You are on page 1of 10

552 J. Phys. Chem.

B 2001, 105, 552-561

Intermolecular Interactions and the Structure of Fatty Acid-Soap Crystals

Matthew L. Lynch,*,† Fred Wireko,† Mounir Tarek,‡,§ and Michael Klein‡


The Procter & Gamble Company, Corporate Research DiVision, Miami Valley Laboratories, Ross, Ohio 45061,
Center for Neutron Research, National Institute of Standards and Technology, Gaithersburg, Maryland 20899,
and Center for Molecular Modeling, Department of Chemistry, UniVersity of PennsylVania,
Philadelphia, PennsylVania 19104-6323
ReceiVed: July 22, 2000; In Final Form: October 30, 2000

Single crystals of NaHP2 (sodium hydrogen dipalmitate) have been prepared from mixtures of NaP (sodium
palmitate) and HP (palmitic acid) in ethanol. The phase compound crystallizes in the P21/a space group, with
a ) 9.906 Å, b ) 7.163 Å, c ) 45.580 Å, β ) 92.78°, and 4 molecules per unit cell. The arrangement of the
headgroups is unique among known soap and fatty acid structures by accommodating both hydrogen bonding
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

and electrostatic interactions. Carboxylate and acid-like pairs couple perpendicular to the bilayer to
accommodate a “short” hydrogen bond, and the sodium is shared among laterally adjacent carboxylate anions
to create a pseudo-six-member ring, which adds to the crystal stability. Molecular dynamic (MD) simulations
Downloaded via COLGATE-PALMOLIVE CO on June 22, 2021 at 22:38:50 (UTC).

establish a set of interaction parameters to describe the structure and energetics of acid-carboxylate bonds,
which accounts for the solid-state miscibility of HP and NaP. Infrared data, collected by ATR-FTIR on very
pure, powdered NaH2P3, Na2H3P5, and NaHP2 acid-soap standards reveal peculiar changes in the carbonyl
and hydroxyl spectral regions. These spectra were analyzed on the basis of crystal structure and MD simulation
data.

1. Introduction Acid-soaps have both commercial and academic importance.


Chevreul is widely recognized as having made the first These compounds are ubiquitous in consumer products. They
reference to acid-soaps in 1823.1 However, it was not until are found in superfatted, transparent, and traditional soap bars,
the work of Ekwall2-4 and McBain5-8 that the existence of these cleaning products, cosmetics, facial cleaners, shaving creams,
compounds was proved by systematic experimental methods. deodorants, and topical-delivered products.25-26 Sodium car-
A summary of subsequent studies is contained in a recent boxylates can be used to form rheological solids with less than
review.9 The miscibility of fatty acid and sodium carboxylate 2% organic material27 and can soften with the incorporation of
in the solid state is postulated to be a consequence of the strong acid-soaps as a consequence of the inherent differences in
hydrogen bond between the fatty acid headgroup and the crystal habit.28 The study of fatty acid-carboxylate bonding is
carboxylate anion. While the fatty acid is the hydrogen-bond also of academic importance. For example, aqueous sodium and
donor, the carboxylate anion is likely a stronger hydrogen-bond potassium hydrogen oleate systems have been employed as
acceptor than the acid. However, the nature of these interactions potential model membranes.29-31 Further, they are excellent
remains elusive as no crystal structures of acid-soaps with long systems for studying the spectroscopy associated with carboxylic
alkyl chains have been reported. acid-carboxylate systems, as the stretching frequencies and
Clues about the structure of acid-soaps have come from peak intensities are dramatically different from those of the
extrapolation to single-crystal data collected for fatty acids,10-14 parent compounds. These findings can possibly be transferred
short-chain acid-soap crystals such as sodium hydrogen diac- to the study of more complex peptide and protein structures
etate,15 and potassium hydrogen phenylacetate16 and to a survey involving, for example, bonding to aspartate and glutamate
of existing single-crystal data of similar compounds.17 Structural residues. Finally, it is likely that acid-soaps will continue to
details are crucial to understanding the “anomalous” infrared18-22 be of interest, as similar hydrogen bonds are a main design
and NMR20,23,24 data on powder acid-soap crystals, which show element in both crystal engineering and supramolecular
that acid-soap crystals maintain some carboxylic acid-like assembly.32-34
character and some carboxylate anion-like character, which are
quantitatively different from the character of either the parent 2. Experimental Section
fatty acid or the sodium carboxylate. The present combined
crystallographic and molecular dynamics (MD) simulation study Preparation of Single-Crystal NaHP2. Single crystals of
has been undertaken to provide the first structural description NaHP2 were prepared from mixtures of NaP and HP in ethanol.
of these long chain length acid-soaps that allows for detailed The phase behavior of HP-NaP-ethanol mixtures was first
understanding of the spectroscopy and solid-state chemistry of established to ensure the formation of NaP2 crystals. The optimal
these compounds. situation entailed mixing 0.45 g of HP and 0.05 g of NaP in 50
mL of ethanol (HPLC grade) in a shallow recrystallization dish.
* Corresponding author. Tel.: (513) 627-0392. Fax: (513) 627-1233. The mixture was gently heated to speed the dissolution of NaP
E-mail: lynch.ml@pg.com. and HP. The dish was fitted with a plastic cover, which
† The Procter & Gamble Company.
‡ University of Pennsylvania. contained a small 1 mm diameter hole in the center. Crystals
§ National Institute of Standards and Technology. began to form after the first few hours. The sample was allowed
10.1021/jp002602a CCC: $20.00 © 2001 American Chemical Society
Published on Web 12/20/2000
Structure of Fatty Acid-Soap Crystals J. Phys. Chem. B, Vol. 105, No. 2, 2001 553

Figure 1. Powder XRD patterns of crystal standards: (A) palmitic acid, (B) sodium palmitate, (C) NaH2P3, and (D) Na2H3P5.

to sit for 2-3 weeks. Remaining ethanol was removed by An isothermal mixing process was developed to generate
filtration. Crystals of optimal size (about 1 mm) were culled high-quality Na2H3P5 and NaHP2 standards. Unlike the prepara-
by looking for sharp extinction in birefringence upon rotation tion of NaH2P3, the preparation of these standards by the
of the crystals in a polarized light microscope. aforementioned heating/cooling method resulted in mixtures of
Preparation of Polycrystalline Acid, Soap, and Acid-Soap crystals. Instead, a 1 g total of NaP and NaH2P3 crystals (ratios
Standards. Palmitic acid was purchased from NuCheck Prep, defined by the stoichiometry needed to form the acid-soaps)
Inc. GC analysis of the methyl ester derivative (prepared using was added to 50 mL of deionized water in a small beaker to
methanesulfonic acid in methanol) indicated the percentage create a slurry (using the NaH2P3 crystals was tantamount to
composition, as determined by peak area, to be 0.03 C7, 0.03 providing solubility of fatty acid in water). A Teflon stir bar
C12, 0.16 C14, 0.08 C15, 99.52 C16, 0.07 C17, and 0.11 C18. The was added to the beaker, and the slurry was stirred for 10-12
acid was melted and cooled to generate form-C crystals. To h. Large chucks of NaH2P3 and NaP crystals evident at the start
prepare sodium palmitate, a 10% w/w palmitic acid-deionized gave way to a fine dispersion of acid-soap crystals. The crystals
water solution was heated to 85 °C and neutralized with 0.1 M were removed by pouring the slurry through filter paper and
sodium hydroxide (standardized with potassium acid phthalate). allowed to air-dry.
The solution was cooled. A Buchner funnel was used to filter The purity of each acid-soap standard was confirmed with
the resulting crystals, which were allowed to air-dry to constant powder X-ray diffraction (XRD) methods by inspecting the
weight (about 1 week). No further effort was undertaken to number of sets of [00L] reflections. These are intense reflections,
remove the water of hydration. which correspond to the bilayer distances. Each acid-soap has
NaH2P3 crystals were created by partial neutralization of a a single set of these reflections based on their unique bilayer d
10% w/w palmitic acid-deionized water solution, as previously spacings (45.6 Å for NaHP2, 42.8 Å for NaH2P3, and 42.2 Å
noted.20 The mixture was heated to 85 °C and partially for Na2H3P3), as predicted by Bragg’s Law, mλ ) 2d sin θ (m
neutralized with 0.1 M sodium hydroxide. The sample was is the number of different orders). All the standards, except for
cooled. Crystals were removed with a Buchner funnel and NaP, are pure in this respect (Figure 1). The NaP sample has
allowed to air-dry to constant weight (about 1 week). These two crystal phases such as the β and δ phases in soap vernacular,
crystals were determined to be anhydrous, which was confirmed and it is difficult to synthesize pure soap phases with any known
by infrared analysis. preparation procedures. The powder XRD pattern for the NaHP2
554 J. Phys. Chem. B, Vol. 105, No. 2, 2001 Lynch et al.

integrated for 2 s at each step. The data were collected and


normalized on an interfaced computer using Philips diffraction
software. Precise crystallographic spacings were made with 1/4°
slits, 0.005 °2θ increments, and 5 s integration times at angles
less than 2° 2θ and with 1/2° slits, 0.01 °2θ increments, and 2
s integration times at angles greater than 15° 2θ. All X-ray
spacings were standardized using a powdered crystalline silicon
standard (NBS 640b).
Infrared Measurements. The FT-IR measurements were made
using a Digilab FTS 7 spectrometer. All spectra were measured
on powder samples using an attenuated total reflection (ATR)
accessory with a zinc selenide internal reflection element (IRE)
to preserve the crystal structure and eliminate perturbation of
the data due to sample preparation. Each spectrum represents
64 scans taken at 4 cm-1 resolution referenced to the spectrum
of a clean zinc selenide IRE. The spectra were not corrected
Figure 2. Powder XRD patterns of NaHP2 standard at 150 K compared for the distortion of peak intensities by ATR. Peak frequencies
with (superimposed lines) a computed powder pattern derived from were determined using the three-center weighted algorithm in
the single-crystal unit cell data at 298 K (see text).
the GRAMS/386 software. Intensities are shown in unnormal-
ized absorption units.
TABLE 1: Unit Cell Data at Different Temperatures
unit cell 3. Results and Discussion
parameter at 150 K at 293 K
Crystal Structure and Hydrogen Bonding in NaHP2. The
a 9.906 Å 9.989 Å structure of NaHP2 is illustrated in Figure 3. The carboxylate
b 7.163 Å 7.379 Å
c 45.58 Å 45.77 Å and acid headgroups are arranged in pairwise fashion to form
β 92.78° 92.78° a bilayer structure. The carboxylate headgroup was differentiated
from the acid headgroup on the basis of carbon-oxygen bond
lengths (Figure 3C and Table 2). The hydrogen atom was fixed
is highlighted in Figure 2, superposed on the intense reflections in a calculated position for illustrative purposes. The carboxylate
(I/Imax > 1%) computed from the single crystal data. It should and acid groups are held together via two interactions: H-
be noted that the original data were collected at 150 K and the bonding and coordination to a common Na atom as part of a
computed reflections did not suitably match the XRD pattern six-member ring defined by hydrogen, O2′, NaB, O1A, C1A,
of ambient temperature. However, the unit cell dimension was and O2A (Figure 3B). The strength of the interaction between
remeasured at 293 K (Table 1). There was a thermal expansion the hydroxyl group of the acid and the carboxylate oxygen
of the unit cell. The 2θ positions, calculated from the ambient manifests itself in the short distance of about 2.44 Å and a
temperature cell parameters, now match the powder diffraction possible angle of about 176° subtended at the hydrogen (Figure
data (Figure 2). The stoichiometry of the crystals was confirmed 3C).
by the elemental analysis of pure powders. The Na ions are sandwiched between the carboxylate and
Single-Crystal XRD Measurements and Analysis. X-ray acid headgroups, but each carboxylate-oxygen interacts sepa-
diffraction data were collected using the Siemens (Bruker) P4- rately with a symmetry-related sodium ion. The immediate
diffractometer with monochromatic CuKR radiation. The XS- coordination sphere of Na is comprised of five oxygens, three
CANS data acquisition software was used. The crystal structure of which are symmetry-related carboxylate oxygens while the
was solved and refined in the SHELXTL suite of programs. other two come from symmetry-related acids. The average
All nonhydrogen atoms were located experimentally and refined O-Na distance is about 2.37 Å, while the closest Na is about
anisotropically. Hydrogen atoms were fixed in a riding model 3.35 Å away.
except the carboxylic acid H, which was put in a calculated The long hydrocarbon chains assume all-trans conformation
position after the refinement. with some twist around the headgroup area. The twist of the
Refinements were carried out on F2 for all reflections. The chain with respect to the mean-plane defined by the carboxylic
weighted R-factor, Rw, and the goodness of fit, S, are based on acid/carboxylate headgroup is 35° for the acid and about 7.5°
F2. The conventional R-factor is based on F, with F set to zero for the carboxylate. Headgroup twist values similar to the above
for reflections with negative F2.35 The observed criterion of F2 have been observed for lauric acid and potassium palmitate.
> 2σ(F2) was used for calculating the R-factor. R-factors based Discussion of the MD Simulations. We used empirical
on F2 are statistically about twice as large as those based on F, interaction potentials with the following form:
and R-factors based on all data are even larger. The relatively
large R-factor for this structure is due to crystal quality and U ) Ubond + Uangles + Utorsions + Unonbonded
possibly the entropy of the long chains. Further data collection,
refinement details, atomic coordinates, and thermal parameter Here, Ubond and Uangles are harmonic terms that account for
tables will be given as Supporting Information. intramolecular bond stretching and bond angle deformations,
Powder X-ray Measurements. The powder XRD studies were and Utorsions is a appropriate periodic torsion term. Unonbonded
done with a Philips PW1830 automated powder diffraction describes the intermolecular potential and is based on pairwise
system. X-rays were generated from a Cu anode tube at 45 kV additive site-site electrostatic and Lennard-Jones contributions.
and 40 mA, and the diffracted X-rays were measured using a To model the hydrocarbon part of the long-chain molecules,
proportional counter. Standard measurements were made from we used the all-atom potential that reproduces the structures of
2.0° to 40.0° 2θ at 0.05° 2θ increments. The signal was solid and liquid alkanes well.36 The sodium ion parameters were
Structure of Fatty Acid-Soap Crystals J. Phys. Chem. B, Vol. 105, No. 2, 2001 555

Figure 3. Structure of the NaHP2 crystal. The top drawing (A) shows the alkyl chain and headgroup arrangement, while the bottom drawing (B)
delineates the headgroup region. The local coordination around sodium should be noted. (C) Illustration of the specific bonding distances (in Å)
associated with an acid-carboxylate pair with the NaHP2 crystal.

taken from Chandrasekhar et al.37 without further modification. the lauric acid C11H23COOH and the B-form of stearic acid
The carboxylate headgroup was modeled using parameters from C17H23COOH. To our knowledge, there is no atomic-resolution
Jorgensen and Gao,38 while for the acid headgroups we used crystallographic data on sodium soaps against which to compare
parameters from Briggs et al.39 MD simulations.
Most of the components of the force field have been tested The crystal structure of the A-super form of the lauric acid
independently and exhibited reasonable agreement with experi- was taken from Goto and Asada11,41 (Figure 4a). The symmetry
mental results for the structural and thermodynamic properties operators of the triclinic P1h space group were used to generate
of model compounds and/or single component systems. How- the atomic coordinates from the asymmetric units. All of the
ever, there is no guarantee that they will work equally well when hydrogens were added in the simulation, including the carboxyl-
combined in the systems under study. To test these models, we group protons, whose positions were set such that the lauric
set up MD simulations for pure acid crystals. Most fatty acids acid molecules form hydrogen-bonded dimers. The MD simula-
are known by powder X-ray diffraction to exist in three tion cell comprised 60 molecules in a 5a × b × c lattice (i.e.,
crystalline polymorphs, named A(triclinic), B (monoclinic), and 2280 atoms). We note here that, as is commonly found in fatty
C (monoclinic),40 all of which are stable. We have generated acids, the molecules are bound together to form dimers through
trajectories at equilibrium for the A-super form of the lauric hydrogen bonds between the carboxyl groups; however, for this
acid C12 and the B-form of stearic acid C18 (see descriptions somewhat unusual structure, the carboxyl groups are not all
of the structures below). These included the A-super form of located in the same plane.
556 J. Phys. Chem. B, Vol. 105, No. 2, 2001 Lynch et al.

molecules (i.e., 6a × 8b × 2c lattice) and 48 sodium ions (i.e.,


about 4800 atoms). The carboxyl-group protons whose positions
were not determined from the X-ray analysis were added in the
simulation. Their positions were set such that two opposing
molecules (in c direction) are hydrogen bonded (Figure 4c).
The MD trajectories were carried out with three-dimensional
periodic boundary conditions. For every system, short equilibra-
tion runs were performed at constant temperature and volume,
followed by a longer run at constant temperature, and pressure
(NPT) using the hybrid MD algorithm developed by Martyna
et al.42,43 The extended system (ES) equations of motion were
integrated by using an iterative Verlet-like algorithm. The
fictitious masses of the ES variables were chosen according to
the prescript ion given by Martyna et al.,44 with time scales of
0.5 ps for the thermostats and 1 ps for the volume and cell
variables. The Nosé-Hoover thermostat chain length was five.
The Ewald method was used to compute the electrostatic
interactions, and the minimum image convention was employed
Figure 4. Initial arrangement of crystal structures used to validate the to calculate the real-space part of the Ewald sum and the van
interaction parameters in the molecular dynamic simulations: (a) the der Waals interactions with simple truncation at 10 Å.45 Bonds
A-super form of lauric acid, (b) the B-form of stearic acid, and (c) the
NaHP2 acid-soap defined by the present crystallographic data.
involving hydrogen atoms were constrained using the SHAKE
algorithm.46
TABLE 2: Carboxylate/Carboxylic Acid Headgroup and The A-super form of the lauric acid remained stable during
H-bond Parameters for NaHP2 the 160 ps simulation. Since we are dealing with solid phases,
Bond Lengths a NPT simulation of a 160 ps is adequate in testing for “local”
CdO 1.199 Å stability. Any tendency to “transform” to a different structure
C-O 1.313 Å usually manifests itself in large fluctuations in unit cell
CdO 1.258 Å
CdO 1.300 Å parameters, which serve as precursors to a transition. In the
O f Na (avg) 2.37 Å present case, no such latent instability was apparent. The overall
agreement of the unit cell parameters with experiment is
H-bond Parameter
satisfactory, with variations less than 3% (Table 3). For stearic
atoms (D-H‚‚‚A)‡ D‚‚‚A (Å) H‚‚‚A(Å) ∠DHA
O2-H‚‚‚O2′a 2.444 Å 1.5 Å 176.2° acid, the crystal remained stable during the 160 ps NPT
simulation. We compare the average cell parameters to the
The initial configuration for the stearic acid crystal was taken experimental values (Goto et al.)12 in Table 4. Overall the
from Goto et al.12 and corresponds to the B-form (Figure 4b). agreement is very good; the most significant deviations are less
The space group P21/a operators and lattice translations were than 2.7% in the a and b dimensions. In Table 5, we report the
used to generate all of the atomic coordinates from the average cell parameters from the MD simulation for the NaHP2
asymmetric units. The carboxyl-group protons whose positions acid crystal; the agreement again with experiment is very
were not determined from the X-ray analysis were added in the satisfactory. The crystal remained stable during the 160 ps NPT
simulation. Their positions were initially set to reproduce the simulation (Figure 5). Taken together, these results validate the
double hydrogen bonding involving the carboxyl dimers. The force field parameters used in the calculation.
MD simulation cell contained 100 molecules in a 5a × 5b × c To compare the headgroup organization of the acid and acid-
lattice (i.e., about 5600 atoms). soap molecules, we display in Figure 4 snapshots of the
For the sodium palmitate-palmitic acid 1:1 acid-soap assemblies for the A-super form of the lauric acid, the B-form
crystal, the initial configuration was taken from the X-ray data of the stearic acid, and the NaHP2 sodium acid-soap. In the
analysis (cf. above). Accordingly, the space group P21/a first two cases, the molecules labeled 1 and 2 form hydrogen-
operators and lattice translations were used to generate all of bonded dimers. In the latter case, an acid molecule (1) forms a
the atomic coordinates from the asymmetric units in a 3a × 4b single hydrogen bond with a palmitate ion (2) and shares a
× c lattice. The resulting simulation box contained 48 × 2 sodium ion with a second neighboring ion (4) while the ionic

TABLE 3: Experimental and Simulation Results for Super-A Form Lauric Acid
a (Å) b (Å) c (Å) R (deg) β (deg) γ (deg)
X-ray 5.415(2) 25.96(1) 35.18(13) 69.82(4) 113.14(4) 121.15(3)
simulation 5.44(3) 25.63(30) 34.66(89) 70.67(73) 112.0(10) 118.0(13)

TABLE 4: Experimental and Simulation Results for Form B Stearic Acid


a (Å) b (Å) c (Å) R (deg) β (deg) γ (deg)
X-ray 5.587(10) 7.386(10) 49.33(8) 90.0 117.2(9) 90.0
simulation 5.44(3) 7.59(5) 49.01(9) 90.01(40) 117.53(35) 89.99(36)

TABLE 5: Experimental and Simulation Results for NaHP2


a (Å) b (Å) c (Å) R (deg) β (deg) γ (deg)
X-ray 9.906(1) 7.163(1) 45.580(6) 90.0 92.78(1) 90.0
simulation 9.80(6) 7.11(4) 45.25(6) 90.00(54) 93.40(31) 89.99(24)
Structure of Fatty Acid-Soap Crystals J. Phys. Chem. B, Vol. 105, No. 2, 2001 557

Figure 5. Unit cell dimensions as a function of time for the NaHP2 crystal as it evolved in the molecular dynamics simulation.

molecule (2) shares another sodium ion with the upper molecule
(3). Comparison of panels a-c of Figure 4 shows that the dimer
1-2, instead of sharing a double hydrogen bond like the acid
crystal, shares only one hydrogen bond and compensates for
the extra (layer-to-layer) binding energy by ionic interactions
with a pair of sodium ions.
It is difficult, considering only the local arrangement of the
dimers or dimers plus first neighboring molecules, to compare
quantitatively the energetics involved. In the acid crystal case,
the headgroup-to-headgroup interaction can be considered
primarily due to the hydrogen-bonded pairs; however, for the
NaHP2 crystal, a likely strong interaction with neighboring ions
is involved in the headgroup region. Rather than isolating
specific pairwise interactions, we compare the energetics of the
acid-like arrangements for the B-stearic acid with that of the
NaHP2 crystal by investigating the leaflet to leaflet binding
energies. To do so, we have calculated, using the force field
parameters that reproduced both crystal structures well, the van
der Waals and electrostatic interaction profiles between the upper
and lower layers as a function of their separations. The two
leaflets are pulled apart a distance r along the c axis, and the
intermolecular nonbonded interactions are estimated for the
whole system while considering periodic boundary conditions.
For the B-stearic acid, the successive configurations, obtained
by increasing r from 0.0 Å (equilibrium) to 2.0 Å, lead to
breaking the hydrogen bonds between molecules in opposite
layers (Figure 6). For the NaHP2 crystal, the positions of the
Figure 6. Energy necessary to separate the acid-soap and fatty acid
sodium ions were held fixed (in contrast to allowing them to as a function of distance.
move with the leaflets). The calculation included the contribution
from the chain-chain interactions within the same leaflet. shallower for the acid crystal compared to that of the sodium
The variation of the individual components contributing to acid-soap, which indicates that in the later case the leaflets
the nonbonded energies is shown as a function of r in Figure 6. are much more strongly bound.
To aid the comparison, we shifted the curves to a common Discussion of Spectroscopy of Acid-Soaps. The infrared
minimum. The results show that the energy variation is much spectra for HP and NaP are generally well understood. For HP
558 J. Phys. Chem. B, Vol. 105, No. 2, 2001 Lynch et al.

Figure 7. Entire midrange IR spectra of the standards: (A) HP, (B) NaP, (C) NaH2P3, (D) Na2H3P5, and (E) NaHP2.

crystals, the CdO stretch is at ∼1690 cm-1 47 (Figures 7A and NaP has a strong asymmetric stretching band (νasym COO-)
8A). However, the frequency of the stretch is known to be center at about ∼1550 cm-1 (Figures 7B and 8B). In general,
susceptible to intermolecular interactions. Fatty acid crystals carboxylate groups also have symmetric stretches (νsym COO-)
contain dimers, in which a pair of acid molecules aligns head- as low as ∼1400 cm-1, which are broad and generally have
to-head to form two hydrogen bonds using the acid proton on two to three peaks. These bands are obscured in the NaP spectra.
one molecule and the carbonyl oxygen on the other. Unasso- Finally, there may be a broad OH stretch envelope between
ciated fatty acids (i.e., not dimerized) can have CdO stretches 3000 and 3600 cm-1, which arises from the water of hydration.
as high as 1760 cm-1. In addition, there is a broad OH stretch Spectra for acid-soap show distinct changes in the carbonyl
envelope for HP between 2200 and 3500 cm-1 (Figure 7A). and the OH stretching bands. The acid-carbonyl stretching band
Structure of Fatty Acid-Soap Crystals J. Phys. Chem. B, Vol. 105, No. 2, 2001 559

Figure 8. Expansion of the carbonyl regions of the standards: (A) HP, (B) NaP, (C) NaH2P3, (D) Na2H3P5, and (E) NaHP2.

moves to higher frequencies: 1722.8, 1725.0, and 1724.0 cm-1 The change in the νasym COO- bands seems to reflect the
for NaH2P3, Na2H3P5, and NaHP2, respectively (Table 6). At change in bond order or strength of the CO bond. By classical
the same time, the asymmetric carboxylate stretching band shifts arguments, the frequency of the vibration is proportional to the
to lower frequencies, about 1520 cm-1 for NaH2P3, Na2H3P5, square root of the force constant.48 It is reasonable to suggest
and NaHP2. The band becomes broad and far less intense. There that longer bond distances imply weaker force constants, and
is an additional carboxylate stretching band in the NaHP2 spectra consequently, the band will shift to lower frequencies. Compar-
at 1623 cm-1. The OH stretching band for acid-soaps shift to ing the CO bond distance of NaHP2 with known potassium soaps
a broad, weak envelope between 2200 and 3000 cm-1. There (Table 7) shows that the bond distances are indeed longer in
has been no reported hydration within acid-soap crystals. the acid-soap. The increase in bond order likely reflects both
560 J. Phys. Chem. B, Vol. 105, No. 2, 2001 Lynch et al.

TABLE 6: Carbonyl Band Frequencies for Sodium of intermolecular interactions in long chain length acid-soaps.
Palmitate, Palmitic Acid, and Acid-Soaps The crystallography data provide succinct evidence for distinct
other acid and carboxylate groups within the acid-soap crystal, which
acid-soap acid-carbonyl carboxylate-carbonyl carboxylate eliminates the conjecture of acid dimers and carboxylate groups
crystal stretch (νasym stretch) stretch and some “blending” arrangement. The advantage of using MD
NaP 1557.1 cm-1 simulation is its ability to incorporate the entire crystal struc-
HP 1696.6 cm-1 tureand thermal effects, rather than just chain fragments.
NaH2P3 1722.8 cm-1 ∼1520 cm-1 (weak) Comparisons between the experimental data and the MD
Na2H3P5 1725.0 cm-1 ∼1520 cm-1 (weak)
NaHP2 1724.0 cm-1 ∼1520 cm-1 ∼1623 cm-1
simulations provide a quantitative test of the carboxylate and
fatty acid interactions. These interaction potentials utilized in
TABLE 7: Comparison of Carboxylate, Carboxylic Acid, the present study indicate that the sum of Coulombic and van
and Acid-Soap Bond Distances (Å) der Waals interactions is stronger in the acid-soap than in the
fatty acid. The present data can also be used to explain the
peculiar changes observed in the IR spectroscopy of acid-soaps.
The presence of definitive acid and carboxylate fractions in these
crystals explains why both acid-like and carboxylate-like
features are observed in the IR spectra. However, subtle
variations in bond distances (and, hence, intermolecular interac-
tions) are reflected in deviations of these spectral features from
the parent acid and soap crystals. Comparisons with other acid-
soaps allow us to make reasonable assessments as to the nature
of all these crystals and thereby broaden our understanding of
the field.

Acknowledgment. The authors would like to acknowledge


the hydrogen bonding and placement of the carbonyl within Robert Laughlin and Curt Marcott, both from Procter and
the lattice. The shift of the νasym COO- in the other acid-soaps Gamble’s Corporate Research Division, for many useful dis-
implies a comparable situation. It is impossible to explain the cussions. The simulation work at the University of Pennsylvania
additional carbonyl peak in the NaHP2 by this argument. was supported by a generous grant from the Procter and Gamble
However, it was previously noted that the νsym COO- band is Company.
often observed as multiple peaks, which would be a reasonable
assignment for this band. Supporting Information Available: Five tables describing
Differences in the acid CdO and O-H stretching bands seem the crystallographic data. This material is available free of charge
to reflect changes in the intermolecular interactions. The acid via the Internet at http://pubs.acs.org.
group is not dimerized in the acid-soap, which, as previously
noted, causes the CO stretching band to shift to higher References and Notes
frequencies. Further, the hydrogen bond in the NaHP2 crystal (1) Chevreul, M. E. Rech. Chim. Corps Gras d'Origine Animal (Paris)
is short (2.444 Å). Short hydrogen bonds are known to shift 1823, 34, 53, 408.
(2) Ekwall, P.; Mylius, W. Ber. Dtsch. Ges. 1929, 62, 1080.
the OH groups to lower frequency.49 So rather than a broad (3) Ekwall, P.; Mylius, W. Ber. Dtsch. Ges. 1929, 62, 2687.
envelop between 3000 and 3600 cm-1, the envelop shifts to (4) Ekwall, P. Z Anorg. Allg. Chem. 1933, 210, 337.
2200-3000 cm-1. A comparison of acid CO bond distances (5) McBain, J. W.; Stewart, A. J. Phys. Chem. 1927, 31, 2.
within NaHP2 and other fatty acid crystal data suggests these (6) McBain, J. W.; Field, M. C. J. Phys. Chem. 1933, 37, 675.
(7) McBain, J. W.; Field, M. C. J. Chem. Soc. 1933, 920.
bond lengths are comparable and, consequently, cannot be used (8) McBain, J. W.; Stewart, A. J. Chem. Soc. 1933, 924.
to explain the shift. Again, it is assumed from parity with other (9) Lynch, M. L. Curr. Opin. Colloidal Interfacial Sci. 1997, 2, 495.
acid-soap spectra that these crystals must have a similar type (10) Lomer, T. R. Acta Crystallogr. 1963, 16, 984.
of arrangement. (11) Goto, M.; Asada, E. Bull. Chem. Soc. Jpn. 1978, 51, 70.
(12) Goto, M.; Asada, E. Bull. Chem. Soc. Jpn. 1978, 51, 2456.
While we have attempted to qualitatively describe the changes (13) Malta, V.; Celotti, G.; Zannetti, R.; Martelli, A. F. J. Chem. Soc.
in the intermolecular vibrational spectra when going from pure B 1971, 548.
acid-and-soap compounds to the acid-soap crystal, it is clear (14) Keneko, F.; Kobayashi, M.; Kitagawa, Y.; Matsuura, Y. Acta
that a quantitative description of the changes observed is much Crystallogr., C 1990, 46, 1490.
(15) Speakman, J. C.; Mills, H. H. J. Chem. Soc. 1961, 1164.
more complicated. Indeed, the vibrational properties (i.e., (16) Speakman, J. C. Nature 1948, 162, 695.
frequencies and intensities) of the carbonyl and carboxylate (17) Goerbitz, C. H.; Etter, M. C. J. Chem. Soc., Perkin Trans. 1992, 2,
moieties are dependent on the interactions of these moieties with 131.
(18) Mantsch, H. H.; Weng, S. F.; Yang, P. W.; Eysel, H. H. J. Mol.
their immediate environment.50,51 The molecular dynamics Struct. 1994, 324, 133.
simulation trajectories can in principle be used to compute the (19) Bian, J.; Weng, S.; Wu, J.; Xu, G. Beijing Daxue Xuebao, Ziran
IR spectra for the crystals understudy. However, the force field Kexueban 1995, 31 (6), 711.
used in this study is not optimized for such an analysis. A (20) Lynch, M. L.; Pan, Y.; Laughlin, R. G. J. Phys. Chem. 1996, 100,
357.
quantitative analysis of the IR spectra therefore necessitates (21) Wen, X.; Franses, E. I. J. Colloid Interface Sci. 2000, preprint.
either the use of a more suitable force field (cf. Ishioka et al.)52 (22) Tandon, P.; Raudenkolb, S.; Neubert, R. H. H.; Retting, W.;
or a quantum calculation. Wartewig, S. J. Phys. Chem. B 2000, preprint.
(23) Etter, M. C.; Ruetzel, S. M.; Vojta, G. M. J. Mol. Struct. 1990,
237, 165.
4. Conclusions (24) Etter, M. C.; Hoye, R. C. Trans. Am. Crystallogr. Assoc. 1986, 22,
31.
The present study employs a combination of crystallography, (25) Porter, M. R. Handbook of Surfactants; Blackie Academic &
MD simulation, and infrared spectroscopy to explore the nature Professional: London, 1997; p 100.
Structure of Fatty Acid-Soap Crystals J. Phys. Chem. B, Vol. 105, No. 2, 2001 561

(26) Bartolo R. G.; Lynch M. L. Kirk-Othmers Encyclopedia of Chemical (38) Jorgensen, W. L.; Gao, J. J. Phys. Chem. 1986, 90, 2174.
Technolgies, 4th ed.; Wiley: New York, 1997. (39) Briggs, J. M.; Nguyen, T. B.; Jorgensen, W. L. J. Phys. Chem.
(27) Kacher, M. L.; Taneri, J. E.; Camden, J. B.; Vest, P. E.; Bowles, 1991, 95, 3315.
S. J. Shaped Solid Made with a Rigid, Interlocking Mesh of Neutralized (40) Francis, F.; Piper, S. H. J. Am. Chem. Soc. 1939, 61, 577. Clara,
Carboxylic Acid. U.S. Patent 5340492, 1994. G. L. Applied X-rays; Mc Graw-Hill: New York, 1955; p 608. Von Sydow,
(28) DeBretteville, A.; Reyer, F. V. J. Phys. Chem. 1944, 48, 154. Ryer, E. Acta Chem. Scand. 1955, 9, 1685; 1959, 13, 984.
F. V. Oil Soap 1946, 3100. (41) Goto, M. I.; Asada, E. Bull. Chem. Soc. Jpn. 1978, 51, 1456.
(29) Cistola, D. P.; Atkinson, D.; Hamilton, J. A.; Small, D. A. (42) Martyna, G. J.; Tobias, D. J.; Klein, M. L. J. Chem. Phys. 1994,
Biochemistry 1986, 25, 2804. 101, 4177.
(30) Cistola, D. P.; Hamilton, J. A.; Jackson, D.; Small D. A. (43) Martyna, G. J.; Tuckerman, M. E.; Tobias, D. J.; Klein, M. L. Mol.
Biochemistry 1988, 27, 1881. Phys. 1996, 87, 1117.
(31) Small, D. M. Handbook of Lipid Research- The Physical Chemistry (44) Martyna, G. L.; Klein, M. L.; Tuckerman, M. J. Chem. Phys. 1992,
of Lipids from Alkanes to Phospholipids; Plenum Press: New York, 1986; 97, 2635.
p 285. (45) Allen, M. P.; Tildesley, D. J. Computer Simulation of Liquids;
(32) Desiraju, G. R. In Material Science Monographs, 54, Crystal Oxford University Press: New York, 1989.
Engineering: The Design of Organic Solids; Elsevier: New York, 1989; (46) Ryckaert, J. P.; Ciccotti, G.; Berendsen, H. J. C. J. Comput. Phys.
Chapter 5. 1977, 23, 327.
(33) Bernstein, J.; Etter, M. C.; Leiserowitz, L. Struct. Correl. 1994, 2, (47) Socrates, G. Infrared Characterization Group Frequencies;
431. Wiley: New York, 1980; Chapter 10.
(34) Videnova-Adrabinska, V. J. Mol. Struct. 1996, 374, 199. (48) Rao, C. N. R. Chemical Applications of Infrared Spectroscopy;
(35) Stout, G. H.; Jensen, L. H. X-ray Structure Determination, A Academic Press: New York, 1963; Chapters 1 and 4.
practical Guide, 2nd ed.; Wiley: New York, 1989. (49) Lord, R. C.; Merrifield, R. E. J. Am. Chem. Soc. 1952, 166.
(36) Tobias, D. J.; Tu, K.; Klein, M. L. J. Chim. Phys. Phys.-Chim. (50) Buckingham, A. D. Proc. R. Soc. 1960, 248 (a), 169.
Biol. 1997, 94 (7-8), 1482. (51) Buckingham, A. D. Trans. Faraday Soc. 1960, 56, 753.
(37) Chandrasekhar, J.; Spellmeyer, D. C.; Jorgensen, W. L. J. Am. (52) Ishioka, T.; Murotani, S.; Kanesaka, I.; Hayashi, S. J. Chem. Phys.
Chem. Soc. 1984, 106, 903. 1995, 103, 1999.

You might also like