ENERGY ANALYSIS AND FABRICATION OF PHOTOVOLTAIC THERMAL ELECTROLYZER AND ION TRANSPORT THROUGH MODIFIED NANOPOROUS MEMBRANES

BY

MUHAMMED ENES ORUC

DISSERTATION

Submitted in partial fulfillment of the requirements for the degree of Doctor of Philosophy in Chemical Engineering in the Graduate College of the University of Illinois at Urbana-Champaign, 2014

Urbana, Illinois

Doctoral Committee:

Professor Hong Yang, Chair Professor Ralph G. Nuzzo, Director of Research Professor Paul J.A. Kenis Assistant Professor David W. Flaherty

Abstract

Hydrogen is an environmentally sustainable energy carrier that can be stored. It is not found naturally and therefore must be artificially produced. We can obtain from renewable energy, such solar and wind energy, which is environmentally clean. One such a promising options is via electrolysis using electricity from a photovoltaic generator. In the first part of the dissertation we studied a microfluidic energy conversion device to produce hydrogen. Particularly, we proposed a new integrated system – a so-called “photovoltaic thermal water electrolyzer (PVTE)” – which consists of

PV cells positioned on top of a planar micro-water electrolyzers in order to harness waste heat as a storable form of energy. The concept of PVTE has the outputs such as electricity and thermal storage, and also it provides hydrogen production efficiently. First, we provided a comprehensive analysis of the overall efficiency of the PVTE system. COMSOL Multiphysics software was used to predict the temperatures for the electrolyte and the PV cells operating at various temperatures and solar fluxes.

Moreover, hourly and monthly efficiency analyses were accomplished for Phoenix, AZ in the year

2010. This new integrated approach is advantageous over conventional PV modules (Chapter 2).

Second, we fabricated a micro-water electrolyzer which utilizes heat from PV cell and works as a heat sink in order to eliminate additional energy input for electrolysis in order to operate at elevated temperatures. We also presented electrode preparation and fabrication of the electrolyzer. The increase in the hydrogen production rate affirms the predictions of our system that utilizes waste heat from PV

(Chapter 3). Finally, we successfully fabricated a new water electrolyzer including hydrophobic porous membrane. This new design allows us to manage gas production and collection in the chamber.

By using this method, we are able to collect gases on the top of the electrolyzer at low flow rates at elevated temperatures (Chapter 4).

ii

Nanoporous membranes have received great attention in the fields of water desalination, biosensing, and chemical separations. Bare nanopores can be used as size-selective filters but if the surface chemistry of a nanopore is modified by coating it with another substance, however, enhanced separations based other properties can be achieved. Many studies have been performed on ion permselectivity across gold-coated charged surfaces and charged nanopores. In the second part of the dissertation, a focus of interfacial transport phenomena is proposed in order to achieve improved- charge selective nanofluidic systems. There have been numerous studies on the quality of organic

SAMs as a blocking mechanism for prevention of ion adsorption. First, we investigated the electrochemical interfacial properties of a well-ordered SAM of 1-undecanethiol (UDT) on evaporated gold surface by EIS in electrolytes without a redox couple. Using a constant phase element (CPE) series resistance model, prolonged incubation times (up to 120 h) show decreasing monolayer capacitance approaching the theoretical value for 1-undecanethiol (Chapter 6). Secondly, we fabricated a membrane permeate flow cell is described with the aim of studying the transport of methyl viologen (paraquat, MV2+) and napathalenedisulfonate disodium salt (NDS2-), using a conductive

NCAM. A polycarbonate track etched (PCTE) membrane was made conductive by sputter coating gold on the membrane surface. Transport studies were done in a voltage range in which faradaic current was minimized at the surface of the gold-coated NCAMs. The goal of the transport studies is to demonstrate improved charge selectivity when a well-grown 1-undecanethiol monolayer is assembled at the surface of the NCAM for a wide range of applied potentials (-400 mV < Vappl < 400 mV). Results show the selectivity of charged analytes through the metallized NCAM can be improved by functionalizing the surface with a self-assembled monolayer (SAM). The selectivity coefficients for

MV2+ and NDS2- increased with functionalization of undecanethiol on the gold-coated NCAM surface

(Chapter 7).

iii

Acknowledgements

I would like to express my gratitude to several people without whom this work would not have been possible. First and foremost, I am profoundly grateful to my advisors, Professor Mark A. Shannon and Professor Ralph Nuzzo, for their patience, guidance, and support. They have been true mentors to me, and have taught me how to do fundamental research in the applied sciences. I admire their creative and inspired way of thinking, their brilliant scientific visions, their excitement in research, their energy in teaching, their effort to direct their work towards environmentally sensitive subjects, and their endless enthusiasm that carried me through challenging times during my graduate career. I am grateful for their guidance and mentorship for pursuing an academic career myself, and the way they trained me through my doctorate for that purpose. Also, I am thankful to my thesis committee, Professor Paul

Kenis, Professor Hong Yang, and Professor David Flaherty, for their support.

I am grateful to Shannon and Nuzzo group members, particularly Damena Agonafer, Edward

Chainani, Evan Erickson and Chris Corcoran, for their scientific insight, editorial assistance, and much needed friendship during the completion of this work.

I would like to thank National Science Foundation WaterCAMPWS, Ministry of National

Education of Turkey Graduate Scholarship, and the Dow Chemical Company Graduate fellowship programs for their support during my PhD Program.

Finally, a special feeling of gratitude to my family. I wish to thank you my parents, Ahmet and Eda whose words of encouragement and push support for all my life, I am thankful to my wonderful wife,

Bilgehan, for giving me her love, encouragement and support during my all my Ph.D. She is an inspiration and source of motivation that leads me to success. Also with the existence of my gorgeous daughters, Zisan and Nurefsan, I was able to forget all challenges in academia and realize real the meaning of the world.

iv

To my lovely wife and my wonderful daughters,

v

Table of Contents

Chapter 1 Photovoltaic Thermal Water Electrolyzer for Hydrogen Production ...... 1 1.1 Global Energy Challenges ...... 1 1.2 Hydrogen Production Technologies from Water Electrolysis ...... 2 1.3 Alkaline Water Electrolyzer ...... 3 1.4 Microfluidic Energy Conversion Devices...... 4 1.5 Photovoltaic Thermal Water Electrolyzer ...... 5 1.6 Dissertation Overview ...... 7 1.7 References ...... 8 1.8 Figures ...... 11 Chapter 2 Comprehensive Energy Analysis of Photovoltaic Thermal Water Electrolyzer ...... 13 2.1 Introduction ...... 13 2.2 Modeling a Photovoltaic Water Electrolyzer Module ...... 15 2.2.1 PVTE Model ...... 15 2.2.2 Theory, Assumptions and Governing Equations ...... 15 2.3 Evaluation of Overall Performance ...... 17 2.3.1 Efficiency Output of PV Module ...... 17 2.3.2 Energetic and Exergetic Performance of the Module ...... 18 2.3.3 Analysis of the Hydrogen Production Efficiency ...... 19 2.4 Results and Discussion ...... 20 2.4.1 COMSOL Simulation and Temperature Distribution ...... 20 2.4.2 Optimization of Chamber Thickness and Fluid Velocity in the Chamber ...... 21 2.4.3 Hourly Variations of the System Efficiency ...... 22 2.4.4 Monthly Variations of the System Efficiency ...... 23 2.5 Conclusion ...... 25 2.6 References ...... 25 2.7 Figures and Tables ...... 28 Chapter 3 Fabrication of Photovoltaic Thermal Water Electrolyzer...... 36 3.1 Introduction ...... 36 3.2 Design of Water Electrolyzer ...... 38 3.3 Experimental ...... 39 3.3.1 Electrode Preparation ...... 39 3.3.2 Fabrication of the Water Electrolyzer ...... 39

vi

3.3.3 Electrochemical Testing ...... 40 3.4 Results and Discussion ...... 40 3.4.1 Electrode characterization...... 40 3.4.2 Effect of Flow Rate on Hydrogen Production Rate ...... 41 3.4.3 Effect of Chamber Thickness on Hydrogen Production Rate ...... 42 3.4.4 Effect of Residence time on Hydrogen Production Rate ...... 43 3.4.5 Effect of Residence time on Total Hydrogen Production ...... 43 3.5 Conclusion ...... 44 3.6 References ...... 45 3.7 Figures and Tables ...... 47 Chapter 4 Gas Management in Photovoltaic Thermal Water Electrolyzer ...... 60 4.1 Introduction ...... 60 4.2 Gas Management in Microchannels ...... 62 4.3 Experimental ...... 62 4.3.1 Fabrication of the Water Electrolyzer ...... 62 4.3.2 Electrochemical Testing ...... 63 4.4 Results and Discussion ...... 64 4.5 Conclusion ...... 65 4.6 References ...... 66 4.7 Figures and Tables ...... 68 Chapter 5 Species Separation through Self-Assembled Monolayers Modified Nanoporous Membranes .. 75 5.1 Ionic and Molecular Transport through Nanoporous Membrane ...... 75 5.2 Specific Ion Adsorption and Properties of well-ordered Self-Assembled Monolayers ...... 76 5.3 Electrochemical Impedance Spectroscopy...... 78 5.4 Dissertation Overview ...... 79 5.5 References ...... 80 5.6 Figures ...... 85 Chapter 6 Insulating Properties of Alkanethiol SAMs Formed Under Prolonged Incubation Using Electrochemical Impedance Spectroscopy ...... 87 6.1 Introduction ...... 87 6.2 Experimental ...... 91 6.2.1 Electrochemical Instrumentation and Measurements ...... 91 6.2.2 Preparation of Monolayer Surfaces ...... 91 6.2.3 Surface Characterization ...... 92 6.3 Results and Discussion ...... 92 6.3.1 Determination of Electrochemical Surface Area ...... 92 6.3.2 AFM measurements ...... 93

vii

6.3.3 EIS and CV measurements ...... 93 6.4 Conclusion ...... 101 6.5 References ...... 103 6.6 Figures and Tables ...... 107 Chapter 7 Study of Ionic Transport through Metalized Nanoporous Membranes Functionalized with Self- assembled Monolayers ...... 113 7.1 Introduction ...... 113 7.2 Experimental ...... 116 7.2.1 Preparation of Gold-Coated NCAMs ...... 116 7.2.2 Preparation of Monolayer Surfaces ...... 117 7.2.3 Preparation of Kapton Electrodes ...... 117 7.2.4 Transport Experiments ...... 117 7.3 Results and Discussion ...... 119 7.3.1 Effects of Transport of NDS2- and MV2+ when NCAM is Functionalized with a SAM ..... 119 7.3.2 Effects of Ion Selectivity when Monolayer Formation is Present ...... 122 7.3.3 Effects of a Charged Nanopore and EDL Thickness on the Transport of Ions ...... 123 7.4 Conclusion ...... 124 7.5 References ...... 125 7.6 Figures and Tables ...... 130

viii

CHAPTER 1

PHOTOVOLTAIC THERMAL WATER ELECTROLYZER FOR HYDROGEN PRODUCTION

1.1 Global Energy Challenges

Global warming, fluctuating oil prices, and the depletion of fossil fuel resources have activated a tremendous research in the development of alternative and sustainable energy sources such as wind, solar and hydropower. Besides, national and international energy and environmental security concerns are motivated a transformation from fossil fuels to secure and clean fuels[1]. Alternative and sustainable energy sources are promising; however, the main challenges of these sources are capital cost and intermittent nature in power production. The power production of these energy sources are fluctuating based on changing of solar radiation, wind with seasons, months, days and hours [2]. In this respect, when high amount of power is produced within these time intervals, a storage medium is crucial to overcome this problem a convenient solution.

Hydrogen is an environmentally friendly and sustainable energy carrier for our economy. We can produce electricity from hydrogen with only the formation of water by using fuel cells or burning hydrogen in conventional combustion engines [3]. Steam reforming of natural gas is the dominant process to generate hydrogen with CO2 as a by-product which is linked to global warming [4]. A sustainable way to produce hydrogen is electrolysis of water from renewable energy sources such as wind and solar. Although this tactic is expensive relative to steam reforming, it is very promising for the future while considering the limitation and adverse environmental impact of fossil fuels [1].

1

1.2 Hydrogen Production Technologies from Water Electrolysis

The most prominent water electrolysis technologies for hydrogen production are: Alkaline electrolysis, proton exchange membrane (PEM) electrolysis and solid oxide (SO) electrolysis.

Alkaline water electrolysis: This type of electrolyzer operates by having two metal electrodes at applied potential in a liquid alkaline electrolyte consisting of a solution with between 20%-40% potassium hydroxide. These electrodes are separated by diaphragm, separating the product gases and transporting the hydroxide ions from one electrode to the other. These electrolyzer is well established and commonly found in the market. Alkaline electrolyzer will be discusses in the next section in details.

Polymer electrolyte membrane electrolysis: Another approach is to use a proton exchange membrane as electrolyte. This perfluorosulfonic acid polymer (also known as Nafion) has been used in electrolysis. PEM electrolysis is a process just reverse of a PEM process. Water is split into , protons and electrons on the anode electrode. Protons pass through the polymer electrolyte membrane and on the cathode combine with electrons to form hydrogen. Passage of protons through the membrane is escorted by water transport The PEM electrolyzer was introduced to overcome the issues of partial load, low current density, and low pressure operation currently plaguing the alkaline electrolyzer. PEM electrolyzers have reached the commercial market, but only for small scale niche applications. PEM electrolyzers are almost exclusively operated at temperatures below 100 °C, at ambient pressure [5].

Solid oxide electrolysis: SO electrolyzers are operated at very high temperatures around 800 °C seems to be most common which results in higher efficiencies compared to alkaline or PEM electrolyzers. These electrolyzers require extremely high heat supply and they are commonly integrated with other energy plants in order to utilize waste heat energy. It operates in the reverse mode of solid oxide fuel cells (SOFCs). Both steam and recycled hydrogen are fed to the cathode, where water is reduced to produce hydrogen. The oxide anions generated in the cathode pass through the solid

2 electrolyte to the anode, where they recombine forming oxygen and closing the circuit with the released electrons. In SOs, the most common electrolyte is yttria (Y2O3) stabilized zirconia (ZrO2) (YSZ) [5].

1.3 Alkaline Water Electrolyzer

In the electrolysis of water, a DC electric current passes across two electrodes separated by electrolyte and the two electrodes are connected through an external circuit. Water decomposes into hydrogen at cathode and oxygen at anode. Most of the commercially available electrolyzers are alkaline and concentrated KOH solutions are used as electrolyte [6]. The half reactions take place on the cathode and the anode are written as below, respectively:

+ − 푐푎푡ℎ표푑푒 ∶ 2퐻 + 2푒 → 퐻2 (1.1)

1 푎푛표푑푒 ∶ 2푂퐻− → 푂 + 퐻 푂 + 2푒− (1.2) 2 2 2

The overall water splitting reaction via electrolysis is written as in Equation 1.3.

1 퐻 푂 → 퐻 + 푂 (1.3) 2 2 2 2

At standard conditions (25ºC and 1 bar) the splitting of water is a non-spontaneous reaction and the minimum amount of work that is required to split water can be obtained from the change in Gibbs free energy (ΔG°=237kJ/mol), or in terms of voltage:

훥퐺° 237000퐽 푈표 = = (1.4) 푟푒푣 푧 ∗ 퐹 2 ∗ 96485퐶

where z is the number of electrons transferred per hydrogen molecule and F is the Faraday’s constant.

Water decomposition is an endothermic reaction where the total amount of energy needed in water electrolysis is equivalent to the change in enthalpy (훥퐻°=286kJ/mol). The rest of the energy to achieve

3

ΔH° can be supplied via heat (푇훥푆° = 훥퐻° − 훥퐺°). Therefore the total energy demand can be expressed by the thermo-neutral cell voltage as below:

훥퐻° 286000퐽 푈표 = = (1.5) 푡ℎ 푧 ∗ 퐹 2 ∗ 96485퐶

표 표 Total energy demand for both 푈푟푒푣 and 푈푡ℎ change slightly with temperature. At standard

표 표 표 conditions (25ºC and 1 bar) 푈푟푒푣 = 1.229푉 and 푈푡ℎ = 1.482푉 whereas at 80°C 푈푟푒푣 = 1.184푉 and

표 푈푡ℎ = 1.473푉.

On the other hand, anodic and cathodic are affected by temperature. As the operating temperature increases at a given current density, the overvoltages decrease due to faster electrode kinetics [7, 8]. Advanced commercial water electrolyzers do not require heating systems since a significant portion of the energy dissipated is used as heat input in order to operate the electrolyzer at adequate temperatures to reduce the cell overvoltages. The heat produced is described as the difference between the actual cell potential and thermoneutral voltage as seen in Equation 1.6 [9].

푄̇𝑔푒푛 = 퐼 ∗ (푈푐푒푙푙 − 푈푡ℎ) (1.6)

The electrolyzer in this work aims to reduce actual cell potential by utilizing the excess heat in PV cells. Therefore, the additional energy input to operate the electrolyzer at elevated temperatures will be eliminated.

1.4 Microfluidic Energy Conversion Devices

Microfluidic energy conversion devices have been widely studied in the literature: common examples are micro fuels cells [10-12], microfluidic biofuel cell [13], and micro flow batteries [14,

15]. Microfluidics are advantageous in comparison to conventional fluidic systems due the high surface area to volume ratio of small components facilitating rapid heat and mass transfer leading to rapid temperature changes and reaction kinetics.

4

One of the major practical challenges in microfluidics is the management of gas bubbles in the microchannels. Although bubbles can provide improved mixing for chemical reactions [16], and enhanced heat and mass transfer [17], they also accumulate and clog the channels [18] or reduce the electrode performance in the microfluidic device. Therefore, many efforts have been devoted to get rid of the gas bubbles in the microfluidics such as locally varying wetting properties of channels and capillarity restricted modifications in the channels [19, 20], or dynamic bubble traps [21].

One approach is diffusion based gas bubble removal by a gas-permeable membrane. A thin PDMS layer [22, 23] or a hydrophobic membrane [17] are used for an active bubble trap and removal. Meng et al. [24] used a hydrophobic membrane providing a gas-venting microchannel that directly removes

CO2 gas bubbles from the electrochemical reactions of fuel cells without leakage. Another application of hydrophobic membranes in microfluidics is vapor venting two-phase microchannel heat exchanger

[17, 25]. When they used a hydrophobic membrane that vents the vapor phase into separate vapor transport channels, they were able to reduce the pressure head with power consumption for high heat flux generating electronics. A similar approach was taken by Winther-Jensen et al. for water electrolyzer [26]. They proposed a breathable electrode design which allows the diffusion of O2 and

H2 gases through a hydrophobic membrane leading to an improvement of the efficiency of the water splitting reaction.

1.5 Photovoltaic Thermal Water Electrolyzer

Solar energy is chief among renewables that has potential for great growth in a sustainable economy [27]. The sun’s energy is often harnessed by using a photovoltaic (PV) module that converts solar energy into electricity or a solar thermal collector which transforms solar energy into heat, e.g. for domestic hot water or room heating [28, 29] (Figure 1.1). A so-called photovoltaic/thermal (PVT) collector combines these two modules together [30]. PV cells generate electricity and also act as an

5 absorber for the thermal collector, which uses the remaining radiation energy for generating heat.

Therefore, a PVT collector brings several advantages as compared to separately installed modules; firstly, it possesses a higher electrical and thermal yield per unit surface area; secondly, need for space on a roof should be reduced; and, thirdly, installations costs needs to be minimized [31].

Heat energy has been conventionally used for domestic hot water and/or room heating; it is, however, highly desirable to convert the released energy into alternative storable forms. Hydrogen is an environmentally benign and a sustainable energy carrier that can be stored. It is not found naturally and therefore must be artificially produced. One such method is to utilize the direct current created by a photovoltaic generator to generate hydrogen. Many studies have been done on coupling together a

PV system and an alkaline water electrolyzer [32-35]. In these studies the systems were separately installed and analyses were done on optimizing the voltage and maximum power output of the PV to the operating voltage of electrolyzer. Currently, PV-assisted water electrolyzers available on the market have PV conversion efficiencies between 15% and 17% [36].

The reason why conversion efficiency of silicon based PV cells is not high is that a portion of solar spectrum is converted into electricity and the energy that is not transformed is dissipated as heat (Figure

1.2). To this end, in order to harness waste heat as a storable form of energy, we have proposed a new integrated system – a so-called “photovoltaic thermal water electrolyzer (PVTE)” – which consists of

PV cells positioned on top of a planar micro-water electrolyzer. Figure 1.3a shows a simple overview of the PVTE system, and Figure 1.3b depicts the working mechanism and design of the electrolyzer.

As can be seen in Figure 1.3b, our electrolyzer design is based microfluidic energy devices which have been investigated extensively in the literature [37, 38]. PV cells generate electricity that is used in electrolysis. The excess heat dissipated from the cells is transferred through the water electrolyzer. The electrolyte in the electrolyzer functions as a heat-transport fluid to remove heat from the water electrolyzer. In this way, the operating temperature of PV panel decreases to yield an increase in the

6 electric efficiency as 0.05/°C [39]. The temperature of the electrolyte in the electrolyzer approaches the temperature of the PV cell. The raised temperature of the electrolyte in the electrolyzer aids to reduce the overpotentials for H2 and O2 gas evolution for a given current flux. Therefore, the energy consumption of the electrolyzer will be reduced with increasing efficiency of electrolysis. Electrolyte at an elevated temperature leaving the electrolyzer is circulated through tubing to transfer heat to an insulated water tank.

1.6 Dissertation Overview

The first part of dissertation will discuss the topics around the photovoltaic water electrolyzer for hydrogen production.

In Chapter 2, we provided a comprehensive analysis of the overall efficiency of the PVTE system.

COMSOL Multiphysics software was used to predict the temperatures for the electrolyte and the PV cells operating at various temperatures and solar fluxes. The electrical, thermal, exergetic, electrolysis, and overall system efficiencies were calculated to optimize the system by varying the chamber thickness and velocities of electrolyte in the electrolyzer. Moreover, hourly and monthly efficiency analyses were accomplished for Phoenix, AZ in the year 2010. This new integrated approach is advantageous over conventional PV modules. Water electrolyzer systems utilize heat into efficient hydrogen production, less space occupation with architectural uniformity, and reduces installation costs.

In Chapter 3, we fabricated a micro-water electrolyzer which utilizes heat from PV cell and works as a heat sink in order to eliminate additional energy input for electrolysis in order to operate at elevated temperatures. Our approach is first to pump and reside the electrolyte in the chamber at specific time intervals in order to transfer heat to the electrolyte. Finally all the electrolyte and evaluated gases are evacuated from the chamber through two outlets by a syringe pump. Herein, we used a three

7 dimensional finite element (COMSOL Multiphysics) analysis of the temperature distribution of electrolyte. We also presented electrode preparation and fabrication of the electrolyzer. We applied a constant temperature on top of the water electrolyzer to see the effect of temperature on hydrogen production in this system. The increase in the hydrogen production rate affirms the predictions of our system that utilizes waste heat from PV.

In Chapter 4, we successfully fabricated a new water electrolyzer including hydrophobic porous membrane. This new design allows us to manage gas production and collection in the chamber. By using this method, we are able to collect gases on the top of the electrolyzer at low flow rates at elevated temperatures.

1.7 References

1. An Integrated Strategic Plan for the Research, Development, and Demonstration of Hydrogen and Fuel Cell Technologies. 2011.

2. Yilanci, A., I. Dincer, and H.K. Ozturk, A review on solar-hydrogen/fuel cell hybrid energy systems for stationary applications. Progress in Energy and Combustion Science, 2009. 35(3): p. 231-244.

3. Nowotny, J. and T.N. Veziroglu, Impact of hydrogen on the environment. International Journal of Hydrogen Energy, 2011. 36(20): p. 13218-13224.

4. Mason, J.E., World energy analysis: H-2 now or later? Energy Policy, 2007. 35(2): p. 1315-1329.

5. Ursua, A., L.M. Gandia, and P. Sanchis, Hydrogen Production From Water Electrolysis: Current Status and Future Trends. Proceedings of the Ieee, 2012. 100(2): p. 410-426.

6. Zeng, K. and D.K. Zhang, Recent progress in alkaline water electrolysis for hydrogen production and applications. Progress in Energy and Combustion Science, 2010. 36(3): p. 307-326.

7. Appleby, A.J., G. Crepy, and J. Jacquelin, HIGH-EFFICIENCY WATER ELECTROLYSIS IN ALKALINE-SOLUTION. International Journal of Hydrogen Energy, 1978. 3(1): p. 21-37.

8. Hug, W., et al., HIGHLY EFFICIENT ADVANCED ALKALINE ELECTROLYZER FOR SOLAR OPERATION. International Journal of Hydrogen Energy, 1992. 17(9): p. 699-705.

9. Dieguez, P.M., et al., Thermal performance of a commercial alkaline water electrolyzer: Experimental study and mathematical modeling. International Journal of Hydrogen Energy, 2008. 33(24): p. 7338-7354.

8

10. Mitrovski, S.M., L.C.C. Elliott, and R.G. Nuzzo, Microfluidic devices for energy conversion: Planar integration and performance of a passive, fully immersed H2-O2 fuel cell. Langmuir, 2004. 20(17): p. 6974-6976.

11. Choban, E.R., et al., Membraneless laminar flow-based micro fuel cells operating in alkaline, acidic, and acidic/alkaline media. Electrochimica Acta, 2005. 50(27): p. 5390-5398.

12. Jayashree, R.S., et al., Air-breathing laminar flow-based microfluidic fuel cell. Journal of the American Chemical Society, 2005. 127(48): p. 16758-16759.

13. Togo, M., et al., Structural studies of enzyme-based microfluidic biofuel cells. Journal of Power Sources, 2008. 178(1): p. 53-58.

14. Ferrigno, R., et al., Membraneless vanadium redox fuel cell using laminar flow. Journal of the American Chemical Society, 2002. 124(44): p. 12930-12931.

15. Kjeang, E., et al., A microfluidic fuel cell with flow-through porous electrodes. Journal of the American Chemical Society, 2008. 130(12): p. 4000-4006.

16. Gunther, A., et al., Transport and reaction in microscale segmented gas-liquid flow. Lab on a Chip, 2004. 4(4): p. 278-286.

17. Xu, J., R. Vaillant, and D. Attinger, Use of a porous membrane for gas bubble removal in microfluidic channels: physical mechanisms and design criteria. Microfluidics and Nanofluidics, 2010. 9(4-5): p. 765-772.

18. Jensen, M.J., G. Goranovic, and H. Bruus, The clogging pressure of bubbles in hydrophilic microchannel contractions. Journal of Micromechanics and Microengineering, 2004. 14(7): p. 876- 883.

19. Hibara, A., et al., Surface modification method of microchannels for gas-liquid two-phase flow in microchips. Analytical Chemistry, 2005. 77(3): p. 943-947.

20. Yang, Z., S. Matsumoto, and R. Maeda, A prototype of ultrasonic micro-degassing device for portable dialysis system. Sensors and Actuators a-Physical, 2002. 95(2-3): p. 274-280.

21. Schonburg, M., et al., Significant reduction of air microbubbles with the dynamic bubble trap during cardiopulmonary bypass. Perfusion-Uk, 2001. 16(1): p. 19-25.

22. Skelley, A.M. and J. Voldman, An active bubble trap and debubbler for microfluidic systems. Lab on a Chip, 2008. 8(10): p. 1733-1737.

23. Sung, J.H. and M.L. Shuler, Prevention of air bubble formation in a microfluidic perfusion cell culture system using a microscale bubble trap. Biomedical Microdevices, 2009. 11(4): p. 731-738.

24. Meng, D.D., et al., A -tolerant gas-venting microchannel for a microdirect methanol fuel cell. Journal of Microelectromechanical Systems, 2007. 16(6): p. 1403-1410.

25. David, M.P., et al., Hydraulic and thermal characteristics of a vapor venting two-phase microchannel heat exchanger. International Journal of Heat and Mass Transfer, 2011. 54(25-26): p. 5504-5516.

9

26. Winther-Jensen, O., et al., Towards hydrogen production using a breathable electrode structure to directly separate gases in the water splitting reaction. International Journal of Hydrogen Energy, 2012. 37(10): p. 8185-8189.

27. Lewis, N.S., Powering the planet. Mrs Bulletin, 2007. 32(10): p. 808-820.

28. Shah, A., et al., Photovoltaic technology: The case for thin-film solar cells. Science, 1999. 285(5428): p. 692-698.

29. Kalogirou, S.A., Solar thermal collectors and applications. Progress in Energy and Combustion Science, 2004. 30(3): p. 231-295.

30. Chow, T.T., A review on photovoltaic/thermal hybrid solar technology. Applied Energy, 2010. 87(2): p. 365-379.

31. van Helden, W.G.J., R.J.C. van Zolingen, and H.A. Zondag, PV thermal systems: PV panels supplying renewable electricity and heat. Progress in Photovoltaics, 2004. 12(6): p. 415-426.

32. Hollmuller, P., et al., Evaluation of a 5 kW(P) photovoltaic hydrogen production and storage installation for a residential home in Switzerland. International Journal of Hydrogen Energy, 2000. 25(2): p. 97-109.

33. Clua, J.G.G., R.J. Mantz, and H. De Battista, Hybrid control of a photovoltaic-hydrogen energy system. International Journal of Hydrogen Energy, 2008. 33(13): p. 3455-3459.

34. Djafour, A., et al., Photovoltaic-assisted alkaline water electrolysis: Basic principles. International Journal of Hydrogen Energy, 2011. 36(6): p. 4117-4124.

35. Gibson, T.L. and N.A. Kelly, Optimization of solar powered hydrogen production using photovoltaic electrolysis devices. International Journal of Hydrogen Energy, 2008. 33(21): p. 5931- 5940.

36. Tyagi, V.V., et al., Progress in solar PV technology: Research and achievement. Renewable & Sustainable Energy Reviews, 2013. 20: p. 443-461.

37. Erickson, E.M., et al., Optimization of a permeation-based microfluidic direct fuel cell (DFAFC). Electrophoresis, 2011. 32(8): p. 947-956.

38. Lopez-Montesinos, P.O., et al., Design, fabrication, and characterization of a planar, silicon- based, monolithically integrated micro laminar flow fuel cell with a bridge-shaped microchannel cross-section. Journal of Power Sources, 2011. 196(10): p. 4638-4645.

39. Skoplaki, E. and J.A. Palyvos, On the temperature dependence of photovoltaic module electrical performance: A review of efficiency/power correlations. Solar Energy, 2009. 83(5): p. 614-624.

10

1.8 Figures

Figure 1.1. The current commercial technologies that harness solar energy: (a) PV module in which Si based materials converts photons into electricity, (b) solar thermal collector which transforms solar energy into heat, e.g. for domestic hot water or room heating (http://www.wikipedia.org/).

Figure 1.2. Solar spectrum of silicon based photovoltaic cells (http://www.vicphysics.org/).

11

(a)

(b)

Figure 1.3. (a) A Simple overview of the PVTE system, (b) Working mechanism and design of the electrolyzer.

12

CHAPTER 2

COMPREHENSIVE ENERGY ANALYSIS OF PHOTOVOLTAIC THERMAL WATER ELECTROLYZER

2.1. Introduction

Hydrogen is an environmentally benign and a sustainable energy carrier that can be stored. It is not found naturally and therefore must be artificially produced. We can obtain hydrogen from renewable energy, such solar and wind energy, which is environmentally clean. One such a promising options is via electrolysis using electricity from a photovoltaic generator. Many studies have been done on coupling together a PV system and an alkaline water electrolyzer [1-4]. In these studies the systems were separately installed and analyses were performed on optimizing the voltage and maximum power output of the PV to the operating voltage of electrolyzer. On the other hand, one drawback of the PV cells is the lack of electrical yield per unit surface area. Currently PV-assisted water electrolyzers available on the market have PV conversion efficiencies between 15% and 17% [5]. Moreover, the energy that is not transformed into electricity is dissipated as heat. It causes an increase in the operating temperature of PV panel and this inclement yields a decrease in the electric efficiency as 0.05/°C [6].

In order to utilize the waste heat of the PV cells, one approach is employing a hybrid photovoltaic/thermal (PVT) collector which combines a photovoltaic (PV) module and a solar thermal collector [7]. The PV module converts solar energy into electricity and also acts as an absorber for the thermal collector which transforms the remaining radiation energy into heat [8, 9]. Heat energy has been conventionally used for domestic hot water and/or room heating; it is, however, highly desirable

13 to convert the released energy into other alternative storable forms. To this end, the aim of this current work is to propose a new integrated system – a so-called “photovoltaic thermal water electrolyzer

(PVTE)” – which consists of PV cells positioned on top of a planar micro-water electrolyzers in order to harness waste heat as a storable form of energy. The concept of PVTE has not only the same outputs of PVT collector as electricity and thermal storage but also provides hydrogen production efficiently.

Figure 2.1.a shows a simple overview of the PVTE system, and Figure 2.1.b depicts the working mechanism and design of the electrolyzer. During the day, the PV cells generate electricity which is used in electrolysis and the rest is sent to the grid. The excess heat dissipated from the cells is captured by the water electrolyzer as a heat sink. The temperature of the electrolyte in the electrolyzer approaches the temperature of the PV cell. The raised temperature of the electrolyte in the electrolyzer aids to reduce the overpotentials for H2 and O2 gas evolution for a given current flux. Therefore, the energy consumption of the electrolyzer will be reduced with increasing efficiency of electrolysis. The electrolyte in the electrolyzer also functions as a heat-transport fluid and the electrolyte at an elevated temperature leaving the electrolyzer is circulated through tubing to transfer heat to an insulated water tank.

Herein, we provided a comprehensive analysis of the overall efficiency of the PVTE system.

COMSOL Multiphysics software was used to predict the temperatures for the electrolyte and the PV cells operating at various temperatures and solar fluxes. The electrical, thermal, exergetic, electrolysis, and overall system efficiencies were calculated to optimize the system by varying the chamber thickness and velocities of electrolyte in the electrolyzer. Moreover, hourly and monthly efficiency analyses were accomplished for Phoenix, AR in the year 2010. The analysis indicated that electrical efficiency gain between the PV modules with and without water electrolyzer is 1%. The thermal and energetic efficiencies of the system vary between 55%-60% and 15%-16%, respectively. The PVTE system made an improvement in the efficiency of electrolysis is around 4%. This new integrated

14 approach is advantageous over conventional coupling of PV modules and water electrolyzer systems as it utilizes heat into efficient hydrogen production, less space occupation with architectural uniformity, and reduces installation costs.

2.2. Modeling a Photovoltaic Water Electrolyzer Module

2.2.1. PVTE Model

In this work, we analyzed a unit PV cell (16 cm x 16 cm) with 16 water electrolyzers underneath it. A model drawing of the module with layers and thicknesses are seen in Figure 2.2.a. The solar cell portion of this module is sandwiched between two layers of ethylene vinyl-acetate (EVA) and a protective glass layer is located on them. Finally, a back layer (Tedlar) is added below the PV cell to protect it from environmental conditions. A cross-section view of an electrolyzer unit cell, in the dimensions of 40 mm length, 40 mm width and about 11 mm height, is shown in Figure 2.2.b. The size of the electrolyzer chamber is 20 mm x 20 mm x 0.4 mm. Two thin-film electrodes that are 1 cm2 in area are located in the chamber and on a non-electrically conductive glass substrate that is 0.5 mm thick. A 7 mm thick PDMS slab was chosen as the electrolyzer chamber material due to its low thermal conductivity. The PDMS reservoir includes an inlet, where electrolyte at ambient temperature enters, and two outlets for H2 and O2 gases with the heated electrolyte to exit the chamber. Properties of each material used in the COMSOL simulation are listed in Table 2.1.

2.2.2. Theory, Assumptions and Governing Equations

The PVTE module was studied using COMSOL Multiphysics software in order to evaluate the temperature of the PV cell, and the temperature of the electrolyte in the electrolyzer and while leaving the module at different ambient temperatures and solar fluxes. Moreover, effects of chamber thickness and electrolyte velocity as a coolant were studied.

15

In the COMSOL simulations, all forms of heat transfer are taken into consideration. A unit cell was taken into consideration and it is assumed that it is infinitely long on each side and insulated.

Absorbed heat within the PV cell is transported via conduction to the electrolyzer. Heat is transferred to the surroundings by convection and thermal radiation.

The absorption factor of the PV cell for both crystalline silicon and thin-film solar cells plays an important role in the thermal efficiency of PVT collector systems. Radiated solar energy on the PV cell is either reflected, absorbed, or transmitted. The absorption factor of the laminate (퐴) is described as the ratio of the incident solar irradiance that is absorbed. All absorbed solar energy that is not converted into electricity is transformed into heat. The fraction of incident solar irradiance that is converted into heat is called the effective absorption factor, which is defines as:

퐴푒푓푓 = 퐴 − 휂푒 (2.1)

where 휂푒 is the electrical efficiency of the PV cell. Typical values for absorption factors and effective absorption factors can vary between 70-90% and 60-80%, respectively. In this model 퐴푒푓푓 is taken as

75% where 휂푒 is 15%.

At steady state, the general heat conduction through the module is given in the Equation below.

훻 ∙ (푘훻푇) = 0 (2.2)

The heat loss by convection from the surfaces of the module to the environment is given in

Equation 2.3. An average wind speed was taken as 1 m/s which means heat transfer coefficient ℎ푐 is

6.5 W/m2/K [10].

푞푐표푛푣 = −ℎ푐 ∙ 퐴 ∙ (푇푠푢푟푓푎푐푒 − 푇푎푚푏) (2.3)

The heat loss by thermal radiation from the PV surface is calculated as given in the Equation below.

4 4 푞푟푎푑 = 휖 ∙ 𝜎 ∙ (푇푠푢푟푓푎푐푒 − 푇푎푚푏) (2.4)

16

Average surface emissivity of silicon ϵ is selected as 0.85 and the Stefan-Boltzmann constant 𝜎 is

5.67x10-8 W/m2/K4.

The conjugate heat transfer module in COMSOL allows analysis of forced convection between the electrolyte as a heat carrier fluid and a high-temperature solar cell as a heat source using the continuity and momentum equations as seen in Equations 2.5 and 2.6. Heat transfer for the electrolyte is given in

Equation 2.7.

훻 ∙ (𝜌푢) = 0 (2.5)

𝜌푢 ∙ 훻푢 = 훻푝 + 훻 ∙ (휇(훻푢 + (훻푢)푇)) (2.6)

푝퐶푝푢 ∙ 훻푇 = 훻 ∙ (푘훻푇) (2.7)

It is assumed that the initial temperature of the electrolyte is equal to the ambient temperature [11].

The flow through the electrolyzer is considered laminar and incompressible.

2.3. Evaluation of Overall Performance

2.3.1. Efficiency Output of PV Module

The operating temperature of the PV cell influences its electrical efficiency. There are many parameters which affect the temperature of the PV system. These parameters include ambient temperature, local wind speed, as well as solar radiation flux, material selection, and design for each

PV system. In the literature, a number of equations and correlation coefficients have been proposed to calculate the electrical efficiency of the PV module depending on its temperature [12, 13]. In this work we use the equation below:

휂푃푉 = 휂푃푉(푟푒푓)[1 − 훽푟푒푓(푇 − 푇푟푒푓)] (2.8)

17 where 휂푃푉(푟푒푓) and 훽푟푒푓 are 0.15 and 0.0041, respectively [12].

2.3.2. Energetic and Exergetic Performance of the Module

The major evaluation approach for overall performance of PVT systems in the literature is to apply the first law of thermodynamics [11, 14]; this is called thermal energy and is described in Equation

2.9:

퐸 휂 = 푇 (2.9) 푡ℎ푒푟푚푎푙 퐸

퐸푇 = 푚푤푎푡푒푟 ∗ 퐶푊푎푡푒푟 ∗ 훥푇푤푎푡푒푟 (2.10)

where 퐸푇 is the thermal output power and 퐸 is the irradiation per unit area.

The first law for thermal energy estimates the amount of energy contained in the system. This fact itself, however, doesn’t give enough information to decide the work potential of the system. A temperature difference between the heat source and the heat sink is necessary in order to extract energy as useful work as known as 2nd law of thermodynamics. This concept is also called exergy or available energy. The exergetic efficiency of system is stated as in Equation 2.11 [15]:

푇푎 휀푡ℎ푒푟푚푎푙 = (1 − ) ∗ 휂푡ℎ푒푟푚푎푙 (2.11) 푇표푢푡

where Ta is the ambient temperature and Tout is the temperature of the electrolyte leaving the electrolyzer.

Therefore the PVT efficiencies based on the 1st (energetic) and 2nd (exergetic) laws are listed below, respectively

휂푃푉푇 = 휂푃푉 + 휂푡ℎ푒푟푚푎푙 (2.12)

휀푃푉푇 = 휂푃푉 + 휀푡ℎ푒푟푚푎푙 (2.13)

18

2.3.3. Analysis of the Hydrogen Production Efficiency

In the electrolysis of water, a DC electric current passes across two electrodes separated by electrolyte and the two electrodes are connected through an external circuit. Water decomposes into hydrogen at cathode and oxygen at anode. Most of the commercially available electrolyzers are alkaline and concentrated KOH solutions are used as electrolyte [16]. The half reactions take place on the cathode and the anode are written as below, respectively:

+ − 푐푎푡ℎ표푑푒 ∶ 2퐻 + 2푒 → 퐻2 (2.14)

1 푎푛표푑푒 ∶ 2푂퐻− → 푂 + 퐻 푂 + 2푒− (2.15) 2 2 2

The overall water splitting reaction via electrolysis is written as in Equation 2.16.

1 퐻 푂 → 퐻 + 푂 (2.16) 2 2 2 2

At standard conditions (25ºC and 1 bar) the splitting of water is a non-spontaneous reaction and the minimum amount of work that is required to split water can be obtained from the change in Gibbs free energy (ΔG°=237kJ/mol), or in terms of voltage:

훥퐺° 237000퐽 (2.17) 푈표 = = 푟푒푣 푧 ∗ 퐹 2 ∗ 96485퐶

where z is the number of electrons transferred per hydrogen molecule and F is the Faraday’s constant.

Water decomposition is an endothermic reaction where the total amount of energy needed in water electrolysis is equivalent to the change in enthalpy (훥퐻°=286kJ/mol). The rest of the energy to achieve

ΔH° can be supplied via heat (푇훥푆° = 훥퐻° − 훥퐺°). Therefore the total energy demand can be expressed by the thermo-neutral cell voltage as below:

훥퐻° 286000퐽 (2.18) 푈표 = = 푡ℎ 푧 ∗ 퐹 2 ∗ 96485퐶

19

표 표 Total energy demand for both 푈푟푒푣 and 푈푡ℎ change slightly with temperature. At standard

표 표 표 conditions (25ºC and 1 bar) 푈푟푒푣 = 1.229푉 and 푈푡ℎ = 1.482푉 whereas at 80°C 푈푟푒푣 = 1.184푉 and

표 푈푡ℎ = 1.473푉.

On the other hand, anodic and cathodic overpotentials are affected by temperature. As the operating temperature increases at a given current density, the overvoltages decrease due to faster electrode kinetics [17, 18]. Advanced commercial water electrolyzers do not require heating systems since a significant portion of the energy dissipated is used as heat input in order to operate the electrolyzer at adequate temperatures to reduce the cell overvoltages. The heat produced is described as the difference between the actual cell potential and thermoneutral voltage as seen in Equation 2.19 [19].

푄̇ 푔푒푛 = 퐼 ∗ (푈푐푒푙푙 − 푈푡ℎ) (2.19)

The electrolyzer in this work aims to reduce actual cell potential by utilizing the excess heat in PV cells. Therefore, the additional energy input to operate the electrolyzer at elevated temperatures will be eliminated.

2.4. Results and Discussion

2.4.1. COMSOL Simulation and Temperature Distribution

Figure 2.3 depicts the results of COMSOL simulations for a 0.4 mm thick chamber at an irradiance of 1000W/m2, ambient temperature of 25°C and the velocity of 0.17 mm/s and 0.68 mm/s. For the velocity of 0.17 mm/s the temperature gradient across the module can be seen in Figure 2.3.a. The highest temperature is 54.8°C at the top of the glass layer and it is around 45°C at the bottom of the module. Figure 2.3.b represents the electrolyte in the chamber with the inlet temperature of 25°C and outlet of 52°C. The average temperature of the electrolyte in the chamber is 53°C. The velocity, as expected, plays an important role in the heat transfer for the PVT systems. Figure 2.3.c and 2.3.d represent the results for the velocity of 0.68 mm/s. The temperature of the solar cell, electrolyte in the

20 chamber and the outlet are 43°C, 42°C and 41.5°C, respectively. Higher velocities provide higher thermal recovery and allow the PV cells operate at lower temperatures with higher electrical efficiency.

Meanwhile, the temperature of electrolyte approaches the ambient temperature and affects adversely the hydrogen production efficiency. At this point it is crucial to figure out the optimum velocity of the electrolyte and the chamber thickness to achieve the highest system efficiency.

2.4.2. Optimization of Chamber Thickness and Fluid Velocity in the Chamber

The overall efficiency of the system depends on the solar radiation, ambient temperature, chamber thickness and fluid velocity in the chamber. The first two parameters depend on the location and time.

The effect of the last two parameters which can be controlled was investigated in the literature [7, 20].

It was reported that when the velocity increases the thermal recovery from the solar cell is improved whereas the temperature difference between the inlet and the outlet decreases. As expected, a larger channel diameter at a constant velocity allows higher heat storage in the fluid. In our design, our approach is to minimize the channel thickness down where the following criteria are provided. i) the temperature difference between the inlet and the outlet of the electrolyte should be larger

than 15°C ii) the maximum overall efficiency is targeted

In order to evaluate the overall efficiency of the system we take into account PVT and electrolysis efficiencies as described in the Equation below:

표 휂푃푉푇퐸 = 휂푃푉푇 ∗ 푈푟푒푣 (2.20)

We simulated the system at different chamber thicknesses and electrolyte velocities under standard test conditions – an irradiance of 1000W/m2 and 25°C – to find out the optimum parameters. In PEM water electrolyzers the thickness of the channel is taken as 1 mm in order to allow better gas removal and management in the stacks [21, 22]. Therefore we limit the maximum thickness as 1 mm and go

21 down to 0.2 mm for our design. Figure 2.4 depicts temperature difference between the inlet and the outlet of the electrolyte and it decreases with increasing flow rate and increasing chamber thickness.

We are able to achieve 25°C at the flow rate of 0.34 mm/s and the thickness of 0.4 mm. When we take the flow rate as 1.02 mm/s and the thickness as 1 mm, the temperature difference dropped to 10°C.

The data points above the dash line in Figure 2.4 obeys the first criteria and these points are marked in orange color in Figure 2.5 which presents the overall efficiency for each chamber thickness at different velocities. The overall efficiency increases with the increase of flow rate and chamber thickness.

Among the overall efficiencies which follows the constraint of temperature difference, when the chamber thickness is 0.4 mm at the velocity of 1.36 mm/s, it reaches its highest value. At this point the temperature difference is more than 15°C and the total efficiency is 43%.

2.4.3. Hourly Variations of the System Efficiency

We investigated how the efficiencies vary hourly in Phoenix, AR in May 2010. The average hourly radiation and ambient temperature data in May is shown in Figure 2.6. We used the optimum parameters for the chamber thickness and velocity as 0.4 mm and 1.36 mm/s, respectively. Figure 2.7 represents the efficiency of a PV panel integrated with an electrolyzer system and a PV panel itself. As stated in the literature [15, 23], the temperature of the PV cells have the highest temperature at noon time and this increase is resulted in the minimum electrical efficiency for both cases. The PV cells with a module always perform better than the stand-alone PV system due to active cooling of the PV cells by electrolyzer. The improvement in the electrical efficiency is significant at noon time. The PV panel temperature decreases from 65°C to 48°C yielding an increase of 1.5% in the efficiency. Although at

8AM and 6PM have the same radiation, the efficiencies are not identical due to the ambient temperature being around 30°C at 6 PM leading PV cells operate at relatively higher temperatures.

The PVT efficiencies calculated based on 1st (Equation 2.12) and 2nd (Equation 2.13) laws can be seen in Figure 2.8. As we stated before, the 1st law represents the heat gain from the system whereas

22

2nd law indicates the work function. Although the heat recovery increases for the solar cell at noon, the energetic PVT efficiency decreases from 60% to 56%. The efficiency drop of 2.5% corresponds to thermal recovery and of 1.5% is for electrical efficiency. On the other hand, exergetic PVT efficiency makes a peak at noon time even the electrical efficiency decreases. The key parameter for exergetic thermal efficiency, the temperature difference between inlet and outlet of the electrolyzer, is expanding when the radiation and ambient values are getting larger as stated in the work by Chow et al [24]. The exergetic PVT efficiency gives an idea for the hydrogen production efficiency which has almost the same trend during the day as seen in Figure 2.9. It depicts the hydrogen production efficiency in the module and at ambient temperature. The solar radiation and ambient temperature have a direct effect on efficiency values. At 1PM the ambient temperature is 29°C and the electrolyte temperature in the chamber is at 48°C. The PVTE brings an efficiency gain from 74% to 78%. At 6PM the efficiency is higher than the value at 8AM due to a higher ambient temperature and higher module temperature at

6PM.

2.4.4. Monthly Variations of the System Efficiency

We analyzed monthly variations in the electrical, thermal, and hydrogen efficiencies for the climate of Phoenix, AZ in 2010. The data represent the average efficiencies in each month which means these values can vary during the day as in Figure 2.7-9. The monthly variations in electrical output of the PV with and without the module are depicted in Figure 2.10. Between May and September there is a 5% increase in the electrical efficiency for the PV device with the module over the PV cells without the module. In the winter months it is 3%, and for the other months it is around 4%. It can be explained as during the summer time the cell temperatures go up to a higher value in which the thermal recovery from the solar cell panel becomes more significant. As is expected during the winter time, the temperature of the PV cell is close to normal standard operating temperatures, providing a higher electrical efficiency.

23

Figure 2.11 represents energetic and exergetic PVT efficiencies for each month. Both of these efficiencies decrease in the summer time due to a higher ambient temperature leading to higher heat energy accumulation in the PV cell. Although there is room to recover more heat from the PV cells, increasing the heat collection will compromise the system's overall performance. Moreover, Figure

2.12 represents the thermal and exergetic gain in terms of kWh for each month. Although PVT efficiencies decrease during the summer, energetic and exergetic energy values increase. The month of May has the highest energy gain from the system. The gains follow the same trend as they increase until May, then they plateau during the summer and start to decrease again. The thermal gain in May is double that of the month of January, whereas the exergetic gain in May is 2.95 times that of the value in January. If we take the month of January as a benchmark, these results suggest that the exergetic energy gain is more significant than energetic energy gain in May due to higher solar intensities and less ambient temperature differences than during the summer months. As a result, our system performs energetic efficiencies between 56% and 59% during the year where Chow et al reported [25] that the energetic PVT efficiency varies between 52% and 58%.

The hydrogen production efficiency is at the highest value during the month of July, as can be seen in Figure 2.13. In month of July it reaches up to 78% whereas it is around 73% in the month of January.

The efficiency gain, relative the electrolysis at ambient temperature, is more than 3% for the months from June to September. Although the radiation and ambient temperature is low during the winter, the efficiency gain is around 3%. These results indicate that our proposed system is able to increase the hydrogen production efficiency while the keeping the energetic PVT efficiency in the range stated in the literature.

24

2.5. Conclusion

This proposed PVTE system has advantages over separate PV module and water electrolyzer systems; not only through utilizing heat into more efficient hydrogen production, but also it has less of a spatial footprint and less installation costs. Based on COMSOL Multiphysics software and numerical models, exergetic and hydrogen production analysis of the PVTE system has been done. Although the heat recovery from the PV cell can be increased by increasing the velocity of the electrolyte, this approach will decrease exergetic and hydrogen production efficiencies. The optimization results indicate that the maximum overall efficiency of the system was achieved when the chamber thickness is 0.4 mm at a velocity of 1.36 mm/s. The average electrolyte temperature difference of the inlet and the outlet of the system is greater than 15°C.

Hourly and monthly efficiency analyses were studied using Phoenix, AZ in the year 2010 as a template. The PVTE system brought a 1% increase in electrical efficiency with respect to the stand- alone PV system. The thermal efficiency changes between 56% and 59% whereas exergetic efficiency varies between 15% and 16%. The main motivation of this system, the water electrolyzer, presents improvements as high as 4% in the efficiency of hydrogen production.

Our future reports will explore the fabrication and characterization of the photovoltaic thermal water electrolyzer. We will focus on the gas management in the electrolyzer. Another interest of our work will be how the performance of the electrolyzer changes at different potentials and temperatures.

2.6. References

1. Hollmuller, P., et al., Evaluation of a 5 kW(P) photovoltaic hydrogen production and storage installation for a residential home in Switzerland. International Journal of Hydrogen Energy, 2000. 25(2): p. 97-109.

2. Clua, J.G.G., R.J. Mantz, and H. De Battista, Hybrid control of a photovoltaic-hydrogen energy system. International Journal of Hydrogen Energy, 2008. 33(13): p. 3455-3459.

25

3. Djafour, A., et al., Photovoltaic-assisted alkaline water electrolysis: Basic principles. International Journal of Hydrogen Energy, 2011. 36(6): p. 4117-4124.

4. Gibson, T.L. and N.A. Kelly, Optimization of solar powered hydrogen production using photovoltaic electrolysis devices. International Journal of Hydrogen Energy, 2008. 33(21): p. 5931- 5940.

5. Tyagi, V.V., et al., Progress in solar PV technology: Research and achievement. Renewable & Sustainable Energy Reviews, 2013. 20: p. 443-461.

6. Skoplaki, E. and J.A. Palyvos, On the temperature dependence of photovoltaic module electrical performance: A review of efficiency/power correlations. Solar Energy, 2009. 83(5): p. 614-624.

7. Charalambous, P.G., et al., Photovoltaic thermal (PV/T) collectors: A review. Applied Thermal Engineering, 2007. 27(2-3): p. 275-286.

8. Shah, A., et al., Photovoltaic technology: The case for thin-film solar cells. Science, 1999. 285(5428): p. 692-698.

9. Kalogirou, S.A., Solar thermal collectors and applications. Progress in Energy and Combustion Science, 2004. 30(3): p. 231-295.

10. Jones, A.D. and C.P. Underwood, A thermal model for photovoltaic systems. Solar Energy, 2001. 70(4): p. 349-359.

11. Fujisawa, T. and T. Tani, Annual exergy evaluation on photovoltaic-thermal hybrid collector. Solar Energy Materials and Solar Cells, 1997. 47(1-4): p. 135-148.

12. Evans, D.L. and L.W. Florschuetz, COST STUDIES ON TERRESTRIAL PHOTOVOLTAIC POWER-SYSTEMS WITH SUNLIGHT CONCENTRATION. Solar Energy, 1977. 19(3): p. 255-262.

13. Chow, T.T., Performance analysis of photovoltaic-thermal collector by explicit dynamic model. Solar Energy, 2003. 75(2): p. 143-152.

14. Bhargava, A.K., H.P. Garg, and R.K. Agarwal, STUDY OF A HYBRID SOLAR-SYSTEM SOLAR AIR HEATER COMBINED WITH SOLAR-CELLS. Energy Conversion and Management, 1991. 31(5): p. 471-479.

15. Rajoria, C.S., S. Agrawal, and G.N. Tiwari, Exergetic and enviroeconomic analysis of novel hybrid PVT array. Solar Energy, 2013. 88: p. 110-119.

16. Zeng, K. and D.K. Zhang, Recent progress in alkaline water electrolysis for hydrogen production and applications. Progress in Energy and Combustion Science, 2010. 36(3): p. 307-326.

17. Appleby, A.J., G. Crepy, and J. Jacquelin, HIGH-EFFICIENCY WATER ELECTROLYSIS IN ALKALINE-SOLUTION. International Journal of Hydrogen Energy, 1978. 3(1): p. 21-37.

18. Hug, W., et al., HIGHLY EFFICIENT ADVANCED ALKALINE ELECTROLYZER FOR SOLAR OPERATION. International Journal of Hydrogen Energy, 1992. 17(9): p. 699-705.

26

19. Dieguez, P.M., et al., Thermal performance of a commercial alkaline water electrolyzer: Experimental study and mathematical modeling. International Journal of Hydrogen Energy, 2008. 33(24): p. 7338-7354.

20. Teo, H.G., P.S. Lee, and M.N.A. Hawlader, An active cooling system for photovoltaic modules. Applied Energy, 2012. 90(1): p. 309-315.

21. Casati, C., et al., Some fundamental aspects in electrochemical hydrogen purification/compression. Journal of Power Sources, 2008. 180(1): p. 103-113.

22. Grigoriev, S.A., et al., Mathematical modeling of high-pressure PEM water electrolysis. Journal of Applied Electrochemistry, 2010. 40(5): p. 921-932.

23. Tiwari, G.N., R.K. Mishra, and S.C. Solanki, Photovoltaic modules and their applications: A review on thermal modelling. Applied Energy, 2011. 88(7): p. 2287-2304.

24. Chow, T.T., et al., Energy and exergy analysis of photovoltaic-thermal collector with and without glass cover. Applied Energy, 2009. 86(3): p. 310-316.

25. Chow, T.T., A review on photovoltaic/thermal hybrid solar technology. Applied Energy, 2010. 87(2): p. 365-379.

27

2.7. Figures and Tables

a

b

Figure 2.1. (a) A Simple overview of the PVTE system, (b) Working mechanism and design of the electrolyzer.

28

a b

Figure 2.2. (a) A model drawing of the module with layers and thicknesses, (b) A cross-sectional view of an electrolyzer unit cell.

29

a b

c d

Figure 2.3. Temperature gradient of the module (a,c) and the electrolyte (b,d) for 0.4 mm thick chamber at the radiation of 1000W/m2 and ambient temperature of 25°C at the velocity of 0.17 mm/s (a,b) and 0.68 mm/s (c,d).

30

30

25 C

° 20

15

Tamb Tamb -

10 Tout

5 0.2 0.4 0.6 0.8 1 0 0.34 0.68 1.02 1.36 1.7 2.04 Flow Rate (mm/s)

Figure 2.4. Temperature difference between inlet and outlet of electrolyte is at different flow rates and chamber thicknesses at 1000W/m2 radiation and 25°C ambient temperature.

0.60

0.50

0.40

0.30 Efficiency (%) Efficiency 0.20 0.2 0.4 0.6 0.8 1 0.10 0.34 0.68 1.02 1.36 1.7 2.04 Flow Rate (mm/s)

Figure 2.5. Overall system efficiency is at different flow rates and chamber thicknesses at 1000W/m2 radiation and 25°C ambient temperature.

31

1400 35

)

2 C)

30 o 1100 25 800 Radiation 20 500 Temperature

15

Temperature ( Temperature Radiation (W/m Radiation 200 10 7 9 11 13 15 17 19 hour of the day

Figure 2.6. The average hourly radiation and ambient temperature data in May, 2010.

0.155 PV with module 0.150 PV alone 0.145 0.140 0.135 0.130 0.125 Electrical efficiency Electrical 0.120 7 8 9 10 11 12 13 14 15 16 17 18 19 hour of the day

Figure 2.7. Average hourly variations in electrical efficiency of PV panel integrated with electrolyzer system and PV panel itself in May, 2010.

32

0.60 0.170 1st Law 0.59 2nd Law 0.165

0.58 0.160

0.57 0.155

PVT efficiency PVT 0.56 0.150

0.55 0.145 7 9 11 13 15 17 19 hour of the day

Figure 2.8. Average hourly variations in PVT efficiency of the system in May, 2010.

0.80 in the electrolyzer 0.78 @ T amb

0.76

0.74

0.72

0.70 production efficiency production

2 7 8 9 10 11 12 13 14 15 16 17 18 19 H hour of the day

Figure 2.9. Average hourly variations in hydrogen production efficiency of the PVTE system in May, 2010.

33

0.155 PV with module 0.150 PV alone 0.145

0.140

0.135

0.130

Electrical efficiency Electrical 0.125

0.120 Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec

Figure 2.10. Average monthly variations in electrical efficiency of PV panel integrated with electrolyzer system and PV panel itself in Phoenix, AR (2010).

0.60 0.175 1st Law 0.59 2nd Law 0.17

0.58 0.165

0.57 0.16

PVT efficiency PVT 0.56 0.155

0.55 0.15 Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec

Figure 2.11. Average monthly variations in PVT efficiency in Phoenix, AR (2010).

34

160 10 1st Law 140 2nd Law 8

2 120 6 100 4

KWh / / m KWh 80

60 2

40 0 Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec

Figure 2.12. Average monthly variations in thermal gain of the PVTE system in Phoenix, AR (2010).

0.80 in the electrolyzer 0.78 @ T amb

0.76

0.74

0.72

production efficiency production 2

H 0.70 Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec

Figure 2.13. Average monthly variations in hydrogen production efficiency of the PVTE system in Phoenix, AR (2010).

Table 2.1. Properties of each material used in COMSOL simulation Density Specific heat Thermal conductivity Layer/material kg/m3 J/kg/K W/m/K Glass 2203 703 1.38 EVA 950 3135 0.23 PV cell (Si) 2330 700 130 Tedlar 1200 1090 0.16 PDMS 970 1460 0.15

35

CHAPTER 3

FABRICATION OF PHOTOVOLTAIC THERMAL WATER ELECTROLYZER

3.1. Introduction

Hydrogen is an environmentally friendly and sustainable energy carrier for our economy [1]. A sustainable way to produce hydrogen is electrolysis of water from renewable energy sources. Among renewable energy sources, solar energy is chief among renewables that show potential for great growth

[2]. Past literature has explored systems coupling together a photovoltaic (PV) system and an alkaline water electrolyzer to optimize the voltage and maximum power output of the PV to the operating voltage of electrolyzer [3-7]. In water electrolyzers, researchers have changed the electrode properties, the type of and concentration of electrolyte, as wells as operating pressure and temperature in order to promote the rate of hydrogen production [8, 9].The temperature has a positive effect on the efficiency of water electrolyzers. As the operating temperature increases at a given current density, the overpotentials at anode and cathode decrease due to faster electrode kinetics [10, 11]. In industrial applications of water electrolysis, operating temperatures vary between 60oC to 80oC under voltages of 1.8-2V in alkaline water electrolyzer [8]. Advanced commercial water electrolyzers do not require heating systems since they are designed to reach and hold this temperature at a specific current load through internal heat generation [12]. In order to operate within elevated temperature range, extra energy load increases operating cost of the system.

36

Currently PV-assisted water electrolyzers available on the market have PV conversion efficiency between 15% and 17%. The energy that is not transformed into electricity is dissipated as heat. We recently reported the concept and energy analysis of photovoltaic thermal water electrolyzer (PVTE) which consists of PV cells positioned on top of a planar micro water electrolyzer in order to harness waste heat as a storable form of energy. The similar concept is found in the market, a so-called photovoltaic/thermal (PVT) collectors which combines PV cells and solar thermal collector together in order to excess heat dissipated from the PV cells [13]. In PVT collector, solar energy is harnessed by using a photovoltaic (PV) module that converts solar energy into electricity or a solar thermal collector which transforms solar energy into heat, e.g. for domestic hot water or room heating [14, 15].

Figure 3.1.a shows a simple overview of the PVTE system, and Figure 3.1.b depicts the working mechanism and design of the electrolyzer. The PVTE design provides a new output in terms of generating hydrogen efficienctly. The harnessed energy increases the temperature of the electrolyte in the electrolyzer leading to overcome the of kinetics. In previous chapter showed that at the radiation of 1000W/m2 and ambient temperature of 35°C, the temperature of the electrolyte reaches to 60°C when we set the flow rate to 0.5 mL\min. It is possible to achieve elevated temperatures under higher radiation and higher ambient temperatures.

In this work, we fabricated a micro-water electrolyzer which utilizes heat from PV cell and works as a heat sink in order to eliminate additional energy input for electrolysis in order to operate at elevated temperatures. Our approach is first to pump and reside the electrolyte in the chamber at specific time intervals in order to transfer heat to the electrolyte. Finally all the electrolyte and evaluated gases are evacuated from the chamber through two outlets by a syringe pump. We also presented electrode preparation and fabrication of the electrolyzer. To use a solar cell in this experimental setup is beyond the scope of this work. Instead we applied a constant temperature on top of the water electrolyzer to

37 see the effect of temperature on hydrogen production in this system. The increase in the hydrogen production rate affirms the predictions of our system that utilizes waste heat from PV.

3.2. Design of Water Electrolyzer

Microfluidic energy conversion devices have been investigated in the field of micro fuels cells

[16-18], microfluidic biofuel cell [19], and micro flow batteries [20, 21]. Microfluidics have high surface area to volume ratio of small components facilitating rapid heat and mass transfer leading to rapid temperature changes and reaction kinetics. Our design is based on a microfluidic chamber which utilizes excess heat of PV cells. The intention of planar design is not only to separate the gases but also to focus on the thermal management aspect. Briefly, a view of an electrolyzer unit cell, in the dimensions of 40 mm length, 40 mm width and about 11 mm height, is seen in Figure 3.1.c. We used

1 mm thick non-electrically conductive glass substrate in which the temperature difference between each side is less than 1oC for 1000 W/m2 heat input. Two thin film electrodes that are 1 cm2 in area are located on the glass substrate and in the chamber. A 7 mm thick Polydimethylsiloxane (PDMS) slab was chosen as the electrolyzer chamber material due to its low thermal conductivity. The PDMS reservoir includes an inlet, where electrolyte at ambient temperature enters, and two outlets apart from each other by a separator for H2 and O2 gases with the heated electrolyte to exit the chamber.

The size of the electrolyzer chamber is 20 mm x 20 mm x 0.4 mm. The thickness of the chamber is crucial for hydrogen production rate and efficiency of the all the system. After the detailed energy analysis of the PVTE system, we found out the optimum flow rate and chamber thickness as

0.65mL/min and 0.4 mm, respectively. The other reason why we take the chamber thickness is a possibility of ohmic losses (IR drop). Ito et al [22] studied cyclic voltammograms of potassium ferrocyanide (0.01mol dm-3) in sodium sulfate buffer (0.5mol dm-3) of planar electrodes at different micro channel thickness (from 30 µm to 520 µm). They reported that the cell potential decreases with

38 increasing the channel thickness as the potential decreases dramatically by 200 µm and changes slightly between 200 µm and 500 µm. The attribution of this behavior is explained by an IR drop because the shallow microchannel increases the electrical resistance. Also in our preliminary experiments we compared the performance of the electrolyzer at the thickness of 200 and 400 um. The data showed that thicker chamber design provides higher hydrogen production at a constant potential and will be discussed in the section 4.3.

3.3. Experimental

3.3.1. Electrode Preparation

Figure 3.2 (a-d) summarizes the electrode preparation. The electrode pattern was created on a glass slide (40mmx40mmx1mm) through photolithography. The thin film electrodes were obtained by electron-beam deposition of a 10 nm Ti adhesion layer, followed by 100 nm Au layer onto the glass slide while maintaining a vacuum of ~2x10-7 mbar. Then, Au deposited 1 cm2 electrodes were electrochemically cleaned by potential cycling from 0.2 to 1.45V (Ag/AgCl) at 25 mV/s in 0.5M H2SO4 solution. After this cleaning step, electrodes were electrochemically platinized at 0.5 V (Ag/AgCl) for

30 s in a 3 wt. % H2PtCl6 ·6H2O (Aldrich) + 0.03% Pb-acetate (Aldrich) solution in water to yield porous Pt electrodes. Following the deposition, the electrodes were thoroughly rinsed with milli-Q water and electrochemically cleaned by potential cycling from -0.2 to 1.1 V (Ag/AgCl) at 25 mV/s in

0.5M H2SO4 solution. The experiments were performed in a conventional three-compartment electrochemical cell (Model 660D, CH Instrument), using a Pt-wire counter electrode and an Ag/AgCl reference electrode.

3.3.2. Fabrication of the Water Electrolyzer

Figure 3.2 (e-f) also explains the fabrication process. Rapid prototyping technique was used for the

PDMS mold. The PDMS base and curing agent (Dow Corning) were mixed at a ratio of 10:1 and

39 poured onto the mold and cured at 70°C for 2 hours. The PDMS reservoir was peeled from the mold and an inlet and two outlet holes were formed by a biopsy punch. UV cleaning for 5 minutes was applied to both the glass slide and PDMS reservoir [16]. Both surfaces were brought into contact immediately to chemically bond them and the integrated of the water electrolyzer is seen in Figure 3.3.

3.3.3. Electrochemical Testing

o All experiments were done after electrolysis in 1 M H2SO4 at 60 C and under applied voltage of

1.8V for 2 hours. To calculate the polarization overpotential, each glass substrate with electrodes was immersed in 0.5 M H2SO4 electrolyte in a conventional three-compartment electrochemical cell with an Ag/AgCl reference electrode. The hot plate and probe were used to control the temperature of the electrolyte. Figure 3.4 depicts the electrolyzer experimental setup. In the experiments, 30 mm x 30 mm polyimide film insulated flexible heater (Omega Engineering) was attached on the glass substrate and a constant temperature was applied on the water electrolyzer through the glass side of the water electrolyzer. 1M H2SO4 at room temperature was sent to the chamber from the reservoir. Total voltage of 1.8V and 2V were applied to the electrodes continuously. After that all the electrolyte and evaluated gases are evacuated from the chamber through two outlets by a syringe pump (Harvard PHD-2000).

3.4. Results and Discussion

3.4.1. Electrode characterization

Figure 3.5 represents the cyclic voltammetry of Au thin film and Pt-black catalyst after electro deposition on Au thin film. The characteristic behavior of Au and Pt was achieved in each case. Figure

3.6 depicts polarization characteristics of the Pt-black electrode in electrolyte of 0.5 M H2SO4 at different temperatures. For the current density of 10 mA/cm2, the efficiency of the electrolysis is 60% when the temperature of the electrolyte is at room temperature. On the other hand, the efficiency goes

40 up to around 70% when the temperature of the electrolyte is increased to 60oC. It can be also seen how the overpotentials of the anode and cathode are positively affected by the increase in the temperature.

Pt-black catalyst loses its activity during the electrolysis over time. Different approaches in the literature were taken such as temperature annealing or electrolysis at close to the desirable operating conditions for a specific time in order to be consistent in the experiments. We operated at 60oC under applied voltage of 1.8V. Figure 3.7 indicates that minimum 2 hour electrolysis is required to eliminate the effect of activity loses for the electrode in order to be consistent in the experiments. In Figure 3.8,

SEM images can be seen before and after electrolysis at 60°C and under applied voltage of 1.8V for 2 hours. The morphology changes after electrolysis resulting in lower surface area and lower hydrogen evaluation.

3.4.2. Effect of Flow Rate on Hydrogen Production Rate

The flow rate of the electrolyte pumping into the electrolyzer plays significant role on the performance of the electrolyzer. Figure 3.9 shows the hydrogen production in terms of current flux at temperature of 25°C and 50°C for different flow rates. For the temperature of 25°C, when the flow rate is 0.5 and 0.65 ml/min, the hydrogen production decreases over time. This result is attribution to the evolved gases are stuck in the chamber leading to loss of interaction between electrode and electrolyte.

It can be seen that when the flow rate is increased to 5 mL/min, the current flux is consistent during the time and the results indicates the gases are not stuck in the chamber. As stated by Kjeang et al. [23] current density increases with flow rate, which indicates that mass transport limitations are significant and that the electrochemical reactions are relatively fast. On the other hand, when the electrolyzer is operating at 50°C, it can be definitely seen that increase in the temperature of the electrolyzer does not have an improvement in the hydrogen production for the flow rate of 5 mL/min due to less time contact for heat transfer. At the flow rate of 0.5 and 0.65 ml/min, the hydrogen production is higher than the temperature of 25°C. Unfortunately, the hydrogen production decreases over time due the stuck of the

41 gases in the chamber as the temperature of 25°C. As a conclusion, at low flow rates gas is stuck in the chamber but at higher flow rates heat recovery is not achieved when the temperature of the electrolyzer is increased.

In order to overcome this challenge, we proposed a new approach as called residence time [24]. As can be seen in Figure 3.10, electrolyte is pumped into the electrolyzer chamber and allow it to reside in the chamber for a specific time interval to recover the heat. Then electrolyte with gases are flushed from the electrolyzer at high flow rate. In this work, we used the flow rate of 10 mL/min with the volume of 0.16 mL that corresponds to the volume of the chamber. Figure 3.11.a depicts the hydrogen production at the temperature of 25°C for the residence time of 60 seconds. It can be seen that after flushing the electrolyte in the chamber the hydrogen production rate regenerates itself and goes down to by the flushing out. When the operating temperature is at 50°C, as can be seen in Figure 3.11.b, after the flushing out the electrolyte in the chamber, three different section is observed for the hydrogen production rate. First the electrolyte temperature increases by the heat recovery. After a time, generated gases start to block the interaction between electrode and the electrolyte and hydrogen production rate starts to decrease dramatically. We can explain this attribute as mass transport limitations as described before [18, 23].

3.4.3. Effect of Chamber Thickness on Hydrogen Production Rate

As the bulk systems, bubble formation on the electrode and inside the microchannel is the major cause of transport resistance. We investigated the effect of channel thickness on hydrogen production rate. Figure 3.12 depicts the hydrogen production current for two different channel thicknesses of 200 and 400 μm. We applied 1.8 V at room temperature when the electrolyte is stationary in the microfluidic device. At 25oC it can be seen that for a constant applied potential 400 μm thick channel device gives higher current. We believe that the difference in hydrogen production between different channel thicknesses is dominantly affected by the mass transfer limitations. Because, Ito et al [22]

42 reported that the potential drop is only 0.01V when they changed the channel thickness from 400 μm to 200 μm. We repeated the same experiment as increasing the temperature of the electrolyte from

25oC to 60oC. Figure 3.12 shows that hydrogen production current is almost same for two different thicknesses within the 5 seconds. In thinner design heat is transferred rapidly and it compensates the hydrogen production over thicker design. After 5 seconds when the temperature of electrolytes reaches to the same level for each thickness and we observe the same trend as at 25oC due to mass transport.

3.4.4. Effect of Residence time on Hydrogen Production Rate

For each temperature and applied voltage, hydrogen production rate will be different and optimization of residence time becomes important. We investigated the hydrogen production rate when the electrolyte resides in the chamber within 60 seconds. The data in Figure 3.13 reveals the behavior of rate at the voltage of 1.8 V and 2 V for different temperatures. When operating conditions are set to the potential of 1.8 V and the temperature is 40°C, the rate increases and is almost constant during the time. The behavior of rate changes when the temperatures are increased to 50°C and 60°C. The rate increases rapidly within 10 seconds and then decreases over the time. As explained in previous section higher temperatures provides higher rate within shorter time intervals. As the operating potential is set to 2 V, the rate reaches peak point within 5 seconds for all temperatures. If we compare the rates in the electrolyzer with polarization curve data (Figure 3.8) the highest rate of the potential of 1.8V within

60 seconds for each temperature is close to data. But for the voltage of 2 V, they are far away from data in Figure 3.8 due to mass transfer limitations. The more hydrogen and oxygen gases are generated, the more mass transfer limitations occur in the chamber of electrolyzer.

3.4.5. Effect of Residence time on Total Hydrogen Production

In order to estimate the amount of hydrogen production for each residence time we calculated the area under curves in Figure 3.13. Figure 3.14 presents the data for the voltage of 1.8 V and 2 V for all

43 temperatures investigated in this work. For the temperature of 40°C and 50°C at 1.8 V, 60 second residence time is necessary. When the electrolyte temperature is 60°C, 30 second residence time gives the highest hydrogen production. For the potential of 2V, hydrogen production reaches the highest amount when the residence time is set to 15 seconds for all temperatures. The extended time will be result in mass transfer limitations in the electrolyzer which leads to decrease in the hydrogen production.

Table 3.1 summarizes the ratio of total hydrogen production with respect to temperature of 25oC for different temperature and different residence time at the potentials of 1.8V and 2 V. As can be seen hydrogen production is achieved 7 times of the case when the temperature of the electrolyte is 25oC at the temperature of 60oC and the potential of 1.8 V for the 30 second residence time. On the other hand, at 2 V and 60oC, hydrogen production is obtained 2.5 times of the case when the temperature of the electrolyte is 25oC when the residence time is set to 15 seconds. Although the ratio is less for the potential of 2 V than of the 1.8, we should pay attention that the total hydrogen production is higher for the potential of 2V.

3.5. Conclusion

We summarize by noting that the micro-water electrolyzer fabrication described here can be integrated with a PV cell to enable utilizing heat recovery from PV cell. It works as a heat sink in order to eliminate additional energy input for electrolysis in order to operate at elevated temperatures. The aim of planar design is not only to separate the gases but also to provide the heat recovery. The challenge of this electrolyzer design is that at low flow rates gas is stuck in the chamber but at higher flow rates heat recovery is not achieved when the temperature of the electrolyzer is increased. Our approach is first to pump and reside the electrolyte in the chamber at specific time intervals in order to

44 transfer heat to the electrolyte then all the electrolyte and evaluated gases are evacuated from the chamber.

We optimized the residence time for the hydrogen production rate within 60 seconds. Whereas, for the voltage of 1.8 V more than 30 second residence time is necessary for all temperature, for the voltage of 2 V less than 15 second residence time is necessary. When the electrolyte temperature is 60°C, 30 second residence time gives the highest hydrogen production. For the potential of 2V, hydrogen production reaches the highest amount when the residence time is set to 15 seconds for all temperatures.

Hydrogen production is achieved 7 times of the case when the temperature of the electrolyte is 25oC at the temperature of 60oC and the potential of 1.8 V for the 15 second residence time. On the other hand, at 2 V and 60oC, hydrogen production is obtained 2.5 times of the case when the temperature of the electrolyte is 25oC when the residence time is set to 15 seconds. The increase in the hydrogen production rate affirms the predictions of our system that utilizes waste heat from PV.

3.6. References

1. Nowotny, J. and T.N. Veziroglu, Impact of hydrogen on the environment. International Journal of Hydrogen Energy, 2011. 36(20): p. 13218-13224.

2. Lewis, N.S., Powering the planet. Mrs Bulletin, 2007. 32(10): p. 808-820.

3. Hollmuller, P., et al., Evaluation of a 5 kW(P) photovoltaic hydrogen production and storage installation for a residential home in Switzerland. International Journal of Hydrogen Energy, 2000. 25(2): p. 97-109.

4. Clua, J.G.G., R.J. Mantz, and H. De Battista, Hybrid control of a photovoltaic-hydrogen energy system. International Journal of Hydrogen Energy, 2008. 33(13): p. 3455-3459.

5. Djafour, A., et al., Photovoltaic-assisted alkaline water electrolysis: Basic principles. International Journal of Hydrogen Energy, 2011. 36(6): p. 4117-4124.

6. Gibson, T.L. and N.A. Kelly, Optimization of solar powered hydrogen production using photovoltaic electrolysis devices. International Journal of Hydrogen Energy, 2008. 33(21): p. 5931- 5940.

7. Maeda, T., et al., Study on control method of the stand-alone direct-coupling photovoltaic - Water electrolyzer. International Journal of Hydrogen Energy, 2012. 37(6): p. 4819-4828.

45

8. Zeng, K. and D.K. Zhang, Recent progress in alkaline water electrolysis for hydrogen production and applications. Progress in Energy and Combustion Science, 2010. 36(3): p. 307-326.

9. Carmo, M., et al., A comprehensive review on PEM water electrolysis. International Journal of Hydrogen Energy, 2013. 38(12): p. 4901-4934.

10. Appleby, A.J., G. Crepy, and J. Jacquelin, HIGH-EFFICIENCY WATER ELECTROLYSIS IN ALKALINE-SOLUTION. International Journal of Hydrogen Energy, 1978. 3(1): p. 21-37.

11. Hug, W., et al., HIGHLY EFFICIENT ADVANCED ALKALINE ELECTROLYZER FOR SOLAR OPERATION. International Journal of Hydrogen Energy, 1992. 17(9): p. 699-705.

12. Dieguez, P.M., et al., Thermal performance of a commercial alkaline water electrolyzer: Experimental study and mathematical modeling. International Journal of Hydrogen Energy, 2008.

13. Chow, T.T., A review on photovoltaic/thermal hybrid solar technology. Applied Energy, 2010. 87(2): p. 365-379.

14. Kalogirou, S.A., Solar thermal collectors and applications. Progress in Energy and Combustion Science, 2004. 30(3): p. 231-295.

15. Shah, A., et al., Photovoltaic technology: The case for thin-film solar cells. Science, 1999. 285(5428): p. 692-698.

16. Mitrovski, S.M., L.C.C. Elliott, and R.G. Nuzzo, Microfluidic devices for energy conversion: Planar integration and performance of a passive, fully immersed H2-O2 fuel cell. Langmuir, 2004. 20(17): p. 6974-6976.

17. Choban, E.R., et al., Membraneless laminar flow-based micro fuel cells operating in alkaline, acidic, and acidic/alkaline media. Electrochimica Acta, 2005. 50(27): p. 5390-5398.

18. Jayashree, R.S., et al., Air-breathing laminar flow-based microfluidic fuel cell. Journal of the American Chemical Society, 2005. 127(48): p. 16758-16759.

19. Togo, M., et al., Structural studies of enzyme-based microfluidic biofuel cells. Journal of Power Sources, 2008. 178(1): p. 53-58.

20. Ferrigno, R., et al., Membraneless vanadium redox fuel cell using laminar flow. Journal of the American Chemical Society, 2002. 124(44): p. 12930-12931.

21. Kjeang, E., et al., A microfluidic fuel cell with flow-through porous electrodes. Journal of the American Chemical Society, 2008. 130(12): p. 4000-4006.

22. Ito, T., et al., Electrochemical behavior of parallel opposed dual electrode in a microchannel. Electroanalysis, 2004. 16(24): p. 2035-2041.

23. Kjeang, E., et al., High-performance microfluidic vanadium redox fuel cell. Electrochimica Acta, 2007. 52(15): p. 4942-4946.

24. Moth-Poulsen, K., et al., Molecular solar thermal (MOST) energy storage and release system. Energy & Environmental Science, 2012. 5(9): p. 8534-8537.

46

3.7. Figures and Tables

a

b c

Figure 3.1. (a) A Simple overview of the PVTE system, (b) Working mechanism and design of the electrolyzer, (c) A cross-sectional view of an electrolyzer unit cell. During the day, the PV cells generate electricity which is used in electrolysis and the rest is sent to the grid. The excess heat dissipated from the cells is captured by the water electrolyzer as a heat sink. The temperature of the electrolyte in the electrolyzer approaches the temperature of the PV cell. The raised temperature of the electrolyte in the electrolyzer aids to reduce the overpotentials for H2 and O2 gas evolution for a given current flux. Therefore, the energy consumption of the electrolyzer will be reduced with increasing efficiency of electrolysis. The electrolyte in the electrolyzer also functions as a heat-transport fluid and the electrolyte at an elevated temperature leaving the electrolyzer is circulated through tubing to transfer heat to an insulated water tank.

47

Figure 3.2. Micro-water electrolyzer fabrication flow scheme: (a)–(d) electrode preparation, (e)–(f) chamber fabrication, (g) integration of glass slide and chamber.

Figure 3.3. The fabricated water electrolyzer cell. 40 mm x 40 mm glass slide, 1 cm2 electrode area (each) and 7 mm thick PDMS chamber with 0.4mm thick chamber height.

48

Figure 3.4. The experimental setup of the system for water electrolyzer. The flow rate is controlled with the syringe pump and the temperature is set with flexible heater with PID system.

49

2.0 0.10 1.5 0.05

1.0

) 2 0.00 0.5

-0.05 0.0

-0.10 -0.5

-0.15 -1.0

current (mA/cm2) current current (mA/cm -1.5 -0.20 -2.0 -0.25 -2.5 -0.30 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 -0.4 -0.2 0.0 0.2 0.4 0.6 0.8 1.0 1.2 E - Ag/AgCl E - Ag/AgCl

Figure 3.5. Cyclic voltammetry of (left) Au thin film catalyst and (right) Pt-black catalyst after electrodeposition on Au thin film in 0.5 M H2SO4, at 50 mV / s.

50

a 2.0 Efficiency ~ 60%

1.9

1.8

Efficiency ~ 70%

V

- 1.7 E

E - V  25oC 1.6  40oC  50oC o 1.5  60 C  70oC

1.4 0 2 4 6 8 10

1.8 b

1.6

o )  25 C 1.4  40oC  50oC

Ag/AgCl o (

 60 C V

- o

E - V (Ag/AgCl)  70 C E

-0.2

0 2 4 6 8 10 i - mA/cm2 i-mA/cm2 Figure 3.6. (a) “Total potential” of the system and (b) ‘potentials’’ of the anode and cathode versus Ag/AgCl reference electrode at constant current densities at temperatures of room, 50 oC, 60 oC, 70 oC, and 80 oC. All experiments were performed after electrolysis at 60oC and under applied voltage of 1.8V for 2 hours.

51

12

10 2

8

6

4 current mA/cm current 2

0 0 1 2 3 4 5 6 7 8 9 10 11 12 time (hour)

Figure 3.7. The loss of Pt-black catalyst activity within 12 hours at 60oC and under applied voltage of

1.8V in the electrolyte of 1M H2SO4.

Figure 3.8. SEM images of Pt-black catalyst (a) before and (b) after electrolysis @60oC and under applied voltage of 1.8V for 2 hours in the electrolyte of 0.5 M H2SO4. The morphology changes after electrolysis resulting in lower surface area and lower hydrogen evaluation.

52

6 25°C "5 mL/min" 5

2 "1 mL/min" 4 "0.65 mL/min" "0.5 mL/min" 3

2 current mA/cm current 1

0 0 10 20 30 40 50 60 time (sec)

6

50°C "5 mL/min" 5

2 "1 mL/min" 4 "0.65 mL/min"

3 "0.5 mL/min"

2 current mA/cm current 1

0 0 10 20 30 40 50 60 time (sec)

Figure 3.9. The effect of flow rate on the hydrogen production at the temperature of 25oC and 50oC at the potential of 1.8V in the electrolyte of 1M H2SO4.

53

6

5

4

3

2 cell potential cell 1

0 0 50 100 150 200 250 300

6 Residence time 5

4

3

flow rate flow 2

1

0 0 50 100 150 200 250 300 time (sec)

Figure 3.10. This illustration represents the constant applied potential to the cell and the pulsed electrolyte velocity in the chamber. Residence time depicts the time interval in which the electrolyte rests in the chamber without any flow. The flow is applied in each 60 seconds at 10mL/min for a volume of 0.16 mL.

54

8 a 7

2 6 5 4 3

2 current mA/cm current 1 0 0 50 100 150 200 250 300 time (sec)

8 b

7 2 6 I II 5 4 3

2 current mA/cm current 1 0 0 50 100 150 200 250 300 time (sec)

Figure 3.11. The residence time effect on the hydrogen production at the temperature of (a) 25oC and (b) 50oC. At 50oC the sections indicate three different behavior during the time; i) heat recovery and rate increases, ii) generated gases block the interaction between electrode and electrolyte, mass transfer limitations. The electrolyte is 1M H2SO4.

55

8 25 oC 400 um 200 um 6

4

2 current mA/cm2 current

0 0 10 20 30 40 50 60 time (sec)

8 60 oC 400 um 200 um

2 6

4

2 current mA/cm current

0 0 10 20 30 40 50 60 time (sec)

Figure 3.12. The hydrogen production current at the temperature of 25oC and 60oC when we apply the potential of 1.8 V for the channel thicknesses of 200 um and 400 um in the micro electrolyzer. The flow rate is set to 0 mL/min. The electrolyte is 1M H2SO4.

56

5 1.8 V "60°C "

2 4 "50°C " "40°C " 3 "25°C "

2

current mA/cm current 1

0 0 10 20 30 40 50 60 time (sec)

10 2 V "60°C "

2 8 "50°C " "40°C " 6 "25°C "

4

current mA/cm current 2

0 0 10 20 30 40 50 60 time (sec)

Figure 3.13. The hydrogen production rate in terms of current within in 60 seconds at different temperatures and at the potential of 1.8V and 2V. The electrolyte is 1M H2SO4.

57

1.5 1.8 V "60°C " "50°C " "40°C " 1 "25°C "

0.5

0

hydrogen production (mL/hr) production hydrogen 0 15 30 45 60 75 residence time (sec)

4 2 V "60°C " "50°C " 3 "40°C " "25°C " 2

1

0

hydrogen production (mL/hr) production hydrogen 0 15 30 45 60 75 residence time (sec)

Figure 3.14. Total hydrogen production for different residence time and different temperatures at the potentials of 1.8 V and 2 V.

58

Table 3.1. The ratio of total hydrogen production with respect to temperature of 25oC for different temperature and different residence time at the potentials of 1.8V and 2 V. Temperature residence time the ratio wrt RT The ratio wrt RT (oC) (sec) @1.8V @2V RT 15-60 1 1 15 2.75 1.8 30 3.135 1.4 40 45 3.23 1.1 60 3.33 1.1 15 4.2 2.3 30 4.7 1.9 50 45 4.85 1.5 60 5 1.3 15 6.4 2.5 30 7 2.0 60 45 6.8 1.8 60 5.6 1.7

59

CHAPTER 4

GAS MANAGEMENT IN PHOTOVOLTAIC THERMAL WATER ELECTROLYZER

4.1. Introduction

Hydrogen is a sustainable energy carrier for our economy [1]. A renewable way to produce hydrogen is electrolysis of water from electricity by photovoltaic cells [2]. We have recently reported the fabrication and characterization of the “photovoltaic thermal water electrolyzer (PVTE)”. In PV cells, the energy that is not transformed into electricity is dissipated and PVTE provides an efficient way to benefit from this energy in hydrogen production. We describe in this chapter the fabrication and performance of a microfluidic water electrolyzer which is comprised of gas permeable membrane allowing gas removal from the chamber.

Microfluidic energy conversion devices have been widely studied in the literature: common examples are micro fuels cells [3-5], microfluidic biofuel cell [6], and micro flow batteries [7, 8].

Microfluidics are advantageous in comparison to conventional fluidic systems due the high surface area to volume ratio of small components facilitating rapid heat and mass transfer leading to rapid temperature changes and reaction kinetics.

One of the major practical challenges in microfluidics is the management of gas bubbles in the microchannels. Although bubbles can provide improved mixing for chemical reactions [9], and enhanced heat and mass transfer [10] , they accumulate and clog the channels [11] or reduce the electrode performance in the microfluidic device. Therefore, many efforts have been devoted to get rid

60 of the gas bubbles in the microfluidics such as locally varying wetting properties of channels and capillarity restricted modifications in the channels [12, 13], or dynamic bubble traps [14].

One approach is diffusion based gas bubble removal by a gas-permeable membrane. A thin PDMS layer [15, 16] or a hydrophobic membrane [10] are used for an active bubble trap and removal. Meng et al. [17] used a hydrophobic membrane providing a gas-venting microchannel that directly removes

CO2 gas bubbles from the electrochemical reactions of fuel cells without leakage. Another application of hydrophobic membranes in microfluidics is vapor venting two-phase microchannel heat exchanger

[10, 18]. When they used a hydrophobic membrane that vents the vapor phase into separate vapor transport channels, they were able to reduce the pressure head with power consumption for high heat flux generating electronics. A similar approach was taken by Winther-Jensen et al. for water electrolyzer [19]. They proposed a breathable electrode design which allows the diffusion of O2 and

H2 gases through a hydrophobic membrane leading to an improvement of the efficiency of the water splitting reaction.

The design of the water electrolyzer proposed previous chapter is challenging as the gas is stuck in the chamber at low flow rates –mass transfer limitations- and heat recovery is not achieved at high flow rates when the electrolyzer is operating at higher temperature increased. In order to overcome this drawback, we proposed a residence time approach. Although this approach shows an improvement in the heat utilization, the system is practically difficult to operate at varying the temperature of the electrolyte during the day. Herein, we propose a new design including a hydrophobic membrane allowing us to collect generated gases on top of the electrodes and overcome the clog of the bubbles in the chamber (Figure 4.1). Our fabrication technique is based on multilayer soft lithography in which a hydrophobic membrane can be located between two chambers of the electrolyzer.

61

4.2. Gas Management in Microchannels

In order to solve the bubble-clogging problem of microfluidic devices, particularly, microfuel cells such as μDMFC, a hydrophobic venting approach has been developed to directly remove gas bubbles from methanol aqueous solutions in the anodic microchannel [17]. Additionally hydrophobic PTFE membrane vents the vapor phase into separate vapor transport channels in microfluidic cooling systems

[18].

In our design, a hydrophobic porous membrane is placed between the chamber and gas collection outlets. As seen in Figure 4.2, the surface tension and capillary actions upon interaction with the hydrophobic membrane prevent the liquid solution from flowing through the membrane pores. Only the generated vapor can escape from the chamber to the gas collection outlets through the hydrophobic membrane, which realizes the liquid–vapor separation. The pressure of the generated vapor is higher than any other bulk fluid pressure leading to a non-equilibrium system: the pressure difference between the solution and vapor side of the membrane acts as a driving force for vapor to escape through the hydrophobic membrane.

4.3. Experimental

4.3.1. Fabrication of the Water Electrolyzer

In the electrolyzer fabrication, the critical step is embedding PFTE hydrophobic membrane of 1

µm pore size (Sterlitech Corporation) into the chamber. It is crucial to have a solid and leak free device.

Our design has two different chamber layers and the membrane is sandwiched between them. A straightforward soft lithography techniques does not work in terms of a leak free electrolyte. Unger et al. [20] developed a fabrication technique based on multilayers of PDMS microchannels for valves and pumps. The principle mechanism of this technique is to partially cure the PDMS base and curing agent at different mixing ratios, bring together and complete the curing. Figure 4.3 explains the fabrication

62 process. The PDMS with a mixing ratio of 15:1 is poured on the bottom mold. The thickness of the

PDMS layer was controlled with a spin-coater. To reach a thickness of 0.45 mm, we spin the PDMS at 120 rpm for a minute. For the top part of the PDMS chamber, the PDMS with mixing ratio of 5:1 is poured on the other mold. The thickness of this part is around 7 mm and it includes two separate

o chambers in 1 mm height for O2 and H2 gases. These two PDMS parts are partially cured at 70 C for

25 minutes. The top part is peeled off from the mold and located on the bottom part including the membrane. The integrated PDMS chamber is cured at 70°C for 2 hours to complete the curing process.

After peeling off from the mold, first the thin layer under the membrane is removed and inlet and outlet holes are formed by a biopsy punch.

More details regarding the Pt black electrode preparation is described previous chapter. Briefly, thin film electrodes obtained by electron-beam deposition of 10 nm Ti adhesion layer, followed by 100 nm Au layer onto the glass slide (40mmx40mmx1mm). Electrodes are then electrochemically platinized at 0.5 V (Ag/AgCl) for 30 s in a 3 wt. % H2PtCl6 ·6H2O (Aldrich) + 0.03% Pb-acetate

(Aldrich) solution in water to yield porous Pt electrodes. The electrode characterization is done in 0.5M

H2SO4 solution in a conventional three-compartment electrochemical cell (Model 660D, CH

Instrument), using a Pt-wire counter electrode and a Ag/AgCl reference electrode.

The glass slide and PDMS reservoir are cleaned with a UV light treatment for 5 minutes [3]. Both surfaces are brought into contact immediately to chemically bond them and the integrated water electrolyzer is seen in Figure 4.4.

4.3.2. Electrochemical Testing

o All experiments were done after electrolysis in 1 M H2SO4 at 60 C and under applied voltage of

1.8V for 2 hours. Figure 4.5 depicts the electrolyzer experimental setup. In the experiments, a polyimide film insulated flexible heater (Omega Engineering) was attached on the glass substrate and a constant temperature was applied on the water electrolyzer through the glass side. 1M H2SO4 at room

63 temperature was sent to the chamber from the reservoir. Total voltages of 1.8V and 2V were applied to the electrodes continuously. After that all the electrolyte and evaluated gases were evacuated from the chamber through two outlets by a syringe pump (Harvard PHD-2000).

4.4. Results and Discussion

Current water electrolyzers are suffering from the coverage of all electrode surfaces with bubbles causing decreased conductivity and increased ohmic drop [21]. Additionally, the low pressure and high current density trigger off larger diameter gas bubbles formation causing a chaotic system [22]. A number of efforts on the bubble management has been investigated such as adding additives to reduce the surface tension of the electrolyte so that the bubble can leave the system easily, modification of the electrode surface geometries and properties to be less attractive to the gas bubbles, and employing mechanical circulation of the electrolyte [23]. The nature of microfluidics system has also bubble problem as discussed previously so the gas management is vital in the micro water electrolyzers.

The flow rate plays a significant role in the utilization of heat from PV cells and prevents bubble clogging in the membraneless electrolyzer chamber. We tested the electrolyzer at temperature of 25oC and 50oC. As can be seen in Figure 4.6 when the flow rate is low, bubble is stuck in the chamber and if we increase the flow rate at that time we do not give a chance to increase in the temperature of the electrolyte.

Figure 4.7 represents the hydrogen production currents for the potential of 1.8V and 2V for the membrane based electrolyzer. In the first section, the temperature is at 25oC and the flow rate is 0.5 mL/min. In the second section, the temperature is elevated to 60oC and current increases because heat transfer is provided between the heat source and electrolyte. When the flow rate is raised to 5 mL keeping the temperature same in the last section, the current goes back to the value in the first section.

These data prove that at higher flow rates, utilization of heat cannot be achieved.

64

Figure 4.8 depicts the hydrogen production at different temperatures and voltages. In all cases, it can be easily seen that the production is consistent with fluctuations related with the growth and the departure of gas bubble on the electrode. Fluctuations are more significant at high temperatures due to the formation of bubbles with large diameters. The current flux reaches to 10 mA/cm2 for the temperature of 60oC at 1.8 V and to 22 mA/cm2 for the temperature of 60oC at 2 V. When we compare the results at the potential of 1.8 with the results of membraneless electrolyzer (Figure 4.6) we showed that hydrophobic membrane aids the gas bubbles to departure from the chamber as the current is stable and it is higher than the membraneless design. de Jong et al. [24] indicated that mass transport control can be provided with gas permeable membrane in microfluidic systems and our results also explains that mass transfer limitations can be minimized by employing hydrophobic membrane in the electrolyzer. Table 4.1 compares the total hydrogen production and comparison with respect to temperature of 25°C. For a potential of 1.8V, the hydrogen production is 1.20 mL/hr at 25oC and 3.41 mL/hr at 60oC. At elevated temperature (60°C) the hydrogen production is 2.84 times higher in comparison to the room temperature. When we apply 2 V, at the temperature of 60oC the hydrogen production is 2.82 times of the temperature of 25oC.

4.5. Conclusion

We summarize the fabrication and the characterization of the membrane based water electrolyzer.

The use of the hydrophobic membrane enables the collect the gas bubbles on top the electrodes leading to operate at low flow rates with an increase in the temperature of electrolyte. We assist the multilayer soft lithography fabrication technique in order to insert the hydrophobic membrane between two

PDMS layers. The partial curing temperature and time are found as 60°C and 25 minutes for our device, respectively.

65

When the electrolyte temperature is 60°C, the hydrogen production current reaches up to 22 mA/cm2 for the potential of 2V. It means we can achieve 2.82 times faster kinetics with respect the temperature of 25°C. In addition, for the potential of 1.8 V at 60°C, the achievement in the hydrogen production is 2.84 times higher of the temperature of 25°C. The increase and the consistency in the hydrogen production rate affirms that the hydrophobic membrane flushes the electrode surface from generated gas bubbles and utilizes waste heat from PV cells.

4.6. References

1. Nowotny, J. and T.N. Veziroglu, Impact of hydrogen on the environment. International Journal of Hydrogen Energy, 2011. 36(20): p. 13218-13224.

2. Tyagi, V.V., et al., Progress in solar PV technology: Research and achievement. Renewable & Sustainable Energy Reviews, 2013. 20: p. 443-461.

3. Mitrovski, S.M., L.C.C. Elliott, and R.G. Nuzzo, Microfluidic devices for energy conversion: Planar integration and performance of a passive, fully immersed H2-O2 fuel cell. Langmuir, 2004. 20(17): p. 6974-6976.

4. Choban, E.R., et al., Membraneless laminar flow-based micro fuel cells operating in alkaline, acidic, and acidic/alkaline media. Electrochimica Acta, 2005. 50(27): p. 5390-5398.

5. Jayashree, R.S., et al., Air-breathing laminar flow-based microfluidic fuel cell. Journal of the American Chemical Society, 2005. 127(48): p. 16758-16759.

6. Togo, M., et al., Structural studies of enzyme-based microfluidic biofuel cells. Journal of Power Sources, 2008. 178(1): p. 53-58.

7. Ferrigno, R., et al., Membraneless vanadium redox fuel cell using laminar flow. Journal of the American Chemical Society, 2002. 124(44): p. 12930-12931.

8. Kjeang, E., et al., A microfluidic fuel cell with flow-through porous electrodes. Journal of the American Chemical Society, 2008. 130(12): p. 4000-4006.

9. Gunther, A., et al., Transport and reaction in microscale segmented gas-liquid flow. Lab on a Chip, 2004. 4(4): p. 278-286.

10. Xu, J., R. Vaillant, and D. Attinger, Use of a porous membrane for gas bubble removal in microfluidic channels: physical mechanisms and design criteria. Microfluidics and Nanofluidics, 2010. 9(4-5): p. 765-772.

66

11. Jensen, M.J., G. Goranovic, and H. Bruus, The clogging pressure of bubbles in hydrophilic microchannel contractions. Journal of Micromechanics and Microengineering, 2004. 14(7): p. 876- 883.

12. Hibara, A., et al., Surface modification method of microchannels for gas-liquid two-phase flow in microchips. Analytical Chemistry, 2005. 77(3): p. 943-947.

13. Yang, Z., S. Matsumoto, and R. Maeda, A prototype of ultrasonic micro-degassing device for portable dialysis system. Sensors and Actuators a-Physical, 2002. 95(2-3): p. 274-280.

14. Schonburg, M., et al., Significant reduction of air microbubbles with the dynamic bubble trap during cardiopulmonary bypass. Perfusion-Uk, 2001. 16(1): p. 19-25.

15. Skelley, A.M. and J. Voldman, An active bubble trap and debubbler for microfluidic systems. Lab on a Chip, 2008. 8(10): p. 1733-1737.

16. Sung, J.H. and M.L. Shuler, Prevention of air bubble formation in a microfluidic perfusion cell culture system using a microscale bubble trap. Biomedical Microdevices, 2009. 11(4): p. 731-738.

17. Meng, D.D., et al., A methanol-tolerant gas-venting microchannel for a microdirect methanol fuel cell. Journal of Microelectromechanical Systems, 2007. 16(6): p. 1403-1410.

18. David, M.P., et al., Hydraulic and thermal characteristics of a vapor venting two-phase microchannel heat exchanger. International Journal of Heat and Mass Transfer, 2011. 54(25-26): p. 5504-5516.

19. Winther-Jensen, O., et al., Towards hydrogen production using a breathable electrode structure to directly separate gases in the water splitting reaction. International Journal of Hydrogen Energy, 2012. 37(10): p. 8185-8189.

20. Unger, M.A., et al., Monolithic microfabricated valves and pumps by multilayer soft lithography. Science, 2000. 288(5463): p. 113-116.

21. Qian, K., Z.D. Chen, and J.J.J. Chen, Bubble coverage and bubble resistance using cells with horizontal electrode. Journal of Applied Electrochemistry, 1998. 28(10): p. 1141-1145.

22. Caboussat, A., et al., LARGE GAS BUBBLES UNDER THE ANODES OF ALUMINUM ELECTROLYSIS CELLS. Light Metals 2011, 2011: p. 581-586.

23. Zeng, K. and D.K. Zhang, Recent progress in alkaline water electrolysis for hydrogen production and applications. Progress in Energy and Combustion Science, 2010. 36(3): p. 307-326.

24. de Jong, J., R.G.H. Lammertink, and M. Wessling, Membranes and microfluidics: a review. Lab on a Chip, 2006. 6(9): p. 1125-1139.

67

4.7. Figures and Tables

Figure 4.1. Working mechanism and design of the membrane based water electrolyzer. The generated gases diffuse through the hydrophobic porous membrane.

Figure 4.2. Bubble extraction working principle: an air–liquid meniscus is pinned at the entrance of the pore and the surface tension holds the pressure difference across the meniscus and prevents water from leaking through the pore

68

Figure 4.3. Fabrication scheme of the micro-water electrolyzer (a)–(b) layers formation with partially curing, (c) integration of layers with fully curing, (d) removing the thin PDMS layer under the membrane.

Figure 4.4. The fabricated water electrolyzer cell. 40 mm x 40 mm glass slide, 1 cm2 electrode area (each), 7 mm thick PDMS chamber.

69

Figure 4.5. The experimental setup of the system for water electrolyzer. The flow rate is controlled with the syringe pump and the temperature is set with flexible heater with PID system.

70

6 25°C "5 mL/min" 5

2 "1 mL/min" 4 "0.65 mL/min" "0.5 mL/min" 3

2 current mA/cm current 1

0 0 10 20 30 40 50 60 time (sec)

6

50°C "5 mL/min" 5

2 "1 mL/min" 4 "0.65 mL/min"

3 "0.5 mL/min"

2 current mA/cm current 1

0 0 10 20 30 40 50 60 time (sec)

Figure 4.6. For the membraneless design electrolyzer; the effect of flow rate on the hydrogen o o production at the temperature of 25 C and 50 C at the potential of 1.8V in the electrolyte of 1M H2SO4.

71

30 I II III 2 V

25 1.8 V 2

20

15

10 current mA/cm current 5

0 0 20 40 60 80 100 120 140 160 180 time (sec)

Figure 4.7. For the membrane based electrolyzer; the hydrogen production current during the time: (I) at 25oC and the flow rate is 0.5 mL/min, (II) at 60oC and the flow rate is 0.5 mL/min, (III) at 60oC and the flow rate is 5 mL/min.

72

15 1.8 V 60°C 50°C 40°C 25°C

12

2 cm 9

6

current mA/ current 3

0 0 10 20 30 40 50 60 time (sec)

30 2 V 60°C 50°C 40°C 25°C

25 2 20

15

10 current mA/cm current 5

0 0 10 20 30 40 50 60 time (sec)

Figure 4.8. For the membrane based electrolyzer; the hydrogen production current during the time at different temperatures and potentials

73

Table 4.1. Hydrogen Production rate and the ratio with respect to respect to 25oC. Potential Temperature Hydrogen Production the ratio wrt 25oC (V) (oC) (mL/hr) 25 1.20 1.00 40 2.03 1.70 1.8 50 2.74 2.28 60 3.41 2.84 25 2.43 1.00 40 3.96 1.63 2 50 5.49 2.26 60 6.85 2.82

74

CHAPTER 5

SPECIES SEPARATION THROUGH SELF-ASSEMBLED MONOLAYERS MODIFIED NANOPOROUS MEMBRANES

Ionic and Molecular Transport through Nanoporous Membrane

Nanoporous membranes have received great attention in the fields of water desalination, biosensing, and chemical separations [1-17]. Bare nanopores can be used as size-selective filters but if the surface chemistry of a nanopore is modified by coating it with another substance, however, enhanced separations based other properties can be achieved. In the work by Savariar, they modified the polymeric membrane surface with SnCl2 leads to separation based not only size but also charge and hydrophobicity (Figure 5.1.a) [18]. In another work, metallized silicon nitride nanopores chemically modified with a type of acetic acid receptors can be used for the stochastic sensing of proteins (Figure 5.1.b) [19].

Many studies have been performed on ion permselectivity across gold-coated charged surfaces and charged nanopores. Many authors have used the electroless gold-coated PCTE membrane to study different mechanisms of ionic transport based on the influence of charge, hydrophobicity, steric effects, and pH [20-32]. Nguyen et al. fabricated ion-tracked polyethylene terephthalate (PET) membranes, which had a surface and pore-wall carboxylate terminal group [3]. The membranes selectively transported cations and prevented diffusion of anions. They were able to reverse the selective properties of the membrane by making an amino terminal group on the surface and pore wall of the membrane.

Other studies have included the study of mass transport of molecules based on the hydrophobicity of

75 the pore, which was made conductive through the electroless gold deposition process. The gold-coated nanotubules were modified with a fluorinated thiol [20]. Zenglian et al. used an electroless gold- deposited PCTE membrane with a modified thiol surface that was hydrophobic to study mass transport experiments of two molecules and the effects of adding alkyl surfactants to change the hydrophobic properties of the membrane surface [21]. Hou and coworkers used an electroless gold deposited PCTE membrane that was modified with carboxylic-acid functionalized thiols onto gold to study the pH dependency of mass transport of molecules through the membrane [22]. In addition to separation based on hydrophobicity and electrostatic interactions, it has been demonstrated that analytes can be excluded based on size. Martin and coworkers were able to use electroless gold-coated nano-capillary-array membrane (NCAMs) to exclude ions by molecular sieving [22-24]. There have been numerous studies for separation of biomolecules and analytes based on charge selectivity [25-30]. It has been demonstrated that proteins can be separated by applying potentials across conductive alumina membranes [31]. Cheow et al. developed a coated nanoporous alumina membrane to demonstrate the permselectivity of proteins by applying different potentials across the membrane [32].

All these studies indicate that surface modification is crucial in order to enhance separation methods with nanoporous membranes.

Specific Ion Adsorption and Properties of well-ordered Self-Assembled

Monolayers

Self-Assembled Monolayers (SAMs) provide a unique way to control and functionalize the interfacial properties at a solid liquid/interface. There have been numerous studies on the quality of organic SAMs as a blocking mechanism for prevention of ion adsorption [33-38], with applications ranging from biosensors to nanoscale transport. The applications of SAMs are often limited by the immersion time allocated for the growth of the SAM. Past work has focused on studying the properties

76 of a well-ordered SAM by electrochemical impedance spectroscopy (EIS) [39-41]. It was found that a minimum of 48 hours is needed to grow a well-ordered SAM [39, 41].

If specific adsorption of a molecule occurs on the surface of a membrane, ion selectivity for separation will be inhibited by the loss of interaction between the membrane surface and analyte as shown in Figure 5.2.a. If the membrane surface is well covered by the SAM in order to minimize ion adsorption, ion selectivity can be enhanced for a specific molecule of interest as can be seen in Figure

5.2.b.

The interaction between the protein and nanopore wall can be quantified by the time constant associated with adsorption and desorption events. Without the presence of a SAM on the surface, non- specific protein adsorption can occur on the walls of the nanopore. This can lead to effective blockage within the nanopore with non-targeted proteins, leading to a smaller measurable current (lower S/N ratio) across the nanopore [42]. Moreover, redox currents typically increase the background noise for electrochemical sensors. If detection can be accomplished with lower applied potentials, S/N ratios can be improved for electrochemical sensing of specific analytes [43].

Studies on charge selectivity have been done on electroless gold coated polycarbontate track etched membranes (PCTE) [44]. An alkanethiol was used to prevent anion (e.g. chloride ions) adsorption to the gold surface. The membrane was immersed in 1 mM thiol solution for 24 hours. However, it was observed that an irreversibility of charge selectivity occurred due to the poor quality of the monolayer over the gold-coated membrane surface. In addition to prevention of ion adsorption, the diffuse layer potential at the monolayer surface can act as a screening mechanism for nanoscale transport applications [45]. The model of a SAM-covered electrode-electrolyte interface include the following features: a potential of the metal (φm), a potential drop across the monolayer (assumed to be linear), and a potential at the surface of the monolayer (φs); and, a potential drop in the diffuse layer, all relative to some remote point in the solution (φref) [36]. For a smooth surface, the zeta potential should be close

77 to, if not coincident with the diffusive double-layer potential [2]. As the ionic strength is increased, the diffuse-layer potential becomes a linear function of the applied potential. A calculation of diffuse layer potential for HO(CH2)14SH monolayer coated Au electrode in 1 mM electrolyte resulted in a range of roughly ±40 mV results from ±300 mV of applied potential [36]. Additional studies have shown the surface charge density of alkane monolayers to be dependent on the pH of the electrolyte solution [46,

47]. The diffuse layer potential decreases for the same applied potential as the ionic strength is increased, due to the compaction of the diffuse layer.

Electrochemical Impedance Spectroscopy

EIS is used to determine the resistive and capacitive nature of a system over a set of frequencies.

EIS measurements are done by applying an AC potential (with known amplitude) over a range of frequencies and measuring the current response of the system. By measuring the current of the electrochemical cell, the overall impedance may be determined by dividing the phasor voltage by the phasor current. The impedance may be represented as a real and imaginary part as shown in Equation

5.1 [48].

퐸푡 퐸표푠푖푛⁡(푤푡) 푠푖푛⁡(푤푡) 푍 = = = 푍표 (5.1) 퐼푡 퐼표푠푖푛⁡(푤푡 + ∅) 푠푖푛⁡(푤푡 + ∅)

The real and imaginary part describes the ohmic and capacitive nature of the system respectively.

Figure 5.3.a-c demonstrates a Bode plot, Nyquist plot, and an equivalent linear circuit model used to fit the mechanistic behavior of the electrochemical system being studies. The Bode plot demonstrates the behavior of an electrochemical system over a range of frequencies. At low frequencies, the two plots exhibit that the system’s impedance is real as a result of ion conduction through a SAM or redox reactions at the working electrode. At higher frequencies, the phase of the system approaches a value of - 90°, displaying the capacitive component of the system. This capacitive nature of the system is a

78 result of the double layer charging at the surface of the electrochemical system. Figure 5.3.c is a

Randles circuit, which models an electrochemical system in which the solution resistance is in series with an RC circuit. The model is used when a SAM is considered to be a “nonperfect” insulator to ion conduction [48].

EIS has been used to characterize ionic transport through nanoporous membranes [49]. EIS has been applied to the detailed analysis of through-film and at-defect electron transport paths in thiol- modified gold electrode in a solution containing a diffusing redox species [34, 37]. Several studies used EIS to show that a defect-free monolayer obeys the Helmholtz ideal capacitor exhibiting a phase shift ≥ 80° [50-56]. Further, the impedance spectra can be fitted to an equivalent circuit of a solution resistance Rs in series with a constant phase element (CPE), which accounts for double-layer capacitance and SAM capacitance. Equation 5.2 is impedance of the CPE model expressed as

1 푍퐶푃퐸 = ⁡ 푎 (5.2) 푌휊(푗푤 )

where 푌휊 is the capacitance (F), ω is the frequency of the applied electric field, and α is the constant which represents the ideality of the capacitor. The constant-phase element (CPE) is analogous to a distributed capacitance, with α being a measure of variation of the dielectric coating thickness, since we assume that the monolayer formation has no effect on gold electrode surface roughness [57].

Dissertation Overview

The second part of dissertation will discuss the topics around the ionic transport across the nanoporous membrane under applied potential and investigation of surface coverage of SAMs on the

Au surface in order to increase reversibilities in charge selectivity.

In Chapter 6, we investigated the electrochemical interfacial properties of a well-ordered SAM of

1-undecanethiol (UDT) on evaporated gold surface by EIS in electrolytes without a redox couple.

79

Using a constant phase element (CPE) series resistance model, prolonged incubation times (up to 120 h) show decreasing monolayer capacitance approaching the theoretical value for 1-undecanethiol. The

EIS data shows that UDT (methylene chain length n=10), incubated for 120 h, forms a monolayer whose critical voltage range extends from -0.3 to 0.5 V vs. Ag/AgCl, previously attained only for alkanethiol at n=15. At low frequencies where ion diffusion occurs, almost pure capacitive phase (-

89°) were attained with lengthy incubation.

In Chapter 7, we fabricated a membrane permeate flow cell is described with the aim of studying the transport of methyl viologen (paraquat, MV2+) and napathalenedisulfonate disodium salt (NDS2-), using a conductive NCAM. A polycarbonate track etched (PCTE) membrane was made conductive by sputter coating gold on the membrane surface. Transport studies were done in a voltage range in which faradaic current was minimized at the surface of the gold-coated NCAMs. The goal of the transport studies is to demonstrate improved charge selectivity when a well-grown 1-undecanethiol monolayer is assembled at the surface of the NCAM for a wide range of applied potentials (-400 mV < Vappl < 400 mV). Results show the selectivity of charged analytes through the metallized NCAM can be improved by functionalizing the surface with a self-assembled monolayer (SAM). The selectivity coefficients for

MV2+ and NDS2- increased with functionalization of undecanethiol on the gold-coated NCAM surface.

References

1. Shannon, M.A., et al., Science and technology for water purification in the coming decades. Nature, 2008. 452(7185): p. 301-10.

2. Holtzel, A. and U. Tallarek, Ionic conductance of nanopores in microscale analysis systems: where microfluidics meets nanofluidics. Journal of separation science, 2007. 30(10): p. 1398-419.

3. Nguyen, Q.H., et al., Charge-selective transport of organic and protein analytes through synthetic nanochannels. Nanotechnology, 2010. 21(36): p. 365701.

4. Yuan, H., et al., Characterization of the electrophoretic mobility of gold nanoparticles with different sizes using the nanoporous alumina membrane system. Sensors and Actuators B 2008. 134: p. 127-132.

80

5. Kim, D.H., A review of desalting process techniques and economic analysis of the recovery of salts from retentates. Desalination, 2011. 270(1-3): p. 1-8.

6. Eijkel, J.C. and A. van den Berg, Nanofluidics and the chemical potential applied to solvent and solute transport. Chemical Society reviews, 2010. 39(3): p. 957-73.

7. Dermentzis, K. and K. Ouzounis, Continuous capacitive deionization–electrodialysis reversal through electrostatic shielding for desalination and deionization of water. Electrochimica Acta, 2008. 53(24): p. 7123-7130.

8. Lindstrand, V., G. Sundstrom, and A.S. Jonsson, Fouling of electrodialysis membranes by organic substances. Desalination, 2000. 128: p. 91-102.

9. Pilat, B., Practice of water desalination by electrodialysis. Desalination, 2001. 139: p. 385-392.

10. Sata, T., Modification of properties of ion exhcange Colloid and Polymer Science, 1978. 256: p. 62-77.

11. Shaposhnik, V.A. and K. Kesore, An early history of electrodialysis with permselective membranes. Journal of Membrane Science, 1997. 136: p. 35-39.

12. Tanaka, Y., Current density distributin and limiting current density in ion-exchange membrane electrodialysis. Journal of Membrane Science, 2000. 173(179-190).

13. Xu, T. and C. Huang, Electrodialysis-based separation technologies: A critical review. AIChE Journal, 2008. 54(12): p. 3147-3159.

14. Chen, X.J., et al., Detection of the Superoxide Radical Anion Using Various Alkanethiol Monolayers and Immobilized Cytochrome c. Anal. Chem, 2008. 80: p. 9622-9629.

15. Davis, K.L., et al., Electron-Transfer Kinetics of Covalently Atached Cytochrome c/SaAM/Au Electrode Assemblies. J. Phys. Chem. C, 2008(112): p. 6571-6576.

16. de Groot, M.T., et al., Electron Transfer and Lingand Binding to Cytochrome c' Immobolized on Self-Assembled Monolayers. Langmuir, 2007. 23: p. 729-736.

17. Prakash, S., M.B. Karacor, and S. Banerjee, Surface modification in microsystems and nanosystems. Surface Science Reports, 2009. 64(7): p. 233-254.

18. Savariar, E.N., K. Krishnamoorthy, and S. Thayumanavan, Molecular discrimination inside polymer nanotubules. Nature nanotechnology, 2008. 3(2): p. 112-7.

19. Wei, R.S., et al., Stochastic sensing of proteins with receptor-modified solid-state nanopores. Nature Nanotechnology, 2012. 7(4): p. 257-263.

20. Velleman, L., J.G. Shapter, and D. Losic, Gold nanotube membranes functionalised with fluorinated thiols for selective molecular transport. Journal of Membrane Science, 2009. 328(1-2): p. 121-126.

21. Yue, Z., et al., Study of transportation of atrazine and paraquat through nanochannels. Journal of Membrane Science, 2010. 356(1-2): p. 117-122.

81

22. Hou, Z., N. Abbott L., and P. Stroeve, Self- Assembled Monolayers on Electroless Gold Impart pH- Responsive Transport of Ions in Porous Membranes. Langmuir, 2000. 16: p. 2401-2404.

23. Kang, M.S. and C.R. Martin, Investigations of Potential-Dependent Fluxes of Ionic Permeates in Gold Nanotubule Membranes Prepared via the Template Method. Langmuir, 2001. 17: p. 2753- 2759.

24. Jirage, K.B., Nanotubule-Based Molecular-Filtration Membranes. Science, 1997. 278(5338): p. 655-658.

25. Han, J., J. Fu, and R.B. Schoch, Molecular sieving using nanofilters: past, present and future. Lab on a chip, 2008. 8(1): p. 23-33.

26. Peng, X., et al., Ultrafast permeation of water through protein-based membranes. Nature nanotechnology, 2009. 4(6): p. 353-7.

27. Cannon, D.M., et al., Nanocapillary Array Interconnects for Gated Analyte Injections and Electrophretic Separations in Multilayer Microfluidic Architectures. Anal. Chem, 2003. 75: p. 2224-2230.

28. Kler, P.A., et al., High performance simulations of electrokinetic flow and transport in microfluidic chips. Computer Methods in Applied Mechanics and Engineering, 2009. 198(30-32): p. 2360-2367.

29. Plecis, A., R.B. Schoch, and P. Renaud, Ionic Transport Phenomnea in Nanofluidics: Experimental and Theoretical Study of the Exclusion-Enrichment Effect on a Chip. Nano letters, 2005. 5(6): p. 1147-1155.

30. Schoch, R., J. Han, and P. Renaud, Transport phenomena in nanofluidics. Reviews of Modern Physics, 2008. 80(3): p. 839-883.

31. Nguyen, B.T., E.Z. Ting, and C.S. Toh, Development of a biomimetic nanoporous membrane for the selective transport of charged proteins. Bioinspiration & biomimetics, 2008. 3(3): p. 035008.

32. Cheow, P., et al., Transport and separation of proteins across platinum-coated nanoporous alumina membranes. Electrochimica Acta, 2008. 53(14): p. 4669-4673.

33. Gupta, C., M.A. Shannon, and P.J.A. Kenis, Electronic Properties of a Monolayer-Electrolyte Interface Obtained from Mechanistic Impedance Analysis. J. Phys. Chem C, 2009. 113(21): p. 9375-9391.

34. Gupta, C., M.A. Shannon, and P.J.A. Kenis, Mechanisms of Charge Transport through Monolayer- Modified Polycrystalline Gold Electrodes in the Absence of Redox-Active Moieties. J. Phys. Chem C, 2009. 113(11): p. 4687-4705.

35. Walczak, M.M., et al., Reductive desorption of alkanethiolate monolayers at gold: a measure of surface coverage. Surface Chemistry and Colloids, 1991. 7(11): p. 2687-2693.

36. Becka, A.M. and C.J. Miller, Electrochemistry at w-hydroxy thiol coated electrodes. 4. Comparison of the double layer at w-hydroxy thiol and alkanethiol monolayer coated Au electrodes. Section Title: Electrochemistry, 1993. 97(23): p. 6233-6239.

82

37. Nahir, T.M. and E.F. Bowden, Impedance spectroscopy of electro-inactive thiolate films adsorbed on gold. Electrochemistry, 1994. 39(16): p. 2347-2352.

38. Finklea, H.O., Electrochemistry of Organized Monolayers of Thiols and Related Molecules on Electrodes. 1996, Dekker: New York. p. 109-335.

39. Agonafer, D.D., Chainani, E., Oruc, M.E., Lee, K.S., Shannon, M.A., Study of Insulating Properties of Alkanethiol Self-Assembled Monolayers Formed Under Prolonged Incubation Using Electrochemcial Impedance Spectroscopy. Journal of Nanotechnology in Engineering and Medicine, 2012. 3(3): p. 1-8.

40. Diao, P., M. Guo, and R. Tong, Characterization of defects in the formation process of self- assembled thiol monolayers by electrochemical impedance spectroscopy. Journal of Electroanalytical Chemistry, 2001. 495: p. 98-105.

41. Boubour, E. and R.B. Lennox, Stability of Omega-Functionalized Self-Assembled Monolayers as a Function of Applied Potential. Langmuir, 2000. 16(19): p. 7464-7470.

42. Sexton, L.T., Horne, L.P., Sherrill, S.A., Bishop, G.W., Baker, L.A., Martin, C.R., Resistive Pulse Studies of Proteins and Protein Antibody Complexes Using a Conical Nanotube Sensor. JACS, 2007. 129: p. 13144-13152.

43. Wilson, R.C.K., et al., Flexible, micron-scaled superoxide sensor for in vivo applications. Journal of Electroanalytical Chemistry, 2011. 662(1): p. 100-104.

44. Kang, M. and C.R. Martin, Investigations of Potential-Dependent Fluxes of Ionic Permeates in Gold Nanotubule Membranes Prepared via the Template Method. Langmuir, 2001. 17: p. 2753- 2759.

45. Nguyen, B.T. and C. Toh, Development of an electrode-membrane system for selective Faradaic response towards charges redox species. Electrochemica Acta, 2009. 54: p. 5060-5064.

46. Tian, C.S. and Y.R. Shen, Structure and charging of hydrophobic material/water interfaces studied by phase-sensitive sum-frequency vibrational spectroscopy. Proceedings of the National Academy of Sciences of the United States of America, 2009. 106(36): p. 15148-53.

47. Wu, Y., et al., Dynamic response of AFM cantilevers to dissimilar functionalized silica surfaces in aqueous electrolyte solutions. Langmuir : the ACS journal of surfaces and colloids, 2010. 26(22): p. 16963-72.

48. Agonafer, D.D., Surface Characterization of Self-assembled Monolayers For Applications of Selective Ionic Transport Through Nanoporous Membranes 2012.

49. Vitarelli, J.M., Jr., S. Prakash, and S.D. Talaga, Determining Nanocapillary Geometry From Electrochemical Impedance Spectroscopy Using Variable Topology Network Circuit Model. Anal. Chem., 2011. 83: p. 533-541.

50. Boubour, E. and R.B. Lennox, Potential-Induced Defects in n-Alkanethiol Self-Assembled Monolayers Monitored by Impedance Spectroscopy. J. Phys. Chem. B, 2000. 104(38): p. 9004- 9010.

83

51. Janek, R.P. and W.R. Fawcett, Impedance Spectroscopy of Self-Assembled Monolayers on Au(111): Evidence for Complex Double-Layer Structure in Aqueous NaClO4 at the Potential of Zero Charge. J. Phys. Chem. B, 1997. 101: p. 8550-8558.

52. Love J., C., Estroff A., Lara, Kriebel K., Jennah, Nuzz G., Ralph, Whitesides M., George, Self- Assembled Monolayers of Thiolates on Metals as a Form of Nanotechnology. Chem Rev, 2005. 105: p. 1103-1169.

53. L.V., P. and F. W.R., Electrochemical Impedance Spectroscopy at Alkanethiol-Coated Gold in Propylene Carbonate. Langmuir, 2002. 18: p. 8933-8941.

54. Wang, B., et al., Dielectric Properties and Frequency Response of Self-Assembled Monolayers of Alkanethiols. Langmuir, 2004. 20: p. 5007-5012.

55. Campuzano, S., et al., Characterization of alkanethiol-self-assembled monolayers-modified gold electrodes by electrochemical impedance spectroscopy. Journal of Electroanalytical Chemistry, 2006. 586(1): p. 112-121.

56. Darwish, N., et al., Reversible potential-induced structural changes of alkanethiol monolayers on gold surfaces. Electrochemistry Communications, 2011. 13(5): p. 387-390.

57. Sahalov, H., et al., Influence of Applied Potential on the Impedance of Alkanethiol SAMs. Langmuir, 2007. 23(9681-9685).

84

Figures

(a) (b)

Figure 5.1. Protein separation based on charge and hydrophobicity after modification of membrane surface with SnCl2 (left) [18] and stochastic sensing of proteins after modified with a type of acetic acid receptors (right) [19].

Figure 5.2. Schematics (a) and (b) of a gold coated NCAM, which illustrates the improvement when a well-grown SAM is attached to the monolayer.

85

Figure 5.3. The plots above are examples of common EIS plots and circuit models used for analyzing EIS data. Figure (a) is a Bode plot, which displays the magnitude of the impedance and phase of the system. Figure (b) is a Randles circuit model, which is used to fit EIS data when the system contains exhibits a resistive behavior either from ion conduction through a monolayer or redox reactions at the surface of the working electrode. Figure (c) is an example of a Randles circuit, which can be used to linearly fit EIS data [48].

86

CHAPTER 6

INSULATING PROPERTIES OF ALKANETHIOL SAMS FORMED UNDER PROLONGED INCUBATION USING ELECTROCHEMICAL IMPEDANCE SPECTROSCOPY

The text and figures in this chapter are reproduced with permission from the work: Damena D. Agonafer, Edward Chainani, Muhammed E. Oruc, Ki Sung Lee, and Mark A. Shannon. "Study of Insulating Properties of Alkanethiol Self-Assembled Monolayers Formed Under Prolonged Incubation Using Electrochemical Impedance Spectroscopy." Journal of Nanotechnology in Engineering and Medicine 3, no. 3 (2012): 031006-031006.

Introduction

Interfacial transport at the micro and nanoscale plays a crucial role due to high surface area to volume ratio of micro and nanosytems [1]. SAMs provide a unique way to control and functionalize the interfacial properties at a solid liquid/interface. There have been numerous studies on the quality of organic self-assembled monolayers (SAMs) as a blocking mechanism for prevention of ion adsorption [2-7], with applications ranging from biosensors to nanoscale transport. Studies on charge selectivity have been done on electroless gold coated polycarbontate track etched membranes (PCTE)

[8]. An alkanethiol was used to prevent anion (e.g. chloride ions) adsorption to the gold surface. The membrane was immersed in 1 mM thiol solution for 24 hours. However, it was observed that an irreversibility of charge selectivity occurred due to the poor quality of the monolayer over the gold- coated membrane surface. In addition to prevention of ion adsorption, the diffuse layer potential at the monolayer surface can act as a screening mechanism for nanoscale transport applications [9]. The model of a SAM-covered electrode-electrolyte interface include the following features: a potential of

87 the metal (φm), a potential drop across the monolayer (assumed to be linear), and a potential at the surface of the monolayer (φs); and, a potential drop in the diffuse layer, all relative to some remote point in the solution (φref) [5]. For a smooth surface, the zeta potential should be close to, if not coincident with the diffusive double-layer potential [10]. As the ionic strength is increased, the diffuse- layer potential becomes a linear function of the applied potential. A calculation of diffuse layer potential for HO(CH2)14SH monolayer coated Au electrode in 1 mM electrolyte resulted in a range of roughly ±40 mV results from ±300 mV of applied potential [5]. Additional studies have shown the surface charge density of alkane monolayers to be dependent on the pH of the electrolyte solution [11,

12]. The diffuse layer potential decreases for the same applied potential as the ionic strength is increased due to the compaction of the diffuse layer.

The applications of SAMs are often limited by the quality of the monolayer. In the biosensor arena,

SAMs on a metal electrode allow the direct electron transfer of an enzyme immobilized on the monolayer, facilitating the investigation of enzyme kinetics using electrochemical techniques [13-18].

Ideally, the catalytic current through the enzyme is detected as increased tunneling current through the monolayer at the enzyme’s Nernst potential, but conduction through pinholes (areas where metal is exposed to solution due to absence of monolayer) and collapsed sites may affect the outcome. Studies have been performed to characterize how the temperature of the thiol solution can affect the growth of the monolayer on a gold substrate. Using STM measurements, it has been demonstrated that the vacancy island area does indeed change over a wide range of temperatures, but the overall defect fraction area does not change [19, 20]. As the temperature is increased, the number of vacancy islands decrease, but the area of each vacancy island increases. This is a result of the Ostwald ripening process, in which larger vacancy islands grow at the expense of smaller gold islands [21, 22]. Some of the contributions that dictate how well a monolayer is grown can be determined by the roughness of the

SAM on the substrate.

88

Douglass Jr. et al. studied the roughness factor for electrodes that had dodecanethiol grown on a gold substrate [23]. It was determined that the macroroughness (≥10nm) of the substrate influenced the capacitive nature of the system, while the microroughness (1-10 nm) had a larger effect on the quality and structure of the SAM. In addition to monolayers containing defect sites (pinholes), monolayers may include collapsed sites that are present on the substrate. Studies have been done to try and distinguish the properties of SAMs when defect or collapsed sites are present. The electron rate constant for a monolayer with a collapsed sites were found to be two orders of magnitude smaller than that for a defect site [24]. This is evidence that tunneling effects occur for regions within the monolayer that contain collapsed sites. The investigation of pH on the variation of charge transfer resistance (Rct) and the apparent rate constant have been done using EIS [25]. Furthermore, EIS measurements that are done over a range of potentials, give insight to the permeability of the monolayer to penetration by solution ions. Ion penetration into the monolayer manifests as a parallel conduction path from solution to the electrode, necessitating a shift in the model from the series Constant Phase Element (CPE)- resistor to a Randles circuit composed of two resistors, one of which is in parallel with a capacitor. The positive and negative potential limits where the shifts occur are termed as the critical voltages Vc.

Studies have shown that the critical voltage range can be dependent on the length of the alkane chain

[26]. Boubour and Lennox studied the critical voltages for linear chain alkane thiols [26], where it was found that the critical voltage range was chain length dependent. For an alkanethiol of chain lengths n=7 to 15, the Vc ranged from -0.15 V—0.3 V, but Vc for chain lengths below n=15 were all less than

-0.3 V. The greater impermeability of longer chains may be ascribed to increased chain-chain interactions [27]. These attractive lateral interactions, driven by van der Waals forces, generally increase with chain length [28]. The use of EIS to study the quality of a monolayer for different immersion times has been done to explore the evolution of SAMs over different time scales. Daio et al. studied self-assembled monolayers of octadecanethiol (ODT) on gold by electrochemical

89 impedance spectroscopy (EIS) with a redox couple, in order to make a quantitative analysis of the change of the fractional coverage, pinhole size and separation as a function of incubation time [29].

The range of incubation times in that study was from 5 seconds to 24 hours. Their results showed that the defective monolayers exhibited Warburg impedances at low frequencies with a redox couple in solution, an indication that the ions were mass transport limited while diffusing through the defect sites. Monolayer formation kinetics were characterized by an initial fast step (chemisorption) followed by a slow step of surface coverage (through lateral bonding), and the size of the pinholes remain the same, but the number density decreases over time. However, because redox species were used, it is difficult to understand the insulative behavior of a monolayer when charge transfer resistance is always present due to tunneling effects. In another investigation utilizing EIS, experiments were performed without a redox couple in the electrolyte to study the quality of a monolayer for different incubation times of a gold electrode immersed in dodecanethiol (CH3(CH2)11SH) solution [30]. A minimum of 40 hours was found necessary to achieve phase shifts φ≥-88°. For an ideal insulating (capacitive) monolayer, φ=-90° is expected. The EIS measurements for the monolayers incubated at various times were done at a single applied potential of 0 V vs. Ag/AgCl. Because ion penetration is exacerbated at anodic and cathodic potentials, it is important to understand how the monolayer incubation time can affect the insulative properties of a monolayer over a range of voltages [2, 3]. Furthermore, as the applied potential approaches the PZC, a field can induce a torque on the headgroup, resulting in conformational changes in the SAM [31]. These conformational changes can lead to a lower barrier for ion penetration through the monolayer to occur. It is therefore imperative to understand the incubation time of a SAM. This work explores how incubation time on growing the monolayer can affect the critical voltage range.

90

Experimental

Electrochemical Instrumentation and Measurements

Electrochemical measurements were performed using a Gamry Reference 600 (Gamry

Instruments, Warminster, PA) employing a three-electrode cell: the monolayer-coated gold surface acted as the working electrode, an Ag/AgCl wire in 3 M NaCl (Bioanalytical Instruments, West

Lafayette, IN) as the reference, and a gold wire as the counter electrode. Gold wire counter electrodes

(Alfa Aesar, Ward Hill, MA) were used for the impedance measurements. Gold-coated microscope slides were mounted in a custom-designed cell. The gold layer was used as the working electrode. All potentials are reported with respect to the Ag/AgCl electrode. Phosphate buffer solutions (PB) were prepared as a ratio of mono- and dihydrogen potassium phosphate salts (Sigma-Aldrich, St. Louis,

MO) in DI water serves as electrolyte. The total concentration and pH of the electrolyte is reported.

The electrolyte solutions were bubbled for 15 minutes with nitrogen to remove oxygen prior to electrochemical characterization. The pH of the electrolyte solution was monitored before and after the experiments. Because each experiment was performed in less than 10 minutes, the pH of the solution did not change within the small time interval.

Preparation of Monolayer Surfaces

1-undecanethiol was purchased from Sigma Aldrich, and dissolved in absolute ethanol (Pharmaco-

Aaper, Shelbyville, KY) at a concentration of 1 mM. SPR-quality glass slides (EMF Corp., Ithaca NY) coated with evaporated titanium (5 nm) and gold (100 nm) thin films were washed in a heated SC-1 bath (100mL of DI water/25 mL of H2O2/2 mL of NH4OH) and rinsed thoroughly with DI water

(resistivity 18 MΩ-cm) and ethanol. The clean gold surfaces were incubated in a sealed Petri dish for immersion times ranging from 3 hours to 5 days. The monolayer gold coated substrates were first electrochemically polished. Electrochemical polishing was done as follows: in 0.1 M NaOH was de-

91 aerated with nitrogen gas for 15 minutes, with two cycles being required. The first was a wide and fast sweep from -1.2 to +1.5 V vs. Ag/AgCl at 1 V/s (21 segments were required to obtain a stable voltammogram). The voltage range chosen was based on previous studies, which ensure a monolayer formation of gold oxide from whose reduction current the electrochemical surface area is determined

[32, 33]. Secondly, without removing the electrode from solution, another (narrow and slow) sweep from -0.8 to 0.6 V vs. Ag/AgCl at 0.05 V/s (21 segments) was performed. Cycling through these potentials ensures removal and formation of an oxide layer, which smoothens out the surface by removing an atomic layer of gold atoms. The gold surface is subsequently rinsed with DI water and absolute ethanol. The Au substrates were then incubated in thiol solutions in the manner described above. All monolayers were grown at room temperature. The monolayer-coated gold were rinsed with absolute ethanol and DI water and blow-dried with N2 before use.

Surface Characterization

AFM measurements of bare gold substrates were made under ambient conditions using an Asylum

MFP-3D (Santa Barbara, CA). AFM probe tips used were from BudgetSensors (Sofia, Bulgaria).

Images were obtained in intermittent contact mode. Roughness was calculated using the built-in software, Igor Pro from Wavemetrics (Portland, OR). Samples were prepared by SC-1 cleaning, followed by electrochemical polishing.

Results and Discussion

Determination of Electrochemical Surface Area

The area of the gold substrate was measured by oxygen adsorption in 0.1 M NaOH, cyclic voltammetry at 50 mV s-1. The amount of charge necessary for complete surface coverage was reported to be 677 µC cm-2 for polycrystalline gold [32]. The Au substrate was SC-1 cleaned for 30 minutes with 100:20:1 of DI water, H2O2, and NH4OH. The solution was de-aerated with nitrogen

92 for 15 minutes. Two CV scans were done between -0.061 V and 1.239 V (vs. Ag/AgCl) with a scan rate of 50 mVs-1. The total charge was then calculated by integrating the oxide peak area. The electrochemical area was calculated to be 0.59 cm2.

AFM measurements

The AFM measurements provided information on surface topography of the gold substrate. The main parameter used in quantitative analysis of the surface is the rms roughness. The rms roughness was measured to be 1.7 nm for electrochemical polished substrates. The peak-to-peak values

(maximum height difference) were 14 nm shown in Figure 6.1. These values are consistent with those found in previous studies for gold coated glass slides with a titanium adhesion layer [34], which were also used as substrates for monolayer formation.

EIS and CV measurements

EIS and CV experiments were done on a 1-undecanethiol monolayer that was immersed in a 1mM thiol solution from 3 to 120 hours. A gold-coated microscope slide was used as the working electrode.

The system was purged with nitrogen for 15 minutes before experiments were started. To rule out the possibility of other redox species, several CV scans were done which showed the absence of faradaic current with scan rates from 100 mVs-1 – 1 Vs-1 and voltage ranges from -0.5 V-0.5 V with only phosphate buffer as the electrolyte solution. The ionic strength (I) of the buffer solution was I =

0.0316 M.

Self-assembled monolayers of alkanethiols on gold have been shown to strongly block electrochemical oxidation of gold as well as electron transfer with redox couples in solution. The behavior of the SAMs has been ascribed to their ability to limit access of solution-phase molecules

(water, electrolyte ions and redox molecules) to the electrode surface. The insulating abilities of thiol monolayers on gold were demonstrated in studies that used methyl-terminated alkanethiols (11, 13, 15

93 and 17 methyl groups) could suppress gold oxide anodic stripping by 3-5 orders of magnitude [4, 24].

The present study has confirmed that cyclic voltammetry of bare gold surfaces in electrolyte containing no redox species shows the peaks where gold oxidation and reduction of the oxide occurs, as well as the double layer region where there is only charging current. The CVs compare well with bulk gold and gold sputtered on glass microscope slides. Upon the adsorption of thiol monolayers on the gold surface, gold oxide formation and reduction is suppressed, as evidenced by the lack of peaks in the CV seen in Figure 6.2, which compares two CV experiments between a bare gold polycrystalline surface and of a gold substrate with a monolayer grown for several days.

Cyclic voltammetry of alkanethiol monolayer-coated electrodes in electrolyte solutions show that the charging current drops dramatically in comparison to the bare metal electrode and becomes constant with electrode potential. This constant potential is consistent with the presence of a low dielectric constant between electrode and electrolyte. Equation 6.1 is an expression for the differential capacitance Cd, which is modeled by a series combination of two capacitances, the monolayer capacitance Cm (which replaces the Helmholtz capacitance CH of the bare metal), and the diffuse layer capacitance CD [35]. Monolayer capacitance is a property of the spacer length, while the diffuse layer capacitance is a function of the electrolyte concentration and applied potential with a minimum at the potential of zero charge (Epzc), such that

1 1 1 = + (6.1) 퐶푑 퐶퐷 퐶푚

To account for the Faradaic currents, investigators postulated that electron transfer could occur at pinholes (sites exposed to electrolyte due to the absence of adsorbate) in the monolayers. Pinholes are thought to occur mainly at grain boundaries, steps and kinks in polycrystalline gold. Even then, at these sites, the gold surface may be partially protected by a collapse of the monolayer, producing a defect in the monolayer where molecules and ions can approach the electrode surface at a distance shorter than

94 the distance of the SAM. The presence of pinholes and defect sites is evidenced by departure from perfect blocking behavior. Potential dependent background currents are observed for gold-monolayer- electrolyte systems, even in the absence of redox species. Such currents are explained by ion migration through pinhole defects in the monolayer film or charge transport, through the monolayer phase [2].

Even in the absence of pinholes and defects, tunneling through the full width of the monolayer contributes to a current. However, it is not easy to assess quantitatively the fraction of the surface that is composed of pinholes or defects [7]. Electrochemical impedance spectroscopy is used to determine the resistive and capacitive nature of a system over a set of frequencies. It has been used to characterize ionic transport through nanoporous membranes [36]. EIS has been applied to the detailed analysis of through-film and at-defect electron transport paths in thiol-modified gold electrode in a solution containing a diffusing redox species [3, 6]. Several studies used EIS to show that a defect-free monolayer obeys the Helmholtz ideal capacitor exhibiting a phase shift ≥ 80° [28, 37-42]. Further, the impedance spectra can be fitted to an equivalent circuit of a solution resistance Rs in series with a constant phase element (CPE), which accounts for double-layer capacitance and SAM capacitance.

Equation 6.2 is impedance of the CPE model expressed as

1 푍퐶푃퐸 = 푎 (6.2) 푌휊(푗푤 )

where 푌휊 is the capacitance (F), ω is the frequency of the applied electric field, and α is the constant which represents the ideality of the capacitor. The constant-phase element (CPE) is analogous to a distributed capacitance, with α being a measure of variation of the dielectric coating thickness, since we assume that the monolayer formation has no effect on gold electrode surface roughness [31].

Several theoretical models have demonstrated that the roughness, or the degree of disorder of a monolayer, can be estimated by the CPE exponent α [43-47]. The proof of the presence of the SAM can be observed from Table 6.1. The capacitance of the bare gold substrate is approximately 1.4 to 60

95 times greater than the capacitances for SAMs grown from 3 hours to 120 hours. In addition, as shown in Table 6.1, the α value increases as incubation time of monolayer growth increases. An ideal capacitor has α=1, while α>0.88 suggests an adequately smooth gold surface. As mentioned above, a CPE model fits better than a simple series RC model. For alkanethiol SAMs with n=9 (1-dodecanethiol), a CPE of

1.36 to 1.67 µF/cm2 was reported over the range -0.2 to 0.4 V (vs. Ag/AgCl). [37]. Above 10-2 M

2 electrolyte concentration the diffuse layer capacitance is greater than 10 µF/cm , so Cm dominates the interfacial capacitance.

Figure 6.3 is a Bode phase plot that shows the insulative behavior of monolayers for two different immersion times. For the monolayer grown for 48 hours, the EIS data shows that the phase angle peaks between 1-10 Hz at -68 to -75 degrees for the electrolyte concentrations of 1 mM and 10 mM. The dependency is due to the contribution of the diffuse layer capacitance, which is strongly dependent on concentration. The model used here is from Gouy-Chapman theory, where the capacitance of the space- charge between the electrolyte and the electrode is at a minimum at the potential of zero-charge and rises at higher and lower potentials. At the minimum, at 25° C with a z:z electrolyte, this capacitance is

CGC (µF/cm-2) = 228 z c*1/2 (6.3)

where c* is bulk electrolyte concentration in mol L-1 [35]. For a 1:1 electrolyte at 50 mM concentration, the double-layer capacitance is calculated to be 50 µF cm-2. This relatively high capacitance is characteristic of the gold-solution interface when it is not blocked from electrolyte by a monolayer, which corresponds to pinhole areas in a SAM.

The monolayer-covered gold/solution interface forms a second kind of interfacial capacitance. This capacitance (neglecting charge on the head group) is dominated by the monolayer capacitance Cm, determined by the relative permittivity of the alkanethiol chain and the film thickness δ of the monolayer, such that

96

휖 휖 퐶 = 표 푟 (6.4) 푚 훿

For a well-ordered monolayer, characterized by a permittivity value of around 2 expected for

-2 hydrocarbon chains and thickness δ ≤ 2 nm, Cm values in the range of 1 to 1.8 µF cm are expected, with longer chain thiols having lower capacitance due to the higher thickness [48]. Since Cm<

Collapse sites where the monolayer thickness is lower than that of the well-ordered monolayer form a third type of capacitor. Since the electrolyte ions are blocked from approaching the gold surface, but can approach the electrode much closer than the full width of the monolayer, the measured CPE here would then be lower than that of a bare electrode/electrolyte but higher than that of the well- ordered monolayer. In the case of incomplete monolayer coverage, with the possibility of pinholes, these three capacitances co-exist. With increased incubation time, it is reasonable to assume that with reorganization of the monolayer results in reduction of coverage by defects. In this case we can expect a decrease in capacitance, as well as an increase in α values with incubation time implying that the monolayer thickness is becoming more uniform.

Therefore, for defect-free monolayers, the phase shift should be independent of concentration due to the monolayer capacitance exhibiting behavior similar to the Helmholtz capacitance, which is independent of electrolyte concentration. A similar gold surface incubated for 5 days exhibited a -88° phase shift at 1 Hz shown in Figure 6.3. At this frequency, the phase angle varied less than 1.2% between 1 and 10 mM for the Au substrate immersed for several days. This is evidence of a monolayer that has minimal pinholes and defects, due to the improved ordering of the monolayer afforded by prolonged incubation in thiol solution. It has been reported that the structure of SAMs do not change

97 significantly when exposed to a 1 mM solution for 12-18 hours [28]. However, evidence has shown that immersion times of 7-10 days can reduce the number of pinholes in the SAM and the conformational defects in the alkane chain to decrease [28]. Thus, for applications such as the study of electron transport by tunneling through monolayers or preparation of surfaces that need to be well insulated from solution redox species, longer immersion times for improved monolayer regularity and stability are needed. It is also important to note that an inductance was present at high frequencies above 10 kHz for the EIS experiments with 1 mM electrolyte concentration. A survey of the literature was made as to find the causes of inductance in a circuit with only capacitances and resistances. These effects have been divided into inductances that appear at low frequencies (<10 Hz) and those that appear at high frequencies (>100 KHz). For low frequencies, these have been ascribed to adsorbed intermediates [49] or relaxation of adsorbed anions on a metal surface [50]. However, these inductances appear at frequencies (1 and 0.1 Hz) much lower than what we have observed (10 kHz). At high frequencies, the usual suspects are stray inductances from wire leads. Indeed at these frequencies (>100 kHz) we do see an inductive phase shift. However, neither of the high or low frequency explanations account for the fact that the inductive phase in our data appears at 10 kHz, a frequency intermediate between low and high frequency situations above. Vanysek and Birss have suggested that this may be partly due to the potentiostat itself (capacitance at the reference input in conjunction with high capacitance and low resistance at the working electrode) at frequencies greater than about 10 kHz [51].

Using various model circuits with the above characteristics, we were unable to recreate the peaked, positive phase shift at 10 kHz by using model circuits, thus we are continuing to investigate the matter.

One other possibility is the coupling of the capacitance of the electrodes and wires to the Faraday cage where the experiments on the monolayers were carried out [52].

To develop models for categorizing SAMs into those that are defect-free or those having induced defects, the corresponding impedance data were fitted to equivalent circuits by using the algorithms

98 built into the Echem Analyst software package (Gamry, Inc.). Models were constructed using the built- in Model Editor. The Levenberg-Marquardt method was used to fit the data to the CPE-resistor circuit shown in Figure 6.4a. The results are shown in Table 6.1. The 1-Hz frequency was chosen for measurements due to the relaxation times in which diffusion related phenomena occur for alkane

SAMs. This has been consistent with previous studies, which have characterized SAM layers on gold using EIS [30, 38].

For a monolayer that does not exhibit purely capacitive behavior, a Randles circuit was used. Figure

6.4b is a Randles circuit, which includes a solution resistor in series with an RC circuit. The RC circuit includes the double layer capacitance of the surface and the charge transfer resistance, due in part to the ionic transport through the monolayer.

Taken as a whole, the results show that the CPE value decreases, while the α parameter approaches unity as the incubation time increases. The high CPE (although lower than the bare gold CPE value) at

3 h, suggests the presence of defect collapsed sites, implying a thinner dielectric. The presence of pinholes leading to α = 0.944 means that the electrolyte has access to the metal, further adding to the overall capacitance. A low α could also imply low uniformity of dielectric thickness. Increasing incubation time anneals the alkanethiolate into a sufficiently uniform, very close to defect-free, blocking monolayer as evidenced by the low CPE value and α approaching unity for 120 h. Figure 6.5 shows the evolution of the SAM surface with incubation time. This is consistent with the results of kinetic studies of alkanethiol adsorption as reviewed by Ulman [53] and Schreiber [27] where two distinct steps in the absorption kinetics were observed: a very fast step and a second slow step, which can be described as a surface crystallization process. The kinetics of the second step is related to chain disorder, chain-chain interaction (van der Waals) and mobility of the surface chains. As previously suggested, the pinhole size remains invariant, while the number of pinholes decrease with increasing immersion time in thiol solution [29]. More importantly, however, the present study demonstrates that

99 beyond 24 h, an improvement of the insulating character of the SAM is possible; and EIS is a sensitive tool for monitoring the state of the monolayer. Figure 6.6 shows a Bode plot of each monolayer at different incubation times. At 3 h, it is apparent that the number of defects and pinholes greatly affects the impedance at frequencies in the 1-10 Hz range in terms of both magnitude and phase.

The results of Table 6.2 also show that beyond the critical voltage, the charge transfer resistance increases as the incubation time of the monolayer increased. Table 6.2 demonstrated that the critical voltage is not only dependent on chain length, but also the amount of time that the monolayer is grown.

For monolayers grown for 120 h, the critical voltage was extended to larger anodic and cathodic potentials, which in this case is -0.3 V to +0.5 V vs. Ag/AgCl. The critical voltage behavior of this shorter chain alkanethiol (n=10) was previously achieved only for longer alkane chain lengths (n=15)

[37].

Figure 6.7 is a plot of the CPE and α values from the fit of the EIS data obtained for various thiol incubation times (15-, 24- and 120- h), and recorded at different DC potentials. The voltage ranges vary for each SAM, since SAMs grown for lower incubation times have lower resistances for ion conduction to occur through the SAM. The motivation behind these two plots is to show the potential in which the SAM behaves as an ideal capacitor. Beyond the critical voltage for each SAM, the CPE model is no longer valid. As incubation time is increased, the critical potentials are extended from -0.1 and +0.2 V (15 h) to -0.3 to +0.5 V at 120 h. This extends the useful range of the Au/SAM electrode for use in biosensor applications where applied potentials are needed to induce signal transduction currents. Figure 6.8 shows the phase shift from EIS measurements performed at a range of voltages.

As can be seen, the monolayer grown for 15 h exhibits a small voltage range before ion transport through the monolayer becomes relevant. The 48 h monolayer exhibits a larger potential range up to -

0.2 V (vs. Ag/AgCl). It can be seen that the phase shift is > -80° for cathodic potentials beyond -0.2 V

(vs. Ag/AgCl). However, the monolayer grown over 5 days exhibits a phase shift ≤ -80 for voltages

100 ranging from -0.3 to 0.5 V (vs. Ag/AgCl). The authors of this study consider this to be evidence that the number of pinholes within the monolayer has been reduced, thus increasing the resistance for ions to transport through the monolayer.

The bi-modal shape of the CPE curve for 120-h incubation is intriguing, as it suggests two different pzc values. This behavior implies two sites governed by different charge environments. The more negative of the two (-0.1 V) probably corresponds to the alkanethiol pzc, however this is expected to be at more negative potentials. The PZC of an n=11 alkanethiol was reported to be in the vicinity of -

0.37 V vs. SCE [54], to -0.49 vs. Ag/AgCl [31, 55] for UDT SAMs on Au(111). More studies are required to understand the reason a shift in PZC is observed. The minimum at +0.2 V probably corresponds to pzc of pinholes or defects where electrolyte can access small areas of bare gold. The

PZC of Au(111) was measured by cyclic voltammetry in the absence of specific adsorption (HClO4 and NaClO4, 0.1M) to be about 0.23 V vs. SCE [38] or 0.275 V vs. Ag/AgCl. Due to specific adsorption of phosphate anions, we expect a shift towards more negative pzc as seen in the plot, where a minimum is seen at 0.1 V for 15 and 48 h incubations. It is interesting to note that unlike the rest of the monolayers, the 120 h incubation positive pzc has shifted from 0.1 V to 0.2 V (vs. Ag/AgCl). At this point, further comment is not warranted without additional data. The measured CPE corresponding to this PZC, however, is not an indication of the surface concentration of pinholes, as its capacitance is expected to be much higher than that of a full monolayer. The high α values and the low CPE, as well as the ion flux for 120 h incubated sample is an indication of a crystalline, uniform monolayer that has the lowest number of defects across the various incubation times.

Conclusion

We have shown that EIS is a sensitive probe of monolayer condition. Defects in the monolayer are readily detected as deviations from capacitive behavior with the appearance of charge transfer

101 resistance and critical voltage behavior, and are evident even at the usual incubation times reported in the literature. These results have ramifications for investigation and application of electrochemical sensors based on immobilized enzymes on SAMs. EIS is a facile method of screening SAM electrodes for defects, especially since the same equipment (potentiostat in CV or CA mode) used in the enzyme electrode studies is also utilized to check the monolayer (potentiostat in EIS mode). Short-chain thiols

(n<7) generally do not form well-ordered monolayers and sensors built on these thiols are expected to have high double-layer charging from pinholes and background currents from redox species in solution, competing with enzyme catalytic current; and thus, reducing the sensitivity of the method

[18, 56]. Sensors made with longer-chain thiols that were not allowed sufficient time to anneal, will likewise suffer from background currents from pinhole conduction [16]. Although the tunneling barrier at long-chain thiols (n≥14) limits the transduction current, the reduction in electron transport rate is not as severe at medium chains and the ability to form highly-ordered layers resulting in decreased background current is advantageous [17]. The larger range of working potentials afforded by well- ordered monolayer resulting from lengthy incubation times allows a larger repertoire of enzymes with a wider range of E0. For transport studies utilizing a SAM on a gold-coated nanoporous membrane,

EIS can be used similarly as a probe of the insulating property. Insulation is especially crucial for short length scales found in microchannels and nanopores. A wider range of electrode potentials can be applied when the monolayer is highly ordered, allowing a greater range of available diffuse layer potentials. Permeation of ions into a well-ordered monolayer is reduced, extending the usable life of the device. The present work has laid the groundwork for finding the limits of applied potentials on a blocking monolayer with minimal faradaic current. In future work, studies of ionic transport under diffuse layer potential control at blocking monolayer-gold-coated nanopillary array membranes

(NCAMs).

102

References

1. Prakash, S., M.B. Karacor, and S. Banerjee, Surface modification in microsystems and nanosystems. Surface Science Reports, 2009. 64(7): p. 233-254.

2. Gupta, C., M.A. Shannon, and P.J.A. Kenis, Electronic Properties of a Monolayer-Electrolyte Interface Obtained from Mechanistic Impedance Analysis. J. Phys. Chem C, 2009. 113(21): p. 9375-9391.

3. Gupta, C., M.A. Shannon, and P.J.A. Kenis, Mechanisms of Charge Transport through Monolayer- Modified Polycrystalline Gold Electrodes in the Absence of Redox-Active Moieties. J. Phys. Chem C, 2009. 113(11): p. 4687-4705.

4. Walczak, M.M., et al., Reductive desorption of alkanethiolate monolayers at gold: a measure of surface coverage. Surface Chemistry and Colloids, 1991. 7(11): p. 2687-2693.

5. Becka, A.M. and C.J. Miller, Electrochemistry at w-hydroxy thiol coated electrodes. 4. Comparison of the double layer at w-hydroxy thiol and alkanethiol monolayer coated Au electrodes. Section Title: Electrochemistry, 1993. 97(23): p. 6233-6239.

6. Nahir, T.M. and E.F. Bowden, Impedance spectroscopy of electro-inactive thiolate films adsorbed on gold. Electrochemistry, 1994. 39(16): p. 2347-2352.

7. Finklea, H.O., Electrochemistry of Organized Monolayers of Thiols and Related Molecules on Electrodes. 1996, Dekker: New York. p. 109-335.

8. Kang, M. and C.R. Martin, Investigations of Potential-Dependent Fluxes of Ionic Permeates in Gold Nanotubule Membranes Prepared via the Template Method. Langmuir, 2001. 17: p. 2753- 2759.

9. Nguyen, B.T. and C. Toh, Development of an electrode-membrane system for selective Faradaic response towards charges redox species. Electrochemica Acta, 2009. 54: p. 5060-5064.

10. Holtzel, A. and U. Tallarek, Ionic conductance of nanopores in microscale analysis systems: where microfluidics meets nanofluidics. Journal of separation science, 2007. 30(10): p. 1398-419.

11. Tian, C.S. and Y.R. Shen, Structure and charging of hydrophobic material/water interfaces studied by phase-sensitive sum-frequency vibrational spectroscopy. Proceedings of the National Academy of Sciences of the United States of America, 2009. 106(36): p. 15148-53.

12. Wu, Y., et al., Dynamic response of AFM cantilevers to dissimilar functionalized silica surfaces in aqueous electrolyte solutions. Langmuir : the ACS journal of surfaces and colloids, 2010. 26(22): p. 16963-72.

13. Davis, K.L., et al., Electron-Transfer Kinetics of Covalently Atached Cytochrome c/SaAM/Au Electrode Assemblies. J. Phys. Chem. C, 2008(112): p. 6571-6576.

14. de Groot, M.T., et al., Electron Transfer and Lingand Binding to Cytochrome c' Immobolized on Self-Assembled Monolayers. Langmuir, 2007. 23: p. 729-736.

103

15. Chen, X.J., et al., Detection of the Superoxide Radical Anion Using Various Alkanethiol Monolayers and Immobolized Cytochrome c. Anal. Chem, 2008. 80: p. 9622-9629.

16. Ge, B. and F. Lisdat, Superoxide sensor based on cytochrome c immobolized on mixed-thiol SAM with a new calibration method. Analytica Chimica Acta, 2001. 454: p. 53-64.

17. Murgida, D.H. and P. Hilderbrandt, Redox and redox-coupled processes of heme proteins and enzymes at electrochemical interfaces. Phys. Chem. , 2005. 7: p. 3773-3784.

18. Wilson, R.C.K., et al., Flexible, micron-scaled superoxide sensor for in vivo applications. Journal of Electroanalytical Chemistry, 2011. 662(1): p. 100-104.

19. Ryo, Y., W. Hiromi, and U. Kohei, Effect of Temperature on Structure of the Self-Assembled Monolayer on Decanethiol on Au (111) Surface. Langmuir, 2000. 16: p. 5523-5525.

20. Poerier, G.E. and M.J. Tarlov, Molecular Ordering and Gold Migration Observed in Butanethiol Self-Assembled Monolayers Using Scanning Tunneling Microscopy. J. Phys. Chem 1995. 99: p. 10966-10970.

21. Lifshitz, I.M. and V.V. Slyozov, The Kinetics Of Precipitation From Supersaturated Solid Solutions. J. Phys. Chem, 1961. 19: p. 35-50.

22. Hoang, T.K.N., et al., Monitoring the Simultaneous Ostwald Ripening and Solbilization of Emulsions. Langmuir, 2004. 20: p. 8966-8969.

23. Douglass, E.F., et al., Effect of Electrode Roughness On the Capacitive Behavior of Self-Assembled Monolayers. Anal. Chem, 2008. 80: p. 7670-7677.

24. Finklea, H.O., M. Lynch, and S. Avery, Blocking Oriented Monolayers of Alkyl Mercaptans on Gold Electrodes. Langmuir 1987. 3: p. 409-413.

25. Mendes, R.K., et al., Characterization of Self-Assembled Thiols Monolayers on Gold Surface by Electrochemical Impedance Spectroscopy. J. Braz. Chem, 2004. 15(849-855).

26. Wang, W., et al., Electric Potential Stability and Ionic Permeability of SAMs on Gold Derived from Bidentate and Tridentate Chelating Alkanethiols. J. Phys. Chem. C, 2009. 113: p. 3717-3725.

27. Schreiber, F., Structure and growth of self-assembling monolayers. Progress in Surface Science, 2000. 65: p. 151-256.

28. Love J., C., Estroff A., Lara, Kriebel K., Jennah, Nuzz G., Ralph, Whitesides M., George, Self- Assembled Monolayers of Thiolates on Metals as a Form of Nanotechnology. Chem Rev, 2005. 105: p. 1103-1169.

29. Diao, P., M. Guo, and R. Tong, Characterization of defects in the formation process of self- assembled thiol monolayers by electrochemical impedance spectroscopy. Journal of Electroanalytical Chemistry, 2001. 495: p. 98-105.

30. Boubour, E. and R.B. Lennox, Stability of Omega-Functionalized Self-Assembled Monolayers as a Function of Applied Potential. Langmuir, 2000. 16(19): p. 7464-7470.

104

31. Sahalov, H., et al., Influence of Applied Potential on the Impedance of Alkanethiol SAMs. Langmuir, 2007. 23(9681-9685).

32. Oesch, U. and J. Janata, Electrochemical Study Of Gold Electrodes With Anoid Oxide Films. Formation And Reduction Behavior Of Anodic Oxides On Gold. Electrochimica Acta, 1982. 28(9): p. 1237-1246.

33. Hamelin, A., et al., Cyclic voltammetric characterization of oriented monocrystalline gold surfaces in aqueous alkaline solution. J. Electroanal. Chem, 1990. 295: p. 291-300.

34. Losic, D., et al., The Influence of the Underlying Gold Substrate on Glucose Oxidase Electrodes Fabricated Using Self-Assembled Monolayers. Electroanalysis, 2001. 13(17): p. 1385-1393.

35. Bard, A.J. and L.R. Faulkner, Double-Layer Structure And Adsorption. Second ed. 2001, New York Chichester Weinheim Brisbane Singapore Toronto: John Wiley & Sons, INC.

36. Vitarelli, J.M., Jr., S. Prakash, and S.D. Talaga, Determining Nanocapillary Geometry From Electrochemical Impedance Spectroscopy Using Variable Topology Network Circuit Model. Anal. Chem., 2011. 83: p. 533-541.

37. Boubour, E. and R.B. Lennox, Potential-Induced Defects in n-Alkanethiol Self-Assembled Monolayers Monitored by Impedance Spectroscopy. J. Phys. Chem. B, 2000. 104(38): p. 9004- 9010.

38. Janek, R.P. and W.R. Fawcett, Impedance Spectroscopy of Self-Assembled Monolayers on Au(111): Evidence for Complex Double-Layer Structure in Aqueous NaClO4 at the Potential of Zero Charge. J. Phys. Chem. B, 1997. 101: p. 8550-8558.

39. L.V., P. and F. W.R., Electrochemical Impedance Spectroscopy at Alkanethiol-Coated Gold in Propylene Carbonate. Langmuir, 002. 18: p. 8933-8941.

40. Wang, B., et al., Dielectric Properties and Frequency Response of Self-Assembled Monolayers of Alkanethiols. Langmuir, 2004. 20: p. 5007-5012.

41. Campuzano, S., et al., Characterization of alkanethiol-self-assembled monolayers-modified gold electrodes by electrochemical impedance spectroscopy. Journal of Electroanalytical Chemistry, 2006. 586(1): p. 112-121.

42. Darwish, N., et al., Reversible potential-induced structural changes of alkanethiol monolayers on gold surfaces. Electrochemistry Communications, 2011. 13(5): p. 387-390.

43. L, N. and P. T, Fractional Dimension and Fractional Power Frequency-Dependent Impedance of Blocking Electrodes. Electrochimica Acta, 1985. 30(11): p. 1533-1540.

44. S.H., L., Fractal Response for the ac Rough Surface of a Rough Interface. Physical Review Letters, 1985. 55(5).

45. Theodore, K. and G. L.J., Effect of disorder on a fractal model for the ac response of a rough interface. Physical Review B, 1985. 32(11).

105

46. Theodore, K. and S.H. Liu Inverse-Cantor-bar model for the ac response of a rough surface. Physical Review B 1986. 34(7).

47. Kaplan, T., L. Gray, and S. Liu, Self-affine fractal model for a metal-electrolyte interface. Physical Review B, 1987. 35(10): p. 5379-5381.

48. Ramirez, P., et al., Determination of the Potential of Zero Charge of Au(111) Modified with Thiol Monolayers. Anal. Chem, 2007. 79: p. 6473-6479.

49. Bai, L. and B.E. Conway, Three-Dimensional Impedance Spectroscopy Diagrans For Processes Involving Electrosorbed Intermediates, Introducing The Third Electrode-Potential Variable- Examination Of Conditions Leading To Pseudo-Inductive Behavior. Electrochimica Acta, 1993. 38: p. 1803-1815.

50. Aoki, I.V., et al., Ac-impedance and Raman spectroscopy study of the electrochemical behavior of pure aluminum in citric acid media. Electrochimica Acta, 2001. 46: p. 1871-1878.

51. Dinh, H.N., P. Vanysek, and V.I. Birss, The Effect of Film Thickness and Growth Method on Polyaniline Film Properties. Journal of the Electrochemical Society, 1999. 146(9): p. 3324-3334.

52. Fleig, J., J. Jamnik, and J. Maier, Inductive Loops in Impedance Spectroscopy Caused by Electrical Shielding. J. Electrochem. Soc, 1996. 143(11): p. 3636-3641.

53. Ulman, A., Formation and Structure of Self-Assembled Monolayers. Chem. Rev., 1996. 96: p. 1533-1554.

54. Rentsch, S., Siegenthaler, Hans, Papastavrou, Georg, Diffuse Layer Properties of Thiol-Modified Gold Electrodes Probed by

Direct Force Measurements. Langmuir, 2007. 23: p. 9083-9091.

55. Iwami, Y., et al., Determination of the potential of zero charge of Au(111) electrodes modified with thiol self-assembled monolayers using a potential-controlled sessile drop method. Journal of Electroanalytical Chemistry, 2004. 564: p. 77-83.

56. Chen, X.J., et al., Detection of the Superoxide Radical Anion Using Various Alkanethiol Monolayers and Immobilized Cytochrome c. Anal. Chem, 2008. 80: p. 9622-9629.

106

Figures and Tables

Figure 6.1. AFM measurement of bare gold substrate obtained in intermittent contact mode. The scan area shown is 1 μm by 1 μm.

12

8

4

)

-2 0

A cm A

I ( -4

-8

-12 -400 -200 0 200 400 600 E (mV vs Ag/AgCl) Figure 6.2. Cyclic voltammograms of a bare polycrystalline gold surface scanned (▼) and an Au substrate with a “defect- free” thiol monolayer (■), which was grown for 5 days. The supporting electrolyte used was 10 mM (I=0.0316 M) phosphate buffer solution, pH 7 and scan rate 1 V s-1.

107

Figure 6.3. Phase angle plot of undecanethiol SAMs formed at different immersion times, with an applied potential of -0.2 V. The phase angle at low frequencies of the 48-hour grown monolayer exhibits a phase angle less than -88 degrees, for 1mM and 10mM concentrations. The substrate immersed in thiol solution for 5 days shows a phase angle of -88°, an indication of a monolayer without defects for 1mM and 10mM concentrations.

(a)

(b)

Figure 6.4. Equivalent circuits of the thiol monolayer: a) CPE in series with a solution resistance, indicative of a “defect-free” SAM and b), an equivalent Randle circuit of a SAM with potential induced defects.

108

Figure 6.5. Schematic of alkanethiolate SAM evolution on a gold surface over a range of incubation times.

Figure 6.6. Bode plot of undecanethiol applied at 0 V (vs. Ag/AgCl) for different incubation times.

109

Figure 6.7. Plot of model data (CPE and α) obtained from fitting the EIS data for various incubation times. The DC potential was varied from -0.3 V (vs. Ag/AgCl) to 0.5 V. For immersion times shorter than 120 h, the CPE model did not hold over the full DC potential range due to ion penetration.

110

Figure 6.8. Plot of phase shift at 1 Hz from EIS measurements performed at voltages ranging from - 0.3 V-0.5 V (vs. Ag/AgCl).

Table 6.1. Model data for at applied potential of 0 V vs. Ag/AgCl at various incubation times.

2 α -2 Immersion Time Rs (Ω • cm ) CPE (µS • s cm ) α (Hours) 0 (Bare Gold) 282 95 ± .002 0.940 ± .009 3 269 68 ± .008 0.944 ± .001 15 277 1.73 ± .02 0.986 ± .004 48 280 1.6 ± .02 0.992 ± .004 120 281 1.58 ± .02 0.995 ± .004

111

Table 6.2. Fitting parameters for the impedance measurements of SAMs incubated at different times. Calculated using the Levenberg − Marquardt method† CPE in series Randles with Resistor Equivalent Immersion E/V Equivalent Circuit b Rion Time (vs. Ag/AgCl) Circuit b CPE α (MΩ/cm2) Cdl (μF/cm2) (μS*sα 8cm-2) 0.5 0.721 ± 0.025 1.60 ± 0.07 0.4 1.28 ± 0.07 1.56 ± 0.07 0.3 1.47 ± 0.010 1.53 ± 0.05 0.2 1.69 ± 0.02 0.984 ±0.004 ‡ ‡ 0.1 1.64 ± 0.02 0.988 ±0.004 ‡ ‡ 0 1.73 ± 0.02 0.986 ±0.004 ‡ ‡ 15 hours -0.1 1.69 ± 0.02 0.986 ±0.004 ‡ ‡ -0.2 2.73 ± 0.02 1.34 ± 0.07 -0.3 0.788 ± 0.03 1.5 ± 0.07 -0.4 0.456 ± 0.02 1.59 ± 0.07 0.5 1.88 ± 0.02 0.972 ± 0.004 ‡ 0.4 1.73 ± 0.02 0.977 ± 0.004 ‡ 0.3 1.74 ± 0.02 0.976 ± 0.004 ‡ 0.2 1.68 ± 0.02 0.983 ± 0.004 ‡ ‡ 0.1 1.61 ± 0.02 0.991 ± 0.004 ‡ ‡ 48 hours 0 1.60 ± 0.02 0.992 ± 0.004 ‡ ‡ -0.1 1.55 ± 0.02 0.992 ± 0.004 ‡ ‡ -0.2 1.56 ± 0.02 0.988 ± 0.004 ‡ ‡ -0.3 1.26 ± 0.02 1.46 ± 0.07 -0.4 0.573 ± 0.02 1.57 ± 0.05 0.5 1.73 ± 0.02 0.988 ± 0.004 ‡ ‡ 0.4 1.63 ± 0.02 0.991 ± 0.004 ‡ ‡ 0.3 1.59 ± 0.02 0.992 ± 0.002 ‡ ‡ 0.2 1.54 ± 0.02 0.994 ± 0.004 ‡ ‡ 0.1 1.55 ± 0.02 0.994 ± 0.004 ‡ ‡ 5 days 0 1.58 ± 0.02 0.995 ± 0.004 ‡ ‡ -0.1 1.53 ± 0.02 0.994 ± 0.004 ‡ ‡ -0.2 1.55 ± 0.02 0.987 ± 0.004 ‡ ‡ -0.3 1.63 ± 0.02 0.973 ± 0.004 ‡ ‡ -0.4 1.26 ± 0.002 1.37± 0.01 † Two equivalent circuit models were used to describe the insulative behavior of the SAMs depending on the potential applied to each SAM. ‡ A CPE model is (Figure 4a) used due to the insulative behavior of the SAM.

112

CHAPTER 7

STUDY OF IONIC TRANSPORT THROUGH METALIZED NANOPOROUS MEMBRANES FUNCTIONALIZED WITH SELF- ASSEMBLED MONOLAYERS

The text and figures in this chapter are reproduced with permission from the work: Damena D.Agonafer, Muhammed E. Oruc, Edward Chainani, Ki Sung Lee, Huan Hu, and Mark A. Shannon. "Study of Ionic Transport through Metalized Nanoporous Membranes Functionalized with Self- assembled Monolayers." Journal of Membrane Science, 461 (2014): 106-113.

Introduction

Nanoporous membranes have received great attention in the fields of water desalination, biosensing, and chemical separations [1-17]. There have been numerous studies for separation of biomolecules and analytes based on charge selectivity [18-23]. It has been demonstrated that proteins can be separated by applying potentials across conductive alumina membranes [24]. Cheow et al. developed a platinum coated nanoporous alumina membrane to demonstrate the permselectivity of proteins by applying different potentials across the membrane [25]. Several studies have also been done demonstrating charge selectivity through nanoporous membranes [26-31]. Studies have included fabrication of ion-tracked polyethylene terephthalate membranes, with the membrane surface terminated with carboxylate terminal groups [3]. The membranes selectively transported cations and prevented migration of anions. Attaching amino moieties on the membranes surface and pore wall proved to reverse the selective properties of the membrane. It has also been demonstrated that functionalizing gold nanotubule membranes with self-assembled monolayers (SAMs) could alter the pH responsiveness of the membrane [32]. Selective transport through nanopores has also been found

113 to be dependent on the hydrophobic interaction between the analyte and membrane surface. Zenglian et al. used an electroless gold-deposited PCTE membrane modified with a hydrophobic terminated thiol surface. The competing effects between the interactions of the entrance and surface of the nanopore were compared. It was concluded the entrance effect had the most influence on the transport of analyte across the nanopores [33].

Several studies have shown that anion adsorption on a conductive NCAM surface can affect the exclusion enrichment effect of coions and counterions. Mun et al. demonstrated charge selectivity using electroless gold coated polycarbonate track etched membranes (PCTE) [34]. The PCTE membranes are fabricated by bombarding a polycarbonate film with a collimation of high-energy nuclear fission fragments, which results in damaged tracks through the film. The damaged tracks are then etched by an aqueous base solution to create cylindrical pores within the polycarbonate film [35].

The PCTE gold-coated surface was functionalized with propanethiol for prevention of anion adsorption

(e.g. chloride ions) on the surface of the membrane. However, it was observed that the permselectivity of the membrane was inhibited when positive potentials (anodic) were applied at the membrane surface. It was concluded that the poor quality of SAM layer led to irreversibilities on the charge selective system as a result of adsorption of anions. Although the membrane was immersed in thiol solution for 24 hours, studies have shown that more allocated time is necessary to grow a well ordered

(SAM) [36-38]. Authors have shown that detection of analytes can be achieved by interactions with the membrane surface/pore wall [39-41]. Schmuhl et al. investigated electric field-driven transport of ions through supported mesoporous γ-alumina membranes, with surfaces metalized with gold [42].

The studies showed that the electrolyte composition could affect the flux of copper ions through a permselective membrane. It was demonstrated that the flux of copper ions (Cu2+) was enhanced by the presence of adsorbed chloride ions on the gold-coated alumina surface. The same experiments were done with solutions containing fluoride (F-) and nitrate (NO3-) ions. Flux enhancement of Cu2+ was not

114 present when electrolyte solutions containing F- and NO3- were used. As a result from XPS analysis, it was found that chemisorption of chloride on to the gold-coated membrane surface occurred, resulting in a negatively charged surface. It was concluded that the negatively charged membrane surface led to a local accumulation of counter ions, which enhanced the flux of Cu2+ ions.

If specific adsorption of a molecule occurs on the surface of a membrane, ion selectivity for separation will be inhibited by the loss of interaction between the membrane surface and analyte as shown in Figure 7.1a. If the membrane surface is well covered by the SAM in order to minimize ion adsorption, ion selectivity can be enhanced for a specific molecule of interest (Figure 7.1b). The applications of SAMs are often limited by the immersion time allocated for the growth of the SAM.

Past work has focused on studying the properties of a well-ordered SAM by electrochemical impedance spectroscopy (EIS) [37, 43]. It was found that a minimum of 48 hours is needed to grow a well-ordered

SAM [37, 38].

Surface functionalization is crucial for bioseparation and biosensing applications. Savariar, et al. functionalized the surface of a polymeric membrane with SnCl2, which led to separation of analytes by electrostatic and hydrophobic interactions [44]. Additional work has focused on studying the transient current response based on the interaction between proteins and walls of single nanopores [41,

45]. Wei et al were able to immobilize proteins by using an OH- terminated thiolated SAM to act as a protein receptor. The interaction between the protein and nanopore wall can be quantified by the time constant associated with adsorption and desorption events. Without the presence of a SAM on the surface, non-specific protein adsorption can occur on the walls of the nanopore. This can lead to effective blockage within the nanopore with non-targeted proteins, leading to a smaller measurable current (lower S/N ratio) across the nanopore [41]. Moreover, redox currents typically increase the background noise for electrochemical sensors. If detection can be accomplished with lower applied potentials, S/N ratios can be improved for electrochemical sensing of specific analytes [46].

115

In this work, we study the transport of methyl viologen (paraquat, MV2+) and napathalenedisulfonate disodium salt (NDS2-) across a conductive NCAM while applying a range of potentials at with respect to the membrane surface at the entrance side. A polycarbonate track etched

(PCTE) membrane was made conductive by sputter coating gold on the membrane surface. Transport studies were done in a voltage range in which faradaic current was minimized at the surface of the gold-coated nano-capillary-array-membranes (NCAMs). The goal of the transport studies is to demonstrate improved charge selectivity when a well-grown 1-undecanethiol monolayer is assembled at the surface of the NCAM for a wide range of applied potentials (-400 mV

Experimental

7.2.1. Preparation of Gold-Coated NCAMs

PVP coated polycarbonate membranes (PCTE) (Osmonics Inc., MN) membranes were sputter- coated with 50 nm of gold on each side of the membrane surface, with a 6 nm thick titanium adhesion layer. The membrane thickness, nominal pore size, and pore size density were 6 µm, 30 nm, and 6 ×

108 cm-2, respectively. The sputtering power was 300 W with a 23 nm/min deposition rate. The chamber pressure during sputtering was 5 mTorr, and the distance between the gold target and sample was approximately 30 cm. To avoid electrical shortage between each side of the membrane, the edges around the permeation area were not coated. After 2 minutes of gold sputtering, the inner pore diameter of the NCAM did not exhibit any noticeable decrease in pore size (Figure 7.2). 11-undecanethiol, purchased from Sigma Aldrich, was dissolved in absolute ethanol (Pharmaco-AAPER, Shelbyville,

KY) in order to prepare 1mM thiol solution. The NCAM was placed in a membrane holder and immersed in undecanethiol solution for 48 hours.

116

7.2.2. Preparation of Monolayer Surfaces

Figure 7.3 shows a summary of steps involved for preparation of a gold-coated NCAM functionalized with a SAM. The steps include sputter-coating gold on the membrane surface followed by functionalizing the gold-coated NCAM with undecanethiol. The monolayer-coated membrane surface is then rinsed with absolute ethanol and DI water, followed by blow-drying the NCAM surface with N2. Prior to experiments, the NCAM was subjected to the channel outgas technique [47] in order to saturate the membrane pores with electrolyte solution. While the membrane was immersed in phosphate buffer solution, the membrane was placed in vacuum conditions for 48 hours in order to release any trapped bubbles within the pores. Prior to experiments, the NCAM was pretreated with a

1mM phosphate buffer solution, in order to establish equilibrium conditions for the transport experiments.

7.2.3. Preparation of Kapton Electrodes

To establish an electrical connection to the conductive membrane surface, two pieces of gold- coated Kapton (Du Pont 500 HN, 127 μm) were used as electrodes to apply a potential at each surface of the membrane. Kapton films were scored in a long rectangular shape with a 5 mm diameter punched permeation hole for ionic transport. The Kapton film was cleaned with solvents followed by 1 min of oxygen plasma treatment at 100 W, which removed organic matter from the surface in order to improve bonding between the adhesion and gold layers. A 5 nm titanium layer was deposited as an adhesion layer, followed by a 100 nm deposited gold layer. Wires were soldered to the electrodes for electrical connection to the bipotentiostat.

7.2.4. Transport Experiments

Figure 7.4 is a schematic of the membrane permeate flow cell (MPFC). The membrane was sandwiched between two gold-coated Kapton electrodes. The Kapton electrodes were then clamped

117 between the feed and MPFC. Ag/AgCl reference and gold counter electrodes were placed in the feed cell. A CBP Pine Instruments bipotentiostat (Pine Instrument Co., Grove City, PA) was used to apply a potential at each side of the gold-coated NCAM. 1mM phosphate buffer solution (PBS) was prepared as a ratio of mono- and dihydrogen potassium phosphate salts (Sigma-Aldrich, St. Louis, MO) in DI water, serving as an electrolyte. The permeate cell initially had 1mM potassium phosphate buffer solution (PBS) with pH of 7. The total dead volume of the circulating solution was 1 mL. A peristaltic pump (Masterflex L/S 7520-47) was used to circulate PBS solution from the MPFC through a Z-cell

(FIAlab Instruments, Bellevue, WA) for UV detection of molecules.

The two molecules being investigated for transport experiments were methyl viologen dichloride

(MV2+) and naphthalene 1,5 napthalenedisulfonic acid disodium salt (NDS2-), each purchased from

Sigma Aldrich. Initial concentrations of each molecule in the feed cell were 1mM in a 1 mM PBS solution of pH 7. The flowrate used for all experiments was 2.16 ml/min. The effective membrane permeation area was 0.2 cm2. Ionic transport through the NCAM was measured by UV absorption spectroscopy using a DH-2000 deuterium/tungsten-halogen light source and Jaz spectrometer, (Ocean

Optics, Dunedin, FL). The concentrations of ionic spices was determined according to wavelengths:

MV2+ at 257 nm and NDS2- at 228 nm, as shown in Figure 7.5.

The potential of the permeate side of the membrane was held at 0V (vs. Ag/AgCl) in all experiments in order to provide consistent electrochemical conditions. The potential of the feed side of the membrane was varied to study the influence of the diffuse layer potential on the transport of charged species.

118

Results and Discussion

7.3.1. Effects of Transport of NDS2- and MV2+ when NCAM is Functionalized with a SAM

Figures. 7.6a and 7.6b are permeation experiments for NDS2- with and without the presence of an alkane-terminated SAM on the NCAM surface. The fluxes for NDS2- are higher for applied potentials of -0.4V, 0V, and +0.4V (vs. Ag/AgCl) when the NCAM is functionalized with undecanethiol. The increase in flux for NDS2- with the presence of a SAM layer on the membrane surface can be attributed to the reduction of anion adsorption (e.g. phosphate, NDS2-). Several studies have confirmed that adsorption of anions on a bare gold surface, which functions as a working electrode, occur over a range of applied electrochemical potentials [48-52]. In those studies, phosphate buffer solutions (PBS) were used as a background electrolyte. While rarely used as supporting electrolyte in electrochemical studies, phosphate buffer solutions are prevalent in physiological and biochemical studies. There are few studies of phosphate anion adsorption on gold [53-57], and even fewer done at close to neutral pH

[55]. Those studies suggest that phosphate adsorption is higher than sulfate adsorption, but is considerably weaker than halide (Cl-, Br-) ion adsorption [57].

Figure 7.7 illustrates the entrance length of the gold-coated membrane layer is approximately 1.5 times greater than the characteristic pore diameter, which means the surface interaction at the membrane surface can have a strong influence on the transport of ionic species. The ratio between the thickness of the gold layer and pore diameter can be thought as an access resistance in terms of electrostatic interaction of ions at the wall entrance of the pore [58, 59]. As the ratio increases, the interaction of ions at the gold layer can become significant; whereas a smaller gold thickness-pore diameter ratio will lead to a less effective entrance effect as a result of a smaller effective surface area for ions to interact with the charged gold layer. As a result, the functionalization of the gold-coated

NCAM surface with an alkane-terminated SAM can play a crucial role for influencing the wall interaction with analytes (e.g. Cl-, NDS2- , phosphate ions) present in the electrolyte solution. The effect

119 of adding SAMs on the NCAM surface is significant at positive applied potentials (Figure 7.6). This suggests that there is specific adsorption of NDS2- as observed in previous work [34]. Although the adsorption of phosphate ions and NDS2- is weaker for neutral and negative applied potentials, there is still a slight increase in flux of NDS2- when functionalized with a SAM. The thickness of the Au coated layer is only a small fraction of the total thickness of the NCAM (50 nm << 6 um). It can be concluded that the thickness of the gold layer does not have any significant impact on the flux of ionic species related to the total thickness of the membrane.

As discussed earlier, anions adsorb on the gold-coated NCAM surface without the presence of an alkane-terminated monolayer. This can lead to specific adsorption of anions (e.g. NDS2- and Cl-) [42,

57]. As a result, the charge from an external applied potential on the membrane surface can be countered by more than an equivalent charge of anion adsorption [60]. The adsorption of anions creates a negative surface charge density, which creates a diffuse layer potential at the membrane-electrolyte interface. This in turn distorts the local concentration of ions at the surface around the pore entrance.

Although the bulk concentration of NDS2- is believed to not change over the time of the experiment, the local concentration of anionic species at the entrance of the membrane pore can decrease, leading to a smaller local concentration gradient across the nanopore. This in turn can reduce the driving mechanism for diffusion of NDS2- across the NCAM. As a result of the absence of a SAM on the membrane surface, NDS2- molecules are screened due to a negatively charged membrane surface. With the functionalization of a SAM on the NCAM surface, the negative surface charge density is reduced, lowering the driving mechanism for enrichment and screening of counter and coions, in order to preserve electroneutrality. This enables NDS2- molecules to transport across the NCAM with less electrostatic interaction ascribed from the surface charge at the membrane surface; which in turn enables NDS2- to permeate with a higher flux across the NCAM. As can be observed from Figures 7.6a

120 and 7.6b, the flux of NDS2- increases when a monolayer is present over a range of potentials applied at the membrane surface as a result of the reduction of irreversibilities from anion adsorption.

Figures 7.8a and 7.8b show that fluxes for MV2+ are higher for applied potentials of +0.4V, -0.4V and 0V (vs. Ag/AgCl) without the presence of a SAM layer on the NCAM. The results are in contrast with transport of NDS2- through the NCAM. The decrease in flux of MV2+ can be attributed to the presence of Cl- counter ions from the dissolution of paraquat molecules. Past work has demonstrated that working electrodes immersed in KBr electrolyte solution can create a surface excess of K+ at anodic and cathodic potentials, due to specific adsorption of Br- on the electrode surface [60]. The excess of K+ ions at positive applied potentials can be elucidated by the necessity to partially compensate specific adsorbed bromide ions [61]. As a result of the absence of a SAM on the gold- coated membrane surface, the negatively charged NCAM can attract counterions (MV2+), which can lead to local accumulation at the membrane surface. This in turn, increases the local concentration of

MV2+ at the entrance of the membrane, resulting in an increase of diffusional flux of MV2+ across the nanopore. As mentioned earlier, the net negative surface charge on the membrane surface is significantly reduced when undecanethiol is functionalized at the NCAM surface. This leads to a reduction of a negative surface charge attributed from the specific adsorption of anions. Reduction of anion adsorption subsequently acts to reduce the accumulation of MV2+ at the membrane surface. The decrease in accumulation of MV2+ leads to a smaller flux across the NCAM compared to the condition when the gold-coated NCAM surface is not functionalized with an alkane-terminated SAM. Although the variation of diffusional flux varies at all applied potentials, the effect of surface functionalization with SAMs is significant for potentials at 0V and +0.4V. In contrast, the addition of a SAM layer has minimal contribution for diffusional flux experiments for NDS2- at an applied potential of 0V. As previously discussed, the different mechanism for ion selectivity between NDS2- and MV2+ can be attributed to the difference in electrolyte composition. Chloride ions are present in the buffer solution

121 of MV2+, but absent in the electrolyte solution containing NDS2-. As also discussed earlier, Cl- ions have a higher affinity to adsorb as NDS2- or phosphate ions. This can lead to a higher impact on minimization of the ion exclusion-enrichment-effect over a wider range of potentials due to the reduction of adsorption with the presence of a SAM.

7.3.2. Effects of Ion Selectivity when Monolayer Formation is Present

Table 7.1 summarizes the flux and selectivity coefficients for transport experiments for NDS2- and

MV2+. For NDS2-, the percentage of increase in flux with the presence of a SAM is approximately 20%,

12%, 73% for potentials of -0.4V, 0V, and 0.4V. The results show that with the addition of an alkane- terminated monolayer, the enrichment effect of NDS2- at positive applied potentials is increased, due to a reduction of irreversibilities on the membrane surface. The addition of the monolayer increases the ability to enhance flux of anions when the contributions of a negative surface charge density are reduced. The percent decrease of flux of MV2+ when a monolayer is added is 0.5%, 26%, and 52% for applied potentials of -0.4V, 0V, and 0.4V respectively.

The selectivity coefficient for anions is defined as:

훼+0.4/−0.4 = (푓푙푢푥 표푓 푎푛푖표푛 푎푡 0.4푉/푓푙푢푥 표푓 푎푛푖표푛 푎푡 − 0.4푉) (7.1)

The selectivity coefficient for cations is defined as:

훼−0.4/+0.4 = (푓푙푢푥 표푓 푐푎푡푖표푛 푎푡 − 0.4푉/푓푙푢푥 표푓 푐푎푡푖표푛 푎푡 0.4푉) (7.2)

The selectivity coefficient increases from 4.3 to 6.2 when undecanethiol is attached to the gold- coated NCAM. This supports the argument that irreversibilities on the surface of a gold-coated NCAM can be reduced when a monolayer is grown on the surface. This can be attributed to the surface charge density formed on the bare gold surface, which aids in screening the NDS2- molecules from transporting across the NCAM. When the alkane-terminated monolayer is added, the electrostatic

122 repulsion towards anions is reduced, leading to a higher flux of NDS2-. With the reduction of specific ionic adsorption, the contributions of ionic transport are mainly from the externally applied potential at the membrane surface and surface charge density at the PVP coated walls, which will be discussed later in the next section. Similarly, MV2+ displays a distinct improvement for selectivity when undecanethiol is functionalized at the surface. As mentioned earlier, the selectivity of methyl viologen is improved with the addition of a monolayer, due to an increased ability of the membrane to selectively screen MV2+ at positive applied potentials.

7.3.3. Effects of a Charged Nanopore and EDL Thickness on the Transport of Ions

The mechanism for ionic transport across the NCAM is by passive transport. At low ionic strength, the exclusion enrichment effect of ionic species can occur as a result of the surface charge at the walls of the nanopore. This can reduce the effective permeability and cross sectional area of the nanopore for transport of coionic species due to the electrostatic forces at the walls of the nanopore. Although there is no electric double layer (EDL) overlap, it has been postulated that EDL overlap is not required in order to distort the local equilibrium concentration within a nanopore [3, 22, 62].

At an ionic strength of 7mM, the EDL covers about 27% of the NCAM with nominal pore size of

30 nm. This means that electrostatic effects within the nanopore cannot be ignored. Equation 7.3 is an index for comparing the ratio size of the nanopore to the Debye length

1 휀2 ∑ 푛∞ 푧2 2 휅푎 = 푎 ( 푖 푖 푖 ) (7.3) 휀표휀푟푘퐵푇

where κ and a are defined as the inverse Debye length and diameter of the nanopore.

The surface charge of PVP-coated PCTE membranes contain a negative surface charge density at a pH level of 7 [63]. As previously discussed, the ionic mobilities of NDS2- and MV2+ are comparable due to similar molecular volumes (0.637 nm3 for NDS2- and 0.680 nm3 for MV2+), hydrophobic

123 properties (each contain two benzyl rings), and same valencies (z=2) with opposite charge. However, as shown in Figure 7.9, the flux of MV2+ at a potential of 0V (vs. Ag/AgCl) is 8 and 12 times larger for NCAMs with and without the presence of a SAM layer. It should also be pointed out that the ratio of flux between NDS2- and MV2+ decreases as the irreversibilities from anion adsorption at the membrane surface is reduced. Although the EDL only covers approximately 27% of the pore area, screening effects of coions still occur. Figure 7.10 shows a schematic of diffusion transport of anionic species through a gold-coated NCAM. Figure 7.10a shows that at high ionic strength, the electrostatic interactions at the pore wall are negligible due to a small Debye length. Figure 7.10b shows that at low ionic strength, the EDL extension across the nanopore creates a non-homogenous concentration of coionic species. It can be concluded that the flux of NDS2- is significantly less than MV2+ due to electrostatic screening from the walls of the PVP-coated PCTE membrane. Although the surface charge density from the surface of the membrane affects ionic transport, the electrostatic effects from the walls of the nanopore prove to be significant. More studies are needed to conclude the relative significance of each electrostatic interaction.

Conclusion

It has been demonstrated that functionalizing a metalized NCAM surface with undecanethiol can enhance charge selectivity. For a well-grown alkane-terminated SAM, ionic species can be electrostatically separated from the diffuse layer potential at the monolayer/electrolyte interface

(neglecting charge on the head group) at the membrane surface. It was demonstrated that entrance effects at low ionic strengths are not negligible and can dictate the transport of analytes across a nanopore. This work opens up the possibility of utilizing SAMs with different charged head groups to manipulate ionic species based on the diffuse layer potential at the surface of the membrane. EIS experiments in previous work have demonstrated monolayers grown for several days can significantly enhance the quality of a SAM [38], which opens the possibility of improving when incubation times

124 for monolayer growth occur longer than 48 hours. Future work should investigate transport mechanisms at lower and higher ionic strengths to find the critical concentration in which entrance effects become negligible. More studies are needed to study the influence of surface charge from the walls and surface of the NCAM effect ionic transport. This may give more insight in how to improve charge selectivity at both anodic and cathodic potentials.

References

1. Shannon, M.A., et al., Science and technology for water purification in the coming decades. Nature, 2008. 452(7185): p. 301-10.

2. Holtzel, A. and U. Tallarek, Ionic conductance of nanopores in microscale analysis systems: where microfluidics meets nanofluidics. Journal of separation science, 2007. 30(10): p. 1398-419.

3. Nguyen, Q.H., et al., Charge-selective transport of organic and protein analytes through synthetic nanochannels. Nanotechnology, 2010. 21(36): p. 365701.

4. Yuan, H., et al., Characterization of the electrophoretic mobility of gold nanoparticles with different sizes using the nanoporous alumina membrane system. Sensors and Actuators B 2008. 134: p. 127- 132.

5. Kim, D.H., A review of desalting process techniques and economic analysis of the recovery of salts from retentates. Desalination, 2011. 270(1-3): p. 1-8.

6. Eijkel, J.C. and A. van den Berg, Nanofluidics and the chemical potential applied to solvent and solute transport. Chemical Society reviews, 2010. 39(3): p. 957-73.

7. Dermentzis, K. and K. Ouzounis, Continuous capacitive deionization–electrodialysis reversal through electrostatic shielding for desalination and deionization of water. Electrochimica Acta, 2008. 53(24): p. 7123-7130.

8. Lindstrand, V., G. Sundstrom, and A.S. Jonsson, Fouling of electrodialysis membranes by organic substances. Desalination, 2000. 128: p. 91-102.

9. Pilat, B., Practice of water desalination by electrodialysis. Desalination, 2001. 139: p. 385-392.

10. Sata, T., Modification of properties of ion exhcange Colloid and Polymer Science, 1978. 256: p. 62-77.

11. Shaposhnik, V.A. and K. Kesore, An early history of electrodialysis with permselective membranes. Journal of Membrane Science, 1997. 136: p. 35-39.

12. Tanaka, Y., Current density distributin and limiting current density in ion-exchange membrane electrodialysis. Journal of Membrane Science, 2000. 173(179-190).

125

13. Xu, T. and C. Huang, Electrodialysis-based separation technologies: A critical review. AIChE Journal, 2008. 54(12): p. 3147-3159.

14. Chen, X.J., et al., Detection of the Superoxide Radical Anion Using Various Alkanethiol Monolayers and Immobilized Cytochrome c. Anal. Chem, 2008. 80: p. 9622-9629.

15. Davis, K.L., et al., Electron-Transfer Kinetics of Covalently Atached Cytochrome c/SaAM/Au Electrode Assemblies. J. Phys. Chem. C, 2008(112): p. 6571-6576.

16. de Groot, M.T., et al., Electron Transfer and Lingand Binding to Cytochrome c' Immobolized on Self-Assembled Monolayers. Langmuir, 2007. 23: p. 729-736.

17. Prakash, S., M.B. Karacor, and S. Banerjee, Surface modification in microsystems and nanosystems. Surface Science Reports, 2009. 64(7): p. 233-254.

18. Han, J., J. Fu, and R.B. Schoch, Molecular sieving using nanofilters: past, present and future. Lab on a chip, 2008. 8(1): p. 23-33.

19. Peng, X., et al., Ultrafast permeation of water through protein-based membranes. Nature nanotechnology, 2009. 4(6): p. 353-7.

20. Cannon, D.M., et al., Nanocapillary Array Interconnects for Gated Analyte Injections and Electrophretic Separations in Multilayer Microfluidic Architectures. Anal. Chem, 2003. 75: p. 2224-2230.

21. Kler, P.A., et al., High performance simulations of electrokinetic flow and transport in microfluidic chips. Computer Methods in Applied Mechanics and Engineering, 2009. 198(30-32): p. 2360-2367.

22. Plecis, A., R.B. Schoch, and P. Renaud, Ionic Transport Phenomnea in Nanofluidics: Experimental and Theoretical Study of the Exclusion-Enrichment Effect on a Chip. Nano letters, 2005. 5(6): p. 1147-1155.

23. Schoch, R., J. Han, and P. Renaud, Transport phenomena in nanofluidics. Reviews of Modern Physics, 2008. 80(3): p. 839-883.

24. Nguyen, B.T., E.Z. Ting, and C.S. Toh, Development of a biomimetic nanoporous membrane for the selective transport of charged proteins. Bioinspiration & biomimetics, 2008. 3(3): p. 035008.

25. Cheow, P., et al., Transport and separation of proteins across platinum-coated nanoporous alumina membranes. Electrochimica Acta, 2008. 53(14): p. 4669-4673.

26. Huang, C., et al., Application of electrodialysis to the production of organic acids: State-of-the-art and recent developments. Journal of Membrane Science, 2007. 288(1-2): p. 1-12.

27. Lin Teng Shee, F. and L. Bazinet, Cationic balance and current efficiency of a three-compartment bipolar membrane electrodialysis system during the preparation of chitosan oligomers. Journal of Membrane Science, 2009. 341(1-2): p. 46-50.

28. Veerman, J., et al., Reverse electrodialysis: evaluation of suitable electrode systems. Journal of Applied Electrochemistry, 2010. 40(8): p. 1461-1474.

126

29. Sadrzadeh, M. and T. Mohammadi, Treatment of sea water using electrodialysis: Current efficiency evaluation. Desalination, 2009. 249(1): p. 279-285.

30. Sadrzadeh, M. and T. Mohammadi, Sea Water desalination using electrodialysis. Desalination, 2008. 221: p. 440-447.

31. Ben Sik Ali, M., et al., Electrodialytic desalination of brackish water: effect of process parameters and water characteristics. Ionics, 2010. 16(7): p. 621-629.

32. Hou, Z., N. Abbott L., and P. Stroeve, Self- Assembled Monolayers on Electroless Gold Impart pH- Responsive Transport of Ions in Porous Membranes. Langmuir, 2000. 16: p. 2401-2404.

33. Yue, Z., et al., Study of transportation of atrazine and paraquat through nanochannels. Journal of Membrane Science, 2010. 356(1-2): p. 117-122.

34. Kang, M.S. and C.R. Martin, Investigations of Potential-Dependent Fluxes of Ionic Permeates in Gold Nanotubule Membranes Prepared via the Template Method. Langmuir, 2001. 17: p. 2753- 2759.

35. Fleischer, R.L., P.B. Price, and R.M. Walker, Nuclear tracks in solids: principles and applications. 1975: Univ of California Press.

36. Love J., C., Estroff A., Lara, Kriebel K., Jennah, Nuzz G., Ralph, Whitesides M., George, Self- Assembled Monolayers of Thiolates on Metals as a Form of Nanotechnology. Chem Rev, 2005. 105: p. 1103-1169.

37. Boubour, E. and R.B. Lennox, Insulating Properties of Self-Assembled Monolayers Monitored by Impedance Spectroscopy. Langmuir, 2000. 16(9): p. 4222-4228.

38. Agonafer, D.D., et al., Study of Insulating Properties of Alkanethiol Self-Assembled Monolayers Formed Under Prolonged Incubation Using Electrochemical Impedance Spectroscopy. Journal of Nanotechnology in Engineering and Medicine, 2013. 3(3): p. 031006-031006.

39. Martin, C.R. and S.Z. Siwy, Learning Nature's Way: Biosensing with Synthetic Nanopores. Science, 2007. 317.

40. Sexton, L.T., et al., An Adsorption-Based Model for Pulse Duration in Resistive-Pulse Protein Sensing. JACS. 132(19).

41. Sexton, L.T., Horne, L.P., Sherrill, S.A., Bishop, G.W., Baker, L.A., Martin, C.R., Resistive Pulse Studies of Proteins and Protein Antibody Complexes Using a Conical Nanotube Sensor. JACS, 2007. 129: p. 13144-13152.

42. Schmuhl, R., et al., Controlling the transport of cations through permselective mesoporous alumina layers by manipulation of electric field and ionic strength. Journal of colloid and interface science, 2004. 273(1): p. 331-8.

43. Diao, P., M. Guo, and R. Tong, Characterization of defects in the formation process of self- assembled thiol monolayers by electrochemical impedance spectroscopy. Journal of Electroanalytical Chemistry, 2001. 495: p. 98-105.

127

44. Savariar, E.N., K. Krishnamoorthy, and S. Thayumanavan, Molecular discrimination inside polymer nanotubules. Nature nanotechnology, 2008. 3(2): p. 112-7.

45. Wei, R., Gatterdam, Volker, Wieneke, Gatterdam, Tampe, Robert, Rant, Ulrich, Stochastic sensing of proteins with receptor-modified solid-state nanopores. Nature nanotechnology, 2012. 7: p. 257- 263.

46. Wilson, R.C.K., et al., Flexible, micron-scaled superoxide sensor for in vivo applications. Journal of Electroanalytical Chemistry, 2011. 662(1): p. 100-104.

47. Monahan, J., A.A. Gewirth, and R.G. Nuzzo, A Method for Filling Complex Polymeric Microfluidic Devices and Arrays. Anal. Chem., 2001. 73: p. 3193-3197.

48. Hamelin, A., Coadsorption of Sulphate Ions and Pyridine on the (111), (110) and (100) Faces of Gold. J. Electroanal. Chem, 1983. 144: p. 365-372.

49. Ikeda, S. and H. Okuda, Preferential Adsorption of Cations and Anions of Mixed Electrolytes on Aqueous Surfaces. J. Phys. Chem., 1985. 89(7): p. 1140-1147.

50. Kolics, A.r., A.E. Thomas, and A. Wieckowski, 36Cl labelling and electrochemical study of chloride adsorption on a gold electrode from perchloric acid media. Journal of the Chemical Society, Faraday Transactions, 1996. 92(20): p. 3727.

51. Pajkossy, T., Capacitance dispersion on solid electrodes: anion adsorption studies ongold single crystal electrodes. Solid State Ionics, 1997. 94: p. 123-129.

52. Tucceri, R.I. and D. Posadas, A Surface Conductance Study of the Anion Adsorption on Gold. J. Electroanal. Chem., 1985. 191: p. 387-399.

53. Lipkowski, J., Shi, Z, Chen, A, Pettinger, B, Bilger, C, Ionic adsorption at the Au(111) electrode. Electrochim Acta, 1998. 43: p. 2875-2888.

54. Weber, M., Nart, F.C., On the adsorption of ionic phosphate species on Au(111)-an in situ FTIR study. Electrochim Acta, 1996. 41: p. 653-659.

55. Niaura, G., Gaigalas, A, Viler, V.L., Surface-Enhanced Raman Spectroscopy of Phosphate Anions: Adsorption on Silver, Gold, and Copper Electrodes. J. Phys. Chem B, 1997. 101: p. 9250-9262.

56. Weber, M., de Moraes, I.R., Motheo, A.J., Nart, F.C., In situ vibrational spectroscopy analysis of adsorbed phosphate species on gold single crystal electrodes. Colloids Surf. Phycicochem. Eng. Aspects, 1998. 134: p. 103-111.

57. Silva, F., Martins, A, Surface structural effects on specific adsorption of oxoanions on gold single crystal electrodes. J. Electroanal. Chem, 1999. 467(335-341).

58. Velleman, L., J.G. Shapter, and D. Losic, Gold nanotube membranes functionalised with fluorinated thiols for selective molecular transport. Journal of Membrane Science, 2009. 328(1-2): p. 121-126.

59. Hall, J.E., Access Resistance of a Small Circular Pore. Journal of General Physiology, 1975. 66: p. 531-532.

128

60. Bard, A.J. and L.R. Faulkner, Double-Layer Structure And Adsorption. Second ed. 2001, New York Chichester Weinheim Brisbane Singapore Toronto: John Wiley & Sons, INC.

61. Kaifer, A.E. and A.J. Bard, Micellar Effects on the Reductive Electrochemistry of Methylviologen. J. Phys. Chem., 1985. 89: p. 4876-4880.

62. Plecis, A., Schoch, R.B., Renaud, P., Ionic Transport Phenomena in Nanufluids: Experimental and Theoretical Study of the Exclusion-Enrichment Effect on a Chip. Nano Letters, 2005. 5(6): p. 1147- 1155.

63. Lettmann, C., D. Mockel, and E. Staude, Permeation and tangential flow streaming potential measurements for electrokinetic characterization of track-etched microfiltration membranes. Journal of Membrane Science, 1999. 159: p. 243-251.

129

Figures and Tables

Figure 7.1. Schematics (a) and (b) of a gold coated NCAM, which illustrates the improvement when a well-grown SAM is attached to the monolayer.

Figure 7.2. SEM image of the surface of a gold-sputtered PCTE 30 nm, PVP coated membrane.

130

Figure 7.3. Fabrication steps for coating the membrane surfaces and functionalizing the surface with an alkane SAM.

Figure 7.4. Schematic of the membrane permeate flow cell connected to the UV spectrometer system and the bipotentiostat. . WE1, WE2, RE and CE represent working electrode #1, working electrode #2, reference electrode, and counter electrode connections to the bipotentiostat.

131

Figure 7.5. UV absorption spectra for NDS2- and MV2+

Figure 7.6. Diffusional flux measurements of NDS2- through a conductive NCAM when the gold surface of the NCAM is not functionalized with a SAM (a), and when functionalized with undecanethiol (b).

132

Figure 7.7. Schematic representation of a 50 nm sputtered-gold layer at the surface of the NCAM. The thickness of the gold layer is the same order of magnitude of the NCAM pore diameter, indicating the relevance of the interaction of analytes at the gold-electrolyte interface.

Figure 7.8. Diffusional flux measurements of MV2+ through a conductive NCAM when the gold surface of the NCAM is not functionalized with a SAM (a), and when functionalized with undecanethiol (b).

133

Figure 7.9. Diffusional flux measurements of NDS2- and MV2+ through a conductive NCAM at 0 V. Figures (a) and (b) are permeability measurements performed when the NCAM was not functionalized (a) and functionalized with undecanethiol (b).

Figure 7.10. Schematic of the EDL effect on anionic transport. At high ionic strength (a), the Debye length is small, therefore anions can diffuse easily through the nanochannel. At low high ionic strength (b), the Debye length is relatively higher then it reduces the flux of counterions through the nanopore.

134

Table 7.1. Summary of Fluxes of NDS2- and MV2+ and Selectivity Coefficients

Flux of NDS2- (nmoles • min-1 • cm-2) Flux of MV2+ (nmoles • min-1 • cm-2) Applied Voltage (V) Unmodified Modified Unmodified Modified 0 1.44*10-2±5.71*10-4 1.61*10-2±1.30*10-3 1.70*10-1±4.42*10-3 1.26*10-1±7.61*10-3 +0.4 4.75*10-2±3.23*10-3 8.24*10-2±2.18*10-3 1.62*10-1±5.71*10-3 7.81*10-2±3.73*10-3 -0.4 1.11*10-2±5.71*10-4 1.33*10-2±8.33*10-4 2.03*10-1±3.65*10-3 2.02*10-1±6.34*10-3 Selectivity Coefficient for NDS2- Selectivity Coefficient for MV2+- Unmodified Modified Unmodified Modified

휶+ퟎ.ퟒ/−ퟎ.ퟒ 4.3 6.2

휶−ퟎ.ퟒ/+ퟎ.ퟒ 1.25 2.6

135