«.9: a... ‘4— “an. i - .5 Ii fag)- -. e;- ’9'"; h. 4:.“ , ; ”57;”? :L‘ ‘f..'l‘!§a?:g ’ m 3~vnfiV?-‘%‘fiv;s “gr .t! :. stratum: & N. . r" .‘ g y 4 w fl-r‘vr *: .rt‘rflifirlflh “9.;qu ‘ a ”41-3. i“ v 5; 3" 45‘ ‘3'“ ' . , .. _ ‘ 2 . “3,-5.1, ' 5%:5'5'5'2552'“ *‘a'r’ 75“" fit? -:'.._...2:?~‘~ "’5'“ IE \ “33$.“ 'vfir ”.3." w 'L n . - X‘. ’. «1153:..uxé—flw fizvfimzuzggmh“ , 2%.;‘..:¢L"'2‘l 32:. 14x1" 3' ~.:fw-‘.~;..kmyn v ' A“. W ‘2.“ ,1; . n; - > - 9v 1‘“ ' AM ' uf tr.» 5% , . . . _ ‘9!" “:7 .33: "-. an. ..-, ,...V W -. A, 9... ‘1‘? ##WA‘: .. .. n.“ ‘ 4.. “3......L " 1.1:" 1.; a» $3.“- ' ,. 5“.” € ,~ ‘- 1w «rd mg.” an 7.2 “r w :m > "1.5.1; . ‘ “ "EasEC-Eh r' . 1'3“, .. («1 , 11“”:- . 1 W ‘4 , '32. . *3: :12 - Sam. “.4... 31H" "‘ “" "’ ; '4" . .. - Efifiénv'ggifix V m f , .. &”’”}Za}%“:?~ Fig“: 'i‘ a £99.; 1 m». $11, an»: 5.3. 1 ‘ 4 ‘ .3 “QE'9m \ .2.- .m-r- ..-.- .w ‘ _ w hcAfgu-iémfi. ‘1“ .31.: " . ' it??? ' v4 ,_.__ " .uits‘; A 1‘1!” 4"- ‘ on: f . X THEScs This is to certify that the thesis entitled Fundamental Studies of Organic/Inorganic Carposite Thin Films As Selective Coatings for Mass Sensors presented by Gregory 0. Noonan has been accepted towards fulfillment of the requirements for M.S. degree in Cherm stry MMMM Ugfééor professor Date @Hk/ g; /7 74 0-7639 MS U is an Affirmative Action/Equal Opportunity Institution Illilllllllllllilllllll'llliillllillll 293 01033 6778 LIBRARY MIchIgan State UnlverSIty PLACE II RETURN BOX to roman this checkout from your record. ‘ TO AVOID FINES return on or before date duo. DATE DUE DATE DUE DATE DUE MSU I'M Affirmative WM OpportunIty Infiltwon 7 7 W pus-9.1 FUNDAMENTAL STUDIES OF ORGANIC/INORGANIC COMPOSITE THIN FILMS AS SELECTIVE COATINGS FOR MASS SENSORS By Gregory 0. Noonan A THESIS Submitted to Michigan State University in partial fulfillment of the requirements for the degreee of MASTER OF SCIENCE Department of Chemistry 1994 ABSTRACT FUNDAMENTAL STUDIES OF ORGANIC/INORGANIC COMPOSITE THIN FILMS AS SELECTIVE COATINGS FOR MASS SENSORS By Gregory 0. Noonan The mass sensing properties of organic/inorganic composite films have been studied through the use of surface acoustic wave (SAW) devices ,quartz crystal microbalances (QCM), and Fourier transform infrared spectroscopy. Zirconium acetate polymer oxide glasses were prepared by hydrolysis of zirconium (IV) n-propoxide in excess acetic acid. X-ray photoelectron spectroscopy (XPS) and FTIR were used to determine structure of the films. The zirconium acetate films consist of Zr-O backbone coordinated with bridging and chelating bidentate ligands. The films responded to alcohols while rejecting alkanes. It also shows molecular sieving properties which leads to the exclusion of larger molecules from the pores of the film. Poly(dimethylmethylsilicone) and poly(phenylmethyldiphenylsilicone) films were compared using SAW devices. FTIR spectroscopy using an attenuated total reflectance attachment was also used to observe viscoelastic changes of both films. The responses of both films was correlated to solvatochromic and chromatographic data. Any deviations from the expected trends were attributed to the FTIR observed viscoelastic changes. In memory of Donald R. Shaw for his support, pride, and encouragement. iii ACKNOWLEDGEMENTS I would like to thank Dr. Jeffrey Ledford for his support and guidance in completing this research. Without his ideas, encouragement, constant focusing of my research, and unyielding patience this body of work would not have been completed. I want also like to thank past and present members of the Ledford group: Kathy, Tom, Paul, Radu, and Jefl‘ R. for both scientific and emotional support during this time. Thank you Per, Mike, Ed, Maria, and the new addition, Sara T., for your fi'iendship, basketball, and home maintenance over the last two years. I also wish to acknowledge the other graduate students Matt, Chris, Sara, Dana, Tim, and Elizabeth for the needed distractions from graduate research. I would like to thank my parents, stepparents, brother, and sister for supporting my decision to return to graduate school. I want thank the Ouimette family for their encouragement in everything I have done. I want to acknowledge the MSU Vet school contingent: Tom and Kelley, Scott, Allison, Lisa, I-Iiedi, Kay and Jeff, and Vicki and Tim for their friendship before and during my program. Thank you Joe and Sue, for a place to live and a “friendly ear” whenever it was needed. I want to recognize Jake for helping me stay sane through walks in the woods and playing ball. You truly are a “man’s best fi’iend”. Finally I would like to thank Jane, a very important friend. It is not an exaggeration to say that without you I would not be writing this acknowledgment. You made many sacrifices, but still gave me the support and encouragement I needed. I have learned many things from you about what is and isn’t important, and I hope that you are proud of me because I am proud of you. This research was sponsored in part by MARTIN MARIETTA ENERGY SYSTEMS, INC., and performed under Subcontract 19K-MP607C for the Oak ridge K- 25 Site, which is managed for the US. DEPARTMENT OF ENERGY under Contract DE-AC05-84OR21400. TABLE OF CONTENTS Page List of Tables ............................................................................................................... vi List of Figures ............................................................................................................. vii Chapter 1: Introduction to the Study of Mass-Based Chemical Sensors ....................... l 1.1. Concepts of Chemical Sensing .............................................................. l 1.2. Overview of Piezoelectric Devices ........................................................ 2 1.2.1. Piezoelectric Effect ........................................................ 2 1.2.2. Quartz Crystal Microbalance (QCM) ............................. 5 1.2.3. Surface Acoustic Wave (SAW) Devices ......................... 10 1.3. Response Mechanisms of Acoustic Wave Devices ................................. 12 1.3.1. Partitioning .................................................................... 12 1.3.2. Viscoelastic Effects on Sensor Response ........................ 15 1.4. Overview of Sol-Gel Processing ........................................................... 17 1.4.1. Sol-Gel Chemistry ......................................................... 17 1.4.2. Preparation of Metal Carboxylate Thin Films ................. 20 1.4.3. Use of Sol-Gel Materials in Mass Sensors ...................... 22 1.5. Characterization of Thin Film Sensors ................................................... 23 1.5.1. X-ray Photoelectron Spectroscopy (XPS) ...................... 24 1.5.2. FTIR/Attenuated Total Reflectance (ATR) .................... 26 1.5.3. Sensor Test Apparatus ................................................... 28 1.6. Focus of Research ................................................................................ 29 References ................................................................................ 31 Chapter 2: Structure and Mass Sensing Properties of Acetate Thin Films .................... 34 2.1. Introduction .......................................................................................... 34 2.2. Experimental ........................................................................................ 35 2.3. Results ................................................................................................. 41 2.4. Discussion ............................................................................................ 50 2.5. Conclusions .......................................................................................... 61 References ................................................................................ 62 Chapter 3: Study of Polysiloxanes as Selective Coatings for SAW-based Sensors ........ 63 3.1. Introduction .......................................................................................... 64 3 .2. Experimental ........................................................................................ 66 3.3. Results ................................................................................................. 70 3.4. Discussion ............................................................................................ 82 3.5. Conclusions .......................................................................................... 87 References ................................................................................ 88 Chapter 4: Future Work .............................................................................................. 89 4.1. Tailoring Metal Carboxylate Polymer Oxide Glasses ............................. 89 4.2. Molecular Recognition Centers in Polymer Oxide Glasses ..................... 93 References ................................................................................ 98 vii Table Table 2.1. Table 2.2 Table 3.1. Table 3.2 List of Tables Page Probe analytes. ........................................................................................ 36 Analyte responses on coated and uncoated QCM. .................................... 47 List of analyte boiling points, dipole moments, and sensor response. ....... 67 Response ratios for probe analytes. .......................................................... 76 viii List of Figures Figure Page Figure 1.1. Fundamental thickness shear mode vibration. ........................................... 6 Figure 1.2. ATR cell used for analyte/film interaction studies ..................................... 28 Figure 1.3. Gas handling and sensor test cell schematic. ............................................. 29 Figure 2.1. Schematic diagram of sensor test apparatus. ............................................ 40 Figure 2.2. FTIR spectra of (a) acetic acid, (b) zirconium (IV) n-propoxide in n- propanol, and (c) zirconium acetate film. ................................................. 43 Figure 2.3. FTIR spectrum of carboxylate region obtained for (a) sodium acetate and (b) zirconium acetate film. ................................................................. 45 Figure 2.4. Variation in QCM response (Hz/mmole) versus analyte boiling point for uncoated quartz crystal. ..................................................................... 46 Figure 2.5. Variation in QCM response (Hz/mmole) versus analyte boiling point for zirconium acetate coated quartz crystal. ............................................. 49 Figure 2.6. Variation in hydroxyl stretch intensity for methanol (—) and ethanol (---) as a function of time following analyte injection. .............................. 51 Figure 2.7. Variation in relative intensity of the hydroxyl stretch for n-propanol (—) and C-H stretch for heptane (--) as a function of time following analyte injection ....................................................................................... 52 Figure 2.8. FTIR spectra of zirconium acetate film (a) before and (b) after heating. ..... 54 Figure 2.9 Figure 2.10. Figure 3.1. Figure 3.2. Figure 3.3 Figure 3.4. Figure 3.5 Figure 3.6. Figure 3.7. Figure 3.8. Figure 3.9. Figure 4.1. Variation in normalized QCM response (zirconium acetate coated crystal/uncoated crystal) versus analyte boiling point. .............................. 59 Variation in normalized QCM response (zirconium acetate coated crystal/uncoated crystal) versus analyte molar volume .............................. 60 General structure of polysiloxane polymer. .............................................. 64 Response curve for blank SAW device ..................................................... 72 Response curve for OV-101 coated SAW device. .................................... 73 Response curve for OV-25 coated SAW device. ...................................... 74 Response ratio curve for OV-25/OV101 comparison. .............................. 77 FTIR/ATR spectra of methanol on OV-lOl. . .......................................... 78 FTIR/ATR spectra of analyte induced morphology changes for nonane on OV-lOl. ................................................................................. 8O FTIR/ATR spectra of analyte induced morphology changes for pyridine on OV—25 ................................................................................... 81 FTIR/ATR spectra of analyte induced morphology changes for nonane on OV-25. ................................................................................... 83 FTIR spectrum of zirconium valerate/benzoate film. ................................ 91 Chapter 1 Introduction to the Study of Mass-Based Chemical Sensors 1.1. Concepts of Chemical Sensing A chemical sensor has been defined as a device which provides direct information about the chemical composition of its environment. It consists of a physical transducer and a chemically selective layer [1]. This definition is often narrowed further to include only devices that reversibly respond to changes in the chemical environment. Using this definition, indicator tubes and test strips are usually not considered to be chemical sensors. Also, analytical instrumentation that requires several discrete steps to obtain data are not within the standard definition of chemical sensors. In the development of chemical sensors it is necessary to consider the operation of the transducer as well as the interactions taking place in the chemically selective coating. The transducer (e. g., mass, optical, thermal, and/or electrochemical device) defines the physical or chemical parameter to be monitored while the chemical layer enhances the selectivity and sensitivity of the device. In many cases, transducers are available that inherently provide the desired level of sensitivity. Understanding chemical interactions and synthesizing chemically selective materials may be considered the frontier of sensor development. Acoustic wave or mass sensors respond to changes at the surface or coating of the device. These changes can be caused by the addition or removal of mass, changes in the density or viscosity of the coating or even changes in the conductivity of the overlayer. The large number of events that can produce a response from a mass sensor can be a disadvantage and a benefit. One major disadvantage is the difficulty in determining the fundamental processes that generate a response. However, it is the multiple response mechanism which allows the use of mass sensors for the study of a variety of analyte/film interaction types and material properties [2]. Other advantages of mass sensors are their simplicity of construction and operation, light weight, and low power requirements. Additionally, mass sensors are relatively inexpensive and have a high degree of sensitivity for a broad range of compounds. This high sensitivity can lead to problems with interferents and thus dictates the development of chemically selective coatings and a detailed understanding of analyte/film interactions. 1.2. Overview of Piezoelectric Devices 1.2.1. Piezoelectric Effect The piezoelectric effect was first reported by Jacque and Pierre Curie in 1880 [3]. They observed that with certain compounds the application of a mechanical force produced an electrical signal. A short time later it was also noted that the application of an electrical signal produced a deformation in the crystal. Since these initial discoveries, the piezoelectric effect has been documented in a large number of compounds such as rochelle salt (NaKC4H406 - 4H20), lead zirconium titanate, and or-SiOZ. To fully explore the piezoelectric effect, it is first necessary to understand the relationship between stress and strain in solids and the displacement of particles within a solid. Displacement is defined as the vector indicating the difference between equilibrium and perturbation. Strain is commonly used to define relative displacement and the two are related by: S=Vs~u (1.1) where S is the strain, u is the displacement vector, and Vs is the symmetric gradient operator. In any elastic solid the application of a strain produces a restoring force defined as stress. In the case of a piezoelectric solid the strain also generates an electric field. Stress and strain are related by: T=cS (1.2) or the inverse relationship: S =sT (1.3) T is a second rank tensor representing stress, c is a matrix of material stifi‘ness constants, and s, in the inverse relationship, is called the compliance matrix. Elements of the compliance matrix are derived from stiffness constants that depend on the symmetry of the solid. In the case of an acoustic wave it is not only necessary to consider a single force, but also a time dependent force based on the propagation of the wave. It is also important in a time dependent situation to account for losses due to dampening of the wave. Accounting for time dependency and dampening effects, equation 1.3 can be modified to: T=cS+ndS/dt (1.4) where r] is a matrix of viscosity constants. Materials which display piezoelectric characteristics do so because of the lack of a center of inversion. At equilibrium, before strain, the dipoles cancel resulting in a net charge of zero within the solid. Application of a strain distortion in the form of an extension or compression of the structure leads to a macroscopic charge separation within the solid. Of the 32 point groups known, 20 lack inversion centers, however not all 20 exhibit appreciable piezoelectric characteristics. The strain and displacement can be equated to the stress and electrical forces acting on the crystal by: D=eE+dT (1.5) s=d'e+sT (1.6) where E is the electrical field gradient, 8 is the electrical permittivity matrix, and d',d are piezoelectric strain coefficient matrices. Equations 1.5 and 1.6 show that the magnitude of the measurable piezoelectric effect is determined by the values of the piezoelectric coefficients. After considering the physical properties leading to piezoelectricity, it is important to consider the behavior of the wave created in these materials. The Christofl‘el equation is used to determine the propagation of waves within solids: k21‘~-v-= p 1.7 u 1 A2”, ( ) I‘ij is the Christoffel matrix, which is a function of the piezoelectric coupling constants and the electrical permittivity of the solid, and vi is the particle velocity. Solution of the Christofi‘el matrix leads to simultaneous equations describing the modes of propagation of the acoustic waves. In anisotropic solids three modes of mutually orthogonal wave propagation are established. Displacements along the direction of propagation are referred to as longitudinal waves, while perpendicular displacements are referred to as shear waves [4]. or-SiOz is the material most often used for mass sensor applications. Quartz is physically and chemically rugged, and is fairly inexpensive. Both quartz crystal microbalances (QCM) and surface acoustic wave (SAW) devices are employed in mass sensing applications. QCMs utilize a thickness shear mode wave to measure the mass change at the crystal surface while SAW devices depend on the propagation of Rayleigh waves. 1.2.2. Quartz Crystal Microbalance (QCM) A quartz crystal will exhibit numerous modes of vibration when placed in an electronic feedback loop. These include, but are not limited to, longitudinal (extensional), lateral (shear and flexural) and torsional (twist) vibrations. There is also the possibility of overtones and the coupling of modes. Removal of unwanted modes and enhancement of waves best suited for sensor applications are afl‘ected by the geometric cut of the crystal and placement of the electrodes. The singly rotated Y-cut enhances the shear wave mode which has the greatest sensitivity to the addition of mass (Figure 1.1). /_ / / 7‘ / / / \ / / / / / / / / / / / / / / / / / / / / / / / / / / / i" 4..— 1 / / \ \ \ ‘\ / \ —-——> / Figure 1.1. Fundamental thickness shear mode vibration. A specific class of rotated Y-cut crystals known as AT-cut are most often used for mass sensor applications. They exhibit the lowest temperature dependence near room temperature [5]. Plating of electrodes on either face of a crystal allows for the application of an oscillating electric field across the device. The electric field generates an acoustic shear wave through the crystal and produces a standing wave with nodes at either face. Since the piezoelectric material, orientation and wave mode fix the velocity, equation 1.8 defines the resonant frequencies that may exist for a given thickness. mt T (1.8) Where k is the wave vector, b is the crystal thickness, and n is an integer. This expression simply indicates that the thickness of a bulk wave device must be an integral number of half wavelengths for an acoustic wave to be resonant. In 1959, Sauerbrey developed an equation to relate the frequency change of a quartz crystal to the mass added at the surface of the crystal [6]. In developing this relationship Sauerbrey assumed that the added material simply increased the thickness of the crystal. The resulting expression (equation 1.9) does not consider the density, viscosity or shear modulus of the added layer. 2 df = —2fq Am/(pqvq) (1.9) pq and vq are the density and shear wave velocity of quartz , and fq is the firndamental frequency of the quartz oscillator. Equation 1.9 is written in many forms including: M=Cfo (1.10) where Cf =2fq2/(pqvq) (1.11) is the mass sensitivity or calibration constant of a QCM, and mf is the mass added to the crystal [4]. Although the assumptions made by Sauerbrey are physically unrealistic, experimenatal data showed the Sauerbrey equation to be accurate for layers no greater than 2% of the crystal mass [7-9]. As the use of QCMs as thickness monitors became more widespread, new expressions were developed to more accurately describe the change in frequency with the increase in mass. The first equation to improve on Sauerbrey's assumptions used the areal density (ml), defined as the mass per unit area of a deposited film, and assumed it was linearly proportional to the frequency change. v mf =.p_q_?.[_l-_.l_:l (1.12) Equation 1.12 can also be written in terms of film thickness and density as follows: Where fc is the fi'equency of the coated crystal. These equations were shown to be accurate for a 10% change in the base fi'equency of the QCM [10,11]. Although empirical data showed the above equations appeared to be more accurate, there was still no consideration of the overlayer physical properties. Lu and Lewis [12] developed an expression which considered the differences in the shear modulus, density and thickness of the quartz and film. The shear mode acoustic impedence was defined as: I Z. 49.11.)? (1.14) Of l Zf=(Pfl»lf)2 (1-15) for quartz and film, respectively. This allowed Lu and Lewis to write: p ,1 f =(pqvq/21thc)tan"{Ztan[1t(fq -f,)/fq]} (1.16) In the case where film and quartz acoustic impedences are matching, Z=1 and equation 1.15 simplifies to equation 1.12. For unmatched impedances and very small mass changes: At“ ”a <M-OH+ROH Once hydroxy groups are generated two competing reactions, olation or oxolation may occur. Oxolation is shown in reaction (2) and forms hydroxy bridges through the elimination of of solvent molecules. M—OH+M-OID(——)M—OH-M+XOH (2) X in the above reaction can be either a hydrogen or an alkyl group. Olation forms oxygen bridges through the the elimination of water or alcohol as shown in reaction 3. M—OH+M-OX——)M-—O-M+XOH (3) As in reaction (2), X can be either hydrogen or an alkyl group. The above three reactions can be described as 8N2 type nucleophilic substitutions. The mechanism involves the addition of a nucleophile to the metal center (a), followed by the transfer of a proton toward an OR group (b), and finally departure of the positively charged leaving group (c). These three steps are outlined in the reaction scheme below: H5+ XOH +MOR —> O-M-OBR —-> x0 —M-O/ -—-> X0 -M + ROH / \R X (a) (b) (c) 19 The rate of the nucleophilic reaction depends on the charge of the metal atom (5M) and the ability of the metal atom to increase its coordination sphere (N). This latter factor accounts for the general rule that the reactivity of analogous metal alkoxides increases going down the periodic table. The size and shape of the alkoxide group also affects the chemical reactivity of the metal precursor. Large or bulky ligands can decrease the reaction rate by sterically hindering attachment of the nucleophile. In addition, the ability of the ligand to form a stable leaving group will also affect the reactivity. Most of the initial sol-gel work utilized fairly unreactive silicon alkoxides. The high electronegativity, small size and fourfold coordination of silicon forces the use of acid and base catalysts to increase the rates of hydrolysis and condensation. On the other hand, the larger size, smaller charge, and ability to increase N causes transition metal alkoxides to be extremely reactive. Thus, steps to slow the rates of hydrolysis and condensation are necessary to avoid the formation of insoluble oxides. Pioneering work in the area of sol-gel synthesis was performed by Yamane [43], Yoldas [44,45], and Hench [46,47]. In general, synthesis involves the mixing of metal alkoxides with alcohol, followed by the addition of excess water and either an acid or base catalyst. This mixture is normally vigorously stirred and refluxed for several hours at slightly elevated temperatures. Through drying, the gel is allowed to form over a period that varies from several hours to several months. Yamane [43] correlated processing conditions with properties of gels and reported that gelation time decreased, while density and pore size increased with increasing ammonia concentrations. Yoldas [44,45] studied alumina gels and determined that acid catalysts must be noncomplexing and sufficiently 20 strong to produce a charge effect. Hench [46,47] showed that the addition of drying control chemical additives (DCCAs) such as, formamide, glycerol, and organic acids, could be used to control pore size distribution in gels. 1.4.2. Preparation of Metal Carboxylate Thin Films As mentioned above the reactions of transition metal alkoxide complexes are rapid compared to the rates of hydrolysis and condensation for silicon alkoxides. The increased rates are attributed to the metals' small charge, large size and ability to extend its coordination. Because of this reactivity, transition metal alkoxides can react to form insoluble metal oxide precipitates. To prevent precipitation chemical additives such as solvents, B-diketonates, or organic acids are normally used. In most instances they are nucleophilic molecules that react with the alkoxide to form new molecular species with difi‘erent structure, functionality, and reactivity. These substitutions produce a decoupling of hydrolysis and condensation leading to the growth of particles and the formation of gels. Carboxylic acids, especially acetic acid, are commonly used chemical control agents. Langkammerer [48] and Haslam [49] first reported the use of carboxylic acids chemical additives to produce metal carboxylate films. More recently carboxylic acid modification has been investigated by Berglund [50,51] and Livage [52]. A variety of reactions may occur upon the addition of carboxylic acids to transition metal alkoxides. Perhaps the most common reaction products are metal alkoxide- carboxylate species; however, additional ligands (e. g., esters, hydroxides, alcohols, and 21 water) may be attached to the metal center [40,42]. In the initial step, the alkoxide ligand is displaced, via a nucleophilic substitution by the carboxylic acid. (no)4 M + R'COOH ——+(rt'coo)1vr(oa)3 + ROH (4) The addition of water to the metal alkoxide, carboxylic acid solution produces hydrolysis shown in reaction 5: (Ir'coo)1vr(oa)3 + 1120———>M(OH)(R'COO)(OR)2 (5) The hydrolysis step is then followed by condensation: 2 M(OH)(R'COO)(OR)2 ——> (R. C00)(OR)2 M - O - M(OR)2 (R' coo) (6) which continues to form polymeric gels. In the presence of excess acid there is no evidence for the presence of esters, alkoxides or alcohols coordinated to the metal. Indeed IR spectra clearly indicate the presence of bidentate chelating/bridging carboxylate ligands [53]. The present synthetic route to metal carboxylate films increases gelation time yielding reproducible films which are fracture free and optically transparent. The incorporation of carboxylic acids also allows greater control over the chemical properties of the final products. For example the addition of acetic acid produces films with difi‘ering solubility properties than those synthesized from valeric acid [50]. Additionally, the use of 22 carboxylic acids is applicable to a wide number of transition metal alkoxide starting materials [50]. 1.4.3. Use of Sol-Gel Materials in Mass Sensors The sol-gel synthetic process is used to produce aerogels, xerogels, monoliths, ceramics and thin films. Films were initially used as protective or reflection resistant coatings. More recently researchers have taken advantage of the porous, inert properties of the films in applying them to the area of chenrical sensing. Most of the research involving sol-gel films as chemical sensors has utilized the glasses as immobilization media for optically active centers [54,55]. However, some work applying sol-gel films to mass sensing has also been carried out. Frye et a1. [56] first used a sol-gel matrix in conjunction with a mass sensor device in 1988. The silicate sol-gel coated SAW devices were not used as chemical sensors, but rather to compare N2 adsorption isotherms between bulk and thin film silicate glasses. Pore size, surface area and percent porosity were different in the two forms. Thin films were shown to be less porous than bulk samples produced using the same synthetic steps. This result is not surprising considering the more rapid solvent removal from a thin film sample. Nevertheless, these results confirmed that physical properties of polymer oxide glasses cannot be inferred from one form to another. Frye et a1. [57] again used SAW devices to study pore size distribution, surface area, diffusion coefficients and aging effects on four component sol-gel films. Aging was used to produce larger polymeric species and thus more porous films. 23 Also in 1988, Bein et a]. [58] embedded zeolite crystals into silicate glasses . The films were densified at 570 K in order to collapse the pore structure of the silicate glass. The destruction of the silicate pore structure produced a film in which only the channels of the zeolite were available for diffusion. This zeolite pore structure imparted molecular sieving properties to the silicate thin film. IR results indicated that pyridine molecules could diffuse into the zeolite and interact with the structural hydroxyls while the larger perfluorotributylamine molecule could not enter the pores. Bein er a]. [59] extended the zeolite membrane work to mass sensing by coating the films onto SAW devices and measuring their response to methanol, propanol and isooctane. The alcohols, having smaller kinetic diameters, adsorbed at levels around 500-800 ng/cm2 while absorption of isooctane was less than 5 ng/cmz. Bein's research established the ability of SAW devices to monitor analyte interactions with sol-gel films. However, it must be noted that none of the intrinsic properties of the sol-gel matrix were examined in this work. The sol-gel film served only to immobilize zeolite crystals. 1.5. Characterization of Thin Film Sensors The intermolecular interactions which contribute to sensor response occur at the surface of the thin film. Therefore, characterisation of the surface is necessary to understand the nature of these interactions. X-ray photoelectron spectroscopy (XPS) is a surface analytical technique which gives quantitative and qualitative information about the sample. Attenuated total reflectance (ATR) is a sampling technique that enhances the ability of IR spectroscopy to examine analyte/thin film interactions. 24 1.5.1. X-ray Photoelectron Spectroscopy (XPS) The firndamental process for X-ray photoelectron spectroscopy involves the irradiation of the sample with photons of energy hv (e.g., 1253.6 or 1486.6 eV). These photons produce photoelectrons with kinetic energy (BK). Since the energy of the incident photon is also known it is possible to determine the binding energy of the electron. The energy conservation equation for this process is: mus; =EK+E{(k) (1.25) where E1]- is the initial state energy, Bil-(k) is the total final state energy after ejection of the photoelectron fi'om the kth level. The binding energy of an electron is defined as the energy necessary to remove it to infinity with a kinetic energy of zero. In XPS measurements EVB(k), the binding energy of an electron in the kth level referred to the vacuum level is defined by: V . EB(k)=E{ +E;. (1.26) Ifwe substitute equation 1.25 into 1.26 we get the photoelectric equation. hv = EK = agar) (1.27) Binding energies in XPS are expressed relative to a reference level. In the case of insulating solids a well defined feature such as the C Is binding energy (284.6 eV) is often used. 25 Non-equivalent atoms of the same element can give rise to core level peaks with measurably different binding energies. These differences, called chemical shifts, arise fi'om difl‘erences in formal charge, molecular environment, and lattice energy. These difl‘erent conditions lead to dissimilar coulombic interactions between electrons and attractive potentials from neighboring nuclei. This redistributes the valence electrons and affects the potential of the core electrons. This potential change alters the binding energies of electrons. Signal intensities, Ii, are of interest in quantitative chemical analysis. Relative concentrations of elements are calculated using expressions similar to equation 1.26: m = Gbelbflbla (1.26) [B] Gagaxanalb where o is the photoelectron cross section, C is the fraction of photoelectric events which take place without intrinsic plasmon excitation, )1 is the mean fi'ee path of the photoelectron in the solid, 1] is the kinetic energy dependent spectrometer transmission function, and I is the area of the photoemission peak. XPS allows for quantitative analysis of the thin film surfaces. In our sol-gel research, these ratios allow us to compare the quantity of metal centers to backbone and ligand oxygen and the number of ligands per metal center. In conjunction with FTIR data these techniques allow for detailed characterization of our sol-gel polymer films. 26 1.5.2. FTIR/Attenuated Total Reflectance (ATR) An attenuated total reflectance (ATR) attachment utilizes total internal reflectance to obtain multiple interactions with the sample, leading to a higher signal-to-noise ratio. Total internal reflection can occur when light traveling through an optically dense medium reaches a boundary with an optically rarer medium. If the angle of incidence (Oi) is greater then the critical angle (9‘), all of the incident radiation will be internally reflected. (9c is determined from equation 1.27: sin e, = “%l (1.27) where nland 112 are the refractive indices of the optically dense and optically rare medium, respectively. The term total internal reflectance suggests that there is zero energy flux into the rarer medium, however a finite decaying electric field extends some distance across the boundary. This field is called the evanescent field or wave and its penetration depth is given as follows: dp = l (1.23) 21:01? sin2 01 41:2, )1/2 The use of FTIR/ATR to examine the structure of thin film polymeric materials is well established [60,61]. ATR has been used to depth profile and characterize polymer blends [62] and study polymer formation reactions [63]. It has also been used to 27 determine reaction rates in a wide variety of biological and chemical systems [64]. In addition, polymer coated internal reflectance elements have been used to extract solutes from aqueous solution for analysis by FTIR [65,66]. ATR has also been utilized to study the molecular level interactions of polymers and surfactants [67]. Rates of diffusion of aqueous samples into polymer films has also been measured [68,69]. All of these diflirsion systems involved difi‘usion into or from aqueous solution; none of the research focused on determining gas phase difi‘usion rates. Fieldson et al. [69] suggested the technique could be used to measure diffusion rates of gas or vapors or to examine polymer changes during difl‘usion. However, they did not establish the ability to investigate such effects. ATR cells have been used to expose vapor phase analytes to a sol-gel film. Dulebohn et al. [70] determined bonding interactions between CO and a rhodium doped titanium sol-gel film. FTIR/ATR spectra confirmed the reversible bonding of CO to the rhodium center. Through application of a pulse injection system, analyte residence time within the ATR cell should also ofi‘er information into analyte difl‘usion and analyte/film interactions. Additionally, band shifts could offer insight into the type of interactions causing retention of the analyte. Finally, viscoelastic changes or morphology changes induced by solvent vapor could affect in the infrared spectra of the polymers. Figure 1.2 shows the ATR cell used for our pulsed experiments. 28 Gas Out Gas In Temperature Control Figure 1.2. ATR cell used for analyte/film interaction studies. 1.5.3. Sensor Test Apparatus There are two common experimental methods to expose the acoustic sensor to the analyte of interest. Analyte can be introduced to the sensor cell as either a constant stream or a pulse. Constant stream methods rely on the establishment of equilibrium before sensor readings are recorded. Pulse measurements normally give a single maximum frequency shift which is recorded as the sensor response. The choice of pulse or constant flow exposure depends on the intended analytical application of the sensing device. The use of pulse injection dictates specific design considerations for a sensor cell and gas handling system. As with gas chromatography, the analyte must be vaporized upon injection into the gas stream and be carried to the sensor test cell as a plug of vapor. To insure this, a heated injector block, heated transfer line and minimum cell volume are necessary. A schematic of the gas system and sensor cell appears in Figure 1.3. 29 Gas Flow —9 Zeolite Injector N2 MassFlow Controller U L I Sensor Cell Figure 1.3. Gas handling and sensor test cell schematic. 1.6. Focus of Research In the study and development of chemical sensor coatings the ability to control the chemical and physical properties of the films is of great utility. Tailoring film structure allows us to investigate specific modifications and relate them to changes in the film/analyte interactions. Polysiloxanes are a classic organicfrnorganic polymer which is available with a large variety of chemical and physical properties. The Si-O-Si backbone can be derivatized with a large number of functional groups to enhance specific solubility 30 interactions. Additionally, data concerning chromatographic, solvatochromic, and thermodynamic characteristitcs of film/analyte interaction are available for comparison to sensor results. Such a large quantity of data is useful in helping to establish the importance of solubility interactions and viscoelastic effects on mass sensor response. More recently our interest has focused on the use of transtion metal carboxylate thin films as coatings for chemical sensors. As noted previously, very little work utilizing sol-gel derived thin films as mass sensor coatings has been reported. Considering the ability to easily alter the ligands and thus the chemical properties of the films, polymer oxide glasses may indeed offer the versatility necessary to produce selective coatings. Additionally, changes in processing conditions, such as aging, can also lead to changes in the physical properties of these films. Sol-gel films offer advantages as matrices for the incorporation of guest molecules. The optical clarity of the metal carboxylate glasses makes them ideal materials for optical sensing applications. While work on silicates as optical matrix is proceeding at a rapid rate, very few researchers have taken advatage of the chemical control possible with sol- gel transition metal film synthesis. Nearly all of the sol-gel research published to date has focused on the use of guest molecules to supply the necessary selectivity to the system. Little if any work has been done to deterrrrine how the matrix affects analyte interactions. Sol-gel synthesis offers the ability to control the synthesis from molecular precursor to final product, but researchers have not taken advantage of this control. 31 References 1. Janata, J .; Bezegh, A. Anal. Chem. 1988, 60, 62R. 2. Frye, G. C.; Martin, S. J. Appl. Spec. Rev. 1991, 26(1&2), 73. 3. Curie, J .; Curie, P. Bull. Soc. Min. Paris 1880, 3, 90. 4. Bastiaans, G. J. in Chemical Sensors“, Edmunds, G. Ed.; Chapman Hall: New York, NY, 1988, 295. 5. Lu, C. in Applications of Piezoelectric Quartz Crystal Microbalances; Lu, C.; Czandema, A. W., Eds; Elsevier: Amsterdam, 1984. 6. Sauerbrey, G. Z. Z Phys. 1959, 115, 206. 7. Mueller, R M.; White, W. Rev. Sci. Instum. 1968, 39, 291. 8. Mueller, R M.; White, W. Rev. Sci. Instum. 1969, 40, 1646. 9. Stockbridge, C. D. in Vacuum Microbalance Techniques, vol. 5; Behrndt, K. H. Ed.; Plenum Press: New York, NY, 1966, 135. 10. Behrndt, K. H. J. Vac. Sci. Technol. 1971, 8, 622. 11. Denison, D. R. J. Vac. Sci. Technol. 1973, 10, 126. 12. Lu, G; Lewis, 0. J. Appl. Phys. 1972, 43, 4385. 13. Kanazawa, K. K.; Gordon, J. G. Anal. Chim. Acta 1985, 175, 99. 14. Kanazawa, K. K.; Gordon, J. G. Anal. Chem. 1985, 57, 1770. 15. Lord Rayleigh London Math. Soc. Proc. 1887, 17, 14. 16. White, R. M.; Voltmer, F. W. Appl. Phys. Lett. 1965, 7, 314. 17. Wohltjen, H.; Dessy, R. Anal. Chem. 1979, 51(9), 1458. 18. Wohltjen, H. Sens. &Act. 1984, 5, 307. 19. King, W. H., Jr. Anal. Chem. 1964, 36, 1735. 20. Grate, J. W.; Snow; A.; Ballantine; D. 8., Jr.; Wohltjen, H.; Abraham, M. H.; McGill, R. A.; Sasson, P. Anal. Chem. 1988, 60(9), 869. 21. Grate, J. W.; Abraham, M. H. Sens. &Act. B. 1991, 3, 85. 22. Abraham, M. H.; Whiting, G. S.; Doherty, R. M.; Shuely, W. J. J. Chem. Soc. Perkin. Trans. 2 1990, 1451. 23. Abraham, M. H.; Whiting, G. S.; Doherty, R. M.; Shuely, W. J. J. Chromatogr. 1990,518,329. 24. Abraham, M. H.; Grellier, P. L.; Prior, D. V.; Morris, J. J.; Taylor, P. J. J. Chem. Soc. Perkin. Trans. 2 1990, 521. 25. Abraham, M. H.; Grellier, P. L.; Prior, D. V.; Duce, P. P.; Morris, J. J.; Taylor, P. J. J. Chem. Soc. Perkin. Trans. 2 1989, 699. 26. Abraham, M. H.; Duce, P. P.; Grellier, P. L.; Prior, D. V.; Morris, J. J.; Taylor, P. J. Tetrahedron Lett. 1989, 29(13), 1587. 27. Abraham, M. H.; Grellier, P. L.; Prior, D. V.; Morris, J. J .; Taylor, P. J .; Laurence, C.; Berthelot, M. Tetrahedron Lett. 1989, 30(19), 2571. 28. Abraham, M. H.; Grellier, P. L.; McGill, R. A.; Doherty, R. M.; Karnlet, M. J.; Hall, T. N.; Taft, R. W.; Carr, P. W.; Koros, W. J. Polymer 1987, 28, 1363. 29. Abraham, M. H.; Grellier, P. L.; McGill, R. A. J. Chem. Soc. Perkin. Trans. 2 1988, 523. 30. 31. 32. 33. 34. 35. 36. 37. 38. 39. 40. 41. 42. 43. 45. 46. 47. 48. 49. 50. 51. 52. 53. 54. 55. 56. 32 Abraham, M. H.; Grellier, P. L.; Prior, D. V.; Duce, P. P.; Morris, J. J.; Taylor, P. J J. Chem. Soc. Perkin. Trans. 2 1989, 699. Kamlet, M. J.; Abboud, J-L, M., Abraham, M. H.; Taft, R. W. J. Org. Chem. 1983, 48, 2877. Patte, F.; Etcheto, M.; Laffort, P. Anal. Chem. 1982, 54, 2239. Patrash, S. J.; Zellers, E. T. Anal. Chem. 1993, 65(15), 2055. Ballantine, D. 8., Jr.; Wohltjen, H. IEEE Ultrason. Sympos. 1988, 559. Bartley, D. L.; Dominquez, D. D. Anal. Chem. 1990, 62(15), 3069. Frye, G. C.; Martin, S. J.; Ricco, A. J. Sens. Mater. 1989, 1-6, 335. Ballantine, D. 8.; Jr.; Rose, 8. L.; Grate, J. W.; Wohltjen, H. Anal. Chem. 1986, 58(14), 3058. Grate, J. W.; Klusty, M.; McGill, R. A.; Abraham, M. H.; Whiting, G.; Andonian- Haftvan, J. Anal. Chem. 1992, 64(6), 610. Ballantine, D. S., Jr. Anal. Chem. 1992, 64(24), 3069. Brinker, C. J.; Scherer, G. W. Sol-Gel Science, The Physics and Chemistry of Sol-Gel Processing Academic Press, Inc.: San Diego, CA, 1990. Livage, J.; Sanchez, C. NewJ. Chem. 1990, 14, 513. Livage, J.; Henry, M.; Sanchez, C. Prog. Solid State Chem. 1988, 18, 259. Yamane, M.; A50, 8.; Sakaino, T. J. Mater. Sci. 1978, 13, 865. Yoldas, B. E. J. Mater. Sci. 1975, 10, 1856. Yoldas, B. E. Ceram. Bull. 1975, 54(3), 289. Hench, L. L. in Science of Ceramic Chemical Processing, Hench, L. L.; Ulrich, D. R., Eds; John Wiley & Sons: New York, NY, 1986, 52. Hench, L. L.; Orcel, G.; Nogues, L. in Better Ceramics Through Chemistry II; Brinker, C. J.; Clark, D. E.; Ulrich, D. R. Eds; Mater. Res. Soc. Symp. Proc., 73: Pittsburgh, PA, 1986, 35. Langkammerer, C. M. US. Patent No. 2 621 193 1952. Haslam, J. H. US. Patent No. 2 621 195 1952. Gagliardi, C. D.; Berglund, K. A. in Processing Science of Advance Ceramics; Aksay, A. 1.; McVay, G. L.; Ulrich, D. R., Eds; Mater. Res. Soc. Symp. Proc., 155: Pittsburgh, PA, 1989, 127. Gagliardi, C. D.; Dunuwila, D.; Berglund, K. A. in Better Ceramics Through Chemistry IV; Zelinsky, B. J. J.; Brinker, C. J.; Clark, D. E.; Ulrich, D. R., Eds; Mater. Res. Soc. Symp. Proc., 180: Pittsburgh, PA, 1990, 801. Sanchez, C.; Livage, J .; Henry, M.; Babonneau, F. J. Non-Cryst. Solids 1988, 100, 65. Doeufi‘, 8.; Henry, M.; Sanchez, C.; Livage, J. J. Non-Cryst. Solids 1987, 89, 206. Lev, O. Analusis 1992, 20, 543. Levy, D. J. Non-Cryst. Solids 1992, 147&148, 739. Frye, G. C.; Ricco, A. J.; Martin, S. J.; Brinker, C. J. in Better Ceramics Through Chemistry 111; Brinker, C. J.; Clark, D. 13.; Ulrich, D. R., Eds; Mater. Res. Soc. Symp. Proc., 121: Pittsburgh, PA, 1988, 349. 57. 58. 59. 60. 61. 62. 63. 65. 66. 67. 68. 69. 70. 33 Frye, G. C.; Martin, S. J.; Ricco, A. J.; Brinker, C. J. in Chemical Sensors and Microinstmmentation; ACS Symp. Ser. 403, 1989, 208. Bein, T.; Brown, K.; Enzel, P.; Brinker, C. J. in Better Ceramics Through Chemistry 111; Brinker, C. J.; Clark, D. E.; Ulrich, D. R., Eds; Mater. Res. Soc. Symp. Proc., 121: Pittsburgh, PA, 1988, 761. Bein, T.; Brown, K.; Frye, G. C.; Brinker, C. J. .I. Am. Chem. Soc. 1989, 111, 7640. Holland-Moritz, K.; Siesler, H. W. Appl. Spectros. Rev. 1976, 11(1), 1. Smith, A. L. Applied Infiared Spectroscopy John Wiley & Sons: New York, NY, 1979. Popli, R.; Dwivedli, A. M. J. Appl. Poly. Sci. 1989, 37, 2469. Kuhn, K. J.; Hahn, B.; Perec, V. Urban, M. W. Appl. Spectrosc. 1987, 41(5), 843. Miller, M. P. Appl. Spectrosc. Rev. 1987, 23(3&4), 329. Heinrich, P.; Wyzgol, R.; Schrader, B.; Hatzilazaru, A.; Lubbers, D. W. Appl. Spectrosc. 1990, 44(10), 1641. Meuse, C. W.; Tomellini, S. A. Anal. Lett. 1989, 22(9), 2065. Evanson, K. W.; Urban, M. W.; J. Appl. Polym. Sci. 1991, 42, 2287. Xu, J. R.; Balik, C. M. Appl. Spectrosc. 1988, 42(8), 1543. Fieldson, G. T.; Barbari, T. A. Polymer 1993, 34(6), 1146. Dulebohn, J. 1.; Haefner, S. C.; Berglund, K. A.; Dunbar, K. R. Chem. Mater. 1992, 4, 506. 34 Chapter 2 Structure and Mass Sensing Properties of Zirconium Acetate Thin Films 2.1. Introduction The sol-gel process is an attractive approach for the preparation of sensing materials because porous, optically transparent, thin films of high purity can be synthesized at low temperatures. Polymer oxide glasses have been used to construct chemical sensors based on optical [1-5], mass [6-10], and conductimetric [11,12] transduction methods. In most cases, porous glass-based sensors have utilized recognition centers such as porphyrins [3], cyclodextrins [13], cryptates [4] and organometallic compounds [2] to prepare selective composite devices. In addition, Bein et al. [9,10] have reported that zeolite/glass composite films exhibit molecular exclusion properties. Much of the previous work with porous glass sensors has considered the gel to be simply a immobilization matrix for a more selective interaction site. Little effort has been devoted to using the versatility of sol-gel chemistry to prepare glass films which can enhance the selectivity, sensitivity, or stability of the recognition center. We have recently reported that variation of the carboxylic acid used to prepare transition metal carboxylate thin films can significantly affect the chemical properties of the glasses [14]. For example, film hydrophobicity can be controlled by selection of the hydrocarbon chain length attached to the carboxylate ligand. This suggests that control of 34 35 sol-gel synthesis conditions can lead to glass films with tunable properties that may complement the selectivity of molecular recognition centers. In this work, Fourier transform infrared spectrosc0py (FTIR) and X-ray photoelectron spectroscopy (XPS) have been used to determine the structure of a zirconium acetate thin film prepared using Zirconium(IV) n-propoxide and acetic acid. The information obtained from spectroscopic measurements has been compared to the mass sensor response of the film to simple alcohols and alkanes to evaluate the effect of film structure on sensor properties. 2.2. Experimental Materials. Zirconium(IV) n-propoxide (70% in n-propanol), acetic acid, and n- propanol were obtained from Aldrich Chemical Company and used without further purification. The set of probe analytes used in this work is given in Table 2.1. Analytes were chosen to study analyte/film interactions for n-alkanes and alcohols with difl‘erent steric limitations. The interaction of water with the film was examined to determine film hydrophilicity. Analytes were obtained from either EM Science or Aldrich Chemical companies and used as received. Film Preparation. The method used to synthesize the film was based on techniques developed in these laboratories for the preparation of transition metal carboxylate thin films [14,16]. Each solution and film was prepared in triplicate. Reactions were carried out at room temperature in capped vials. Molar ratios of acetic acidzwaterzalkoxide of 3:13.921 were used. The synthesis involved mixing acetic acid with the zirconium n—propoxide, diluting the acetic acid/alkoxide solution with n-propanol (12 36 Table 2.1. Probe analytes. Boiling Point Molar Volume Analyte Symbol (°C) (cm3/mole)3 methanol 0 65 40.4 hexane El 69 130.5 ethanol 0 78 58.4 iso-propanol $ 82 76.8 tert-butanol 0 82 94 n-propanol O 97 74.7 heptane El 98 146.5 sec-butanol $ 99 92 water A 100 18 iso-butanol O 108 92.8 n—butanol O 1 17 91 .5 octane Cl 126 162.3 a - Reference [15]. 37 mL added to ~ 2 mL of solution), and finally adding water. A vortex mixer was used to vigorously stir solutions following the addition of each reactant. The reaction between acetic acid and zirconium n-propoxide was exothermic and the resulting zirconium solutions were clear and colorless. Films were immediately spray coated onto the desired substrate (quartz, KBr, or AMT IR) using the n-propanol diluted solutions. The films were air-dried overnight and stored in a desiccator. All films prepared in this study were transparent, colorless, and exhibited complete and uniform coverage of all substrates. Fourier Transform Infrared Spectroscopy (FTIR). FTIR spectra of films spray coated onto KBr windows were obtained using a Mattson Instruments Galaxy 3020 Fourier transform infi'ared spectrometer. The mid-IR region (4000-400 cm’l) was examined with a resolution of 2 cm'l. Data acquisition and processing were performed using an Enhanced First software package. FTIR/ATR experiments were performed using a Mattson Instruments Galaxy 5020 FTIR spectrometer equipped with a wideband mercury cadmium telluride (MCT) detector. A commercial attenuated total reflectance (ATR) accessory (Specac, Inc.) designed for HPLC detection has been modified to perform analyte/film studies. The attachment consists of a 500 11L cell in contact with an AMTIR internal reflectance element (IRE), contained in a temperature controlled water bath held at 30°C. The fittings of the cell have been changed to accommodate 1/4" o.d. (1/8” i.d.) stainless steel tubing. The inlet of the ATR cell is connected to a gas chromatograph injector by a 4" long, 1/8" i.d. stainless steel transfer line. The injector and transfer line were heated to 250°C. Analytes were injected as 0.5 11L pulses into a 20 cm3/min flow of dry nitrogen 38 (99.95%, AGA Gas Co.). Analyte/film interaction studies were performed using a polymer film coated onto the AMTIR IRE. The thinnest coating that could be detected by the FTIR was used. FTIR/ATR spectra were collected with 4 cm'1 resolution. A software package supplied by Mattson (FIRST Macros) was modified to collect time dependent FTIR/ATR data. The program uses a spectrum of the polymer coated ATR element as the background for firrther data collection. Eighty spectra (each spectrum is the sum of 10 scans taken at one second intervals) were obtained following the injection of each analyte. Each spectrum is integrated over a wavenumber interval characteristic of the analyte. Plots of integrated peak area versus elapsed time are used to determine the kinetics of analyte/polymer interaction. X-ray Photoelectron Spectroscopy (XPS). Films cast on quartz slides (used as received) were analyzed with a Perkin-Elmer surface science instrument equipped with a model 10-3 60 precision energy analyzer and an omnifocus small spot lens. All spectra were collected using a non-monochromatized Mg anode (1253.6 eV) operated at a power of300 W (15 kV and 20 mA) with an analyzer pass energy of50 eV. The C Is, 0 1s and Zr 3d were scanned for each film. Binding energies were referenced to ligand alkyl and adventitious carbon (C ls = 284.6 eV). Empirically derived sensitivity factors [17] were used for quantitative XPS calculations involving the C15 and Ols. For the Zr 3d, sensitivity factors were derived from the analysis of Zr02 (Aldrich Chemical Company, 99.99+%). XPS peaks were fitted with 20% Lorentzian-Gaussian mix Voigt functions using a non-linear least squares curve fitting program [18]. Each reported result is based 39 on the analysis of at least three films. It must be noted that the use of XPS to evaluate film composition assumes that the films are homogeneous. Sensor Test Apparatus. A schematic diagram of the sensor test apparatus is shown in Figure 2.1. The approximate volume of the cell is 8 mL. Analytes were delivered to the cell as a pulse using a gas chromatograph injector operated at 250°C. Typical operating conditions are 20 cm3/min flow of dry nitrogen (99.95%, AGA Gas Co.) and test cell temperature of 30°C. The temperature of the test cell was controlled by resistance heating the transfer line to 150°C. The transfer line was constructed from a 3.5" long, 1/16" i.d. stainless steel tube. The injection volume used for the set of analytes varied from 0.1 1.1L to 3.0 11L in order to maintain frequency shifts of less than 150 Hz. At least three injections of each analyte were made to ensure reproducibility. Analyte injection order was varied to minimize the effect of analyte order on sensor response. The peak shape for each analyte consists of a sharp leading edge that appears 15 seconds after analyte injection, a narrow peak maximum (approximately 2 sec), and a tailing return to baseline. The sharp rising edge indicates that each analyte reaches the sensor cell as a pulse. The time required for the frequency to return to baseline was less than 2 minutes for the alkanes and approximately 6 minutes for the alcohols. Quartz Crystal Microbalance (QCM). Bulk acoustic wave measurements were performed using 6 MHz quartz crystal microbalances fi'om Maxtek Inc. (Torrance, Calif). A model TM-100 Thickness Monitor (Maxtek Inc.) was used to determine the mass of the polymer coating and the response of the coated device to the probe analytes. The output voltage of the TM-100 thickness monitor (3.06 mV = 1 Hz) was recorded on a 4o fi 1 fi 1 Gas Handling . System Injector 1 1 L o Teflon Aluminum Gasket\/< >/ / Blocks C 0 > 1 Data Acquisition J Figure 2.1. Schematic diagram of sensor test apparatus. 41 Hewlett Packard 3394A integrator. The sensitivity of the QCM is 53 Hz/ug. Quartz oscillators were soaked in iso-propanol, rinsed with methanol, and dried in air prior to use. Cleaned devices were placed in the oscillator circuit and the thickness monitor set to zero. The oscillators were spray coated with the polymer solution and allowed to dry. Films of approximately 0.4 um thickness (~ 61 ug,) were used in this study. Mass sensor responses are reported as the maximum frequency shift in hertz per micromole of analyte injected (Hz/umole) to normalize the response. Responses are plotted against analyte boiling point to facilitate the study of condensation effects on sensor response. A smooth increase in response versus analyte boiling point is expected for simple condensation, while selective analyte/film interactions result in deviations fiom the smooth trend. 2.3. Results FTIR Spectra. The FTIR spectra of acetic acid, zirconium(IV) n-propoxide (in n- propanol), and the zirconium acetate film are shown in Figure 2.2. The assignment of spectral features for acetic acid (Figure 2.2a) is straightforward [19,20]. Weak methyl C- H stretching absorbances are observed near 2900 cm'1 superimposed on an O-H absorbance centered at 3000 cm'l. The strong absorbance observed at 1711 cm'1 is due to a dirneric carboxylic C=O asymmetric stretch. Asymmetric and symmetric CH3 deformation modes are observed at 1414 and 1280 cm'l, respectively. Bending and stretching vibrations of the C-O-H groups are observed between 1440 and 900 cm'l. Skeletal vibrations are evident at 1382 and 625 cm'l. 42 The FTIR spectrum obtained for zirconium(IV) n-propoxide in n-propanol is shown in Figure 2.2b. Spectral features of the alkoxide are similar to those of the parent alcohol with many absorbances shifted to lower fi'equency and additional features present due to (C-O)Zr and (Zr-O)C stretches [21]. The major absorbances are the C-H stretching vibrations at 2960, 2934, and 2874 cm'1 and a feature at 1133 cm'1 which we attribute to a combination of (C-O)Zr and skeletal stretches. The peak observed at 1005 cm'1 is due to (C-O)Zr stretching and features observed at 597, 549, and 505 cm'1 are attributed to alkoxide (Zr-O)C stretches. These features are similar to those reported by Barraclough [21] for zirconium(IV) iso-propoxide: a C-O absorbance at 1011 cm'1 and features between 590 and 520 cm'1 due to alkoxide Zr-O stretches. Interference from n- propanol in the zirconium(IV) n-propoxide spectrum does not allow the assignment of features to bridging OR groups attached to the metal center. The peak observed at 3340 cm'1 is the O-H stretch of the n-propanol solvent. Figure 2.2c shows the FTIR spectrum measured for the zirconium acetate film. Three weak features characteristic of methyl stretches of the acetate ligand are observed at 3017, 2933, and 2980 cm]. Less intense bands observed between 1400 and 700 cm‘1 are due to C-C stretching and C-H bending vibrations of the acetate ligand. A broad band due to hydroxyl groups is observed near 3400 cm]. The low fiequency region of the spectrum shows two major features at 645 and 466 cm'1 which we attribute to Zr-O stretches. Atik and Aegerter [22] have attributed bands at 666 and 363 cm"1 to Zr-O stretches in a zirconium acetate film. Figure 2.3 shows the carboxylate stretch region of the FTIR spectra measured for the zirconium acetate film and sodium acetate. The asymmetric and symmetric COO 43 S ‘33? ”H 8 m V <1- m on ‘9 to m V no (c) a l m m 8 2 3“: (D N cu o c (U .o S. o u) (b) .o < H r-I l\ r-I O m 8 fl 8 2.5 H m ,2, s (a) l l I l l 3600 2500 1600 900 400 Frequency (cm'l) Figure 2.2. FTIR spectra of (a) acetic acid, (b) zirconium (IV) n-propoxide in n- propanol, and (c) zirconium acetate film. 44 stretch fi'equencies for sodium acetate (1580 and 1418 cm'l, Av = 162 cm’l) are similar in relative intensity and position to those reported by Spinner [23]. In the film spectrum, the symmetric COO stretch is as broad as the asymmetric stretch; however, only one well- resolved maximum is observed at 1449 cm’l. The separation between the asymmetric and symmetric COO stretches (Av = 113 cm‘l) is reported with respect to this maximum which is on the high energy side of the peak. XPS Spectra. The XPS Zr 3d,/2 binding energy measured for the zirconium film (182.2 eV) is typical of values measured for tetravalent zirconium [24]. The XPS C Is spectra measured for the film consists of peaks at 284.6 eV (reference peak) and 288.9 eV due to alkyl and carboxylate carbon, respectively. No features due to alkoxide or ether carbon (~ 286 eV) are observed in the spectra. The XPS 0 1s spectra measured for the film consists of peaks at 530.0 eV and 531.4 eV that may be attributed to backbone and ligand (acetate, hydroxyl, or water) oxygen, respectively. The carboxyl carbon/zirconium, backbone oxygen/zirconium, and ligand oxygen/zirconium atomic ratios calculated from C Is, 0 Is, and Zr 3d XPS peak intensities were 1.1, 1.5, and 2.5, respectively. QCM Response. Uncoated Crystal. Figure 2.4 and Table 2.2 show the variation in sensor response (I-Iz/umole) as a function of analyte boiling point for the uncoated quartz oscillator. The response of the uncoated device increases with increasing analyte boiling point. Zirconium Acetate Film. The variation in sensor response as a firnction of analyte boiling point for the zirconium acetate film is shown in Figure 2.5. In contrast to the uncoated crystal results, the response for the zirconium acetate film does not increase simply with analyte boiling point. The fi'equency shifts observed for the unbranched 45 N Os 6 s 3 (b) E .. a 3 oo E (a) l l l l 1764 1600 1444 1296 Frequency (cm") Figure 2.3. FTIR spectrum of carboxylate region obtained for (a) sodium acetate and (b) zirconium acetate film. 46 6 o 5- o 4‘ E" e '3 E 3- ‘6“ o 8 a e m 2_ . 6‘2 0 . :1 A 1‘ o q 0 D e 0 . . . , . , . 60 80 100 120 140 Analyte Boiling Point (°C) Figure 2.4. Variation in QCM response (Hz/umole) versus analyte boiling point for uncoated quartz crystal. 47 Table 2.2 Analyte responses on coated and uncoated QCM. Response Response Ratio (Hz/umole) (coated/uncoated) Analyte Blank Acetate Normalized methanol 0.49 23 46 hexane 0.33 ' 0.17 0.52 ethanol 0.89 25 28 iso—propanol 0.42 12 29 tert-butanol 2. l 3 .8 l .8 n-propanol 1 .8 24 13 heptane 1.6 1.1 0.69 sec-butanol 3.8 18 4.7 water 1 .3 26 20 iso-butanol 2.7 19 7.0 n-butanol 5.5 29 5.3 octane 3.6 2.3 0.64 48 primary alcohols (open circles) increase by less than 30% from methanol (b.p. 65°C) to n- butanol (b.p. 117°C). It should be noted that for the uncoated crystal the response for n- butanol is over a factor of ten greater than the methanol response. The response measured for iso-butanol (open circle, b.p. 108°C), a branched primary alcohol, is lower than that observed for the other primary alcohols. The primary alcohol responses are also significantly higher than the values observed for alkanes (open squares) with comparable boiling points. For the secondary alcohols, iso-propanol and sec-butane] (crossed circles), the measured responses are significantly lower than the values obtained for unbranched primary alcohols of comparable boiling point. The response for tert-butanol (solid circle) is lower than all the other alcohols. Alkane responses increase from 0.2 (hexane) to 2.3 (octane) Hz/umole. In addition, the alkane fiequency shifts observed for the zirconium acetate film are approximately 30% lower than the values recorded for the uncoated crystal. The water response of the acetate film (open triangle) is similar to that observed for the primary alcohols and significantly higher than alkanes of comparable boiling point. This response is considerably enhanced compared to the uncoated crystal. FTIR/ATR Studies. For the uncoated IRE, FTIR/ATR kinetics plots observed for the analytes exhibit a steady increase in purge time and integrated peak area with increasing analyte boiling point. This is the trend expected for condensation of analytes. Figure 2.6 shows the variation in integrated area of the hydroxyl region for methanol (—) and ethanol (--) on a zirconium acetate film versus purge time. The area of the absorbance peak measured for methanol is 5 times the area determined for ethanol. In 49 30 O 25- O A . O O A 20- % 9 0 g ., a 15- 0 a r 9 o o. 8 10- M 51 O ‘ U U o-——EL , . T . , . 60 80 100 120 140 Analyte Boiling Point (°C) Figure 2.5. Variation in QCM response (HzJumole) versus analyte boiling point for zirconium acetate coated quartz crystal. 50 addition, the purge time of methanol is twice that observed for ethanol. For alcohols with boiling points higher than ethanol (> 78°C), the peak areas and purge times increase with increasing alcohol boiling point. This may be attributed to condensation effects noted for the uncoated IRE. Normalized peak areas of the hydroxyl stretch of n-propanol and the CH stretch of heptane observed on the zirconium acetate film are shown as a function of time in Figure 2.7. The purge time observed for n-propanol (b.p. 97°C) is twice that recorded for heptane (b.p. 98°C). The purge time obtained for heptane is representative of the values observed for hexane and octane. 2.4. Discussion Composition of Zirconium Acetate Films. The variety of reactions that occur in the polymer solutions could result in the coordination of several different ligands (e.g.,alkoxides, carboxylates, esters, hydroxides, alcohols and water) to the metal centers in the cast films. The predominance of carboxylate ligand features observed in the FTIR spectra of the film (Figure 2.2c) suggests that most of the ligands are acetate groups. This is consistent with qualitative XPS results that show C Is peaks due only to alkyl and carboxylate carbon. In addition, no features attributable to ester or alkoxide ligands are observed in the FTIR spectrum of the film. It should be noted that previous work on zirconium acetate gels [22] showed that acetate and alkoxide ligands were bound to the polymer backbone. It is possible that the excess acid used in our preparation increases the reactivity of the alkoxide ligands during the hydrolysis and condensation steps of the reaction. The enhanced reactivity would lead to a decrease in the number of alkoxide 51 f 1 ii Integrated Peak Area \ o I I I I k U I I l 0 50 100 150 200 Time (see) Figure 2.6. Variation in hydroxyl stretch intensity for methanol (—) and ethanol (--) as a function of time following analyte injection. 52 I I ' I 1.0 b 3 0.8 r ‘. - ‘ié 0.6 e .' - a) . a. ‘. E 0.4 - .‘ 3 o ; . z . ‘1 0.2 L .' - 0.0 a L 1 a I 0 50 100 Time (sec.) Figure 27. Variation in relative intensity of the hydroxyl stretch for n-propanol (——) and OH stretch for heptane (---) as a function of time following analyte injection. 53 ligands present in the cast films. Doeuff et al. [25] reported that gelatinous precipitates prepared from excess acetic acid and titanium butoxide (Ac/T i = 10) contained no alkoxide ligands. The atomic ratios derived from XPS intensity ratios provide a quantitative assessment of the number and type of ligands present in the film. The extent of carboxylate incorporation observed for zirconium acetate film (1.1 acetates/zirconium) is similar to results obtained fi'om TGA analyses of zirconium-based acetate gels [26]. Since each carboxylate carbon is associated with two ligand oxygens, a film containing only carboxylate ligands is expected to have a ligand oxygen/carboxyl carbon ratio close to two. The ligand oxygen/zirconium atomic ratio calculated for the zirconium acetate film (2.5) suggests that there are approximately 0.3 oxygen ligands per zirconium center that cannot be attributed to acetate. FTIR and XPS analyses indicate that the only other possible oxygen containing ligands present in the film are hydroxyl ligands or molecular water. Water has a characteristic absorbance near 1640 cm'1 [19]. Unfortunately, the strong carboxylate absorbances in this region make it impossible to use this feature to distinguish between hydroxyl ligands and molecular water. Figure 2.8 compares the spectra of the zirconium acetate film before (a) and after (b) heating at low temperature (< 100°C). It is obvious that a significant portion of the hydroxyl features are lost. Since it seems unlikely that hydroxyl ligands would be lost at such low temperatures, we attribute the excess ligand oxygen primarily to water absorbed in the film. This is consistent with previous work which indicated that zirconium acetate films were hydrophilic and water soluble [l6]. 54 Absorbance E (a) l l l l 4000 3000 2000 1000 Frequency (cm'l) Figure 2.8. FTIR spectra of zirconium acetate film (a) before and (b) after heating. 55 Zirconium/Acetate Coordination. Previous studies of metal carboxylate coordination have compared the splitting between asymmetric and symmetric COO stretches with the splitting observed in the ionic form of the ligand to ascertain carboxylate coordination modes [27]. Splitting much larger than that observed for the ionic species tend to indicate monodentate ligands while smaller separations are considered characteristic of bidentate ligands. It has also been suggested that splitting significantly smaller than the ionic value are indicative of bidentate chelating coordination while separations closer to ionic values are typical of bidentate bridging coordination [28]. This reasoning has been used to detemrine acetate ligand coordination in the evolution of titanium xerogels [25] and in titanium, zirconiurrr, and hafirium acetate solutions before and after the addition of water [16]. The asymmetric-symmetric splitting observed for the acetate film (Av = 113 cm'l) is lower than the separation measured for sodium acetate (Av = 162 cm'l). This indicates that a majority of the acetate ligands are bidentate. Since the asymmetric and symmetric peaks are fairly broad, both bidentate bridging and chelating ligands are probably present. Atik and Aegerter [22] report values of 1578 and 1452 cm‘1 (Av = 126 cm'l) for the asymmetric and symmetric COO stretches of a zirconium acetate film, a separation also indicative of bidentate ligands. Structure of Zirconium Acetate Films. Using the XPS estimates of backbone oxygen and carboxylate ligand concentrations in the film, a general picture of the film structure can be proposed. Our results suggest that the zirconium acetate film consists of highly crosslinked networks of zirconium and oxygen, with most zirconium centers 56 covalently bonded to three backbone oxygen atoms (assuming a backbone oxygen coordination number of 2). In addition, if we assign a -2 charge to backbone oxygen and a -1 charge to the carboxylate group, the atomic ratios calculated from XPS lead to a total negative charge of -4.2. With experimental error, this is the value expected for a neutral polymer involving a Zr4+ metal center. This further validates the use of XPS to determine polymer composition. It must be noted that film stoichiometry determined by XPS cannot be used to determine the zirconium coordination number. It is likely that the acetate films have significant interchain interactions that lead to a coordination number greater than 5. Effect of Coating on Sensor Response. Uncoated Crystal. The increase in response observed for the uncoated quartz oscillator with increasing analyte boiling point (Figure 2.4) indicates that the uncoated crystal does not selectively interact with any analyte class. These results also suggest that physisorption (condensation) of the analytes significantly influences the sensor response trend. This is consistent with results reported by King [29] which showed that a larger amount of partitioning would occur for higher boiling analytes exposed to coated quartz crystals. While partitioning cannot occur on the uncoated device, condensation of the higher boiling analytes is expected to yield similar results. Thus in our discussion of zirconium acetate film response, the film will be considered selective for a particular analyte only if the response is significantly difi‘erent from that expected for condensation. Zirconium Acetate Coatings. Comparison of the frequency shifts observed for primary alcohols and normal alkanes (Figure 2.5, Table 2.2) suggests that the acetate film interacts preferentially with alcohols. We attribute this enhanced alcohol response to 57 hydrogen bonding interactions between the alcohol hydroxyl group and the O atoms (Zr-O backbone and acetate ligands) in the film. The relatively small number of hydroxyl groups present in the fihn may also participate in hydrogen bonding interactions. The enhanced interaction between alcohols and the polymer may serve as the driving force for migration of alcohols into the pores of the film. Further discussion of the alcohol responses observed for the zirconium acetate film is facilitated by correcting the responses to compensate for increasing analyte boiling point. We behave that this correction can be accomplished by dividing the response of the acetate film by the values obtained for the uncoated crystal. Figure 2.9 and Table 2.2 show the variation in the acetate coated/uncoated crystal response ratio as a fimction of analyte boiling point. It should be noted that the alkane response ratios are essentially independent of hydrocarbon boiling point. This suggests that the procedure has minimized the effect of condensation on our evaluation of sensor response. With the exception of tert-butanol and sec-butanol, the ratios calculated for the alcohols decrease dramatically from methanol to n-butanol. We attribute this trend to molecular sieving properties of the zirconium acetate film. Larger alcohols cannot diffuse into the pores of the film as readily as smaller alcohols. Further evidence of the molecular sieving effect is obtained by examining the variation in response ratio as a function of molar volume (cm3/mole) shown in Figure 2.10. Indeed, with the exception of iso-propanol, the response ratio is inversely proportional to the molar volumes of the primary alcohols (Table 2.1). Note that the molar volume of tert-butanol is larger than any other alcohol examined in this study. The assertion that pore size effects dominate the alcohol response of the acetate film is 58 consistent with FTIR/ATR results for methanol and ethanol (Figure 2.6) which show that significantly more methanol diffirses into the film compared to ethanol. In addition, the longer residence time measured for methanol may indicate that methanol diffuses further into the film than ethanol. The response ratio calculated for iso-propanol is significantly higher than the value obtained for n—propanol. In part we attribute this deviation to the low response recorded for iso—propanol on the uncoated crystal. This low response would tend to artificially increase the ratio response reported for iso-propanol. The low alkane response observed for the zirconium acetate film may be attributed to the limited interaction possible between the polar acetate film and nonpolar hydrocarbons. The suppressed alkane response, compared to the uncoated crystal, also suggests that the hydrocarbons do not diffuse into the acetate film. This is consistent with the molecular sieving properties proposed to explain the alcohol responses. Note that the molar volumes of the alkanes are significantly larger than the values reported for the alcohols examined in this study (Table 2.2). In addition, FTIR/ATR results obtained for heptane and n-propanol (Figure 2.7) show that the residence time for heptane is significantly shorter than that observed for n-propanol. Thus we believe that the alkanes condense on the surface of the acetate film, but do not migrate into the pores of the glass. 59 50 O , 4o - 30 g 7 O 9 M o d a o g 20- M , O 10 '- O . a O 0 n . 1:1 1:1 I I ' I ' l 60 30 100 120 Analyte Boiling Point (°C) Figure 2.9 Variation in normalized QCM response (zirconium acetate coated crystal/uncoated crystal) versus analyte boiling point. 60 50 404 30- O 9 Response Ratio 10- o - 0 e 0 1'1 El [:1 r ' r ' r f r 1 f 1 ' r 40 60 80 100 120 140 160 Analyte Molar Volume (cm3/mole) Figure 2.10. Variation in normalized QCM response (zirconium acetate coated crystal/uncoated crystal) versus analyte molar volume. 61 2.5. Conclusions Zirconium acetate thin films have been prepared using sol-gel techniques. FTIR and XPS results indicate that the polymer is a highly crosslinked network consisting of a Zr-O backbone modified by approximately one bidentate acetate ligand per Zr center. The zirconium acetate film selectively interacts with alcohols through hydrogen bonding interactions involving the alcohol hydroxyl group and the oxygen atoms in the acetate polymer. The responses for most alcohols can be explained in terms of pore size efi‘ects of the zirconium acetate film. 62 References 1. Avnir, D.; Levy, D.; Reisfeld, R. J. Chem. Phys. 1984, 88, 5956. 2. Dulebohn, J .; Haefirer, S.; Berglund, K.A.; Dunbar, K. Chem. Mater. 1992, 4, 506. 3. Lessard, R.B.; Wallace, M.M.; Oertling, W.A.; Chang, C.K.; Berglund, K.A.; Nocera, D.G. in Processing Science of Advanced Ceramics; Aksay, A.I.; McVay, G.L.; Ulrich, DR, Eds; Mater. Res. Symp. Proc. 155: Pittsburgh, PA, 1989, 109. 4. Dulebohn, J.I.; VanVlierberge, B.; Berglund, K.A.; Lessard, R.B.; Yu, J .; Nocera, D.G. in Better Ceramics Through Chemistry IV; Zelinsky, B.J.J.; Brinker, C.J.; Clark, BE; Ulrich, D.R., Eds; Mater. Res. Soc. Proc. 180: Pittsburgh, PA, 1990; 733. 5. Newsham, M.D.; Cerreta, M.K.; Berglund, K.A.; Nocera, D.G. in Better Ceramics Through Chemistry 111; Brinker, 01.; Clark, BE; Ulrich, D.R., Eds; Mater. Res. Soc. Proc. 121: Pittsburgh, PA, 1988; 627. 6. Yan, Y.; Bein, T. J. Chem. Phys. 1992, 96, 9387. 7. Frye, G.C.; Martin, S.J.; Ricco, A.J.; Brinker, CI. in Chemical Sensors and Microinstrumentation; ACS Symp. Ser. 403, ACS: Washingston, DC, 1989, 208. 8. Frye, G.C.; Martin, S.J.; Ricco, A.J.; Brinker, CI in Better Ceramics Through Chemistry III; Brinker, C.J.; Clark, BE; Ulrich, D.R., Eds; Mater. Res. Soc. Proc. 121: Pittsburgh, PA, 1988;349. 9. Bein, T.; Brown, K.; Frye, G.C.; Brinker, C]. J. Amer. Chem Soc. 1989, 111, 7640. 10. Bein, T.; Brown, K.; Enzel, P.; Brinker, CI. in Better Ceramics Through Chemistry III; Brinker, C.J.; Clark, BE; Ulrich, D.R., Eds.;Mat. Res. Soc. Symp. Proc., 121: 1988, 761. 11. Yoshimura, N.; Sato, S.; Itoi, M.; Taguchi, H. Sozai Busseigahu Zasshi 1990, 3, 47. 12. Inubushi, A.; Masuda, S.; Okubo, M.; Matsumoto, A.; Sadamura, H.; Suzuki, K. in High Tech Ceramics; Vincenzini, P., Eds; Elsevier Science Publishers: Amsterdam, 1987, 2165. 13. Berglund, K.A.; Nocera, D.G. unpublished work. 14. Gagliardi, C.D.; Dunuwila, D.; Berglund, K.A. in Better Ceramics Through Chemistry IV; Zelinsky, B.J.J.; Brinker, C.J.; Clark, BE; Ulrich, D.R., Eds; Mater. Res. Soc. Proc. 180: Pittsburgh, PA, 1990; 801. 15. Handbook of Solubility Parameters and Other Cohesion Parameters;2nd Ed.; Barton, A.F.M., Ed; CRC Press: Boca Raton, Florida, 1991. 16. Gagliardi, C.D.; Berglund, K.A. in Processing Science of Advanced Ceramics; Aksay, A.I.; McVay, G.L.; Ulrich, D.R., Eds, Mater. Res. Symp. Proc. 155: Pittsburgh, PA, 1989 127. 17. Wagner, C.D.; Davis, L.E.; Zeller, M.V.; Taylor, J.A.; Gale, L.H. Surf Interface Anal. 1981, 3, 211. 18. 19. 20. 21. 22. 23. 24. 25. 26. 27. 28. 29. 63 Software provided by Dr. Andrew Proctor, University of Pittsburgh, Pittsburgh, PA. Colthup, N.B.; Daly, L.H.; Wiberley, S.E. Introduction to Infiared and Raman Spectroscopy; 3rd Ed. Academic Press: San Diego, 1990. Silverstein, R.M.; Bassler, C.G.; Morrill, T.C. Spectrometric Identification of Organic Compounds; 3rd Ed. John Wiley & Sons, Inc.: New York, 1974. Barraclough, C.G.; Bradley, D.C.; Lewis, J.; Thomas, I.M. J. Chem. Soc. 1961, 2601. Atik, M.; Aegerter, MA. J. Non-Cryst. Sol. 1992, 147&148, 813. Spinner, E. J. Chem. Soc. 1964, 4217. Handbook of X-ray Photoelectron Spectroscopy; Wagner, C.D.; Riggs, W.M.; Davis, L.E.; Joulder, J.F.; Muilenburg, G.E., Eds; Perkin Elmer, 1979. Doeuff, S.; Henry, M.; Sanchez, C.; Livage, J. J. Non-Cryst. Solids 1987, 89, 206. Leaustic, A.; Riman, RE. J. Non-Cryst. Sol. 1991, 135, 259. Mehrotra, R.C.; Bohra, R. Metal Carboxylates, Academic Press: London, 1983. Deacon, G.B.; Phillips, R.J. Coord Chem. Rev. 1980, 33, 227. King, W.H.,Jr. Anal. Chem. 1964, 36, 1735. 64 Chapter 3 Study of Polysiloxanes as Selective Coatings for SAW-Based Sensors 3.1. Introduction Chromatographic stationary phases are commonly used coatings for mass sensor applications [1]. The phases are readily available with a wide variety of chemical and physical properties. Their synthesis is normally well controlled, leading to reproducible molecular weight and chain length distributions between samples. Perhaps most importantly, there exists a large body of data on the chromatographic selectivity of these phases. Allowing some predictive capability for their application as mass sensor coatings. Polysiloxanes are one class of organic polymers traditionally used as gas chromatographic stationary phases [2] and more recently utilized as sensor coatings [1,3,4]. The general siloxane structure is shown in Figure 1. r ” r B E - J X .. a Y Figure 3.1. General structure of polysiloxane polymer. Where A, B, D, E can represent the same or different fimctional groups. The ability to easily change functional groups makes it possible to tailor the physical and chemical 64 65 properties of the polymer. Additionally, the extensive use of siloxane polymers as stationary phases has prompted research devoted to understanding how their chemical composition influences polymer selectivity [5-8]. This is useful when researching the interactions which contribute to mass sensor response. Mass sensor responses cannot be solely attributed to the addition of mass at the surface of the crystal. Polymer morphology changes caused by changing temperatures or solvent induced swelling can lead to a change in the amplitude or velocity of the propagating wave [9-11]. However, a useful quantitative model for determining viscoelastic effect does not yet exist. Therefore, recent work describing the response of SAW devices has focused on using linear solvation energy relationships (LSERs) to describe and predict analyte responses [3,4]. These studies have ignored the contribution of viscoelastic changes to the response of the sensor. However, morphology changes may interfere with accurate determination of LSER and parameters for more elastic films. In this study responses from surface acoustic wave devices coated with OV-101 (dirnethylmethylsilicone oil), a low molecular weight methylsilicone, and OV-25 (phenylmethyldiphenylsilicone), the most polar of the phenyl silicones, are compared. In addition, a novel method for qualitatively monitoring the presence of analyte induced morphology changes is presented. FTIR/ATR studies illustrate numerous morphology changes for OV-25, while few changes are noted for OV-101. Deviations from the chromatographic [5] and solvatochromic [7,12] predicted trends are explained in light of these findings. 66 3.2. Experimental Materials. OV-101 (M.W. 30,000) and OV-25 (M.W. 10,000) were obtained from Anspec Chemical Company. Table 3.1 shows the boiling points (vide infra) for the analytes examined in this study. Analytes were chosen to study analyte-film interactions for molecules with a wide range of chemical functionalities, reactivities, molecular weights, and boiling points. The analyte/film interactions for a series of homologous hydrocarbons (n-hexane to n-nonane) were examined to establish the effect of analyte boiling point on sensor response. Thiophene, and pyridine, were examined to determine the selectivity for aromatic heterocycles. Benzene, 2-pentanone, pyridine, nitropropane, and 1-butanol were studied because they are used to determine McReynolds constants for chromatographic stationary phases. Analytes were obtained from Aldrich, EM Science or J. T. Baker Chemical Companies and used as received. Sensor Test Apparatus. A schematic diagram of the 13.5 mL cell and experimental apparatus has been shown previously [13]. Analytes were delivered to the cell as a pulse using a Varian autosampler Series 8000 and a Varian gas chromatograph injector operated at 300°C. The temperature of the test cell was controlled by resistance heating of the transfer line. Typical transfer line temperatures are in the range of 3 15-33 5° C. Typical operating conditions are 40 cc/min flow of dry N2 and test cell temperatures of 60°C. Normal autoinjection volume range of analytes are limited to 1.0 11L to 4.0 11L based on the expected sensor response. Normally, six injections of each analyte were made on three coated SAW devices to ensure reproducibility. 67 Table 3.1. List of analyte boiling points, dipole moments and sensor response. Response (I-Iz/umole) Boiling Dipole Point Moment Analyte (°C) (debyes) Blank SAW OV-101 OV-25 pentane 31 0 9 39 41 methanol 65 1.70 40 18 56 hexane 69 0 29 185 297 ethanol 78 1 .69 98 53 167 benzene 80 0 75 345 909 iso-propanol 82 1.66 104 86 221 thiophene 84 0.55 144 373 1212 n-propanol 97 1 .68 289 175 485 heptane 98 0 124 456 744 2-pentanone 100 2.69 333 502 1760 toluene l 1 1 0.36 305 846 2682 pyridine l 16 2.19 1177 1226 3 508 n-butanol 118 1.66 1947 1211 2191 nitropropane 121 3.66 1587 1742 5130 octane 126 0 451 1344 1532 chlorobenzene 131 1.69 2185 2324 5799 nonane 151 0 1827 5004 6024 68 Surface Acoustic Wave (SAW) Device. Data were obtained using a 200 MHz SAW resonator array microbalance (F emtometrics, Costa Mesa, CA). The 200 MHz device has been described previously [14]. The array device consists of 6 individual resonators mounted on a single platform that is directly inserted into the circuit board. The difference between each resonator and a separate reference device is determined and this beat frequency is the measured signal. Power is supplied and frequency counted by an external unit provided by the manufacturer. Injection System. A Varian series 8000 autoinjector system was mounted on a steel plate above a Varian injector that was removed from a gas chromatograph that was subsequently mounted beneath the same steel plate. The pneumatic pressure is provided from a standard laboratory gas cylinder, while purified gas with controlled flow rates are supplied to the injector by Maxtech Mass Flow Controllers. Both the actuators and the cell are using nitrogen gas. Data Acquisition and Manipulation. The fi'equency of the resonators were collected using a digital counting board (MetraByte, Taunton, MA) and stored in a Gateway 2000 486-66 personal computer. The software used to collect the data was written by the manufacturer, and is capable of collecting data for all 6 channels at resolutions of 1-100 Hz (0.2-6 sec). The software also allows for the displaying and printing of data, and for manual peak determinations. Additionally, the standard software has been modified by F emtometrics to allow for timed, multiple runs to allow for the full utilization of the Varian autoinjector capabilities. In this study, data analysis was performed using custom software written using LabView for Windows (v2.52, National 69 Instruments, Austin, TX). This software takes the data output file and separates it into each individual channel, and then simultaneously displays the response of all 6 channels on a single page. Additionally, the software counts peaks, determines peak heights, peak areas, and the relevant analytical information. It also allows for the user to select either Savitsky—Goolay smoothing or median filtering of the data, and some control over the sensitivity of the peak determination. An output file that contain peak height, area, analytical information, and relevant processing information (such as filtering, sensitivity, etc.) can be automatically generated whenever the user is satisfied with the level of data analysis. Data presented in this paper were collected with a resolution of 10 Hz, which provides 5-7 points across the peak maximum Coating Procedure. OV-lOl and OV-25 films are prepared by spray casting from dilute toluene and acetone solutions (~200 ppm), respectively. The resonators were cleaned by soaking in iso-propanol, and then air dried. An airbrush is used to deliver the polymer solution to the entire surface of the crystal. The films were then allowed to air dry and stored in a desiccator. The change in frequency was constantly monitored during the coating procedure. The coating thicknesses used in this study were 150 kHz i 10 kHz. Fourier Transform Infrared Spectroscopy (FTIR). FTIR/ATR experiments were performed using a Mattson Instruments Galaxy 5020 FTIR spectrometer equipped with a wideband mercury cadmium telluride (MCT) detector. A commercial attenuated total reflectance (ATR) accessory (Specac, Inc.) designed for I-IPLC detection has been modified to perform analyte/film studies. The attachment consists of a 500 uL cell in 70 contact with a zinc selenide (ZnSe) internal reflectance element (IRE), contained in a temperature controlled water bath held at 60°C. The fittings of the cell have been changed to accommodate 1/8" stainless steel inlet and outlet fittings that are connected to a gas handling system. The inlet of the ATR cell is connected to a gas chromatograph injector by a 4" long, 1/8" id stainless steel transfer line. The injector and transfer line were heated to 250°C. Analytes were injected as 0.5 11L pulses. The polymers were prepared by drop coating 400 11L of 1000 ppm polymer solution on to the IRE and allowing the solvent to air dry. The resulting polymer films were approximately 0.5 pm thick. FTIR/ATR film spectra were collected with 4 cm'1 resolution. A software package supplied by Mattson (FIRST Macros) was modified to collect time dependent FTIR/ATR data. The program uses a spectrum of the polymer coated IRE as the background for firrther data collection. Eighty spectra (each spectrum is the sum of 10 scans taken at one second intervals) were obtained following the injection of each analyte. Each spectrum can be integrated over a wavenumber interval characteristic of the analyte. Plots of integrated peak area versus elapsed time are used to determine the kinetics of analyte-polymer interaction. Additionally, changes in analyte absorption bands or polymer background may indicate modes of analyte/film interaction or changes in polymer structure or morphology. 3.3. Results SAW Response. Uncoated Device. Figure 3.2 (Table 3.1) shows the variation in sensor response (Hz/umole) as a function of analyte boiling point for the uncoated device. 71 Frequency change per mole was chosen to account for differences in the volume of analyte injected. While analyte boiling point emphasizes the effect of physisorption on sensor response. The uncoated SAW device exhibits an increase in response with increasing analyte boiling point. Alkanes produce slightly lower responses than analytes with comparable boiling points. Analyte response, with the exception of octane, increase dramatically at pyridine (b.p. 116°C) and above. OV-101 Film. Figure 3.3 shows the variation in sensor response as a function of analyte boiling point for the OV-101 coated SAW device. As with the uncoated device, the response of the OV-101 film increases with increasing analyte boiling point. However, it exhibits a smoother trend, and does not show the sudden increase noted for the uncoated device. The alcohols, with the exception of n-butanol, define the lower limit of the smooth increase. All of the analytes, with the exception of the alcohols, produce a greater response on the OV-101 coated device. OV-25 Film. Figure 3.4 shows the response of OV-25 film plotted versus analyte boiling point. As with the uncoated and OV-lOl coated sensors there is a general increase in response with increasing analyte boiling point. However, benzene and thiophene (b.p. 80°C and 84°C, respectively) exhibit significant enhanced responses compared to analytes of similar boiling points. Alcohols and alkanes produce responses slightly lower than the remaining analytes. The response of the OV-25 coated sensor is greater for every analyte than that of the uncoated device. OV-25/OV-101 Ratios. Table 3.2 lists the response ratio for OV-25/OV-101, while Figure 3.5 shows the ratios plotted versus analyte boiling point. The alkanes (squares) 2500 2000 - : fl Response (Hz/umole) 500 - Figure 3.2. 72 ‘s” l s—s O O O 1 El El / O-M ' u r 20 40 60 80 100 U 1 r 120 Analyte Boiling Point (°C) Response curve for uncoated SAW device. ' U 140 160 73 5000 - O 4000 - 1.3 . 0 § 3000- ;“ . 8 0 g- 2000 - d) M O . /\O (D 1000 - o/ / . (I) d) o \o 0 -—r0Fi_'rO-I—cp s ' 1 I 1 l s 20 40 60 80 100 120 140 160 Analyte Boiling Point (°C) Figure 3 .3. Response curve for OV-101 coated SAW device. 74 7000 6000- A/ 5000 - l A § 1 D) O O O 1 Response (Hz/umole) \ 2000 - A l . A A 1000 - \ - A AA 0 - —FA-1—-—_I—-_l— t F I I U I f I I 20 4O 60 80 100 120 140 160 Analyte Boiling Point (°C) Figure 3.4. Response curve for OV-25 coated SAW device. 75 exhibit the lowest increase from OV-101 to OV-25 with a ratio of 1.3 i 0.3. The alcohol (circle) response ratios are similar within experimental error except for n-butanol (b.p. 117°C). n-Butanol exhibits a ratio of 1.81. All of the other analytes (triangles) produce response ratios of 2.5 or greater. FTIR/ATR. The diffirsion time of analytes through the ATR cell was not significantly difl‘erent for the coated or uncoated IRE. Therefore, differences in diffusion rates were not determined fi'om the FTIR/ATR results. The residence time of the analyte seemed to be governed by analyte boiling point and not analyte/film interactions. Although analyte kinetics were not evident from the ATR experiments, changes in polymer film absorption bands were observed for some analytes. These changes were evident after analyte injection, and were attributed to vapor induced polymer morphology changes. OV-101 FTIR/A TR. OV-101 exhibited the smallest amount of polymer changes. Indeed, the majority of analytes did not produce any changes in the polymer spectrum. Figure 3.6 shows a series of spectra collected during the injection of methanol across an OV-101 coated IRE. Figure 3.6a is the initial spectrum taken before methanol has been injected. After injection methanol absorption bands are evident in spectrum 2 which is shown in 3.6b. By spectrum 3, which is shown in 3.6c, methanol has completely purged from the cell. Certain analytes did produce OV-101 polymer changes, these included pyridine, nitropropane, octane, chlorobenzene, and nonane. Figure 3.7 illustrates a series of spectra taken during injection of nonane across a OV-101 coated IRE. Figure 3.7a is simply a reference spectrum of OV-101 taken prior to any analysis. Figure 3.7b is the spectrum Table 3.2. 76 Response ratios for probe analytes. Response Ratio Analyte Uncoated/OV- 1 01 OV—25/OV- 1 01 pentane 4.43 1 .05 methanol 0.45 3.1 1 hexane 6.3 1 1 .61 ethanol 0.54 3.14 benzene 4.61 2.64 iso-propanol 0.82 2.57 thiophene 2.59 3.25 n-propanol 0.60 2.78 heptane 3.68 1 .63 2-pentanone 1.51 3.51 toluene 2.77 3. 17 pyridine 1.04 2.86 n-butanol 0.62 1.81 nitropropane 1 . 10 2.94 octane 2.98 1 . l4 chlorobenzene 1.06 2.50 nonane 2.74 1.20 77 3.5- A Response Ratio 1.5- E El l ' l ' l l 0 - D . q- I I I I I I I I 0 40 60 80 100 120 140 160 2 Analyte Boiling Point (°C) Figure 3.5. OV25/OV-101 response ratio. 78 (C) Absorbance (b) i (a) 1 l 1 4000 3000 2000 Frequency (cm-1) Figure 3.6. FT IR/ATR spectra of methanol on OV-101. 1 1000 79 before the injection of nonane. It is important to remember that the polymer coated IRE is used as the background for the entire run. Therefore, any changes in the polymer structure should produce absorption bands in the following scans. Adsorption bands at 3000 cm“1 in figure 3.7c indicate nonane in present in the ATR cell. Also in this spectrum, changes in polymer morphology are evident from the negative features around 1050 cm". Figure 3.7d shows the polymer film spectrum returned to original features once the nonane is purged from the cell. All of the analytes that produced morphology changes on OV-101 exhibited similar profiles to that of nonane. OV-25 F TTR/A TR. Unlike OV-101, OV-25 exhibited polymer absorption changes for a majority of the compounds injected. Indeed, the only analytes that did not affect the polymer were pentane, methanol, hexane, ethanol, and iso-propanol. Of the analytes which showed viscoelastic effects all, except nonane, produced positive shifts in the polymer absorption bands. Additionally, in contrast to OV-101, the polymer changes persisted after the analytes had purged fi'om the cell. An example of the changes noted for OV-25 is shown for pyridine in figure 3.8a-f. Figure 3.8 (a) and (b) are reference spectra of OV-25 and pyridine, respectively. Although certain bands at 1800 cm'1 and below do overlap, their shape and relative responses do differ between OV-25 and pyridine. Figures 3.8 (c) through (1) are spectrum 1, 2, 4, and 80 of the pyridine kinetic run. In 3.8d the relative peak absorbances match those of pyridine and are therefore attributed to the analyte. However, by the fourth spectrum (figure 3.8e), the band profiles and ratios are identical to those of the OV-25 film. Finally, at spectrum 80, nearly 15 minutes after the injection of pyridine, the polymer bands are still present. 80 (C) (b) Absorbance 1M f m 1 1 1 1 4000 3000 2000 1000 Frequency (cm'l) Figure 3.7. FTIR/ATR spectra of analyte induced morphology changes for nonane on OV-lOl. 81 1(I) l I I 1 L M (e) 1... M (d) 3., M1 (C) Absorbance W Quintin (a) #11».— 1 1 1 4000 3000 2000 1000 Frequency (cm'l) Figure 3 .8. FTIR/ATR spectra of analyte induced morphology changes for pyridine on OV-25. 82 As mentioned above, nonane does not produce positive peaks on OV-25 as noted for the other analytes. In fact, nonane produces negative polymer peaks while present in the cell. However, these peaks reduce in magnitude and eventually reverse as nonane purges fi'om the ATR cell. This change in the polymer background is illustrated in Figure 3.9a-d. Figure 3.9a shows the initial spectrum of the nonane kinetics run. A slight negative peak is visible just above 1000 cm". Upon injection of luL of nonane this negative band greatly increases. The presence of nonane is evident from the features just below 3000 cm'l. By spectrum 4 of the series (figure 3%) the nonane is beginning to purge item the cell and the negative polymer peaks are decreasing in magnitude. Spectrum 20 (figure 3.9d) shows all the nonane purged form the cell, but the polymer feature is now positive. Polymer peaks slowly return toward baseline in the continuing spectra. 3.4. Discussion Effect of Coating on Sensor Response. Uncoated Crystal. The general increase in response observed for the uncoated SAW device with increasing analyte boiling point (Figure 3.2) indicates that physisorption (condensation) of the analytes significantly influences the sensor response trend. This is consistent with results reported by King [15] which showed that a larger amount of partitioning would occur for higher boiling analytes exposed to coated quartz crystals. While partitioning cannot occur on the uncoated device, condensation of the higher boiling analytes is expected to yield similar results. 83 f- I £93-1L, - II_ (C) 5 Absorbance ,0», A) u, £9111_ 1 I I I I 1 J ATV?“ 4000 3000 2000 1000 Frequency (cm'l) Figure 3.9. FTIR/ATR spectra of analyte induced morphology changes for nonane on OV-25. 84 Notwithstanding, physisorption is not the only interaction contributing to the uncoated sensor response. If the response profile were due solely to condensation we would not expect the sharp rise at pyridine and above, or the slight enhancement of other more polar analytes over alkanes. Indeed, we would expect a trend more similar to that observed for the OV-101 coated device. The sharp rise at pyridine is attributed to a combination of increased physisorption and the more polar analytes interacting more strongly with the polar quartz surface. This type of interaction with an uncoated SAW device has been observed previously [16]. OV-101 Coated. As stated above the OV-101 coated SAW device exhibits the smoothest increase in response with increasing analyte boiling point. OV-101 is considered to be a nonpolar stationary phase which separates compounds based solely on boiling point [2]. Solvatochromic parameters calculated from chromatographic data and linear solvation energy relationship calculations firrther establish that PDMS interact solely through dispersion forces [7,12]. This type of interaction should produce a response curve based almost entirely on analyte boiling point due to physisorption. With the exception of small deviations this is exactly what is observed for the OV-101 coated device. The slight deviations from this trend are caused by the alcohols defining the lower limit of response on the OV-101 coated sensor. This is actually in agreement with chromatographic findings which suggest that alcohols do not show eflicient partitioning on OV-101 [6]. In addition, comparison of retention volume of ethanol and a series of alkanes shows extremely poor retention of the ethanol [17]. 85 OV-25 Coated. OV-25 is the most polar of the phenyl siloxanes, however chromatographic data suggests that a phenyl content of no greater than 50 % is sufficient for better separation. Decrease in retention volume has been observed with increasing phenyl content above 50%. However, solvatochromic investigations do show an increase in the 7t* character of the stationary phase with increasing phenyl content. This increased p* should lead to an increase in the response to analytes containing this parameter. The OV-25 coated device only shows enhanced responses for benzene and thiophene and not for the other aromatic. The lack of response maybe due to an artifact of condensation effects. The response due to condensation may be large compared to that of the 11:“ interaction. Therefore, any possible enhancement is masked by the already large response. Condensation is an important factor in the response of the OV-25 coated sensor. However, the response of OV-25 is cannot be completely explained by physisorption. The alkane series defines the response expected for simple condensation, while the remaining non-alcohols exhibit enhanced response. All of the analytes which display enhanced response exhibit considerable polarizability coefficients [18-23]. Solvatochromic parameters determined fiom chromatographic data suggest OV-25 interacts through dispersion forces, polarizability and even hydrogen bond accepting characteristics [7,12]. This suggests that polarizability is a major contributor to the response of our OV-25 coated SAW device. The alcohol response is somewhat puzzling since alcohol have polarizability factors comparable to those exhibited by the higher responding analytes. However, the response is not contrary to data observed during chromatographic separations of alcohols. Indeed retention volume of ethanol falls considerably below that of other comparable analytes [17]. Alcohols also tend to exhibit anomalous results 86 during separation by OV-25 stationary phase. The remaining analytes exhibit considerable polarizability character. OV-101 and OV-25 Comparison. The comparison of OV-25 with OV—101 sensor responses (Figure 3.5) indicates that OV-25 exhibits larger responses for the entire analyte set. This increased response is fully expected for the analytes which interact through induced dipole forces. However, alkanes which interact solely through dispersion forces do not exhibit efficient partitioning in OV-25 during chromatographic separation [24]. The alkanes Indeed, retention times of n-alkanes decrease dramatically from OV-101 to OV-25 [17]. Therefore we expect OV-101 to exhibit larger responses to the alkane series than the OV-25 coated device. Although in need of firrther study, we attribute the increase in response of the OV- 25 not only to increased retention of analyte, but also to larger viscoelastic changes. Our FTIR/ATR results display significant changes in the polymer structure during the presence of the analyte and even after it has purged from the cell. Grate et al. [25] proposed that SAW responses were predominantly due to polymer morphology changes. Grate also showed that an increase in polymer modulus should lead to an increase in the frequency shift of the SAW device. Therefore, a polymer which exhibits a greater viscoelastic effect should produce a greater to nonselectively retained analytes. Ballantine [26], and Bartley and Dominguez [10] observed little or no viscoelastic responses for the systems they examined. These conflicting results suggest that responses due to polymer morphology changes are film and analyte dependent. Therefore, it is not surprising that large changes were observed for OV-25, but OV-101 exhibited swelling 87 only for analytes that showed a large amount of physisorption. Indeed, Patrash and Zellers recently [4] reported a greater relaxation time for the OV-25 polymer. 3.5 Conclusions The selectivity of OV-101 and OV-25 polysiloxane films have been investigated using SAW devices and FTIR/ATR spectroscopy. The results for OV-101 and OV-25 agree fairly well with what is predicted from solvatochromic and chromatographic data. The responses which do not agree are attributed to the analyte induced viscoelastic changes observed in OV-25. FTIR/ATR analysis is able to observe these polymer morphology changes. 88 References 1. Guilbault, G. G.; Jordan, J. M. CRC Crit. Rev. Anal. Chem. 1988 19, 1, 1-28. 2. Haken, J. K.; J. Chromatogr. 1977, 300, 247-288. 3. Grate, J. W.; Abraham, M. H. Sens. &Act. B. 1991, 3, 85-11. 4. Patrash, S. J.; Zellers, E. T. Anal. Chem. 1993, 65(15), 2055-2066. 5. Yancey, J.A. J. Chromatog. Sci. 1985, 23, 161. 6. Yancey, J .A. J. Chromatog. Sci. 1985, 24, 117. 7. Brady, J. A.; Bjorkman, D.; Herter, C. D.; Carr, P. W. Anal. Chem. 1984, 56, 278-283. 8. Ito, M. M.; Kato, J.; Takagi, S.; Nakashiro, E.; Sato, T.; Yamada, Y.; Saito, H.; Narniki, T.; Takamura, 1.; Wakatsuki, K.; Suzuki, T, Endo, T. J. Am. Chem. Soc. 1988, 110, 5147. 9. Ballantine, D. 8., Jr.; Wohltjen, H. IEEE Ultrason. Syrnpos. 1988, 559. 10. Bartley, D. L.; Dominquez, D. D. Anal. Chem. 1990, 62(15), 3069. 11. Frye, G. C.; Martin, S. J.; Ricco, A. J. Sens. Mater. 1989, 1-6, 335. 12. Li, J.; Zhang, Y.; Carr, P. W. Anal. Chem. 1992, 64, 210. 13. Townsend, E. B.; Ledford, J. S. submitted Anal. Chem. 1994. 14. Bowers, W.D; Chuan, R.L.; Duong, T.M. Rev. Sci. Instrum. 1991, 62(6), 1624- 1629. 15. King, W.H.,Jr. Anal. Chem. 1964, 36, 1735. 16. Martin, S. J.; Ricco, A. J.; Ginley, D. S.; Zipperian, T. E. IEEE Trans. Ultra, Ferroe. Freq. Contr. 1987, 34(2), 142. 17. Parcher, J. F.; Hansbrough, J. R.; Koury, A. M. J. Chromatogr. Sci. 1978, 16, 183. 18. Abraham, M. H.; Whiting, G. S.; Doherty, R. M.; Shuely, W. J. J. Chem. Soc. Perkin. Trans. 2 1990, 1451. 19. Abraham, M. H.; Whiting, G. S.; Doherty, R. M.; Shuely, W. J. J. Chromatogr. 1990, 518, 329. 20. Abraham, M. H.; Grellier, P. L.; Prior, D. V.; Morris, J. J.; Taylor, P. J. J. Chem. Soc. Perkin. Trans. 2 1990, 521. 21. Abraham, M. H.; Grellier, P. L.; Prior, D. V.; Duce, P. P.; Morris, J. J.; Taylor, P. J. J. Chem. Soc. Perkin. Trans. 2 1989, 699. 22. Abraham, M. H.; Duce, P. P.; Grellier, P. L.; Prior, D. V.; Morris, J. J.; Taylor, P. J. Tetrahedron Lett. 1989, 29(13), 1587. 23. Li, J.; Dallas, A. J.; Carr, P. W. J. Chromatogr. 1990, 517, 103. 24. Kong, I. M.; Hawkes, S. J. J. Chromatogr. Sci. 1976, 14, 279. 25. Grate, J. W.; Klusty, M.; McGill, R. A.; Abraham, M. H.; Whiting, G.; Andonian- Haftvan, J. Anal. Chem. 1992, 64(6), 610. 26. Ballantine, D. 8., Jr. Anal. Chem. 1992, 64(24), 3069. Chapter 4 Future Work 4.1. Tailoring Metal Carboxylate Polymer Oxide Glasses When choosing polymers for use as chemical mass sensor coatings one needs to consider the properties of the target analytes. Polymers containing functional groups that enhance the chemical interactions with the probe compounds should lead to a more selective response. Additionally,if polymers possess similar physical properties differences in sensor response are more readily attribited chemical effects intead of physical changes in the polymer film. Transition metal carboxylate films are relatively new materials which are easily synthesized. The straightforward synthetic process allows chemical composition and physical properties of the films to be readily altered. It is this ability to tailor the metal carboxylate characteristics that make them usefirl, interesting materials for sensor development. As mentioned previously, production of metal carboxylate films through carboxylic acid modification of the alkoxide results in high quality films. The final polymer consists of a metal-oxygen backbone with bidenate chelating/bridging carboxylate groups attached to the metal center. Carboxylate films have been synthesized using carboxylic acids of increasing chain length up to oxanoic acid [1]. It has been shown that the choice of acid 89 90 can effect properties ( e. g., solubility and hydrophobicity) of the final films [1]. However, the abiliy to tailor the chemical properties of metal carboxylate films using various carboxylate ligands has not been fully explored. We intend to develop transition metal carboxylate chemistry necessary to produce thin films with ligands that emphasize specific intermolecular interactions. The carboxylic acids we intend to incorporate are p-arninobenzoic, p-fluorobenzoic, glutaric, valeric, and 4-oxopentanoic acid. Although each acid exhibits a number of solubility interactions, each was chosen to enhance a specific chemical property. For example p-fluorobenzoic acid was chosen for its large polarizability value although it also exhibits dispersion and dipole interactions. Recently we have attempted synthesizing zirconium benzoate films fi'om zirconium n-propoxide and benzoic and valeric acid in the usual method [Chapter 2]. Benzoic acid was used in a 1:1 ratio with the alkoxide, and the valeric acid was used in excess to produce the conventional 9:1 acid ratio. The method produced a clear colorless solution which was cast to form a crack free colorless film. FTIR spectra of the dried film (Figure 4.1) confirmed the presence of the benzoate ligand, and showed no indication of dimerized carboxylic acid. Unfortunately, preliminary sensor results indicate little difference between films containing valerate and those containing benzoate ligands. This suggests that the valeric acid dominates the chemical properties of the film containing benzoate ligands. Our benzoic acid results suggest that films must be synthesized using only a single carboxylic acid. Initial results using only benzoic acid indicate that the synthetic route used for valeric acid films will not be suitable for other carboxylic acids. Benzoic acid, 91 4000 3000 2000 1 000 Figure 4.1. FTIR spectrum of zirconium valerate/benzoate film. 92 when mixed with zirconium n-propoxide in n-propanol produces a white precipitate on addition of water. Attempts to cast films without the addition of water, with a R3 of 1, produced fractured films and a white precipitate upon evaporation of the n-propanol. FTIR results suggest some substitution by benzoate ligands does occur, however the extent of ligation is still under investigation. These results indicate that synthesis of metal caboxylate thin films will require alternative routes. One alternative procedure which may lead to the production of high quality fihns is the use of additional solvents in the hydrolysis and condensation steps. Hench [1,2] has established that drying can be controlled through the addition of drying control chemical additives (DCCAs). These additives reduce the drying rate allowing the polymer network to form before the solvent evaporates. This procedure has been successfully used for the synthesis of monolithic silicates. Lessard et al. [3] modified the Hench procedure, using different solvent additives, to produce silicate monoliths of greater density and, good optical quality. In addition, different molecular precursors have been obtained depending on the solvent used. Zirconium propoxide dissolve in cyclohexane leads to the formation of monolithic gels [4]. The cyclohexane slows hydrolysis and condensation rates allowing the gels to form. The addition of solvents to metal alkoxides can assist in forming gels by altering the rates of hydrolysis and condensation. However, a major difficulty in the synthesis of metal carboxylate thin films is not synthesizing the polymer, but producing a high quality fiacture free film. Shrinking of the films during drying can often cause them to crack and flake. For the production of monoliths this cracking is can be controlled by controlling the 93 evaporation rate of the solvents. However, for thin films evaporation is not easily controlled. Instead, we propose allowing the gels to dry rapidly and then redissolving them in an appropriate solvent. Once dissolved the polymers could be recast as thin films. This method is commonly used for organic polymers. Although the physical properties of the polymers maybe altered their chemical composition remains unchanged. The overall goal of this portion of research is to synthesize films with difi‘erent solubility parameters. Each acid was chosen to enhance a single property in the solvatochromic parameters used in LSER. This systematic study should allow a better understanding of the importance of these interactions in mass sensor response. In addition to calculating LSER, the production of the films should lead to some important insight on the nature and importance of the carboxylic acid. Presently little is known about the exact nature of this additive. Film physical properties such as pore size, polymer chain length, and average molecular weight are also of interest in such a study. Attempts to understand and control these properties will also be made since each property will have influence the response of mass sensors. 4.2. Molecular Recognition Centers in Polymer Oxide Glasses Tailoring the backbone and organic ligands of sol-gel films will allow some control of the selectivity. However, it is unlikely that analyte specificity is obtainable through this method of sol-gel modification. This is especially true in the area of mass sensors. Since mass sensor response is largely influenced by dispersion forces, diffirsion, and 94 physisorption. Therefore, it is necessary to incorporate a more selective recognition center into the design of a chemical sensor. Molecular recognition has been simply defined as a high level of control over chemical interactions [5]. However, a more accurate definition is a process involving both binding and selection of substrate by a given receptor molecule, and possibly a given function. It implies a structurally well defined pattern of intermolecular interactions [6]. Molecular inclusion compounds, a specific type of molecular recognition center, are able to surround all or part of the guest, establishing numerous non-covalent binding interactions. This arrangement generally achieves higher recognition through the large contact area of the host and guest. A large variety of inclusion and recognition centers have been developed [67-8]. The porosity, optical clarity and non-reactivity of sol-gel glasses makes them ideal matrices for the incorporation of host molecules. Indeed, a variety of optical centers, molecular recognition and molecular inclusion compounds have been entrapped in sol-gel glasses [9,10]. Much of the research has focused on probing the properties of the gels with the entrapped species [1 1-1213]. However, the addition of compounds such as enzymes [14,15], cryptate complexes [16], porphyrins [17] and cyclodextrins [18] concentrated on the development of chemical sensors. The first successful addition of an organic molecule to a sol-gel matrix was performed by Avnir et al. [19]. Researchers entrapped rhodamine 6G (R6G), an organic laser dye, in a silica sol-gel glass to investigate its use as a solid state laser matrix. R6G was simply added to the glass before gelation, and the glass dried as usual. Avnir showed 95 that the incorporated dye maintained its absorption and emission properties within the silica matrix. Attempts to leach the dye by soaking the gel in methanol and water established that the dye was not simply adsorbed at the exposed walls of the pores, but trapped within the rigid glass, most likely within the pore structure. The retention of the optical properties of the dye was encouraging for a wide range of applications. More recent work focused on the development of chemical sensor has established the utility of sol-gel matrices as supports for chemical sensors. Dulebohn et al. [16] entrapped an europium cryptate complex in a titanium film. Using the target analyte as a light harvesting center (LHC), he demonstrated the ability to use absorption energy transfer emission (AETE) to produce a signal in sol-gel films. Titanium films have also been used as the matrix for porphyrins in order to detect aqueous carbon monoxide [17]. Narang et al. [18] incorporated dansyl-glysine-B—cyclodextrin into a silica glass to study the interaction with the target analyte bomeol. Dansyl, a fluorophore, is highly fluorescent in nonpolar enviromnents was tethered to the rim of the CD and resides within the CD cavity. Exposure to bomeol displaces the dansyl from the cavity which led to a dirrrinished fluorescence signal. As mentioned above, Narang et al. [18] established the ability to utilize inclusion phenomena in sol-gel matrices. However, the use of transition metal films in place of silicates offers greater versatility and control of film properties. This work will focus on the incorporation of CD complexes into transition metal carboxylate thin films. Unlike the silicate glasses, the incorporation of CD into these polymers is not straightforward. Past attempts at entrapment have deactivated the inclusion properties of the cyclodextrin by 96 filling the cavity with carboxylic acid during synthesis [20]. Since the CD is simply entrapped and not covalently bound to the metal any attempt at rinsing the excess valeric acid also removes the CD. Attempts at binding the CD to the zirconium center once the film has been cast show that although the CD does adsorb to the film, it is a reversible process. We hope to covalently attach the CD to the metal center during polymerization of the oxide glass. Although the CD may still become filled during hydrolysis and condensation it should be possible to empty the cavity while retaining the CD in the film. The easiest method of attaching the CD to the metal center is through derivatization of the CD's bottom rim. Ideally, the modification of the secondary alcohol to a carboxylic acid group should provide a tethering site to the inorganic backbone. However, if this proves too difficult the use of the tosylate has been proposed for reaction in the sol-gel matrix [21]. Most recognition centers are not perfectly selective, interferences fi'om similarly sized or charged species often produce false signals. The use of sol-gel films as matrices for the inclusion compounds could lead to a reduction in interferences. By controlling the solubility properties of the films it is possible to incorporate a second dimension of selectivity. Another level of confidence is presently being explored in the areas of array devices and pattern recognition algorithms. SAW devices are presently available as six array devices. Coating each array with films of different solubility properties or containing 97 different recognition centers leads to differences in response. It is the development and characterization of film-analyte interactions that will allow the use of sensor array devices. 98 References 1. Hench, L. L. in Science of Ceramic Chemical Processing, Hench, L. L.; Ulrich, D. R., Eds; John Wiley & Sons: New York, NY, 1986, 52. 2. Hench, L. L.; Orcel, G.; Nogues, L. in Better Ceramics Through Chemistry II; Brinker, C. J.; Clark, D. E.; Ulrich, D. R. Eds; Mater. Res. Soc. Symp. Proc. Vol. 73, Pittsburgh, PA, 1986, 35. 3. Lessard, R. B.; Wallace, M. M.; Oertling, W. A.; Chang, C. K.; Berglund, K. A.; Nocera, D. G. in Processing Science of Advanced Ceramics; Aksay, A. I.; McVay, G. L.; Ulrich, D. R., Eds; Mat. Res. Soc. Symp.Proc. 155: Pittsburgh, PA, 1989; 109. 4. Kundu, D.; Ganguli, D. J. Mater. Sci. Lett. 1986, 5, 293. 5. Bein, T. in Supramolecular Architecture, Bein, T. Ed.; ACS Symp. Ser. 499: Washington, DC, 1992, 1. 6. Lehn, J-H. Chemica Scripta 1988, 28, 237. 7. Galan, A.; Mendoza, J.; Toirin, C.; Bruix, M.; Deslongchamps, G.; Rebek, J. J. Am. Chem. Soc. 1991, 113, 9424. 9. Katz, H. E. in Inclusion Compounds; Acedemic Press, Orlando, FL. 1991, 391. 9. Lev, O. Analusis 1992, 20, 543. 10. Avnir, D.; Braun, S.; Ottolenghi, M in Suprarnolecular Architecture, Bein, T. Ed.; ACS Symp. Ser. 499: Washington, DC, 1992, 385. 11. Avnir, D.; Kaufinan, V. R. Langmuir 1986, 2, 717. 12. Kaufman, V. R.; Avnir, D.; Pines-Rojanski, D.; Huppert, D. J. Non-Cryst. Solids 1988, 99, 379. 13. Pouxviel, J. C.; Dunn, B.; Zink, J. I. J. Phys. Chem. 1989, 93, 2134. 14. Braun, S.; Shtelzer, S.; Rappaport, 8; Avnir, D.; Ottolenghi, M. J. Non-Cryst. Solids 1992, 147&148, 739. 15. Ellerby, L. M.; Nishida, C. R.; Nrshida, F.; Yamnaka, S. A.; Dunn, B.; Valentine, J. S.; Zink, J. 1. Science 1992, 255, 1113. 16. Dulebohn, J. 1.; Van Vlierberge, B.; Berglund, K. A.; Lessard, R. B.; Yu, J.; Nocera, D. G. in Better Ceramics Through Chemistry IV; Zelinsky, B. J. J .; Brinker, C. J.; Clark, D. E.; Ulrich, D. R. Eds; Mat. Res. Soc. Symp.Proc. 180: Pittsburgh, PA, 1989; 733. 17. Gagliardi, C. D.; Dunuwila, D.; Chang, C. K.; Berglund, K. A. in Better Ceramics Through Chemistry V; Hampden-Srnith, M. J .; Klemperer, W. G.; Brinker, C. J ., Eds; Mat. Res. Soc. Symp. Proc., 271: Pittsburgh, PA, 1992; 645. 18. Narang, U.; Dunbar, R. A.; Bright, F. V.; Prasad, P. N. Appl. Spectrosc. 1993, 47, 1700. 19. Avnir, D.; Levy, D.; Reisfeld, R J. Phys. Chem. 1984, 88, 5956. 20. Private communication with Dr. K. A. Berglund. 21. Technical Proposal CFMR 1992-1993. onN STATE UNIV. LIBRARIES lllllllHIIHHlIlIHlIlllllllllllllllllll 12 3010336778 | 9 nICHI E3