2 '. .. ‘ A} Lfliklp'} ' 1.3 ”1‘,“th 3w" , ‘ 2 ' .: I- 0‘ HAT-‘3: .3 .M‘l‘n :1"; ‘ i This is to certify that the thesis entitled Thermodynamic and Kinetic Characterization of Porous Graphitic Carbon in Reversed-Phase Liquid Chromatography presented by Yuening Zhang has been accepted towards fulfillment of the requirements for the Master of degree in Chemistry Science “Wax m Btmm Major Professor’s“ SllH’LOO‘i’ Date MSU is an Afiinnative Action/Equal Opportunity Institution LIBRARY Michigan State University PLACE IN RETURN BOX to remove this checkout from your record. To AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE 6/01 c-JClRClDateDucpss-pts ___.__.__.t_._. .. —_ . --_.__ ....--- Thermodynamic and Kinetic Characterization of Porous Graphitic Carbon in Reversed-Phase Liquid Chromatography By Yuening Zhang A THESIS Submitted to Michigan State University In partial fulfillment of the requirements For the degree of MASTER OF SCIENCE Department of Chemistry 2004 Abstract THERMODYNAMIC AND KINETIC CHARACTERIZATION OF POROUS GRAPHITIC CARBON IN REVERSED-PHASE LIQUID CHROMATOGRAPHY By Yuening Zhang In this dissertation, the thermodynamics and kinetics of solute transfer for alkylbenzenes and methylbenzenes on porous graphitic carbon were studied. In the thermodynamic studies, the retention factor increases with an increase in the number of methylene or methyl groups. The data indicate that increases in the number of methylene or methyl groups result in more negative changes in molar enthalpy (AHsm). Hence, the transition from mobile to stationary phase is a more energetically favorable exothermic process with each methylene or methyl group added. The data also show that the change in molar volume (AVsm) is close to zero, which indicates that the solute molecules adsorb on the surface of porous graphitic carbon during retention. In the kinetic studies, the rate of solute transfer increases with increasing number of methylene and methyl groups. The activation enthalpy from the stationary phase to transition state (AHg) was found to increase with increasing number of methylene and methyl groups. But the activation enthalpy from the mobile phase to transition state (AHtm) was much smaller than AH”. The volume barriers AV” and Avtm again are very close to zero. These results provide a thorough and detailed characterization of the retention mechanism for porous graphitic carbon, which is based on adsorption. Acknowledgements At this moment when I finished my dissertation, I am very grateful to my advisor, Dr. Victoria L. McGuffin, for her guidance in my study and research, and her support for me all the time. Many thanks to my committee members, Dr. Bruening and Dr. Jackson in the Chemistry Department of Michigan State University. Thanks to all the group members, Sam, Carl, Xiaoping, Rashad, Melissa, and Amber, who help me to overcome many difficulties. Thanks to my parents, brother, Zheng and our son Dylan, whose love supports me through all the hard times. iii TABLE OF CONTENTS LIST OF TABLES LIST OF FIGURES Chapter 1: Introduction and Background 1.1 Porous Graphitic Carbon (PGC) 1.1.1 Structure and Method of Preparation 1.1.2 Physical and Chemical Properties 1.1.3 Applications of PGC in Separations 1.2 Thermodynamic and Kinetic Theory 1.2.1 Thermodynamics 1.2.2 Kinetics 1.3 Previous Investigations of Thermodynamics and Kinetics in Reversed-Phase Liquid Chromatography 1.3.1 Thermodynamics 1.3.2 Kinetics 1.4 Conclusions 1.5 References Chapter 2: Retention of Aromatic Hydrocarbons on Porous Graphitic Carbon 2.1 Introduction 2.2 Experimental Methods 2.2.1 Solutes 2.2.2 Chromatographic system 2.2.3 Data Treatment and Analysis 2.4 Results and Discussion 2.4.1 Thermodynamic Behavior 2.4.1 .1 Retention Factor 2.4.1.2 Molar Enthalpy 2.4.1.3 Molar Volume 2.4.2 Kinetic Behavior 2.4.2.1 Rate Constant 2.4.2.1 Activation Enthalpy 2.4.2.1 Activation Volume 2.5 Conclusions 2.6 References Chapter 3: Conclusions and Future Directions 3.1 Results 3.2 Future Directions 3.3 References iv 25 25 28 28 29 32 35 35 35 39 46 51 51 55 62 66 67 68 68 70 72 Table 2.1: Table 2.2: Table 2.3: Table 2.4: Table 2.5: Table 2.6: Table 2.7: Table 2.8: Table 2.9: Table 2.10: Activation enthalpy for alkylbenzenes on porous graphitic carbon LIST OF TABLES Corresponding pressure and length of restrictor in the chromatographic system Retention factors (k) for alkylbenzenes on porous graphitic carbon Retention factors (k) for methylbenzenes on porous graphitic carbon Molar enthalpy for alkylbenzenes on porous graphitic carbon Molar enthalpy for methylbenzenes on porous graphitic carbon Molar volume for alkylbenzenes on porous graphitic carbon Molar volume for methylbenzenes on porous graphitic carbon Kinetic rate constants for alkylbenzenes on porous graphitic carbon Kinetic rate constants for methylbenzenes on porous graphitic carbon Table 2.1 1: Activation enthalpy for methylbenzenes on porous graphitic carbon Table 2.12: Activation volume for alkylbenzenes on porous graphitic carbon Table 2.13: Activation volume for methylbenzenes on porous graphitic carbon 31 36 37 42 43 49 50 52 53 57 58 65 LIST OF FIGURES Figure 1 .1: Atomic structure of graphite: Bemal structure of perfect three- dimensional graphite with ABAB... layer registration 3 Figure 1.2: Atomic structure of graphite: Warren structure of two-dimensional turbostratic graphite with no layer registration 3 Figure 1.3: Scanning electron micrograph: spherical particle of PGC (black band represents 2 pm) 5 Figure 1.4: Scanning electron micrograph: High magnification of surface (white horizontal line represents 400 nm) 5 Figure 1.5: Transmission electron micrograph: Spherical particle of PGC 6 Figure 1.6: High-resolution electron micrograph of PGC after heating to 2340 °C 6 Figure 1.7: Energy coordinate diagram 11 Figure 2.1: Structure of alkylbenzenes used to study retention on PGC 26 Figure 2.2: Structure of methylbenzenes used to study retention on PGC 27 Figure 2.3: Schematic diagram for the chromatographic system 30 Figure 2.4: Regression results with the Gaussian equation (A) and the EMG equation (B). Experimental data: (—), regression: ( - - - ). 34 Figure 2.5: Representative graph of the retention factor (k) versus inverse temperature (K'I) used to calculate the change in molar enthalpy according to Equation 1.8. Column: porous graphitic carbon. Mobile phase: methanol, 1.21X107 Pa (1760 psi), 0.955 mL/min. Solutes: toluene (o), m-xylene (I), 1,3,5-trimethylbenzene (A), pentamethylbenzene (A), hexamethylbenzene (a) 40 Figure 2.6: Differential change in molar enthalpy (AAH) (lemol) versus carbon number for alkylbenzenes at P = 5.24%106 Pa (760 psi) and T = 295 to 327 K 44 Figure 2.7: Differential change in molar enthalpy (AAH) (kJ/mol) versus carbon number for methylbenzenes at P = 5.24X10 Pa (760 psi) and T = 295 to 327 K 45 vi Figure 2.8: Representative graph of the retention factor versus pressure used to Figure 2.9: calculate the change in molar volume according to Equation 1.10. Column: porous graphitic carbon. Mobile phase: methanol, 327 K, 1.38 mUmin. Solutes: toluene (o), m-xylene (I), 1,3,5- trimethylbenzene (A), pentamethylbenzene (A), hexamethylbenzene (a) 47 Representative graph of the rate constant (ks...) versus inverse temperature (K‘) used to calculate the change in activation enthalpy according to Equation 1.14. Column: porous graphitic carbon. Mobile phase: methanol, 2760 psi, 0.956 mUmin. Solutes: toluene (o), m-xylene (I), 1,3,5—trimethylbenzene (A), pentamethylbenzene (A), hexamethylbenzene (a) 56 Figure 2.10: Differential change in activation energy from the stationary phase to transition state AAHts (kJ/mol) versus carbon number for alkylbenzene at 5.24x1o‘5 Pa (760 psi) 59 Figure 2.1 1: Differential change in activation energy from the stationary phase to transition state AAHn (kJ/mol) versus carbon number for methylbenzene at 5.24x1o6 Pa (760 psi) 61 Figure 2.12: Representative graph of the rate constant versus pressure used to calculate the change in activation volume according to Equation 1.16. Column: porous graphitic carbon. Mobile phase: methanol, 327 K, 1.38 mUmin. Solutes: toluene (o), m-xylene (I), 1,3,5- trimethylbenzene (A), pentamethylbenzene (A), hexamethylbenzene (c1) 63 vii Chapter 1: Introduction and Background High-perfonnance liquid chromatography (HPLC) is one of the most widely used techniques in the analytical sciences. By relying on the subtle interplay of solute interactions with mobile and stationary phases, HPLC is capable of separating a variety of complex mixtures, including neutrals, charged species, enantiomers, and biological materials. However, to separate a wide range of sample types, different stationary phases are often required. During the past few years, there has been an intense search for new adsorbents because of the rapid proliferation of HPLC and the realization of the disadvantages of most commonly used silica-based materials. Porous graphitic carbon (PGC) is a two-dimensional graphite with a unique sponge-like structure, which gives it good mechanical strength while maintaining an adequate surface area (around 150 mzlg) and porosity (around 80%) [1]. Porous graphitic carbon has unique chromatographic properties, being the only truly hydrophobic adsorbent. It shows strong retention of non-polar, polar, and ionic organic compounds as well as novel stereoselectivity. It gives chromatograms of high efficiency. This chapter presents a description of the PGC stationary phase that was used in this study, as well as the relevant thermodynamic and kinetic theory. 1.1 Porous Graphitic Carbon (PGC) 1.1.1 Structure and Method of Preparation There are, in fact, three distinct forms of carbon to which the term “graphitic" can reasonably be applied and which have well-defined crystal structures. Two of these forms are shown in Figures 1.1 and 1.2. The Bemal structure of perfect three-dimensional graphite consists of layers of carbon atoms, organized in a hexagonal array and ordered ABABAB.... This form of graphite is termed hexagonal graphite. A second, rarer form of three-dimensional graphite, the Lipson and Stokes form, also exists, but here the layers are ordered ABCABC.... This is termed rhombohedral graphite. In both crystalline forms, the layer spacing is 3.354 A and the atomic spacing within the layer is 1.42 A. Most synthetic carbons, when heated to about 3000 °C, assume the second structure shown in Figure 1.2. This is two-dimensional graphite in which graphitic sheets are randomly oriented relative to one another. In two- dimensional graphites, the layer spacing is slightly greater than in three- dimensional graphites at 3.40-3.43 A, and the atomic spacing within the layers is slightly less. The template required for the production of PGC should be a well-bonded silica gel of high porosity. The spherical silica is specially prepared and has a pore volume of 1.4 cm3/g with a specific surface area of 50 mzlg. The mean particle size diameter is 7 pm. This template is impregnated with a melt of phenol and hexylamine in a 6:1 weight ratio. The impregnated material is heated gradually to 150 °C to form phenol-formaldehyde resin within the pores of the silica gel. am 2-13A Figure 1.1: Atomic structure of graphite: Bemal structure of perfect three- dimensional graphite with ABAB. .. layer registration [1] ii I I \ am 2-10A Figure 1.2: Atomic structure of graphite: Warren structure of two-dimensional turbostratic graphite with no layer registration [1] This silica-polymer is then heated slowly to 900 °C in a stream of oxygen- free nitrogen in a specially designed rotary furnace. Approximately 50% of the weight of the polymer is thereby lost, and the density of what remains is increased to about 2 g/cm3. The silica-carbon particles are then treated with hot aqueous potassium hydroxide to dissolve the silica template. The remaining porous glassy carbon has a BET (Brunauer-Emmet-Teller) surface area of 450- 600 mzlg and a pore volume of 2.0 cm3lg, corresponding to a particle porosity of about 80%. This carbon is then heated to 2500 °C in oxygen-free argon. The final carbon has a surface area of about 150 mzlg and a pore volume similar to that of the carbon before high-temperature treatment. The carbon retains the spherical form and the particle size of the original silica template [1]. 1.1.2 Physical and Chemical Properties The structure of the porous carbon has been studied in detail by X-ray diffraction, electron diffraction, scanning electron microscopy, and high-resolution electron microscopy. These techniques provide a self-consistent picture. The scanning electron micrograph in Figure 1.3 clearly shows the spherical shape of the particles and their sponge-like structure [1]. The strong sponge—like structure is capable of withstanding considerable shearing forces, such as those met with in high-performance liquid chromatography. Figure 1.4 is the high magnification of the surface [1]. The porous two-dimensional graphite (PGC) consists of 3-10 pm spherical particles. Figure 1.5 is the transmission electron micrograph of a spherical particle of PGC. Figure 1.6 shows the high-resolution electron micrograph (HREM) of PGC Figure 1.3: Scanning electron micrograph: Spherical particle of PGC (black band represents 2 pm) [1] Figure 1.4: Scanning electron micrograph: High magnification of surface (white horizontal line represents 400 nm) [1] Figure 1.5: Transmission electron micrograph: Spherical particle of PGC Figure1.6: High-resolution electron micrograph of PGC after heating to 2340 °C [1] after heating to 2500 °C [1]. Particularly notable are the straight, parallel lines, indicating ordered graphitic sheets. Also seen are a number of rounded corners and several groups of curved or bent sheets. There is also some evidence of surface defects. The high-resolution electron micrograph (HREM) indicates that PGC consists of intertwined ribbons of carbon, where each ribbon consists of some 30 discrete layers or sheets about 3.4 A apart. It is this ribbon form that accounts for the high mechanical strength of PGC. The crystallites comprise tens of layers of graphitic sheets, thus giving stack heights of 30-60 A, in agreement with the X- ray evidence. Carbon in the form of graphite is one of the most attractive packing materials for HPLC, as it should be free from the disadvantages peculiar to silica gels, such as the limited pH of the mobile phase and the strongly adsorptive silanol groups. The reason is the following. Pure graphite is a crystalline material made up of sheets containing hexagonally arranged carbon atoms linked by the same conjugated 1.5-order bonds that are present in any large polynuclear aromatic hydrocarbon. There are, in principle, no adventitious functional groups on the surface because the aromatic carbon atoms have all valencies satisfied within the graphitic sheets. Perfect graphite is therefore an intrinsically reproducible material with a completely uniform surface, free from any functional groups. Furthermore, graphite is one of the most unreactive substances known and can withstand any mobile phase that will not attack the HPLC equipment itself. Problems, however, may arise when a material of high specific surface area is required. Surface defects must therefore be present at the edges of such crystallites and the ideality of bulk graphite is lost to some extent. However, it is still possible to provide high-surface-area graphites with minimal adventitious surface groups. As an example, a square array of graphitic carbon atoms with a side of 10 nm will contain 40,000 C atoms, of which 160 will be at the edge of the sheet. The edge atoms will thus comprise about 0.4% of the total C atoms. In fact, characterization using X-ray photoelectronic spectroscopy (XPS) of the graphitic carbon spheres detected only the C(1s) band, pointing to a surface devoid of any appreciable oxygen-containing functionalities (estimated detection limit, 0.2 atomic %) [2]. If these oxyen-containing functionalities are present, they will produce much stronger retention for the solutes than graphitic carbon. 1.1.3 Applications of PGC in Separations There are several main factors that influence retention by graphite in HPLC [31- 1. Mobile phase-solute attraction These attractive interactions include dispersion, dipole-dipole, dipole-induced dipole, ion-ion, ion-dipole, ion-induced dipole, and hydrogen bonding forces. These interactions discourage retention by keeping the solute in the mobile phase. 2. Mobile phase-solute repulsion These repulsive interactions arise from disruption of the structure of the hydrogen-bonded mobile phase by non-hydrogen-bonding solutes. These hydrophobic interactions occur between an aqueous or aqueous-organic mobile phase and any nonpolar segments of the solute. These interactions encourage retention by expelling the solute from the mobile phase. 3. Stationary phase-solute interaction These interactions arise from dispersion forces of the London type between the graphite surface and the solute. These interactions are balanced to a greater or lesser extent by similar interactions between the graphite surface and the mobile phase. The net effect may either encourage or discourage retention, but may have an important effect on selectivity. In addition, charge-induced interactions between the stationary phase and solute can promote the retention of polar compounds. These charge-induced interactions are strongest when the polar groups of the solute are forced into direct contact with the graphite surface by the stereochemistry of the solute molecule. In such cases, the additional interactions resulting from a polar group can be so strong that they more than compensate for the increased solute-mobile phase interactions. When stereochemistry does not force direct contact of the polar groups with the surface, the effect is less strong but still significant. The overall effect of these competing interactions is that an increase in the hydrophobicity of the solute, say by adding alkyl groups, always increases retention, as expected in a typical reversed-phase mode. However, an increase in the polarity of the solute by adding groups that can either donate or accept electrons or can polarize the graphite surface may also increase retention, particularly if these groups are constrained to be in close contact with the graphite surface. Porous graphitic carbon shows separations of nonpolar organic compounds [4], highly polar organic compounds [5], geometric isomers [6], and ionized organic compounds [7]. Porous graphitic carbon has proven to have a number of unsuspected properties that have substantially enlarged its area of applications and opened up entirely new possibilities. 1.2 Thermodynamic and Kinetic Theory In order to characterize the solutes undergoing adsorption on the PGC stationary phase, it is necessary to utilize a mathematical system by which they can be compared. The relationships of chemical thermodynamics and kinetics provide such a system. For this purpose, traditional thermodynamic and transition state theories need to be synthesized first. In Figure 1.7, the retention behavior is depicted by using an energy coordinate diagram when the solute moves from mobile (X...) to stationary phase (X3). The thermodynamic changes in molar enthalpy (AH...) and molar volume (AVsm) represent the difference between the final and initial states. If the molar enthalpy (AHsm) is positive, it means the solute gains enthalpy when it transfers from mobile to stationary phase. The chromatographic process is thus unfavorable. Conversely, if the molar enthalpy is negative, the chromatographic process is energetically favorable. If the change in molar volume (AV...) is positive, it implies the solute occupies a larger volume when it is adsorbed on the stationary phase than in the mobile phase. The reverse is true too. 10 Energy Reaction Coordinate Figure 1.7: Energy coordinate diagram depicting the transfer of solute from mobile phase (X...) to stationary phase (X3) via a high-energy transition state (xi). Other thermodynamic and kinetic parameters as described in the text. 11 The kinetic aspects of the retention behavior are characterized by a fast equilibrium between the mobile phase (X...) and transition state (X;) with equilibrium constant Km, followed by a rate-limiting step between the transition state and the stationary phase with rate constant k3,... The corresponding changes in activation enthalpy and activation volume are represented by 13th and Avg", respectively. A similar diagram can be drawn to illustrate the transition from stationary to mobile phase, in which the rate constant km... activation enthalpy AH]... and activation volume Avg are used. 1.2.1 Thermodynamics The equilibrium distribution of the solute (X) between mobile (m) and stationary phases (5) can be characterized by X... <—> Xs This process can be mathematically defined by using the equilibrium constant expression =a_s~_[Xi_ (1.1) a... “' [X]... where a... and a. are the activities, and [X]... and [X]s are the molar concentrations of the solute in the mobile and stationary phases, respectively. The physical characteristics of the system in which molecules enter equilibrium is characterized by the phase ratio ([3), which is the ratio of the surface area of the stationary phase (As) and volume of the mobile phase (Vm): A8 l3=7 (1.2) m 12 The equilibrium constant and phase ratio are important because they are utilized in the calculation of thermodynamic parameters that characterize solutes as they undergo adsorption. The retention factor (k) is a function of the retention times of a solute (t.) as well as an unretained compound (to): t —to to k=5 I3 r k: (1.3) (1.4) The change in Gibbs free energy (AG) can describe whether or not an equilibrium transfer is spontaneous. When applied to the chromatographic process, the Gibbs free energy is an indicator of how favorable is a transition from the mobile to the stationary phase. Traditionally defined as G = H - TS (1.5) this equation describes the inherent energy of a molecule in some defined state where H is the enthalpy, T the absolute temperature, and S the entropy. Given that a single molecule in a chromatographic system can exist in one of two mutually exclusive phases, it is necessary to define each phase with its own energy (G... or 6.). As the molecule equilibrates between these two states, the change in energy (AGsm) can be measured. This change is defined by AGsm = AH,,,, -A(TS)sm (1.6) This energy can be directly related to the equilibrium constant described in Equation 1.1 through the van’t Hoff equation: 13 mK=1§§1-UJ) RT The importance of this equation is that it directly links the equilibrium constant to the Gibbs free energy, thus correlating a molecule’s affinity for a given phase to a thermodynamic value that describes how energetically favorable is a transition into that phase. Instead of K, AHS... is used as a quantitative measure of the equilibrium event. A relationship between the retention factor (k) and phase ratio ([3) allows for the calculation of AH... By substituting Equation 1.4 into Equations 1.6 and 1.7, the resulting equation is AH AS 5m 3m -| 1.8 RT + R "B( ) mk=‘ Equation 1.8 allows for the calculation of AH... from the slope of a graph of In k vs. 1/T at constant pressure. The assumption within this calculation is that neither AH... nor ASS... are temperature dependent. The intercept contains information about the change in molar entropy as well as the phase ratio. Because it is not known how the phase ratio [3 changes with temperature and pressure, the change in molar entropy ASS... cannot be reliably calculated. In addition to the energetic parameters that describe a system, the change in molar volume (Avg...) can also be calculated. When applied to the chromatographic system, AV... is the physical change a molecule undergoes as it tranfers from mobile to stationary phase. This transition can be described through a model of solvation. In the mobile phase, a solute molecule is surrounded by a given number of solvent molecules. As the solute transfers into 14 the stationary phase, these solvent molecules may be stripped away. When the transition is complete, the molecule is solvated by the stationary phase molecules if it has partitioned, or has lost some of its hydration shell if adsorbed. As a result, AV... is an indicator of physical change as the molecule moves between the mobile and stationary phases. In order for Avg... to be calculated, enthalpy must be further defined as a function of the internal energy (AE), pressure (P), and volume (V). When using these three fundamental variables of a system, the change in enthalpy (AH...) is defined as AHsm = AEsm +A(PV)3m (1.9) By substituting Equation (1.9) into Equations (1.6) and (1.7), under constant temperature, the retention factor k can be related to the pressure: A5,,“ +TASsm - PAVsm) RT mk=( -In[3 (1.10) Equation (1.10) allows for the change in molar volume to be extracted from the slope of a graph of In k vs. P at constant temperature by assuming that neither AEs... nor ASS... are pressure dependent. Using these graphical methods of analysis, it is possible to quantify the physical change of a solute as it transfers between phases. Thermodynamics provide a simple method by which physical and energetic properties can be calculated. While the parameters defined above allow for the comparison of solutes, they do not provide insight into the rate at which a given transition may occur. As a result, transitions that may be 15 energetically favorable (i.e. large negative AG...) may be either slow or fast. Since thermodynamics is incapable of describing the rate at which a transfer between mobile and stationary phases will occur, kinetic studies need to be conducted. 1.2.2 Kinetics Kinetic theory requires an initial definition of the transitions a molecule undergoes, where kms is the rate constant for transfer from stationary to mobile phase and K...“ is the rate constant from mobile to stationary phase. The first step in elucidating kinetic data is the determination of the rate constants. By using the retention factor (Equation 1.3), the rate constants are defined by _ 2kto ms 2 T _ 2k2to k =kkm, 12 (1.11) . k (1.12) am where ‘t is the exponential component. These rate constants quantitatively define the transfer, thereby allowing for comparison between solutes, mobile phases, and stationary phases. However, a more in depth treatment of kinetics using transition state theory allows for other rate parameters to be calculated. Transition state theory is an expansion on the kinetic process shown above. The crux of the theory is based upon the existence of a transition state: a short-lived, high-energy state on the reaction coordinate diagram (Figure 1.7), where K]... is an equilibrium constant for the solute from mobile phase to transition state reaction, k3... is a rate constant, and X; is the transition state. This conformation can be defined more clearly by relating the rate constants km and k... to transition state energy (AEp... AEp), enthalpy (AHm, AH”), and volume 16 (AV... AV”). To demonstrate the mathematical relationships, k... will be used as an example. In order to calculate the three kinetic parameters, the Arrhenius equation is first used. Defined by the form ksm = Atmexp(—AEtm IRT) (1.13) this equation relates the rate constant to the internal energy necessary to cross the barrier on the reaction coordinate, where A... is the pre-exponential factor and AE... is the activation energy. By algebraic rearrangement, the Arrhenius equation becomes AE lnksm=lnAtm- R?“ (1.14) Thus, AEt... can be calculated from the slope of a plot of In k... vs. 1/T at constant pressure under the assumption that A... and AE;.. are temperature independent. In order for activation enthalpy to be calculated, a relationship between the energy and the enthalpy must be defined. Given the form A5,", = AH.“ + RT- A(Pv,m) (1.15) Equation (1.15) directly relates the activation energy to the activation enthalpy at a given temperature and pressure. As illustrated, though, the change in activation volume (AV...) as a molecule moves between the mobile phase and the transition state must first be calculated. Using a similar method to the thermodynamic equations, A5,", + TAStm -PAvm) Ink5m =( RT (1.16) 17 AV.. is calculated from the slope of a plot of In k... vs. P at constant temperature. Once again an implicit assumption is that AE... and AS... are pressure independent. As a result, it is possible to calculate the activation enthalpy (AH...) by using the activation volume and activation energy by means of Equation 1.15. By using the above method, the thermodynamic and kinetic parameters that characterize the adsorption events in reversed-phase liquid chromatography can be determined in the following chapter. 1.3 Previous Investigations of Thermodynamics and Kinetics in Reversed-Phase High-Performance Liquid Chromatography A number of very unusual separations can be successfully performed with PGC as stationary phase. These separations are difficult or impossible with more conventional reversed-phase materials, such as octadecylsilica (ODS) or polymeric materials. Porous graphitic carbon is thus a totally unique chromatographic material which will, no doubt, be the subject of much further research. At this stage, the retention mechanism on PGC is not well understood. A detailed thermodynamic and kinetic study of the solutes on PGC will help to understand the retention better. Before this, a brief summary of thermodynamic and kinetic studies on conventional reversed-phase materials will be summarized for comparison with PGC. 1.3.1 Thermodynamics A number of homologous series have been used to study the effect of repeating structural units on chromatographic retention behavior. Such series 18 have included fatty acids, alkylbenzenes, and polycyclic aromatic hydrocarbons (PAHs). These series can offer a simple relationship between the retention factor (k) and the number of units, such as the methylene (CH2) groups. Early studies showed that an increase in methylene unit number results in a logarithmic increase in the retention factor [8]. The effect of stationary phase chain length was studied by Berendsen and de Galan for a number of solutes with varying methylene unit numbers [9]. The data from their study indicated that there is a critical chain length, beyond which retention does not change. According to the authors, the limiting chain length lies between 6 and 14 carbons for the alkylsilica materials. The critical chain length is independent of mobile phase composition, but increases with increasing length of the solute alkyl chain. Similarly, Tchapla et al. investigated seven series of alkyl-substituted solutes including alkylbenzenes, alkylchlorides, and carboxylic acids on different stationary phases. They concluded that the solute alkyl chains intercalate into the alkylsilica stationary phase [10]. In addition to studies of retention, the molar enthalpy and molar entropy have also been studied for homologous series. lssaq and Jaroniec calculated the molar enthalpy using Equation 1.8 for alkylbenzenes using binary mobile phases and alkylsilica stationary phases (C1, C4, C3, C18) [11]. They concluded that the change in molar enthalpy and molar entropy depend linearly on the number of methylene units in the alkyl chains of the homologous series, not the bonded stationary phases. McGuffin and Chen studied the retention of a series of even- 19 numbered fatty acids (C10 through 022) not only as a function of temperature, but also pressure [12]. From their experiments, they quantitated the changes in molar enthalpy (AAH) and molar volume (AAV) of —3.5 kcal/mol and —14.1 mL/mol, respectively, for methylene units. In addition to the studies of retention as a function of methylene unit number, series containing phenyl units have also been investigated. Most notably, PAHs have been used since they have repeating phenyl units in different configurations. Whereas alkylated solutes can be characterized as a function of the number of methylene units, PAHs have been studied as a function of ring number, annelation structure (i.e. degree of ring fusion), and planarity [13-16]. Studies have shown that an increase in the ring number leads to a logarithmic increase in the retention factor. An increase in temperature results in a decrease in the retention of PAHs. By contrast, the effect of pressure is much smaller. McGuffin and Howerton studied the thermodynamics of solute transfer for a series of PAHs as a function of ring number, annelation structure, planarity, temperature, pressure, and bonding density using octadecylsilica (ODS) [17,18]. The data from their study indicated that an increase in ring number results in more negative changes in molar enthalpy (AH...) and molar volume (AV...) For a series of isomers, solutes with condensed annelation structure exhibited less negative changes in molar enthalpy and molar volume than the more linear solutes. Nonplanar solutes demonstrated changes in molar enthalpy and molar volume that are less negative than would be expected based upon ring number alone. These data indicated that the condensed PAHs, as well as the nonplanar 20 PAHs, interacted with the first few carbons near the distal terminus of the ODS phase. But the more linear solutes as well as those with more rings interacted with the more ordered regions of the ODS closer to the proximal terminus. 1.3.2 Kinetics Although the thermodynamics of retention have been extensively probed, less work has been conducted into the kinetics of retention. Work on kinetics has generally fallen into two classes based on the techniques employed: non- chromatographic and chromatographic. Nonchromatographic experiments have utilized pressure, temperature, and dipole jump methods [19-24], as well as fluorescence correlation spectroscopy [25,26] to investigate chromatographic materials. However, the inherent difference between these systems and a packed chromatography bed provide limited information about the kinetics of an actual separation. A greater problem is that the temperature or pressure jump actually changes the kinetics of the system, according to Equations 1.14 and 1.16. To circumvent these difficulties, several chromatographic techniques have been used, including frontal analysis and pulse methods [27-29]. Frontal analysis involves injecting a large concentration of the solute onto the chromatography column, and then analyzing the resultant curve. Also know as a breakthrough curve, this curve provides information about the kinetics of sorption. However, it does not represent the most common type of chromatography, which involves injecting smaller concentrations of material onto the column. 21 Pulse methods better approximate a typical chromatographic separation since they involve the injection of smaller volumes/concentrations. However, the current method to elucidate the kinetics is also questionable. Thus the current chromatographic methods are not ideal for characterizing the kinetics of retention since these methods rely upon a number of assumptions and approximations. Using the exponentially modified Gaussian model (EMG), McGuffin and Howerton have studied the kinetics of PAHs in ODS [17,18]. The rate constants decreased with increasing ring number and less condensed annelation structure. The enthalpic and volume barriers were found to be very large. These barriers increased as a function of ring number, but decreased with more condensed annelation structure. 1.4 Conclusions Although significant efforts have been invested in the applications of PGC in both chromatography and extraction, the retention mechanism is far from being understood. Limitations in both theory and experimental design have hindered definitive quantitative analysis. In this dissertation, the thermodynamic and kinetic contributions to retention of two series of aromatic compounds on PGC will be explored using the EMG model. These studies will help to understand the retention mechanism on PGC. 22 1.5 References [1] J. H. Knox, B. Kaur, G. R. Millward, J. Chromatogr., 352 (1986) 3-25. [2] R. S. Deinhammer, E. Ting, M. D. Porter, Anal. Chem., 67 (1995) 237-246. [3] J. H. Knox, P. Ross, in Advances in Chromatography, P. R. Brown, E. Grushka, Marcel Dekker, New York, NY, vol. 37, pp 73-119 (1997). [4] M. C. Hennion, V. Coquart, S. Guenu, C. Sella, J. Chromatogr. A, 712 (1995) 287-301 . [5] G. Machtalere, V. Pichon, M. C. Hennion, J. High Resol. Chromatogr., 23 (2000) 437-444. [6] M. C. Pietrogrande, A. Benvenuti, S. Previato, F. Dondi, Chromatographia, 52 (2000) 425-432. [7] M. C. Hennion, ESPCI (Ecole Supérieure de Physique et de Chimie Industrielles), private communication, 2003. [8] E. Grushka, H. Colin, G. Guiochon, J. Chromatogr., 248 (1982) 325-339. [9] G. E. Berendsen, L. de Galan, J. Chromatogr., 196 (1980) 21-37. [10] A. Tchapla, S. Heron, H. Colin, G. Guiochon, Anal. Chem., 60 (1988) 1443- 1448. [11] H. J. Issaq, M. Jaroniec, J. Liq. Chromatogr., 12 (1989) 2067-2082. [12] V. L. McGuffin, S. H. Chen, J. Chromatogr. A, 762 (1997) 35-46. [13] L. A. Cole, J. G. Dorsey, Anal. Chem., 64 (1992) 1317-1323. [14] K. B. Sentell, J. G. Dorsey, J. Chromatogr., 461 (1989) 193-207. [15] M. Olsson, L. C. Sander, S. A. Wise, J. Chromatogr., 477 (1989) 277-290. [16] L. C. Sander, M. Pursch, S. A. Wise, Anal. Chem., 71 (1999) 4821-4830. [17] S. B. Howerton, V. L. McGuffin, J. Chromatogr. A, 1030 (2004) 3-12. [18] S. B. Howerton, V. L. McGuffin, Anal. Chem., 75 (2003) 3539-3548. [19] D. B. Marshall, J. W. Burns, D. E. Connolly, J. Chromatogr., 360 (1986) 13- 24. 23 [20] S. W. Waite, J. M. Harris, E. H. Ellison, D. B. Marshall, Anal. Chem, 63 (1991) 2365-2370. [21] S. W. Waite, D. B. Marshall, J. M. Harris, Anal. Chem, 66 (1994) 2052-2061. [22] F. Y. Ren, J. M. Harris, Anal. Chem, 68 (1996) 1651-1657. [23] J. M. Harris, D. B. Marshall, J. Microcolumn Sep., 9 (1997) 185-191. [24] S. R. Shield, J. M. Harris, Anal. Chem, 74 (2002) 2248-2256. [25] R. L. Hansen, J. M. Harris, Anal. Chem, 70 (1998) 2565-2575. [26] R. L. Hansen, J. M. Harris, Anal. Chem, 70 (1998) 4247-4256. [27] K. Miyabe, G. Guiochon, Anal. Chem, 72 (2000) 5162-5171. [28] K. Miyabe, S. Sotoura, G. Guiochon, J. Chromatogr. A., 919 (2001) 231-244. [29] K. Miyabe, G. Guiochon, Anal. Chem, 74 (2002) 5754-5765. 24 Chapter 2: Retention of Aromatic Hydrocarbons on Porous Graphitic Carbon 2.1 Introduction As depicted in Figures 2.1 and 2.2, eighteen aromatic hydrocarbons have been chosen to investigate the effect of length of alkyl chain and number of methyl substituents on the thermodynamics and kinetics of retention on porous graphitic carbon (PGC). Alkyl-substituted benzenes were chosen as the solutes for two reasons. First, they make a good homologous series in establishing retention trends. Second, for environmental and toxicity concerns, BTEX levels in oil are a very important parameter. BTEX is the collective name of benzene, toluene, ethylbenzene, and the xylene isomers (o-, m-, and p-xylenes). BTEX compounds are the most common aromatic compounds in petroleum, comprising up to a few percent of the total weight for some crude oils. BTEX constitutes the most soluble, most mobile fraction of crude oil and many petroleum products. BTEX compounds have many chemical applications and are widely used in the manufacture of paints, synthetic rubber, agricultural chemicals, and chemical intermediates. The concentration of BTEX in petroleum can directly affect the physical and chemical properties of petroleum such as density, viscosity, flash points, dispersibility, emulsion stability, solubility, and weathering processes. On the other hand, BTEX compounds frequently enter soil, sediments, and groundwater because of accidental oil spill, leakage of gasoline and other petroleum fuels from underground storage tanks and pipelines, and improper waste disposal practices. 25 I F‘fi Figure 2.1: Structure of alkylbenzenes used to study retention on porous graphitic carbon. 26 Figure 2.2: Structure of methylbenzenes used to study retention on porous graphitic carbon. 27 In these cases, BTEX compounds can also be used as markers in identifying the presence and determining the composition of these weathered fuels originating from leaking underground fuel tanks or pipelines. BTEX are hazardous carcinogenic and neurotoxic compounds and are classified as priority pollutants regulated by the US. Environment Protection Agency (EPA) [1], especially benzene, which has a very low tolerance standard in drinking water of 5 pg/L. 2.2 Experimental Methods 2.2.1 Salutes The alkylbenzenes (Figure 2.1) were chosen as a homologous series to study the effects of methylene addition on retention. Benzene, ethylbenzene, propylbenzene, butylbenzene, pentylbenzene and hexylbenzene were obtained from Aldrich. The methylbenzenes (Figure 2.2) were chosen as a homologous series to study the effects of methyl substituent addition on retention. Toluene, o-xylene, m-xylene, p-xylene, 1,2,3-trimethylbenzene, 1,2,4-trimethylbenzene 1,3,5- trimethylbenzene, 1,2,3,5-tetramethylbenzene, 1,2,4,5-tetramethylbenzene, pentamethylbenzene, and hexamethylbenzene were obtained from Aldrich. Mixed standard solutions were prepared around 10‘3 M in high-purity methanol (Burdick and Jackson, Baxter Healthcare) such that there was no overlap of solute zones at any temperature and pressure. A nonretained marker, acetonitrile, was added to each solution at a concentration of 1 M. 28 2.2.2 Chromatographic System The chromatographic system is shown schematically in Figure 2.3. It includes a single-piston reciprocating pump (Model 114 M, Beckman Instruments) operating in constant pressure mode (:t 15 psi) over the range of 6.90 x 106 Pa (1000 psi) to 3.45 x 107 Pa (5000 psi). The pump allows for the mobile phase to be supplied continuously without the need for depressurization during the refill cycle. At all pressures, the flowrate was maintained at approximately 1 mUmin at room temperature. Samples are introduced via a manual injection valve (20 pL, Model 7125, Rheodyne Instruments). Once the sample is introduced, it is carried by the methanol mobile phase to the column. The Hypercarb column is supplied by Shandon HPLC (100 x 4.6 mm id, mean particle diameter 7 pm). At the end of the column, a 100 pm i.d. fused-silica capillary (Polymicro Technologies) is attached to serve as a restrictor. The length of the restrictor is reduced as the inlet pressure is decreased in order to maintain a constant pressure drop along the column (3.31x 105 Pa (48 psi)/cm). The length of the restrictor and corresponding pressures and flow rate are listed in Table 2.1. The injector, column, and restrictor are all housed within an oven (Model SM900, Anspec Instruments) that maintains a constant temperature (10.1 K) through the use of resistive heating coils. The oven was operated through the temperature range of 296 to 326 K. 29 Pump ® UV-Visible @— Column Absorbance Detector Oven Computer Figure 2.3: Schematic diagram for the chromatographic system. I: injection valve, R: restrictor. 30 Table 2.1: Corresponding pressure, flow rate, and length of restrictor in the chromatographic system. Pressure (psi) Length of restrictor (m) Flow rate (mL/min)" 1000 1.450 0.965 2000 3.777 0.955 3000 5.982 0.956 4000 8.194 0.949 5000 10.345 0.993 A at room temperature 31 The spectroscopic technique used for detection is ultraviolet-visible absorbance. As shown in Figure 2.3, a commercially available unit (Model UVlDEC-100—V, JASCO) is directly coupled to the column such that the effluent is directed to a capillary flow cell. The flow cell is a transparent fused-silica capillary with a 180-pm i.d., able to hold high pressure in the chromatographic system. The wavelength for detection (210 nm) is chosen based upon the absorption maximum of the solutes. The output from the detector is directed to a data acquisition board (National Instruments) and computer for further data analysis. The data collecting program was written in Labview (National Instruments, v3.1). 2.2.3 Data Treatment and Analysis Since the desired data are zone profiles contained within a multicomponent chromatogram, the first challenge for the data analysis is the reproducible extraction of the individual profiles. The integration limits will affect the accuracy of the analysis results [2]. Once removed, the zone profiles are iteratively fit using a commercially available program (Peakfit v3.18, SYSTAT Software). Since there are many functions available for the analysis of zone profiles, the second obstacle to the analysis is choice of an appropriate function. The accurate description of asymmetrical peaks usually requires a large number of statistical moments and a large number of data points [3]. In addition, noise can affect the values for the moments [4]. As a result of these limitations, alternative equations such as the exponentially modified Gaussian (EMG) [5], the 32 nonlinear chromatography (NLC) [6], and the Haarhoff-Van der Linde function [7] have been developed for evaluating non-Gaussian zone profiles. The EMG and NLC models provide variables that can be directly related to experimental parameters. As a result, these two functions were initially chosen to study chromatographic zones. The results of nonlinear regression of a typical experimental zone profile with the Gaussian function and the EMG equation are shown in Figure 2.4. The Gaussian equation (Figure 2.4A) produces nonrandom residuals with a relatively small value for the square of the correlation coefficient (R2 = 0.987). The EMG equation (Figure 248) provides significantly better fit with only small residuals around the maximum of the zone profile (R2 = 0.998). Moreover, the EMG equation yields better statistical results than the NLC and other models mentioned above. On the basis on these results, the EMG model was used throughout the studies in this dissertation. The EMG is a convolution of a Gaussian and an exponential function. The form of the equation is _A i tG—t t—tG_ o C(t)—21exp|:212+-——t :l[erf[ J‘z‘. JEJHI (2.1) where A is the area, ts is the mean of the Gaussian component, a is the standard deviation of the Gaussian component, and 17 is the exponential component. Over the past thirty years, several authors have sought to characterize the use of the EMG function for chromatographic analysis. While Grushka [8,9] and Yau [10] presented some early work, Foley and Dorsey have published two seminal papers reviewing the EMG function in the intervening years [11,12]. 33 Figure 2.4 140 145 150 155 160 165 170 175 180 Time (s) Figure 2.4: Regression results with the Gaussian equation (A) and the EMG equation (8). Experimental data: (—), regression: ( - - - ). These reviews have demonstrated the use of the EMG function for a variety of chromatographic applications. More recent studies have compared the EMG to other equations for the analysis of zone profiles [13,14]. The studies have concluded that high-order polynomials that exceed the four parameters used in the EMG function can describe asymmetrical peaks very well. However, the complex nature of such functions makes them impractical for common use. In addition, the parameters for these high-order polynomials may not have any physical significance as do those calculated using the EMG function. Within this thesis, the EMG is employed predominantly to calculate the thermodynamics and kinetics of retention. The thermodynamic and kinetic parameters can be calculated by substituting the measured values of t6 and 1 into Equations 1.3, 1.11, and 1.12 of Chapter 1, where t. = ta + T (2.2) 2.4 Results and Discussion 2.4.1 Thermodynamic Behavior 2.4.1.1 Retention Factor The retention factor (k) is the simplest of the measurable parameters that allow for the comparison of the relative strength of interaction. A larger retention factor indicates that the solute has greater affinity for the stationary phase. This comparison is useful to evaluate the effect of different solute structural features. Using Equation 1.3, the retention factor for each solute on PGC was calculated. The results of these calculations are summarized in Tables 2.2 and 2.3. For the homologous series of alkylbenzenes, the retention factor increases 35 Table 2.2: Retention factors (k) for alkylbenzenes on porous graphitic carbon. Solute k“ k3 Aklk k° k° Aklk (%) (%) Benzene 0.16 0.16 49 0.16 0.16 1.90 Toluene 0.36 0.31 -16.2 0.34 0.33 -119 Ethylbenzene 0.37 0.30 -17.2 0.34 0.33 -0.89 Propylbenzene 0.53 0.43 -18.4 0.49 0.49 -0.41 Butylbenzene 0.80 0.62 -23.1 0.71 0.71 -0.14 Pentylbenzene 1.43 1.04 -27.8 1.23 1.24 1.30 Hexylbenzene 2.34 1.56 -33.3 1.90 1.93 1.53 A Retention factor calculated at 3.28 x 107 Pa (4760 psi) and 298 K B Retention factor calculated at 3.28 x 107 Pa (4760 psi) and 326 K C Retention factor calculated at 5.24 x 106 Pa (760 psi) and 309 K D Retention factor calculated at 3.28 x 107 Pa (4760 psi) and 309 K 36 v. 8... 96 .6... 8.... an. .o. x mm.” 6 86.8.8 666. 8:86". a v. 8.. .26 8.. RE 6.. .o. x cm... 6 86.8.8 666. 8:86”. o v. cum 6.6 86 8.... 6n. .o. x mm... 6 86.8.8 666.. 8:86”. m v. new 6.6 8.. 8.... 6n. .o. x mm... 6 86.8.8 6.6.. 8.86m < .3 on... 8.2 «6.1 8.8 8.8. 6888.2656on m... 8.8 8.8 0.8- .3. atom 6888.2656666 or... 86 8.6 6.6. we... a... 68.86.268.666{N. o... 86 $6 6.... on... 8.. 68.82.856.63.3.. x... 8... an. 36. mm. 8.. ocoueoa.>..6&_..-m.m.. new «.6 8a 6.....- w... New 68.82.568.64... as... 86 men can- 86 8d 68883565...-QN. an... 8... 8.0 new- NS cad 683x... 6.5. «No 6.... o.-- 8... Ned 68.32.. 8.... .6... .6... sea. a... 8... 68.38 a..- «no one «6.- .2. cm... 886. 8.. a... c... a... c... 0.... 888m 3... .52 ox ox 3... 35.4 cc. 5omecc .3. 652335688ch .2. 6523352566.. .AIV 9.63.3... .3 9.622 6928 SEE... mmmd .cma 8...: mm .o. x .m... .9556... ”woman .2522 .658 enigma 36.8 EEEoO .m. cozmaam 9 9.5.0066 365:3 5.0:. 5 39.626 65 3.6528 2 now: .3: 6.58953. 3.62: m...m.m> c: .905 55:96. 65 .6 38¢ 65623683”. ”Wu 2:2". AL: 0.3.6.8Eot. 38...... 8.8.... 8.8.... $8.... $8.... $8... 8.8.... .38.. 86.8.. I 4 1 . . I . rd 9 n I . TI 8 m. 9 U m. . c. U a T .0. O J all . 8. 8e. em 6.8... 40 positive. A linear graph indicates that the change in molar enthalpy is constant with temperature. A positive slope is demonstrative of a negative change in molar enthalpy, which indicates that the transition from the mobile to stationary phase is an energetically favorable exothermic process. Molar enthalpies for alkylbenzenes and methylbenzenes at different pressures are summarized in Tables 2.4 and 2.5. For the alkylbenzenes (Table 2.4), the change in molar enthalpy with pressure is very small. At constant pressure, the homologous series illustrates a trend of decreasing molar enthalpy, where retention becomes more exothermic as the number of methylene groups increases. The differential change in molar enthalpy (AAH...) versus carbon number for alkylbenzenes is shown in Figure 2.6. From benzene to toluene, AH... changes about -3 kJ/mol. But from toluene to propylbenzene, the change in AH... is very small (—0.5 kJ/mol). This small change indicates the addition of one methylene group at this stage does not significantly affect the retention. Thereafter, the change in AH... becomes relatively constant at approximately -2 kJ/mol. The methylbenzenes (Table 2.5) also illustrate decreasing molar enthalpy as the number of methyl groups increases. Solutes with the same number of methyl substituents have similar molar enthalpies. For example, o-, m-, and p- xylenes have comparable molar enthalpies of -8 kJ/mol. The same is true for 1,2,3-trimethylbenzene, 1,2,4-trimethylbenzene, and 1,3,5-trimethylbenzene with molar enthalpies of —11 kJ/mol. The differential change in molar enthalpy versus carbon number for methylbenzenes, shown in Figure 2.7, is relatively constant at 41 V. 8.. o. .8 u . e8 .6... 8.3 6.. .2 x 8... u n. 6 86.8.8 .513 .3658 66.2 w v. .8 o. 68 u . e8 .6. 8.... 6.. .o. x 8.8 u n. 6 86.8.8 .513 .3658 66.... o v. 88 o. 68 u . c8 .6. 8.8. 6.. .2 x 8.. u n. 6 86.8.8 6.13.1658 66.2 c v. .8 o. 68 u -. e8 .6. 8... 6.. .e. x 8.. u n. 6 868.8 .513 .3658 66.2 m v. .8 o. 68 u . c8 .6... 8.. 6n. m.... x 8.6 u n. 6 86.8.8 .513 .5658 6.22 < we... a 8. . .- 8... a 8. . .- 1.... a 8....- 8... a 8. . .- 8... a 8. . .- 68.8....on- 8... a 6...- .m... a 8...- 8... u. .6...- ...o a 8..- .m... a .8...- 68.85.58,. 8... a 8..- 8... a .6..- 8... a 8..- .4... a ....- $8 a 8..- 68.86.58 8... a o...- 8... a 8...- 8... a 8...- 8... a 8...- 8... a 8...- 68.8.5.8... 8... u :2.- c.... u. .6...- m..... a 8...- 8... u. .8- 8... a 8...- 68.82.56 - 8... a 8.6 .6... a .31 6.... a 8.1 .8... u. 8...- .6... u. :6- 68:6. .8... a 8. .- 8... a 8..- 8... a 8... .6... a 8...- 8... u. 8.8- 68.86 m.65.}. ._..o......... 08.5.... .5.95.... (8.5.... 5.86 .61.. 2.16 5.1.. 5.14. 2.1.. .5563 35.5.65 3:90.. :o m3:3~:3£>x_6 .o. 385:6 .6_os_ 3N 636... 42 v. ..8 o. .8 u . e8 .65 8.3 6.. .... x 8... u 5 6 86.8.8 .513 5658 66.2 w v. .8 o. 88 u .- .26 .65 8.... 6.. .... x 8.8 u n. 6 86.8.8 .513 5658 6.6.2 a v. 8.. o. 8. u . e8 .65 8.3 65 .... x 8.. u 5 6 86.8.8 .513 .5658 665. e v. .8 o. 88 u . ..8 ...-... 8... 6.. .... x .N. u n. 6 86.8.8 .513 5658 66.2 m v. .8 o. 88 u e ..8 .65 8.. 6n. m.... x 8...- u 5 6 86.8.8 .513 .5658 66.2 < 8... a 8....- .~.o « 8.5.- 6.... a 8....- .m... a 8.8- 8... u 8.8- 68.85.568.61 N... . .6... 8... a 8....- ...... a 8....- 8... 8......- 8... a 6......- 88.85.568.85 6.... a 8....- 8... a 8.....- .... a 8.6.- 8... a 8.6.- 8... a 8....- o..o~..o......_.c8..o.-m....m.. 8... a 8....- 8... a 86.- 8... a 8....- 8... a 8.8.- ...... a 8....- o..o~..o........o8.6.-m...-.8.. ...-.... a .6...- 8... u." .3...- m... a 8.... 8... a 8...- ...5 « ......- o..o~8.......6......-m.m.. ...... a 8....- 8... a 8. . .- ...... a .... . .- 6.... a 8...- 8... « ... . .- 68.85.565.61... 8... a 8....- ..8 a 8. . .- 8... a 3.. . .- 8... .. ... . .- ...... a .... . .- c..o~8........oe_..-m.~.. 8... a 8.5- 6.... a 8...- N... u 8...- 8... a 8...- 8... a 8...- 88.3.... 8... « .....- 8... a 8..- .... « ......- 8... u. 8..- 8... . ....- 683..-... 8... a 8..- ...V... a 8..- 6.... . .....- ...... . ......- o... u. 8...- o8...x-o 8... a 8.... ...... « ......- c.... a 8.6 ...... u 8..- ..c... a ...m- 68:6. .8... a 8. .- 8... a 8..- 8... a ....m- ...... a 8.8- 8... a 8.8- cam-.86 m...o......... n...o.....-... o85...... m.5952. «8.5.3.. ...-.666 51a 51< 51< 514 51< 53.63 353565 636.65 :6 63:3~:6n_>53E .o. 385:6 .665. "m." 636... 43 .v. .8 5 8. u . ..8 .65 8.. 65 m.... x vmh u a 6. 3135233 .0. 85:5: :36... 389 1252. .13.. .8855 BEE :_ mmcmco 83:93.5 6.N 239". .3832 .8960 N. m m V m N _. o ‘0. “3 IO flow/m) Aduamua JBIOW u! afiuaqo lama-lama 9. ‘7‘ 0. F I ‘0. c? o.o .... 285 .x .8 o. 8.. ... . ..8 .65 8.. 65 6... x vm... u n. .6 $523350... .0. .35.... 558 many-...; 99...}. .13. .6855 .66... 5 mmcmco 551905... ...u 952“. .3832 23.60 h m m V m N _, o . . . .II J . 0.? o...”- mN- me. o...- o 8‘ (low/N) “lat-DUB JaIOW u! 95U9UO lenueJe-um m6. 0.0 ..N 5595 45 around —2.5 kJ/mol. This value is slightly larger than that for the alkylbenzenes when the carbon number is greater than three. The constant change suggests that the addition of each methyl group produces a similar effect on retention. The methylbenzenes also demonstrate very small differences in molar enthalpies at different pressures. For the high-density octadecylsilica stationary phase, the change in molar enthalpy (AH-m.) ranges from —44 kJ/mol to —128 kJ/mol for fatty acids ranging from C10 to C22, respectively, at 303 K and 1.01 x 107 Pa (1470 psi) [15]. The differential change in molar enthalpy per ethylene group remains constant for solutes C10 to C13 with an average value of -15 kJ/mol [15]. Thus, when PGC is compared with octadecylsilica, the addition of each methylene group has less effect on retention. 2.4.1.3 Molar Volume Similar to the molar enthalpy, the molar volume also provides insight into the effects that arise from structure. Using Equation 1.10 and the methods described in Section 1.2.1, the change in molar volume was calculated. A representative graph used in the calculation of the change in molar volume is depicted in Figure 2.8. For all solutes at all temperatures, the data are relatively linear. However, because of the small change in retention factor with pressure, the square of the correlation coefficient is small (R2 = 0.549-0.889). A linear graph indicates that the molar volume is constant with pressure. The slope of the line is slightly positive or negative. A negative slope results in a positive change in molar volume, which indicates that the molecule occupies more space in the 46 ..0. 55255355555; .3. 55525535559555 ..4. 5c5~5n355E_...m.m.. .... 553...... .... 5522 .55....ow .553... mm... .x .mm .6555... ”555.... 5.30.2 .5555 5555.5 55.3 ...ano .0... 55539... 0. 9.5.8.5 5523 .56... c. 555.... 55 5.53255 o. 555: 5.3555... m:m.5> .255. 555.5. 55 .o 55.5 555555.55”. .5.N 5.52". .55.. 5.335... no+momd ho+woo.n ho+momd ho+moc.N no+womé ho+wooé oo+moo.m oo+woo.o 5 q 5 J J . v.0 ? A ? loll, l6 ll lIT $l T d. P {n {N If la. [G a w W m. u I: 3 P m o. oo. ...m 5.3.... 47 stationary phase than in the mobile phase. Conversely, a positive slope indicates that the transition from the mobile to stationary phase results in a smaller volume. Representative data are contained in Tables 2.6 and 2.7 Overall, the change in molar volume is very small on PGC, compared with that of polycyclic aromatic hydrocarbons (PAHs) on octadecylsilica, which is around -2 cm3/mol [16]. The change in molar volume for methylene homologues on high—density octadecylsilica ranges from -40 to -110 cm3lmol [15], which is much greater than that on PGC. Considering the uncertainty in measurement, the change in molar volume on PGC is very close to zero. This is consistent with the retention mechanism on PGC, which is based on adsorption and desorption of the solute molecules from the solid surface, rather than partition of the solute molecules into the stationary phase as on octadecylsilica. 48 Table 2.6: Molar volume for alkylbenzenes on porous graphitic carbon. AVsm AVsm Solute (cm3lmol)A (cm3lmol)8 Benzene -0.24 :1: 0.22 0.23 :l: 0.20 Toluene 0.25 1 0.19 0.16 1 0.09 Ethylbenzene 0.18 1 0.17 -0.25 1 0.21 Propylbenzene -0.35 1 0.21 0.25 1 0.19 Butylbenzene 0.23 1 0.20 0.29 1 0.15 Pentylbenzene -0.29 1 0.15 0.36 1 0.26 Hexylbenzene 0.33 1 0.26 -0.24 1 0.20 A Molar volume (AV...) calculated at T = 296 K and P = 5.24 x 10° Pa (760 psi) to 3.28 x 107 Pa (4760 psi) '3 Molar volume (Av...) calculated at T = 327 K and P = 5.24 x 106 Pa (760 psi) to 3.23 x 107 Pa (4760 psi) 49 Table 2.7: Molar volume for methylbenzenes on porous graphitic carbon. AVsm Avsm Solute (cm3/mol)A (cm3/mol)B Benzene -0.24 1 0.22 0.23 1 0.20 Toluene 0.25 1 0.19 0.16 1 0.09 o-Xylene 0.28 1 0.19 -0.22 1 0.21 m-Xylene -0.35 1 0.21 0.23 1 0.19 p-Xylene 0.24 1 0.20 0.21 1 0.15 1,2,3-Trimethylbenzene 0.29 1 0.15 -0.36 1 0.26 1,2,4-Trimethylbenzene 0.33 1 0.26 0.24 1 0.20 1,3,5—Trimethylbenzene -0.34 1 0.23 0.29 1 0.20 1,2,3,5-Tetramethylbenzene 0.36 1 0.26 -0.46 1 0.39 1,2,4,5-Tetramethylbenzene 0.34 1 0.17 0.37 1 0.21 Pentamethylbenzene -0.55 1 0.21 0.61 1 0.36 Hexamethylbenzene 0.63 1 0.28 0.69 1 0.34 A Molar volume (AV-m.) calculated at T = 296 K and P = 5.24 x 106 Pa (760 psi) to 3.23 x 107 Pa (4760 psi) '3 Molar volume (Av...) calculated at T = 327 K and P = 5.24 x 106 Pa (760 psi) to 3.28 x 107 Pa (4760 psi) 50 2.4.2 Kinetic Behavior 2.4.2.1 Rate Constants Although the thermodynamic data demonstrate the steady-state aspects, they do not fully explain the mechanism of retention. Using Equations 1.11, 1.12, 1.14, and 1.16 and the method described in Section 1.2.2, the pseudo-first-order rate constants, activation enthalpies, and activation volumes were calculated. These values help to quantify the kinetic aspects of solute transfer between the mobile and stationary phases. Representative values of the rate constants are summarized in Tables 2.8 and 2.9. For the alkylbenzenes (Table 2.8), the structures with longer alkyl chains result in larger rate constants for the mobile to stationary phase (km), and stationary to mobile phase (km) transitions. From benzene to butylbenzene, the rate-limiting step is the transfer from mobile to stationary phase (i.e., kms > ksm) because the retention factor, which is the ratio of km and kms, is less than 1. 51 Table 2.8: Kinetic rate constants for alkylbenzenes on porous graphitic carbon. Solute kms (s“)A ksm (s")B Benzene 5.9 1.0 Toluene 13.1 4.9 Ethylbenzene 1 1.8 4.5 Propylbenzene 1 5.0 8.4 Butylbenzene 16.7 14.0 Pentylbenzene 16.3 24.5 Hexylbenzene 1 5.4 37.4 A Rate constants for the stationary to mobile phase transition at 5.24 x 106 Pa (760 psi) and 295 K. 3 Rate constants for the mobile to stationary phase transition at 5.24 x 106 Pa (760 psi) and 295 K. 52 Table 2.9: Kinetic rate constants for methylbenzenes on porous graphitic carbon. Solute kms (5'1)A ksm (8-1)8 Benzene 5-9 1'0 Toluene 13.1 4.9 o-Xylene 13.9 14.2 m-Xylene 16.3 14,0 prlene 15.0 14.9 1,2,3-Trimethylbenzene 14.5 44.9 1 ,2,4-Tnmethylbenzene 145 33.1 1 ,3,5-Trlmethylbenzene 20.4 38.4 1,2,3,5-Tetramethylbenzene 3 5 25 5 1,2,4,5-Tetramethylbenzene 16 0 120 6 Pentamethylbenzene 3 5 99 o Hexamethylbenzene 2 6 292 o A Rate constants for the stationary to mobile phase transition at 5.24 x 10'5 Pa (760 psi) and 295 K '3 Rate constants for the mobile to stationary phase transition at 5.24 x 106 Pa (760 psi) and 295 K 53 However, for pentylbenzene and hexylbenzene, the rate-limiting step is the transfer from stationary to mobile phase (i.e., kms < k3...) because the retention factor is greater than 1. For the methylbenzenes (I' able 2.9), generally the molecules with more methyl substituents produce larger rate constants for the mobile to stationary phase (km), and stationary to mobile phase (km) transitions. For benzene, toluene, and the xylenes, the rate-limiting step is the transfer from mobile to stationary phase (i.e., kms > ksm) because the retention factor is less than 1. But for the trimethylbenzenes, tetramethylbenzenes, pentamethylbenzene, and hexamethylbenzene, the rate-limiting step is the transfer from stationary to mobile phase (i.e., kms < ksm) because the retention factor is greater than 1. In addition, an increase in temperature yields an increase in the rate constants for all molecules. For example, the rate constant (km) for hexylbenzene ranges from 15.4 to 32.8 s'1 for temperatures from 295 to 327 K, respectively, at 5.24 x 106 Pa (760 psi). In contrast, an increase in pressure yields only slight changes. The rate constant (km) for hexylbenzene ranges from 23.6 to 20.4 s'1 for pressures from 5.24 x 106 to 3.28 x 107 Pa (760 to 4760 psi), respectively, at 309 K. Overall, the data indicate that the rate constants (km and ksm) increase as the retention factor increases, which is the reverse order of the rate constants on octadecylsilica for PAHs [16]. 54 2.4.2.2 Activation Enthalpy Using Equation 1.14 and 1.15 and the methods described in Section 1.2.2, the activation enthalpies were calculated. Figure 2.9 illustrates a typical graph of the natural logarithm of the rate constant versus inverse temperature used in the calculation of the activation enthalpy. For all solutes at all pressures, the data are linear (R2 = 0443-0869) and the slope of the line is negative. A linear graph indicates that the change in activation enthalpy is constant with temperature. A negative slope is demonstrative of a positive change in activation enthalpy, which is the energy barrier for the transfer from mobile or stationary phase to the transition state. Tables 2.10 and 2.11 contain the activation enthalpies for alkylbenzenes and methylbenzenes. As demonstrated for the alkylbenzenes (Table 2.10), an increase in the number of methylene groups results in an increase in the activation enthalpy. The data indicate that the energy barrier from the stationary phase to the transition state AH;s increases with the addition of every methylene group. The differential change in AH” versus carbon number for the alkylbenzenes is shown in Figure 2.10. From benzene to toluene, AAHts is about 3 lemol. But from toluene to propylbenzene, AAHts increases only slightly at 0.5 kJ/mol. When the carbon number is greater than three, AAH¢3 is relatively constant at around 2 kJ lmol, indicating that the effect of each methylene group on activation enthalpy is relatively constant. 55 .on 33:33:3me0; AS ocoNcoe>£mEman A: ocouco£>£oEtfmd€ .3 oco_>x.E .3 9.3.9 $320.0, .EEDE ommd .cma omRV an. N3x23 .6559: ”$ch 250.2 .598 0:236 328 “c6200 .34 cozmacm 9 mEEooom €355 congfim c_ 09.9.0 9: 05.3.8 2 cow: 6: 938389 0995 use"; Assoc ESmcoo Em. 9: co coma 026E393”. 6N 2:2". AL: 23235.net? modmvd moflotm mo.mmm.m mofioflm moflmmd moflomd moflmfim moflofim mo-wmo.m mo-moo.m _ u . 1 u _ a 4 F 9 9 T1 + o z 2 w I! + ll + I M O T 14 'il iii” l‘ W: U «I N m u . i z 2: mt ml lml in + b coo? mm 9:9“. 56 .5: 85 an. .2 x 30 a 85328 90; .513 0.0.0 533:0: 9 000...: 0.59: 05 So: 0:0 .313 0.0.0 53.20:. 2 000:0 30:95.0 0.: Eo: 00.2055 :o=0>=o< n ...ma 8: an. .2 x v3 .0 86.328 90; 313 0.0.» 53.20: 2 000:: 0.59: 05 So: 0:0 .313 0.0.» 53.30: o. 000:: tacos—30 05 Ea: 006.055 5302.3. < mod « mud $3 a $3 5.0 a 2d mod « 5.2 05283me 8.0 a 80 23 «EN: 0: a B: «3 a o... z 2022.35: 3.0 a 8.... who “am. 3 «no « oz” 3.0 a 0:: 0523350 3.: a Se 86 a mg 1.: a new Re a new 0520239: 8.0 a 03 one a 2.0 03 « Rd F? a one 0523350 Bo 0 Ed 8; a 8.... 93 a m: :3 a 3.0 0:023 8... a one who a 3.” Ed a 91 N3 « Ed 0:250 .65.}. 2952. .655. (.655. 228 .55. 3.2 gf< 3% doemo 03.2005 0:92. :o 00:05:85.0 ho. 3.05:0 82028.0. ”Emu 030:. 57 000 0000. 0n. .0, x 00.0 00 00.0.8.8 00; 31$ 0050 550:0: 8 0005.. 0.50.: 05 So... 0:0 .313 05050 :o=.0:0= on 0005. E05500 05 Eo: 005.055 55028.0. 0 .000 000. 0.. 005 x 00.0 .0 00.0.8.8 203313 00.050 3.00:0: B 0005. 0.59: 05 So... 0:0 .313 0050 55.00.05 2 000.5 205.00 05 So: 005.055 5.0280. < 00.0 a 00.00 00.0 a 00.00 0.0 a 3.00 00.0 a 00.0... 00000003505000: 0: a 00.0 0.? a 00.00 03 a 0.0 00.? a 00.00 00000020505900; 00.? 0 00.0 00.? a 00.0. 00.? a .00 0: a 00.0, 00020030500000{0.r 0.0 a 00.0 00.0 a 00.0, 0.0 0 00.0 00.0 a 00.: 00000003505008-00.0.F 03 a 00.0 0.0.. a 2.2 :5 a 00.0 0: 0 00.3 00000003505500; 0: a 0:0 3.. 0 v0.0, 03 a 00.0 00; a 00.2 00000003005510; 00., a 00.0 00.0 a 0.0? 5.0 0 00.0 00.? 0 00.2 00000003505500; 00.? a 00.0 00.? a 0.0 00; a 00.0 05 0 2.0? 000.33 00.0 a 00... 00.? a 00.: 00.? a 00.0 00.0 a 00.2 000.020 0.0.0 0 00.0 00.0 0 00.0? 00.0 a 00... 00.0 a 00.0F 000.300 00.0 0 v0.0 00.? a 00.0 00.0 a 0: 0.0 0 00.0 0000.00 00.0 a 00.? 00.0 a .00 .00 a 0.1 00.? a 00.0 0000000 00.53.. 0005.30 005.30 (005.30 20.00 5:0 4:0 5:0 0:0 .:o€0o 055006 0:92. :o 00:00:05..»50... .8 3.05:0 550590. 3. ...N 050... 58 :00 cot 0n. 09 x 000.0 00 0:00:0n_>v..0 :8 .0080: :0500 000.01.080.05 0".....3 00000 55.0:05 2 0005. E05530 05 82.. 3.05:0 :050500 :_ 00:05. 05:90.50 5..." 0.59... 03:52 :0500 o... m.—. 0N (low/m) Adlemua uouamov U! efiueuo lauuwomo m.m 00.0 050E 59 For the methylbenzenes (Table 2.11), the activation enthalpy from the stationary phase to the transition state AH” increases with increasing number of methyl substituents. This result suggests that with the addition of each methyl group, the energy barrier from the stationary phase to the transition state increases. Figure 2.11 shows the differential change in AHp versus carbon number for the methylbenzenes. From benzene to tetramethylbenzenes, the differential change in activation enthalpy AAHts is relatively constant at around 3 kJ lmol. But for pentamethylbenzene and hexamethylbenzene, AAHt. is much greater. For both alkylbenzenes and methylbenzenes, the change in activation enthalpy from the mobile phase to transition state AH;m is nearly constant, considering the uncertainty in the kinetic data. But AH;m for pentamethylbenzene and hexamethylbenzene is greater than that of other solutes in the same homologous series. Comparison between AHB and AH;m indicates that the retention on PGC is controlled by desorption, which is decided by the molecular structures of the solutes. Thus, the structure of the molecules affects the energy needed to transfer between mobile and stationary phases. Comparing the activation enthalpy with the absolute value of molar enthalpy, the trend is AH” > AHsm > AHtm. As expected, AH,"m is nearly equal to the difference between AH” and AHtm. Comparing with the activation enthalpy for PAHs on octadecylsilica [16], the activation enthalpy for PGC is much smaller. This indicates that the energy barrier for retention on PGC is much smaller than that on octadecylsilica. 60 :00 act 0n. 00.. x 0.0.0 .0 00:00:003505 :8 .0080: :oE0o 0:0:0> cog—é .013 05000 55.055 2 0005. E03590 05 Eo: 3.05:0 550500 :_ 00:05 05:90.05 2. ...N 0.52“. 03:52 :3000 n m m v m N F o i o a. 0 w 9 W O W u e 0 u m w m. 9. 3 u m m. M m 00 0:. mm :0 050.“. 61 2.4.2.3 Activation Volume Using Equation 1.16 and the methods described in Section 1.2.2, the activation volumes were calculated. A representative graph of the natural logarithm of the rate constant versus pressure is depicted in Figure 2.12 and resulting values for the activation volume are presented in Tables 2.12 and 2.13. The slope of the line is either slightly positive or slightly negative. The activation volumes for the stationary phase to transition state (Av...) and the mobile phase to transition state (Av...) appear to be larger than the change in molar volume (Avg...) because of greater uncertainty of the kinetic data. However, these values are also nearly zero. They are much smaller than the activation volumes reported for methylene and benzene homologues on octadecylsilica [15]. This suggests that the volume barrier on PGC is much smaller than that on octadecylsilica stationary phase. 62 .00. 0:00:0n.>50E0x0: .3. 0520035050050 .2. 0:00:00_>50E.5.m.m.0 .3 05.02: .... 0:022 ”005.00 .:.EDE 00.0 .v. Rm ..o:050E ”000...: 0.322 .:o€00 05.5.0.0 0:92. “:ano .0: :05000m. 2 0508000 9:29 550500 :_ 00:0:0 05 20.00.00 2 000: 050005 0:0:0> E00050 000: 05 .0 5.000 050.500.0000 ”00.0 0.50.“. .00. 05000.0 no+momd ho+mood ho+momd ho+mood no+mom. F ho+moo. F 00+woo.m oo+mood . . l . . 1 F f + O 4* L o r H I F A [T I ”4 O O U .5 T L1] Gr 11 L m. 00. all if [T T 4 00. ( T + 1m T in coo F 00.0 0.00.... 63 Table 2.12: Activation volume for alkylbenzenes on porous graphitic carbon. Av... Av...n Av... AV¢m Solute (cm3/mol)A (cm3lmol) (cm3/mol)l3 (cm3lmol) Benzene -0.94 :l: 0.92 0.93 1: 0.70 -1.24 :h 0.92 1.23 :I: 1.20 Toluene 1.15 :l: 0.69 0.76 :l: 0.89 1.25 :l: 1.19 -1.16 :l: 0.99 Ethylbenzene 1.18 :l: 0.97 -1.05 :l: 0.91 -0.98 :l: 0.87 -0.95 :l: 0.91 Propylbenzene 1.35 :t: 1.21 0.95 :l: 0.69 -1.15 1 1.21 1.25 :l: 0.89 Butylbenzene 1.33 :1: 0.90 -1.29 :l: 0.85 0.99 :1: 0.80 0.73 :l: 1.15 Pentylbenzene -1.29 :t 0.85 0.96 :1: 0.76 -1.36 :l: 1.25 0.99 :l: 0.76 Hexylbenzene 1.33 :l: 0.76 -0.84 :l: 0.70 1.33 :l: 0.96 -1.24 :l: 0.90 A Activation volumes for the stationary phase to transition state (Ava), and mobile phase to transition state (Av...) were calculated at T = 296 K and P = 5.24 x 106 Pa (760 psi) to 3.28 x 107 Pa (4760 psi) 3 Activation volumes for the stationary phase to transition state (AV;), and mobile phase to transition state (AVm) were calculated at T = 327 K and P = 5.24 x 10‘5 Pa (760 psi) to 3.28 x 107 Pa (4760 psi) .0 0.0 0.000 0_ 000000 0 .< 0.0 0.000 0. 000000 < 00.. 0 00. .- 3.0 0 00.0 00.0 0 00.0 00.. a 000. 0000000350530... 000 0 ....0- .0. 0 00.0 0.... 0 .0. .00 0 00.0 - 00000020505950 .00 0 00.. 0.0 0 00.0 00.. 0 00.. .00 0 00.0- 00000025020000-......0.. 00.0 0 .00 00.0 a 00.0- 00.. 0 00. .- 00.0 0 00.0 0000000350500000.00.. 00.. 0 00.. 00.. 0 0.0- 00.. 0 00.. 00. a 00.0 0000:00_0...0E_.0-0.0.. 00.. a ...... 00.0 0 00.0- 0.... 0 00. .- 00.0 0 00.0 0000000_0£0E_.0-0_0.. 00.. ... 00.. 0.0 0 00.0- 0.. 0 00.. 00.. 0 00.0- 0:00:00_0£0EE-0.0.. 00.. ... ..0. 0.0 u .00 0... 0 .0..- 00. u ..0. 00053 .0. 0 00.. - 0.... 0 00.. 0... 0 00.. .0. 0 00..- 0003.-.: 00.0 0 00.0- .00 0 00. .0. 0 00. .- 00.0 0 00.. 0005.0 00.0 0 0...- 0... H 00.. 00.0 0 00.0 00.0 ... 0... 000200 00.. 0 00.. 00.0 0 ..0. .- 00.0 0 00.0 00.0 0 00.0- 0000000 0000.50. fi>< 0002.00.30.20 095.50. fi>< (020.020. a>< 0.200 .5500 05.5.05 0:92. :o 00:00:3350... :00 0E:.o> 55050.... "2.0 0.0.0... 65 2.5 Conclusions To illustrate the full implications of the data presented above, it is helpful to describe the entire retention event for a single solute. For example, as hexylbenzene transfers from the mobile to stationary phase, it passes through a high-energy transition state. The creation of this transition state requires an enthalpy change of 2 lemol and a volume change of -0.84 cm3lmol. This transfer occurs with a rate constant of 37 s". As hexylbenzene transfers from the stationary to mobile phase, it passes through another high-energy transition state. This transition state requires an enthalpy change of 14 kJ/mol and a volume change of 1.33 cm3/mol. This transfer occurs with a rate constant of 15 s", which is only half of the rate constant for mobile to stationary phase transfer. Hence, the rate-limiting step is the transfer from stationary to mobile phase. This description illustrates that both the thermodynamics and kinetics are necessary to fully describe the retention event. By using the necessary theoretical foundation and instrumentation, this dissertation demonstrates a comprehensive study of retention for two series of known pollutants on porous graphitic carbon. 66 2.6 References [1] List of lists, U.S. Environmental Protection Agency, July 1987. [2] S. B. Howerton, C. Lee, V. L. McGuffin, Anal. Chim. Acta, 478 (2003) 99-110. [3] S. N. Chesler, S. P. Cram, Anal. Chem., 43 (1971) 1922-1929. [4] J. V. H. Schudel, G. Guiochon, J. Chromatogr., 457 (1988) 1-12. [5] W. E. Barber, P. W. Carr, Anal. Chem., 53 (1981) 1939-1942. [6] J. L. Wade, A. F. Bergold, P. W. Carr, Anal. Chem., 59 (1987) 1286-1295. [7] P. C. Haarhoff, H. J. Van der Linde, Anal. Chem., 38 (1966) 573-582. [8] E. Grushka, Anal. Chem., 44 (1972) 1733-1738. [9] E. Grushka, J. Phys. Chem., 76 (1972) 2586-2593. [10] W. W. Yau, Anal. Chem., 49 (1977) 395-398. [11] J. P. Foley, J. G. Dorsey, J. Chromatogr. Sci., 22 (1984) 40-46. [12] M. S. Jeansonne, J. P. Foley, J. Chromatogr. Sci., 29 (1991) 258-266. [13] M. L. Phillips, R. L. White, J. Chromatogr. Sci., (1997) 75-81. [14] J. R. Torres-Lapasio, J. J. Baeza-Baeza, M. C. Garcia-Alvarez-Coque, Anal. Chem., 69 (1997) 3822-3831. [15] V. L. McGuffin, S.-H. Chen, J. Chromatogr. A, 762 (1997) 35-46. [16] S. B. Howerton, V. L. McGuffln, Anal. Chem., 75 (2003) 3207-3216. 67 Chapter 3: Conclusions and Future Directions Porous graphitic carbon (PGC) is a nonpolar reversed-phase adsorbent. Wrth aqueous or aqueous-organic mobile phases, porous graphitic carbon has stronger retention for the same solutes than the alkyl-bonded silica gels. Increasing the hydrophobicity of a solute by the addition of a methylene group or other nonpolar groups increases retention. In addition, by virtue of its great adsorptive strength, PGC can be used with nonpolar mobile phases for the retention of polar analytes. Porous graphitic carbon has proven to have a number of unsuspected properties, which have substantially enlarged its area of applications and opened up entirely new possibilities. However, the investigations of molecular contributions to retention on PGC have been limited primarily to selectivity factor [1,2]. Virtually no experimentation has sought to characterize the change in molar enthalpy and molar volume, which provide insight into how molecules interact with the stationary phase. In addition to thermodynamic information, the kinetics of retention have also been largely overlooked. The kinetics of retention are important because they provide insight into the rate at which transitions occur, as well as the energy barriers that exist for a molecule to move from one phase to another. 3.1 Results In the research presented in this dissertation, solute transfer of two series of aromatic hydrocarbons (alkylbenzenes and methylbenzenes) has been studied as a function of molecular structure. By quantitating the thermodynamics and 68 kinetics together, a better description of retention on PGC is made, providing a more accurate comparison of different solutes and stationary phases. In the thermodynamic studies, the retention factor of alkylbenzenes increases with increasing number of methylene groups, and the retention factor of methylbenzenes increases with increasing number of methyl substituents. For the same solute, the retention factor (k) decreases with an increase in temperature, but does not change significantly with pressure. The data from this study indicate that increases in the number of methylene and methyl groups result in more negative changes in molar enthalpy (AHsm). The more negative change in AH8m suggests that the transition from the mobile to stationary phase is a more energetically favorable exothermic process with a methylene or methyl group added. The data also show that the change in molar volume (AV...) is close to zero on PGC, which indicates that during the retention, the solute molecules adsorb on the surface of PGC instead of penetrating further into the stationary phase, as in octadecylsilica. In the kinetic studies, the rate constants of alkylbenzenes and methylbenzenes increase with increasing number of methylene and methyl groups. The enthalpic barriers between the stationary phase and transition state (AH...) were found to increase with increasing number of methylene and methyl groups as well. But the enthalpic barriers between the mobile phase and transition state (AHm) were much smaller than AH;s and, at the same time, very similar for all the solutes. The values of enthalpic barriers suggest that the retention on PGC is controlled by the adsorption and desorption of the solutes on 69 the stationary phase. The volume barriers AV... and AV,m again are very close to zero due to the adsorption mechanism on porous graphitic carbon. 3.2 Future Directions In order for a complete and quantitative description of retention to be presented, a series of experiments beyond those presented herein can be done. The first series of experiments that could lead to a more comprehensive explanation of retention would be a systematic study of the effect of mobile phase composition, such as methanol/water or acetonitrile/water, on the derived values. Similar to varying the mobile phase, a series of stationary phases, such as phenylsilica, octadecylsilica, and polystyrene divinylbenzene, could be used to compare the effect of solute structure on retention. Porous graphitic carbon exhibits unexpected and largely unexplained retentive properties for molecules containing polar groups. This property makes PGC particularly useful for the separation of highly polar compounds that would be difficult to retain on octadecylsilica phases. To help understand the retention of polar solutes on PGC better, a number of environmentally important polar solutes, such as atrazine and its polar metabolites can be used [3]. Porous graphitic carbon also shows unusual discrimination of closely related stereoisomers [4]. Chiral separations can be achieved when the molecule is very flexible and can adapt to the flat surface of PGC. The study of the retention of some stereoisomers on PGC can help to better understand the 70 ‘1- 1 ”:3 m A El“! 2 ”£542 ' retention mechanism. These position isomers can be the o-, m-, p- substituted benzene derivatives. In conclusion, the adsorption retention mechanism on PGC is totally different from that of octadecylsilica, the most widely used stationary phase in liquid chromatography. A better understanding of the retention mechanism on PGC will advance its applications in separation. 71 3.3 References [1] J. Kriz, E. Adamcova, J. H. Knox, J. Hora, J. Chromatogr. A, 663 (1994) 151- 161. [2] Q. H. Wan, P. N. Shaw, M. C. Davies, D. A. Barrett, J. Chromatogr. A, 697 (1995) 219-227. [3] M. C. Hennion, V. Coquart, S. Guenu, C. Sella, J. Chromatogr. A, 712 (1995) 287-301 . [4] M. Josefsson, B. Carlsson, B. Norlander, Chromatographia, 37 (1993) 129- 138. 72 -1 ‘ “ ‘ ' .l "u .‘ l\‘ -'.Ilu llilililllliillilflllljflijii