6
1052 LCGC NORTH AMERICA VOLUME 30 NUMBER 12 DECEMBER 2012 www.chromatographyonline.com T (°C) ln k 210 7 6 5 4 3 2 200 190 180 170 160 150 140 1/T (K -1 ) 0.00207 0.00217 0.00227 0.00237 David S. Jensen * , Thorsten Teutenberg , Jody Clark , and Matthew R. Linford * * Department of Chemistry and Biochemistry, Brigham Young University, Provo, Utah; Institut für Energie- und Umwelttechnik e. V., Duisburg, Germany; and Selerity Technologies Inc., Salt lake City, Utah. Direct correspondence to: [email protected]. Elevated Temperatures in Liquid Chromatography, Part III: A Closer Look at the van ’t Hoff Equation Part I of this article series discussed some of the advantages of and practical considerations for elevated temperature separations in liquid chromatography (LC) (1). Part II reviewed some of the basic thermodynamics of chromatography and elevated temperature separations, which included a brief derivation and discussion of the van ’t Hoff equation (2). This third and final part continues the exploration of elevated temperatures in LC with a more detailed discussion of the van ’t Hoff equation, exploring its usefulness and relevance using various examples from the literature. V an ’t Hoff plots can be a useful and interesting part of data anal- ysis for high-temperature liquid chromatography (LC). Here, in part III of this series, we take a closer look at the van ’t Hoff equation using various exam- ples from the literature. Review of the Advantages of Elevated- Temperature Separations Elevated temperatures offer a number of benefits in LC. One such benefit is that they facilitate retention mapping in which the retention factor, k, is measured at dif- ferent temperatures so that the values of k over a range of temperatures can be pre- dicted (3,4). Retention mapping can also include the probing and predicting of k at different mobile-phase compositions and is widely used in method development (3,5,6). Another benefit is that selectiv- ity (α) may change with temperature, which is important for retention map- ping and is another parameter that can be considered in method development (3,5,6). Also, increasing temperatures can improve sample throughput because the van Deemter minima shifts to higher flow rates; that is, the optimal efficiency for a separation shifts to a higher mobile- phase velocity (7–9). Finally, a decrease in the organic modifier is possible due to the change in the polarity of water at increasing temperatures; water behaves more like an organic solvent at elevated temperatures. Furthermore, a more aque- ous mobile phase is considered “greener” because of a reduction in the amount of organic modifier used (10–14). Clearly, there are good reasons for considering the use of elevated temperatures in LC. Now, let’s discuss various aspects of high-tem- perature separations in the context of the van ’t Hoff equation. Review of the van ’t Hoff Equation The van ’t Hoff equation is derived from the following two basic thermodynamic equations: ΔG 0 = ΔH 0 TΔS 0 [1] and ΔG 0 = -RT ln K [2] where ΔG 0 is the Gibbs free energy, ΔH 0 is the enthalpy of transfer, T is the absolute temperature in kelvins, ΔS 0 is

van't hoff

Embed Size (px)

DESCRIPTION

vant hoff

Citation preview

  • 1052 LCGC NORTH AMERICA VOLUME 30 NUMBER 12 DECEMBER 2012 www.chromatographyonline.com

    T (C)

    ln k

    210

    7

    6

    5

    4

    3

    2

    200 190 180 170 160 150 140

    1/T (K-1)

    0.00207 0.00217 0.00227 0.00237

    David S. Jensen*, Thorsten Teutenberg, Jody Clark, and Matthew R. Linford**Department of Chemistry and Biochemistry, Brigham Young University, Provo, Utah; Institut fr Energie- und Umwelttechnik e. V., Duisburg, Germany; and Selerity Technologies Inc., Salt lake City, Utah. Direct correspondence to: [email protected].

    Elevated Temperatures in Liquid Chromatography, Part III: A Closer Look at the van t Hoff Equation

    Part I of this article series discussed some of the advantages of and

    practical considerations for elevated temperature separations in

    liquid chromatography (LC) (1). Part II reviewed some of the basic

    thermodynamics of chromatography and elevated temperature

    separations, which included a brief derivation and discussion of the

    van t Hoff equation (2). This third and final part continues the

    exploration of elevated temperatures in LC with a more detailed

    discussion of the van t Hoff equation, exploring its usefulness and

    relevance using various examples from the literature.

    Van t Hoff plots can be a useful

    and interesting part of data anal-

    ysis for high-temperature liquid

    chromatography (LC). Here, in part III

    of this series, we take a closer look at the

    van t Hoff equation using various exam-

    ples from the literature.

    Review of the

    Advantages of Elevated-

    Temperature Separations

    Elevated temperatures offer a number of

    benefits in LC. One such benefit is that

    they facilitate retention mapping in which

    the retention factor, k, is measured at dif-

    ferent temperatures so that the values of k

    over a range of temperatures can be pre-

    dicted (3,4). Retention mapping can also

    include the probing and predicting of k at

    different mobile-phase compositions and

    is widely used in method development

    (3,5,6). Another benefit is that selectiv-

    ity () may change with temperature,

    which is important for retention map-

    ping and is another parameter that can

    be considered in method development

    (3,5,6). Also, increasing temperatures

    can improve sample throughput because

    the van Deemter minima shifts to higher

    flow rates; that is, the optimal efficiency

    for a separation shifts to a higher mobile-

    phase velocity (79). Finally, a decrease

    in the organic modifier is possible due

    to the change in the polarity of water at

    increasing temperatures; water behaves

    more like an organic solvent at elevated

    temperatures. Furthermore, a more aque-

    ous mobile phase is considered greener

    because of a reduction in the amount of

    organic modifier used (1014). Clearly,

    there are good reasons for considering the

    use of elevated temperatures in LC. Now,

    lets discuss various aspects of high-tem-

    perature separations in the context of the

    van t Hoff equation.

    Review of the

    van t Hoff Equation

    The van t Hoff equation is derived from

    the following two basic thermodynamic

    equations:

    G0 = H0 TS0 [1]

    and

    G0 = -RT ln K [2]

    where G0 is the Gibbs free energy,

    H0 is the enthalpy of transfer, T is the

    absolute temperature in kelvins, S0 is

    ES160069_LCGC1212_1052.pgs 11.28.2012 23:01 ADV blackyellowmagentacyan

  • 1054 LCGC NORTH AMERICA VOLUME 30 NUMBER 12 DECEMBER 2012 www.chromatographyonline.com

    the entropy of transfer, R is the gas con-

    stant, and K is the equilibrium constant.

    When we equate these two equations

    and solve for ln K we achieve the van t

    Hoff equation:

    ln K = - H0/RT + S0/R [3]

    As discussed in part II (2), ln K = ln k,

    where k is the retention factor and is

    the phase ratio (VM/VS) the ratio of

    the mobile-phase volume and stationary-

    phase volume. By substituting k for K in

    equation 3 we obtain the van t Hoff equa-

    tion as it is commonly encountered in LC:

    ln k = - H0/RT + S0/R ln [4]

    Note that sometimes is used instead

    of , where = 1/ = VS/VM. Thus, an

    equivalent form of equation 4 is:

    ln k = - H0/RT + S0/R + ln [5]

    Unfortunately, both and are referred

    to as the phase ratio.

    The van t Hoff Equation

    in Retention Mapping

    Van t Hoff plots are often linear, which

    makes them useful in retention map-

    ping. For this purpose, a simpler, but

    mathematically equivalent, version of

    equation 4 can be used (3,15):

    log k = A + B/T [6]

    where plots of log k vs. 1/T are generated

    under isocratic conditions (6,1618) and

    A and B either have values as given by

    equation 5 or may be viewed empirically.

    While to some degree this semi-empirical

    equation conceals the underlying ther-

    modynamics, in most cases this is not the

    primary concern. Of course, retention

    mapping may be conveniently performed

    using commercially available software.

    When using temperature as a variable

    to optimize a separation, chromatogra-

    phers can create a series of van t Hoff

    plots for various analytes to determine

    the best temperature for the separation.

    To demonstrate this optimization pro-

    cess, Figure 1 shows a series of van t Hoff

    plots for various drugs. In this example,

    some of the drugs, such as in the circled

    lines in the plot, show different slopes

    and change elution order (reverse selec-

    tivity) where their lines cross. Obviously,

    the temperature at this crossing point

    would be a very poor choice for separa-

    tion conditions because the peaks would

    coelute, that is, = k2/k1 = 1. Thus, in a

    separation involving multiple analytes, a

    series of van t Hoff plots can be used to

    optimize the separation.

    Thermodynamics of

    Linear van t Hoff Plots

    Linear van t Hoff plots such as those in

    Figure 1 suggest that the retention mecha-

    nisms for the analytes are constant; that

    is, the values for H0, S0, and for the

    analytes are constant over the temperature

    range under consideration. Of course there

    is the possibility that H0, S0, and are

    mutually changing so that the net effect is

    a linear relationship, but this will be dis-

    cussed below. If the retention mechanism

    is constant with temperature it may be

    possible to compare the enthalpies (H0)

    and entropies (S0) of similar analytes on

    the same column, or of a single analyte

    on different columns. It should be noted

    again that the H0 transfer of the analyte

    from the mobile phase to the stationary

    phase can be derived from the slope of its

    van t Hoff plot (see equation 4) (2). When

    H0 is negative, which is typically the case

    in reversed-phase chromatography, trans-

    fer of the analyte from the mobile phase

    to the stationary phase is favored and

    exothermic. Clearly, the more negative

    H0 is the more favorable the interaction

    between the analyte and the stationary

    phase is, which generally leads to larger

    values of k. For example, in reversed-

    phase chromatography, the H0 values

    for a homologous series of increasingly

    hydrophobic analytes such as the alkyl

    benzenes with longer and longer alkyl

    chains should steadily become more nega-

    tive (exothermic). This effect is shown in

    Figure 2: The alkyl benzenes with longer

    alkyl chain lengths (more hydrophobic)

    have larger slopes than those with shorter

    chain lengths (less hydrophobic) (19).

    Nonlinearities in

    van t Hoff Plots

    Because of Phase Transitions

    As a corollary to the previous state-

    ments, a nonlinear van t Hoff plot

    shows that H0, S0, or are chang-

    ing, that is, the retention mechanism

    for the analyte is not constant over the

    temperature range under consideration.

    A possible explanation for a nonlinear

    van t Hoff plot is a phase transition in

    the stationary phase; at lower tempera-

    tures the stationary phase will generally

    be in a solid-like conformation and at

    T (C)190

    7

    6

    5

    4

    3

    2

    1

    0

    -1

    0.00211

    Aminohippuric acid Aminobenzoic acidCaffeineTheophyllineAminoantipyrine

    ParacetamolPhenacetinAntipyrineHydroxyantipyrine

    ln k

    0.00231 0.002511/T (K-1)0.00271 0.00291 0.00311

    170 150 130 110 90 70 50 30

    Figure 1: Van t Hoff plots for test probes showing linear relationships between the natural logs of the retention factors vs. 1/T for these compounds. The inset shows the point at which the elution is reversed for aminoantipyrine and caffeine. From left to right, the data points correspond to 180, 150, 120, 90, 60, and 40 C. Adapted from reference 10 with permission.

    ES160068_LCGC1212_1054.pgs 11.28.2012 23:01 ADV blackyellowmagentacyan

  • DECEMBER 2012 LCGC NORTH AMERICA VOLUME 30 NUMBER 12 1055www.chromatographyonline.com

    higher temperatures it will adopt a more

    liquid-like conformation (13,20,21).

    Thus, a nonlinear van t Hoff plot may

    indicate that the thermodynamic inter-

    actions between the analyte and sta-

    tionary phase change when the station-

    ary phase undergoes a phase transition.

    The possibility of a phase transition in

    a C18 phase seems reasonable because

    long chain hydrocarbons have melting

    points in the range of 2050 C, for

    example, the melting point of octadec-

    ane is approximately 28 C. For silica-

    based C18 stationary phases, this phase

    transition may occur in the range of

    2050 C (2224). Of course, the melt-

    ing transition of a C18 stationary phase

    is much more complicated than the

    simple melting of a pure hydrocarbon

    because of the tethering of the chains

    in the stationary phase, the density or

    packing of the chains, and the presence

    of other chemical groups in the film

    such as endcapping agents.

    Other phase transitions at higher

    temperatures have also been reported.

    For example, a phase transition around

    100 C was found for a silica-based hybrid

    C18 column (Figure 3). This transition was

    attributed to a change in the conformation

    of the stationary phase in the presence of

    the mobile phase as a function of tem-

    perature (19). This idea was substantiated

    by solid-state nuclear magnetic resonance

    (NMR) spectroscopy, which showed, over

    a temperature range of 30150 C, that

    the dry stationary phase did not undergo

    any conformational changes. Differential

    scanning calorimetry (DSC) was also per-

    formed on the stationary phase under con-

    ditions that mimicked typical LC mobile

    phase conditions. DSC showed thermal

    desorption of the mobile phase (70:30

    wateracetonitrile) from the stationary

    phase around the phase transition point

    (circa 97 C), suggesting a change in the

    conformation of the stationary phase when

    the mobile phase was present.

    The phase transition in Figure 3 pro-

    duces two linear van t Hoff plots: region

    I at lower temperatures (3297.3 C) and

    region II at higher temperatures (97.3

    200 C). Interestingly, H0 in region II

    is approximately twice that of region I.

    Coym and Dorsey (25) discussed this

    possibility of an increase in H0 follow-

    ing a phase transition:

    It may seem odd that the enthalpy of transfer (retention) at high tem-perature is more favorable than at low temperature, because retention is greater at low temperature. The ther-modynamic quantity that governs retention is the free energy (G0), which has an entropy component (S0). Because of the change in hy-drogen bond structure of water with temperature, the entropy change as-sociated with retention changes with temperature. At lower temperatures, where the mobile phase is hydrogen bonded, there is a favorable entropy change upon retention. This is com-monly referred to as hydrophobic effect. However, at high tempera-tures, where there is little or no hy-drogen bonding, the entropy change would be expected to be much less. As a result, although the enthalpy (H0) of retention is more favorable at high temperature, it is outweighed by the entropic (S0) contribution.

    Irregularities in van t Hoff

    Plots Because of pH Effects

    Unusual van t Hoff plots may be

    observed in the separation of acids and

    T (C)ln

    k210

    7

    6

    5

    4

    3

    2

    200 190 180 170 160 150 140

    Toluene Ethyl benzene Propyl benzene Butyl benzene Amyl benzene

    1/T (K-1)

    0.00207 0.00217 0.00227 0.00237

    T (C)

    1/T (K-1)

    ln k

    175 125 75 25

    3.5

    2.5

    1.5

    0.5

    0

    0.00205 0.00245 0.00285 0.00325

    Region I

    Region II 32 C to 97.3 C

    H = -4.2 kcal/mol

    H = -8.0 kcal/mol

    97.3 C to 200 C

    1

    2

    3

    Figure 2: Van t Hoff plots for a homologous series of alkyl benzenes. Increases in alkyl character result in larger slope values indicating more negative H0 values. From right to left, the data points correspond to 200, 190, 180, 170, 160, and 150 C. Adapted from reference 19 with permission.

    Figure 3: Van t Hoff plot for toluene demonstrating curvilinear behavior around 100 C. Adapted from reference 19 with permission.

    ES160066_LCGC1212_1055.pgs 11.28.2012 23:01 ADV blackyellowmagentacyan

  • 1056 LCGC NORTH AMERICA VOLUME 30 NUMBER 12 DECEMBER 2012 www.chromatographyonline.com

    bases (26) because a change in the tem-

    perature, and therefore polarity, of the

    mobile phase can affect the pKa and pKb

    values of weak acids and bases. Obvi-

    ously, a change in the ionization state of

    an analyte or buffer in a mobile phase

    can alter retention (27), and selectiv-

    ity changes associated with temperature

    changes are larger for polar and ioniz-

    able analytes than for nonpolar analytes

    (2832). In particular, an analytes pKa

    value can shift approximately 0.03 pKa

    units per degree Celsius (33,34). These

    effects are complex and are usually

    strongly manifested when pH pKa; that

    is, where both the weak acid and conju-

    gate base have appreciable concentrations

    (29). The success of these types of separa-

    tions depends on the nature of the buf-

    fer and analyte (30). Two examples are

    shown in Figures 4 and 5. Figure 4 shows

    an increase in retention of protriptyline,

    a tricyclic antidepressant, with increasing

    temperature on two different columns.

    Figure 5 also shows analytes that exhibit

    (unusual) negative slopes in their van t

    Hoff plots. As a side note, temperatures

    can also affect large molecules for

    example, proteins may undergo confor-

    mational changes with temperature (31).

    Confirming the Linearity

    of van t Hoff Plots and

    Evaluating Changes in Entropy

    Chester and Coym (35) explored the

    possibility of changing during a van t

    Hoff analysis and noted that, at least in

    theory, a change in could compensate

    for changes in H0 or S0, leading to an

    (apparently) linear van t Hoff relation-

    ship. This statement is consistent with

    some of the concerns raised by Gritti and

    Guichon (36,37); that is, different com-

    pensating or canceling factors may lead

    to the linearity often observed in van t

    Hoff plots. To eliminate this possibility,

    Chester and Coym (35) noted a slightly

    more advanced use of van t Hoff analy-

    sis in which one plots ln vs. 1/T, where

    is the selectivity (k2/k1) between two

    analytes. This relationship is obtained

    from equation 4 by subtracting the van t

    Hoff relationship for the second analyte

    from the van t Hoff relationship for the

    first, leading to the following equation:

    ln k2 - ln k1 = ln(k2/k1) =

    ln = (1/RT )(H02 H01) + (1/R)

    (S02 S01) [7]

    which can also be expressed as

    l n = (1/RT ) H 02,1 + (1/R)

    S02,1 [8]

    If the individual van t Hoff relation-

    ships for the first and second analytes

    are linear, and the van t Hoff relation-

    ship in equation 7 or 8 for their selectiv-

    ity is also linear, then there is a higher

    probability that H0, S0, and are

    constant (or at least not substantially

    changing) over the temperature range

    in question. Thus, if one wishes to use a

    van t Hoff analysis to extract H0 and

    S0 values for an analyte, it would prob-

    ably be advisable to apply this additional

    check on the data. If the resulting plot

    of equation 7 or 8 is linear, it would add

    credence to any claim that meaningful

    thermodynamic information could be

    extracted from the analysis. In addition,

    if the plot of ln vs. 1/T is nonlinear

    then the stationary phase may undergo a

    conformational change in the tempera-

    ture range studied (3841).

    As a corollary to these last points,

    using b1 = S01/R ln for the first

    analyte and b2 = S02/R ln for the

    second analyte under the same condi-

    tions or same column, we can calculate

    the difference in entropies of transfer

    for the two analytes as S02,1 = S02

    S01 = R(b2 b1), where this latter term

    is the gas constant multiplied by the dif-

    ference between the y-intercepts of the

    two van t Hoff plots for the two ana-

    lytes. Note that the phase ratio, which

    we often do not know, has canceled,

    leaving us with the difference between

    the two entropies of transfer. This anal-

    ysis can be useful on a series of com-

    pounds, where they are all compared to

    one member of the series.

    T (C) T (C)

    1/T (K-1) 1/T (K-1)

    ln k

    ln k

    75

    (a) (b)

    65 55 45 35 25 75 65 55 45 35 251.7 0.75

    0.65

    0.55

    0.45

    0.35

    0.25

    1.6

    1.5

    1.4

    1.3

    1.2

    1.1

    10.0029 0.0030 0.0031 0.0032 0.0033 0.0029 0.0030 0.0031 0.0032 0.0033

    T (C) T (C)

    1/T (K-1) 1/T (K-1)

    ln k

    ln k

    55

    (a) (b)

    2

    1.5

    0.5

    -0.5

    -1.5

    -1

    0

    1

    2

    1.5

    0.5

    -0.5

    -1.5

    -1

    0

    1

    50 45 40 35 30 25 20 55 50 45 40 35 30 25 20

    0.00305 0.00315 0.00325 0.003250.00335 0.003350.00305 0.00305

    Figure 4: Van t Hoff plots of protriptyline obtained at pH 7.8. Column: (a) Inertsil ODS 3V, (b) X-Terra RP18. Flow rate: 1.0 mL/min. Temperature increases from right to left. Adapted from reference 33 with permission.

    Figure 5: Van t Hoff plots of acidic and basics analytes. (a) Phosphate buffer pH(25 C) = 8.10, and (b) tris + HCl buffer pH (25 C) = 8.09; mobile phase contains 50% (v/v) methanol. Analytes: = 2,4-dichlorophenol, = 2,6-dichlorophenol, = ben-zylamine, = benzyldimethylamine. Adapted from reference 29 with permission.

    ES160067_LCGC1212_1056.pgs 11.28.2012 23:01 ADV blackyellowmagentacyan

  • DECEMBER 2012 LCGC NORTH AMERICA VOLUME 30 NUMBER 12 1057www.chromatographyonline.com

    Concerns About the

    van t Hoff Equation

    A careful study of the van t Hoff equa-tion and its use in elevated-temperature LC suggests that there is some question regarding its fundamental accuracy (36,37). For example, the van t Hoff relationship assumes that the stationary phase is homogeneous, which is clearly not the case in general, a stationary phase will contain different types of sites, which will have different affini-ties for a given analyte. The van t Hoff equation also assumes that both the stationary phase and the mobile phase remain constant as a function of tem-perature. Neither will be entirely true. The adsorption and absorption (parti-tioning) of mobile-phase components in the stationary phase, which will alter the properties of the stationary phase, will vary with temperature, and the mobile phase will also change with temperature; for example, the static per-mittivity (dielectric constant) of water will change with temperature. Perhaps a measured view of these concerns is to acknowledge their validity, while also noting that in many circumstances it appears that these effects are not so extreme that useful information cannot be obtained by van t Hoff analysis.

    Conclusions

    Van t Hoff plots can be a useful and interesting part of data analysis for high temperature LC. They are a valuable tool in an empirical sense for retention mod-eling, and thermodynamic data may be extracted from them. They can reveal phase transitions in stationary phases, and the changes in pKa values of analytes with temperature. Plots of ln vs. 1/T can help confirm that H0 and S0 are constant with temperature. It should be understood that the underlying assump-tions of the van t Hoff equation are not entirely correct.

    References

    (1) D.S. Jensen, T. Teutenberg, J. Clark, and

    M.R. Linford, LCGC North Am. 30(9),

    850862 (2012).

    (2) D.S. Jensen, T. Teutenberg, J. Clark, and

    M.R. Linford, LCGC North Am. 30(11),

    992998 (2012).

    (3) J.W. Dolan, J. Chromatogr., A 965(12),

    195205 (2002).

    (4) S. Wiese, T. Teutenberg, and T.C. Schmidt, J.

    Chromatogr., A 1218(39), 68986906 (2011).

    (5) J.W. Dolan, L.R. Snyder, N.M. Djordjevic,

    D.W. Hill, D.L. Saunders, L. Van Heukelem,

    and T.J. Waeghe, J. Chromatogr., A 803(12),

    131 (1998).

    (6) R.G. Wolcott, J.W. Dolan, and L.R. Snyder,

    J. Chromatogr., A 869(12), 325 (2000).

    (7) G. Vanhoenacker, F. David, and P. Sandra,

    Chromatogr. Today 3(3), 1416 (2010).

    (8) Y. Xiang, B. Yan, B. Yue, C.V. McNeff,

    P.W. Carr, and M.L. Lee, J. Chromatogr., A

    983(12), 83-89 (2003).

    (9) G. Vanhoenacker and P. Sandra, Anal. Bio-

    anal. Chem. 390(1), 245248 (2008).

    (10) A.M. Edge, S. Shillingford, C. Smith, R.

    Payne, and I.D. Wilson, J. Chromatogr., A

    1132(12), 206210 (2006).

    (11) C.M. Hong and C. Horvth, J. Chromatogr.,

    A 788(12), 5161 (1997).

    (12) J.V. Tran, P. Molander, T. Greibrokk, and

    E. Lundanes, J. Sep. Sci. 24(12), 930940

    (2001).

    (13) S. Heinisch and J.-L. Rocca, J. Chromatogr.,

    A 1216(4), 642658 (2009).

    (14) J. Bowermast and H.M. McNair, J. Chro-

    matogr. Sci. 22, 165170 (1984).

    (15) L.R. Snyder, J.J. Kirkland, and J.W. Dolan,

    Introduction to Modern Liquid Chromatog-

    raphy (Wiley, Hoboken, New Jersey, 2010).

    (16) P.L. Zhu, L.R. Snyder, J.W. Dolan, N.M.

    Djordjevic, D.W. Hill, L.C. Sander, and T.J.

    Waeghe, J. Chromatogr., A 756(12), 2139

    (1996).

    (17) P.L. Zhu, J.W. Dolan, and L.R. Snyder, J.

    Chromatogr., A 756(12), 4150 (1996).

    (18) R.G. Wolcott, J.W. Dolan, L.R. Snyder, S.R.

    Bakalyar, M.A. Arnold, and J.A. Nichols, J.

    Chromatogr., A 869(12), 211230 (2000).

    (19) Y. Liu, N. Grinberg, K.C. Thompson, R.M.

    Wenslow, U.D. Neue, D. Morrison, T.H.

    Walter, J E. OGara, and K.D. Wyndham,

    Anal. Chim. Acta 554(12), 144151 (2005).

    (20) J.F. Wheeler, T.L. Beck, S.J. Klatte, L.A.

    Cole, and J.G. Dorsey, J. Chromatogr., A

    656(12), 317333 (1993).

    (21) K.B. Sentell and A.N. Henderson, Anal.

    Chim. Acta 246(1), 139149 (1991).

    (22) L.A. Cole and J.G. Dorsey, Anal. Chem.

    64(13), 13171323 (1992).

    (23) L.A. Cole, J.G. Dorsey, and K.A. Dill, Anal.

    Chem. 64(13), 13241327 (1992).

    (24) D. Morel and J. Serpinet, J. Chromatogr., A

    214(2), 202208 (1981).

    (25) J.W. Coym and J.G. Dorsey, J. Chromatogr.,

    A 1035(1), 2329 (2004).

    (26) T. Teutenberg, Anal. Chim. Acta 643(12),

    112 (2009).

    (27) B. Hemmateenejad, J. Chemom. 19(1112),

    657667 (2005).

    (28) S. Pous-Torres, J.R. Torres-Lapasi, J.J.

    Baeza-Baeza, and M.C. Garca-lvarez-

    Coque, J. Chromatogr., A 1163(12), 4962

    (2007).

    (29) C.B. Castells, L.G. Gagliardi, C. Rfols,

    M. Ross, and E. Bosch, J. Chromatogr., A

    1042(12), 2336 (2004).

    (30) L.G. Gagliardi, C.B. Castells, C. Rfols,

    M. Ross, and E. Bosch, J. Chromatogr., A

    1077(2), 159169 (2005).

    (31) R.C. Chloupek, W.S. Hancock, B.A. Mar-

    chylo, J.J. Kirkland, B.E. Boyes, and L.R. Sny-

    der, J. Chromatogr., A 686(1), 4559 (1994).

    (32) L.G. Gagliardi, C.B. Castells, C. Rfols, M.

    Ross, and E. Bosch, Anal. Chem. 78(16),

    58585867 (2006).

    (33) S.M.C. Buckenmaier, D.V. McCalley, and

    M.R. Euerby, J. Chromatogr., A 1060(12),

    117126 (2004).

    (34) S.M.C. Buckenmaier, D.V. McCalley, and

    M R. Euerby, J. Chromatogr., A 1026(12),

    251259 (2004).

    (35) T.L. Chester and J.W. Coym, J. Chromatogr.

    A 1003(12), 101111 (2003).

    (36) F. Gritti and G. Guiochon, J. Chromatogr., A

    1099(12), 142 (2005).

    (37) F. Gritti and G. Guiochon, Anal. Chem.

    78(13), 46424653 (2006).

    (38) F. Wang, T. OBrien, T. Dowling, G. Bicker,

    and J. Wyvratt, J. Chromatogr., A 958(12),

    6977 (2002).

    (39) T. OBrien, L. Crocker, R. Thompson, K.

    Thompson, P.H. Toma, D.A. Conlon, B.

    Feibush, C. Moeder, G. Bicker, and N. Grin-

    berg, Anal. Chem. 69(11), 19992007 (1997).

    (40) E. Papadopoulou-Mourkidou, Anal. Chem.

    61(10), 11491151 (1989).

    (41) R.K. Gilpin, S.E. Ehtesham, and R.B. Greg-

    ory, Anal. Chem. 63(24), 28252828 (1991).

    David S. Jensen and Matthew R. Lin-ford are with the Department of Chem-istry and Biochemistry at Brigham Young University in Provo, Utah. Direct correspon-dence to: [email protected].

    Thorsten Teutenberg is with the Institut fr Energie- und Umwelttechnik e. V. in Duisburg, Germany.

    Jody Clark is with Selerity Technologies Inc., in Salt Lake City, Utah.

    For more information on this topic,

    please visit

    www.chromatographyonline.com/Linford

    ES160070_LCGC1212_1057.pgs 11.28.2012 23:01 ADV blackyellowmagentacyan

  • Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.