Chem Educator

download Chem Educator

of 4

Transcript of Chem Educator

  • 7/31/2019 Chem Educator

    1/4

    Chem. Educator2007,12, 327330 327

    Variation of the Critical Micelle Concentration with Surfactant Structure:

    A Simple Method To Analyze the Role of AttractiveRepulsive Forces on

    Micellar Association

    D. Lpez-Daz and M. M. Velzquez*

    Departamento de Qumica Fsica, Facultad de Ciencias Qumicas. Universidad de Salamanca, 37008-Salamanca, Spain, [email protected]

    Received September 3, 2006. Accepted July 1, 2007.

    Abstract: In this laboratory experiment students analyze the role of attractive interactions between thehydrocarbon tails and repulsive interactions between the surfactant head groups on the micellar association

    process by determining the CMC of two surfactant classes. Using electrical conductivity measurements and

    pyrene fluorescence emission, the CMCs of two homologous series of sulfobetaines, zwitterionic surfactants, and

    alkyl trimethyl ammonium bromide, cationic surfactants, were determined. In each surfactant family we use

    surfactants with 12, 14, and 16 carbon atoms. From the correlation of the head group charge to the methylene

    group contribution to CMC for the two surfactant classes, students can clearly analyze the influence of the

    electrostatic repulsions on the self-assembly processes.

    Traditionally undergraduate physical chemistry courses

    briefly discuss surface and interfacial phenomena; however,

    intensive research of recent years in the structure and

    properties of microheterogeneous systems, in molecular

    recognition at the membrane interface, and in the transport

    across membrane shows the importance of the interfacial

    phenomena. Consequently, colloid and surface chemistry must

    be incorporated to the chemistry and biochemistry curriculum.

    Experiments with colloids and surface chemistry can be

    incorporated in physical chemistry laboratory. They can also

    be designed as a complement to the concepts studied in the

    traditional physical chemistry experiments. Experimentsfocusing on micellar association introduce students to the

    concept of colloid stability and forces involved in molecular

    aggregation. In addition, the formation of micellar aggregates

    causes significant changes on a larger number of physical

    properties, such as conductivity, molecular fluorescence, and

    surface tension; therefore, with experiments focusing on the

    determination of the concentration at which a surfactant forms

    a micelle, called the critical micelle concentration or CMC,

    students also learn about electrochemical or spectroscopic

    techniques. Several experiments to determinate the CMC have

    been published using several of these methods [15].

    Micelles are formed and stabilized by a balance of forces;

    the insolubility of the alkyl tail promotes aggregation

    (hydrophobic forces), and the electrostatic repulsions of the

    ionic head groups inhibit aggregation. The effect of a small

    change in these forces can be seen in the experimental data as

    changes in the CMC values; therefore, it is possible to analyze

    the role of these forces on the micellar aggregation process

    studying the effect of both, the electrical charge of the

    surfactant head group and the hydrocarbon chain length on the

    CMC. With this objective in mind we have designed one

    laboratory experiment focusing to study the role of the

    different forces responsible of micelle formation by analyzing

    the CMC values of two families of surfactants of different head

    groups: alkyl trimethyl ammonium bromide, cationic

    surfactants and alkyl dimethyl ammonium propane sulfonate,

    zwitterionic surfactants. The hydrocarbon chain length varies

    between 12 and 16 carbon atoms.

    Students can use electrical conductivity measurements to

    obtain the CMC and the ionization degree of cationic micelles.

    This methodology is widely used to characterize ionic

    surfactants [3, 5]; however, it cannot be used for zwitterionic

    surfactants because they do not conduce the electrical current.

    In this case, the CMC is determined by measuring the change

    in the fluorescence emission spectrum of pyrene monomers [4

    6]. This method is based on the changes on the intensity of the

    vibrational bands of pyrene emission caused by changes on thepolarity in the environment of the probe [7].

    Experimental

    The surfactants dodecyldimethylammonium propane sulfonate,

    DDPS; tetradecyldimethylammonium propane sulfonate, TDPS;

    hexadecyldimethyl ammonium propane sulfonate, HDPS;

    dodecyltrimethylammonium bromide, DTAB; and

    tetradecyltrimethylammonium bromide, TTAB, and the fluorescence

    probe, pyrene, were from Sigma-Aldrich. Methanol and the surfactant

    hexadecyltrimethylammonium bromide, CTAB, were from Merck.

    Conductivity Measurements. Prepare 50 mL of the each cationic

    surfactant in deionized water. A minimum of twenty surfactant

    solutions, ten above and ten below the CMC of each surfactant are

    necessary to the correct determination of CMC and ionization degree.These solutions are placed on a constant temperature bath at least 20

    minutes before measurements.

    The electrical conductivity was measured with a conductometer,

    model 727 from Metrohm, operated at 2.4 kHz. A Metrohm Herisau

    conductivity cell, model AG 9101, was used. The cell constant, 0.847

    cm1

    , was obtained by calibration with potassium chloride standards

    (0.0100 and 0.0050 M).

    Fluorescence Measurements. The solubilization of pyrene in

    micelles was carried out as follows: 5 L of a solution of 0.002 M

    pyrene dissolved in methanol, solution A, was placed into a 10-mL

    volumetric flask and the solvent was evaporated till dryness by slow

    passage of N2. The surfactant solution of each surfactant

    concentration, solution B, was added to the evaporated residue and the

    2007 The Chemical Educator, S1430-4171(07) 52075-X, Published on Web 9/25/2007, 10.1333/s00897072075a, 12070327mv.pdf

  • 7/31/2019 Chem Educator

    2/4

    328 Chem. Educator, Vol. 12, No. 5, 2007 Velzquez et al.

    Figure 1. Variation of the electrical conductivity with concentration

    for cationic surfactants.

    resulting solution was stirred until pyrene was solubilized. Thus, in all

    the surfactant solutions, pyrene concentration was kept at 1 M. The

    pyrene concentration has to be less than 5mM to avoid excimer

    formation. In general, the excimer molecules are formed between

    electronically excited and other ground state molecules. Pyrene shows

    a characteristic excimer emission at 480 nm [8].

    Solution A: Prepare 25 mL of a 0.002 M of pyrene in methanol(~10 mg of pyrene). Solution B: Prepare 25 mL of each zwitterionic

    surfactant in deionized water.

    The surfactant concentration range was between 8 104

    M and 6.4

    103

    M for DDPS; 1.4 105

    M and 7 104

    M for TDPS and 3

    106

    M and 1.5 104

    M for HDPS. A minimum of twelve surfactant

    solutions, below and above the CMC, are necessaries to calculate the

    CMC values.

    The experimental conditions to obtain the emission spectrum of

    pyrene were the following: the excitation and emission slits used gave

    a bandwidth of 2.5 nm and the excitation wavelength was 320 nm.

    The wavelength emission range was between 350 and 440 nm. The

    fluorescence spectra of pyrene were recorded in a Perkin Elmer

    spectrofluorometer model LS-50B.

    Results and Discussion

    CMC and Ionization Degree Determination for Cationic

    Micelles: Electrical Conductivity Measurements. Figure 1

    shows the variation of the electrical conductivity with the

    cationic surfactant, DTAB, TTAB, and CTAB, concentration.

    The CMC is obtained from the interception of conductivity

    lines above and below the CMC.

    It is well accepted that below the CMC there are no micelles

    in solutions; thus, the conductivity of an aqueous ionic

    surfactant, SC, where S represents the surfactant ion and C the

    corresponding counter-ion, is due to the independent

    contribution of these ions. If the aqueous surfactant solutions

    obey the Kohlrauschs law [9], the conductivity can be written

    as:

    = S(c + s) (1)

    where c ands are themolar ionic conductivity of the counter-

    ion and the surfactant, respectively; and S is the surfactant

    concentration. Equation 1 explains the linear dependence

    between conductivity and S below the CMC. The slope of this

    line, s1, represents (c + s). Above the CMC, further additionof surfactant results in an increase in micelle concentration

    while the monomer concentration remains constant in a value

    close to the CMC. The ionic mobility of the micelle is very

    different to that of the monomer molecule and, even though the

    conductivity linearly increases with surfactant concentration,

    the slope of this line is smaller than s1. The conductivity of

    surfactant solutions at concentrations above the CMC is from

    three different contributions: the independent ions S and C at

    the CMC, the micelle conductivity and the counterions

    unbonded in the micelle. Thus, the conductivity is given by:

    [ ] ( )C S C( )CMC micelles CMCmic S = + + + (2)

    Taking into account that [micelles] = (S CMC)/N, whereN

    is the micelle aggregation number, and assuming that the

    micelle conductivity is the same that the conductivity of all

    monomers with electrical charge in the micellar aggregate, that

    is, mic = SN, eq 2 can be rearranged:

    ( ) ( ) ( )C S C S o 2CMC 1 ( ) ( )S s = + + + = + S (3)

    where s2 is the slope of the linear plot ofversus S above the

    CMC. Consequently, the s2/s1 ratio represents the micelle

    ionization degree, , [10], which is the fraction of surfactant

    molecules in the micellar aggregate that do not have bound

    counter-ions.Because the error of the slope and ordinate of these lines are

    smaller than the conductivity uncertainty, one can consider the

    error on the conductivity measurements (1.3%) as responsible

    of the error in both, the CMC and values.

    The values obtained in the work are collected in Table 1. As

    can be seen the CMC values agree very well with values on

    literature also presented in a table

    The CMC values found in this work are in excellent

    agreement with data in the literature [11]. The ionization

    degree for cationic surfactants decreases from 0.26 to 0.24 in

    going from hexadecyl trimethyl ammonium to dodecyl

    ammonium bromide. This behavior was reported elsewhere for

    2007 The Chemical Educator, S1430-4171(07) 52075-X, Published on Web 9/25/2007, 10.1333/s00897072075a, 12070327mv.pdf

  • 7/31/2019 Chem Educator

    3/4

    Variation of the Critical Micelle Concentration with Surfactant Structure Chem. Educator, Vol. 12, No. 5, 2007 329

    Table 1. CMC and Values Found in This Laboratory Experiment

    for Cationic Surfactants

    Surfactant 103 CMC

    M

    103CMCbib

    M

    a

    DTAB 15.6 0.2 0.261 0.003 16 -

    TTAB 3.76 0.05 0.252 0.003 3.5 0.27

    CTAB 0.924 0.001 0.245 0.003 0.92 0.24

    aFrom reference 12

    360 380 400 420 4400

    100

    200

    300

    400

    500

    5

    4

    32

    1

    I/a.u

    / nm

    Figure 2. Fluorescence spectra of pyrene dissolved in aqueous TDPSsolutions of different concentrations: (--) 7.0 10

    5M, () 1.4 10

    4

    M, (--) 2.1 104

    M, (__

    ) 2.8 104

    M.

    0 1 2 3 4 5 6 71.3

    1.4

    1.5

    1.6

    1.7

    CMC

    I1/I

    3

    103

    [DDPS] / M

    0 1 2 3 4 5 6 7 81.3

    1.4

    1.5

    1.6

    1.7

    1.8

    I1/I3

    104

    [TDPS] / M

    CMC

    0.0 0.4 0.8 1.2 1.6

    1.3

    1.4

    1.5

    1.6

    1.7

    1.8

    CMCI1/I

    3

    104

    [HDPS] /M

    Figure 3. Variation of the I1/I3 ratio with the zwitterionic surfactant

    concentration.

    alkyl sulfates or alkylcarboxilates and it was related with

    changes of surface areas per head group [13]. Thus, in the case

    of surfactants with great surface areas, shorter hydrocarbon

    tails, geometric constraints prevent the approach of the

    counterion to the head group increasing the ionization degree.

    Determination of CMC Values for Sulfobetaines Micelles

    by Using Pyrene Fluorescence Probing. Because zwitterionicsurfactants do not conduct electrical current, the CMCs of

    these compounds have to be determinate by an alternativemethod. A widely used methodology is pyrene fluorescence

    probing. This method is based on changes in the intensity of

    the vibrational bands of pyrene solubilized in water and in

    micellar medium.

    Figure 2 presents the pyrene emission spectra of solubilized

    aqueous surfactant solutions containing surfactant

    concentrations below and above the CMC.

    Figure 2 clearly shows that the vibrational structure of

    fluorescence spectra depends on the surfactant concentration.

    This is because the fluorescence of pyrene at low

    concentrations in homogeneous solutions possesses fine

    structure whose relative peak intensity undergoes significant

    perturbation upon going from polar to nonpolar solvents. The

    ratio of the first vibrational band (372 nm), the highest energyvibrational band, to the fluorescence intensity of the third

    vibrational band (385 nm) has been shown to correlate with

    solvent polarity [7]. For example in hydrocarbon solventI1/I3 =

    0.6 and in water is around 1.6.

    It is well established that in surfactant solutions the plot of

    I1/I3 versus surfactant concentration shows a typical sigmoid

    shape. Figure 3 shows the results found for sulfobetaine

    zwitterionic surfactants.

    Below the CMC the I1/I3 ratio corresponds to a polar

    microenvironment; when the surfactant concentration increases

    the ratio decreases rapidly as a consequence of the more

    hydrophobic environment of pyrene. Above the CMC theI1/I3

    ratio reaches a constant value due to the incorporation of

    pyrene into the hydrophobic region of the micelle [6]. TheCMC is obtained from the interception of the horizontal and

    the steep parts of the curve. The CMC values found in this

    work are 3.3 103

    M for DDPS, 3.1 104

    M and 3.5 105

    M for TDPS and HDPS, respectively. These values agree with

    data in the literature [14].

    It is well known that the CMC has a strong dependence on

    the alkyl chain length of a surfactant, Nc. This dependence can

    be described by eq 3, Klevens rule [15, 16].

    Log CMC =A BNc (3)

    In this equation, A depends on the surfactant head group,

    temperature, and the addition of inert electrolytes and B

    represents the contribution of each methylene group in the

    lowering of the CMC by the tail. Figure 4 shows the CMC

    values represented according to Klevens equation. Results are

    in excellent agreement with eq 3.

    The A and B values were calculated to be: sulfobetaines,A =

    3.3 0.3 and B = 0.48 0.02, and for alkyl

    trimethylammonium bromide surfactantsA = 1.77 0.02 andB

    = 0.299 0.001. TheB values are in very good agreement with

    data in the literature [17]. For the zwitterionic surfactants theB

    value is higher than that for the cationic ones. This fact

    indicates that in the balance of forces present during micelle

    aggregation, the ability of the alkyl chain to lower CMC

    2007 The Chemical Educator, S1430-4171(07) 52075-X, Published on Web 9/25/2007, 10.1333/s00897072075a, 12070327mv.pdf

  • 7/31/2019 Chem Educator

    4/4

    330 Chem. Educator, Vol. 12, No. 5, 2007 Velzquez et al.

    11 12 13 14 15 16 17-5.0

    -4.5

    -4.0

    -3.5

    -3.0

    -2.5

    -2.0

    -1.5

    logCMC

    Nc

    Figure 4. Dependence of the CMC on the alkyl tail for: (circles) alkyl

    trimethyl ammonium bromide; (squares) sulfobetaines.

    depends on the magnitude of the charge on the head group. A

    larger value ofB indicates that each additional methylene has a

    great effect on lowering the CMC. Thus, for the zwitterionic

    surfactants, no head group charge, B is large, while, for the

    cationic surfactants,B decreases due to the repulsions between

    the positive charges of the head group. Consequently the CMC

    values of zwitterionic surfactants are usually smaller than theCMCs of ionic ones.

    From the B values students can obtain the charge on the

    head group by using the following linear correlation betweenB

    and the square of the charge of the head group, q2, [16]:

    B = (0.499 0.007) (0.234 0.011)q2

    (4)

    From results obtained in this work we found a q2

    of 0.12 and

    0.85 for sulfobetaines and alkyltrimethyl ammonium

    surfactants, respectively. These values are in good agreement

    with the charge distribution estimated using semi-empirical

    quantum chemical methods [16]. It is interesting to note that

    the charge of the head group of the cationic surfactants, q =

    0.92 is not exactly 1. This fact was predicted by semi-empiricalquantum chemical methods. Results obtained from these

    methods show that the charge of the head group in ionic

    surfactants is partially distributed to the rest of the molecule,

    with significant charge on the -methylene group and a partial

    charge on the remaining alkyl tail. Thus, for dodecyl trimethyl

    ammonium surfactants, the charge of the combined head group

    and the -methylene group obtained by semi-empirical

    methods is around 0.89 and the partial charge on the surfactant

    tail is 0.11 [16]. If one compares this value with that obtained

    in this laboratory experiment, 0.92, it can be concluded that the

    -methylene group is part of the head group of surfactants as

    several studies have suggested [16].

    Conclusion

    In the design of this laboratory experiment we had several

    goals in mind. First, we wanted to introduce colloidal

    chemistry into the physical chemistry laboratory curriculum. In

    addition, we wanted to introduce students to the concepts of

    molecular photochemistry and electrochemical measurements.

    This laboratory experiment is intended for the physical

    chemistry laboratory curriculum. There are several ways to

    organize the experiments depending on the time and the

    equipment available. Assuming that the detailed experimental

    procedure is provided in advance, students should be able to

    carry out this experiment individually in two four-hour

    laboratory periods. This requires groups of at least fivestudents. Each group can determine the CMC of two

    surfactants using conductivity and fluorescence measurements,

    for instance, dodecyl dimethyl ammonium propane sulfonate

    (fluorescence) and dodecyl trimethyl ammonium bromide

    (conductivity). Finally, the students analyze all results.

    Acknowledgment. This work was financially supported bythe Ministerio de Ciencia y Tecnologa (BQU 2001-1507) and

    the Ministerio de Educacin y Ciencia (MAT 2004-04180). D.

    Lopez wishes to thank Ministerio de Educacin y Ciencia of

    Spain for the grant AP2002-1734.

    References and Notes

    1. Furton, K. G.; Norelus, A.J. Chem. Educ. 1993, 70, 254257.

    2. Goooling, K.; Johson, K.; Lefkowictz, L.; Williams, B. W.J. Chem.

    Educ. 1994, 71, A8A12.

    3. Bachofer, S. J.J. Chem. Educ. 1996, 73, 861864.

    4. Stam, J.; Depaemelaere, S.; De Schryver, F. C.J. Chem. Educ. 1998,

    75, 9398.

    5. Domnguez, A.; Fernndez, A.; Gonzlez, N.; Iglesias, E.;

    Montenegro, L.;J. Chem. Educ. 1997, 74, 12271231.

    6. Kalyanasundaram, K. Photochemistry. in Microheterogeneous

    Systems; Academic Press: Orlando FL, 1987.

    7. Kalyanasundaram, K.; Thomas, J., K., J. Am. Chem. Soc. 1977, 99,

    20392044.

    8. Turro, N, J. Modern Molecular Photochemistry; Benjamin

    Cummings: San Francisco CA, 1978, pp 137143.9. See for example Atkins P. W. Physical Chemistry, 4th ed.; Oxford

    University Press: Oxford, 1990; p 750.

    10. Weers, J. G.; Rathman, J. F.; Axe, F. U.; Crichlow, C. A.; Foland, L.

    D.; Scheuing, D. R.; Wiersema, R. J.; Zielske, A. G.Langmuir, 1991,

    7, 854867.

    11. Rosen, M. J. Surfactants and Interfacial Phenomena, Wiley: New

    York, 1978; p 97.

    12. Sepulveda L., Corts, J. J. Phys. Chem.1985, 89, 53225324.

    13. Zana, R. J. Colloid Interface Sci. 1980, 78, 330337

    14. Evans, H. C.J. Chem. Soc. 1956, 579586.

    15. Klevens, H. B.J. Am. Oil Chem. Soc. 1953, 30, 7479

    16. Huibers, P. D. T.Langmuir1999, 15, 75467550.

    17. Rosen, M. J. Surfactants and Interfacial Phenomena, 2nd ed.; Wiley:New York, 1989.

    2007 The Chemical Educator, S1430-4171(07) 52075-X, Published on Web 9/25/2007, 10.1333/s00897072075a, 12070327mv.pdf