Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Langmuir 1985, 1, 45-52 45

Spontaneously Organized Molecular Assemblies. 1.


Formation, Dynamics, and Physical Properties of
n -Alkanoic Acids Adsorbed from Solution on an Oxidized
Aluminum Surface
David L. Allara*
Bell Communications Research, Murray Hill, New Jersey 07974

Ralph G . Nuzzo*
AT&T Bell Laboratories, Murray Hill, New Jersey 07974

Received August 16,1984

This study shows that closest packed, oriented monolayers of n-alkanoic acids can be formed on oxidized
aluminum substrates by adsorption from dilute solution. The formation of these organic surface phases
is characterized by complicated kinetics in which surface and/or monolayer defects, as well as impurities,
seemingly play important roles. The structures so obtained are dynamic in nature in that they undergo
rapid exchange with ligands in solution. The data further suggest an important limitation to self-assembly
in this system which is directly related to the length of the n-alkanoic acid tail. The influences of experimental
variables on the formation of oriented monolayers is discussed.

Introduction of quasi-two-dimensionalorganic assemblies and the pro-


cesses by which they can be prepared. As regards the latter
The adsorption and spontaneous organization of am- point, spontaneous adsorption methodologies are to be
phiphilic molecules, especially the n-alkanoic acids, on contrasted with the Langmuir-Blodgett transfer process'*
smooth, ambient' metal surfaces, are the subject of a vast in which ordered, multilayer assemblies can be produced
l i t e r a t ~ r e . ~ - 'The
~ interest in these surface adsorbate for the longer chain acids &e., fatty acids), particularly
structures traditionally has been motivated by their rele- when the total number of carbon atoms is greater than 18.
vance to such important technological areas as lubrication, In the case of spontaneous adsorption, no preassembling
corrosion, and catalysis. In recent years, speculative of the adsorbate molecules is carried out; the substrate is
technological applications, involving such areas as optics exposed directly to the adsorbate in solution or the gas
and microelectronics,16 as well as the development of phase and any ordering that occurs is thus considered to
surface-modified electrodes for electrochemical applica- be spontaneo~s.'~The questions of and related to the
' ~ generated a renewed interest as to the nature
t i o n ~ have adsorption mechanism(s) (reflecting kinetics) and equi-
librium structures (reflecting thermodynamics)are of great
interest, but past studies have been intended to concen-
(1)The term ambient is used to indicate that moat attention has trate almost exclusively on the latter. It is not clear in all
centered on metals having oxide overlayers which may or may not have cases, however, whether true equilibrium structures, com-
contained various levels of impurities and/or contaminants.
(2)A review of earlier work can be found in: Bowden, F. P.; Tabor, prised solely of the intended adsorbate(s), in fact have been
D."The Friction and Lubrication of Solids", Oxford University Press: studied.
London; 1968; Part 11, Chapter 19. Most studies of the adsorption of carboxylic acids on
(3)(a) Chapman, J. A.; Tabor, D. R o c . R. SOC.1967,London, Ser. A
242, 96-107. (b) Brockway, L. 0.; Karle, J. J. Colloid Sci. 1947, 2, metals have involved surfaces exposed to air prior to ad-
277-287. (c) Menter, J. W.; Tabor, D.R o c . R. SOC.London, Ser. A 1961, sorption and thus largely refer to adsorption on native
204,514-524. (d) Sagiv, J. J. Am. Chem. SOC.1980,102,92-98. oxide overlayers. These studies, which largely complement
(4)Walker, D.C.; Ries, H. E. Adv. Chem. Ser. 1964,No. 43,295-301. the voluminous data on bulk oxide powders, can be divided
(5)Doyle, W. P.; Ellison, A. H. Adu. Chem. Ser. 1964,No. 43,268-274.
(6)Gaines, G. L. J. Colloid Sci. 1960,15, 321-339. into three general functional classes: (1)acids chemisorbed
(7)Timmons, C.0.; Zisman, W. A. J. Phys. Chem. 1965,69,984-990. to form metal carboxylate salts, processes that may involve
(8)Timmons, C. 0.; Patterson, R. L.; Lockhart, L. B. J. Colloid In- substrate corrosion and multilayer formation; (2) acids
terface Sci. 1968,26, 120-127. chemisorbed via proton transfer to a lattice oxygen atom
(9)Bornong, B. J. Surf. Sci. 1969,16,321-330.
(10)Allara, D.L.; Tompkins, H. G. J . Colloid Interface Sci. 1974,49,
to yield monolayer structures; (3) acids chemisorbed
410-421. (perhaps physisorbed is a more appropriate description)
(11)Boerio, F. J.; Chen, S. L. J. Colloid Interface Sci. 1980, 73, with no proton transfer. Representative examples of these
176-185. three are carboxylic acid chemisorption on copper,6,10J1
(12)Brown, N. M. D.; Floyd, R. B.; Walmsley, D. G. J. Chem. SOC., aluminum,'2-16 and siliconFOrespectively. Although the
Faraday Trans. 2 1979,75,261-270.
(13)Lewis, B. F.; Mosesman, M.; Weinberg, W. H. Surf. Sci. 1974.41, cases of multilayer formation are quite important and of
142-164. significant interest, the less complicated cases of monolayer
(14)Golden, W. G.; Snyder, C.; Smith, B. J. Phys. Chem. 1982,86, chemisorption are more attractive as beginning points for
4675-4678.
(15)Cass, D.A,; Strauss, H. L.; Hansma, P. K. Science (Washington,
D.C.)1976,192,1128-1130. (18)For example, see: Gaines, G. L. 'Insoluble Monolayers at Liq-
(16)For example, see: (a) Vincett, P. S.; Roberts, G. C. Thin Solid uid-Gas Interfaces"; Interscience: New York, 1966.
film 1980,68,135-171.(b) Special issue on molecular electronics: Proc. (19)As we shall detail, spontaneous, in this context, cannot be used
IEEE 1983,230,197-263. interchangeably with the word instantaneous.
(17)For example, see: Faulkner, L. R. Chem. Eng. News 1984,28-45 (20)Iler, R. K. "The Chemistry of Silica"; Wiley: New York, 1979;
and references therein. Chapter 6.

0743-7463/85/2401-0045$01.50/0 0 1985 American Chemical Society


46 Langmuir, Vol. 1, No. 1, 1985 Allara and Nuzzo
careful structural studies of the first formed layer. wetting data, an improved understanding of the structure
A concern with regard to many of the studies cited and dynamic character of the self-assembly of n-carboxylic
above, in particular those that provide contrasts with acid monolayers on aluminum has been obtained. Our
Langmuir-Blodgett techniques, is whether the term self- results suggest an extremely important role for kinetics
assembling is in fact appropriate. These concerns focus in the formation of equilibrium closest packed structures.
on two issues. First, are monolayers with oriented (or Through an examination of both a homologous series of
perhaps ordered), closest-packed structures of low surface high-purity acids (C,-c24) and vinyl- and propargyl-ter-
free energy indeed formed? Second, to what extent do the minated C22acids, the influence of adsorbate structure on
various experimental variables-solvent, temperature, monolayer structure has been detailed. Our results further
carbon-chain length, substrate, and concentration as rep- demonstrate that these monolayer structures are capable
resentative examples-influence the nature of the surface of undergoing extremely facile exchange reactions with
phases that are formed? materials in solution. Taken together, these results develop
The classic study of Timmons and Zisman7concluded a rich though complicated coordination chemistry for
that for chain lengths of C14 and longer, close-packed carboxylic acids chemisorbed on this (and presumably
monolayers can be prepared on Pt; on NiO substrates, other) oxide surface.
chain lengths as low as Clo were successfully employed. An
earlier study of Zisman21 presented the interesting ob- Experimental Section
servation that the coefficient of friction of a glass surface, Materials. The n-alkanoic acids from C6 and above were
covered with a homologous series of fatty acids, drops to obtained from Sigma Chemical Co. and were used without further
a minimum (and nominally constant) value for chain purification. GC analyses were obtained for the liquid acids and
lengths of C14or greater. Borno~$ observed that coverages showed purities of >99.9%. The melting points of the higher acids
of fatty acids (C6--c30)adsorbed from the melt on chro- were checked and all showed melting ranges of less than 2 "C.
mium substrates were less than that expected for a The terminally substituted acids CH2=CH(CH2)lgC02H and
Langmuir-Blodgett-type film; no discontinuities of the CH=C(CH2),,C02H were synthetic samples supplied by Prof.
type described above were observed. Various studies of George Whitesides, Harvard University, and had melting points
of 70.0-70.5 and 78.5-79.5 "C, respectively. Hexadecane, obtained
specific longer chain acids (CI6and above) seem to offer from Aldrich Chemical Co., had a nominal purity of 99%. It was
conflicting results as to whether submono-, mono-, and/or further purified by slow elution through a column of activity 1
multilayer coverages are obtained for ambient metal sur- alumina and stored under nitrogen. Deuterated alkanoic acids
faces such as copper416J1and aluminum.6J4 For example, were obtained from Cambridge Isotope Laboratories (Cambridge,
Gaines reported that, for the latter, greater than monolayer MA) and were used without further purification. Acetic acid was
coverages were obtained for stearic acid (C1J in sharp obtained from J. T. Baker Chemicals as the Ultrex grade (99.9%)
contrast to the results of Golden, Snyder, and Smith14 and was used as received. Aluminum used for the substrate
which show submonolayer coverages for arachidic acid preparation was 99.999% purity. Tritiated acetic anhydride was
(C20). The literature is repleat with such discrepancies obtained from New England Nuclear Co. (Boston, MA).
Sample Preparation and Treatment. Aluminum substrates
despite an impressive array of experimental techniques were prepared by the evaporation of aluminum from resistively
(electron diffraction, reflection infrared spectroscopy, ra- heated tungsten boats onto 2-in.-diameter single-crystal silicon
dioisotopic labeling, etc.) which have been brought to bear wafers which had been polished to C40-A surface roughness.
on this problem. Thus, while many additional comparisons Depositions were carried out in the lo4 to torr range in an
can be made, no general conclusions can be drawn. These oil diffusion pumped system equipped with a liquid nitrogen trap
continuing significant and puzzling differences, in our assembly. After melting the aluminum, a shield was removed,
opinion, reflect the need for a detailed, self-consistent thus exposing both the samples and a quartz crystal thickness
reexamination of this classic and widely reported chemi- monitor which was placed nearby in order to determine the
sorption system. thickness of the deposited film. Approximately 200 nm of A1 was
It has been our contention22that spontaneous adsorp- deposited and the chamber then backfilled with high-purity re-
search-grade oxygen to 1 atm. The samples were removed and
tion, under the appropriate conditions, out to be capable treated in one of two ways: (1) the samples were examined quickly
of producing monolayers with structural features akin to by ellipsometry and, within several minutes, placed into the
those of the Langmuir-Blodgett type. This contention is adsorbate solution; (2) the samples were placed directly into an
not new to us, having been suggested by previous authors%3 ethanol solution of acetic acid (0.05 M) for -15 min, removed,
and recently broadly demonstrated by Hubbard et al.23*24 rinsed with pure ethanol on a spinner (Headway photoresist type),
for the adsorption of nonamphiphilic molecules (differing examined by ellipsometry, and then placed in the adsorbate
structurally from the type reported herein) on metal solution. The acid solutions were made by dissolving the weighed
surfaces. In the present and the following paperz5 we portion of acid in hexadecane contained in a Pyrex flask which
report the results of our detailed reexamination of the had been precleaned in ethanolic KOH followed by repeated water
and methanol rinses. For the acids C18 and above, solution only
adsorption of carboxylic acids on native aluminum oxide could be obtained by warming the contents to -35-40 "C. The
surfaces. From reflection infrared spectroscopy,z6ellip- solutions were transferred to wide-mouth, screw-top polypropylene
sometry, isotopic labeling (both 2H and 3H), X-ray containers, precleaned with chloroform and methanol washes, and
photoelectron spectroscopy (XPS), and contact angle maintained a t 25.0 f 0.3 "C during sample immersion. The
solutions of acids CISand above tended to remain supersaturated
~~~ ~ ~~ ~
at 25 "C although occasionally one would form crystals and require
(21)Zisman, W.A. In "Friction and Wear"; Davies, R., ed.; Elsevier: rewarming. Following immersion, the samples were removed and
New York, 1959;pp 110-148. the surface rinsed quickly on the spinner with HPLC grade
(22)Nuzzo, R. G.;Allara, D. L. J. Am. Chem. SOC.1983, 105, hexane.
4481-4483.
(23)Soriaga, M. P.;Hubbard, A. T. J. Am. Chem. SOC.1982,104, Infrared Measurements. IR spectra of adsorbates were taken
3937-3945 and references cited therein. by reflection using a Digilab 15B spectrometer with modified
(24)Hubbard, A. T.Acc. Chem. Res. 1980,13, 177-184. optics. Details are given elsewhere.25
(25)Allara, D.L.;Nuzzo, R G. Langmuir, following paper in this isaue. Ellipsometry Measurements. Ellipsometric measurements
(26)For discussions of the application of infrared spectroscopy to the were made using a Rudolph 423 null ellipsometer equipped with
determination o f alkyl chain orientation and ordering in monolayer
structures, see: Allara, D. L.; Swalen, J. D. J . Phys. Chem. 1982,86, a Babinet-Solet compensator set a t 7r/4 retardation and with the
2700-2704. Rabolt, J. F.; Burns, F.C.; Schlotler, N. E.; Swalen, J. D. J.
Chem. Phys. 1983,78,946-952. -
optical axes at 45" to the instrument verical axis. Measurements
were made with an 1-mm beam at 623.8 nm (He-Ne laser) and
Spontaneously Organized Molecular Assemblies. 1 Langmuir, Vol. I , No. 1, 1985 47
at 442.0 nm (He-Cd laser) using a 70" angle of incidence at the
sample. Little difference was seen, however, in the results at the
two wavelengths, and for convenience,most measurements were
done at 632.8 nm. Spatial variations across the sample were - II
usually within the experimental errors of m-fl.5 A in terms of I
the final film thicknesses calculated from the readings. I
Contact Angle Measurements. Static contact angles were 30 - 0 I 0
measured with a Ram6-Hart goniometer. Drops were placed on
the surface in the ambient environment and read several times
on both sides of the drop. Angles were read to -&loand were
reproducible from sample to sample to <f3".
Radioisotopic Labeling. Tritiated acetic anhydride (5 mCi;
50.0 mCi/mmol) was dusolved in a sufficient quantity of absolute 20 -
ethanol under nitrogen to give a specific activity of acetic acid
(arising from esterolytic cleavage) of 25.0 mCi/mmol with a
concentration of 0.0050M. Sampleswere incubated by contacting
only the aluminum surface with the radiolabel. Calculations were 0 8.1
based on the measured activity of aliquots of known volume which 10 - 0

t
1
I
had been removed from the exchange or sample digestion media.
Aquasol-2 scintillation cocktail (NEN) and a Beckman 300
scintillation counter were used in all experiments.
Photoemission Measurements. A Kratos XSAMSOO pho-
- . a 'i
8
',e
I

toemission spectrometer utilizing a hemispherical analyzer and I


operating in a fixed analyzer transmission mode was employed 0 I , I I I I
for data acquisition. The instrumental resolution was 1 eV, using
1-mm mechanical apertures in conjunction with a Mg Ka X-ray
source. Digital data were acquired by using 0.05-V steps. The
base operating pressure was <5 X lo-" torr. CARBON NUMBER
Figure 1. Ellipsometry data showing the limiting f i i thicknesses
Results of n-alkanoic acids adsorbed on an oxidized aluminum substrate
Calculation of Film Thicknesses from Ellipsometry [RC02H]= 0.0050 M in n-hexadecane, 25 "C.
Data. In order to transform the raw data of the ellipso-
metry measurements into film thicknesses, the standard
parallel layer, classical electromagnetic model was cho-
sen.25p27 In this model one assumes that the system is
reasonably described as a flat sample consisting of an in-
finite substrate with a parallel overlayer of thickness d.
All phases are considered to be of homogeneous compo-
sition and isotropic, with the optical behavior of each phase
being accurately described by a single dielectric function.
A more complete description of this approach is given
elsewhere,25but specific relevant details are given here.
The measured analyzer and polarizer angles, A and P ,
respectively, are related to the complex electric fields of
incident and reflected light via the following equations:
6
a
/o:
/ 0
rp/r, = tan $eiA 0
0
.8

$ = A(10) (3) /: 0
a
a

+
A = 2P ( x / 2 ) (4) /
/
m a
a

/
The El's are the complex electric fields of the incident (+) ' a
and reflected (-) light beams of p and s polarization. The /
terms rp and rs are the standard Fresnel coefficients and
a
are functions of both the amplitude ratio tan $ and the
optical phase shift A. These latter parameters are calcu-
0 5 10 15 20 25
lated directly from the raw ellipsometry data. From the
theory mentioned above, r /rs can be related to the optical CARBON NUMBER
properties of the ample!^^^^^ With this approach it is Figure 2. Ellipsometry data showing the film thicknesses of
possible to determine a complex refractive index which n-alkanoic acids adsorbed on an oxidized aluminum substrate:
describes a multilayered sample as a pseudo-single-phase [RC02H]= O.oOo50 M in n-hexadecane,25 "C. Incubation times
are as follows: ( 0 )2, (m) 28, (0)
51, (V)73, (A)384 h. The dotted
material. In the case of a "clean" substrate, this provides line represents the best fit to the data in Figure 1for n-alkanoic
a reference value. For the same substrate, but with an acids of 12 carbons or greater.
overlayer of refractive index, iio = no (assuming no ab-
sorption term for the overlayer at the probe wavelength), calculating no for monolayer samples are such that it is not
the thickness and no in principle can be calculated by possible to routinely obtain reliable values of both no and
numerical iteration techniques. In practice, the errors in d. For this reason, and because the actual value of no for
a close-packed monolayer differs only slightly from the
(27) Aspnes, D. E. In 'Opical Properties of Solids";Seraphin, B. O., bulk material, we assign the bulk value to noand calculate
Ed.; North Holland Amsterdam, 1975. d. The errors inherent in this procedure are estimated to
48 Langmuir, Vol. I , No. I, 1985 Allara and Nuzzo
be no greater than -4~0.5 A for reasonable errors of
f0.05 in no (typical value -1.50).
- samples or solutions (see Experimental Section).
Figure 2 shows data that further indicate the importance
Film Thickness Measurements. Monolayer films of of kinetics in the formation of self-assembling monolayers.
linear carboxylic acids (prepared as described in the ex- In this series of experiments (similar to that above), the
perimental techniques section) were examined by ellipso- concentration of the fatty acids in hexadecane was de-
metry in order to explore qualitatively the effects of ex- creased to 5 X M, a 10-fold dilution. The substrates
perimental variables on the nature of the layers which for this experiment were all evaporated and prepared si-
form. Film thicknesses, d, are presented in Figures 1 and multaneously, and replicate runs of a given acid were in-
2 as plots of d vs. the number of carbons in the n-alkanoic cubated in the same solutions. Several features of the
acid. The values of d were calculated assuming a refractive results obtained deserve specific comment. First, consid-
index of 1.5 for the organic overlayer, a reasonable value erable variation in formation kinetics is seen for duplicate
for a hydrocarbon material with a density close to that of samples in the same solution. Second, the mean variation
the bulk solid. For films of these thicknesses, the exper- between samples decreases as they approach their limiting
imental errors were too large to permit an accurate cal- thickness. Third, limiting film thicknesses are found which
culation of n as an independent variable (see above). correspond exactly to those obtained at higher concen-
These errors notwithstanding, the data could be fit rea- tration. Fourth, the time needed to reach a maximum
sonably using a value of n = 1.50 f 0.05. The largest single packing density is now of the order of 2 weeks. Fifth,
source of variability in the ellipsometry measurements was multilayers are still seen for acids within the chain lengths
found to result from spurious contamination of the alu- below C12.
minum-aluminum oxide substrate during the determina- We note that while most of these data were obtained
tion of the “clean” substrate constants. In order to provide using samples pretreated with acetic acid, these results are
a uniform and displaceable initial overlayer, we have generally predictive of adsorption behavior on this oxide
preadsorbed acetic acid on most of the substrates exam- material. Acetate merely provides a uniform perturbation
ined. After determining the optical constants for these of substrate constants, thus allowing the results of different
coordinated surfaces, monolayers were prepared by experiments to be correlated and compared. The obser-
exchanging acetate with a longer n-alkanoic acid. The data vation of large variations in the kinetics of assembly in
given in Figures 1 and 2 are not corrected for the contri- both the presence and absence of acetate suggests im-
bution of the displaced acetate monolayer (see below). portant constraints on self-assembly other than the pres-
Control experiments indicate that this corection, in the ence of contaminating overlayers on the oxide support.
absence of other variables, would add an additional 2-3
A to the measured thicknesses. Control experiments also Studies of Surface-Exchange Reactions. In the
showed that over the course of time required for sample chemistry described above, the importance of ligand-ex-
incubation, oxide thickening almost exactly compensated change reactions in the formation of self-assembling
the apparent lost thickness due to acetate displacement. monolayers was suggested. The high sorption activity of
The effect of surface roughness on the thickness calcu- the native aluminum oxide surface virtually guarantees
lated from the ellipsometry measurements is expected to that some material from the laboratory ambient will be
be a small but systematic source of error in the data re- adsorbed on the substrate before it is immersed in the
ported. Scanning electron micrographs show that the n-alkanoic acid solution. These spurious ligands-water,
substrate oxide topography is comprised of shallow, rolling carbonate, air-borne organics, etc.-or even the acetate
ligands described above, can and do greatly influence the
-
hills of -50-200-A periodicity. Roughness of this order
( 1.2 to 1.3 times the geometric area) is expected to yield
a systematic variation of much less than 10% of the total
kinetics of closest packed monolayer formation. The sim-
ple notion of a surface formation constant, in analogy to
film thickness. transition-metal coordination chemistry in solution, re-
Taken together, the data presented in Figures 1and 2 quires that these ligands be displaceable by a long chain
demonstrate the importance of kinetics and competitive carboxylic acid of a given concentration if a good mono-
adsorption in the formation of self-assembling monolayers. layer is to be formed. To examine the dynamic charac-
At higher concentrations of fatty acids in hexadecane (-5 teristics of such surface coordination reactions and further
X M), the limiting film thicknesses described in Figure develop our understanding of the mechanismb) involved,
1were obtained. Such values usually required incubation we have conducted three independent labeling studies. In
times of the order of several days to be reached. I t is of the first, we have examined the displacement of tritiated
interest to note that the thickness of most films increased acetate ( [2-3H])from native aluminum oxide substrates
significantly (220%) after they had become autophobic by Cg, CIS,and Czon-alkanoic acids. In the second, X-ray
(i.e., had nonzero contact angles with hexadecane). It is photoelectron spectroscopy (XPS) was used to follow the
clear from the data that there exists a continuity of gross comparable displacement of trifluoroacetic acid by this
structure28and packing density for acids of chain length same series of acids. The third study, by reflection IR,
C12or greater. Below C12,there was a pronounced tend- examined the competitive formation of and displacement
ency for the films to form multilayers. The kinetics of from closest packed monolayers using protio and per-
multilayer formation were extremely variable-hence the deuterated stearic acids. Taken together, these results
range of values seen in Figure 1-and few samples ever allow strong inferences to be made as to the nature of the
were found to develop nonzero contact angles with hexa- monolayers formed and point out several complicated,
decane. Careful control experiments and analysis of the though little appreciated, characteristics of the substrate
acids by GC-mass spectroscopy did not suggest that this used. Each is discussed in turn.
behavior was related to any atypical impurities in the (a) Radioisotopic Labeling Studies. When aluminum
oxide substrates (which had been previously incubated in
a 0.005 M solution of [2-3H]aceticacid in ethanol) were
(28) As discussed above, the most significant sources of error are immersed in hexadecane solutions of Cg, C18,and Cz0 n-
systematic in nature. Thus, for organic overlayers of well-known bulk alkanoic acids (0.0050 M), significant quantities of the
refractive index, ellipsometry measurements, when carefully made, can
provide a simple but powerful method for determining the manner in tritiated label were displaced from the surface into the
which the apparent thickness of an adsorbate layer varies with structure. solution. this process was slow; after several days, a lim-
Spontaneously Organized Molecular Assemblies. 1 Langmuir, Vol. 1, No. 1, 1985 49

m *

*\

BINDING ENERGY (ev) 0.10


. .
0 20 40 60 80 100 120 (40
Figure 3. X-ray photoelectron spectra showing (a) the Cls core
level spectra which result from treating an oxidized aluminum TIME (hr)
substrate with a 0.050 M solution of CFBCOzHin ethanol at 25 Figure 4. First-order kinetic plot of the fractional coverage of
"C for 15 min and (b)this same sample after immersion in a 0.0050 the substrate surface by stearic-do acid. In this experiment, a
M solution of stearic acid in n-hexadecane. closest packed monolayer of the doacid has been placed in 0.0050
M solution of the perdeuterated (d,) acid in n-hexadecane at 25
iting activity was counted which corresponded to the av- "C.The two symbols (m, 0 ) represent replicate experiments on
erage displacement by the higher acid of 6.7 X 1014acetic one and three samples, respectively.
acid molecules per square centimeter of substrate surface
area. This value is larger than that expected for a closest interface may be an important process.29
packed monolayer (-5.1 X 1014molecules/cm2)6but is (c) Kinetics of Surface-Exchange Reactions in
remarkably close when this measured site density is cor- Closest Packed Monolayers. The data described above
rected for substrate surface roughness (the real surface area suggest the importance of kinetics in determining the
is -1.3 times the geometric value). As had been observed structural nature of the monolayers obtained by adsorp-
in the ellipsometry experiments, significant quantities of tion. There further exists a strong indication that tradi-
radiolabel, 520% of the total, were displaced by the C18 tional macroscopic indicators of monolayer "perfection",
and Czoacids after the samples had developed nonzero such as wetting, are not sufficiently sensitive to monitor
contact angles with the hexadecane solution. the approach to equilibrium, closest packed structures. To
Digestion of these films and their substrates (after further explore the dynamics of the formation of self-as-
washing with hexane) in concentrated aqueous sodium sembling monolayers, we have followed the exchange of
hydroxide yielded a surprising result. The aqueous solu- stearic acid for its perdeuterated analogue by reflection
tion thus produced contained an activity that corresponded infrared spectroscopy. These data were determined from
formally to an additional 7.2 X 1014acetic acid molecules the ratio of the integrated areas of the most intense bands
per square centimeter of substrate surface area being left in the CH and CD stretching-mode regions, respectively.
on the support. We will defer our discussion of this result The spectral characteristics in each of these regions were
to the following section and the companion paper to this found to be insensitive to the degree of coverage of each
study. isotopic component. These results provide an insight into
(b) XPS Study of the Displacement of Trifluoro- the reactive coordination chemistry relevant to this oxide
acetic Acid by Stearic Acid. Aluminum substrates at high packing densities.
which had been incubated in 0.05 M solutions of tri- Figure 4 shows a first-order plot of the fractional cov-
fluoroacetic acid in ethanol showed complex Cls core level erage of the surface by the protio (initial) acid as a function
spectra. At least four discrete peaks were observed (Figure of time. For the formal reaction described by eq 5, where
3), only two of which are assignable to the labeled marker
(BE = 293.7 and 290.6 eV). The large peak at 285.1 eV A + B*SE
kl
ASS+ B
k-1
(5)
was almost entirely absent on control samples (510% of
the intensity found here). Incubation in a 0.005 M hexa- A and B are the labeled and unlabeled exchange pair re-
decane solution of stearic acid for 5 days resulted in the spectively and S is the substrate, simple approach to
loss of most of the trifluoracetate marker. A new, intense equilibrium kinetics is predicted. In the range of con-
peak at 286.4 eV was observed. The large peak observed centrations relevant to this experiment ([A] = M,
initially a t 285.1 eV is still seen and shows little change
in absolute integrated intensity. These results, and those (29) For example, the intense peak seen in the XPS experiments at
obtained above using tritiated acetic acid, argue strongly a binding energy of 285.1 eV is shifted 1.3 eV to lower binding energy than
for the absorption and incorporation of small molecules the Cls core level emission for the stearic acid hydrocarbon tail (no
correction for charging). Such relative shifts generally indicate a highly
into the oxide overlayer. We further suspect that reduction reduced form of carbon. Angular-dependentXPS measurements are also
of these small organics at or near the metal-metal oxide consistent with the suggestions given in the text.
50 Langmuir, Vol. 1, No. 1, 1985 Allara and Nuzzo

ADSORBATE

CARBON NUMBER
Figure 5. Contact-angle data for n-alkanoic acid monolayers
adsorbed on an oxidized aluminum substrate [RC02H]= 0.0050
M in n-hexadecane,25 OC. The immersion times in the n-alkanoic
acid solutions were as follows: (a)15, (m) 72 h. The three wetting
liquids were from top to bottom water, methylene iodide, and
n-hexadecane.
kl[A] >> k-,[B]), this should reduce to pseudo-first-order
~~

2700 2800 2900 3000 3100


behavior. As shown in Figure 4, this clearly is not ob- WAVENUMBERS
served. These data, representing two independent runs
with four separate samples, suggest both nonideality and Figure 6. Infrared spectra of n-C19H&02H. The ordinate is
adsorption defined as -log (Illo)where Z is the intensity of de-
remarkable self-consistency. While the precision of the tected light for the sample and Io that of a reference. Lower:
data precludes a detailed analysis (the apparent two com- transmission spectrum of the bulk acid dispersed in KBr. Upper:
pleting first-order processes evidenced in the plot may be glancing angle reflection spectrum of a monolayer of the acid
an artifact), the general conclusion is clear. Simply stated, solution adsorbed onto an oxidized aluminum film. The features
exchange in closest packed monolayers on native aluminum at 2851 and 2919 cm-' are assigned to the C-H stretching modes
oxide substrates is not kinetically simple. of the CH2groups. The features at 2879 cm-l and those in the
2938-2966-cm-' region are assigned to the C-H stretching modes
(d) Formation of Mixed Monolayers of Stearic-do of the CH, groups.
Acid and Stearic-d36Acid. When an aluminum sub-
strate was incubated in a solution containing a mixture of terminated with an acetylene and vinyl functional group
deuterated and nondeuterated stearic acids in n-hexade- (see Experimental S e ~ t i o n )both
~ ' have nonzero contact
cane (2.5 X M in each), a distinct preference for the angles with hexadecane. The observed value of 36' for the
adsorption of the protio acid was observed by reflection vinyl-terminated derivative is remarkably close to that of
infrared spectroscopy. We estimate, based on a compar- the methyl-terminated acid of the same chain length.
ison of the spectra of pure do and d, monolayers to that Finally, the wetting data do not suggest a significant
obtained above, that the observed composition reflects an odd-even progression in surface tension for acids C12and
equilibrium ratio of do/d35 of the order of 1.4 f 20% at above. This is a likely consequence of the near-orthogonal
25 OC. orientation of the adsorbate to the surface (see below).
Wetting Characteristics of Adsorbed Monolayer
Films. The wetting characteristics of our monolayer films, Discussion
as judged by contact angle data, strongly parallel the classic The ellipsometry data in Figure 1, given the known
data in the literature on the properties of oleophobic optical constants of the fatty acid series, argue strongly
monolayer film^.^ In our studies, we have examined three that the acids of chain length CI2 or longer are at or near
liquids: water, methylene iodide, and n-hexadecane. The bulk packing density. This suggestion is fully supported
data in Figure 5 illustrate the influence of two experi- by independent radioisotopic labeling and
mental variables-immersion time and carbon chain our own results on the displacement of [2-3H]CH3C02H
length-on the wetting properties of n-alkanoic acids ad- from this substrate. Taken together, these data strongly
sorbed from n-hexadecane solutions (0.005 M). The times suggest the formation of highly oriented, closest packed
chosen (16- and 72-h incubations in the n-alkanoic acid monolayer structures. In Figure 6, we compare the infrared
solutions) correspond roughly to the times when consid- spectra, in the C-H stretching region, of arachidic acid
erable thickening was noted by ellipsometry under com- C
(), taken under two sets of experimentalconditions. The
parable experimental conditions. First, the contact angle lower portion of the figure, taken in transmission, shows
data does not seem to mirror this effect. Second, there the isotropic spectrum of the free acid dispersed in a KBr
does appear to be a discontinuity of behavior below CI2, pellet. The upper portion shows a glancing angle reflection
a feature similar to that seen in the ellipsometry experi- spectrum of the limiting (Le., highest coverage seen by
ments. Third, for carboxylic acids with chain lengths XlZ, ellipsometry) monolayer of this same acid.33 It is imme-
materials that consistently yield oleophobic films, some diately apparent, even on a qualitative level, that these
variation is noted. In general, across this series the average
contact angles increase slightly and reach limiting values,
-
depending on the wetting liquid, at carbon chain lengths
of 18-19. We note, with some interest, that the Czzacids
(31) These materials are discussed in detail: Allara, D. L.; Nuzzo, R.
G. Langmuir, following paper in this issue.
(32) Whitesides, G . M., unpublished results.
(33) The accompanying paper describes the vibrational spectroscopy
(30) For leading references, see: "Advances in Chemistry Series 43"; of n-alkanoic acid monolayers, especially as it relates to the determination
Gould, R. F., Ed.; American Chemical Society: Washington, D.C., 1964. of structure and orientation, in depth.
Spontaneously Organized Molecular Assemblies. 1 Langmuir, Vol. 1, No. 1, 1985 51

spectra show significant differences in relative band in- surface, we feel, make it an extremely difficult material
tensity. The symmetric and asymmetric methylene to handle in the laboratory ambient. It was our experience
stretching modes (2851 and 2919 cm-l, respectively) are, that some form of reversible chemical passivation-in this
as expected, the predominant bands in the isotropic case adsorption of acetic acid-was absolutely necessary
spectrum. In the monolayer spectrum, these features are to ensure that closest packed monolayers would be ob-
lower in intensity than those assigned to the methyl C-H tained in reasonable lengths of time and that the data from
stretching modes, seen here at 2879 and 2938-2966 cm-1.33 such simple probes as ellipsometry could be routinely
As we shall show in the accompanying paper, the alkyl tails compared.
in these acids are best modeled as being at, full chain ex- Implicit in the above discussion is the most significant
tension, showing little tilt from the surface normal. In this aspect of this study as regards the nature of adsorbed
context then, we discuss the data presented above. carboxylic acid monolayers on native aluminum oxide
One of the most striking features of this study was the overlayers. Simply stated, these are dynamic structures;
observation of a minimum chain length for the formation ligands, once adsorbed, can be displaced by other materials
of well-oriented, closest packed structures. As evidenced in solution and presumably in the laboratory ambient as
by the ellipsometry and wetting data, closest packed, low well. The observed exchange of stearic acid by its per-
surface free energy phases were obtained only for chain deuterated analogue deserves specific comment in this
lengths greater than ell. The reasons underlying this regard. First, the exchange reaction does not appear to
observation are not immediately evident. Careful control require that the exchanging ligand have a much higher
experiments did not reveal any obvious or atypical im- formation constant for coordination to the surface. In this
purities in these lower carbon number preparations. It is system, where the ligands are structurally similar, the
significant to note that, even above C12,the contact angle process is driven simply by mass action. Qualitative ob-
data shows a pronounced trend toward higher values, in- servations suggest that this is a general property of the
dicating the formation of “denser’! (presumably lower system for most if not all pairings of the linear carboxylic
defect density) structures, as one adsorbs chains of in- acids studied; longer acids will displace shorter ones and
creasing length. Thus, it would appear that the absolute visa versa as long as the concentrations of the displacing
magnitude of the enthalpy associated with chain-chain ligand in solution are high enough. Second, the exchange
packing interactions is an important effect, a conclusion process is not kinetically simple. As we have described
that also can be strongly inferred from the Langmuir- above, a simple analysis predicts that the kinetics of the
Blodgett literature.ls The observation of nonstatistical approach to equilibrium should exhibit first-order be-
selection of stearic-do acid over stearic-d,, acid further havior. This clearly is not consistent with the observed
illustrates this point. If the structure of the adsorbed layer data.36 If we assume that the observed exchange kinetics
were determined solely by the interaction of the head truly reflect two competing first-order processes, two
group with the substrate (as a result of electrostatic in- reasonable explanations come to mind. First, the exchange
teractions or a preferred surface bonding site), it is hard mechanism may reflect differing rates related to a heter-
to imagine how either chain length or isotopic substitu- ogeneous distribution of bonding sites on the support (that
t i ~ could
n ~ ~provide a significant perturbation to gross is, some sites are “stickier” than others). A second pos-
structure.35 Our feeling, at this time, is that the chain- sibility is that the exchange mechanism is dominated by
length discontinuity discussed above reflects in part the surface or monolayer defects. Exchange of material at or
complex balance between the head-group bonding and near these defects is expected to be fast. The diffusion
tail-packing interactions, forces which we believe tend to of material further removed from these selective desorption
expand and contract the layer, respectively. sites might be slow, and thus kinetically distinguishable,
The formation kinetics we have observed illustrates at these high surface coverages. We cannot, at present,
another interesting and technically important feature of distinguish these and other possible mechanisms. Our
self-assembly. First and most significant is that these inclination, given the complicated microscopic topography
kinetics can be extremely slow, viz., on the order of days. of the substrate, is to conclude that structural defects must
The tremendous inconsistency seen for similar samples in play an important mechanistic role in the exchange reac-
the same solutions (Figure 2) argues strongly for an im- tions d e s ~ r i b e d . ~ ’ ~ ~ ~
portant role for surface and/or monolayer defects, the
nature of which we are not able to determine. Our
qualitative observation is that the slowest and most un- (36) Attempts to fit the data to second- and zero-order rate depen-
predictable part of the assembly occurs during the incor- dences yielded equally poor correlations. Analyses based on a postulated
rate-determining dissociation of the carboxylic acid dimer in hexadecane
poration of the last 20-25% of the monolayer. These were found to be incompatible with the kinetic data.
findings must also be viewed in the context of ligand-ex- (37) Assuming first-order rate behavior and an Arrhenius preexpo-
change chemistry. While most of our work involved the nential factor of 1013s-1,the activation energy for exchange in the limiting
slow rate regime is calculated to be -15 kcal/mol, a value consistent with
preadsorption of acetate (see above), the final results we selective ’displacement” a t defects.
obtain, in terms of packing and structure, in no way de- (38) The work of Torney and McConnell (Torney, D. C.; McConnell,
pend upon this added step. Rather, it was noticed that H. M. Proc. R. SOC.London, Ser. A 1983, 387, 147-170) suggests an
intriguing mechanistic possibility which, unfortunately, the precision of
the kinetics of assembly became much more erratic in the our data prevents us from explicitly testing. If, in fact, exchange a t
absence of a specific preadsorbed monolayer. The highly defects were preferred, then the observed rate of exchange may be dom-
active sorption characteristics of the clean native oxide inated both by the density of these defects as well as the rate of diffusion
of the adsorbate along the surface to these same sites. Processes of this
sort,involving mass-transport-limited rates in two dimensions, need not
(34) Isotope effects on crystal packing are well documented in the be characterized by a time-independent rate constant. Thus, the data
polymer literature. For example, deuterated polyethylene does not co- do not necessarily have to reflect complex, multiprocess kinetics. The
crystallize with ita protio analogue (Schelten, J.; Wignall, G. D.; Ballard, fact that diffusion in two-dimensional organic assemblies can be ex-
D. G. H.; Longman, G. W. Polymer 1977,18,1111-1120). We have not tremely varied-behaviors ranging from liquid- to solidlike-has been
seen any evidence that militates either for or against phase separation broadly shown, for example, in lipid multibilayers and monolayers. See:
in these mixed monolayers. Seul, M.; Weis, R. M.; McConnell, H. M. Biophys. J. l983,41,212a. Wu,
(35) One experimental observation we have made offers a perplexing E.; Jacobson, K.; Papahadjopoulos, D. Biochemistry 1977,16,3936-3941.
and incompletely understood contradiction to this suggestion. Hexa- Smith, B. A.; McConnell, H. M. Proc. Nutl. Acud. Sci. U.S.A. 1978, 75,
noic-dll acid consistently yielded oleophobic monolayers, a property 2759-2763. Suel, M.; Eisenberger, P.; McConnell, H. M. Proc. Nutl. Acud.
rarely observed for any acid below CI2. Sci. U.S.A. 1983, 80, 5795-5797 and references cited therein.
52 Langmuir 1985,1, 52-66
Taken together, the data show that closest packed, or- portant limitation to closest-packed, self-assembly on this
iented monolayers of n-alkanoic acids can be formed by substrate which is related directly to the length of the alkyl
adsorption from dilute solution. The formation of these chain. The data further show that oleophobicity is an
quasi-two-dimensional organic surface phases is charac- insufficiently sensitive probe to determine the formation
terized by complicated kinetics in which surface and/or of equilibrium closest packed structures. In a companion
monolayer defects, as well as impurities, seemingly play paper we describe these structures and their relationship
important roles. The structures so obtained are dynamic to many of the above observations in detail.
in nature in that they undergo rapid exchange with com-
parable ligands in solution. These results, and those Registry No. n-Cl9H&O2H, 506-30-9; aluminum oxide,
presented in the following paper, further suggest an im- 1344-28-1; stearic acid, 57-11-4; hydrogen, 1333-74-0.

Spontaneously Organized Molecular Assemblies. 2.


Quantitative Infrared Spectroscopic Determination of
Equilibrium Structures of Solution-Adsorbed n -Alkanoic
Acids on an Oxidized Aluminum Surface
David L. Allara*
Bell Communications Research, Murray Hill, New Jersey 07974
Ralph G. Nuzzo*
AT&T Bell Laboratories, Murray Hill, New Jersey 07974
Received August 16, 1984

Infrared reflection spectroscopy has been applied to the determination of the structures of adsorbed
monolayer films of n-alkanoic acids on oxidized aluminum substrates. Acids of 16-22 carbons, terminated
by methyl, vinyl, or propargyl groups, were absorbed from hexadecane solution at 25 O C , using immersion
times up to several days. These results,together with those for perdeuterated acids, indicate that close-packed
-
assemblies are formed with extended alkyl tails oriented with their chain axes tilted away from normal
to the surface with a value of 10” for the longer chains. Frequency shifts to higher values are observed
for CH stretching modes of the vinyl and propargyl terminal groups measured in monolayer films compared
to the bulk acids. These shifts are of similar magnitude to gas-liquid phase shifts for these groups and
imply the environment of the ambient-monolayer interface exhibits very diminished intermolecular in-
teractions for these terminal groups compared to the bulk. The spectra clearly show that chemisorption
occurs by proton dissociationto form carboxylatespecies. Examination of both line width and peak positions
of the carboxylate stretching modes suggests a variety of binding geometries of the carboxylate groups
to the surface exists. These results lend support to the general concept of forming stable, oriented and
ordered two-dimensional organic films by solution adsorption and show the utility of applying surface
vibrational spectroscopy to structural determination in these films.

1. Introduction cated molecular organic films. A very elementary problem


The structures, formation, and properties of organic is that most popular surface techniques require impinging
molecular monolayer films are the subjects of much current ions or electrons as probes and this usually causes damage
interest. One system of recent particular interest to us has to organic molecules. The techniques that have been
been spontaneously adsorbed n-alkanoic acid monolayer generally found useful for organic structures are those
films on ambient metal surfaces. In a previous paper’ we using photon probes, in particular, X-ray photoelectron,
presented extensive results on the formation, properties, infrared, and Wan spectroscopies. Additionally, inelastic
and dynamics of solution-adsorbed n-alkanoic acid films electron tunneling and electron energy loss spectroscopies
on oxidized aluminum substrates. Central to a compre- have had success for organic monolayers. Past studies of
hensive understanding of those results is the development the adsorption of medium- to long-chain n-alkanoic acids
of the details of molecular structure. Such detail includes on ambient metal surfaces have utilized electron diffrac-
the nature of the individual molecular species, e.g., ion t i ~ n ,inelastic
~ - ~ electron tunneling (IETS),
formation, surface orientations, and surface-adsorbate
bonding, and the nature of the total monolayer assembly,
(2) A review of earlier work can be found in: Bowden, F. P.; Tabor,
especially as regards ordered packing arrangements. While D. “The Friction and Lubrication of Solids”; Oxford University Press:
a number of surface analysis tools are now available due London, 1968; Part 11, Chapter 19.
to extensive developments in the surface physics commu- (3) Menter, J. W.; Tabor, D. Proc. R . SOC.London, Ser. A 1957,204,
514-524.
nity, in actual fact, few are applicable to the analysis of (4) Chapman, J. A,; Tabor, D. Proc. R. SOC.London, Ser. A 1957,242,
the subtle features of the chemical structures of compli- 96-107.
(5) Brockway, L. 0.; Karle, J. J. Colloid Sci. 1947,2, 277-287.
(6) Cass, D. A.; Straws, H. L.; Hansma, P. K. Science (Washington,
(1) A h a , D. L.; Nuzzo, R. G. Langmuir, preceding paper in this issue. D.C.)1976, 192, 1128-1130.

0743-7463/85/2401-0052$01.50/0 0 1985 American Chemical Society

You might also like