3. . A .wwfiafi .. 9.2, 3.15., H...» .. ..., . ,3 i 5,. , . 1w..." 3%..“ n5. Lhasa“... .... #1.}. A an: ".5“le g1 . .2...- . Ifl‘llciusmtv . Bur vfimufi.‘ . “3...... n. ad. .n.. 0!? 51...)! tit:: 10.... 3.1. Rail". WA»: 7.- .. , V . 3.: $31.. a . s .... m9: 43km .. 12 v. .. . . .flmfififififl? .. I I ”a fit .mw. “Havana“... at . . . . 1:1..th It’l.\:lv.t 5.....«sllatsvbz 1;! tintvkal 9:...niivaln 410. 1.2.3.3.! 30.3%: it: .63.: n .. J. .1 Eli .- qui..ial : 5134...: $.31 a? ££txhfi|flhflil§§ . Z :1Ilivvl-‘l4lI‘ 211:5). i1 1 5,001.81: .lxlizzll 5131:»..‘3‘ 35-9.. 14!: ’3': $11!. . is. méwuffiiufi E17,. 2 a 543%.; . . kin: P23 £53., €1,152 L! r ... . rt... 3 :4 .t. ¢ K . V. . . .a A. L1 : . In... ..'J :3. L: . . 1.. LIBPARY Michigan State University This is to certify that the dissertation entitled Formation and Characterization of Air Stable Phospholipid Adlayers on Modified Substrates presented by Benjamin Oberts has been accepted towards fulfillment of the requirements for the Doctoral degree in Chemistq % gflrz ( VMajor Professor’s Signature ‘3/ ’%W Date MSU is an Affinnative Action/Equal Opportunity Employer is checkout from your record. T0 AVOlD FINES return on or before date due. PLACE IN RETURN BOX [0 remove [h MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE FORMATION AND CHARACTERIZATION OF AIR STABLE PHOSPHOLIPID ADLAYERS ON MODIFIED SUBSTRATES By Benjamin Oberts A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Chemistry 2009 ABSTRACT FORMATION AND CHARACTERIZATION OF AIR STABLE PHOS- PHOLIPID ADLAYERS ON MODIFIED SUBSTRATES By Benjamin Oberts Plasma membranes are essential to the function of cellular systems. As a con- sequence of this fact many proteins and other biological species are not active outside of their native environment. It would be ideal to create an artificial structure that could house selected biomolecules in their active forms, enabling their use in applications such as bio-sensing, for example. As the understanding of these complex and dynamic bilayer structures increases, it has become clear that their stability in air is not sufficient for use in many chemical sensing applications. For this reason, the work presented here was undertaken to find ways to optimize interactions between lipids and planar substrates, enabling the formation of air-stable lipid adlayers. Several phospholipids were explored; 1,2-dimyristoyl-sn-glycer0-3-phosphatidic acid (DMPA), l,2—dimyristoyl-sn—glycero-3- phosphocholine (DMPC), 1,2-dimyristoyl-sn-glycero-3—phosphoethanolamine (DMPE), l,2-dimyristoyl-sn-glycero-3-[phospho-rac-( 1 -glycerol)] (DMPG), and 1,2-dimyristoyl- sn-glycero—3-[phospho-L—serine] (DMPS). These lipids, in vesicle form, were exposed to chemically modified substrates and underwent vesicle fusion to create lipid adlay- ers. The Au substrates on which the adlayers were deposited were modified to interact with the phospholipid headgroups. The resulting lipid adlayer was stable with respect to transport across the water/air boundary. The chemical modification of the interface was A arson actcn. reacts funcli mlh [ into [1’ lipid a 3M sim in lechm Self-(l: blill}' l accomplished using 6—mercapto-1-hexanol to form a self-assembled monolayer char- acterized by a polar, hydrophilic interface. The 6—mercapto-l-hexanol monolayer was reacted with POCl3 and water to create a phosphate-terminated interface. The phosphate functionalities were populated with Zr4+ ions, rendering them capable of complexation with phospholipid phosphate moieties. Other metal salts were also used to gain insight into the effect of metal ion identity on the binding of phospholipids. To characterize the lipid adlayers, time correlated ellipsometry, water contact angle, cyclic voltametery, XPS, 31P-NMR and F TIR measurements where used. All of the analyses performed were ex- situ from the lipid deposition vessel, demonstrating air-stable adlayer formation. Each technique interrogated a different adlayer property. The discovery of a novel family of self-assembling adlayer comprised of biologically important molecules opens the possi- bility of future success in the creation of robust biomimetic interfacial structures. wan: 7 "mm mi“? »- -‘ Copyright by Benjamin Oberts 2009 To my wife Carrie and our cat Lucky Acknowledgements I would like to first thank my advisor Gary Blanchard for his support and guid- ance through out my time in graduate school. He has been a guide through these tough years and has always been there to point me in the right direction or supply a new idea. I have learned a lot from his teachings and will carry with me the experience and con- fidence that he has instilled in me. As he once said to me “the best way to learn how something works it to take it apart and put it back together.” This holds true and I will apply this mentality were ever I end up and toward any problem I face. I have been very fortunate to have known him and work with him. I would also like to thank my friends and fellow group mates as they have always been there for me through failed experiments and new discoveries. I will miss our week- ly lunches out as well as the time we spent together in the lab. Being friends with and working with such talented and knowledgeable people made it a joy to get in the lab and made it easier to work the long hours that some experiments required. I would like to say thanks to my parents Chuck and Leslie Oberts, who have always been there to support me through all my endeavours. They have always gone far beyond what was necessary to help me. Much of what I am today I owe to them and their continued support. Finally, my greatest thanks goes to my wife Carrie. From moving to Michigan to constantly reassuring me that I will figure out the next problem she has been my rock. Her patience and constant belief in me, even when I did not believe in myself, has been the driving force behind my success. I truly would not have gotten my degree with out her. Out of all the things in my life I am most proud and thankful to be her husband and to have her at my side through it all. vi TABLE OF CONTENTS List of Tables viii List of Figures ix Chapter 1 - ' “ J 1 1-1: Understanding Lipid Bilayers 2 1-2: Phospholipid Vesicles 3 1-2. 1: F luidity and Impact of Impurities on Lipid Vesicles ....................... 5 1—2.2: Lipid Rafts 7 1—3: Supported Lipid Bilayers 10 1-3. 1: SLB Formatinn 11 l-3.2: Inter-Layer Exchange (T ‘ ‘h-n) 13 1-4: Controlling Vesicle Fusion 14 1-5: Introduction of Performed Work 16 Chapter 2 - Formation of Air-Stable Supported Lipid Monolayers and Bilayers ............. 19 2-1: Introduction to Air-Stable SLRs 19 2-2: Materials and Instrumentation Utili7ed 21 2-3: Experimental Setup 21 2—4; Results 25 241: Air Stable Supported Bilayer F ormatinn 25 2-4.2: Air-Stable Lipid Monolayer F ormminn ' 30 2-4.3: NMR Analysis 34 2-5: I." ' 36 2-6: How Mono- and Bilayers Form on the Modified Au Substrates .................. 37 2-7: Kinetics of Lipid Adlayer F nrmatinn 38 2—8: C ‘ ' 39 Chapter 3 - Phospholipid Headgroup Dependent Assembly of Lipid Adlayers on Zirco- nium Phosphate-Terminated Interfaces 42 3-1: Introduction to Surface Modification 42 3-2: Experimental Set up and Materials Used 45 3-2.1: Sample F r " 46 3-3: Individual Lipid Results 47 3-3.1: DMPA 48 3-3.2: DMPF 50 3—3.3: DMPF 52 3—3.4: DMPG 53 3—3.5: DMPS 55 3—4: Overall Lipid Observations 56 3-4.1: Lipids with Strong Headgroup ' ‘ " 58 3-4.2: Lipids with Weak Headgroup ' ‘ " 59 3-5; Diemieeinn 59 3-6; (‘nnr‘lncinns 62 Chapter 4 - Ionic Binding of Phospholipids to Interfaces: Dependence on Metal Ion Identity 64 4-1: ' ‘“ J " 64 4—2: Metals for Modification of Au Substrate-e 66 4-3: Experimental Setup and Instrumentation Utili7ed 67 4-3.1: Substrate Preparation 68 4-3.2: DMPA Vesicle P r " 69 4-3.3: Adlayer Formation 69 4-42 Results Of DMPA Fansnre 70 4-4.1: 7 ' ' 70 4-4.2: Iron ll] 73 4-3.3: Nickel 74 4-4.4: 7 im‘ 76 4—4.5: Calcium 77 4-4.6: “ a ' 78 4—4.7: Copper 1 80 4-4.8: Copper 11 82 4-5: Comparison of Metal-ion Modified Interfaces 83 4-6: F ' ' 85 Chapter 5 - F ‘ ' 87 Literature Cited 93 viii Table 2-1. Table 3-1. Table 4-1: LIST OF TABLES Calculated activation energy for lipid flipping as a function of Arrhenius pref- actor Electrochemical data for Ru(NH ) 3+l2+ and Fe(CN) ”4' probes for interfaces studied here. The SAM indicated in the Table is 6-mercapto—1-hexanol. End- point values are reported vs. Ag/AgCl reference electrode ........................... 60 Data for the interfaces examined in this work. In the first column, metal ion- surface coverage reported by the ratio of the XPS M+n to Au4f signal inten- sities; second column, ellipsometric thickness in A; Water contact angle of DMPA-terminated interfaces, in degrees; water contact angle hysteresis in de- grees; Ru+3/+4 CV peak splitting for metal ion-terminated in-terface; Ru+ CV peak splitting for DMPA-terminated interfaces; F e+2/ +3 CV ratio of non- F aradaic current for metal-terminated interface to DMPA-terminated interface. All CV data has an associated 0.5 mV error 85 Fig Fig Fig F igi FigL LIST OF FIGURES Figure 1-1: 1,2-Distearoyl-sn-Glycero-3-Phosphocholine (DPPC) ................................... 2 Figure 1-2: Model vesicle structure 3 Figure 1-3: Fluorescence micrographs of GUVs at 25°C and compositions of 30% choles- terol mixed with 2:1 DOPC/DPPC, 1:1 DOPC/DPPC, and 1:2 DOPC/DPPC. The scale bars indicate 20 um. Figure adapted from Veatch and coworkers.4. 8 Figure 1-4: a) Langmuir-Blodgett and Langmuir-Schaefer lipid deposition. b) Vesicle fireinn 12 Figure 2-1. Schematic of the Teflon® flow-cell used for the deposition of DMPC adlay- ers by vesicle fusion. The depth of the reservoir is ca. 2 mm ...................... 23 Figure 2-2: F TIR spectra of (a) a 6-mercapto-l-hexanol self-assembled monolayer de- posited on a gold surface. (b) a DMPC adlayer deposited on the 6-rnercapto- l-hexanol-modified gold inter-fem 26 Figure 2-3: Cyclic voltammograms of a 6-mercapto-1-hexanol self-assembled mono- layer deposited on a gold surface (solid line) and of the same interface with an adlayer of DMPC deposited (dashed line), recorded using (a) 1 mM K4[Fe(CN)6] in 0.1 M LiClO4 and (b) 1 mM [Ru(NH3)6]Cl3 in 0.1 M KCl as electrochemical probes 27 Figure 2-4: a) Ellipsometric thickness measurements for a 6-mercapto-l-hexanol-mod- ified Au surface, acquired at a series of reaction times to follow DMPC depo- sition. b) Water contact angle measurements acquired at a series of reaction times to follow DMPC deposition the same inter-face. For both panels, initial data points correspond to t=0 29 Figure 2-5: (a) Ellipsometric thickness measurements for a zirconated 6-mercapto-1- hexanol/Au self-assembled monolayer, acquired at a series of reaction times to follow DMPC deposition. (b) Water contact angle measurements acquired at a series of reaction times for the same interface. For both panels, initial data points correspond to t=0 30 Figure 2-6. F TIR spectra of (a) a zirconated 6-mercapto-l-hexylphosphate self-assem- bled mono-layer deposited on a gold surface. (b) a DMPC adlayer deposited on the zirconated 6-mercapto-l-hexylphosphate-modified gold interface... 31 '6‘ —.In'h Figure 2-7: Figure 2-8: Figure 3-1: Figure 3-2: Figure 3-3: Figure 3-4: Figure 3-5: Cyclic voltammograms of a 6— --mercapto l -hexanol/Au self-assembled mono- layer reacted with POCl3 and ZrOCl to zirconate the interface (solid line) and of the same interface with an adlzayer of DMPC deposited (dashed line), recorded using (A) 1 mM K [Fe(CN)6 ] in 0.1 M LiClO4 and (B) 1 mM [Ru(NH3 )6]C13 in 0. l M KCl as electrochemical probes4 ............................... 32 3 IP MAS NMR spectra. (A) Silica gel coated with POC13, (B) surface shown in panel a exposed to ZrOC12(aq), (C) surface shown in panel b exposed to DMPC 35 Phospholipid headgroups used as well as the overall acyl chain. Base acyl chain does not change for each lipid with the R group being the different D r DMPA results. A) Ellipsometric thickness measurement in time. B) Water contact angle in time with the solid line being the initial drop angle, the dashed line is advancing, and the dotted line is receding. C) CV of DMPA monolayer (solid line) and blank substrate (dashed line) when exposed to Ru(NH )6 C13. D) CV of DMPA monolayer (solid line) and blank substrate (dashed36 lme)3 when exposed to K3Fe(CN)b 49 DMPC results. A) Ellipsometric thickness measurement in time. B) Water contact angle in time with the solid line being the initial drop angle, the dashed line is advancing, and the dotted line is receding. C) CV of DMPC monolayer (solid line) and blank substrate (dashed line) when exposed to Ru(NH )6 Cl. D) CV of DMPC monolayer (solid line) and blank substrate (dashedaé lrne)3 when exposed to K3Fe(CN)0 51 DMPE results. A) Ellipsometric thickness measurement in time. B) Water contact angle in time with the solid line being the initial drop angle, the dashed line is advancing, and the dotted line is receding. C) CV of DMPE monolayer (solid line) and blank substrate (dashed line) when exposed to Ru(NH )6 Cl. D) CV of DMPE monolayer (solid line) and blank substrate (dashed36 lrne)3 when exposed to K3Fe(CN)O 53 DMPG results. A) Ellipsometric thickness measurement in time. B) Water contact angle in time with the solid line being the initial drop angle, the dashed line is advancing, and the dotted line is receding. C) CV of DMPG monolayer (solid line) and blank substrate (dashed line) when exposed to Ru(NH )6 C13. D) CV of DMPG monolayer (solid line) and blank substrate (dashed36 11ne)3 when exposed to K 3OFe(CN) 54 xi Figure 3-6: DMPS results. A) Ellipsometric thickness measurement in time. B) Water Figure 3-7: Figure 4-1: Figure 4-2: Figure 4-3: Figure 4-4: contact angle in time with the solid line being the initial drop angle, the dashed line is advancing, and the dotted line is receding. C) CV of DMPS monolayer (solid line) and blank substrate (dashed line) when exposed to Ru(NH )6 Cl. D) CV of DMPS monolayer (solid line) and blank substrate (dashed,6 lrne) when exposed to K3Fe(CN)O 56 A) Overlay of all thickness measurements in time with DMPA (solid line), DMPC (dashed line), DMPE (dash-dot-dot-dash line), DMPG (dotted line), DMPS (dash-dot-dash line). B) Overlay of initial drop water contact angle measurements with DMPA (solid line), DMPC (dashed line), DMPE (dash- dot- dot-dash line), DMPG (dotted line), DMPS (dash-dot-dash line). C) Overlay of Ru(N H 36) Cl probe CVs with Blank (solid line), DMPA (dashed line), DMPC (bold 3solidaline), DMPE (dotted line), DMPG (dash-dot-dot—dash line), DMPS (dash- dot-dash line). D) Overlay of K F e(CN) probe CVs with Blank (solid line), DMPA (dashed line), DMPC (bo d solid lrne), DMPE (dot- ted line), DMPG (dash-dot-dot-dash line), DMPS (dash-dot—dash line) ......... 58 a) XPS spectrum of Zr+4-modified thiol/gold substrate. b) CV of Ru(NH3)6C13 for a DMPA monolayer on Zr+4-terminated interface (dashed line) and for the Zr +4-terminated inter-face with no adlayer (solid line). 0) CV of K F e(CN) for a DMPA monolayer on Zr+4 -terrninated interface (dashed line) and for the Zr+4-terminated interface with no adlayer (solid line) 71 a) XPS spectrum of Fe+3 -mod1fied thiol/gold substrate. b) CV of Ru(NH3)6 Cl3 for a DMPA monolayer on Fe -terminated interface (dashed line) and for 3the Fe+ 3’-terminated inter-face with no adlayer (solid line). c) CV of K 36Fe(CN) for a DMPA monolayer on Fe+ 3-terminated interface (dashed line) and for the F e 3 -term1nated interface with no adlayer (solid line) 73 3) XPS spectrum of Ni+2-modified throl/ gold substrate. b) CV of Ru(NH3)6Cl3 for a DMPA monolayer on Ni +2-terrninated interface (dashed line) and for the Ni 2-terminated inter-face with no adlayer (solid line). c) CV of K F e(CN) for a DMPA monolayer on Ni+2- terminated interface (dashed Fine) and6 for the Ni 2-terrninated interface with no adlayer (solid line) 75 a) XPS spectrum of Zn+2-modified thiol/gold2 substrate. b) CV of Ru(NH3 )6C1 for a DMPA monolayer on ZnJr 2-terrninated interface (dashed line) and for the Zn 2-terminated inter-face with no adlayer (solid line). c) CV of K F e(CN) for a DMPZA monolayer on Zn+ 2-terminated interface (dashed Fine) and 6for the ZnJr 2-terminated interface with no adlayer (solid line) 76 ll 1 hi! .1; i Fist 5 f'a-JUF: f‘m‘ .. ru‘ : R Figure 4—5: Figure 4—6: Figure 4—7: Figure 4—8: a) XPS spectrum of Ca+2-modified thiol/gold substrate. b) CV of Ru(NH3)6 C13 for a DMPA monolayer on Ca 2-terminated interface (dashed line) and for the Ca 2-terminated inter-face with no adlayer (solid line). c) CV of K F e(CN) for a DMPA monolayer on Ca+ 2-terminated interface (dashed Fine) and6 for the Ca +2-terrninated interface with no adlayer (solid line) _ 78 a) XPS spectrum of Mg+2—modified thiol/gold2 substrate. b) CV of Ru(NH3 ) 6Cl3 for a DMPA monolayer on Mg+ 2-terminated interface (dashed line) and 6for the Mg 2-terminated inter-face with no adlayer (solid line). 0) CV of K Fe(CN)6 for a DMPA monolayer on Mg+ 2-terminated interface (dashed fine) and 6for the Mg 2-terminated interface with no adlayer (solid line) 79 a) XPS spectrum of Cu+ -modified thiol/gold substrate. Inset shows Cu2P spectral region. b) CV of Ru(NH3 )6 C13 for a DMPA monolayer on Cu+ -ter- minated interface (dashed line) and 63for the Cu +-terminated interface with no adlayer (solid line). c) CV of K F e(CN) for a DMPA monolayer on Cu+ - terminated interface (dashed line) and for the Cu+-terminated interface with no adlayer (solid line) a) XPS spectrum of Cu 2-modified thiol/gold substrate. Inset shows Cu2P spectral region. b) CV of Ru(NH3 )6Cl3 for a DMPA monolayer on Cu+ 2-ter- minated interface (dashed line) an3d for the Cu +2-terminated interface with no adlayer (solid line). 0) CV of Fe(CN) for a DMPA monolayer on Cu +2- terminated interface (dashed line) and for the Cu 2-terrninated inter—face with no adlayer (solid line) 83 xiii Chapter 1 Introduction Biological systems are complex and dynamic, and consequently it is difficult to gain molecular scale insight into their behavior. The cellular level is the typical start- ing point for the evaluation of biological function, but progress in our ability to measure complex molecular structures at low concentrations has lead to further questions on how such molecular complexity serves to produce functional biological systems. Mammalian cells contain components such as mitochondria, a nucleus (except for erythrocytes), and a cell membrane, among others. Each of the sub-structures within a cell is character- ized by significant molecular complexity and there is much current research activity in the area of detailed characterization of the component parts of cellular systems. Plasma membranes serve as an excellent example of this complexity. The lipid bilayer structure that defines the plasma membrane serves a critical role in supporting trans-membrane proteins, which regulate cellular functions such as transport of ions and other species into and out of the cell. Perhaps the most basic function of the plasma membrane is to maintain cellular integrity and serve as an impediment to pathogens. Mammalian plasma membranes are composed of more than 100 different components,1 including glycero- phospholipids, sphingolipids and cholesterol, which make up ~50% of the cell wall. The balance of the plasma membrane is comprised of integral and peripheral proteins, and a detailed inventory of plasma membrane components varies according to cell fimction.2’3 The complexity of the plasma membrane poses a substantial challenge to achieving a detailed understanding of membrane form and function. It is thought that this composi- tional complexity plays a role in stabilizing the folding of transmembrane proteins and thus mediating their function. It is clear that the relationship between molecular-scale membrane composition and dynamics and transmembrane protein function is a prerequi- site to the broad-based use of synthetic lipid bilayers in applications ranging from cellular function to chemical and/or biological sensing. It is not the purpose of this dissertation to achieve a connection between bilayer composition and transmembrane protein function. The focus of this work is on the more fundamental issue of how bilayer composition and interactions with interfaces can serve to mediate lipid dynamics and organization. 1-1: Understanding Lipid Bilayers A prerequisite for the study of lipid bilayer structures is the ability to form such structures in a reproducible manner. Given the complexity of plasma membranes and the desire to construct biomimetic structures that are compositionally simpler, the primary focus of this work will be on phospholipids. The choice of phospholipids has been made because this family of molecules comprises the largest fraction of plasma membranes and because phospholipids can form bilayer structures without the addition of other compo- nents, such as sterols, for example. Phospholipids are amphipathic, containing two hydrophobic carbon tails that are bound to any of several hydrophilic head group func— tionalities through a phosphate linkage. Fig. 1-1 shows a phosphocholine as an example Fig. 1-1: 1,2-Distearoyl-sn-Glycero-3-Phosphocholine (DSPC) phospholipid. The head group (choline is shown in Fig. 1.1) can vary in size, polarity, and charge while the acyl chains vary in length and the presence of unsaturation(s). The acyl chains can either be identical or different, both in length and in the presence of unsaturations. This structural variability allows for complex mixtures of phospholipids in l‘ill org; izcc Sign diw are i 158 l llOlCL int-err 1‘13"; these l-I. P' bilayer structures, yielding a commensurate complexity in terms of understanding bilayer organization and stability. Lipid bilayers are quasi-two dimensional structures character- ized by compositional heterogeneity and fluidity, and it is this latter property that adds significant complexity to the study of these systems.4'l6 The work presented in this dissertation focuses on comparatively simple lipid mono- and bilayer structures, which are intended as starting points in developing supported biomimetic structures for potential use in areas such as chemical sensing and selective lipid detection.17 It should also be noted that while this dissertation focuses on planer supported lipid bilayers on modified interfaces, it is expected that the dynamics and molecular interactions of non-supported bilayers are going to carry over to supported bilayers. It is then important to first discuss these interactions. 1—2: Phospholipid Vesicles Phospholipid vesicles where devised as a possible means of simulating a plasma membrane. In an aqueous medium, phospholipids will self-assemble to form a bilayer structure called a vesicle (Fig. 1-2). The key structural features of the vesicle are that the lipid tail region for each lipid layer is directed inward, toward the center of the bilayer, with the hydrophilic head groups oriented outward to be in contact with the (aqueous) solution, and that the bilayer structure closes on itself to preclude exposure of the acyl chain region to an aqueous environment. The presence of the inner lipid layer differentiates a vesicle from a micelle structure. Through the study of bilayer structures, Fig 1-2: Model vesicle structure mostly in the form of vesicles, it has been found Kim " Ls we that even for binary and ternary systems, phase separation between the constituents can occurg’15 '18'20 The phase separation of different lipids and/or lipids and sterols plays a major role in determining the organization and fluidity of the vesicles. For such compara- tively simple systems to provide meaningful insight into organization in the substantially more complex plasma membranes, it is important to understand first the organizational behavior of the model system. It is thought that the phase separation and fluidity that is characteristic of plasma membranes mediate the function of transmembrane protein.2l’22 The issue of bilayer composition is inherently related to bilayer fluidity. Bilay- ers exhibit phase transitions between a variety of phases, depending on the composition of the bilayer and the temperature. It has been observed that several phase transitions occur with increasing temperature. At a low temperature for a given lipid, a bilayer will exist in a crystalline gel phase (LC), which undergoes a transition to lamellar gel phase (LB)’ which can undergo a further transition to a rippled gel phase (PB) with increasing temperature.23 The gel—to—fluid phase transition, which occurs with a further increase in temperature, is labeled the gel-to-fluid phase transition temperature (Tm), and is thought to proceed because of the acyl chains undergoing a structural change from predominantly all-trans to a conformation with a significant contribution fiom trans-gauche conform- ers.24 The rationale for phase transitions in bilayer systems is the balance between thermal energy and the attractive inter-chain interactions that operate in the acyl chain region of the bilayer. It is clear on physical grounds that the value of Tm, or any of the preceding transitions will be influenced by the length and degree of unsaturation of the lipid acyl chains. The measurement of Tm is a well established means of characterizing bilayer structures.21’22’25'27 The relative ease of measuring Tm compared to other phase transitions, and the sensitivity of the transition temperature to impurities within the bi- layer, provides an important gauge of lipid purity and suggests that impurities may play -— an important if underestimated role in determining the range of literature values that have been reported for a given system. 1-2. 1: Fluidity and the Effect of Impurities on Lipid Vesicles Koan and coworkers28 examined the effect of impurities on the phase transition temperature of the C14 phosphocholine DMPC using time-resolved fluorescence measure- ments. These experiments reported on the anisotropy decay dynamics of a chromophore inserted into the acyl chain region of the vesicles as function of the amount of impurity (14:1 PC) present in the bilayer. It is useful to review this measurement to understand how Tm is extracted from the anisotropy decay data. The induced orientational anisotro- py function, which is a measurement of the orientational relaxation of molecules excited by polarized light to a random distribution, can be performed using time-correlated single photon counting (TCSPC) spectroscopy. Data collected from this measurement contains information pertaining to the motion of a chromophore within the lipid bilayer, which is present in this case in the form of a vesicle. The goal of anisotropy decay measure- ments is to acquire molecular reorientation information, specifically the reorientation time constant(s), which depend on the identity of the chromophore and the environ- ment in which it resides. 1n the case of this work, temperature-dependent changes in the reorientation time constant reveal the temperature, Tm, at which the gel-to-fluid phase transition occurs. The chromophore local environment will affect the functional form of induced orientational anisotropy decay. Calculation of I”(t) — Ii (t) r = -l) the anisotropy decay function with equation 1-1 is accom- Ill“) ‘1' 21 J. (t ) plished by taking the normalized difference between the r(t) = r(0)exp[t——t] (1-2) 0R flzL (1-3) kBTS 6D fluorescence transients polarized parallel and perpendicular to the incident vertically polarized excitation pulse. In its t OR = simplest interpretation, the anisotropy decay function is fit to equation 1-2 to extract the reorientation time, Tor. In equation 1-2, r(0) is the initial anisotropy, which is related to the angle between the excited and emittng transition dipole moments, and can range from -0.2 to 0.4. Measurement of the reorientation time allows the calculation of the viscos- ity h according to the modified Debye-Stokes-Einstein equation (equation 1~3).29'31 In this equation, Vis the hydrodynamic volume of the chromophore, calculated according to Edward’s formulation,3 2 k1) is the Boltzmann constant, T is the temperature in K, f is the “friction” coefficient for the solvent-solute interaction boundary condition, and S is the shape factor to account for non-spherical shapes, which ranges from 0 to 1. This data yields a general picture of the system. Koan et al,28 used perylene as the “probe” chromophore. Perylene has a well characterized linear response and is a planar polycyclic aromatic hydrocarbon, a structure useful for the interrogation of the lipid bilayer acyl chain region. A time-correlated single photon counting apparatus was used and the results from this system allowed detailed information to be obtained on the chromophore local environment.33'34 The bilayer sys- tem used was composed of two different lipids; 1,2-dimyri.,:v ,1 5:1 _ ‘ r‘ “I ‘ ' " (14:0 PC) and an “impurity” of l,2-dimyristoyl-sn-glycero—3-phophocholine (14:1 PC). Fluorescence lifetime and anisotropy data was collected for perylene in vesicles contain- ing controlled ratios of 14:0 PC to 14:1 PC at several temperatures. These data exhibited a discontinuous change in tor at Tnn which is interpreted in the context of a change in the molecular scale organization of the lipid bilayers. Results from this work indicate that impurities can significantly influence the phase transition temperature Tm.28 Vesicles of 14:0 PC exhibited Tm at 24°C, in agree- ment with the literature.35 Upon addition of 14:1 PC “impurity”, Tm decreased consider- ably. F or 0.3 mol% of 14:1 PC, Tm decreased by ~15°.28 Increasing the concentration of 14:1 PC further lowered Tm, but the change in Tm with 14:1 PC not as great as it was for the initial addition. Tm changes with increasing 14:1 PC concentration exhibit a monotonic decrease, demonstrating that a relatively small amount of impurity gives rise to large changes in the organization of the bilayer. It is interesting to note that while the value of Tm changed significantly with the presence of the impurity, the viscosities sensed by the chromophore did not. For the fluid phase viscosities were in the range of 8.5 :L- 1.5 cP and the gel phase viscosities lied in the range of 14.5 i 2.5 CP.27 While the impurities may perturb the organization of the bilayers, the interactions responsible for maintaining bilayer structure do not change. The overall conclusion to be drawn by this work is that even the simplest systems pose problems in understanding their intrinsic complexity at the molecular scale, and further work is necessary to develop an understanding before it will be possible to create systems that can function as biomimetic plasma membrane structures. While this study focused on the bilayers in the form of vesicles, the interactions probed are pertinent to supported bilayer structure described in this work. l-2.2: Lipid Rafts Lipid heterogeneity in vesicle systems has lead to further observations and chal- lenges in regard to the complexity of these systems. Binary mixtures of cholesterol and phospholipids are found to be miscible however, when cholesterol is mixed with two phospholipids of varying melting temperature (Tm) to form a ternary mixture, cholesterol can complex preferably with one phospholipid creating “rafis” or condensed complexes that phase separate.”39 Rafts have been observed as distinct domains with reduced diffusion within cell membranes.2’8 These physical properties may suggest that these organized regions could be locations for specific proteins to interact with the membrane, which could lead to further bilayer satiability. It has also been theorized that some pro- teins require the presence of raft structures and, as such, pure phospholipid bilayers will not provide the required biomimetic environments for functional biomolecules.3’8 These theories suggest that lipid rafts are important due to their putative role in mediating pro- tein functionality as well as other cellular functionsm’39 It has been proposed that the cholesterol rafts will form between mixtures of cholesterol and saturated lipids because cholesterol is immiscible in unsaturated lip- ids.38 Mixtures of cholesterol and two lipids with different melting points have exhibited two liquid phases, particularly if the Tm’s for the two different lipids are widely differ- ent.“”’4042 To observe and characterize these lipid-cholesterol domains, Veatch and co- workers43 utilized lH-NMR techniques and fluorescence microscopy. Giant unilarnellar vesicles (GUVs) were created that were composed of 30% cholesterol and varying ratios of 1,2-dioleoyl-sn-glycero-3-phosphocholine (DOPC) and l,2-dipaplmitoyl-sn-glycero- 3-phosphocholine (DPPC). 1H-NMR data reveal that not only are domains being formed but it is possible to map the phase boundary of the domains from one relatively homog- enous domain in the vesicles, to two liquid domains containing lipid and cholesterol or lipid only. Fluorescence microscopy was utilized to image the GUVs for that investiga- tion. Fig. 1-3 shows the fluorescence micrographs of the GUVs at different ratios of lip- ids.43 It is clearly visible from Fig. 1-3 that lipid domains are being formed, as well as upon larger concentrations of the Fig. 1—3: Fluorescence micrographs of GUVs at 25°C miscible lipid DPPC, the do- and compositions of 30% cholesterol mixed with 2:1 DOPC/DPPC, 1:1 DOPC/DPPC, and 1:2 DOPC/DPPC. mains become larger until they The scale bars indicate 20 um. Figure adapted from encom ass the ma‘ ' Veatch and coworkers.42 p jorrty 0f the GUV, leaving DOPC minor domains. Both the 1H-NMR and fluorescence data indicate the formation of two phase domains. From this work it was found that the domains can be as small as ca. 80 nm and projected phase diagrams are consistent with other studies.“ To quantify the results reported by Veatch and coworkers,43 a thermodynamic model was developed by Radhakrishnan and McConnell.36’37 In their model it is as- sumed that there are four different species in the liquid phase of the bilayer at equilibri- um. These consist of un-reactive lipids (DOPC), reactive lipids (DPPC), free cholesterol, and the reactive lipid-cholesterol complex. Assuming all species are present in equimolar 0 amounts, the Gibbs free energy can be express as Equations 1-4.37 In this equation ”1 _ 0 0 G _Zix,(m +kT1nxi)+2kZi 400 mV) of a weak wave seen for the lipid adlayer, suggesting very slow electron transfer kinetics at this interface. The presence of the wave following the addition of the lipid adlayer, which is not present with the SAM alone, implies that the adlayer is perturbing the organization of the SAM to some extent upon deposition. The corresponding data for the Ru(NH3)63+/2+ probe reveal the presence of a redox wave for both adlayers as well as for the SAMs on which the adlayers are de- posited (Figs. 2-3b, 2-7b). It is not surprising to observe this electrochemical behav- ior because lipid adlayers are likely to be characterized by some defect sites, and the Ru(NH3)63+/4+ probe is more sensitive to interface defect sites. It is clear for both in- terfaces that the electron transfer kinetics are slowed significantly by the presence of the lipid adlayer. As noted above, for the 6-mercapto-l-hexanol SAM, the peak separation 33 inter same ltise With 1 lipid zircoi acid lion 2 gm ‘tt for the Ru(NH3)63+/2+ waves is ca. 120 mV with a midpoint potential of -154 mV, and this separation increases to ca. 255 mV (and the midpoint potential shifts to -200 mV) with the addition of the lipid bilayer (Fig. 2-3b). For the zirconated interface (Fig. 2-7), the redox peak splitting is ca. 117 mV (midpoint potential -153 mV) and the addition of the lipid adlayer increases the splitting to ca. 143 mV (midpoint potential -l79 mV). For a reaction that is reversible on the time scale of the potential scan, the expected (Epa-Epc) splitting should be 59/n mV at 25° C. The values obtained for these interfaces are all in excess of 59 mV, indicating that the kinetics of the electron transfer are mediated by the interfacial adlayers in all cases. The splitting and midpoint potentials are essentially the same for both the 6-mercapto-l-hexanol and zirconated SAMs, which is not surprising. It is also not surprising that the lipid bilayer impedes the electron transfer kinetics to a greater extent than the lipid monolayer, as manifested by a comparison of the peak split- ting and midpoint potential data. The electrochemical data are in qualitative agreement with the ellipsometric and contact angle data, and are consistent with the formation of a lipid bilayer structure on the 6-mercapto-1-hexanol SAM and a lipid monolayer on the zirconated SAM. 2—4.3: NMR Analysis The ZP-lipid interactions that are central to understanding the formation of the monolayer data can be characterized using 31P NMR measurements. 31P NMR measure- ments have been used in the past to characterize ZP multilayer structures, and this work follows those same procedures.87 All NMR spectra were referenced to 85% phosphoric acid (5=0ppm). Fig. 2-8 shows the 31P NMR spectrum for each step of interface prepara- tion and DMPC exposure. For this work silica gel was used as the substrate for interface growth because of the surface area advantage. As seen in Fig. 2-8a, after exposure of 34 I ~—-‘\ v ~ --~-_—— "r- alv'u'm‘ mwmarv‘ - 3". ’ \—,I 's. - - “ --fi ,1" v “4", 10 0 O - 80 PP“ 140 100 60 140 100 '60 .r—_--.__‘_r.- affirm-Jr. ’1‘. “,‘I r.“ 1. I 140 100 so ' 20 20'1' .‘ x , ,t,‘ I o”350" ! t i 0'lid"w' pp- | . ‘L”‘1}VI"."" r . r ~-- ‘- ‘.-'.- k—r , Lu ‘—'.,_.-.4‘ -60 -100' bps Fig. 2-8: 3 IP MAS NMR spectra. (a) Silica gel coated with POCl , ZrOCl (aqi, DMP . (b) surface shown in panel a exposed to (c) surface shown in panel b exposed to the silica gel to POCI3 a strong resonance is observed at 8=-l .1 ppm, and a smaller resonance is seen at 6=-14.9 ppm, cor- responding to physisorbed and chemically bound phosphate at ' .-60.....-._ 2136“”pr the silica gel surface, respective- ly. The resonance at -1.1 ppm is extremely strong because the physisorbed phosphate cannot be removed easily. The remain- ing bands are due to the spectra not being proton decoupled, and the presence of spinning side bands.87 On exposure to ZrOClz, the spectra exhibit a characteristic shift (Fig. 2-8b).87 Poorly resolved peaks at 8=-4.7, -12.0, and -l9.7 ppm result from zirconium coordination with a phosphate. The prominent peak at -l .1 ppm is eliminated because the Zr+4 in solution complexes the physisorbed phosphate, removing it fi'om the silica gel surface. The spectrum shown in Fig. 2-8c was recorded following exposure of the zirconated surface to DMPC. The 35 v --u- p3..qetn{sv_1(“.1".MAI'lifleig-ii'fl‘“ ’ " “ ‘ ‘ broad peaks at 5=-12.1 and -l9.4 ppm correspond to the zirconium-phosphate complex, and the sharp resonance at -O.5 ppm is due to zirconium complexation by the phospho- choline headgroup of DMPC. This assignment is made by comparing these data to a solution phase DMPC spectrum, which has a single peak at 6=-0.32 ppm (not shown). Washing the silica gel with water does not remove the resonance at -0.5 ppm, indicating that the DMPC phosphocholine moiety complexes with the Zr+4 present at the interface. While this is not a surprising result when viewed in the context of the literature on ZP 78,79,81-86,100-103 chemistry, it is one of the first reports that is available showing a direct complexation between a phospholipid and a zirconated substrate. 2-5: Discussion The FTIR, ellipsometry, electrochemistry, contact angle and 31P NMR data pro- vide a self-consistent picture of these interfaces. A stable DMPC lipid bilayer is formed by vesicle fusion on the 6-mercapto-1-hexanol-coated Au substrate and a DMPC lipid monolayer is ultimately formed on the zirconated substrate, with the phospholipid head- group interacting strongly with the Zr+4. The F TIR data confirm the presence of phos- pholipids on the surfaces through the carbonyl stretches, which could be there only as a result of phospholipid deposition. Ellipsometric thickness data are consistent with the presence of a lipid bilayer on the mercaptohexanol-terminated surface and a lipid mono- layer on the zirconated phosphate surface. These data however, do not provide explicit information on the organization of the phospholipids. The F TIR data suggest that the acyl chains of DMPC are not in a fully crystalline state, as gauged by the band positions of the asymmetric and symmetric CH2 stretches. The cyclic voltammetry data, using two different electrochemical probes, demonstrate that the DMPC mono- and bilayers are not free of defects, but they do cover the interface to a significant extent. Water contact angle 36 data are complementary to the ellipsometric and electrochemical data and show that the bilayer is hydrophilic while the monolayer is hydrophobic. Taken collectively, the F TIR, ellipsometry, contact angle, and electrochemical data confirms that a DMPC bilayer is being formed on the 6-mercapto-l-hexanol coated substrate and a monolayer is formed on the zirconated substrate and that these interfacial structures remain even after being removed from the aqueous environment in which they are formed. The 31P NMR data demonstrate that there is a measurable interaction between the DMPC head group and the zirconated interface, consistent with the known chemistry of Zr—bisphosphonate systems and the ellipsometric and contact angle data presented here. There is limited precedent for the formation of air-stable bilayers. The approach to creating air-stable bilayers described here relies on the formation of a well organized interface on which the phospholipid mono- or bilayer can form. The mercaptohexanol base layer is comparatively well organized and tightly packed, based on the CV data (Figs. 2-3). Such an interface allows for substantial interaction of the terminal —OH fimctionalities with the phospholipid head group. For the case of the zirconated interface, the dominant chemical interaction is the formation of a ZP-like complex with the phos- pholipid headgroup, as confirmed by 3 1P NMR measurements. For both interfaces, the interactions between the terminal chemical functionality and the phospholipid headgroups are sufficiently strong to allow the formation of structures that are stable in a range of environments. 2-6: How Mono- and Bilayers Form on the Modified Au Substrates The interpretation of this data in the context of a lipid bilayer forming on a 6-mer- capto-l-hexanol SAM and a lipid monolayer ultimately forming on a zirconated SAM requires a consideration of how each of these structures can form and has implications 37 in terms of the mass of phospholipid that ultimately deposits at the interface. For lipid adlayer formation on the 6-mercapto-l-hexanol SAM, vesicle fusion gives rise to the for- mation of a bilayer structure. When a vesicle contacts the interface, then spreads on that interface, there is the opportunity for lateral mobility of the bilayer constituents owing to the nature of the interface-lipid interactions. Because of the translational mobility of the planar bilayer on the 6-mercapto-l-hexanol SAM, vesicle fiision can proceed on open regions of the SAM until an essentially complete bilayer forms. It is asserted that the physical “picture” for the zirconated interface is fimda— mentally different. While vesicle fusion proceeds on the zirconated interface, once the phospholipid headgroups of the bottom leaflet of the planar bilayer contact the zirconated interface, they bind and are not free to execute translational motion. Thus the interface coverage is heterogeneous initially, and the interstices between covered regions are, of necessity, left open once they become smaller than an area that can accommodate vesicle deposition and fusion. Over time, mediated by translational motion of the top lipid leaflet and translocation of the top leaflet constituents once they reach the edges of the bottom- leaflet “islands”, the open interstitial regions are filled in with phospholipid molecules and a lipid monolayer results. The coverage of a planar surface achieved by vesicle fusion is difficult to model, but if it is assumed that the spherical vesicle fuses to form a circular bilayer which is not free to translate on the surface, one can estimate the maximum achievable surface cover- age. F or a hexagonal close-packed array of circles, the maximum coverage would be 94% and for a face-centered cubic arrangement, coverage would be ca. 78%, based on simple geometric models. Because of the nominally random deposition of the vesicles on the zirconated surface, it is improbable that such high initial surface coverage is achieved. There may ultimately be a slight excess of lipid molecules present by the end of the par- 38 . v .- r"o.|1:!.VDQP[4{ "A". hfl‘t!" "01".“ tial bilayer deposition and monolayer formation process. Any excess “top leaflet” mole- cules will likely dissolve into the bulk solvent and/or reincorporate into lipid vesicles that remain in solution. 2-7: Kinetics of Lipid Adlayer Formation With this model of the interface in mind, one can consider the energetics and kinetics of lipid monolayer formation on the zirconated interface. From the data shown in Fig. 2-5a, there appears to be a time constant associated with the formation of the lipid monolayer. Initially it appears that a partial lipid bilayer structure forms on the surface, and over time the partial bilayer converts to a monolayer. Through the use of a zircon- ated interface, a condition has been established where a lipid initially in the top leaflet will bind to the zirconated surface essentially irreversibly once it migrates to an edge of the lipid island and executes a translocation to the bottom leaflet. The energetics associ- ated with this process based on the time constant observed for the evolution of the mono- layer structure can be estimated. Using the ansatz that, for the interactions of lipids with zirconated surfaces, the relevant unimolecular reactionlo’65’72 is k N top lop—bottom > Nbonom (2_ 1 ) While it is possible that there is some dissociation from the lipid-ZP complex, the rate of dissociation is expected to be slow. The energy of formation for a Zr-bisphosphonate complex, which is similar to the lipid-ZP complex, is known to be 2250 kJ/mol, cor- responding to a dissociation constant at room temperature of ~3x10'44 for 1:1 lipid:Zr stoichiometry.80 The decay of interface thickness shown in Fig. 2-2a corresponds to the time constant for lipid top-to-bottom flipping, k = 1:"1 = 8.3 (+125, «3. 1) x10'4 top-bottom 39 s'1 , a value that is similar to literature reports for lipid translocation.10 By treating this interface evolution reaction as an activated process, the Arrhenius prefactor associated with such a reaction will lie within the range of 1010 — 1015 Hz. Because temperature- dependent data for the formation of these interfaces has not been examined, it is not A (5") Ba (kJ/mol) 1010 73.3 Table 2-1: Calculated activation en- ” ergy for lipid flipping as a fimctron of 10 79'0 Arrhenius prefactor. 1012 84.6 1013 90.2 ktlolp-bo om zt-le'texPGEa/Rfll} w ere t 4,0“ = 8.3 x10-4 s , 10l4 95'8 T = 300 ie, ananit = 8.314 J/mol-K. 1015 101.4 possible to determine the prefactor experimentally. One can, however, determine the range of activation energies consistent with the unimolecular prefactor range (Table 2-1). Calculating the activation energies for DMPC lipid trans-leaflet migration using an Arrhe- nius prefactor in this range and with the experimental time constant measured, activation energies in the range of 73 — 101 kJ/mol are obtained. It is noted that the values for Ba are in relatively close correspondence with those reported by Kornberg and McConnell for translocation of a tagged lipid,10 while Liu and Conboy72 report Ea values on the order of 200 kJ/mol or more for lipid translocation. The fact that phosphocholines interact strongly with a zirconated interface sug- gests the ability to scavenge certain phospholipids from multi-component solutions using the appropriate interface chemistry. At the present time it is not clear that all phospholip- ids will interact equally well with the zirconated interface, and the process of understand- ing the role of phospholipid headgroup identity on the strength of lipid binding to zircon- ated interfaces is discussed in the following chapter. The phospholipid headgroup is not all that is anticipated that regulates the formation of lipid monolayer structures, and this points the way toward evaluating the strength of interaction of phospholipids with other 40 metal ions as well. 2-8: Conclusion Here was observed a means of forming lipid mono- and bilayer structures that can cross the air/water interface and remain intact. The interfaces formed have been charac- terized using F TIR spectroscopy, cyclic voltammetry, optical ellipsometry, contact angle measurements and 31P NMR data. Reported here is the formation of interfacial lipid layers on substrates that are either polar (hydroxythiol on Au) or are capable of binding the phospholipid headgroup (phosphated and zirconated interface). Complexation of the DMPC phosphocholine headgroup to the zirconated interface was confirmed by 31P NMR data, demonstrating for the first time, at the time of this writing, knowledge of the com- plexation of a phospholipid in such a manner. It hasbeen found that for a DMPC bilayer on the zirconated surface, the time constant for partial bilayer-to-monolayer conversion is ca. 20 minutes, and for a unimolecular reaction,lo’65’72 this time constant is consistent with an activation energy between 75 and 100 kJ/mol. Given the interaction between phospholipids and ZP-terminated interfaces, it is important to understand how the interac- tion varies with the identity of the phospholipid headgroup, which is discussed next. 41 Chapter 3 Phospholipid Headgroup Dependent Assembly of Lipid Adlayers on Zirco- nium Phosphate-Terminated Interfaces 3-1: Introduction to Surface Modification The basis for the formation of a lipid bilayer structure is the balance of intermo- lecular interactions between the lipid nonpolar acyl chain regions and the polar head- group interactions with the (aqueous) medium with which the bilayers are in contact. Lipid bilayers that are present in biological systems are comprised of many constituents and are structurally complex. It is thought that this complexity plays a role in stabilizing the folding of transmembrane proteins and thus mediating their function. There is a significant research effort involved with chemical sensing based on the use of biomolecules as the chemically selective elements. To succeed in using certain biomolecules as chemical sensing elements, an interface is required that can stabilize the structure of the biomolecule and at the same time fimction as part of a transduction sys- tem to relay the chemical signal of interest to instrumentation. Supported lipid bilayers are an appropriate choice for such purposes. The bilayer composition and the manner in which the bilayer interacts with the interface to which it is bound need to be investigated as an initial step in this effort. It is therefore of interest to bind selected phospholipids to chemically modified interfaces, and one way to perform this binding is through interac- tions between the phospholipid headgroup moieties and the supporting surface similarly to that discussed in chapter 2. The work discuss here however, is focused on the depo- sition and characterization of lipid monolayers, not bilayers of several phospholipids. Investigation in to lipid monolayers was pursued because they provide an interesting opportunity to develop an understanding of the interaction(s) between the substrate and 42 the lipid headgroups. This is valuable because it is these interactions that ultimately will govern the bilayer integrity and physical properties. Once the substrate-lipid interactions are understood, it is possible to add an outer phospholipid leaflet by Langmuir-Schaefer deposition,m4'108 for example. Zr-bisphosphonate and Zr—bisphosphate (ZP) chemistry is again suitable for this work as it is a type of self-assembly that has been used in the formation of interfacial ad- layers for some time.79’8]'84’8738'109'I 15 The primary motivation for the use of ZP chem- istry is that the Zr—phosphate/phosphonate association is energetically very favorable,80 resulting in an essentially irreversible complexation that is characterized by fast reaction kinetics. It has been demonstrated recently that ZP complexation chemistry can be used to form phosphocholine lipid adlayers on surfaces terminated with a zirconium phosphate moiety. 1 16 In that work, 1,2-dimyristoyl-sn-glyceroé3'-phosphocholine (DMPC) com- plexed with Zr+4 bound to surface phosphate functionalities to produce a self-assembling lipid adlayer.116 31P NMR data demonstrated that the lipid-interface interaction was through the phospholipid headgroup phosphate moiety. The resulting adlayer remained intact after removal from the solution in which it was deposited, and was stable in air, suggesting its use as a foundation for biomimetic films. The only phospholipid inves- tigated in that work was DMPC. Because of the compositional complexity of plasma membranes,117 it is of interest to understand how the structure of the phospholipid head- group influences the self-assembly of lipid-ZP complexes at planar interfaces. In addition to the importance of binding phospholipids to interfaces as a step in the creation of biomimetic interfaces, understanding such interactions may have immedi- ate utility in characterizing lipid profiles in selected biological systems. Specifically, if an interface can be identified that binds one or more types of phospholipids selectively, its use would facilitate the rapid characterization of plasma membranes. The initial step 43 in this endeavor is to determine what intrinsic, chemical selectivity for phospholipids is manifested by the ZP interface. Phospholipid headgroups vary significantly in size, po- larity, and charge. These factors can and will affect the binding affinity of the phosphate moiety to D“, and the chemical headgroup-dependent binding efficiency of selected lipids is discussed in this work. The use of gold-thiol self-assembled monolayer chemistry 95,118-125 to build a monolayer on a gold surface that can be modified subsequently to bind selected phos- pholipids was utilized again in this study. Au substrates are first exposed to 6-mercapto- l-hexanol to form a self-assembled monolayer (SAM), followed by the reaction of the SAM terminal -OH group with POCl3, H20 and Zr+4.80’87’89 The result- ing zirconated surface has been shown to bind DMPC,l l6 and the focus of this work is on under- standing the chemical and steric factors that are important to this lipid- binding process. Exam- ined here are the affinity of lipids possessing the R: —H DMPA R: \/\,!.0/ DMPC R: m3», DMPE O DMPS R Fig. 3-1: Phospholipid headgroups used as well as the overall acyl chain. Base acyl chain does not change for each lipid with the R group being the different headgroups. 44 selected headgroups, including l,2-dimyristoyl-sn-glycero-3-phosphatidic acid (DMPA), l,2-dimyristoyl-sn-glycero-3-phosphocholine (DMPC), l,2-dimyristoyl-sn-glycero-3- phosphoethanolamine (DMPE), l ,2-dimyristoyl-sn-glycero-3-[phospho-rac-( 1 - glycero 1)] (DMPG), and l,2-dimyristoyl-sn-glycero-3-[phospho-L-serine] (DMPS). The structures of these lipids are shown in Fig. 3-1. Acyl chains for all of the lipids were C14 with no unsaturations, allowing evaluation of the role of lipid headgroup structure on the surface- binding process. The results for lipid-binding to a zirconated surface are presented here, including cyclic voltammetry, time resolved ellipsometry, and water contact angle data to elucidate the formation and to a limited extent, the organization of the monolayers. The data are consistent with structurally-based expectations, where steric factors and intermolecular interactions such as hydrogen-bonding can play a significant role in mediating surface- binding phenomena. 3-2: Experimental Set up and Materials Used Materials utilized in these studies include the following and are similar to those used in the work discussed in chapter 2. They include: 1,2-dimyristoyl-sn-glycero-3- phosphatidic acid (DMPA, monosodium salt), l,2-dimyristoyl-sn-glycero-3-phospho- choline (DMPC) dissolved in chloroform, l,2-dimyristoyl-sn-glycero-3-phosphoetha- nolamine (DMPE), l,2-dimyristoyl-sn-glycero-3-[phospho-rac-(1-glycerol)] (sodium salt) (DMPG), and l,2-dimyristoyl-sn-glycero-3-[phospho-L-serine] (sodium salt) (DMPS), all in a mixture of chloroform, methanol and water, and are obtained from Avanti Polar Lipids, Inc. The solvents acetonitrile, ethanol (100%) and ethyl acetate, as well as the reagents 6-mercapto-l-hexanol, phosphorus oxychloride (POC13), zirconyl chloride octahydrate (ZrOC1208 H20), 2,4,6-collidine, potassium ferrocyanide trihydrate, 45 lithium perchlorate, hexaminerutheniumflll) chloride, and potassium chloride, are all ob- tained from Sigma-Aldrich in the highest purity grade available. 18 M0 Water was used for all of the experiments and all reagents were used as received, without fiuther pmifica- tion. The instumentaion again was the same as the previous study as well and includes: a CH Instruments 650 electrochemical analyzer for electrochemical analysis, a J. A. Woollam Co., Inc. spectroscopic ellipsometer model EC110 with a wavelength range of 185-1100 nm, utilizing 44 wavelengths simultaneously for optical ellipsometry measure- ments, and an ACT Products Inc. VCA 200 video contact angle system for all water con- tact angle measurements. Unless noted otherwise, experiments were performed at 20°C . Two electrochemically active probes were used to characterize the interfaces; K3Fe(CN)6-3 H20 (1.32 mM) in 0.1 M LiClO4 and 'Ru(NH_,))6Cl3 (1.00 mM) in 0.1 M KCl. These two probes were chosen because of their different electron transfer kinet- ics and ionic charges.88 Cyclic voltammetry (CV) was performed with each probe being cycled three times at a scan rate of 0.1 V/s. The F e(CN) 63 44' probe was scanned from -0.1 to 0.5 V and the Ru(NH 3H?“ probe was scanned from -0.4 to 0.1 V vs. Ag/AgCl 3% using a Pt counter electrode. It should be noted that due to calculation error, 10 times the iron probe concentration was used initially leading to larger then expected currents. 3—2. 1: Sample Preparation Gold substrates were prepared using a procedure described previously.89 Briefly, the substrates are rinsed with water and ethanol, cleaned in a UV—cleaner for 15 min., then exposed to 10 mM 6-mercapto-l-hexanol in ethanol for 6 hrs. The resulting interface is then rinsed with ethanol and ethyl acetate, then dried under a stream of N2(g). For Zr- modified interfaces, the 6-mercapto-l-hexanol monolayer is reacted with POCl3 (0.4 mL) 46 Uidr} phosr N,(gl zncor soluni VCSlCIt r0f0rn' sohnio dncdli ndclTi 11111th 1': lipids p; $011579. lSpassc lfincsro in dry acetonitrile (10 mL), and catalyzed with 2,4,6-collidine (0.4 mL) for 3 hrs. The phosphate-modified monolayer formed is then rinsed with ethanol and water, dried with N2(g), and exposed to 5 mM ZrOCl2 in a 60:40 ethanol/water solution for 12 hrs. The zirconated monolayer is then finaly dried under N2(g), then is ready for exposure to the solution containing lipid unilamellar vesicles. Unilamellar vesicles of each lipid are prepared as described previously.60 The vesicles used here are comprised of phospholipid only, with no other constituents. Chlo- roform or the chloroformzmethanolzwater ternary system is first evaporated from the lipid solution using a N2 stream, and any remaining solvent was removed under vacuum. The dried lipid is next dissolved in a 10 mM tris(hydroxymethyl)-aminomethane hydrochlo- ride (Tris®, Aldrich) pH 7.5 buffer solution to a final concentration of 1 mg/mL. The so- lution is treated with five freeze-thaw-vortex cycles to ensure thorough suspension of the lipids prior to extrusion.90 To form unilamellar vesicles with a narrow size distribution, a syringe-based mini-extruder is used (Avanti Polar Lipids, Inc.).9l'93 The lipid suspension is passed through a polycarbonate filter with an average pore diameter of 400 nm eleven times to produce unilamellar vesicles of that diameter. Planar lipid adlayers are formed by spontaneous fusion of unilamellar vesicles.75 The modified gold substrates are placed in a custom-made Teflon® flow cell (Fig. 2-1) with an approximate volume of 1 mL that has been described in detail in the previous chapter. 1 16 The flow cell is used to ensure the lipid vesicle suspension was in full contact with the substrate during adlayer formation. Tris® buffer is flowed over the substrate at ca. 5 mL/min. prior to lipid deposition, followed by the lipid suspension, flowed through the cell at the same rate, until the buffer solution is displaced. The vesicle-containing solution is then allowed to remain in contact with the substrate for a fixed period of time. For the electrochemical experiments, the vesicle-containing solution is in contact with the 47 substrate for at least six hours to allow for the maximum adlayer formation. After expo- sure to vesicle solution, the substrates are washed with water at the same flow rate. The water is then aspirated from the cell at 1 mL/min, and the substrate is removed from the flow cell and allowed to air dry by hanging vertically. 3-3: Individual Lipid Results The primary purpose of this work is to evaluate the affinities of selected phospho- lipid headgroups for a zirconated surface and thus gauge the extent to which lipid adlayer self-assembly occurs. Considered first is the experimental data for the different phospho- lipids individually, then by comparing these results, it is possible to assess which factors are of primary importance in determining lipid-interface interactions. 3-3.1: DMPA The planar substrate used in this work are Au that is first reacted with a 6-mercap- to-l-hexanol to form a hydroxyl-terminated SAM. The resulting interface is subsequent- ly reacted with ZrOCl2 to produce a zirconated surface. The zirconated substrate is then exposed to a solution containing DMPA vesicles, and optical ellipsometry is used to mea- sure the thickness of the resulting adlayer ex situ as a function of vesicle exposure time. DMPA exhibits a rapid build-up to a thickness of 3012 A (Fig. 3-2a), consistent with the formation of a lipid monolayer.96 Once the lipid adlayer formed, the thickness remained constant as a function of vesicle exposure time. Water contact angle measurements are performed on these same interfaces, providing some insight into their polarity and homo- geneity.72’97'99 The water contact angle for a DMPA adlayer is 104°, with a hysteresis (the difference between advancing and receding contact angles) of ca. 7° (Fig. 3-2b). The value of 104° indicates that the chemical functionality of the adlayer in contact with the 48 40 a 300 C 35 ' T “A 200 . ' E ’ 30 _ Vixi 8 [00 _ /,# < R 3 <_‘ A i 1].;// 1 / .4: 25 - v 0— / , / ’ v b /' / . / , / . a 20 - a .100 - / _. D E / / .' r) ”T / . § 15 - 'c -.00 » \\ ,1 O 0— ix: . .... c .. \ _, J: 10 — o -300 - ... I: 5 - 8 400 - 0 I L (w I l A I ”5 50 mo 150 200 250 300 .04 03 02 —Ol 00 01 0.2 IIO - f3 b n" '00 d _ . E _ i05 - 4“; s s , g - , a -- 7 Q A £9 1’ ,,,,,, " ...... < 75 ch '00 ' Y . Y 1 o ‘f ’ v '0 95 .1 _.T J... v .».Lr I I 3:? 50 _ a) 90 f3 w 30 85 5 g '0 25 » ,_, so ‘5 o 0 5 , ,4” N 75 > I: 0 79,; «2 ’:v ’ H f ::_::.:.—3/"“ S 70 - a r! 7 . ° 65 - .25 60 I I l I J 1 A I A l I I I I 0 50 I00 i50 200 250 300 or 0.0 0.1 0.2 0.3 0.4 0.5 time (min.) potential vs. Ag/AgCl (V) Fig. 3-2: DMPA results. A) Ellipsometric thickness measurement in time. B) Water contact angle in time with the solid line being the initial drop angle, the dashed line is advancing, and the dotted line is receding. C) CV of DMPA monolayer (solid line) and blank substrate (dashed line) when exposed to Ru(NH ) Cl . D) CV of DMPA mono- layer (solid linc) and blank substrate (dashed line) when exposed to K3Fe(CN)6. water droplet is nonpolar, suggesting the lipid acyl chains are the outermost component of the adlayer. By comparison, the blank ZP surface is characterized by a contact angle of ca. 72°. The hysteresis seen for the DMPA adlayer suggests modest spatial heterogeneity in the organization of the adlayer. Cyclic voltammetry of electroactive probes in solution over the adlayers is also performed to evaluate interface uniformity. Two electrochemical probes are utilized, with CV data for Ru(NH3)6Cl3 (Ru probe) shown in Fig. 3-2c and K3Fe(CN)6 (Fe probe) in Fig. 3-2d. For adlayers and probes where redox waves were detectable, the reactions are found to be reversible, redox waves for both probes, with peak splitting for both probes 49 being consistent with literature reports.57 For the zirconated adlayer, probe access to the electrode is limited for the Ru probe and is heavily attenuated for the Fe probe. The measured peak splitting is found to be 128 mV with a midpoint potential of 190 mV vs. Ag/AgCl for the Fe probe with the ZP treated 6-mercapto-1-hexanol SAM surface. For the DMPA-terminated interface, it is observed that little to no Fe probe electrochemi- cal response is found. The Ru probe yields a peak splitting of 164 mV and a mid-point potential of -1 70 mV vs. Ag/AgCl for the ZP treated 6-mercapto-1-hexanol SAM. The Ru probe electrochemical signal is attenuated slightly for the DMPA-terminated interface, yielding a splitting of 212 mV and a midpoint potential of -l67 mV vs. Ag/AgCl. These results are compiled for both probes and all adlayers studied in Table 3-1 which is dis- cussed later in this chapter. The splitting data for all measurements suggests interfacial adlayer mediation of the probe electron transfer kinetics. A peak splitting of 59 mV is expected for a fiilly reversible reaction with fast electron transfer kinetics. This increase in peak splitting data could also contain a contribution from hindered diffusion, but this is believe to be less likely to account for these reported findings than mediation of electron transfer kinetics by the adlayer. This statement is based on the fact that the diffusional properties of both probes are similar, and the electron transfer kinetics for the Ru probe are somewhat faster than for the Fe probe. If hindered diffusion accounted for these data, greater similarity in the peak splitting for both probes, and a dependence of the peak split- ting on adlayer identity which follows the ellipsometry and/or contact angle data would be expected. This is not observe for either of these trends (Table 3-1). For the Ru probe, it is clear that the lipid adlayer is slowing the electron transfer kinetics to a greater extent than for the 6-mercapto-1-hexanol SAM-terminated interface alone. The electrochemi- cal and water contact angle data point collectively to the DMPA monolayer containing a measurable quantity of defects. There is precedent for the formation of a lipid monolayer 50 mm -."..._-. ' at a zirconated interface. Recent work on DMPC interactions with a zirconated interface show that a monolayer does form, with the dominant chemical interaction being shown by 31P NMR to be the complexation of the Zr+5 by the lipid phosphocholine group. 1 16 3-3.2: DMPC DMPC forms a stable adlayer on the zirconated Au surface, following a ca. 20 minute equilibration time. 1 '6 The ellipsometric thickness of the DMPC adlayer was measured to be 28i3 A,116 (Fig. 3-3a) similar to that found for DMPA and consistent with a lipid monolayer being deposited at the interface. The water contact angle for the result- 50 » a 300 C l .r‘ l 45‘ it 5 200+ 40 \ 4 - [T < 100- / \ \g, 35 i 1 _, ,7, A l, ‘ V .7 / / ’ o< 30 , ‘ k I// ” } b 0» / / v ,x \ \ Y ”v.1. a 25 i i . 5 -mo- 5. , o x A .5 20- , “U ,,00_ / O—l a \ ' / O : -‘\\ / .— I5 - 0 \ 74/ '5 I: .300~ l0 - 8 5 _ .400 ~ 0 A I l I I I l l I Sm L I I I A 1 l 5 0 so 100 ISO 200 250 300 350 400 .0.4 03 .02 -0.I 0.0 0.1 0.2 iio m" 100 d . E _ I 105 Q 1" 2:. we 2% .. .. . t v 3 ‘ i .33 50 Q) 90‘ K m '3“) 35 "i 5 5 ‘ l! i ‘ . 1: 25- 80 - H 5 E «r 75- l‘: 0- . 4mg a" " "‘ _. .7. 7,, r:;~‘;—‘—"":: , 'r’ 8 7o— 3 , ° 65- 25 (,0 1 I L I 1 r r I J J r I I L I —50 0 50 100 150 200 250 300 350 400 -0.l 0.0 0.: 0.2 0.3 0.4 0.5 time(min.) potential vs. Ag/‘AgCl (V) Fig. 3-3: DMPC results. A) Ellipsometric thickness measurement in time. B) Water contact angle in time with the solid line being the initial drop angle, the dashed line is advancing, and the dotted line is receding. C) CV of DMPC monolayer (solid line) and blank substrate (dashed line) when exposed to Ru(NH3) 6C13. D) CV of DMPC mono- layer (solid line) and blank substrate (dashed line) when exposed to K3Fe(CN)6. 51 ing interface is ca. 90° (Fig. 3-3b) with a hysteresis of 5°. While the water contact angle is slightly less than that seen for the DMPA adlayer, it is consistent with the lipid acyl chains being the moiety which defines the outer portion of the adlayer. The comparative- ly low hysteresis is consistent with the existence of a slightly more homogeneous adlayer than was seen for DMPA, in keeping with the electrochemical data. Cyclic voltammetry data for the DMPC interface using the Ru and Fe electrochemical probes (Figs. 3-3c and 3-3d, respectively) indicate that the DMPC adlayer effectively blocks access of the Fe probe to the Au surface, while allowing the Ru probe access. For the Ru probe, the ZP treated 6-mercapto-l-hexanol interface yields a splitting of 166 mV with a midpoint po- tential of -164 mV vs. Ag/AgCl (Fig. 3-3c). For the DMPC-terminated interface, we ob- serve 167 mV of peak splitting and a midpoint of -1 70 mV vs. Ag/AgCl. Both interfaces mediate the electron transfer kinetics of the probe to the electrode, and given the similar- ity of the data, it appears that DMPC does not influence the organization of the underly- ing 6-mercapto-1-hexanol SAM adversely, as is seen for the DMPE-terminated interface (vide infra). These electrochemical findings are also consistent with a well organized adlayer, because the Ru probe is energetically favored over the Fe probe in terms of being able to penetrate the interfacial adlayer.88 3-3.3: DMPE Ellipsometry and water contact angle measurements for the zirconated substrate exposed to DMPE unilamellar vesicles yielded data consistent with a somewhat less organized adlayer than that of either DMPA or DMPC. The ellipsometric thickness was measured to be l6il A (Fig. 3-4a) suggesting either sub-monolayer coverage or a relatively uniform lipid adlayer exhibiting a ca. 45° tilt angle with respect to the surface normal. Water contact angle data for the DMPE adlayer reveals a hydrophobic interface 52 40 a A 300 C 35 - g 200 - \ /./’ \._ 30 — i 100 . / ; , / '3; 25 E‘ 0 " / .-"/ w 20 E .100 - ,/ .7/ 8 o /' .' / E "U , /’ ' r / . .3 , _ x ,1 g 15 T t: 200 “ \ 77/ - ‘ a) \ .' 5 l0 — . E .300 — f 5 ° 400 0 I r I r_ L . I 1 5“) I r I I J L J 0 so l00 150 200 250 300 .04 —0.3 -0.2 -0.l 0.0 0.1 0.2 NS - b A d 110 - NE 100 - T 105 - ‘ at r»? 2 ’5 ”A A V, i A 75 _ no '00 v7? . iv "' J? J; 3 .8 95 3 .Y T : , y . t v dI *1 I Q 50 _ a) 90 ‘ o . m '3'” 85 - ' ‘ { ° 1 ’i’ 5 g 1 . ’ l -o 25 — ,. so . I 0-0 _ .7/ e E S 75 ' I: 0 - ,4 A». ,L 7:153” 5 70 - g 2'" 7' ’ o . 65 - 25 ’ w I I I I I I l I l I I I I 0 so 100 150 200 250 300 —0.l 0.0 0.: 0.2 0.3 0.4 0.5 time (min.) potential vs. Ag/AgCl (V) Fig. 3-4. DMPE result. A) Ellipsometric thickness measurement in time. B) Water contact angle in time with the solid line being the initial drop angle, the dashed line is advancing, and the dotted line is receding. C) CV of DMPE monolayer (solid line) and blank substrate (dashed line) when exposed to Ru(N H3) 6C13. D) CV of DMPE mono- layer (solid line) and blank substrate (dashed line) when exposed to K3Fe(CN)6. with a contact angle of ca. 103° (Fig. 3-4b). The contact angle hysteresis for the DMPE interface is >10°, a value consistent with a heterogeneous surface and arguing for sub- monolayer coverage. The ellipsometry and water contact angle data are both consistent with a DMPE interfacial adlayer that is measurably less well organized than either the DMPA or DMPC adlayers, and likely present as a partial adlayer. Cyclic voltammetry data for the DMPE interface (Figs. 3-4c and d) reveal a more prominent Ru redox wave than was seen for DMPA or DMPC, but for the Fe probe, no signal is seen for either the ZP treated SAM or the DMPE-terminated interface. The Ru probe data exhibit a peak splitting of 164 mV and a midpoint potential of -l70 mV vs. Ag/AgCl for the ZP treated 53 SAM, and a peak splitting of 151 mV with a midpoint potential of -124 mV vs. Ag/AgCl for the DMPE-terminated interface. The addition of the DMPE adlayer apparently facili- tates the electron transfer kinetics compared to the 6-mercapto-l-hexanol SAM, implying that the DMPE disrupts the organization of the underlying SAM. These data point col- lectively to the existence of a heterogeneous interface, consistent with partial monolayer coverage by DMPE. 3-3.4: DMPG Adlayers formed using DMPG exhibited properties that are significantly differ- ent from those observed for DMPA and DMPC adlayers, and more akin to that seen for 40 a ”A 35 E 30 i V A ‘35 25 *3." m 20 ' 8 8 13 5 I5 *" ' p s S w 5 5 O 0 I L l A l L ‘ L l J 0 50 mo I50 200 250 300 IIS b A (‘0 no -, 5 I05 -rw-I'” 2 ab I00 WY .................................... g e 8 95 ‘i Q 3‘ V 2 m i a— DD 85 E 5 so 1:: ‘5 E 3 75 o g 70 5 O 65 60. 4 L l A l e + l ‘ l ‘ l ‘ 0 50 I00 I50 200 250 300 time (min.) 300 200 ~ _ c I I A 4 OJ 0.2 * d flat-z” ‘* "r“ I . r A J _ n A 1 1 + J # -0.| 0.0 or 0.2 0.3 0.4 05 potential vs. Ag/AgCl (V) Fig. 3-5: DMPG results. A) Ellipsometric thickness measurement in time. B) Water contact angle in time with the solid line being the initial drop angle, the dashed line is advancing, and the dotted line is receding. C) CV of DMPG monolayer (solid line) and blank substrate (dashed line) when exposed to Ru(NH3)6Cl3. D) CV of DMPG mono- layer (solid line) and blank substrate (dashed line) when exposed to K3Fe(CN)6. 54 Matt “7'— ”I m — DMPE. The ellipsometric thickness of the DMPG adlayers is 1628 A, requiring an hour to form (Fig. 3-5a). This thickness value is similar to that seen for DMPE and is consis- tent either with a partial, spatially heterogeneous adlayer, or an adlayer that displays a ca. 45° tilt angle relative to the interface normal for its acyl chains. The water contact angle value for the DMPG adlayer is ca. 101° (Fig. 3-5b), with a hysteresis of ca. 10°. The interface is clearly nonpolar but the magnitude of the hysteresis indicates the presence of spatial heterogeneity in the adlayer, consistent with fractional coverage. Cyclic voltam- metry data for the Ru and Fe probes indicate that the addition of the lipid adlayer can diminish the organization of the thiol SAM (Fig. 3-5d). It is found that for the Ru probe there is a peak splitting of 166 mV for the ZP treated SAM, with a midpoint potential of -l69 mV vs. Ag/AgCl. With the addition of the DMPG adlayer, an increase in splitting to 254 mV and a midpoint potential of -l43 mV vs. Ag/AgCl are observed. It is clear that the presence of the DMPG adlayer is mediating the electron transfer kinetics at this interface, but perhaps more important is the observation that the magnitude of the cur- rent is the same for both interfaces. In all cases, the Ru probe has significant access to the electrode. For the Fe probe, it is found that the thiol adlayer yields a small Fe redox wave, with a peak splitting of 152 mV and a midpoint potential of 225 mV vs. Ag/AgCl. In this case, the presence of the DMPG adlayer hinders access of the Fe probe to the electrode surface. These data indicate, collectively, that there can be significant access of the electrochemical probe to the electrode, implying a disordered, incomplete adlayer that can influence the organization of the 6-mercapto-1-hexanol SAM is formed by DMPG. 3-3.5: DMPS DMPS is also used to evaluate its propensity for adlayer formation on a zirconated interface. Ellipsometric thickness measurements yielded erratic results with data ranging 55 thickness (A) contact angle (deg) a 50 100 150 200 250 300 . t l 1 50 I00 I50 200 250 300 time (min.) current density (uA/cmz) § § current density (uA/cmz) at" N ~l VI LII I N La 3%. O I as: g r O .,1 -0.4 -0.3 -0.2 0 I 0.0 0 I 0.2 . (’7’ -0. I 0.0 0. I 0.2 0.3 0.4 0.5 potential vs. Ag/AgCl (V) Fig. 3-6: DMPS results. A) Ellipsometric thickness measurement in time. B) Water contact angle in time with the solid line being the initial drop angle, the dashed line is advancing, and the dotted line is receding. C) CV of DMPS monolayer (solid line) and blank substrate (dashed line) when exposed to Ru(N H3) 6C13. D) CV of DMPS mono- layer (solid line) and blank substrate (dashed line) when exposed to K3Fe(CN)6. from 4 A to ca. 10 A (Fig. 3-6a). These data indicate the formation of what could opti- mistically be termed a partial adlayer. Water contact angle data on these same interfaces indicated the existence of a hydrophobic adlayer being formed (Fig. 3-6b), with a contact angle of ca. 100° and a hysteresis of ca. 15°. Such a large hysteresis indicates a highly nonuniform interface, consistent with the ellipsometric thickness data. The interaction of DMPS with a zirconated interface is not sufficiently favorable to produce an identifiable adlayer, and what does adsorb onto the zirconated interface does so only afler extended exposure time. Cyclic voltammetry data using the Ru and Fe probes indicate an adlayer with fractional interface coverage. Redox waves are seen for the Ru probe but not for Fe probe (Fig. 3-6c,d). For the ZP treated SAM, the Ru probe produces a peak splitting 56 of 166 mV and a midpoint potential of -l7l mV vs. Ag/AgCl. With the addition of the DMPS adlayer, the splitting is 162 mV and the midpoint potential shifts to -128 mV vs. Ag/AgCl. Given the comparative magnitudes of the voltammograms for the SAM and DMPS interfaces and peak splitting data, it is possible that lipid deposition disrupts the organization of the zirconated SAM. 3-4: Overall Lipid Observations With an overview of the ellipsometry, water contact angle and CV data for each of the lipid adlayers that have been studied, the next consideration is how these data com- pare to one another. Specifically, of concern is how the phospholipid headgroup identity mediates lipid adlayer formation. All data collected for each phospholipid headgroup is overlaid in Fig. 3-7 in order to gain further insights on the variations between head- groups. The data point to the division of the lipids into two broad categories; those phos- pholipids characterized by comparatively strong headgroup interactions with the zirc0n- ated interface (DMPA, DMPC), and those phospholipids that interact only to a limited extent with the zirconated interface (DMPE, DMPG, DMPS). This draws on the asser- tion that there is both a steric component and an “interaction” component (e. g. hydrogen bonding) that account for these findings. A word is in order at this point on the formation of lipid monolayers rather than bilayers in this study. The physical picture for monolayer formation as discused in chap- ter 2, is believed to proceed as follows: initially, vesicle fusion proceeds on the zircon- ated interface, and once the phospholipid headgroups of the bottom leaflet of the planar bilayer contact the zirconated interface, they bind in an essentially irreversible manner, provided the headgroup structure allows for this interaction to occur. The bottom leaflet is thus not free to execute translational motion. It is believed that the interface coverage 57 fl) - 60 — 9.9 ‘0- ” ~ I '. . ‘ A 20— % If M - 5 a .. : ,.. E 20_ 3 _0 g _ : g m . 2 _, f 0 Mb .0 . “ 1:3} - IIIIII ‘ . g A.“ (to- 10 - Ll: ..... .. . 5 . 0 —so. -IOO ‘ 0.2 d 20- j 3 A '5 // 3 3 , v «- IO 1/ '3 I / a g 5 ' l/ / .1 0 / ‘,/.. -_'_‘_'.. ..///. ..... ,»._,,, 0 . .. .. ;§:—.":‘-";‘v,;;-—-, 5133:5557: -~ 0 -s. v/ 0 50 100 I50 21” 250 300 4).] 0.0 0.1 0.2 03 04 0‘5 - ( I ) potential“. Ag/Agcl (v) Fig. 3-7: A) Overlay of all thickness measurements in time with DMPA (solid line), DMPC (dashed line), DMPE (dash-dot-dot-dash line), DMPG (dotted line), DMPS (dash- dot-dash line). B) Overlay of initial drop water contact angle measurements with DMPA (solid line), DMPC (dashed line), DMPE (dash-dot-dot-dash line), DMPG (dotted line), DMPS (dash-dot-dash line). C) Overlay of Ru(NH3) 6C1, probe CVs with Blank (solid line), DMPA (dashed line), DMPC (bold solid line), DMPE (dotted line), DMPG (dash- dot-dot-dash line), DMPS (dash- dot-dash line). D) Overlay of K F e(CN) probe CVs with Blank (solid line), DMPA (dashed line), DMPC (bold solid line), DMPE (dotted line), DMPG (dash-dot-dot-dash line), DMPS (dash-dot-dash line). of lipids is heterogeneous initially, and the interstices between covered regions remain open if they are smaller than the area required for vesicle adsorption and fusion. Subse- quent to initial deposition, translational motion of the top lipid leaflet, and translocation of those lipids, once they reach the edges of the bottom-leaflet islands, serves to fill in the open interstitial regions with phospholipid molecules, resulting in a lipid monolayer. In cases where the phospholipid headgroup interaction is weak (vide infra), the formation of a partial adlayer is achieved. 58 3-4. 1: Lipids with Strong Headgroup Interactions The phospholipids DMPA and DMPC both exhibit strong interactions with the zirconated interface. Both adlayers produce ellipsometric thicknesses of ca. 30 A, con- sistent with a monolayer of C14 phospholipid based on molecular mechanics calculations and experimental data.96 While all phospholipids yielded a water contact angle of 2 90°, implying a very hydrophobic interface, contact angle hysteresis has proven to be more informative in terms of adlayer quality. The magnitude of the contact angle hysteresis scales with interface heterogeneity, with a hysteresis of 5° or less implying a compara- tively homogeneous interface, and a hysteresis of 10° or more, implying significant struc- turalheterogeneity.126 The DMPA and DMPC adlayers yield the lowest contact angle hysteresis of all the lipid adlayers studied, and these findings are in qualitative agreement with the electrochemical data, which indicate that these two adlayers are the least perme- able to the Ru(NH3)6Cl3 and K3Fe(CN)6 probes. 3-4.2: Lipids with Weak Headgroup Interactions The second group of lipids, DMPE, DMPG and DMPS, are all characterized by comparatively thin (ca. 15 A) adlayer thicknesses, and contact angle hysteresis of 10° or more. The CV data point to a measurable ability of the electrochemical probes to un- dergo electron transfer with the Au electrode. In some cases (e.g. DMPE), the addition of the lipid adlayer appeared to diminish the integrity of the underlying SAM. Taken col- lectively, the data point to a heterogeneous interface with sub-monolayer coverage of the lipids. Considered next is the physical and chemical basis for these findings. 3-5: Discussion There is an important feature of the electrochemical data which can be understood 59 in terms of access of the electrochemical probes to the zirconated layer (Table 3-1). It is noted that the midpoint potential for all of the zirconated interfaces is located at ca. -170 mV vs. Ag/AgCl. For a zirconated interface, it is expected that the Zr+4 will be coordi- nated by either OH- or Cl- ions in solution, and it is also likely that non-stoichiometric 127"” These ligands must be water will be associated with the interface terminal group. displaced upon complexation of the phospholipid headgroup to the Zr”. It has been es- tablished previously that the presence of Zr-phosphate at an electrode interface can shift the midpoint potential of a redox-active species to more positive values if the redox- Ru(N l13)63 ”2+ probe F e(CN) 63 ”l 4' probe Lipid SAM SAM+lipid SAM SAM+lipid Splitting E _ Splitting E _ Splitting E . Splitting E . (mV) "’3' (mV) m." (mV) mi' (mV) m." (FRI? 3m 2%? (€30 DMPA 164 -l70 212 -167 128 190 -- -- DMPC 166 -164 167 -170 -- -— -- -- DMPE 164 -l70 151 -124 -- -- -- _- DMPG 166 -169 254 -143 152 225 -- -- DMPS 166 -171 162 -128 -- -- -- -- Table 3-1. Electrochemical data for Ru(NH )63+/2+ and F e(CN)63'/ 4' probes for interfaces studied here. The SAM indicated in the Tab e is 6-mercapto-l-hexanol. Endpoint values are reported vs. Ag/AgCl reference electrode. active species can gain direct access to the ZP moiety. 129 These experimental midpoint potential data show no change for DMPA and DMPC interfaces and a ca. 40 mV positive shift for DMPE, DMPG and DMPS interfaces. For structurally heterogeneous, partial lipid adlayers, such a shift is expected, and this finding is consistent with the ellipsometry and contact angle results. When possible, the Zr—bisphosphate (ZP) structure will form because of the substantial thermodynamic driving force for this reaction.130 The formation of the ZP structure can be precluded by either steric interference or by competition from other intermolecular interactions. As mentioned above, it is known that ZP interfaces with Zr+4 60 " Kim I- _- I. _S E5. as the topmost layer will attract non-stoichiometric water, presumably due to the hydro- .127’128 When a ligand complexes with the Zr+4 ion, it must philic nature of the Zr+4 ion first displace the water surrounding the metal ion. If the ligand also has a propensity for interaction with water, the water may not be displaced or may be competitively bound by the ligand, leading to diminished complexation between the metal and the phosphate moiety. This situation could lead to structural disruption of the ZP interface. Consider- ing the phOSpholipids in the context of their propensity for H-bonding interactions, it is anticipated that the phospholipids DMPA, DMPE, DMPG and DMPS can all undergo extensive H-bonding with water due to the presence of phosphate, amine, hydroxyl and carboxylate moieties, respectively. In the case of DMPC, analogous H-bonding with wa- ter is hindered by the terminal trimethylamine moiety, and experimentally it is found that the existence of the choline substituent on the lipid phosphate group does not sterically preclude formation of the ZP complex. For DMPA, the formation of a ZP complex with a sterically unhindered phosphate can occur, displacing H20 in the process. For the sub- stituted phospholipids DMPE, DMPG and DMPS, the side groups are capable of H-bond formation with any water in the vicinity of the ZP group, and can thus maintain the phos- phonate moieties physically separate from the Zr+4 . For these lipids, it is possible that the formation of a partial adlayer disrupts the organization of the underlying SAM, leading to enhanced exposure of the ZP moieties to the electrochemical probes, as manifested by the observed ca. 40 mV positive shifi in the midpoint potential. For phospholipids that either do not participate substantially in aqueous H-bonding (DMPC) or for those where there is no steric issue with respect to access to the phosphate moiety (DMPA), we ob- serve comparatively strong chemical interactions with the zirconated surface, resulting in the formation of a structure that resembles a monolayer of lipid. Under these conditions, electrochemical probe access to the ZP moieties and underlying Au electrode surface is 61 precluded by the presence of the lipid and the midpoint potential seen for these inter- faces resembles that seen for the (presumably well organized) ZP-terminated interface, where the ZP moieties are likely coordinated to OH' and/or Cl'. For phospholipids with phosphate pendant functionalities capable of significant H-bonding, less well organized adlayers are observed. The data for the phospholipids that fall into this group (DMPE, DMPG, DMPS) suggest the formation of a spatially heterogeneous partial adlayer on the zirconated interface. The results obtained for the DMPS adlayer is somewhat surprising in light of the fact that Zr-phosphate-carboxylate (ZPC) complexation is known to occur.89’l3 1'13 2 While this complexation is not as energetically favorable as ZP complexation, the pres- ence of two functionalities in the DMPS headgroup offers the opportunity for multiple types of complexation, neither of which are found to contribute significantly, based on our experimental results. 3-6: Conclusions In this study the time-dependent ellipsometric thickness and contact angle of a se- ries of phospholipid adlayers bound to a zirconated interface have been measured. These data, in concert with cyclic voltammetry data using two electrochemical probes of the resulting interfaces, serve to characterize the comparative binding efficiency and quality of the lipid adlayers. The data reveal that DMPA and DMPC form the most organized adlayers with DMPE, DMPG, and DMPS producing adlayers characterized by spatial heterogeneity and sub-monolayer coverage. The adlayer properties are then related to the ability of the phospholipid substituted phosphate moiety to interact with the zircon- ated interface. Both steric issues and the propensity of the phospholipid headgroups to H-bond with water and possibly other lipids in the proximity of the interface likely play 62 - . $02“!!! amt-wtnfiflkflnflv'r~ roles in determining the quality of the phospholipid adlayers. One issue that remains to be investigated is whether or not the means of lipid adlayer deposition influences the observed interfacial film thickness and uniformity. Lipid fusion is used to effect the self-assembly process. It is possible that the pre-assembly of lipid adlayers as Langmuir- Blodgett films could influence the properties of films deposited on the same substrates used here, and an effort is underway to explore this possibility. It is clear that the understanding of this novel class of self-assembling adlayers would benefit from imaging measurements, and this is an effort that is ongoing. It is hoped that these adlayers will find use in the formation of supported lipid bilayer struc- tures where the extent of interaction between the lipid adlayer and the support can be adjusted chemically through the composition of the lipids used. It is important however, just like with the varying lipid headgroups, to explore varying substrates. By changing the metal on the surface from Zr to other various met- als, different lipid-metal interactions could be observed. This could then translate to the ability to have multi-metal surfaces that have tailored effects on mixed lipid solutions. Initially however, it is important to gain an understanding of how different metals interact with a lipid headgroup, which is discussed in the following chapter. 63 Chapter 4 Ionic Binding of Phospholipids to Interfaces: Dependence on Metal Ion Identity 4-1 : Introduction The goal of creating bilayer systems is to utilize them to simulate the plasma membrane in a manner that allows for the presence of trans-membrane proteins in their active form(s).57 Success in this area requires that the lipid bilayer and the immediate environment on both sides of the bilayer be sufficiently hydrophilic to mimic a cellular system, and this issue has led to the design of bilayer structures that reside on hydrophilic underlayers, for example.59’74’77 In addition to these structural requirements, there are the issues of lipid bilayer fluidity and the manner in which the bilayer is bound to the underlayer many of these issues already have been addressed. Simple physisorption of bilayers onto most substrates yields an interface that is not sufficiently robust to main- tain its structural integrity in the long term. It is thus important to identify ways to make more robust the lipid bilayer interaction(s) with the support on which they reside. In the previous chapters this has been explored, with this work showing that Zr+4 can interact ”6"33 and there is anecdotal evidence that Ca+2 with phospholipid phosphate headgroups, is required to achieve a high quality lipid bilayer under conditions where the bilayer is physisorbed to the interface. For these reasons it is also of interest to explore the strength of interactions between other interfacial metal ions and selected phospholipids, and these findings are reported here. To reiterate, mammalian plasma membranes are complex systems that are com- prised of more than 100 different components.1 This compositional complexity is thought to be essential for housing transmembrane proteins as well as making the bilayer 64 structure sufficiently robust that it is capable of maintaining its structural integrity upon exposure to air.75'77’1 16 It is found in the earlier work presented that it is possible to cre- ate a hydrophilic interfacial adlayer with a high density of surface hydroxyl groups, and this interface can support a physisorbed phospholipid bilayer. The resulting interface is important for the creation of a well organized and robust biomimetic interface.116 By modifying the hydroxylated interface to create a Zr-phosphate (ZP) fimctionality, it was determined that structurally robust lipid monolayers were formed because of the interac- tion of phosphocholine head groups with the surface-bound ZP functionality. The forma- tion of a Zr-bisphosphate complex was verified by 31P NMR measurements. 1 16 The use of Zr+4-phosphate/phosphonate complex formation to create organized mono- and mul- tilayer interfacial structures is well known, and the success of this approach to controlled adlayer formation is based on the essentially irreversible Zr+4 interaction with the phos- phate moieties.89’134 With the establishment of phospholipid binding to zirconated interfaces, 3 key issue to evaluate was the role of phospholipid headgroup identity in mediating the com;- plexation process. Utilizing the gold-thiol self-assembled monolayer chemistry used in earlier studiesgs’1 18"” to build a monolayer on a gold surface. This can be modified sub- sequently to bind selected phospholipids. Au substrates are first exposed to 6-mercapto- l-hexanol to form a self-assembled monolayer (SAM), followed by reaction of the SAM terminal —OH group with POCl3, H20, and then ZrOClz. It was found that phospho- choline and phosphatidic acid lipids complexed with Zr+4 ions strongly, while phospho- ethanolamine, phosphoglycerol and phosphoserine lipids did not form organized lipid adlayers.133 Contact angle and optical ellipsometry data indicate that the adlayers formed using these lipids were incomplete and spatially heterogeneous. These findings are un- derstood in the context of the propensity of the lipid headgroups to hydrogen bond with 65 water in the vicinity of the Zr-phosphate interface. Lipids with H-bonding headgroups do not complex substantially with the surface-bound Zr+4 because of competitive interac- tions with nonstoichiometric water in the vicinity of the interface. In this next group of experiments, phosphatidic acid was chosen to eliminate issues that could be related to steric contributions to ZP complex formation. 4-2: Metals for Modification of Au Substrates It is clear that, for reasons of the phospholipid headgroup structure, there is intrin- sic chemical selectivity associated with the formation of lipid adlayers in this manner. It is also important to consider whether the metal ion used in the formation of the interfacial metal-phosphate structure will play a role in mediating the interface-lipid interactions. Metal-phosphates are known for a range of metal iOns.78’135'142 Most of the metal ions tested have been divalent transition metals, and Mg2 and Ca+2 have also been used.136' 139 Following the same methods used in the modification of Au substrates with Zr”,116 the examination of the ability of several metal ions, some with biological significance, to form interfacial complexes with phosphatidic acid has been chosen. Metal ions Ca”, Mg”, Zn”, Ni+2 and Cu+2 are chosen based on their known propensity for interactions with phosphates. F e+3 was chosen because iron coordinates phosphate strongly.136 Be- cause of the oxidative instability of F e+2, it was necessary to work with F e+3 due to the fact that experiments are performed in air or in an aqueous medium where no effort had been made to deoxygenate the solution. To gain firrther insight the phosphate interactions with Cu+ was examined in an attempt to understand whether or not metal ionic charge played a significant role in the formation of the supported lipid adlayer. Since lipid inter- actions with Zr+4 have been characterized,133 the data reported here are compared to the Zr+4-modified interface. 66 pl llll The results for lipid binding to a metal modified surface include X-ray photo- electron spectroscopy (XPS), cyclic voltammetry, optical ellipsometry and water contact angle data, to elucidate the formation and, to a limited extent, the organization of the lipid adlayers. The data indicate that lipid-metal coordination is a complex process that is me- diated by the identity and loading density of the metal ion that coordinates to the surface- bound phosphate groups. 4-3: Experimental Setup and Instrumentation Utilized The lipid chosen for these experiments was l,2-dimyristoyl-sn—glycero-3-phos- phatidic acid (DMPA, monosodium salt) and was obtained fi'om Avanti Polar Lipids, Inc. The solvents and reactants: Acetonitrile, ethanol (100%), ethyl acetate, 6-mercapto- l-hexanol, phosphorus oxychloride (POCl3), zirconyl chloride octahydrate (2100; 8 H20), 2,4,6—collidine, lithium perchlorate, potassium chloride, as well as the electro— chemical probes potassium ferrocyanide trihydrate and hexamineruthenium(III) chloride were obtained from Sigma-Aldrich in the highest purity grade available. The metal salts calcium chloride (CaClz), zinc chloride (ZnClZ), nickel chloride (NiClz), magnesium chloride (MgClz) and ferric chloride (F eCl3) were obtained fi'om Spectrum Chemicals. While cupric chloride (CuClz) was obtained from J .T. Baker Inc. and cuprous chloride (CuCl) obtained from Mallinckrodt. All metal salts were purchased in the highest purity grade available and used as received. 18 M9 Water was obtained from an in-house Bam- stead system and used for all experiments. The instrumentation utilized is as follows: all electrochemical data were acquired using a CH Instruments 650 electrochemical bench. Optical ellipsometry measurements were performed using a J. A. Woollam Co., Inc. model EC] 10 spectroscopic ellipsometer with a wavelength range of 185-1100 nm, utilizing 44 wavelengths simultaneously. The 67 water contact angle measurements were performed on an ACT Products Inc. VCA 200 video contact angle system. XPS measurements were performed on a Perkin Elmer Phi 5400 instrument equipped with a Mg-Ka X-ray source. Samples were analyzed at pres- sures between 10'9 and 10'8 Torr with a pass energy of 29.35 eV and a take-off angle of 45°. The spot size is ca. 250 m2. Atomic concentrations were determined using known sensitivity factors. All peaks were referenced to the C 1 5 peak associated with adventi- tious C at 284.6 eV. Unless noted otherwise, experiments were performed at 20°C. The electrochemical measurements were performed similarly to the experiments in chapter 2 and 13.116’13 3 Two electrochemically active probes were used to character- ize the interfaces that were studied; K3Fe(CN)6- 3 H20 (1.32 mM) in 0.1 M LiClO4 and Ru(NH3)6Cl3 (1.00 mM) in 0.1 M KCl. These two probes were chosen because of their different electron transfer kinetics across alkanethiol SAMs and their different ionic charges.88 Cyclic voltammetry (CV) was performed with each probe being cycled three times at a scan rate of 0.1 V/s. The Fe(CN)63'/4’ probe was scanned from -0.1 V to +0.5 V vs. Ag/AgCl and the Ru(NH 3+/2+ probe was scanned from -0.4 V to +0.1 V vs. Ag/ 3)6 AgCl, using a Pt counter electrode. 4-3. 1: Substrate Preparation Gold substrates were prepared using a procedure described previously.89 Briefly, the substrates were rinsed with water and ethanol, cleaned in a UV-cleaner for 15 min., then exposed to 10 mM 6-mercapto-1-hexanol in ethanol for 6 hrs. The resulting inter- face was rinsed with ethanol and ethyl acetate, then dried under a stream of N2(g). For metal modified interfaces, the 6-mercapto-l-hexanol monolayer was reacted with POCI3 (0.4 mL) in dry acetonitrile (10 mL), and catalyzed with 2,4,6-collidine (0.4 mL) for 3 hrs. The phosphate-modified monolayer was rinsed with ethanol and water, dried with 68 N2(g), and exposed to 5 mM concentrations of one metal salt in a 60:40 ethanol/water solution for 12 hrs. For each metal ion used, the substrate was prepared in the same man- ner utilizing metal chloride salts (except for Zr”, where ZrOCl2 was used). The resulting metal ion-containing monolayer was dried under N2(g), then exposed to a solution con- taining a DMPA unilamellar lipid vesicles. 4-3.2: DMPA Vesicle Preparation. Unilamellar vesicles of DMPA were prepared as described previously.60 The vesicles were comprised of the phospholipid only, with no other constituents. The chlorofonnzmethanolzwater solvent system was first evaporated from the lipid solution using a N2 stream. The lipid was then exposed to vacuum to remove any remaining solvent. The dried lipid was dissolved in a 10 mM‘ tris(hydroxymethyl)-aminomethane hydrochloride (Tris®, Aldrich) pH 7.5 buffer solution to a final concentration of 1 mg/ mL. The solution was then mixed using five freeze-thaw-vortex cycles to ensure suspen- sion of the lipids prior to extrusion.90 A syringe-based mini-extruder was used to form unilamellar vesicles with a narrow size distribution (Avanti Polar Lipids, Inc)”93 The lipid suspension was then passed through a polycarbonate filter (average pore diameter 400 nm) eleven times to produce unilamellar vesicles of that diameter. 4-3.3: Adlayer Formation Planar DMPA adlayers were formed by spontaneous fusion of unilamellar ves- icles.75 The modified gold substrates were placed in a custom-made Teflon® flow cell that has been described in chapter 2.116 The flow cell was used to ensure the lipid ves- icle-containing solution was in fill] contact with the substrate during bilayer formation. Tris® buffer was flowed over the substrate at ca. 5 mL/min. prior to DMPA deposition, 69 ‘ .-_ '2"'!t'£~1t-'¢?' 931'...“ u tiflx°n*-tvf‘ul§ewcmfi_ then the vesicle-containing solution was flowed through the cell at the same rate until the buffer solution was displaced, and this solution remained in contact with the substrate for two hours. After exposure to the vesicle-containing solution, the substrate was washed with water. Following washing, the water was aspirated fiom the cell. The substrate was then removed from the flow cell and allowed to dry in air while being held vertically. 4-4: Results of DMPA Exposure The primary purpose of this work is to evaluate the interactions between the DMPA headgroup and selected metal ions bound to surfaces through a phosphate group, and thus gauge the extent to which lipid adlayer self-assembly proceeds. First the ex- perimental data for the metal ions individually is discussed, then the comparison of these results to assess which metal ions give rise to phospholipid self-assembly, and which factors are of primary importance in determining the lipid-interface interaction is looked into. 4-4.1 : Zirconium As noted above, phosphate-terminated SAMs as the substrate for vesicle deposi- tion were used. By reacting the phosphate-terminated interface with ZrOClz, it is pos- sible to produce a Zr+4-terminated surface. XPS was used to determine Zr-surface cover- age. Analysis of the ratio of ZrzAu4f concentrations yields a value of 0.34 (Fig. 4-la), which we take to indicate substantially complete surface coverage based on the known strongly favored complex formation behavior of Zr+4 with ROPO3'2.80’134 It should be recognized that this concentration ratio is not quantitative due to the fact that the signal from a monolayer (or less) of metal ions is compared to the signal from a comparatively thick Au layer, but these ratio data for the different metal ions serve as a useful com- 70 parison. The zirconated sub- strate was exposed to a solution containing DMPA vesicles, and optical ellipsometry was used to measure the thickness of the resulting lipid adlayer ex situ (30i2 A), consistent with the formation of a lipid monolayer.96 Water contact angle measure- ments were performed on these same interfaces, providing insight into their polarity and homoge- neity.72'97'99 The water contact angle for a DMPA adlayer is 104°, with a hysteresis (the dif- ference between advancing and receding contact angles) of ca. 7°. The value of 104° indicates that the chemical functionality of the adlayer in contact with the water droplet is nonpolar, con- sistent with the lipid acyl chains being the outermost component of the adlayer.133 The hysteresis seen for the DMPA adlayer sug- . xl 0A5 a 3 v -Au4f7. ZSi {>15 -Zr3pl l 1035 ‘-Zcr3p 2 F ‘ O—KLL 1S—Zr3d -CKLL 01. 5V"! W ._. 1 gm 60 _ M“ 40— /'\\ .. 2o- z. -- ------ / 3 °' / ’/ a -2o- ; 0 LZ>