nitrate to nitrogen gas proceeds through a series of four reactions represented by 4, 19, 36, 84, 140, 181, 191, 192:

¯ ¯ NO3 NO2 NO N2O N2

55

The anaerobic electron transport pathway includes several branches to the individual reductases 139, 191 and the denitrifying enzymes are ‗plugged‘ into the respiratory chain, shown in Figure 2, at the level of either ubiquinone or cytochrome c

42, 171.

FIGURE 2: ANAEROBIC ELECTRON TRANSPORT CHAIN IN 191 PARACOCCUS DENITRIFICANS MODIFIED FROM

- NO3 nitrate reductase

- NO2

dehydrogenase UQ bc1 c nitrite reductase NO nitric oxide reductase

1/2 N O nitrous oxide 2 reductase

1/2 N2

Ubiquinol transfers electrons directly to nitrate reductase, bypassing the bc1 complex and the c cytochromes 32, 76, 171, 191. Electron flow from ubiquinol to nitrite, nitrous, and nitric reductases proceeds via the cytochrome bc1 complex to cytochrome c 11, 19, 42, 77, 139, 171, 180, 181. There are two coupling sites in the anaerobic electron 56

transport pathway. Site one is the NADH dehydrogenase and site two is the bc1 complex.

The uncoupling activity of anthraquinones in sulfate-reducing bacteria 29 and mitochondria 10 combined with their low systemic toxicity 62 suggest that these compounds may have some utility for biomass reduction in wastewater treatment systems. The purpose of this study is to elucidate the effect of

1-hydroxyanthraquinone (OHAQ) on a model system consisting of Paracoccus denitrificans. The activity of the compound was monitored by nitrogen generation and biomass generation in batch anaerobic cultures. The site of activity was evaluated by using nitrate or nitrite as the terminal electron acceptor. 57

Materials and Methods

Organism and Growth Conditions

Paracoccus denitrificans (ATCC 19367) was obtained from the American Type

Culture Collection, Rockville, MD and maintained on Nutrient Agar or Tryptic Soy

Agar plates. The microbial culture was grown in BTZ-3 130 minimal medium at 35-

37°C, 200 RPM, in an environmental chamber (Barnstead Thermolyne Corporation,

Dubuque, IA) until late exponential or early stationary phase (approximately 18-24

hours).

In these denitrification studies, the medium was supplemented with 76 mM of

NaNO3 or 15 mM of NaNO2 and was purged with oxygen-free helium (Linde Gas,

Cleveland, OH) for a minimum of 15 minutes prior to inoculation. Paracoccus

cultures was adapted to anaerobic conditions by the transferring the culture to 120 ml

borosilicate glass serum bottles (Wheaton, Millville, NJ) with chlorobutyl rubber

stoppers (Wheaton) containing helium purged BTZ-3 medium. The adapted cultures

were transferred to fresh medium a minimum of two times prior to their use in

denitrification studies. In these studies, the nitrogen concentration of the headspace

was used to evaluate the effect of the test compounds on the overall denitrification

process and not on a particular step in the pathway; the concentration of the

intermediates such as nitrite was not determined.

58

Chemicals and Stock Solutions

1-hydroxyanthraquinone (OHAQ) (Aldrich 1,926-4, purity >97%) was a gift

courtesy of J. Martin Odom (DuPont, Wilmington, DE). The other test compounds

were obtained from the following commercial sources: 2,4-dinitrophenol (DNP)

(Fluka 42160, purity >99%), carbonyl cyanide m-chlorophenyl hydrazone (CCCP)

(Fluka 21855, purity ≥98%), carbonyl cyanide-p-trifluromethoxyphenyl hydrazone

(FCCP) (Fluka 21857, purity ≥99%), Tinopal (Sigma F6259, purity 90%), 7-chloro-

4-nitrobenzo-2-oxa-1,3-diazole (NBD-Cl) (Sigma C5261, purity >99%), N,N‘-

dicyclohexylcarbodiimide (DCC) (Sigma D3128, purity >95%), rotenone (Sigma R-

8875, purity 95-98%), antimycin (mixture of compounds A1, A2, A3 and A4, Fluka AG

CH-94 50626-785, purity 90%).

Stock solutions of test compounds were prepared in acetone (J.T. Baker JT

9010-1, purity 99.5%) due to their low aqueous solubility, with the exception of

Tinopal which was soluble in aqueous solutions. The stock solutions were prepared at

an approximate concentration of 1000 mg/L to ensure the acetone concentration did not

exceed 1% (V/V) in any of the experiments. The test compounds, evaluated at a test

concentration of 10 mg/L, were added to the test vessels using a 25G tuberculin syringe

(Becton Dickinson & Company, Franklin Lakes, NJ) to minimize introduction of

oxygen. 59

Experimental Procedures

Paracoccus cultures were adapted to growth under anaerobic conditions by the

transferring aerobically grown cells to serum bottles containing helium purged BTZ-3

medium supplemented with 76 mM of NaNO3 or 15 mM of NaNO2. The bottles were

incubated at 35-37 C, 200 RPM, in an environmental chamber. Late exponential or

early stationary phase (Abs590 3.40-3.80) cultures (approximately 72 hours) were

transferred to fresh medium containing either the test compound (at a concentration of

10 mg/L) or acetone for the control bottles. The nitrogen concentration of the

headspace was determined using a Hewlett Packer 5890 gas chromatograph (GC)

(Aligent, Fort Collins, CO) equipped with a thermal conductivity detector (TCD).

Oxygen and nitrogen were separated using a molecular sieve column (Supleco 5A,

30m x 0.53mm Catalog 2-5463) with ultra pure helium as the carrier gas. The carrier,

auxiliary and reference gas flow rates were maintained at 3, 7 and 15 ml/minute,

respectively. The temperature of the oven was maintained at 35°C and the injector and

detector temperatures were 230 C. Gas samples (200 l) manually injected into the

GC using gas tight syringes (Hamilton Company, Reno, NV) and the nitrogen

concentration was manually integrated from calibration curve.

Calibration standards were prepared by addition of ultrapure N2 (Linde Gas,

Cleveland, OH) to helium purged serum bottles containing BTZ-3 medium. The

calibration consisted of duplicate injections using a minimum of eight data points 60

ranging from 37.2 to 2232 nMoles N2/ injection. Correlation coefficients for the calibration curves were generally > 0.99.

The growth of the Paracoccus cultures was monitored at 590 nm using a Hach

DR/3000 (Loveland, CO) or Pharmacia LKB UV-Vis (Thermo Fisher Scientific

Waltham, MA) spectrophotometer. The relationship between Abs590 and dry weight was determined and found to be linear up to an OD of approximately 2.4. Biomass concentrations were determined by filtration on 0.2 or 0.45 M pre-weighed filters.

The filters were dried to constant weight in a 100 C oven after washing twice with deionized water.

61

Results and Discussion

Comparison of 1-Hydroxyanthraquinone with respiratory inhibitors and uncouplers of oxidative phosphorylation on denitrifying cultures of Paracoccus denitrificans

The effect of respiratory inhibitors and uncouplers of oxidative phosphorylation

on anaerobic respiration was evaluated in batch culture studies using Pa.denitrificans.

The compounds used in the studies (Table 1) included several inhibitors of electron

transport and ATPase as well as the highly active uncouplers of oxidative

phosphorylation, CCCP (carbonyl cyanide m-chlorophenyl hydrazone) and FCCP

(carbonyl cyanide-p-trifluromethoxyphenyl hydrazone). Rotenone 46, 71, 110, 160, 184 and

Tinopal AN, (1,1-bis- (3,N-5-dimethylbenzoxazol-2-yl)-methine p-toluene sulphonate)

144, 145 directly inhibit NADH denhydrogenase while antimycin A inhibits the

79, 144, 162, 166 respiratory chain electron transport between cytochromes b and c1 .

Dicyclohexylcarbodiimide (DCC) 44, 103 and 7-chloro-4-nitrobenzo-2-oxa-1, 3-diazole

(NBD-Cl) 41, 76, 92 specifically react with ATPase and inhibit both the hydrolysis and

synthesis of ATP.

62

TABLE 1: RESPIRATORY INHIBITORS AND UNCOUPLERS OF OXIDATIVE PHOSPHORYLATION STUDIED IN PA. DENITRIFICANS UNDER DENITRIFYING CONDITIONS WITH NITRATE AS THE ELECTRON ACCEPTOR

Inhibitor Active Site Electron Transport Inhibitors Rotenone Iron-suphur centre N-2 of NADH dehydrogenase

Tinopal AN After cluster N-2 of NADH dehydrogenase

Antimycin A Cytochrome bc 1

ATPase Inhibitors NBD-Cl (7-chloro-4-nitrobenzo-2-oxa-1,3-diazole) Modification of phenolic oxygen groups of an essential

tyrosine residue in F1

DCC (N,N'-dicyclohexylcarbodiimide)

F0 component in membrane,

Uncouplers CCCP (Carbonyl cyanide m-chlorophenyl hydrazone) FCCP (Carbonyl cyanide-p-trifluromethoxyphenyl hydrazone) DNP (2,4-Dinitrophenol)

The addition of all test compounds with the exclusion of OHAQ and DNP resulted in the partial or complete inhibition of respiratory activity in denitrifying cultures of Pa. denitrificans (Table 2). The lag phase in the rotenone cultures under denitrifying conditions was significantly longer than the 1 to 2 hour lag reported in Pa. denitrificans growing aerobically on succinate after which growth was restored with a 63

maximum specific growth rate ( max) which was 40% of that before rotenone addition

109, 146. The electron transport inhibitors, rotenone, Tinopal or antimycin A resulted in the immediate inhibition of nitrogen production (Table 2). Nitrogen production was partially restored in the rotenone cultures after a 68 hour lag phase, but the culture did not fully recover by the termination of the experiment (130 hours). The average nitrogen production rate of the rotenone cultures was 0.86 Moles N2/hr, 20% the rate measured in the control bottles.

TABLE 2: NITROGEN PRODUCTION (MEAN MMOLES N2/BOTTLE) OF PA. DENITRIFICANS WITH 10 MG/L OHAQ, ELECTRON TRANSPORT INHIBITORS (ROTENONE, NBD-CL, AND ANTIMYCIN), ATPASE INHIBITORS (DCC, TINOPAL) AND UNCOUPLERS OF OXIDATIVE PHOSPHORYLATION (DNP, FCCP AND CCCP).

Mean Moles N2/bottle Time Acetone Control OHAQ DNP FCCP CCCP 4 11.5 14.1 14.1 14.5 74.9 22 26.8 15.0 15.9 17.0 24.0 45 39.8 187 169 29.9 25.6 68 139 403 408 22.2 61.6 130 401 414 417 81.7 110.8

Time H2O Control Tinopal Rotenone Antimycin DCC NBD-Cl 4 11.8 15.7 16.2 18.2 19.8 42.4 22 24.5 27.1 16.4 21.5 19.6 22.1 45 59.8 69.0 22.9 23.7 26.0 25.7 68 125 52.4 42.2 25.2 44.1 56.1 130 397 57.2 95.7 42.1 102 105

64

The nitrogen production in the bottles receiving Tinopal or antimycin was completely inhibited at a test concentration of 10 mg/L and the cultures did not recover by the termination of the experiment at 130 hours. Similar results have been reported for aerobically grown Pa. denitrificans 147 where the addition of 1, 5 and 10 M reduced and 20 M Tinopal immediately inhibited growth of the organism. The lack of nitrogen production in cultures containing 10 mg/L antimycin may be the result of the complete inhibition of electron transfer between cytochrome bc1 and cytochrome c or the inhibition of denitrification by intermediates, such as nitrite, in the pathway.

Concentrations significantly above 10 mM nitrite are inhibitory to denitrifying microorganisms 72 and in particular toxic to Paracoccus 108, therefore, the conversion of nitrate to nitrite in the first step of the pathway might result in inhibition in the growth and respiration of the test organism. In these studies, the nitrogen concentration of the headspace was used to evaluate the effect of test compounds on the overall denitrification process and not on a particular step in the pathway; the concentration of the intermediates such as nitrite was not determined.

The ATPase inhibitors, DCC or NBD-Cl, produced effects similar to the electron transport inhibitors and significantly decreased the rate of nitrogen production in the batch anaerobic cultures of Pa. denitrificans (Table 2). The mean nitrogen production rate in the DCCD and NBD-Cl bottles at a test concentration of 10 mg/L was 0.79 Moles N2/hr and 1.33 Moles N2/hr, respectively, 20 to 30 % of the 4.4

Moles N2/hr production rate for the control bottles. The nitrogen production rate remained relatively constant in the DCCD and NBD-Cl bottles over the duration of the 65

experiment and at 130 hours (the termination of the experiment) the bottles had produced approximately 25% of the total nitrogen production of the control bottles.

The activity of OHAQ was compared to uncouplers of oxidative phosphorylation; CCCP (Carbonyl cyanide m-chlorophenyl hydrazone), FCCP

(Carbonyl cyanide-p-trifluromethoxyphenyl hydrazone) and DNP (2,4-Dinitrophenol).

The addition of OHAQ or DNP resulted in the stimulation and CCCP or FCCP the inhibition in the nitrogen production in denitrifying Pa. dentificans (Tables 2 and 3,

Figure 3). Thus, the addition OHAQ resulted in classical uncoupling behavior characterized by a decrease in biomass production and increase in the respiration rate.

At a test concentration of 10 mg/L, OHAQ addition increased the nitrogen production rate 45.7 % (p= 2.29 x 10-2) and decreased the biomass production 24.5 % (p=1.33 x

10-3) compared to the control bottles. The mean nitrogen production rate was increased from 4.4 Mole N2/hr for the control bottles to 8.2 Mole N2/hr for the OHAQ bottles without affecting the overall nitrogen production. The biomass production at the termination of the experiment was decreased 24.5 % from 1,765 mg/L for the control bottles to 1,333 mg/L for OHAQ.

66

FIGURE 3: THE EFFECT OF OHAQ AND UNCOUPLERS ON NITROGEN PRODUCTION IN ANAEROBIC CULTURES OF PA. DENITRIFICANS WITH NITRATE AS THE TERMINAL ELECTRON ACCEPTOR

500

/bottle 400 2 300

200 Moles N 100

0 Mean Mean 4 22 45 68 130 Time, hours

Control OHAQ DNP FCCP CCCP

Mean nitrogen production from batch anerobic studies using Pa. denitrificans with nitrate as the terminal electron acceptor and acetone (!) (in the control vessels), OHAQ (!), DNP (!), FCCP (!) or CCCP (!) at an initial test concentration of 10 mg/L. Results are the mean of three replicate test vessels.

67

TABLE 3: THE EFFECT OF 1-HYDROXYANTHRAQUINONE (OHAQ) AND 2,4-DINITROPHENOL (DNP) ON BIOMASS PRODUCTION AND NITROGEN RESPIRATION RATE IN ANAEROBIC CULTURES OF PA. DENITRIFICANS WITH NITRATE AS THE TERMINAL ELECTRON ACCEPTOR.

Mean Moles % Nitrogen t-Test Mean Biomass Biomass t-Test p

N2/hr Increase p value mg/L Reduction % value Control 4.4 1765 OHAQ 8.2 46.7 2.29E-02 1333 24.5 1.33E-03 DNP 10.5 58.4 1.97E-02 1833 -3.9 3.66E-01

In this study, DNP, at a concentration of 13 mg/L, significantly increased the

nitrogen production rate (p=1.97 x 10-2) but, did not decrease the biomass production

rate (Table 3). The mean nitrogen production rate in the DNP bottles was increased

58.4% from the average control production rate of 4.4 Mole N2/hr to 10.5 Mole

N2/hr. As observed in the OHAQ test bottles, the overall nitrogen production in the

control and DNP bottles was comparable indicating that the test compounds did not

inhibit the metabolism of the organism.

Effect of 1-Hydroxyanthraquinone, 9,10-Anthraquinone and 2,4-Dinitrophenol on anaerobic cultures of Paracoccus denitrificans using nitrate or nitrite as the terminal electron acceptor

Nitrate or nitrite was used as electron acceptors to determine if the

anthraquinones interacting specifically with the portion of the electron transport chain 68

involving nitrate reductase. The electrons for nitrate reductase are transferred directly

171, 191 from ubiquinol while electrons for the other reductases are supplied via the bc1 complex and cytochrome c 181. If the anthraquinones specifically interact with nitrate reductase or the transfer of electrons from ubiquinol to nitrate reductase, the cultures would not display the same behavior when nitrite only was supplied. Concentrations significantly above 10mM nitrite are inhibitory to denitrifying microorganisms 72 and in particular toxic to Paracoccus 182, therefore, the level of nitrite in the medium was reduced from the nitrate test concentration of 76 mM to approximately 15 mM to decrease the potential inhibitory effects.

The addition OHAQ to a denitrifying culture of Pa. denitrificans with nitrate as the terminal electron acceptor resulted in classic uncoupling behavior characterized by a decrease in biomass production and increase in the respiration rate. At a test concentration of 10 mg/L, OHAQ addition increased the nitrogen production rate

-3 -5 64.5% (p= 6.66 x 10 ) and decreased the biomass production 30.7% (p=2.06 x 10 ) compared to the control bottles (Table 4, Figure 4). The mean nitrogen production rate was increased from 4.2 Mole N2/hr for the control bottles to 11.8 Mole N2/hr for the

OHAQ bottles without affecting the overall nitrogen production. The biomass production at the termination of the experiment was decreased 30.7 % from 1,900mg/L for the control bottles to 1,317 mg/L for OHAQ. The increase in respiration rate was not observed in cultures treated with the parent compound, 9,10-anthraquinone (AQ), which lacks the hydroxyl group in position 1, but at a test concentration of 10 mg/L the biomass production was reduced by 9.2% (p=7.68 x 10-3). In this study, 2,4- 69

dinitrophenol (DNP), at a concentration of 13 mg/L significantly increased the nitrogen production rate (p=2.18 x 10-5) but as observed in previous studies did not affect the biomass production rate. The mean nitrogen production rate in the DNP bottles was increased from the mean control production rate of 4.2 Mole N2/hr to 14.6 Mole

N2/hr. As in the OHAQ test bottles, the overall nitrogen production in the control and

DNP bottles was similar.

TABLE 4: THE EFFECT OF 1-HYDROXYANTHRAQUINONE (OHAQ), 9,10-ANTHRAQUINONE (AQ) AND 2,4-DINITROPHENOL (DNP) AT A TEST CONCENTRATION OF 10 MG/L ON BIOMASS PRODUCTION AND NITROGEN RESPIRATION RATE IN ANAEROBIC CULTURES OF PA. DENITRIFICANS WITH NITRATE AS THE ELECTRON ACCEPTOR.

70

FIGURE 4: THE EFFECT OF 1-HYDROXYANTHRAQUINONE (OHAQ), 9,10-ANTHRAQUINONE (AQ) AND 2,4-DINITROPHENOL (DNP) AT A TEST CONCENTRATION OF 10 MG/L ON NITROGEN PRODUCTION IN ANAEROBIC CULTURES OF PA. DENITRIFICANS WITH NITRATE AS THE ELECTRON ACCEPTOR.

500 /bottle

2 400 300

200 Moles N 100 0

Mean Mean 0 23 48 72 Time, hours

Con OHAQ AQ DNP

Mean nitrogen production from batch anerobic studies using Pa. denitrificans with nitrate as the terminal electron acceptor and acetone (") (for the control vessels), OHAQ (!), AQ (#) or DNP (%) at an initial test concentration of 10 mg/L. Results are the mean of three replicate test vessels.

The behavior of OHAQ was similar in the both the nitrite and nitrate studies.

The addition of OHAQ to a denitrifying culture of Pa. denitrificans using nitrite as the electron acceptor resulted in a significant increase in the rate of nitrogen production and a 29% decrease in the biomass production compared to the control bottles (Figure

5, Table 5). During the 24 to 48 hour time period, the OHAQ bottles, at a test concentration of 10 mg/L, produced an average of 2.9 Moles N2/hr and the nitrogen production was complete at the 72 hour time point. The biomass production was 71

decreased from 767 mg/L in the control bottles to 544 mg/L (p= 5.84 x 10-4). In contrast, the control bottles exhibited a significant lag in nitrogen production and at 24 hours the cultures had a mean nitrogen production rate of 0.05 Moles N2/hr. After the

96 hour lag phase, the nitrogen production rate in the control bottles increased to 1.8

Moles N2/hr, but the rate never achieved the production rate seen in the OHAQ bottles. The unsubstituted anthraquinone, AQ, had similar nitrogen production rates as the control bottles and did not decrease the biomass production. The mean nitrogen production rate in the AQ bottles was 0.1 Moles N2/hr between 24 and 48 hours and increased to 2.0 Moles N2/hr after the 72 hour sampling interval. The lag in nitrogen production was similar to that seen in the control bottles. 2,4-Dinitrophenol produced similar results to those obtained in the nitrate study. DNP significantly increased the nitrogen production rate but did not result in a significant reduction in the biomass production. The nitrogen production rate during the 24 to 48 hour time period was an average of 2.6 Moles N2/hr and was complete by the 72 hour sampling point (Figure

4). DNP did result in 14% reduction of the biomass, from a mean control value of 767 mg/L to 658 mg/L, but this reduction was not significant (p=1.91 x 10-1). At the termination of the nitrite experiment, the total nitrogen production was similar in the control and experimental bottles.

72

TABLE 5: THE EFFECT OF 1-HYDROXYANTHRAQUINONE (OHAQ), 9,10-ANTHRAQUINONE (AQ) AND 2,4-DINITROPHENOL (DNP) AT A TEST CONCENTRATION OF 10 MG/L ON BIOMASS PRODUCTION AND NITROGEN RESPIRATION RATE IN ANAEROBIC CULTURES OF PA. DENITRIFICANS WITH NITRITE AS THE ELECTRON ACCEPTOR.

Mean Mean Moles % Nitrogen Biomass Biomass t-Test p

N2/hr Increase mg/L Reduction % value Control 0.05 767 OHAQ 2.9 98.2 544 29.0 5.84E-04 AQ 0.1 62.2 825 -7.6 3.52E-01 DNP 2.6 97.9 658 14.1 1.91E-01

73

FIGURE 5: THE EFFECT OF 1-HYDROXYANTHRAQUINONE (OHAQ), 9,10-ANTHRAQUINONE (AQ) AND 2,4-DINITROPHENOL (DNP) AT A TEST CONCENTRATION OF 10 MG/L ON NITROGEN RESPIRATION IN ANAEROBIC CULTURES OF PA. DENITRIFICANS WITH NITRITE AS THE ELECTRON ACCEPTOR.

160 /bottle 2 120

80

Moles N 40

0

Mean Mean 20 70 120 170 Time, hours

Con OHAQ AQ DNP

Mean nitrogen production from batch anerobic studies using Pa. denitrificans with nitrite as the terminal electron acceptor and acetone (") (for the control vessels), OHAQ (!), AQ (#) or DNP (%) at an initial test concentration of approximately 10 mg/L. Results are the mean of three replicate test vessels.

In anaerobic cultures using either nitrate or nitrite as the terminal electron acceptor, the addition of OHAQ resulted in the uncoupling of metabolism as evidence by the increased respiration rate and decreased biomass production. The lack of uncoupling activity with the unsubstituted parent compound, AQ, may be the result of the lack of a hydroxyl group on the ring system. The addition of a hydroxyl functional group to AQ has been shown to increase its activity in mitochondria 183 and 74

chloroplasts 132 and that the position of the functional group or number of functional groups affects the activity. The effect of ring substitutions on the uncoupling efficiency of phenolic compounds has been reported in a variety of systems. Clowes &

Krahl 26, 86, in 1936, noted the location of the substitution in nitro- and halogenated phenols directly affected uncoupling activity in Arbacia eggs and that the presence of a phenolic hydroxyl group was essential for biological activity. Betina and Kuzela 10 found that anthraquinone compounds bearing a free hydroxyl group at the -position exhibited uncoupling behavior in rat liver mitochondria and compounds, lacking a - hydroxyl group, were ineffective. This requirement for a -hydroxyl group in uncoupling was confirmed by Kawai et al 82 who found that 1,2-DiOH, which contains a -hydroxyl group, was the only compound that exhibited significant uncoupling behavior.

In aerobic studies with Pa. denitrificans, this laboratory has demonstrated the activity of the compound was related to both the positional and electronic properties of the substitution on the anthraquinone ring system (Chapter 4). Electron donating groups (EDG) in position 1 (ortho) of the aromatic ring system resulted in a larger biomass reduction than substitutions in either position 2 (meta) or multiple (ortho, meta and para) substitutions. Electron donating groups in position 1 were not inhibitory to

Paracoccus and did not adversely affect the aerobic respiration of the cultures. In the group of compounds tested, a hydroxyl group in position 1 had a larger effect on biomass reduction than an amino group substitution, although both groups resulted in 75

significant biomass reductions. Position 2 EDGs did not result in significant biomass reductions and were, in general, inhibitory to the culture.

The substitution results in the aerobic studies suggest that the difference in the activity of the compounds with substitutions at positions 1 or 2 may be the result of different electron density of the compounds. EDGs have at least one pair of non- bonded electrons on the hetroatom next to the aromatic ring structure 102 which donate electrons onto the ring structure 28. Substitutions in position 1 donate electrons to the ring structure according to the proposed mechanism 1, Figure 6, and result in an electrically neutral compound with partial negative and partial positive charges in close proximity 65. While substitutions in position 2 donate electrons according to the proposed mechanism 2, Figure 6, resulting in the generation of a zwitterion 65, a compound which is electrically neutral but carries formal positive and negative charges

193. 76

FIGURE 6: DISTRIBUTION OF ELECTRONS IN SUBSTITUTED ANTHRAQUINONE COMPOUNDS WITH ELECTRON DONATING GROUPS

Mechanism 1

- + O OH O OH

O O

Mechanism 2

O O

OH OH+

- O O

A difference in the electron distribution may interfere with diffusion of the

compound across or within the membrane and/or the release of protons. The activity of

the substituted anthraquinone compounds utilized in this study indicates that the charge

distribution of the compounds plays a key role in their ability to reduce biomass

production in microbial systems. 77

The activity OHAQ was unlike the respiratory inhibitors, CCCP or FCCP; that partially or completely inhibited respiration in the test organism. Uncouplers have been reported to inhibit nitrate, nitrite and nitrous oxide reductions in intact cells 87, 89,

186, inhibit ATP synthesis in membrane vesicles 137, and eliminate the P-dependent

- 13, 155, 195 transport of NO3 across the membrane . In these studies, DNP stimulated the nitrogen production rate but did not significantly decrease the biomass production compared to the controls. The DNP concentration used in the current studies was significantly lower than the concentration reported to inhibit denitrification in

Psuedomonas denitrificans 186.

While uncoupler studies in denitrification systems have focused primarily on the inhibition rather than the stimulation of respiration by uncouplers, these studies demonstrate that OHAQ uncouples denitrification in anaerobic cultures of

Pa.denitrificans. The uncoupling of denitrification was not sensitive to the terminal electron acceptor used in the study; the compound was active in cultures using either nitrate or nitrite as the terminal electron acceptor. The compound has activity distinctly different from the respiratory inhibitors or uncouplers of oxidative phosphorylation used in the study and anthraquinones have an advantage over other uncouplers of oxidative phosphorylation because of their low systemic toxicity 62, 153. Miscellaneous inhibitory effects on particular enzyme systems have been reported, but the overall lack of toxicity of anthraquinones is supported by their natural occurrence in plants, their widespread use as dyes and their use until recently as laxatives 131. The uncoupling activity combined with the low toxicity of compounds suggests that these compounds 78

may have the potential for biomass reductions in commercial bioprocesses in addition to traditional wastewater treatment systems. 79

References

1. Alefounder, P.R., et al., Selection and Organisation of Denitrifying Electron- Transfer Pathways in Paracoccus denitrificans. Biochimica et Biophysica Acta, 1983. 724: p. 20-39.

2. Anke, H., et al., Metabolic Products of Microorganisms. 185*. The Anthraquinones of the Aspergillus glaucus Group. I. Occurrence, Isolation, Identification and Antimicrobial Activity. Archives of Microbiology, 1980. 126: p. 223-230.

3. Bakola-Christianopoulou, M.N., L.B. Ecateriniadou, and K.J. Sarrris, Evaluation of the Antimicrobial Activity of a New Series of Hydroxy-Quinone Chelates of Some Transition Metals. European Journal of Medicinal Chemistry - Chim. Ther, 1986. 21(5): p. 385-390.

4. Betina, V. and S. Kuzela, Uncoupling effect of fungal hydroxyanthraquinones on mitochondrial oxidative phosphorylation. Chem.-Biol. Interact., 1987. 62(2): p. 179-89.

5. Boogerd, F.C., H.W. van Verseveld, and A.H. Stouthamer, Electron Transport to Nitrous Oxide in Paracoccus denitrificans. FEBS Letters, 1980. 113(2): p. 279- 284.

6. Boogerd, F.C., K.J. Appeldoorn, and A.H. Stouthamer, Effects of electron transport inhibitors and uncouplers on denitrification in Paracoccus denitrificans. FEMS Microbiol. Lett., 1983. 20(3): p. 455-60.

7. Boogerd, F.C., et al., Reconsideration of the efficiency of energy transduction in Paracoccus denitrificans during growth under a variety of culture conditions. A new approach for the calculation of P/2e- ratios. Arch. Microbiol., 1984. 139(4): p. 344-50. 80

8. Boos, K.S. and E. Schlimme, Anthraquione Dyes: A New Class of Potent Inhibitors of Mitochrondrial Adenine Nucleotide Translocation and Oxidative Phosphorylation. FEBS Letters, 1981. 127(1): p. 40-44.

9. Brown, J.P., A Review of the Genetic Effects of Naturally Occurring Flavonoids, Anthraquinones and Related Compounds. Mutation Research, 1980. 75: p. 243- 277.

10. Bryan, B.A., Physiology and Biochemistry of Denitrification, in Denitrification, Nitrification, and Atmospheric Nitrous Oxide, C.C. Delwiche, Editor. 1981, John Wiley & Sons: New York. p. 67-84.

11. Calder, K., K.A. Burke, and J. Lascelles, Induction of Nitrate Reductase and Membrane Cytochromes in Wild Type and Chlorate-Resistant Paracoccus denitrificans. Archives of Microbiology, 1980. 126: p. 149-153.

12. Chudoba, P., J. Chudoba, and B. Capdeville, The Aspect of Energetic Uncoupling of Microbial Growth in the Activated Sludge Process-OSA System. Water Science & Technology, 1992. 26(9-11): p. 2477-2480.

13. Chung, R.H., Anthraquinone, in Kik-Othmer Enclycopedia of Chemical Technology. 1978, John Wiley & Sons: New York. p. 700-757.

14. Clowes, G.H.A. and M.E. Krahl, Studies on Cell Metabolism and Cell Division I. On the Relation between Molecular Structures, Chemical Properties, and Biological Activities of the Nitrophenols. Journal of General Physiology, 1936. 20: p. 145-171.

15. Constable, S.W., Solids disposal costs. 1995, Personal Communication.

16. Cook, S.R., Steve's place-Reaction Mechanisms. 2002, http://www.steve.gb.com/science/reaction_mechanisms.html.

17. Cooling III, F.B., et al., Inhibition of Sulfate Respiration by 1,8- Dihydroxyanthraquinone and Other Anthraquinone Derivatives. Applied and Environmental Microbiology, 1996. 62(8): p. 2999-30004. 81

18. Craske, A. and S.J. Ferguson, The Respiratory Nitrate Reductase from Paracoccus denitrificans. Journal of Biochemistry, 1986. 158: p. 429-436.

19. Cudlin, J., et al., Biological activity of hydroxyanthraquinones and their glucosides toward microorganisms. Folia Microbiol., 1976. 21(1): p. 54-7.

20. Das, A. and L.G. Ljungdahl, Composition and Primary Structure of the F1F0 ATP Synthase from the Obligately Anaerobic Bacterium Clostridium thermoaceticum. Bacteriology, 1997. 179(11): p. 3746-3755.

21. Delwiche, C.C., Denitrification, in Inorganic Nitrogen Metabolism: Function of Metallo-Flavoprotein, W.D. McElroy and B. Glass, Editors. 1956, John Hopkins Press: Baltimore. p. 233-256.

22. Ferguson, S. The Redox Reactions of the Nitrogen and Sulphur Cycles. in Nitrogen and Sulphur Cycles Symposium 42. 1988: Society for General Microbiology.

23. Ferguson, S.J., et al., Selective and Reversible Inhibition of the ATPase of Micrococcus denitrificans by 7-Chloro-4-nitrobenzo-2-oxa-1,3 diazole. Biochimica et Biophysica Acta, 1974. 357: p. 457-461.

24. Ferguson, S.J., Denitrification: a question of the control and organization of electron and ion transport. Trends Biochem. Sci., 1987. 12(9): p. 354-7.

25. Fillingame, R.H., Biochemistry and Genetics of Bacterial H+ - Translocating ATPases. Bioenergetics, 1981. II: p. 35-106.

26. Fillingame, R.H., Membrane Sectors of F- and V-type H+-transporting ATPases. Current Opinion in Structural Biology, 1996. 6: p. 491-498.

27. Fukami, H. and M. Nakajima, Rotenone and the Rotenoids, in Naturally Occurring Insecticides, M. Jacobson and D.G. Crosby, Editors. 1971, Marcel Dekker, Inc.: New York. p. 71-97. 82

28. Gage, D.J. and F. Neidhardt, Adaptation of Escherichia coli to the Uncoupler of Oxidative Phosphorylation 2,4 - Dinitrophenol. Journal of Bacteriology, 1993. 175(21): p. 7105-7108.

29. Garcia-Lopez, P.M., L. Kung Jr, and J.M. Odom, In Vitro Inhibition of Microbial Methane Production by 9,10-Anthraquinone. Journal of Animal Science, 1996. 74: p. 2276-2284.

30. Gaudy Jr, A.F. and E.T. Gaudy, Mixed Microbial Populations. Advances in Biochemical Engineering, 1972: p. 97-143.

31. Gogol, E.P., et al., Ligand-dependent structural variations in Escherichia coli F1 ATPase revealed by cryoelectron microscopy. Proceeding of the National Academy of Sciences, 1990. 87: p. 9585-9589.

32. Haddock, B.A. and C.W. Jones, Bacterial Respiration. Bacteriological Reviews, 1977. 41(1): p. 47-99.

33. Hall, N., Anthraquinone dyes containing a fluorosulfonyl group and their use. 1995, Zeneca Ltd., UK: PCT Int. Appl. p. 17.

34. Harem, G.N., Investigation of Chemical Uncouplers to Reduce Activated Sludge Yield, in Civil Engineering. 1995, University of Wahington.

35. Harold, F.M., Conservation and Transformation of Energy by Bacterial Membranes. Bacteriological Reviews, 1972. 36(2): p. 172-230.

36. Health, N.I.o., Hazardous Substances Databank Trichlorophenol. 2000, National Library of Medicine's TOXNET System.

37. Health, N.I.o., Hazardous Substances Databank Anthraquinone. 2000, National Library of Medicine's TOXNET System.

38. Health, N.I.o., Hazardous Substances Databank Dinitrophenol. 2000, National Library of Medicine's TOXNET System. 83

39. Hennessy, M.J., The activity of electron donating substitutions on aromatic structures. 2007, Personal Communication.

40. Heytler, P.G., Uncouplers of Oxidative Phosphorylation. Methods in Enzymology, 1979. LV: p. 462-473.

41. Hilgert, I., et al., Antitumor and immunosuppressive activity of hydroxyanthraquinones and their glucosides. Folia Biol., 1977. 23(2): p. 99-109.

42. Hinkle, P.C. and R.E. McCarty, How Cells Make ATP. Scientific American, 1978. 238: p. 104-123.

43. House, R.V. and J.H. Dean, Structural requirements for suppression of cytotoxic T lymphocyte induction by polycyclic compounds. In Vitro Toxicol., 1987. 1(3): p. 149-62.

44. Imai, K., A. Asano, and R. Sata, Oxidative phosphorylation in Micrococcus denitrificans I. Preparation and properties of phosphorylating membrane fragments. Biochim. Biophys. Acta, 1967. 143: p. 462-476.

45. Ingraham, J.L., Microbiology and Genetics of Denitrifiers, in Denitrification, Nitrification, and Atmospheric Nitrous Oxide, C.C. Delwiche, Editor. 1981, Wiley-Interscience Publication: Davis. p. 45-65.

46. John, P., Aerobic and Anaerobic Bacterial Respiration Monitored by Electrodes. Journal of General Microbiology, 1977. 98: p. 231-238.

47. John, P. and F.R. Whatley, The Bioenergetics of Paracoccus denitrificans. Biochemica et Biophysica Acta, 1977. 463: p. 129-153.

48. Kaniuga, Z., J. Bryla, and E.C. Slater, Mechanism of Interaction between Antimycin and the Respiratory Chain. FEBS Symposium, 1969. 19: p. 285-295.

49. Karukstis, K.K., et al., Alternative Measures of Photosystem II Electron Transfer Inhibition in Anthraquinone-Treated Chloroplasts. Photochemistry and Photobiology, 1992. 55(1): p. 125-132. 84

50. Kavanagh, F., Activities of Twenty-Two Antibacterial Substances Against Nine Species of Bacteria. Journal of Bacteriology, 1947. 54: p. 761-766.

51. Kawai, K., et al., The inhibition of mitochondrial respiration by anthraquinone mycotoxins. The relationship between the position of hydroxyl group on the anthraquinone nucleus and the uncoupling effect on oxidative phosphorylation. Proc. Jpn. Assoc. Mycotoxicol., 1987. 26: p. 43-6.

52. Knowles, R., Denitrification. Microbiological Reviews, 1982. 46(1): p. 43-70.

53. Krahl, M.D. and G.H.A. Clowes, Studies on Cell Metabolism and Cell Division II. Stimulation of Cellular Oxidation and Reversible Inhibition of Cell Division by Dihalo and Trihalophenols. Journal of General Physiology, 1936. 20: p. 173-184.

54. Kristjansson, J.K., B. Walter, and T.C. Hollocher, Respiration-Dependent Proton Translocation and the Transport of Nitrate and Nitrite in Paracoccus denitrificans and Other Denitrifying Bacteria. Biochemistry, 1978. 17(23): p. 5014-5019.

55. Kucera, I. and P. Kaplan, A Study on the Transport and Dissimilatory Reduction of Nitrate in Paracoccus denitrificans using Viologen Dyes as Electron Donors. Biochimica et Biophysica Acta, 1996. 1276: p. 203-209.

56. Linnett, P.E. and R.B. Beechey, Inhibitors of the ATP Synthetase System. Methods in Enzymology, 1979. LV: p. 472-519.

57. Liu, Y., Effect of Chemical Uncoupler on the Observed Growth Yield in Batch Culture of Activated Sludge. Water Research, 2000. 34(7): p. 2025-2030.

58. Liu, Y. and J.-H. Tay, Strategy for Minimization of Excess Sludge Production from the Activated Sludge Process. Biotechnology Advances, 2001. 19: p. 97-107.

59. Liu, Y., Chemically Reduced Excess Sludge Production in the Activated Sludge Process. Chemosphere, 2003. 50: p. 1-7. 85

60. Low, E.W. and H.A. Chase. The Use of Uncouplers to Inhibit Biomass Production in Wastewater Treatment. in Icheme Research Event. 1996. Europe: Institute of Chemical Engineers.

61. Low, E.W. and H.A. Chase, The Use of Chemical Uncouplers for Reducing Biomass Production During Biodegradation. Water Science and Technology, 1998. 37(4/5): p. 399-402.

62. Low, E.W. and H.A. Chase, Reducing Production of Excess Biomass During Wastewater Treatment. Water Research, 1999. 33(5): p. 1119-1132.

63. Low, E.W., et al., Uncoupling of Metabolism to Reduce Biomass Production in the Activated Sludge Process. Water Research, 2000. 34(12): p. 3204-3212.

64. Macmillan, J.G., Electrophilic Aromatic Substitution. 1999, http://faculty.cns.uni.edu/~macmilla/mcmurry/mcmurry_chapter_16/sld.

65. Maloney, P.C., Energy Coupling to ATP Synthesis by the Proton-Translocating ATPase. Journal of Membrane Biology, 1982. 67: p. 1-12.

66. Mayhew, M. and T. Stephenson, Biomass Yield Reduction: Is Biochemical Manipulation Possible Without Affecting Activated Sludge Process Efficiency? Water Science & Technology, 1998. 38(8-9): p. 137-144.

67. Meijer, E.M., et al., Energy Conservation during Aerobic Growth in Paracoccus denitrificans. Archives of Microbiology, 1977. 112: p. 25-34.

68. Meijer, E.M., et al., Effects Induced by Rotenone During Aerobic Growth of Paracoccus denitrificans in Continuous Culture. Microbiology, 1978. 119: p. 119-127.

69. Meijer, E.M., et al., Effects Induced by Rotenone during Aerobic Growth of Paracoccus denitrificans in Continuous Culture. Archives of Microbiology, 1978. 119: p. 119-127. 86

70. Millis, N.F., Effect of Uncoupling Oxidative Phosphorylation on the Economics of the Activated Sludge Process. Chem. Eng. Common., 1986. 45: p. 135-144.

71. Mitchell, P., Chemiosmotic Coupling in Oxidative and Photosynthetic Phosphorylation. Biological Reviews, 1966. 41: p. 445-502.

72. Mitchell, P. and J. Moyle, Proton Translocation Coupled to ATP Hydrolysis in Rat Liver Mitochondria. European Journal of Biochemistry, 1968. 4(4): p. 530- 539.

73. Mitchell, P. and J. Moyle, Estimation of Membrane Potential and pH Difference across the Cristae Membrane of Rat Liver Mitochondria. European Journal of Biochemistry, 1969. 7(4): p. 471-484.

74. Neidhardt, F.C., J.L. Ingraham, and M. Schaechter, Physiology of the Bacterial Cell. 1990, Sunderland, Massachusetts: Sinauer Associates, Inc. Publishers.

75. Nicholls, D.G. and S.J. Ferguson, Bioenergetics 2. 1992, New York: Academic Press. 255.

76. Nishio, A. and E.M. Uyeki, Cellular uptake and inhibition of DNA synthesis by dihydroxyanthraquinone and two analogs. Cancer Res., 1983. 43(5): p. 1951-6.

77. Odom, J.M., et al., Immunological Cross-Reactivities of Adenosine-5'- Phosphosulfate Reductases from Sulfate-Reducing and Sulfide-Oxidizing Bacteria. Applied and Environmental Microbiology, 1991. 57(3): p. 727-733.

78. Odom, J.M., Anthraquinone inhibition of methane production in methanogenic bacteria, in 42 pp. 1994, Bio-Technical Resources LP, USA: PCT Int. Appl.

79. Oettmeier, W., K. Masson, and A. Donner, Anthraquinone inhibitors of photosystem II electron transport. FEBS Letters, 1988. 231(1): p. 259-62.

80. Parsonage, D. and S.J. Ferguson, Reassessment of pathways of electron flow to nitrate reductase that are coupled to energy conservation in Paracoccus denitrificans. FEBS Lett., 1983. 153(1): p. 108-12. 87

81. Payne, W.J., Reduction of Nitrogenous Oxides by Microorganisms. Bacteriological Reviews, 1973. 37(4): p. 409-452.

82. Payne, W.J., Denitrification. TIBS, 1976: p. 220-222.

83. Phillips-Jones, M.K. and M.E. Rhodes-Roberts, Studies of Inhibitors of Respiratory Electron Transport and Oxidative Phosphorylation. Bacteriology, 1991. 27: p. 203-224.

84. Phillips, M.K. and D.B. Kell, A benzoxazole inhibitor of NADH dehydrogenase in Paracoccus denitrificans. FEMS Microbiology Letters, 1981. 11: p. 111-113.

85. Phillips, M.K. and D.B. Kell, A Novel Inhibitor of NADH Dehydrogenas in Paracoccus denitrificans. FEBS Letters, 1982. 104(2): p. 248-250.

86. Phillips, M.K., D.B. Kell, and M.E. Rhodes-Roberts, The Antibacterial Action of Tinopal. Journal of General Microbiology, 1984. 130: p. 1999-2005.

87. Powell, A.K., Growth-Inhibiting Action of Some Pure Substances. Nature, 1944. 153: p. 345.

88. Prestera, T., H.J. Prochaska, and Talalay, Inhibition of NAD(P)H:(Quinone- Acceptor) Oxidoreductase by Cibacron Blue and Related Anthraquinone Dyes: A Structure-Activity Study. Biochemistry, 1992. 31: p. 284-293.

89. Richardson, M.L. and S. Gangolli, eds. The Dictionary of Substances & Their Effects. 1994, The Royal Society of Chemistry.

90. Rocher, M., et al., Towards a Reduction in Excess Sludge Production in Activated Sludge Processes: Biomass Physicochemical Treatment and Biodegradation. Applied Microbiology and Biotechnology, 1999. 51: p. 883-890.

91. Rowe, J.J., et al., NarK is a nitrite-extrusion system involved in anaerobic nitrate respiration by Escherichia coli. Molecular Microbiology, 1994. 12(4): p. 579- 586. 88

92. Scholes, P.B. and L. Smith, Composition and Properties of the Membrane-Bound Respiratory Chain System of Micrococcus dentrificans. Biochimica et Biophysica ACTA, 1968. 153: p. 363-375.

93. Seddon, B. and J. McVittie, The Effect of Inhibitors on the Electron-transport chain of Bacillus brevis. Evidence for Branching of the NADH Oxidase Respiratory Chain. Journal of General Microbiology, 1974. 84: p. 386-390.

94. Singer, T.P., Mitochondrial Electron-Transport Inhibitors. Methods in Enzymology, 1979. LV: p. 454-463.

95. Soboll, S., Regulation of Energy Metabolism in Liver. Bioenergetics and Biomembranes, 1995. 27(6): p. 571-582.

96. Stouthamer, A.H., Metabolic Regulation Including Anaerobic Metabolism in Paracoccus denitrificans. Journal of Bioenergetics and Biomembranes, 1991. 23(2): p. 163-185.

97. Strand, S.E., G.N. Harem, and H.D. Stensel, Activated-Sludge Yield Reduction Using Chemical Uncouplers. Water Environment Research, 1999. 71(4): p. 454- 458.

98. Swanbeck, G., Interaction Between Deoxyribonucleic Acid and Some Anthracene and Anthraquinone Derivatives. Biochim. Biophys. Acta, 1966. 123: p. 630-633.

99. Terada, H., The Interaction of Highly Active Uncouplers with Mitochondria. Biochimica et Biophysica Acta, 1981. 639: p. 225-242.

100. Terada, H., Uncouplers of Oxidative Phosphorylation. Environmental Health Perspectives, 1990. 87: p. 213-218.

101. Urata, K. and T. Satoh, Mechanism of Nitrite Reduction to Nitrous oxide in a Photodenitrifier, Rhodopseudomonas sphaeroides forma sp. denitrificans. Biochimica et Biophysica Acta, 1985. 841: p. 201-207. 89

102. Van Spanning, R.J.M., et al., Regulation of oxidative phosphorylation: the flexible respiratory network of Paracoccus denitrificans. J. Bioenerg. Biomembr., 1995. 27(5): p. 499-512.

103. Van Verseveld, H.W., E.M. Meijer, and A.H. Stouthamer, Energy conservation during nitrate respiration in Paracoccus denitrificans. Arch. Microbiol., 1977. 112(1): p. 17-23.

104. Verhaeren, E.H.C., The Uncoupling Activity of Substituted Anthracene Derivative on Isolated Mitochondria from Phaseolus aureus. Phytochemistry, 1980. 19: p. 501-503.

105. Vignais, P.M., et al., The electron transport system and hydrogenase of Paracoccus denitrificans. Curr. Top. Bioenerg, 1981. 12: p. 115-96.

106. Walter, B., et al., Inhibition of Denitrification by Uncouplers of Oxidative Phosphorylation. Biochemistry, 1978. 17(15): p. 3039-3045.

107. Weber, J. and A.E. Senior, Catalytic mechanism of F1-ATPase. Biochimica et Biophysica Acta, 1997. 1319: p. 19-58.

108. Weimer, P.J., et al., Anthraquinones as inhibitors of sulfide production by sulfate- reducing bacteria in sewage, oil wells, process tanks, or biomass fermentation, in 22 pp. Cont.-in-part of U.S. Ser. No. 652,406, abandoned. 1993, du Pont de Nemours, E. I., and Co., USA: U.S.

109. Wheatley, A.D., Effluents, in Biodeterioration 7, D.R. Houghton, R.N. Smith, and H.O.W. Eggins, Editors. 1988, Cranfield Institute of Technology: Bedford. p. 193-206.

110. White, D., The Physiology and Biochemistry of Prokaryotes. 1995, New York: Oxford University Press. 378.

111. Wiesmann, U., Biological Nitrogen Removal from Wastewater. Advances in Biochemical Engineering Biotechnology, 1994. 51: p. 113-154. 90

112. Wikipedia, Wikpedia Free Encyclopedia. 2007, http://en.wikipedia.org/wiki/Zwitterion.

113. Wood, N.J., et al., Two domains of a dual-function NarK protein are required for nitrate uptake, the step of denitrification in Paracoccus pantotrophus. Molecular Microbiology, 2002. 44(1): p. 157-170.

114. Yang, X.-F., M.-L. Xie, and Y. Liu, Metabolic Uncouplers Reduce Excess Sludge Production in an Activated Sludge Process. Process Biochemistry, 2003. 38: p. 1373-1377.

115. Yasuda, R., et al., Rotation of the Subunit in F1-ATPase; Evidence That ATP Synthase Is a Rotary Motor Enzyme*. Bioenergetics and Biomembranes, 1997. 29(3): p. 207-209.

116. Ye, F.X., D.S. Shen, and Y. Li, Reduction in Excess Sludge Production by Addition of Chemical Uncouplers in Activated Sludge Batch Cultures. Journal of Applied Microbiology, 2003. 95: p. 781-786.

117. Ye, F.X. and Y. Li, Reduction of Excess Sludge Production by 3,3',4',5- Tetrachlorosalicylanilide in an Activated Sludge Process. Applied Microbiology and Biotechnology, 2005. 67: p. 269-274.

118. Ye, F.X. and Y. Li, Uncoupled Metabolism Stimulated by Chemical Uncoupler and Oxic-Settling-Anaerobic Combined Process to Reduce Excess Sludge Production. Applied Biochemistry and Biotechnology, 2005. 127(3): p. 187-199.

119. Yoshimi, N., et al., The mRNA overexpression of inflammatory enzymes, phospholipase A2 and cyclooxygenase, in the large bowel mucosa and neoplasms of F344 rats treated with naturally occurring carcinogen, 1- hydroxyanthraquinone. Cancer Lett., 1995. 97(1): p. 75-82.

91

Chapter 4 Effect of 1-hydroxyanthraquinone on biomass accumulation and oxygen utilization in Paracoccus denitrificans and in mixed wastewater cultures

Lynn Ann Arlauskas-Dekleva and Michael A. Gealt

Abstract

The activity of 1-hydroxyanthraquinone (OHAQ) on biomass production and oxygen utilization in Paracoccus denitrificans and mixed cultures derived from a municipal wastewater treatment system was evaluated. The addition of 10 mg/L of OHAQ resulted in a significant increase in the oxygen utilization rate and decrease in biomass production in batch respirometry studies. In continuous studies, OHAQ resulted in an increase in the respiratory activity that was sustained over the duration of the experiment. The activity of the compound was related to both the positional and electronic properties of the substitution on the anthraquinone ring system. The uncoupling activity combined with the low toxicity of OHAQ suggests that this compound may be used to reduce both the capital and operating costs of biological wastewater treatment systems by simultaneously improving the metabolic capacity and decreasing the biomass productions rates.

Introduction

The primary purpose of biological treatment processes is the removal of organic

material 51, thereby decreasing its environmental burden before its discharge. The

activated sludge process is the most common biological treatment process for domestic

and industrial wastewater 97, 154 and has proven to be a reliable and efficient treatment 92

technology capable of producing high-quality effluents 173. Wastewaters typically contain a complex mixture of components which are degraded by a mixed microbial population in biochemical reactions 99 that convert the organic material into new cell mass (biomass) and respiration products (carbon dioxide and water) 94, 98, 99. The excess biomass produced in the activated sludge process can be significant and requires additional processing and disposal 173. The costs associated with the handling and disposal of biomass from biological treatment processes can represent from 40 to 60% of a wastewater treatment plant operating costs 27, 99, 190 and may require significant capital expenditures.

The metabolism of organic wastes, in catabolism, provides the energy required for cell synthesis (anabolism) in the form of an energy rich molecule, ATP 24, 97, 112.

The relationship between ATP formation by catabolism and its use in biomass synthesis 24, is shown in Figure 1.

93

FIGURE 1: SIMPLIFIED RELATIONSHIP BETWEEN CATABOLISM 24 AND ANABOLISM (ADAPTED FROM REFERENCE )

Biomass

Substrate ATP

Catabolism Anabolism

Substrate

CO + H O 2 2 ADP + Pi

Energy

In most organisms that have been studied, the majority of the ATP is synthesized in a process known as respiration 127, 167 that utilizes a sequence of oxidation-reduction reactions to transfer electrons from the substrate to oxygen 60 or in the absence of oxygen to inorganic acceptors, e.g. nitrate 19. The electrons travel through a series of membrane-associated electron carrier proteins and lipids 191, known as the electron transport chain or respiratory network 181. The electron transport chain is localized in a membrane essentially impermeable to most ions, including OH(-) and

H(+) 61. In prokaryotes, the membrane is associated with the outer cellular membrane.

In eukaryotic systems the membrane is the inner membrane of the mitochondrion. The flow of electrons is coupled to the translocation of protons across the membrane that 94

generates an electrical or membrane potential ( ) and proton gradient ( pH) on the opposite sides of the membrane 40, 57, 61, 68, 114, 125.

The energy contained in the pH and known collectively as the electrochemical gradient or proton-motive force ( P) 57, 61, 68, 114-116, is recovered by the cell as the protons return across the membrane 103. The cell exploits this movement of

57, 61, 114-116 protons at the cell membrane to synthesize ATP from ADP and Pi in a reversible reaction 35, 45, 53, 187, 200. ATP provides the energy required for cell synthesis

112 and serves as a link between substrate oxidation and biomass synthesis 24.

One approach for decreasing biomass production involves the use of chemical compounds, known as uncouplers, which disrupt the direct energetic link between catabolism (oxidation of organic matter) and anabolism (ATP generation) 105, therefore limiting the cells ability to capture energy from substrate oxidation 98, 203. Uncouplers transport protons across biological membranes, dissipating the proton gradient across the membrane and decreasing the magnitude of P 49, 144. The rate of ATP synthesis and biomass production are reduced 66, 175, 203 while respiration, and therefore oxygen consumption, continues in the presence of substrate 66, 98, 144, 176.

Chemical uncouplers are a structurally diverse group of molecules 95, 144. In general, they are moderately weak lipophilic acids with an acid dissociation constant

176 (pKa) in the range or 5-7 . Uncouplers generally contain an acidic hydrogen, an

66, 176 electron-withdrawing moiety (-NO2, -CF3, -CN) and bulky hydrophobic group(s) 95

and at physiological pH the compounds will exist in both the neutral (HA) and the anionic (A-) forms 61.

Laboratory studies have shown that the application of the uncoupler technology to activated sludge units can result in a significant reduction in the biomass production rate 93, 99, 101, 199, 201-203. The most active uncoupler compounds are usually the most toxic and a serious problem with the uncoupler technology in wastewater treatment is the potential presence of the compound in the effluent where it might be toxic to the environment. A number of uncoupler compounds, (e.g. DNP, chlorinated phenols) are designated as toxic pollutants subject to effluent limitations 63, 64; impeding their utility at a full-scale wastewater treatment facility.

Anthraquinones, are found widely dispersed in the environment 204, occurring as the result of both natural and commercial processes 7, 18, 50, 204. They are found as raw materials 25 in the textile 58, plastics, cosmetic, 204 and medical 5, 67, 70, 128, 174 industries. They exhibit a wide variety of biological effects 7, 9, 10, 16, 18, 29, 34, 67, 70, 80-82,

131, 132, 148, 149, 189 including inhibition of electron transport in chloroplasts 80, 132, uncoupling of mitochondrial 10, 16, 82 and sulfate respiration29, and inhibition of bacterial growth 7, 9, 34, 81 and respiration 29, 131, 189.

The uncoupling activity of anthraquinones in sulfate-reducing bacteria 29 and mitochondria 10 combined with their low systemic toxicity 62 suggest that these compounds may have some utility for biomass reduction in wastewater treatment systems. Previous laboratory studies have focused primarily on the application of 96

anthraquinones for growth inhibition of aerobic and anaerobic microorganisms 29, 50, 131,

189 rather than its potential use as an uncoupler.

The purpose of this study is to elucidate the effect of 1-hydroxyanthraquinone

(OHAQ) on a complex biological system by analyzing not only batch and continuous aerobic cultures of Paracoccus denitrificans but also cultures derived biomass obtained at a municipal wastewater treatment facility. Paracoccus was chosen as the model system because of its ability to grow aerobically or under anaerobic conditions with nitrate, nitrite, or nitrous oxide as the terminal electron acceptor 21, 77, 171, 191. Formerly known as Micrococcus denitrificans 184, Pa. denitrificans is a gram negative 171, 181, non-fermenting facultative anaerobe, 15, 57, 77, 184 commonly found in soil and sludge 181.

The composition of the aerobic electron transport chain of Paracoccus denitrificans, 57, 75, 119, 181, 191 (Figure 2), closely resembles the four enzymatic complexes (I-IV) of a mitochondrial electron transport chain21, 57, 74, 75, 83 Yagi, 2003 #1157,

119, 127, 161, 181.

97

FIGURE 2: AEROBIC ELECTRON TRANSPORT PATHWAY IN 57, 75, 119, 181, PARACOCCUS DENITRIFICANS (MODIFIED FROM REFERENCES 191).

cyt o O2

dehydrogenase UQ bc1 c cyt aa3 O2

dehydrogenase

methanol

In Paracoccus, the electrons from NADH flow to oxygen via the NADH dehydrogenase enzyme complex, quinones, cytochrome bc1 complex, cytochrome c and a terminal cytochrome c oxidase 181. Most bacteria, unlike the mitochondrial respiratory chain, contain a number of alternative pathways in which electrons are transferred from specific dehydrogenases to a variety of terminal oxidases. The

Paracoccus aerobic pathway, branches at ubiquinone (UQ) leading either through the two coupling sites, the bc1 complex and cytochrome aa3 (cyt aa3) generating a proton- motive force, or bypassing the bc1 complex to cytochrome o (cyt o), a non-coupling site 191. The extent to which the o-type cytochrome is active under aerobic conditions is thought to vary with the different Paracoccus strains and growth conditions 160, 179,

184 and enables the bacterium to respond to a variety of environmental stimuli 8, 181. 98

The uncoupling activity of anthraquinones in sulfate-reducing bacteria 29 and mitochondria 10 combined with their low systemic toxicity 62 suggest that these compounds may have some utility for biomass reduction in wastewater treatment systems. The purpose of this study is to elucidate the effect of

1-hydroxyanthraquinone (OHAQ) on batch aerobic cultures of Paracoccus denitrificans. The activity was also evaluated in activated sludge cultures derived from a municipal wastewater treatment and continouous aerobic cultures of Paracoccus to assess the effectiveness of OHAQ in mixed culture and high cell density systems. 99

Materials and Methods

Organisms and Growth Conditions

Paracoccus

Paracoccus denitrificans (ATCC 19367) was obtained from the American Type

Culture Collection, Rockville, MD and maintained on Nutrient Agar or Tryptic Soy

Agar plates (Becton Dickinson & Company, Franklin Lakes NJ). Plates were stored at

room temperature and transferred on a weekly basis. The microbial culture was grown

in BTZ-3 130 minimal medium at 35-37 C, 200 RPM, in an environmental chamber

(Barnstead Thermolyne, Corporation Dubuque, IA) until late exponential or early

stationary phase.

Activated Sludge Cultures

Mixed liquor suspended solids (MLSS) were obtained from the secondary

aeration unit at Meadowview Municipal Wastewater treatment facility in Elkton, MD.

The culture was maintained in an aerated 1-liter fill and draw reactor. The laboratory

reactor was constructed from a 1-liter glass graduated cylinder. There were two

sampling ports, one located near the bottom of the reactor (approximately at the 250 ml

level) and at the midpoint of the unit. The unit was aerated using house air and a air 100

diffusion tube at a rate sufficient to maintain the oxygen concentration in the reactor at

5 mg/L. Mixing in the reactor was accomplished by the aeration system. Neither,

significant solids settling during the aeration period nor wall growth were not observed

during the study. On a daily basis, the supernatant was replaced with plant influent

after settling for 1 hour. The biomass concentration in the reactor was maintained at

approximately 4 g/L by wasting excess sludge as needed.

Chemicals and Stock Solutions

1-hydroxyanthraquinone (OHAQ, Aldrich 1,926-4, >97% pure),

1,8-dihydroxyanthraquinone (1,8-DiOH Aldrich D108103, 96% pure),

1,2-dihydroxyanthraquionone (1,2-DiOH Aldrich 122777), 2-aminoanthraquinone (2-

Amino Aldrich 16,554-9, technical grade), 2-ethylanthraquinone (2-Ethyl Aldrich

E-1,226-6, 97% pure), 2-methylanthraquinone (2-Methyl Aldrich 10,964-9, 95% pure)

, 2-hydroxymethylanthraquinone (2-OHMethyl Aldrich 22,652-1, 97% pure) were the

gift of J. Martin Odom (DuPont, Wilmington, DE). The following test compounds

were obtained from commercial sources: 1-Aminoanthraquinone (1-Amino Sigma

Aldrich 3909, 97% pure), purpurin (1,2,4-trihydroxyanthraquinone 1,2,4-TriOH

Aldrich 229148), 2-hydroxyanthraquinone (2-OHAQ ABCR Chemicals TCH1012,

95% pure) . 101

Stock solutions of test compounds were prepared at an approximate

concentration of 1000 mg/L in acetone (J.T. Baker JT 9010-1, purity 99.5%) due to

their low aqueous solubility. The test compounds, evaluated at a test concentration of

10 mg/L, were added to the test vessels using a syringe to minimize the introduction of

oxygen. The acetone concentration in the control and experimental test systems did

not exceed 1% (V/V).

Experimental Procedures

Respirometry Studies

Aerobic respirometer studies were performed in a closed test system using a

Columbus Instruments Micro-Oxymax respirometer (Columbus Instruments, Inc.

Columbus, OH). The test vessels consisted of a 1000 ml tissue culture bottle (Corning

Pyrex ® Corning, NY) modified with a crimp top addition tube (Bellco Glass,

Vineland NJ) and butyl rubber septum (Wheaton Millville, NJ) to permit sampling and

the addition of reagents while maintaining a closed system. The test bottles containing

the culture medium were aseptically connected to the respirometer and verified for

leaks prior to initiation of the experiment. The test compound or acetone (control) was

added using a 25G tuberculin syringe (Becton Dickinson & Company, Franklin Lakes,

NJ) and the test vessels were mixed for approximately 5 minutes at 200 RPM using an

orbital shaker (Barnstead Thermolyne, Corporation Dubuque, IA). The test vessels 102

were inoculated with an overnight culture of Paracoccus or biomass obtained from the laboratory activated sludge reactor to obtain the desired initial biomass concentration.

In most studies, the initial biomass concentration was approximately 60 mg/L for the

Paracoccus batch studies and 300 mg/L in the mixed culture studies.

The volume of the headspace in the test vessels was measured by the respirometer at the initiation of the experiment. The carbon dioxide and oxygen concentration of the headspace was determined initially and at two hour intervals using a single-beam infrared carbon dioxide sensor and an electrochemical oxygen sensor.

The carbon dioxide and oxygen sensors had operating ranges of 0-1.0% and 10-21%, respectively. The gas from the test vessels was passed through a magnesium perchlorate drying column prior to the sensors and returned to the test vessels. The headspace was refreshed after two test intervals. The sensors were purged with laboratory air after each sample interval and were calibrated using house nitrogen and a gas mixture consisting of 0.65 % CO2 and 20.6 % O2 (Linde Gas, Cleveland, OH).

Calibration of the gas sensors was performed at the initiation of the studies and checked on a monthly interval or if there was a perturbation of the system. The magnesium perchlorate in the gas drying columns was changed when the column was

50% utilized.

The respirometer studies were performed at ambient laboratory temperatures

(20-25 C) and the test vessels were agitated using a Lab Line orbital shaker

(BarnsteadThermolyne Corporation, Dubuque IA). The studies were run in triplicate and the results were averaged. Biomass determinations were performed at the 103

termination of the experiment. The duration of the experiments varied but was

generally 48 hours.

Paracoccus Continuous Studies

Continuous Paracoccus studies were performed in a 2-liter Braun Biotech

Biostat B fermentor (Sartorius, Bethlehem, PA) using a 1-liter working volume. The

BTZ-3 medium was fed to and effluent drawn from the reactor using a MasterFlex L/S

Digital Economy Drive (Cole Parmer Instrument Company, Vernon Hills, IL) and

PharMed tubing (Cole Parmer Instrument Company). The volume of the reactor was

maintained at a constant level by adjusting the height of the effluent tube prior to the

initiation of the studies. The temperature of the reactor was maintained at 25°C, the

temperature used in the batch respirometry studies. The aeration of the reactor was

controlled by a Brooks 5850 series mass flow controller (Emerson Process

Management, Pittsburgh, PA) and maintained at one standard liter per minute (SLPM)

over the duration of the experiment. Dissolved oxygen was measured by a

polarographic oxygen electrode (Mettler Toledo, Columbus, OH) and maintained at a

minimum of 20% by the speed of the impeller. The concentration of nitrogen, oxygen

and carbon dioxide in the off-gas was monitored and recorded at 3 minute intervals by

a VG Gas (Thermo Fisher Scientific, Waltham, MA) Prima 600S mass spectrometer. 104

Camile TG (developed by Guzdial, Georgia Institute of Technology, Atlanta, Georgia) and DuPont proprietary software was used for process control and data acquisition.

The continuous studies were performed consecutively using the same experimental apparatus. The reactor was operated was operated at a hydraulic residence time (HRT) of 11 hours by feeding the BTZ-3 medium containing 1% acetone for the control studies or BTZ-3 containing 10 mg/L OHAQ for the experimental studies at a rate of 1.5 mL/min The control reactor was maintained for a minimum of five residence times before changing to the experimental conditions

(control study) or OHAQ (OHAQ study), prior to initiating the medium feed to the reactor.

Paracoccus was established in the reactor containing 1-liter of sterilized BTZ-3 medium by growing under batch conditions for approximately 19 hours. The reactor was bulk dosed with acetone (10 ml) and fed BTZ-3 medium containing 1% acetone for a minimum of 55 hours (5 residence times) prior to sampling for biomass determinations. The reactor was converted from the control to test reactor by bulk dosing with a stock solution of OHAQ in acetone (final OHAQ concentration of 10 mg/L) and changing the feed to BTZ-3 containing 10 mg/L OHAQ. The reactor was sampled and experiment terminated after a minimum of five residence times.

The oxygen uptake rate (OUR), specific oxygen uptake rate (SOUR) and carbon dioxide evolution rate (CER) were calculated from the weight of the reactor and the oxygen and carbon dioxide concentrations of the inlet and outlet gas streams. The respiratory quotient (RQ) was calculated from the ratio of CER to OUR. 105

Biomass Determinations

The growth of the Paracoccus cultures was monitored at 590 nm using a Hach

DR/3000 (Loveland, CO) or Pharmacia LKB UV-Vis (Thermo Fisher Scientific

Waltham, MA) spectrophotometer. The relationship between Abs590 and dry weight

was determined and found to be linear up to an Abs590 of approximately 2.4.

Biomass concentrations (dry weight per liter of culture) of both the Paracoccus

and activated sludge cultures were determined by filtration on 0.2 or 0.45 M pre-

weighed filters. The filters were dried to constant weight in a 100 C oven after two

deionized water washes. 106

Results

Effect of OHAQ on Paracoccus

The effect of OHAQ was evaluated in aerobic batch respirometry studies using

Paracoccus as the model system. The oxygen uptake rates were used as a measure of

biological activity and were compared in the control and experimental bottles.

Biomass production was evaluated at the termination of the experiment. The addition

of OHAQ to aerobic batch cultures of Pa. denitrificans resulted in an increase in the

initial oxygen uptake rate indicating the stimulation of biological activity in the culture

(Table 2, Figure 3). At a test concentration of 10 mg/L, OHAQ increased the initial

-4 oxygen utilization rate 38.5% (p= 3.05 x 10 ) from 0.09 mg O2/hr for the control

bottles to 0.14 mg O2/hr for the OHAQ bottles. This initial increase in biological

activity did not affect the cumulative oxygen utilization but was associated with a

significant reduction in the biomass production. At 10 mg/L OHAQ, the biomass

production was reduced 39.3% (p = 3.13 x 10-2), from 3,983 mg/L for the control

bottles to 2,417 mg/L for 10 mg/L OHAQ. Similar results in respiratory activity and

biomass production were observed with OHAQ at test concentrations of 1 mg/L and 5

mg/L (Table 1, Figure 3) but not at test concentration of 0.1 mg/L OHAQ (data not

shown), this indicates there is a threshold concentration associated with OHAQ

activity.

107

TABLE 1: THE EFFECT OF 1, 5 AND 10 MG/L OHAQ ON BIOMASS PRODUCTION AND OXYGEN UTILIZATION RATE IN BATCH AEROBIC CULTURES OF PA. DENITRIFICANS.

Mean O2 Utilization Rate

O2 Utilization

Mean Biomass Biomass t-Test Mean O2 / unit Biomass/ O2 Dose Biomass SD Reduction p value 7-9 hrs 27-40 hrs Utilization biomass Utilization

mg O2 *L/ mg biomass/

mg/L mg/L % mg O2/hr mg O2/hr mg mg biomass L*mg O2 Control 3983 629 0.085 1.84 37.1 0.009 107 10 2417 548 39.3 3.13E-02 0.139 1.82 39.5 0.016 61.1 5 2433 369 38.9 2.12E-02 0.166 1.87 40.7 0.017 59.8 1 2075 530 47.9 3.96E-02 0.147 1.77 39.2 0.019 52.9

108

FIGURE 3: THE EFFECT OF OHAQ ON OXYGEN UTILIZATION IN AEROBIC BATCH CULTURES OF PA. DENITRIFICANS

10

8

6

4

2 Mean Oxygen Utilization,Mean Oxygen mg 0 2 4 7 9 11 14 16 18 20 Time, hours

Control 10 mg/L 5 mg/L 1 mg/L

Cumulative oxygen utilization from batch respirometry studies using Pa. denitrificans at an initial biomass concentration of 60 mg/L and acetone (") (for the control vessels) or 1-hydroxyanthraquinone at an initial test concentration of 10 mg/L (!), 5 mg/L (%) or 1 mg/L(#). Results are the mean of three replicate test vessels.

A comparison of oxygen utilization and biomass production at the termination of the experiment demonstrates that the oxygen utilization per unit biomass was significantly higher in the cultures containing OHAQ compared to the controls. This increase in biological activity combined with the significant decrease in biomass production establishes that OHAQ is an effective chemical uncoupler in Paracoccus batch cultures. 109

Paracoccus Substitution Studies

The relationship between the hydroxyl group and uncoupling activity of OHAQ

was investigated using ten anthraquinone compounds with amino-, methyl-, ethyl-,

hydroxymethyl- or multiple hydroxyl substitutions (Figure 4) in aerobic batch cultures

of Pa. dentrificans.

110

FIGURE 4: ANTHRAQUINONE TEST COMPOUNDS

R8 O R1

R7 R2

R6 R3

R5 O R4

Subsitution Test Compound Compound ID R1 R2 R3 R4 R5 R6 R7 R8

1-Hydroxyanthraquinone OHAQ -OH -H -H -H -H -H -H -H 1,8-Dihydroxyanthraquinone 1,8-DiOH -OH -H -H -H -H -H -H -OH

1-Aminoanthraquinone 1-Amino -NH2 -H -H -H -H -H -H -H 1,2,4-Trihydroxyanthraquinone 1,2,4-TriOH -OH -OH -H -OH -H -H -H -H

2-Hydroxymethylanthraquinone 2-OH Methyl -H -CH2OH -H -H -H -H -H -H

2-Ethylanthraquinone 2-Ethyl -H -CH2CH3 -H -H -H -H -H -H 1,2-Dihydroxyanthraquinone 1,2-DiOH -OH -OH -H -H -H -H -H -H

2-Aminoanthraquinone 2-Amino -H -NH2 -H -H -H -H -H -H 2-Hydroxyanthraquinone 2-OH -H -OH -H -H -H -H -H -H 2-Methylanthraquinone 2-Methyl -H -CH -H -H -H -H -H -H 3

These ten compounds were evaluated for their ability to uncouple microbial

respiration as measured by cumulative oxygen utilization and biomass reduction. The

compounds were chosen on the basis of substitution moity, substitution location on

ring system, commercial availability and potential application in wastewater treatment 111

systems. Halogen and nitro substituted compounds were not included in these evaluations because of the strict effluent discharge limits associated with these compounds. The evaluations were performed in a series of experiments that included not only controls but OHAQ to ensure that different series of experiments were comparable. The octanol/water partition coefficient (Log P), acid dissociation constant

(pKa) and water solubility of the chemicals were compiled from the literature and/or estimated using the EPA‘s EPIsuite software 39.

The results from these experiments, presented in Table 2, showed that biomass reductions did not correlate to either Log P or water solubility of the compounds, as correlation coefficients for the relationship were R2 = 0.1002 and R2 = 0.3263, respectively. The pKa of test compounds has been identified as an important parameter in their activity as uncouplers 199. However, the results obtained with the substituted anthraquinone compounds in this study did not show any direct relationship between the pKa and the observed biomass reductions. The test compounds used in this series had pKa‘s ranging from -0.051 for 1-Amino to 14.04 for 2-DiOH. The greatest biomass reductions were obtained with 1-OHAQ (pKa 7.15), 1,8-DiOH ( pKa 6.26) and

2-OH Methyl (pKa 14.04). Additionally, 1,2,4-TriOH with a reported pKa of 7.05, had no effect on the microbial respiration or biomass production.

112

TABLE 2: BIOMASS REDUCTIONS, LOG P, PKA AND WATER SOLUBILITIES OF TEST COMPOUNDS EVALUATED IN BATCH AEROBIC CULTURES OF PA. DENITRIFICANS .

Water Water Compound Biomass Solubilityd Solubilityd CAS d d Test Compound ID Reduction t Test Log P Log P pKa1 pKa2 (exp) (est) % (exp) (est) mg/L mg/L 1-Hydroxyanthraquinone OHAQ 40 3.52 3.64 7.0a 10.6a 8.5 9.5 129-43-1 1,8-Dihydroxyanthraquinone 1,8-DiOH 52.2 3.94 8.1a 10.7a 3.5 117-10-2 1-Aminoanthraquinone 1-Amino 22.2 3.74 3.53 -0.51b 0.3 1.6 72-48-0 1,2,4-Trihydroxyanthraquinone 1,2,4-TriOH 3.9 NS 3.46 7.05b 6.4 7.2 82-45-1 2-Hydroxymethylanthraquinone 2-OH Methyl 54 2.43 14.04b 58.4 117-79-3 2-Ethylanthraquinone 2-Ethyl 18.7 NS 4.37 4.38 0.4 17241-59-7 1,2-Dihydroxyanthraquinone 1,2-DiOH Variable 3.16 7.28a 10.5a 15.8 84-51-5 2-Aminoanthraquinone 2-Amino Inhibition 2.43 0.97b 0.2 21.7 84-54-8 2-Hydroxyanthraquinone 2-OH Inhibition 2.86 7.32a 11.1a 1.1 34.5 81-54-9 Slight Inhibition 605-32-3 2-Methylanthraquinone 2-Methyl 33.3 3.89 1.2

= significant, NS=Not significant , a Gorelikm M. et.al (1989) Zhurnal Organichesko Khimii 25(10):2046-2055 b Calculated using Advanced Chemistry Development (ACD/Labs) Software V8.14 for Solaris ( © 1994-2006 ACD/Labs) reported in Registry Database (Copyright © 2006 ACS) c Kirk-Othermer Encyclopedia of Chemical Technology 4th Edition Vol 2 John Wiley & Sons, New York 1992 d Data provided by EPA's EPIsuite estimation software http://www.epa.gov/oppt/exposure/pubs/episuitedl.htm

The results indicate that biomass reduction was related to both the positional

and electronic properties of the substitution(s). Electron donating groups (EDG), such

as hydroxyl or amino substitutions, in position 1 (ortho) resulted in a significant

reduction in biomass production. Conversely, substitution of either a hydroxyl or

amino group at position 2 (para) resulted in almost the complete inhibition of growth

and respiration.

The activity of compounds with multiple hydroxyl substitutions was found to

be dependent on the number and location of the substitutions. Hydroxyl substitution at

both positions 1 and 8 resulted in a 52.2% reduction (p=2.33 x 10-4) in biomass

production while substitution at both positions 1 and 2 lead to partial inhibition of the 113

culture and highly variable results. Additional substitutions by EDGs in position 4

(meta), as in the test compound 1,2,4-TriOH, resulted in the complete loss of uncoupling activity.

To investigate if the inhibition observed with position 2 substituted compounds was the result of the electron donating properties of the substitution, compounds with methyl, ethyl, or hydroxymethyl substitutions at position 2 were evaluated. Methyl and ethyl groups are not EDGs and the methylene spacer between the oxygen and aromatic ring in 2-hydroxymethyl interrupts electron flow between the oxygen and ring system, preventing the electron density from distributing across the ring system. As anticipated, the test compounds did not inhibit the respiration but rather resulted in a reduction in the biomass production. The biomass production was reduced by 33.3%

(p=1.40 x10-2) and 54.0% (p=2.37 x10-3) for the methyl and hydroxymethyl substituted compounds, respectively. The ethyl substituted compound reduced the biomass production by 18.7% (p=1.75 x10-1), which was not significant. The results obtained with these compounds demonstrate that the inhibition is the result of the electron donating properties of the substitution at position 2 of the ring system.

The results also indicate that biomass reduction was related to both the positional and electronic properties of the substitution(s) of the anthraquinone ring system. EDGs in position 1 resulted in uncoupling while EDGs in position 2 resulted in the inhibition of the microorganism. The hydroxyl substituted compounds had the greatest effect on biomass reductions when the compounds were substituted at position 114

1 or both positions 1 and 8. The addition of a hydroxyl group at position 2 to OHAQ

resulted in highly variable experimental results.

Paracoccus Continuous Culture Studies

The use of chemical uncouplers for biomass reduction in wastewater treatment

facilities requires that the compound is effective at a high population density under

continuous growth conditions. Therefore, the activity of OHAQ in continuous cultures

with Paracoccus, the , was evaluated. The reactor was operated at a

residence time of 0.09hr-1, which was slightly below the observed growth rate for the

organisms in the batch respirometry studies. The carbon dioxide evolution rate (CER),

oxygen utilization rate (OUR) and specific oxygen utilization rate (OUR) was

continuously monitored and final biomass concentration determined after a minimum

of five residence times.

The addition of OHAQ to a continuous Paracoccus culture resulted in an

increase in the CER, OUR and specific oxygen uptake rate (SOUR) but did not

decrease the biomass production rate (Table 3). The CER was increased 10.4 % from

27.6 mM/L·hr for the control reactor to 30.8 mM/L·hr for the OHAQ reactor.

Similarly, there was a 10.8 % increase in the SOUR from a control value of 3.98

mM/mg·hr to 4.46 mM/mg·hr for the OHAQ reactor. The stir speed of the reactor

increased from the steady state control rate of 610 RPM to 700 RPM in the presence of 115

OHAQ to compensate for the increased OUR. The growth of the organism was not adversely affected by the presence of OHAQ and the difference in the biomass concentrations of the control and OHAQ reactors was not significant. The final biomass concentration in the control and OHAQ studies was 9,050 ± 328 mg/L and

9,550 ± 471 mg/L, respectively.

The studies demonstrate that OHAQ increases the respiratory activity of the biomass, as evidence by the increases in CER, OUR and SOUR, under conditions growth conditions. The growth rate of the organism was not adversely affected by the presence of the test compound and was controlled by the reactor operating conditions.

The activity of OHAQ on the microbial biomass was immediate and was sustainable.

TABLE 3: SUMMARY OF THE CONTINUOUS PA. DENITRIFICANS REACTOR RESULTS FOR THE CONTROL AND OHAQ STUDIES. REACTOR WAS FED BTZ-3 MEDIUM CONTAINING ACETONE (FOR THE CONTROL) OR 10 MG/L OHAQ AT A RATE OF 1.5 ML/MIN.

Control OHAQ Final Biomass mg/L 9050 9550 CER mM/L*hr 27.6 30.8 CO2 out % 0.57 0.63 OUR mM/L*hr 35.9 42.6 SOUR mM/mg*hr 3.98 4.46 RQ 0.77 0.72 pH 7.2 7.1 Stir Speed RPM 610 700 CER= carbon dioxide evolution rate, OUR= Oxygen Uptake Rate, SOUR= Specific Oxygen Uptake Rate, RQ=Respiratory Quotient=CER/OUR

116

Mixed Culture Batch Studies

The application of the chemical uncouplers for biomass reduction in wastewater

treatment facilities requires that the compound is active on a mixed population without

adversely affecting the degradation capabilities of that population. Therefore, the

activity of OHAQ was evaluated in batch studies using mixed cultures derived from a

municipal wastewater treatment facility. The oxygen utilization rates were used as a

measure of biological activity and were compared in the control and experimental

bottles. Biomass production was evaluated at the termination of the experiment. The

addition OHAQ of mixed cultures resulted in an increase in the initial oxygen

utilization rate and a reduction in the overall biomass production (Table 4). The overall

oxygen utilization was not adversely affected by OHAQ. In test vessels containing 300

mg/L MLSS, 10 mg/L OHAQ increased the initial oxygen consumption rate 16.9%

(p=3.32 x 10-3) and decreased the biomass production 10.8 % (p = 7.76 x 10-3)

compared to the control bottles (Table 4). The mean oxygen utilization rate was

increased initially from 0.024 mg O2/hr for the control bottles to 0.028 mg O2/hr for the

OHAQ bottles between 7 and 9 hours followed by a decrease to 0.021 mg O2/hr over

the 29 to 34 hour test interval. The oxygen utilization rate in the control bottles was

constant throughout the experiment. The cumulative oxygen utilization in the control

and OHAQ bottles was comparable at the termination of the experiment while the 117

biomass production was decreased from 1083 mg/L for the control bottles to 967 mg/L

for the OHAQ bottles.

TABLE 4: THE EFFECT OF 1-HYDROXYANTHRAQUINONE (OHAQ) AT A TEST CONCENTRATION OF 10 MG/L ON OXYGEN CONSUMPTION RATE AND BIOMASS PRODUCTION IN AEROBIC MIXED CULTURES DERIVED FROM MUNICIPAL WASTEWATER TREATMENT FACILITY.

Mean O2 Utilization Rate

Mean Biomass t-Test Mean O2 O2 Utilization / Biomass / O2 Biomass SD Reduction p value 7 to 9 hrs 29 to 34 hrs Utilization biomass Utilization mg O2 *L/ mg mg biomass/

Sample ID mg/L % mg O2/hr mg O2/hr mg biomass L*mg O2 Control 1083 28.9 0.024 0.023 1.19 1.10E-03 910 OHAQ 967 28.9 10.8 7.76E-03 0.028 0.021 1.21 1.25E-03 799

A comparison of oxygen utilization and biomass production at the termination

of the experiment demonstrates that the oxygen utilization per unit biomass was

significantly higher in the cultures containing OHAQ compared to the controls (p=2.27

x 10-3). This increase in biological activity combined with the significant decrease in

biomass production establishes that OHAQ is an effective chemical uncoupler in mixed

cultures derived from municipal wastewater treatment facilities.

118

Discussion

The addition of 10 mg/L OHAQ to Pa. denitrificans and activated sludge

cultures resulted in a significant increase in the oxygen utilization rate compared to the

controls. In batch studies, this increase was observed in the initial oxygen utilization

rate and was associated with a significant decrease in the final biomass concentration.

The biomass reductions in the Paracoccus batch studies were consistently in excess of

40% throughout the entire experimental program. OHAQ was equally effective over a

test concentration range of 1 to 10 mg/L but had no significant effect at a test

concentration of 0.1 mg/L.

In the continuous Paracoccus studies, OHAQ addition resulted in an increase in

the respiratory activity compared to the steady state control values and was sustained

over the duration of the experiment. The addition of OHAQ did not significantly

reduce the biomass production in the continuous study but increased the SOUR

approximately 10%. These results suggest that the compound could increase the

capacity of culture treatment facilities without modification in or addition to existing

equipment.

The biomass reduction in the mixed culture batch studies was less than that

obtained in the Paracoccus batch studies. The average biomass reduction with 10

mg/L OHAQ was 10.8% in mixed cultures with an initial biomass concentration of 300

mg/L. This value although lower than that obtained in the pure culture studies, is

significant in regards to the relative biomass production produced in wastewater 119

treatment facilities. An industrial facility with a hydraulic flow rate of 1100 GPM with an organic loading of 20,000mg/L biological oxygen demand (BOD5) produces approximately 79,200 lbs of biomass per day 150 and a moderate flow municipal wastewater treatment facility with a hydraulic flow rate of 6319 GPM and organic

111 loading of 187mg/L BOD5 generates 13,300 lbs of biomass per day .

The relationship between the uncoupling activity of the compounds and the uncoupler to biomass ratio (Cu/Xo) has been reported for both pure and mixed cultures

93, 133, 172, 201. The results obtained in the batch and continuous studies verified that the activity of OHAQ on biomass production rates was related to the initial OHAQ concentration to initial biomass ratio, Cu/Xo, rather than the initial test concentration.

There was a strong correlation (R2= 0.9979) between the OHAQ to biomass ratio and the biomass reduction obtained with the Paracoccus and mixed culture studies (Figure

5). The average biomass reductions obtained over the experimental program were in excess of 40% when evaluated at a Cu/Xo of 0.17. A decrease in the Cu/Xo to 0.001, did not result in a reduction in the biomass production in either the batch or continuous

Paracoccus studies.

120

FIGURE 5: THE EFFECT OF OHAQ TO BIOMASS RATIO ON BIOMASS REDUCTIONS IN BATCH AND CONTINUOUS PARACOCCUS AND BATCH MIXED CULTURE STUDIES.

50 y = 408.57x - 1.1742 2 40 R = 0.9979

30

20 Biomass Reduction, % Reduction, Biomass 10

0 0.00 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.10 OHAQ/Biomass Ratio

In the mixed culture batch studies, a Cu/Xo OHAQ of 0.03, resulted in a lower but significant reduction in the overall biomass production. The lower activity of

OHAQ in the mixed culture studies was attributed to the diverse microbial population present in wastewater treatment systems. Chemical uncouplers may affect the growth rates of the individual species in the population differently and result in a change in the population dynamics of the system 100.

The effect of ring substitutions on the uncoupling efficiency of phenolic compounds has been reported in a variety of systems. Clowes & Krahl 26, 86, in 1936, noted the location of the substitution in nitro- and halogenated phenols directly affected 121

uncoupling activity in sea urchin (Arbacia) eggs and that the presence of a phenolic hydroxyl group was essential for biological activity. Betina and Kuzela 10 found that anthraquinone compounds bearing a free hydroxyl group at the -position exhibited uncoupling behavior in rat liver mitochondria and compounds, lacking a -hydroxyl group, were ineffective. This requirement for a -hydroxyl group in uncoupling was confirmed by Kawai et al 82 who found that 1,2-DiOH, which contains a -hydroxyl group, was the only compound that exhibited significant uncoupling behavior.

In this study, there was a strong correlation between biomass reductions and the electronic substitution and its position. Electron donating groups (EDG) in position 1

(ortho) of the aromatic ring system resulted in a larger biomass reduction than substitutions in either position 2 (meta) or multiple (ortho, meta and para) substitutions.

Electron donating groups in position 1 were not inhibitory to Paracoccus and did not adversely affect the respiration of the cultures. In the group of compounds tested, a hydroxyl group in position 1 had a larger effect on biomass reduction than an amino group substitution, although both groups resulted in significant biomass reductions.

Position 2 EDGs did not result in significant biomass reductions and were, in general, inhibitory to the culture.

The substitution results suggest that the difference in the activity of the compounds with substitutions at positions 1 or 2 may be the result of different electron density of the compounds. EDGs have at least one pair of non-bonded electrons on the hetroatom next to the aromatic ring structure 102 which donate electrons onto the ring 122

structure 28. Substitutions in position 1 donate electrons to the ring structure according

to the proposed mechanism 1, Figure 6, and result in an electrically neutral compound

with partial negative and partial positive charges in close proximity 65. While

substitutions in position 2 donate electrons according to the proposed mechanism 2,

Figure 6, resulting in the generation of a zwitterion 65, a compound which is electrically

neutral but carries formal positive and negative charges 193.

FIGURE 6: DISTRIBUTION OF ELECTRONS IN SUBSTITUTED ANTHRAQUINONE COMPOUNDS WITH ELECTRON DONATING GROUPS

Mechanism 1

- + O OH O OH

O O

Mechanism 2

O O

OH OH+

- O O

123

The difference in the charge distribution of the ortho and meta substituted

anthraquinone compounds may be the basis for the difference in biomass reductions in

these studies. The proposed mechanism of proton translocation mediated by an

uncoupler assumes that the cell membrane is impermeable to most ions and the

chemical compounds are soluble in the membrane in both the neutral (HA) and anionic

(A-) forms 107, 175, 176. The anions (A-) move from right to left through the interior of

the membrane driven by the higher negative interior potential (Figure 7).

FIGURE 7: THE MOLECULAR MECHANISM OF H+ TRANSLOCATION MEDIATED BY A PROTONOPHORIC UNCOUPLER 107 175 176 (ADAPTED FROM )

A- A-

H+ H+

HA HA + — Membrane

This increase in the concentration of anions at the left-hand interface drives the heterogenous reaction: 124

(-) (+) A if + H HAif

- where an anion at the interface (A if) combines with a proton from the aqueous

+ phase (H ) to form the undissociated acid (HAif ). The neutral form of the uncoupler diffuses through the membrane to the interior surface, driven by the concentration gradient of the compound across the membrane. The compound encounters the more alkaline environment at the interior interface, deprotonates, and releases a proton into the cytoplasm. The anion then returns electrophoretically to the positively charged outside from the negatively charged inside of the energized membrane. Chemical uncouplers do not directly interact with the components of the electron transport chain

175 but rather decrease the rate of ATP synthesis and biomass production 66, 98, 144, 176, 203 by disrupting the link between substrate oxidation and biomass synthesis 24. The net result of the uncoupler cycle is the transport of H(+) into the cell and dissipation of both the proton gradient and P 107, 175, 176. A larger fraction of the substrate consumed in respiration is in support of proton translocation to maintain the proton gradient rather than the production of ATP resulting in a decrease in biomass production 175.

A difference in the electron distribution may interfere with diffusion of the compound across or within the membrane and/or the release of protons. The activity of the substituted anthraquinone compounds utilized in this study indicates that the charge distribution of the compounds plays a key role in their ability to reduce biomass 125

production in microbial systems. The mechanism of uncoupling outlined above suggests that the polar nature of some of the substituted compounds may interfere with cellular metabolism as a result of their inability to diffuse across the cell membrane.

The studies in this report demonstrate that the activity of OHAQ are the result of the ―uncoupling‖ of aerobic respiration. This laboratory has demonstrated the compound also possesses uncoupling activity under denitrifying conditions (Chapter

3).

Biological treatment by the activated sludge process has become the major technology for treating both municipal and industrial wastewaters worldwide 203.

Previously, the volume of wastewater treated, maximum substrate removal, and effluent quality were the main criteria for the activated sludge process. However, recently the rising cost for the treatment and disposal of excess sludge produced and restrictive legislation have emphasized the need to minimize biomass generated from activated sludge operations 202. The concept of uncouplers for biomass reduction in activated sludge systems is not new, and while there is a large amount of literature on uncouplers, reports on their application in activated sludge systems is limited. The resulted reported here indicate that OHAQ has the potential to reduce both the capital and operating costs of biological wastewater treatment systems by simultaneously improving the metabolic capacity and decreasing the biomass production rates.

Anthraquinones have an advantage over other uncouplers of oxidative phosphorylation because of their low systemic toxicity 62, 153. Miscellaneous inhibitory effects on particular enzyme systems have been reported, but the overall lack of 126

toxicity of anthraquinones is supported by their natural occurrence in plants, their widespread use as dyes and their use until recently as laxatives 131. The uncoupling activity combined with the low toxicity of compounds suggests that these compounds may be used for reducing biomass production in commercial bioprocesses in addition to traditional wastewater treatment systems. The need for efficient, cost-effective and environmentally sound waste treatment in both the municipal and industrial markets warrants the further investigation of the use of this chemical uncoupler for enhancement of metabolic capacity and biomass reductions. 127

References

1. Almasifar, D., et al., Spectrophotometric Determination of Acidity Constants of Some Recently Synthesized Anthraquinones in Methanol + Water. Journal of Chemical Engineering Data, 1997. 42: p. 1212-1215.

2. Anke, H., et al., Metabolic Products of Microorganisms. 185*. The Anthraquinones of the Aspergillus glaucus Group. I. Occurrence, Isolation, Identification and Antimicrobial Activity. Archives of Microbiology, 1980. 126: p. 223-230.

3. Anraku, Y., Bacterial Electron Transport Chains. Annual Review of Biochemistry, 1988. 57: p. 101-32.

4. Bakola-Christianopoulou, M.N., L.B. Ecateriniadou, and K.J. Sarrris, Evaluation of the Antimicrobial Activity of a New Series of Hydroxy-Quinone Chelates of Some Transition Metals. European Journal of Medicinal Chemistry - Chim. Ther, 1986. 21(5): p. 385-390.

5. Betina, V. and S. Kuzela, Uncoupling effect of fungal hydroxyanthraquinones on mitochondrial oxidative phosphorylation. Chem.-Biol. Interact., 1987. 62(2): p. 179-89.

6. Boogerd, F.C., et al., Reconsideration of the efficiency of energy transduction in Paracoccus denitrificans during growth under a variety of culture conditions. A new approach for the calculation of P/2e- ratios. Arch. Microbiol., 1984. 139(4): p. 344-50.

7. Boos, K.S. and E. Schlimme, Anthraquione Dyes: A New Class of Potent Inhibitors of Mitochrondrial Adenine Nucleotide Translocation and Oxidative Phosphorylation. FEBS Letters, 1981. 127(1): p. 40-44.

8. Brown, J.P., A Review of the Genetic Effects of Naturally Occurring Flavonoids, Anthraquinones and Related Compounds. Mutation Research, 1980. 75: p. 243- 277. 128

9. Bryan, B.A., Physiology and Biochemistry of Denitrification, in Denitrification, Nitrification, and Atmospheric Nitrous Oxide, C.C. Delwiche, Editor. 1981, John Wiley & Sons: New York. p. 67-84.

10. Calder, K., K.A. Burke, and J. Lascelles, Induction of Nitrate Reductase and Membrane Cytochromes in Wild Type and Chlorate-Resistant Paracoccus denitrificans. Archives of Microbiology, 1980. 126: p. 149-153.

11. Chudoba, P., J. Chudoba, and B. Capdeville, The Aspect of Energetic Uncoupling of Microbial Growth in the Activated Sludge Process-OSA System. Water Science & Technology, 1992. 26(9-11): p. 2477-2480.

12. Chung, R.H., Anthraquinone, in Kik-Othmer Enclycopedia of Chemical Technology. 1978, John Wiley & Sons: New York. p. 700-757.

13. Clowes, G.H.A. and M.E. Krahl, Studies on Cell Metabolism and Cell Division I. On the Relation between Molecular Structures, Chemical Properties, and Biological Activities of the Nitrophenols. Journal of General Physiology, 1936. 20: p. 145-171.

14. Constable, S.W., Solids disposal costs. 1995, Personal Communication.

15. Cook, S.R., Steve's place-Reaction Mechanisms. 2002, http://www.steve.gb.com/science/reaction_mechanisms.html.

16. Cooling III, F.B., et al., Inhibition of Sulfate Respiration by 1,8- Dihydroxyanthraquinone and Other Anthraquinone Derivatives. Applied and Environmental Microbiology, 1996. 62(8): p. 2999-30004.

17. Cudlin, J., et al., Biological activity of hydroxyanthraquinones and their glucosides toward microorganisms. Folia Microbiol., 1976. 21(1): p. 54-7.

18. Das, A. and L.G. Ljungdahl, Composition and Primary Structure of the F1F0 ATP Synthase from the Obligately Anaerobic Bacterium Clostridium thermoaceticum. Bacteriology, 1997. 179(11): p. 3746-3755. 129

19. EPA, EPI Suite. 2000, http://www.epa.gov/opptintr/exposure/pubs/episuite.htm.

20. Ferguson, S. The Redox Reactions of the Nitrogen and Sulphur Cycles. in Nitrogen and Sulphur Cycles Symposium 42. 1988: Society for General Microbiology.

21. Fillingame, R.H., Membrane Sectors of F- and V-type H+-transporting ATPases. Current Opinion in Structural Biology, 1996. 6: p. 491-498.

22. Gage, D.J. and F. Neidhardt, Adaptation of Escherichia coli to the Uncoupler of Oxidative Phosphorylation 2,4 - Dinitrophenol. Journal of Bacteriology, 1993. 175(21): p. 7105-7108.

23. Garcia-Lopez, P.M., L. Kung Jr, and J.M. Odom, In Vitro Inhibition of Microbial Methane Production by 9,10-Anthraquinone. Journal of Animal Science, 1996. 74: p. 2276-2284.

24. Gaudy Jr, A.F. and E.T. Gaudy, Mixed Microbial Populations. Advances in Biochemical Engineering, 1972: p. 97-143.

25. Gogol, E.P., et al., Ligand-dependent structural variations in Escherichia coli F1 ATPase revealed by cryoelectron microscopy. Proceeding of the National Academy of Sciences, 1990. 87: p. 9585-9589.

26. Haddock, B.A. and C.W. Jones, Bacterial Respiration. Bacteriological Reviews, 1977. 41(1): p. 47-99.

27. Hall, N., Anthraquinone dyes containing a fluorosulfonyl group and their use. 1995, Zeneca Ltd., UK: PCT Int. Appl. p. 17.

28. Harem, G.N., Investigation of Chemical Uncouplers to Reduce Activated Sludge Yield, in Civil Engineering. 1995, University of Wahington.

29. Harold, F.M., Conservation and Transformation of Energy by Bacterial Membranes. Bacteriological Reviews, 1972. 36(2): p. 172-230. 130

30. Health, N.I.o., Hazardous Substances Databank Anthraquinone. 2000, National Library of Medicine's TOXNET System.

31. Health, N.I.o., Hazardous Substances Databank Trichlorophenol. 2000, National Library of Medicine's TOXNET System.

32. Health, N.I.o., Hazardous Substances Databank Dinitrophenol. 2000, National Library of Medicine's TOXNET System.

33. Hennessy, M.J., The activity of electron donating substitutions on aromatic structures. 2007, Personal Communication.

34. Heytler, P.G., Uncouplers of Oxidative Phosphorylation. Methods in Enzymology, 1979. LV: p. 462-473.

35. Hilgert, I., et al., Antitumor and immunosuppressive activity of hydroxyanthraquinones and their glucosides. Folia Biol., 1977. 23(2): p. 99-109.

36. Hinkle, P.C. and R.E. McCarty, How Cells Make ATP. Scientific American, 1978. 238: p. 104-123.

37. House, R.V. and J.H. Dean, Structural requirements for suppression of cytotoxic T lymphocyte induction by polycyclic compounds. In Vitro Toxicol., 1987. 1(3): p. 149-62.

38. Jassal, B., Electron Transport Chain. 2005. http://www.reactome.org/cgi- bin/eventbrowser?DB=gk_current&ID=163200&ZOOM=2

39. John, P. and F.R. Whatley, Paracoccus denitrficans and the Evolutionary Origin of the Mitochondrion. Nature, 1975. 254: p. 495-498.

40. John, P. and F.R. Whatley, The Bioenergetics of Paracoccus denitrificans. Biochemica et Biophysica Acta, 1977. 463: p. 129-153. 131

41. Karukstis, K.K., et al., Alternative Measures of Photosystem II Electron Transfer Inhibition in Anthraquinone-Treated Chloroplasts. Photochemistry and Photobiology, 1992. 55(1): p. 125-132.

42. Kavanagh, F., Activities of Twenty-Two Antibacterial Substances Against Nine Species of Bacteria. Journal of Bacteriology, 1947. 54: p. 761-766.

43. Kawai, K., et al., The inhibition of mitochondrial respiration by anthraquinone mycotoxins. The relationship between the position of hydroxyl group on the anthraquinone nucleus and the uncoupling effect on oxidative phosphorylation. Proc. Jpn. Assoc. Mycotoxicol., 1987. 26: p. 43-6.

44. King, M.W., Oxidative Phosphorylation. 2006, Indiana State University School of Medicine.

45. Krahl, M.D. and G.H.A. Clowes, Studies on Cell Metabolism and Cell Division II. Stimulation of Cellular Oxidation and Reversible Inhibition of Cell Division by Dihalo and Trihalophenols. Journal of General Physiology, 1936. 20: p. 173-184.

46. Liu, Y., Effect of Chemical Uncoupler on the Observed Growth Yield in Batch Culture of Activated Sludge. Water Research, 2000. 34(7): p. 2025-2030.

47. Liu, Y. and J.-H. Tay, Strategy for Minimization of Excess Sludge Production from the Activated Sludge Process. Biotechnology Advances, 2001. 19: p. 97- 107.

48. Liu, Y., Chemically Reduced Excess Sludge Production in the Activated Sludge Process. Chemosphere, 2003. 50: p. 1-7.

49. Low, E.W. and H.A. Chase. The Use of Uncouplers to Inhibit Biomass Production in Wastewater Treatment. in Icheme Research Event. 1996. Europe: Institute of Chemical Engineers.

50. Low, E.W. and H.A. Chase, The Use of Chemical Uncouplers for Reducing Biomass Production During Biodegradation. Water Science and Technology, 1998. 37(4/5): p. 399-402. 132

51. Low, E.W. and H.A. Chase, Reducing Production of Excess Biomass During Wastewater Treatment. Water Research, 1999. 33(5): p. 1119-1132.

52. Low, E.W., et al., Uncoupling of Metabolism to Reduce Biomass Production in the Activated Sludge Process. Water Research, 2000. 34(12): p. 3204-3212.

53. Low, E.W., et al., The Use of Chemical Uncouplers for Reducing Biomass Production in the Activated Sludge Process. Water Research, 2000. 12: p. 3204- 3212.

54. Macmillan, J.G., Electrophilic Aromatic Substitution. 1999, http://faculty.cns.uni.edu/~macmilla/mcmurry/mcmurry_chapter_16/sld.

55. Maloney, P.C., Energy Coupling to ATP Synthesis by the Proton-Translocating ATPase. Journal of Membrane Biology, 1982. 67: p. 1-12.

56. Mayhew, M. and T. Stephenson, Biomass Yield Reduction: Is Biochemical Manipulation Possible Without Affecting Activated Sludge Process Efficiency? Water Science & Technology, 1998. 38(8-9): p. 137-144.

57. McLaughlin, S.G. and J.P. Dilger, Transport of Protons Across Membranes by Weak Acids. Physiological Reviews, 1980. 60(3): p. 825-863.

58. Miller, S., Municipal plant operations. 2006, Personal Communication.

59. Millis, N.F., Effect of Uncoupling Oxidative Phosphorylation on the Economics of the Activated Sludge Process. Chem. Eng. Common., 1986. 45: p. 135-144.

60. Mitchell, P., Chemiosmotic Coupling in Oxidative and Photosynthetic Phosphorylation. Biological Reviews, 1966. 41: p. 445-502.

61. Mitchell, P. and J. Moyle, Proton Translocation Coupled to ATP Hydrolysis in Rat Liver Mitochondria. European Journal of Biochemistry, 1968. 4(4): p. 530- 539. 133

62. Mitchell, P. and J. Moyle, Estimation of Membrane Potential and pH Difference across the Cristae Membrane of Rat Liver Mitochondria. European Journal of Biochemistry, 1969. 7(4): p. 471-484.

63. Moat, A.G. and J.W. Foster, Microbial Physiology. Second Edition ed. 1988, New York: John Wiley & Sons. 597.

64. Neidhardt, F.C., J.L. Ingraham, and M. Schaechter, Physiology of the Bacterial Cell. 1990, Sunderland, Massachusetts: Sinauer Associates, Inc. Publishers.

65. Nicholls, D.G. and S.J. Ferguson, Bioenergetics 2. 1992, New York: Academic Press. 255.

66. Nishio, A. and E.M. Uyeki, Cellular uptake and inhibition of DNA synthesis by dihydroxyanthraquinone and two analogs. Cancer Res., 1983. 43(5): p. 1951-6.

67. Odom, J.M., et al., Immunological Cross-Reactivities of Adenosine-5'- Phosphosulfate Reductases from Sulfate-Reducing and Sulfide-Oxidizing Bacteria. Applied and Environmental Microbiology, 1991. 57(3): p. 727-733.

68. Odom, J.M., Anthraquinone inhibition of methane production in methanogenic bacteria, in 42 pp. 1994, Bio-Technical Resources LP, USA: PCT Int. Appl.

69. Oettmeier, W., K. Masson, and A. Donner, Anthraquinone inhibitors of photosystem II electron transport. FEBS Letters, 1988. 231(1): p. 259-62.

70. Okey, R.W. and H.D. Stensel, Uncouplers and Activated Sludge - The Impact on Synthesis and Respiration. Toxicological and Environmental Chemistry, 1993. 40: p. 235-254.

71. Phillips-Jones, M.K. and M.E. Rhodes-Roberts, Studies of Inhibitors of Respiratory Electron Transport and Oxidative Phosphorylation. Bacteriology, 1991. 27: p. 203-224.

72. Powell, A.K., Growth-Inhibiting Action of Some Pure Substances. Nature, 1944. 153: p. 345. 134

73. Prestera, T., H.J. Prochaska, and Talalay, Inhibition of NAD(P)H:(Quinone- Acceptor) Oxidoreductase by Cibacron Blue and Related Anthraquinone Dyes: A Structure-Activity Study. Biochemistry, 1992. 31: p. 284-293.

74. Reich, R.A., "Typical" Industrial Wastewater Treatment Facility. 2006, Personal Communication.

75. Richardson, M.L. and S. Gangolli, eds. The Dictionary of Substances & Their Effects. 1994, The Royal Society of Chemistry.

76. Rocher, M., et al., Towards a Reduction in Excess Sludge Production in Activated Sludge Processes: Biomass Physicochemical Treatment and Biodegradation. Applied Microbiology and Biotechnology, 1999. 51: p. 883-890.

77. Scholes, P.B. and L. Smith, Composition and Properties of the Membrane-Bound Respiratory Chain System of Micrococcus dentrificans. Biochimica et Biophysica ACTA, 1968. 153: p. 363-375.

78. Schultz, B.E. and S.I. Chan, Structures and Proton-Pumping Strategies of Mitochondrial Respiratory Enzymes. Annual Review of Biophysics and Biomolecular Structure, 2001. 30: p. 23-65.

79. Soboll, S., Regulation of Energy Metabolism in Liver. Bioenergetics and Biomembranes, 1995. 27(6): p. 571-582.

80. Stouthamer, A.H., Metabolic Regulation Including Anaerobic Metabolism in Paracoccus denitrificans. Journal of Bioenergetics and Biomembranes, 1991. 23(2): p. 163-185.

81. Strand, S.E. and H.D. Stensel, Sludge Yield Reduction Using Chemical Uncouplers. 1994, Electric Power Research: Seattle.

82. Strand, S.E., G.N. Harem, and H.D. Stensel, Activated-Sludge Yield Reduction Using Chemical Uncouplers. Water Environment Research, 1999. 71(4): p. 454- 458. 135

83. Swanbeck, G., Interaction Between Deoxyribonucleic Acid and Some Anthracene and Anthraquinone Derivatives. Biochim. Biophys. Acta, 1966. 123: p. 630-633.

84. Terada, H., The Interaction of Highly Active Uncouplers with Mitochondria. Biochimica et Biophysica Acta, 1981. 639: p. 225-242.

85. Terada, H., Uncouplers of Oxidative Phosphorylation. Environmental Health Perspectives, 1990. 87: p. 213-218.

86. Trumpower, B.L., The Three-Subunit Cytochrome bc1 Complex of Paracoccus denitrificans. Its Physiological Function, Structure, and Mechanism of Electron Transfer and Energy Transduction. Journal of Bioenergetics and Biomembranes, 1991. 23(2): p. 241-255.

87. Van Spanning, R.J.M., et al., Regulation of oxidative phosphorylation: the flexible respiratory network of Paracoccus denitrificans. J. Bioenerg. Biomembr., 1995. 27(5): p. 499-512.

88. Vignais, P.M., et al., The electron transport system and hydrogenase of Paracoccus denitrificans. Curr. Top. Bioenerg, 1981. 12: p. 115-96.

89. Weber, J. and A.E. Senior, Catalytic mechanism of F1-ATPase. Biochimica et Biophysica Acta, 1997. 1319: p. 19-58.

90. Weimer, P.J., et al., Anthraquinones as inhibitors of sulfide production by sulfate- reducing bacteria in sewage, oil wells, process tanks, or biomass fermentation, in 22 pp. Cont.-in-part of U.S. Ser. No. 652,406, abandoned. 1993, du Pont de Nemours, E. I., and Co., USA: U.S.

91. Wheatley, A.D., Effluents, in Biodeterioration 7, D.R. Houghton, R.N. Smith, and H.O.W. Eggins, Editors. 1988, Cranfield Institute of Technology: Bedford. p. 193-206.

92. White, D., The Physiology and Biochemistry of Prokaryotes. 1995, New York: Oxford University Press. 378. 136

93. Wikipedia, Wikpedia Free Encyclopedia. 2007, http://en.wikipedia.org/wiki/Zwitterion.

94. Yang, X.-F., M.-L. Xie, and Y. Liu, Metabolic Uncouplers Reduce Excess Sludge Production in an Activated Sludge Process. Process Biochemistry, 2003. 38: p. 1373-1377.

95. Yasuda, R., et al., Rotation of the Subunit in F1-ATPase; Evidence That ATP Synthase Is a Rotary Motor Enzyme*. Bioenergetics and Biomembranes, 1997. 29(3): p. 207-209.

96. Ye, F.X., D.S. Shen, and Y. Li, Reduction in Excess Sludge Production by Addition of Chemical Uncouplers in Activated Sludge Batch Cultures. Journal of Applied Microbiology, 2003. 95: p. 781-786.

97. Ye, F.X. and Y. Li, Uncoupled Metabolism Stimulated by Chemical Uncoupler and Oxic-Settling-Anaerobic Combined Process to Reduce Excess Sludge Production. Applied Biochemistry and Biotechnology, 2005. 127(3): p. 187-199.

98. Ye, F.X. and Y. Li, Reduction of Excess Sludge Production by 3,3',4',5- Tetrachlorosalicylanilide in an Activated Sludge Process. Applied Microbiology and Biotechnology, 2005. 67: p. 269-274.

99. Yoshimi, N., et al., The mRNA overexpression of inflammatory enzymes, phospholipase A2 and cyclooxygenase, in the large bowel mucosa and neoplasms of F344 rats treated with naturally occurring carcinogen, 1- hydroxyanthraquinone. Cancer Lett., 1995. 97(1): p. 75-82.

137

Chapter 5 The economic implications of metabolic uncoupling in wastewater treatment facilities

Lynn Ann Arlauskas-Dekleva and Michael A. Gealt

The primary purpose of biological treatment processes is the removal of organic

material 51, thereby decreasing its environmental burden before its discharge. The

activated sludge process is the most common biological treatment process for domestic

and industrial wastewater 97, 154 and has proven to be both a reliable and efficient

treatment technology capable of producing high-quality effluents 173. Wastewaters

typically contain a complex mixture of components which are degraded by a mixed

microbial population in biochemical reactions 99 that convert the organic material into

new cell mass (biomass) and respiration products (carbon dioxide and water) 94, 98, 99.

The excess biomass produced in the activated sludge process can be significant and

requires additional processing and disposal 173. The costs associated with the handling

and disposal of biomass from biological treatment processes typically represent from

40 to 60% of a wastewater treatment plant operating costs 27, 99, 190 and may require

significant capital expenditures.

In the United States, approximately 55% of the 7.2 millon dry metric tons of

biosolids produced in 2004 were applied to soils for agronomic, silvicultural, and/or

land restoration purposes 123. The remaining 45% were disposed of in municipal solid 138

waste (MSW) landfills (63%), surface disposal units (4%), and/or incineration facilities

(33%) 123. Although composted biosolids were applied to the White House lawn, the

Washington Monument and the National Mall Constitution Gardens 178, the overall picture for beneficial use of biosolids is hampered by the political, economic and regulatory pressures 54. Concerns about pollutants, odors 38, and allegations of illness and death related to land application 52 163 combined with projections of an increase in biosolid production rates 38 emphasizes the need for technologies that minimize biomass production in wastewater treatment facilities.

There are many technologies that can reduce biomass production rates but one technology in particular, chemical uncoupling, has shown considerable promise.

Chemical uncouplers disrupt the link between the catabolic (energy yielding oxidation) step and the formation of ATP and the production of new cells in anabolism 134 therefore limiting the cells ability to capture energy from substrate oxidation 98, 203.

Laboratory studies have shown that the application of the uncoupler technology to activated sludge units can result in a significant reduction in the biomass production rate 93, 99, 101, 199, 201-203. The most active uncoupler compounds are usually the most toxic and a serious problem with the uncoupler technology in wastewater treatment has been the potential presence of the compound in the effluent where it might be toxic to the environment 95. A number of uncoupler compounds, (e.g. DNP, chlorinated phenols) are designated as toxic pollutants subject to effluent limitations 63, 64; impeding their utility at a full-scale wastewater treatment facility. The uncoupling activity of anthraquinones in sulfate-reducing bacteria 29 and mitochondria 10 combined 139

with their low systemic toxicity 62 suggest that these compounds may have some utility for biomass reduction in wastewater treatment systems.

Reports of 40%, 50% even 60% reductions in biomass production in the presence of uncouplers are not uncommon in the literature. The viability of the technology at a full-scale treatment facility however, depends on the ability of the compounds to reduce operating costs above the chemical costs associated with the technology. Although critical, only a few reports have mentioned the economic impact associated with application of the uncoupler technology.

An initial economic assessment was performed that evaluated the impact of uncouplers on an industrial and two municipal wastewater treatment systems. In this analysis, operating values were developed for the industrial and municipal facilities 111,

150 (Table 1) that are similar to these facilities standard operating conditions and, using these values, the economic impact of the uncoupler was determined.

TABLE 1: OPERATING CONDITIONS FOR INDUSTRIAL AND MUNICIPAL WASTEWATER TREATMENT FACILITIES USED IN ECONOMIC 17, 111, 150 ANALYSIS. INFORMATION FROM

Facility Industrial Municipal 1 Municipal 2 Average flow rate (GPM) 1,100 6,319 817

BOD/COD loading (BOD5 lbs/day) 264,000 33,360 1572 Biomass generation/BOD or COD treated (lbs dry * biomass/lb BOD5 removed) 0.3 0.4 0.4 Biomass Disposal Landfill Land Application Incineration Disposal Costs ($/dry ton) $100 $50 $200 Excess plant capacity No Yes No Biomas Production lbs dry/day 79,200 13,344 572 140

The industrial facility is a completely mixed activated sludge unit composed of three 1.2 million gallon secondary aeration tanks. The hydraulic flow rate is 1,100 gallons per minute (GPM) (1.6 million GPD) with an organic loading of 20,000 mg/L

BOD5. The plant generates 79,200 lbs of dry biomass per day based on a yield of 0.3 pounds dry biomass per pound BOD5. The solids are separated by using a membrane filter press and are sent to a landfill. The plant does not have any excess capacity and is considering a capital expansion project 150.

Two municipal cases were considered in this analysis. The first municipal facility has a hydraulic flow rate of 6,319 GPM with an organic loading of 187 mg/L

BOD5. The plant generates 13,334 lbs biomass per day based on a yield of 0.4. The solids are sent to a digester that generates methane used to supplement on-site power requirements and then land applied 111. The second municipal facility has a hydraulic flow rate of 817 GPM with an average organic loading of 160 mg/L BOD5. The plant generates 572 lbs biomass per day based on a yield of 0.4. The solids generated from this facility are sent for incineration. The plant is operating at maximum capacity and is currently sending a portion of its wastewater to a local industrial facility for treatment 17.

The economic analysis initially considered a 10% reduction in biomass production rate and an estimated solids disposal cost of $100 per ton for the industrial

150, $50 per ton for the large municipal 129 and $200 per ton of dry solids for the small 141

municipal case 20. The dry solids contain approximately 55% moisture; therefore one ton of dry biomass was equivalent to two tons of dry solids or 4.44 tons of wet cake 150.

The chemical uncoupler, 1-hydroxyanthraquinone (OHAQ), supplied from the manufacturer (Epochem Company, Shanghai, China) in 50KG containers at a cost of

$20/KG OHAQ (purity 97%), was used at a dose of 10 mg/L wastewater, a dose shown in our studies to be effective in both anaerobic and aerobic microcosms.

Although the oxygen utilization rate in the presence of chemical uncouplers is increased, the oxygen consumed per unit BOD5 is constant. Therefore, the aeration requirement per BOD5 loading remained constant in this evaluation. The capital investment for the equipment (dry powder storage facility, slurry tank, pumps) and facilities (outbuilding) was estimated at $250,000 and $50,000, respectively. The project costs associated with front end loading (FEL) or engineering design of $75,000 and yearly maintenance ($5,000) were included. The front end loading of a project is the fee charged at the time of the initial purchase for an investment and is used as an estimate for initial engineering design of the project 177.

Net Present Value (NPV), a standard method for financial evaluations, was used to determine the impact of the uncoupler technology. NPV is used to analyze the profitability of an investment or project and is the difference of cash inflows and cash outflows 73 194. The cash flows of the project are discounted into present value (PV) amounts using a discount rate that represents the projects cost of capital and its risk 73.

The investment‘s future positive cash flows are reduced into one present value number 142

which is subtracted from the initial cash outlay required for the investment 73 194 according to:

t NPV = ∑ Ct/(1+r) – C0

where C is the cash flow, t is the time of cash flow, r is the is discount rate, C0 is initial cash investment. The discount rate is the rate at which the capital needed for the project would earn if invested in an alternative venture 194. This analysis used 1% for the selling price, 3% for variable cost and 3% fixed cost escalation factors and a 35%

Federal and 3% State/Local Tax rate 73.

The Net Present Value for the industrial treatment facility was calculated for biomass reductions ranging from 10% to 50% (Table 2). Although, the annual savings in solids disposal for a 10% reduction in biomass was $492,000 the NPV for the project was negative indicating the project would result in a negative cash flow. The

NPV is positive for a biomass reduction of 11% but the time required for the cash inflows to equal the original outlay is greater than 10 years. The NPV and discounted payback become more favorable as the biomass reductions exceed 15%. For biomass reductions of 20% or the NPV is $1,042,000 and the project moves to a positive cash flow basis after only 2 years.

143

TABLE 2: ECONOMIC ANALYSIS FOR INDUSTRIAL WASTEWATER TREATMENT FACILITY

Biomass Reduction Baseline 10% 11% 12% 15% 20% 30% 40% 50%

Yield 0.300 0.270 0.267 0.264 0.255 0.240 0.210 0.180 0.150

Biomass Production lbs/day 79,200 71,280 70,488 69,696 67,320 63,360 55,440 47,520 39,600

tons/day 29.6 26.6 26.3 26.0 25.1 23.7 20.7 17.7 14.8 Reagent Costs $1000/yr 448 448 448 448 448 448 448 448 448 Disposal costs $/day 13,126 11,813 11,682 11,551 11,157 10,501 9,188 7,876 6,563 $1000/yr 4,791 4,312 4,264 4,216 4,072 3,833 3,354 2,875 2,396 Annual Savings

$1000/yr 479 527 575 719 958 1,437 1,916 2,396 NPV ($1000) (-192) 55 302 1,042 2,271 4,735 7,199 9,667 Discounted Payback (yr) > 10 9 5 2 1 0 0 0

The results are very different for the municipal wastewater treatment facility

cases. In the larger municipal facility, the reagent costs are approximately $2.5 million

while the estimated cost savings from the solids disposal range from $40,000 to

$202,000 for biomass reductions of 10% and 50%, respectively (Table 3). In addition,

the municipal facility analyzed uses biomass from the facility to generate a portion of

the electrical requirements of the facility. It follows that a reduction in biomass to the

digesters would reduce the gas generation rate resulting in an increase in the amount of

electricity that must be purchased.

144

TABLE 3: ECONOMIC ANALYSIS FOR A MUNICIPAL WASTEWATER TREATMENT FACILITY WITH A HYDRAULIC FLOW RATE OF 6,319 GPM.

Biomass Reduction Baseline 10% 20% 30% 40% 50% Yield 0.400 0.360 0.320 0.280 0.240 0.200

Biomass Production lbs/day 13,344 12,010 10,675 9,341 8,006 6,672 tons/day 5.0 4.5 4.0 3.5 3.0 2.5 Reagent Costs $1000/yr 2,574 2,574 2,574 2,574 2,574 2,574 Disposal costs $/day 1,106 995 885 774 663 553 $1000/yr 404 363 323 283 242 202 Annual Savings $1000/yr 40 81 121 161 202

The lower flow rate municipal facility (817 GPM) resulted in encouraging results. The reagent costs for the 817 GPM facility are approximately $333,000/yr while the estimated cost savings from the solids disposal range from $341,000 to

$369,000 for biomass reductions of 10% and 50%, respectively (Table 4). The overall savings to the plant range from $8,000/yr to $36,000/yr not considering the cost of equipment and labor associated with the project.

145

TABLE 4: ECONOMIC ANALYSIS FOR A MUNICIPAL WASTEWATER TREATMENT FACILITY WITH A HYDRAULIC FLOW RATE OF 817 GPM.

Biomass Reduction Baseline 10% 20% 30% 40% 50% Yield 0.400 0.360 0.320 0.280 0.240 0.200

Biomass Production lbs/day 572 515 458 401 343 286 tons/day 0.2 0.2 0.2 0.1 0.1 0.1 Reagent Costs $1000/yr 333 333 333 333 333 333 Disposal costs $/day 190 171 152 133 114 95 $1000/yr 69 62 55 48 42 35 Annual Savings $1000/yr 341 348 355 362 369

Total Cost Savings $1000/yr 8 15 22 29 36

The results of the economic analysis indicate that the application of the uncoupler technology is dependent on the hydraulic flow and BOD5 loading as well as the reagent costs. The cost savings for moderate flow, high BOD5 wastewaters was sensitive to the reagent costs. In the industrial case used in this study, the uncoupler technology required a minimum biomass reduction of approximately11% for the cost savings to exceed the cost to implement the technology. In contrast, moderate flow and low BOD5 wastewaters required large amounts of reagent and resulted in minimal cost savings associated with the reduction in biomass production rates. In the moderate flow municipal case analyzed in this study, the application of the uncoupler technology would result in a deficit in excess of $2 million annually. The low flow and low BOD5 wastewaters in the second municipal case showed favorable economics at the 10% 146

biomass reduction level. In this case the biosolids were sent for incineration at a relatively high cost to the facility.

The results suggest that further improvements are required in uncoupler technology to achieve positive economics in all municipal wastewater treatment settings. If the base reagent cost or the required treatment dose can be decreased, a positive cash flow basis can be obtained for moderate flow municipal facilities (Figure

1). The implementation of the uncoupler technology generates a positive cash flow basis at a treatment dose of 1 mg/L at all reagent costs except $20.5/KG. Positive cash flows are also seen with reagent doses of 2.5 mg/L and 5 mg/L for reagent costs of

$2.5/KG and 2.5 mg/L for reagent costs of $5.1/KG.

This economic analysis used a very conservative approach and did not consider the potential of an increase in revenue for the facility. The enhanced degradation per unit biomass during uncoupled growth can increase the capacity of the facility without capital expenditures. These results suggest that the use of uncouplers for biomass reductions in wastewater treatment facilities has the potential to significantly decrease the operating expenses and eliminate capital expansion of capacity limited facilities. 147

FIGURE 1: THE YEARLY REAGENT COSTS FOR REAGENT DOSES FOR REAGENT DOSES OF 1 TO 10 MG/L AND REAGENT COSTS OF $2.5/KG TO $20.5/KG IN A MODERATE FLOW MUNICIPAL WASTEWATER TREATMENT FACILITY.

500 400 300 200 $1000/yr 100 0 Yearly Reagent Cost Cost Reagent Yearly 1 2.5 5 10 Reagent Dose, mg/L

$2.5/KG $5.1/KG $10.3/KG $20.5/KG

Yearly reagents costs in $1000/yr for a moderate flow municipal facility using uncoupler treatment doses of 1, 2.5, 5.0 and 10 mg/L and reagent costs of $2.5/KG (!) , $5.1/KG(!), $10.3/KG (!), and $20.5/KG (!). Cost savings for solids disposal for 50% biomass reduction identified with .

148

References

1. Betina, V. and S. Kuzela, Uncoupling effect of fungal hydroxyanthraquinones on mitochondrial oxidative phosphorylation. Chem.-Biol. Interact., 1987. 62(2): p. 179-89.

2. Bowman & Company, The Carneys Point Township Sewerage Authority, Report of Audit for the Fiscal Years Ended November 2005 and 2004. 2006, Bowman & Company: Woodbury. p. 41.

3. Burrowes, P. and T. Bauer. Energy Consideration with Thermal Processing of Biosolids. in Bioenergy Workshop. 2004: Water Environment Federation, Alexandria.

4. Constable, S.W., Solids disposal costs. 1995, Personal Communication.

5. Cooling III, F.B., et al., Inhibition of Sulfate Respiration by 1,8- Dihydroxyanthraquinone and Other Anthraquinone Derivatives. Applied and Environmental Microbiology, 1996. 62(8): p. 2999-30004.

6. EPA, Biosolids Generation, Use and Disposal in The United States. 1999, United States Environmental Protection Agency, Municipal and Industrials Solid Waste Division, Office of Solid Waste: Washington, DC. p. 74.

7. Gaudy Jr, A.F. and E.T. Gaudy, Mixed Microbial Populations. Advances in Biochemical Engineering, 1972: p. 97-143.

8. Gibb, T., A terrible waste gets long look, in Post-Gazette. 2000: Pittsburgh.

9. Goldstein, N., The State of Biosolids in America. Biocycle 2000. 41(1212): p. 50- 56.

10. Health, N.I.o., Hazardous Substances Databank Dinitrophenol. 2000, National Library of Medicine's TOXNET System. 149

11. Health, N.I.o., Hazardous Substances Databank Trichlorophenol. 2000, National Library of Medicine's TOXNET System.

12. Health, N.I.o., Hazardous Substances Databank Anthraquinone. 2000, National Library of Medicine's TOXNET System.

13. Investopedia, Net Present Value-NPV. 2007, Investopedia.com http://www.investopedia.com/terms/n/npv.asp.

14. Liu, Y., Effect of Chemical Uncoupler on the Observed Growth Yield in Batch Culture of Activated Sludge. Water Research, 2000. 34(7): p. 2025-2030.

15. Liu, Y. and J.-H. Tay, Strategy for Minimization of Excess Sludge Production from the Activated Sludge Process. Biotechnology Advances, 2001. 19: p. 97-107.

16. Liu, Y., Chemically Reduced Excess Sludge Production in the Activated Sludge Process. Chemosphere, 2003. 50: p. 1-7.

17. Low, E.W. and H.A. Chase. The Use of Uncouplers to Inhibit Biomass Production in Wastewater Treatment. in Icheme Research Event. 1996. Europe: Institute of Chemical Engineers.

18. Low, E.W. and H.A. Chase, The Use of Chemical Uncouplers for Reducing Biomass Production During Biodegradation. Water Science and Technology, 1998. 37(4/5): p. 399-402.

19. Low, E.W. and H.A. Chase, Reducing Production of Excess Biomass During Wastewater Treatment. Water Research, 1999. 33(5): p. 1119-1132.

20. Low, E.W., et al., Uncoupling of Metabolism to Reduce Biomass Production in the Activated Sludge Process. Water Research, 2000. 34(12): p. 3204-3212.

21. Miller, S., Municipal plant operations. 2006, Personal Communication. 150

22. NEBRA, A National Biosolids Regulation, Quality, End Use & Disposal Survery Preliminary Report. 2007, North East Biosolids and Residuals Association http://www.nebiosolids.org.

23. O'Dette, R.G. Determining the Most Cost Effective Option for Biosolids and Residual Management. in 10th Annual Residuals and Biosolids Management Conference: 10 Years of Progress and a Look Toward the Future. 1996: Water Environment Federation, Alexandria.

24. Okey, R.W. and H.D. Stensel, Uncouplers of Oxidative Phosphorylation. Modeling and Predicting Their Impact on Wastewater Treatment and the Environment, in ACS Symposium Series Emerging Technologies in Hazardous Waste Management V, D.W. Tedder and F.G. Pohland, Editors. 1995, American Chemical Society: Washington. p. 284-309.

25. Reich, R.A., "Typical" Industrial Wastewater Treatment Facility. 2006, Personal Communication.

26. Rocher, M., et al., Towards a Reduction in Excess Sludge Production in Activated Sludge Processes: Biomass Physicochemical Treatment and Biodegradation. Applied Microbiology and Biotechnology, 1999. 51: p. 883-890.

27. Seif, J.M., Report on the Investigation into the Application of Biosolids at the Al Hamilton Top Mine Site and the Death of Tony Behun. 2000, Pennsylvania Department of Environmental Protection http://www.dep.state.pa.us/dep/biosolids/Testimony/BiosolidsReport0500.htm.

28. Strand, S.E., G.N. Harem, and H.D. Stensel, Activated-Sludge Yield Reduction Using Chemical Uncouplers. Water Environment Research, 1999. 71(4): p. 454- 458.

29. TheFreeDictionary, Front End Load. 2007, The Free Dictionary. com http://financial-dictionary.thefreedictionary.com/Front-End+Load.

30. Thompson, J.W. and K. Sorvig, Sustainable Landscape Construction. 2000, Washington, D.C.: Island Press. 372. 151

31. Wheatley, A.D., Effluents, in Biodeterioration 7, D.R. Houghton, R.N. Smith, and H.O.W. Eggins, Editors. 1988, Cranfield Institute of Technology: Bedford. p. 193-206.

32. Wikipedia, Net present value. 2007, Wikipedia http://.en.wikipedia.org/wiki/Net_present_value.

33. Yang, X.-F., M.-L. Xie, and Y. Liu, Metabolic Uncouplers Reduce Excess Sludge Production in an Activated Sludge Process. Process Biochemistry, 2003. 38: p. 1373-1377.

34. Ye, F.X., D.S. Shen, and Y. Li, Reduction in Excess Sludge Production by Addition of Chemical Uncouplers in Activated Sludge Batch Cultures. Journal of Applied Microbiology, 2003. 95: p. 781-786.

35. Ye, F.X. and Y. Li, Uncoupled Metabolism Stimulated by Chemical Uncoupler and Oxic-Settling-Anaerobic Combined Process to Reduce Excess Sludge Production. Applied Biochemistry and Biotechnology, 2005. 127(3): p. 187-199.

36. Ye, F.X. and Y. Li, Reduction of Excess Sludge Production by 3,3',4',5- Tetrachlorosalicylanilide in an Activated Sludge Process. Applied Microbiology and Biotechnology, 2005. 67: p. 269-274.

152

Chapter 6 Summary and Conclusions

The addition of 1-hydroxyanthraquinone (OHAQ) to aerobic and anaerobic batch cultures of Pa. denitrificans resulted in classic ―uncoupling‖ activity characterized by an increase in the respiration rate and decrease in biomass production.

The compound was active in aerobic and anaerobic cultures of Paracoccus, batch continuous Paracoccus cultures and in mixed cultures derived from a municipal wastewater treatment facility.

At a test concentration of 10 mg/L, OHAQ addition increased the nitrogen production rate 45.7 % (p= 2.29 x 10-2) and decreased the biomass production 24.5 %

(p=1.33 x 10-3) in anaerobic cultures using nitrate as the terminal electron acceptor.

The activity was comparable using either nitrate or nitrite as the terminal electron acceptor. In cultures using nitrite as the terminal electron acceptor, OHAQ addition

-3 increased the nitrogen production rate 64.5% (p= 6.66 x 10 ) and decreased the biomass production 30.7% (p=2.06 x 10-5) compared to the control. The unsubstituted parent compound, which lacks a hydroxyl group at position 1 of the ring system, did not result in a decrease in biomass production or increase in nitrogen production rates.

In aerobic batch cultures, OHAQ at a test concentration of 10 mg/L, increased the initial oxygen utilization rate 38.5% (p= 3.05 x 10-4) and decreased the biomass production 39.3% (p = 3.13 x 10-2). In studies using anthraquinone compounds with different ring substitutions, the biomass reduction was found to be related to both the 153

positional and electronic properties of the substitution(s). Electron donating groups

(EDG), such as hydroxyl or amino substitutions, in position 1 (ortho) resulted in a significant reduction in biomass production. Conversely, substitution of either a hydroxyl or amino group at position 2 (para) resulted in almost the complete inhibition of growth and respiration.

OHAQ addition to a continuous Paracoccus culture resulted in an increase in the respiratory activity compared to the steady state control values and was sustained over the duration of the experiment. There was a 10.4% in the carbon evolution rate

(CER) and a 10.8 % increase in the specific oxygen utilization rate (SOUR). The growth of the organism was not adversely affected by the presence of OHAQ and the difference in the biomass concentrations of the control and OHAQ reactors was not significant. These results suggest that the compound could increase the capacity of wastewater treatment facilities without modification in or addition to existing equipment.

The addition OHAQ to a mixed cultures derived from a municipal wastewater treatment facility resulted in an increase in the initial oxygen utilization rate and a reduction in the overall biomass production. The overall oxygen utilization was not adversely affected by OHAQ. In test vessels containing 300 mg/L MLSS, 10 mg/L

OHAQ increased the initial oxygen consumption rate 16.9% (p=3.32 x 10-3) and decreased the biomass production 10.8 % (p = 7.76 x 10-3) compared to the control bottles. The biomass reduction in the mixed culture batch studies was less than that obtained in the Paracoccus batch studies. This value, although lower than that 154

obtained in the pure culture studies, is significant in regards to the relative biomass production produced in wastewater treatment facilities.

These studies demonstrate that the activity of OHAQ is the result of the

―uncoupling‖ of aerobic and anaerobic respiration. The resulted reported here indicate that OHAQ has the potential to reduce both the capital and operating costs of biological wastewater treatment systems, especially in the industrial setting, by simultaneously improving the metabolic capacity and decreasing the biomass production rates. The uncoupling activity of anthraquinones combined with the low toxicity of compounds suggests that these compounds may be used for reducing biomass production in commercial bioprocesses in addition to traditional wastewater treatment systems. The need for efficient, cost-effective and environmentally sound waste treatment in both the municipal and industrial markets warrants the further investigation of the use of this chemical uncoupler for enhancement of metabolic capacity and biomass reductions.

155

Chapter 7 Barriers to Implementation

The concept of uncouplers for biomass reduction in activated sludge systems is

not new, and while there is a large amount of literature on uncouplers, reports on their

application in activated sludge systems is limited. The results from laboratory

investigations indicate that OHAQ has the potential to reduce both the capital and

operating costs of biological wastewater treatment systems by simultaneously

improving the metabolic capacity and decreasing the biomass production rates.

Although uncouplers have the potential to significantly reduce biomass production, the

barriers to their implementation at the plant scale include: lack of pilot scale

demonstrations and long term operating data, bacterial acclimation and potential

discharge of the chemical into the environment.

Lack of Pilot Scale Demonstrations

There is an abundance of literature on the topic of uncouplers but just a handful

of investigations into the use of uncouplers in WWT systems. The lack of information

on the use of uncouplers at the pilot or full scale is at best discouraging. Although

there have been reports on the ―evidence‖ or the ―likelihood‖ of uncoupling in full-

scale treatment systems, most of the information is speculative and lacks a solid 156

scientific investigation. For example, Okey and Stensel 134 observed behavior

indicative of uncoupling at a full-scale activated sludge system but could not detect any

organic uncouplers at a high enough concentration to produce these effects. In their

analysis of 23rd Avenue Treatment Plant in Phoenix, they found a decrease in the

biomass production from 0.25-0.35 lb MLSS/lb COD to zero and a dramatic increase

in the oxygen requirement while the COD removal remained the same or was slightly

increased. Reports of reductions in substrate uptake rates 95, 101, 133, 203, population

shifts, and reduced settling characteristics 23, 101, 203 that adversely affect laboratory

reactor performance merit further investigations on the use of these compounds for

biomass reduction. In addition, the lack of long term pilot scale demonstrations

impedes the implementation of this technology at either an industrial or municipal

facility.

Acclimation

The primary objective of the uncoupler technology is a stable long-term

biomass reduction in a biological treatment system. However, there have been reports

in the literature describing the adaptation and/or acclimation of pure and mixed cultures

to uncouplers. Gage and Neidhardt 49, in experiments with E. coli, found DNP

exposure initially leads to a decrease in the growth rate but that the cultures eventually

recovered to near normal rates. Krulwich et al 88 found mutants of the Bacillus species, 157

originally isolated in the presence of CCCP, were able to synthesize substantial amounts of ATP at uncoupler-reduced P levels that would not support ATP synthesis by the wild type. In continuous flow experiments using activated sludge, Strand and

Stensel 172 observed reductions in yields declined significantly when microbial populations developed which were resistant to the uncoupler, TCP. Harem 60 found the effluent soluble TCP concentration dropped from 2mg/l to an undetectable level between day 75 and 82 in the laboratory reactor study, indicating the development of uncoupler-degrading capability in the bacterial consortium. In their studies of nitrite inhibition of denitrification in Pseudomonas fluorescens, Almeida et al 6 observed uncoupling by nitrous acid in batch systems but not in continuous systems and concluded there was a change in the physiology of the cells with time. The potential for acclimation to chemical uncouplers may be circumvented by alternative dosing regiments or alternating chemical compounds. While it is possible that long term exposure to chemical compounds would encourage the development of a resistant or uncoupler-insensitive population a protocol of bulk or intermittent dosing or alternating of chemical uncoupler compounds may discourage the development of an acclimated population. 158

Effluent Discharges to Environment

The most active uncoupler compounds are generally the most toxic and a

serious problem with the uncoupler technology is the potential presence of the

compound in the effluent. A number of uncoupler compounds, (e.g. DNP, chlorinated

phenols) are designated as toxic pollutants and are subject to effluent limitations under

section 307 of the Clean Water Act 63, 64.

Harem 60 in his studies with 2,4,5-trichlorophenol (245TCP) observed that the

soluble uncoupler concentration was an important operational parameter. The required

effluent concentration for uncoupling with 245TCP was 2 mg/L, which is

approximately the recommended water quality criterion of 2.6 ml/L for the protection

of human health 64. Shah et al 164 found a DNP concentration of 0.92 mg/L was

required for uncoupling in a continuous treatment system. The treatment standard

suggested by Federal guidelines is 70 g/L 63, an order of magnitude below Shah‘s

concentration. 2,4-Dinitrophenol is highly toxic to humans with a lethal oral dose of

14 to 43 mg/kg and has a high acute toxicity, LD50 (rats) of 30 mg/kg, based on short-

term animal tests 153, 168. The toxic nature of DNP and its stringent effluent standard

hamper its utility as an uncoupler at a full-scale wastewater treatment facility.

The anthraquinones have uncoupling activity that resembles that of the classic

10 62 uncouplers but have a low systemic toxicity . The LD50 for anthraquinone is 159

reported to be greater than 5000 mg/kg and in a 90 day feeding trial in rats receiving up to 15mg anthraquinone/kg resulted in no ill-effects153. The low toxicity of the anthraquinone compounds combined with uncoupling activity in bacterial systems may provide an advantage over the classic uncouplers in large scale systems. The laboratory investigations in aerobic and anaerobic test systems suggest that these compounds may be used for reducing biomass production in commercial bioprocesses in addition to traditional wastewater treatment systems. The need for efficient, cost- effective and environmentally sound waste treatment in both the municipal and industrial markets warrants the further investigation of the use of this chemical uncoupler for enhancement of metabolic capacity and biomass reductions. 160

References

1. Ahmad, Z. and A.E. Senior, Identification of Phosphate Binding Residues of Escherichia coli ATP Synthase. Journal of Bioenergetics and Biomembranes, 2005. 37(6): p. 437-440. 2. Alberts, B., et al., Molecular Biology of the Cell. Second Edition ed. 1989, New York: Garland Publishing, Inc. 1219. 3. Alefounder, P.R. and S.J. Ferguson, The Location of Dissimilatory Nitrite Reductase and the Control of Dissimilatory Nitrate Reductase by Oxygen in Paracoccus denitrificans. Biochem. J., 1980. 192: p. 231-240. 4. Alefounder, P.R., et al., Selection and Organisation of Denitrifying Electron- Transfer Pathways in Paracoccus denitrificans. Biochimica et Biophysica Acta, 1983. 724: p. 20-39. 5. Almasifar, D., et al., Spectrophotometric Determination of Acidity Constants of Some Recently Synthesized Anthraquinones in Methanol + Water. Journal of Chemical Engineering Data, 1997. 42: p. 1212-1215. 6. Almeida, J.S., et al., Nitrite Inhibition of Denitrification by Pseudomonas fluorescens. Biotechnology and Bioengineering, 1995. 46(3): p. 194-201. 7. Anke, H., et al., Metabolic Products of Microorganisms. 185*. The Anthraquinones of the Aspergillus glaucus Group. I. Occurrence, Isolation, Identification and Antimicrobial Activity. Archives of Microbiology, 1980. 126: p. 223-230. 8. Anraku, Y., Bacterial Electron Transport Chains. Annual Review of Biochemistry, 1988. 57: p. 101-32. 9. Bakola-Christianopoulou, M.N., L.B. Ecateriniadou, and K.J. Sarrris, Evaluation of the Antimicrobial Activity of a New Series of Hydroxy-Quinone Chelates of Some Transition Metals. European Journal of Medicinal Chemistry - Chim. Ther, 1986. 21(5): p. 385-390. 10. Betina, V. and S. Kuzela, Uncoupling effect of fungal hydroxyanthraquinones on mitochondrial oxidative phosphorylation. Chem.-Biol. Interact., 1987. 62(2): p. 179-89. 11. Boogerd, F.C., H.W. van Verseveld, and A.H. Stouthamer, Electron Transport to Nitrous Oxide in Paracoccus denitrificans. FEBS Letters, 1980. 113(2): p. 279- 284. 161

12. Boogerd, F.C., H.W. Van Verseveld, and A.H. Stouthamer, Respiration-Driven Proton Translocation with Nitrite and Nitrous Oxide in Paracoccus denitrificans. Biochimica et Biophysica Acta, 1981. 638(2): p. 181-191. 13. Boogerd, F.C., K.J. Appeldoorn, and A.H. Stouthamer, Effects of electron transport inhibitors and uncouplers on denitrification in Paracoccus denitrificans. FEMS Microbiol. Lett., 1983. 20(3): p. 455-60. 14. Boogerd, F.C., H.W. Van Verseveld, and A.H. Stouthamer, Dissimilatory nitrate uptake in Paracoccus denitrificans via a .DELTA.~.mu.H+-dependent system and a nitrate-nitrite antiport system. Biochim. Biophys. Acta, 1983. 723(3): p. 415- 27. 15. Boogerd, F.C., et al., Reconsideration of the efficiency of energy transduction in Paracoccus denitrificans during growth under a variety of culture conditions. A new approach for the calculation of P/2e- ratios. Arch. Microbiol., 1984. 139(4): p. 344-50. 16. Boos, K.S. and E. Schlimme, Anthraquione Dyes: A New Class of Potent Inhibitors of Mitochrondrial Adenine Nucleotide Translocation and Oxidative Phosphorylation. FEBS Letters, 1981. 127(1): p. 40-44. 17. Bowman & Company, The Carneys Point Township Sewerage Authority, Report of Audit for the Fiscal Years Ended November 2005 and 2004. 2006, Bowman & Company: Woodbury. p. 41. 18. Brown, J.P., A Review of the Genetic Effects of Naturally Occurring Flavonoids, Anthraquinones and Related Compounds. Mutation Research, 1980. 75: p. 243- 277. 19. Bryan, B.A., Physiology and Biochemistry of Denitrification, in Denitrification, Nitrification, and Atmospheric Nitrous Oxide, C.C. Delwiche, Editor. 1981, John Wiley & Sons: New York. p. 67-84. 20. Burrowes, P. and T. Bauer. Energy Consideration with Thermal Processing of Biosolids. in Bioenergy Workshop. 2004: Water Environment Federation, Alexandria. 21. Calder, K., K.A. Burke, and J. Lascelles, Induction of Nitrate Reductase and Membrane Cytochromes in Wild Type and Chlorate-Resistant Paracoccus denitrificans. Archives of Microbiology, 1980. 126: p. 149-153. 22. Capaldi, R.A. and B. Schulenberg, The subunit of bacterial and chloroplast F1F0 ATPases Structure, arrangement, and role of the subunit in energy coupling withing the complex. Biochimica et Biophysica Acta, 2000. 1458: p. 263-269. 23. Chen, Y.X., F.X. Ye, and X.S. Feng, The use of 3,3',4',5-tetrachlorosalicylanilide as a chemical uncoupler to reduce activated sludge yield. Journal of Chemical Technology and Biotechnology, 2004. 79: p. 111-116. 162

24. Chudoba, P., J. Chudoba, and B. Capdeville, The Aspect of Energetic Uncoupling of Microbial Growth in the Activated Sludge Process-OSA System. Water Science & Technology, 1992. 26(9-11): p. 2477-2480. 25. Chung, R.H., Anthraquinone, in Kik-Othmer Enclycopedia of Chemical Technology. 1978, John Wiley & Sons: New York. p. 700-757. 26. Clowes, G.H.A. and M.E. Krahl, Studies on Cell Metabolism and Cell Division I. On the Relation between Molecular Structures, Chemical Properties, and Biological Activities of the Nitrophenols. Journal of General Physiology, 1936. 20: p. 145-171. 27. Constable, S.W., Solids disposal costs. 1995, Personal Communication. 28. Cook, S.R., Steve's place-Reaction Mechanisms. 2002, http://www.steve.gb.com/science/reaction_mechanisms.html. 29. Cooling III, F.B., et al., Inhibition of Sulfate Respiration by 1,8- Dihydroxyanthraquinone and Other Anthraquinone Derivatives. Applied and Environmental Microbiology, 1996. 62(8): p. 2999-30004. 30. Cox, C.D. and W.J. Payne, Separation of Soluble Denitrifying Enzymes and Cytochromes from Pseudomonas perfectomarinus. Canadian Journal of Microbiology, 1973. 19: p. 861-872. 31. Cox, R.B. and J.R. Quayle, The Autotrophic Growth of Micrococcus denitrificans on Methanol. Biochemical Journal, 1975. 150: p. 569-571. 32. Craske, A. and S.J. Ferguson, The Respiratory Nitrate Reductase from Paracoccus denitrificans. Journal of Biochemistry, 1986. 158: p. 429-436. 33. Cross, R.L., The Mechanism and Regulation of ATP Synthesis by F1-ATPases. Biochemistry, 1981. 50: p. 681-714. 34. Cudlin, J., et al., Biological activity of hydroxyanthraquinones and their glucosides toward microorganisms. Folia Microbiol., 1976. 21(1): p. 54-7. 35. Das, A. and L.G. Ljungdahl, Composition and Primary Structure of the F1F0 ATP Synthase from the Obligately Anaerobic Bacterium Clostridium thermoaceticum. Bacteriology, 1997. 179(11): p. 3746-3755. 36. Delwiche, C.C., Denitrification, in Inorganic Nitrogen Metabolism: Function of Metallo-Flavoprotein, W.D. McElroy and B. Glass, Editors. 1956, John Hopkins Press: Baltimore. p. 233-256.

37. Duncan, T.M., Rotation of subunits during catalysis by Escherichia coli F1- ATPase. Proceeding of the National Academy of Sciences, 1995. 92(10964- 10968). 38. EPA, Biosolids Generation, Use and Disposal in The United States. 1999, United States Environmental Protection Agency, Municipal and Industrials Solid Waste Division, Office of Solid Waste: Washington, DC. p. 74. 163

39. EPA, EPI Suite. 2000, http://www.epa.gov/opptintr/exposure/pubs/episuite.htm. 40. Ferguson, S. The Redox Reactions of the Nitrogen and Sulphur Cycles. in Nitrogen and Sulphur Cycles Symposium 42. 1988: Society for General Microbiology. 41. Ferguson, S.J., et al., Selective and Reversible Inhibition of the ATPase of Micrococcus denitrificans by 7-Chloro-4-nitrobenzo-2-oxa-1,3 diazole. Biochimica et Biophysica Acta, 1974. 357: p. 457-461. 42. Ferguson, S.J., Denitrification: a question of the control and organization of electron and ion transport. Trends Biochem. Sci., 1987. 12(9): p. 354-7. 43. Ferraro, D.J., L. Gakhar, and S. Ramaswamy, Rieske business: Structure-function of Rieske non-heme oxygenases. Biochemical and Biophysical Research Communications, 2005. 338: p. 175-190. 44. Fillingame, R.H., Biochemistry and Genetics of Bacterial H+ - Translocating ATPases. Bioenergetics, 1981. II: p. 35-106. 45. Fillingame, R.H., Membrane Sectors of F- and V-type H+-transporting ATPases. Current Opinion in Structural Biology, 1996. 6: p. 491-498. 46. Fukami, H. and M. Nakajima, Rotenone and the Rotenoids, in Naturally Occurring Insecticides, M. Jacobson and D.G. Crosby, Editors. 1971, Marcel Dekker, Inc.: New York. p. 71-97. 47. Futai, M., Reconstitution of ATPase Activity From the Isolated , , and Subunits of the Coupling Factor, F1, of Escherichia coli. Biochemical and Biophysical Research Communications, 1977. 79(4): p. 1231. 48. Futai, M., et al., Synthase (H+ ATPase): coupling between catalysis, mechanical work and proton translocation. Biochimica et Biophysica Acta, 2000. 1458: p. 276-288. 49. Gage, D.J. and F. Neidhardt, Adaptation of Escherichia coli to the Uncoupler of Oxidative Phosphorylation 2,4 - Dinitrophenol. Journal of Bacteriology, 1993. 175(21): p. 7105-7108. 50. Garcia-Lopez, P.M., L. Kung Jr, and J.M. Odom, In Vitro Inhibition of Microbial Methane Production by 9,10-Anthraquinone. Journal of Animal Science, 1996. 74: p. 2276-2284. 51. Gaudy Jr, A.F. and E.T. Gaudy, Mixed Microbial Populations. Advances in Biochemical Engineering, 1972: p. 97-143. 52. Gibb, T., A terrible waste gets long look, in Post-Gazette. 2000: Pittsburgh.

53. Gogol, E.P., et al., Ligand-dependent structural variations in Escherichia coli F1 ATPase revealed by cryoelectron microscopy. Proceeding of the National Academy of Sciences, 1990. 87: p. 9585-9589. 164

54. Goldstein, N., The State of Biosolids in America. Biocycle 2000. 41(1212): p. 50- 56. 55. Grady Jr, C.P.L. and H.C. Lim, Biological Wastewater Treatment: Theory and Applications. Pollution Engineering and Technology, ed. P.N. Cheremisinoff. Vol. 12. 1980, New York: Marcel Dekker, Inc. 963. 56. Groth, G., Molecular Models of the Structural Arrangement of Subunits and the Mechanism of Proton Translocation in the Membrane Domain of F1F0 ATP synthase. Biochimica et Biophysica Acta, 2000. 1458: p. 417-427. 57. Haddock, B.A. and C.W. Jones, Bacterial Respiration. Bacteriological Reviews, 1977. 41(1): p. 47-99. 58. Hall, N., Anthraquinone dyes containing a fluorosulfonyl group and their use. 1995, Zeneca Ltd., UK: PCT Int. Appl. p. 17. 59. Haraux, F. and Y. De Kouchkovsky, Energy Coupling and ATP Synthase. Photosynthesis Research, 1998. 57(3): p. 231-251. 60. Harem, G.N., Investigation of Chemical Uncouplers to Reduce Activated Sludge Yield, in Civil Engineering. 1995, University of Wahington. 61. Harold, F.M., Conservation and Transformation of Energy by Bacterial Membranes. Bacteriological Reviews, 1972. 36(2): p. 172-230. 62. Health, N.I.o., Hazardous Substances Databank Anthraquinone. 2000, National Library of Medicine's TOXNET System. 63. Health, N.I.o., Hazardous Substances Databank Dinitrophenol. 2000, National Library of Medicine's TOXNET System. 64. Health, N.I.o., Hazardous Substances Databank Trichlorophenol. 2000, National Library of Medicine's TOXNET System. 65. Hennessy, M.J., The activity of electron donating substitutions on aromatic structures. 2007, Personal Communication. 66. Heytler, P.G., Uncouplers of Oxidative Phosphorylation. Methods in Enzymology, 1979. LV: p. 462-473. 67. Hilgert, I., et al., Antitumor and immunosuppressive activity of hydroxyanthraquinones and their glucosides. Folia Biol., 1977. 23(2): p. 99-109. 68. Hinkle, P.C. and R.E. McCarty, How Cells Make ATP. Scientific American, 1978. 238: p. 104-123. 69. Hochstein, L.I. and G.A. Tomlinson, The Enzymes Associated with Denitrification. Annual Review of Microbiology, 1988. 42: p. 231-261. 70. House, R.V. and J.H. Dean, Structural requirements for suppression of cytotoxic T lymphocyte induction by polycyclic compounds. In Vitro Toxicol., 1987. 1(3): p. 149-62. 165

71. Imai, K., A. Asano, and R. Sata, Oxidative phosphorylation in Micrococcus denitrificans I. Preparation and properties of phosphorylating membrane fragments. Biochim. Biophys. Acta, 1967. 143: p. 462-476. 72. Ingraham, J.L., Microbiology and Genetics of Denitrifiers, in Denitrification, Nitrification, and Atmospheric Nitrous Oxide, C.C. Delwiche, Editor. 1981, Wiley-Interscience Publication: Davis. p. 45-65. 73. Investopedia, Net Present Value-NPV. 2007, Investopedia.com http://www.investopedia.com/terms/n/npv.asp. 74. Jassal, B., Electron Transport Chain. 2005. 75. John, P. and F.R. Whatley, Paracoccus denitrficans and the Evolutionary Origin of the Mitochondrion. Nature, 1975. 254: p. 495-498. 76. John, P., Aerobic and Anaerobic Bacterial Respiration Monitored by Electrodes. Journal of General Microbiology, 1977. 98: p. 231-238. 77. John, P. and F.R. Whatley, The Bioenergetics of Paracoccus denitrificans. Biochemica et Biophysica Acta, 1977. 463: p. 129-153. 78. Jones, R.W., A. Lamont, and P.B. Garland, The Mechanism of Proton Translocation Driven by the Respiratory Nitrate Reductase Complex of Escherichia coli. Biochem. J., 1980. 190: p. 79-94. 79. Kaniuga, Z., J. Bryla, and E.C. Slater, Mechanism of Interaction between Antimycin and the Respiratory Chain. FEBS Symposium, 1969. 19: p. 285-295. 80. Karukstis, K.K., et al., Alternative Measures of Photosystem II Electron Transfer Inhibition in Anthraquinone-Treated Chloroplasts. Photochemistry and Photobiology, 1992. 55(1): p. 125-132. 81. Kavanagh, F., Activities of Twenty-Two Antibacterial Substances Against Nine Species of Bacteria. Journal of Bacteriology, 1947. 54: p. 761-766. 82. Kawai, K., et al., The inhibition of mitochondrial respiration by anthraquinone mycotoxins. The relationship between the position of hydroxyl group on the anthraquinone nucleus and the uncoupling effect on oxidative phosphorylation. Proc. Jpn. Assoc. Mycotoxicol., 1987. 26: p. 43-6. 83. King, M.W., Oxidative Phosphorylation. 2006, Indiana State University School of Medicine. 84. Knowles, R., Denitrification. Microbiological Reviews, 1982. 46(1): p. 43-70. 85. Kotlyar, A.B., Albracht, Simon P.J., and R.J.M. van Spanning, Comparison of Energizatin of Complex I in Membrane Particles from Paracoccus denitrificans and Bovine Heart Mitochondria. Biochimica et Biophysica Acta, 1998. 1365: p. 53-59. 166

86. Krahl, M.D. and G.H.A. Clowes, Studies on Cell Metabolism and Cell Division II. Stimulation of Cellular Oxidation and Reversible Inhibition of Cell Division by Dihalo and Trihalophenols. Journal of General Physiology, 1936. 20: p. 173-184. 87. Kristjansson, J.K., B. Walter, and T.C. Hollocher, Respiration-Dependent Proton Translocation and the Transport of Nitrate and Nitrite in Paracoccus denitrificans and Other Denitrifying Bacteria. Biochemistry, 1978. 17(23): p. 5014-5019. 88. Krulwich, T.A., P.G. Quirk, and A.A. Guffanti, Uncoupler - Resistant Mutants of Bacteria. Microbiological Reviews, 1990. 54(1): p. 52-65. 89. Kucera, I. and P. Kaplan, A Study on the Transport and Dissimilatory Reduction of Nitrate in Paracoccus denitrificans using Viologen Dyes as Electron Donors. Biochimica et Biophysica Acta, 1996. 1276: p. 203-209. 90. Lam, Y. and D.J.D. Nicholas, Nitrite reductase with cytochrome oxidase activity from Micrococcus denitrificans. Biochim. Biophys. Acta, 1969. 180(3): p. 459- 72. 91. Link, T.A., The role of the 'Rieske' iron sulfur protein in the hydroquinone oxidation (QP) site of the cytochrome bc1 complex. FEBS Letters, 1997. 412: p. 257-264. 92. Linnett, P.E. and R.B. Beechey, Inhibitors of the ATP Synthetase System. Methods in Enzymology, 1979. LV: p. 472-519. 93. Liu, Y., Effect of Chemical Uncoupler on the Observed Growth Yield in Batch Culture of Activated Sludge. Water Research, 2000. 34(7): p. 2025-2030. 94. Liu, Y. and J.-H. Tay, Strategy for Minimization of Excess Sludge Production from the Activated Sludge Process. Biotechnology Advances, 2001. 19: p. 97-107. 95. Liu, Y., Chemically Reduced Excess Sludge Production in the Activated Sludge Process. Chemosphere, 2003. 50: p. 1-7. 96. Loomis, W.F. and F. Lipmann, Reversible Inhibition of the Coupling Between Phosphorylation and Oxidation. Biological Chemistry, 1948. 173: p. 807-808. 97. Low, E.W. and H.A. Chase. The Use of Uncouplers to Inhibit Biomass Production in Wastewater Treatment. in Icheme Research Event. 1996. Europe: Institute of Chemical Engineers. 98. Low, E.W. and H.A. Chase, The Use of Chemical Uncouplers for Reducing Biomass Production During Biodegradation. Water Science and Technology, 1998. 37(4/5): p. 399-402. 99. Low, E.W. and H.A. Chase, Reducing Production of Excess Biomass During Wastewater Treatment. Water Research, 1999. 33(5): p. 1119-1132. 167

100. Low, E.W., et al., The Use of Chemical Uncouplers for Reducing Biomass Production in the Activated Sludge Process. Water Research, 2000. 12: p. 3204- 3212. 101. Low, E.W., et al., Uncoupling of Metabolism to Reduce Biomass Production in the Activated Sludge Process. Water Research, 2000. 34(12): p. 3204-3212. 102. Macmillan, J.G., Electrophilic Aromatic Substitution. 1999, http://faculty.cns.uni.edu/~macmilla/mcmurry/mcmurry_chapter_16/sld. 103. Maloney, P.C., Energy Coupling to ATP Synthesis by the Proton-Translocating ATPase. Journal of Membrane Biology, 1982. 67: p. 1-12. 104. Mayhew, M. and T. Stephenson, Low Biomass Yield Activated Sludge: A Review. Environmental Technology, 1997. 18: p. 883-892. 105. Mayhew, M. and T. Stephenson, Biomass Yield Reduction: Is Biochemical Manipulation Possible Without Affecting Activated Sludge Process Efficiency? Water Science & Technology, 1998. 38(8-9): p. 137-144. 106. McEwan, A.G., et al., Nitrous Oxide Reduction by Members of the Family Rhodospirillaceae and the Nitrous Reductase of Rhodopseudomonas capsulata. Journal of Bacteriology, 1985. 164(2): p. 823-830. 107. McLaughlin, S.G. and J.P. Dilger, Transport of Protons Across Membranes by Weak Acids. Physiological Reviews, 1980. 60(3): p. 825-863. 108. Meijer, E.M., et al., Energy Conservation during Aerobic Growth in Paracoccus denitrificans. Archives of Microbiology, 1977. 112: p. 25-34. 109. Meijer, E.M., et al., Effects Induced by Rotenone During Aerobic Growth of Paracoccus denitrificans in Continuous Culture. Microbiology, 1978. 119: p. 119-127. 110. Meijer, E.M., et al., Effects Induced by Rotenone during Aerobic Growth of Paracoccus denitrificans in Continuous Culture. Archives of Microbiology, 1978. 119: p. 119-127. 111. Miller, S., Municipal plant operations. 2006, Personal Communication. 112. Millis, N.F., Effect of Uncoupling Oxidative Phosphorylation on the Economics of the Activated Sludge Process. Chem. Eng. Common., 1986. 45: p. 135-144. 113. Mitchell, P. and J. Moyle, Stoichiometry of Proton Translocation Through the Respiratory Chain and Adenosine Triphosphatase Systems of Rat Liver Mitochondria. Nature, 1965. 208: p. 147-151. 114. Mitchell, P., Chemiosmotic Coupling in Oxidative and Photosynthetic Phosphorylation. Biological Reviews, 1966. 41: p. 445-502. 168

115. Mitchell, P. and J. Moyle, Proton Translocation Coupled to ATP Hydrolysis in Rat Liver Mitochondria. European Journal of Biochemistry, 1968. 4(4): p. 530- 539. 116. Mitchell, P. and J. Moyle, Estimation of Membrane Potential and pH Difference across the Cristae Membrane of Rat Liver Mitochondria. European Journal of Biochemistry, 1969. 7(4): p. 471-484. 117. Mitchell, P., Cation-Translocating Adenosine Triphosphate Models: How Direct is the Participation of Adenosine Triphosphate and its Hydrolysis Products in Cation Translocation? FEBS Letters, 1973. 33(3): p. 267-274. 118. Mitchell, P., A Chemiosmotic Molecular Mechanism for Proton-Translocating Adenosine Triphosphatases. FEBS Letters, 1974. 43(2): p. 189-194. 119. Moat, A.G. and J.W. Foster, Microbial Physiology. Second Edition ed. 1988, New York: John Wiley & Sons. 597. 120. Moreau, F. and J.D. De Virville, Stoichiometry of Proton Translocation Coupled to Substrate Oxidation in Plant Mitochondria. Plant Physiology, 1985. 77: p. 118- 123. 121. Muneyuki, E., et al., F0F1-ATP synthase: general structural features of "ATP- engine" and a problem on free energy transduction. Biochimica et Biophysica Acta, 2000. 1458: p. 467-481. 122. Namslauer, A., et al., Redox-Coupled Proton Translocation in Biological Systems: Proton Shuttling Cytochrome c Oxidase. Proceeding of the National Academy of Sciences, 2003. 100(26): p. 15543-15547. 123. NEBRA, A National Biosolids Regulation, Quality, End Use & Disposal Survery Preliminary Report. 2007, North East Biosolids and Residuals Association http://www.nebiosolids.org. 124. Negrin, R.S., D.L. Foster, and R.H. Fillingame, Energy-transducing H+-ATPase of Escherichia coli. The Journal of Biological Chemistry, 1980. 255(12): p. 5643- 5648. 125. Neidhardt, F.C., J.L. Ingraham, and M. Schaechter, Physiology of the Bacterial Cell. 1990, Sunderland, Massachusetts: Sinauer Associates, Inc. Publishers. 126. Neijssel, O.M., The Effect of 2,4-Dinitrophenol on the Growth of Klebsiella aerogenes NCTC 418 in Aerobic Chemostat Cultures. FEMS Letters, 1977. 1: p. 47-50. 127. Nicholls, D.G. and S.J. Ferguson, Bioenergetics 2. 1992, New York: Academic Press. 255. 128. Nishio, A. and E.M. Uyeki, Cellular uptake and inhibition of DNA synthesis by dihydroxyanthraquinone and two analogs. Cancer Res., 1983. 43(5): p. 1951-6. 169

129. O'Dette, R.G. Determining the Most Cost Effective Option for Biosolids and Residual Management. in 10th Annual Residuals and Biosolids Management Conference: 10 Years of Progress and a Look Toward the Future. 1996: Water Environment Federation, Alexandria. 130. Odom, J.M., et al., Immunological Cross-Reactivities of Adenosine-5'- Phosphosulfate Reductases from Sulfate-Reducing and Sulfide-Oxidizing Bacteria. Applied and Environmental Microbiology, 1991. 57(3): p. 727-733. 131. Odom, J.M., Anthraquinone inhibition of methane production in methanogenic bacteria, in 42 pp. 1994, Bio-Technical Resources LP, USA: PCT Int. Appl. 132. Oettmeier, W., K. Masson, and A. Donner, Anthraquinone inhibitors of photosystem II electron transport. FEBS Letters, 1988. 231(1): p. 259-62. 133. Okey, R.W. and H.D. Stensel, Uncouplers and Activated Sludge - The Impact on Synthesis and Respiration. Toxicological and Environmental Chemistry, 1993. 40: p. 235-254. 134. Okey, R.W. and H.D. Stensel, Uncouplers of Oxidative Phosphorylation. Modeling and Predicting Their Impact on Wastewater Treatment and the Environment, in ACS Symposium Series Emerging Technologies in Hazardous Waste Management V, D.W. Tedder and F.G. Pohland, Editors. 1995, American Chemical Society: Washington. p. 284-309.

135. Omote, H., et al., The -subunit rotation and torque generation in F1-ATPase from wild-type or uncoupled mutant Escherchia coli. Proceeding of the National Academy of Sciences, 1999. 96: p. 7780-7784.

136. Pacheco-Moises, F., et al., Sulfite Inhibits the F1F0-ATP Synthase and Activates the F1F0-ATP of Paracoccus dentrificans. Journal of Bioenergetics and Biomembranes, 2002. 34(4): p. 269-278. 137. Parsonage, D. and S.J. Ferguson, Reassessment of pathways of electron flow to nitrate reductase that are coupled to energy conservation in Paracoccus denitrificans. FEBS Lett., 1983. 153(1): p. 108-12. 138. Parsonage, D., A.J. Greenfield, and S.J. Ferguson, The high affinity of Paracoccus denitrificans cells for nitrate as an electron acceptor. Analysis of possible mechanisms of nitrate and nitrite movement across the plasma membrane and the basis for inhibition by added nitrite of oxidase activity in permeabilized cells. Biochim. Biophys. Acta, 1985. 807(1): p. 81-95. 139. Payne, W.J., Reduction of Nitrogenous Oxides by Microorganisms. Bacteriological Reviews, 1973. 37(4): p. 409-452. 140. Payne, W.J., Denitrification. TIBS, 1976: p. 220-222. 141. Pedersen, P.L., Transport ATPases: Structure, Motors, Mechanism and Medicine: A Brief Overview. Journal of Bioenergetics and Biomembranes, 2005. 37(6): p. 445-449. 170

142. Perez, J.A. and S.J. Ferguson, Kinetics of Oxidative Phosphorylation in Paracoccus denitrificans. 1. Mechanism of ATP Synthesis at the Active Sites(s) of F0F1-ATPase. Biochemistry, 1990. 29: p. 10503-10518. 143. Perlin, D.S., et al., H+/ATP Stoichiometry of Proton Pumps from Neurospora crassa and Escherichia coli. Archives of Biochemistry and Biophysics, 1986. 248(1): p. 53-61. 144. Phillips-Jones, M.K. and M.E. Rhodes-Roberts, Studies of Inhibitors of Respiratory Electron Transport and Oxidative Phosphorylation. Bacteriology, 1991. 27: p. 203-224. 145. Phillips, M.K. and D.B. Kell, A benzoxazole inhibitor of NADH dehydrogenase in Paracoccus denitrificans. FEMS Microbiology Letters, 1981. 11: p. 111-113. 146. Phillips, M.K. and D.B. Kell, A Novel Inhibitor of NADH Dehydrogenas in Paracoccus denitrificans. FEBS Letters, 1982. 104(2): p. 248-250. 147. Phillips, M.K., D.B. Kell, and M.E. Rhodes-Roberts, The Antibacterial Action of Tinopal. Journal of General Microbiology, 1984. 130: p. 1999-2005. 148. Powell, A.K., Growth-Inhibiting Action of Some Pure Substances. Nature, 1944. 153: p. 345. 149. Prestera, T., H.J. Prochaska, and Talalay, Inhibition of NAD(P)H:(Quinone- Acceptor) Oxidoreductase by Cibacron Blue and Related Anthraquinone Dyes: A Structure-Activity Study. Biochemistry, 1992. 31: p. 284-293. 150. Reich, R.A., "Typical" Industrial Wastewater Treatment Facility. 2006, Personal Communication. 151. Renner, R., Scientist debate fertilizing soils with sewage sludge. Environmental Science & Technology, 2000. 34(11): p. 242A-243A. 152. Rich, L.G. and J. Oscar W. Yates, The Effect of 2,4-Dinitrophenol on Activated Sludge. Applied Microbiology, 1955. 3: p. 95-98. 153. Richardson, M.L. and S. Gangolli, eds. The Dictionary of Substances & Their Effects. 1994, The Royal Society of Chemistry. 154. Rocher, M., et al., Towards a Reduction in Excess Sludge Production in Activated Sludge Processes: Biomass Physicochemical Treatment and Biodegradation. Applied Microbiology and Biotechnology, 1999. 51: p. 883-890. 155. Rowe, J.J., et al., NarK is a nitrite-extrusion system involved in anaerobic nitrate respiration by Escherichia coli. Molecular Microbiology, 1994. 12(4): p. 579- 586. 156. Sabbert, D., S. Engelbrecht, and W. Junge, Functional and idling rotatory motion with F1-ATPase. Proceeding of the National Academy of Sciences, 1997. 94: p. 4401-4405. 171

157. Sapshead, L.M. and J.W.T. Wimpenny, The Influence of Oxygen and Nitrate on the Formation of the Cytochrome Pigments of the Aerobic and Anaerobic Respiratory Chain of Micrococcus denitrificans. Biochimica et Bioophysica Acta, 1972. 267: p. 388-397. 158. Schagger, H., Respiratory Chain Supercomplexes. IUBMB Life, 2001. 52: p. 119- 128. 159. Schagger, H., Respiratory chain supercomplexes of mitochondria and bacteria. Biochimica et Biophysica Acta, 2002. 1555: p. 154-159. 160. Scholes, P.B. and L. Smith, Composition and Properties of the Membrane-Bound Respiratory Chain System of Micrococcus dentrificans. Biochimica et Biophysica ACTA, 1968. 153: p. 363-375. 161. Schultz, B.E. and S.I. Chan, Structures and Proton-Pumping Strategies of Mitochondrial Respiratory Enzymes. Annual Review of Biophysics and Biomolecular Structure, 2001. 30: p. 23-65. 162. Seddon, B. and J. McVittie, The Effect of Inhibitors on the Electron-transport chain of Bacillus brevis. Evidence for Branching of the NADH Oxidase Respiratory Chain. Journal of General Microbiology, 1974. 84: p. 386-390. 163. Seif, J.M., Report on the Investigation into the Application of Biosolids at the Al Hamilton Top Mine Site and the Death of Tony Behun. 2000, Pennsylvania Department of Environmental Protection http://www.dep.state.pa.us/dep/biosolids/Testimony/BiosolidsReport0500.htm. 164. Shah, P.M., et al., Enhancement of Specific Waste Water Treatment by the Uncoupler 2,4-dinitrophenol. Food Science, 1975. 40(2): p. 302-305. 165. Sholtz, K., I. Gorskaya, and A. Kotelnikova, The Stoichiometry of Proton Translocation through H+-ATPase of Rat-Liver Mirochondria. European Journal of Biochemistry, 1983. 136(1): p. 129-134. 166. Singer, T.P., Mitochondrial Electron-Transport Inhibitors. Methods in Enzymology, 1979. LV: p. 454-463. 167. Soboll, S., Regulation of Energy Metabolism in Liver. Bioenergetics and Biomembranes, 1995. 27(6): p. 571-582. 168. Standards, E.O.o.A.Q.P., 2,4-Dinitrophenol. 1998, United States Environmental Protection Agency Office of Air Quality Planning & Standards. 169. Steinrucke, P., et al., The Cyctochrome c Reductse/Oxidase Respiratory Pathway of Paracoccus denitrificans: Genetic and Functional Studies. Journal of Bioenergetics and Biomembranes, 1991. 23(2): p. 227-239. 172

170. Stockdale, M. and J. Sewyn, Effects of ring substitutions on the activity of phenols as inhibitors and uncouplers of mitochondrial respiration. European Journal of Biochemistry, 1971. 21: p. 565-574. 171. Stouthamer, A.H., Metabolic Regulation Including Anaerobic Metabolism in Paracoccus denitrificans. Journal of Bioenergetics and Biomembranes, 1991. 23(2): p. 163-185. 172. Strand, S.E. and H.D. Stensel, Sludge Yield Reduction Using Chemical Uncouplers. 1994, Electric Power Research: Seattle. 173. Strand, S.E., G.N. Harem, and H.D. Stensel, Activated-Sludge Yield Reduction Using Chemical Uncouplers. Water Environment Research, 1999. 71(4): p. 454- 458. 174. Swanbeck, G., Interaction Between Deoxyribonucleic Acid and Some Anthracene and Anthraquinone Derivatives. Biochim. Biophys. Acta, 1966. 123: p. 630-633. 175. Terada, H., The Interaction of Highly Active Uncouplers with Mitochondria. Biochimica et Biophysica Acta, 1981. 639: p. 225-242. 176. Terada, H., Uncouplers of Oxidative Phosphorylation. Environmental Health Perspectives, 1990. 87: p. 213-218. 177. TheFreeDictionary, Front End Load. 2007, The Free Dictionary. com http://financial-dictionary.thefreedictionary.com/Front-End+Load. 178. Thompson, J.W. and K. Sorvig, Sustainable Landscape Construction. 2000, Washington, D.C.: Island Press. 372. 179. Trumpower, B.L., The Three-Subunit Cytochrome bc1 Complex of Paracoccus denitrificans. Its Physiological Function, Structure, and Mechanism of Electron Transfer and Energy Transduction. Journal of Bioenergetics and Biomembranes, 1991. 23(2): p. 241-255. 180. Urata, K. and T. Satoh, Mechanism of Nitrite Reduction to Nitrous oxide in a Photodenitrifier, Rhodopseudomonas sphaeroides forma sp. denitrificans. Biochimica et Biophysica Acta, 1985. 841: p. 201-207. 181. Van Spanning, R.J.M., et al., Regulation of oxidative phosphorylation: the flexible respiratory network of Paracoccus denitrificans. J. Bioenerg. Biomembr., 1995. 27(5): p. 499-512. 182. Van Verseveld, H.W., E.M. Meijer, and A.H. Stouthamer, Energy conservation during nitrate respiration in Paracoccus denitrificans. Arch. Microbiol., 1977. 112(1): p. 17-23. 183. Verhaeren, E.H.C., The Uncoupling Activity of Substituted Anthracene Derivative on Isolated Mitochondria from Phaseolus aureus. Phytochemistry, 1980. 19: p. 501-503. 173

184. Vignais, P.M., et al., The electron transport system and hydrogenase of Paracoccus denitrificans. Curr. Top. Bioenerg, 1981. 12: p. 115-96. 185. Vik, S.B. and R.R. Ishmukhametov, Structure and Function of Subunit a of the ATP Synthase of Escherichia coli. Journal of Bioenergetics and Biomembranes, 2005. 37(6): p. 445-449. 186. Walter, B., et al., Inhibition of Denitrification by Uncouplers of Oxidative Phosphorylation. Biochemistry, 1978. 17(15): p. 3039-3045.

187. Weber, J. and A.E. Senior, Catalytic mechanism of F1-ATPase. Biochimica et Biophysica Acta, 1997. 1319: p. 19-58. 188. WEF, Biosolids Fact Sheets. 1998, Water Environment Federation. 189. Weimer, P.J., et al., Anthraquinones as inhibitors of sulfide production by sulfate- reducing bacteria in sewage, oil wells, process tanks, or biomass fermentation, in 22 pp. Cont.-in-part of U.S. Ser. No. 652,406, abandoned. 1993, du Pont de Nemours, E. I., and Co., USA: U.S. 190. Wheatley, A.D., Effluents, in Biodeterioration 7, D.R. Houghton, R.N. Smith, and H.O.W. Eggins, Editors. 1988, Cranfield Institute of Technology: Bedford. p. 193-206. 191. White, D., The Physiology and Biochemistry of Prokaryotes. 1995, New York: Oxford University Press. 378. 192. Wiesmann, U., Biological Nitrogen Removal from Wastewater. Advances in Biochemical Engineering Biotechnology, 1994. 51: p. 113-154. 193. Wikipedia, Wikpedia Free Encyclopedia. 2007, http://en.wikipedia.org/wiki/Zwitterion. 194. Wikipedia, Net present value. 2007, Wikipedia http://.en.wikipedia.org/wiki/Net_present_value. 195. Wood, N.J., et al., Two domains of a dual-function NarK protein are required for nitrate uptake, the step of denitrification in Paracoccus pantotrophus. Molecular Microbiology, 2002. 44(1): p. 157-170. 196. Yagi, T., Bacterial NADH-Quinone Oxidoreductases. Journal of Bioenergetics and Biomembranes, 1991. 23(2): p. 211-225. 197. Yagi, T., et al., NADH Dehydrogenases: From Basic Science to Biomedicine. Journal of Bioenergetics and Biomembranes, 2001. 33(3): p. 233-242. 198. Yagi, T. and A. Yagi-Matsuno, The Proton-Translocating NADH-Quinone Oxidoreductase in the Respiratory Chain: The Secret Unlocked. Biochemistry, 2003. 42: p. 2266-2274. 174

199. Yang, X.-F., M.-L. Xie, and Y. Liu, Metabolic Uncouplers Reduce Excess Sludge Production in an Activated Sludge Process. Process Biochemistry, 2003. 38: p. 1373-1377. 200. Yasuda, R., et al., Rotation of the Subunit in F1-ATPase; Evidence That ATP Synthase Is a Rotary Motor Enzyme*. Bioenergetics and Biomembranes, 1997. 29(3): p. 207-209. 201. Ye, F.X., D.S. Shen, and Y. Li, Reduction in Excess Sludge Production by Addition of Chemical Uncouplers in Activated Sludge Batch Cultures. Journal of Applied Microbiology, 2003. 95: p. 781-786. 202. Ye, F.X. and Y. Li, Reduction of Excess Sludge Production by 3,3',4',5- Tetrachlorosalicylanilide in an Activated Sludge Process. Applied Microbiology and Biotechnology, 2005. 67: p. 269-274. 203. Ye, F.X. and Y. Li, Uncoupled Metabolism Stimulated by Chemical Uncoupler and Oxic-Settling-Anaerobic Combined Process to Reduce Excess Sludge Production. Applied Biochemistry and Biotechnology, 2005. 127(3): p. 187-199. 204. Yoshimi, N., et al., The mRNA overexpression of inflammatory enzymes, phospholipase A2 and cyclooxygenase, in the large bowel mucosa and neoplasms of F344 rats treated with naturally occurring carcinogen, 1- hydroxyanthraquinone. Cancer Lett., 1995. 97(1): p. 75-82.

205. Zharova, T.V. and A. Vinogradov, Energy-dependent Transformation of F0F1- ATPase in Paracoccus denitrificans Plasma Membranes. The Journal of Biological Chemistry, 2004. 279(13): p. 12319-12324.

206. Zhou, Y., T.M. Duncan, and R.L. Cross, Subunit rotation in Escherchia coli F0F1- ATP synthase during oxidative phosphorylation. Proceeding of the National Academy of Sciences, 1997. 94: p. 10583.