Вы находитесь на странице: 1из 16

Viewpoints: Chemists on Chemistry

The Flexible Surface


The Flexible Surface:
Molecular Studies Explain the
Extraordinary Diversity of Surface
Chemical Properties

About Viewpoints

Outline

Historical Perspective
External Surfaces
Surface Concentration
Clusters and Small Particles
Thin Films

Internal SurfacesMicroporous Solids


Clean Surfaces
Interfaces
Adsorption

Techniques of Surface Science


Phenomena Discovered by Molecular
Surface Chemistry
Surface Structure Is Different from Bulk Structure
Adsorbate-Induced Restructuring of Surfaces
Rough Surfaces Do Chemistry
Clusterlike Bonding of Adsorbed Molecules
The Flexible Surface

Technological Impact of Molecular Surface


Chemistry
Future Directions in Surface Chemistry
Coadsorption on Surfaces
High-Pressure Surface Science
Monitoring Surface Chemistry at Ever-Improving
Spatial Resolution and Time Resolution
Studies of the Buried interfaces, SolidLiquid
and SolidSolid
Achieving 100% Selectivity in Surface
Reactionsthe Environmental Imperative
Nanoparticles: Surfaces in Three Dimensions

Other Material on
Surface Chemistry in This Issue
On the Surface
Mini-Activities Exploring Surface
Phenomena: JCE Classroom Activity #6
Flying over Atoms CD-ROM
JCE Software Special Issue 19,
by John R. Markham

Gabor A. Somorjai and Gnther Rupprechter

162

176A

iewpoints is a major feature of the celebration of the Journal


of Chemical Educations 75th year. It is being supported by The
Camille and Henry Dreyfus Foundation, Inc., which recently celebrated its own 50th anniversary. Each paper in the Viewpoints series will be written by a chemist or group of chemists with special
expertise in a particular field, with the aim of providing an overview
of that fields accomplishments, importance, and prospects. The goal
is to reflect on developments during the past 50 years and to predict
how each field will evolve over the next 25 years. The total perspective encompassed by Viewpoints corresponds with the 75 years of
this Journals lifetime and reflects its comprehensive interest in all of
chemistry. The 50-year retrospective view of each field corresponds
with the period during which the Camille and Henry Dreyfus Foundation has been supporting the chemical sciences.
Authors of Viewpoints papers will provide perspectives on what
chemistry has done during the lifetime of the Dreyfus Foundation
and to set the stage for what chemistry will become well into the
next century. The papers will be written at a level appropriate for
upper-division undergraduate chemistry students and will extend and
enhance the Journals role as, in the words of an early editor, a living textbook of chemistry. In addition, they will be published in
electronic format via JCE Online (whose founding was also supported
by the Dreyfus Foundation). In the Journal, Viewpoints papers will
take full advantage of color graphics, which will also appear in the
electronic version. In JCE Online there also will be links to the authors and other related Web sites, and video and animations when
relevant.
The Viewpoints series begins this month with The Flexible Surface, a paper from the laboratories of Gabor A. Somorjai of the University of California, Berkeley. Somorjai and his co-author, Gnther
Rupprechter, discuss concepts related to external and internal surfaces and interfaces and the dynamic nature of surfaces during chemical processes. A broad overview of the most frequently used surface
science techniques is also included, along with references for each
technique. Somorjai and Rupprechter close with a discussion of the
present and future technological impact of surface chemistry on catalysis and semiconductor devices, and the chemists role in surface
science in the coming years.
Glenn T. Seaborg, Chair of the Viewpoints Editorial Board

247

JChemEd.chem.wisc.edu Vol. 75 No. 2 February 1998 Journal of Chemical Education

161

Viewpoints

The Flexible Surface: Molecular Studies Explain


the Extraordinary Diversity of Surface Chemical Properties
Gabor A. Somorjai and Gnther Rupprechter
Department of Chemistry, University of California at Berkeley, and Materials Sciences Division, E. O. Lawrence Berkeley
National Laboratory, Berkeley, CA 94720; http://www.cchem.berkeley.edu/~gasgrp

Historical Perspective
Surface science in general and surface chemistry in particular have a long and distinguished history. The spontaneous spreading of oil on water was described in ancient times
and was studied by Benjamin Franklin. The use of surfacechemical processes on a large industrial scale began in the
early part of the 19th century. The application of catalysis
started with the discovery of the platinum-surface-catalyzed
reaction of H2 and O2 in 1823 by Dbereiner. Dbereiner
used this reaction in his portable flame source, of which he
sold a large number. By 1835 the discovery of heterogeneous
catalysis was complete, thanks to the studies of Kirchhoff,
Davy, Henry, Philips, Faraday, and Berzelius (1). It was about
this time that the Daguerre process was introduced for photography. The study of tribology, which includes friction, lubrication, and adhesion, also started around this time; this coincides with the industrial revolution, as machinery with moving parts became prevalent (although some level of understanding of friction appears in the work of Leonardo da Vinci).
Surface-catalyzed-chemistry-based technologies first appeared
in the period of 1860 to 1912, starting with the Deacon process (2HCl + 1/2 O2 H2O + Cl2), SO2 oxidation to SO3
(Messel, 1875), the reaction of methane with steam to produce
CO and H2 (Mond, 1888), ammonia oxidation (Ostwald,
1901), ethylene hydrogenation (Sabatier, 1902), and ammonia
synthesis (Haber, Mittasch, 19051912). Surface tension
measurements and recognition of equilibrium constraints on
surface chemical processes led to the development of the thermodynamics of surface phases by Gibbs (1877).
The existence of polyatomic or polymolecular aggregates
that lack crystallinity and diffuse slowly (gelatine and albumin, for example) was described in 1861 by Graham, who
called these systems colloids. Polymolecular aggregates that
exhibit internal structure were called micelles by Nageli,
and stable metal colloids were prepared by Faraday. The colloid
subfield of surface chemistry gained prominence in the beginning of the 20th century with the rise of the paint industry
and the preparation of artificial rubbers. Studies of high-surfacearea gas-absorber materials for gas masks and other gas-separation
technologies and investigations of the lifetime of the light bulb
filament led to the determination of the dissociation probability and adsorption probability of many diatomic molecules
on surfaces as a function of gas pressure (adsorption isotherm)
and temperature (Langmuir 1915). The properties of chemisorbed and physisorbed monolayers, adsorption isotherms,
dissociative adsorption, energy exchange, and sticking upon
gassurface collisions were studied. Studies of electrode surfaces in electrochemistry led to detection of the surface space
charge (2). Surface diffraction of low-energy electrons was
discovered by Davisson and Germer in 1927, and atom diffraction (helium) from surfaces, somewhat later. Major aca162

demic and industrial laboratories focusing on surface studies


have been formed in Germany (Haber, Polanyi, Farkas,
Bonhoefer), the United Kingdom (Rideal, Roberts, Bowden),
the United States (Langmuir, Emmet, Harkins, Taylor, Ipatief,
Adams), and many other countries. They have helped to bring
surface chemistry into the center of development of chemistry,
both because of the intellectual challenge to understand the
rich diversity of surface phenomena and because of its importance in chemical and energy conversion technologies.
Up to the 1950s, studies of surfaces were mostly on the
macroscopic scale. Then the rise of the solid-state-devicebased electronics industry and the availability of economical
ultra-high-vacuum systemsdeveloped by research in space
sciencesprovided surface chemistry with new challenges
and opportunities and resulted in explosive growth of the discipline. Clean surfaces of single crystals could be studied for
the first time and the development of a large number of new
techniques (cf. section on Techniques of Surface Science) from
the 1960s onwards made possible the investigation of surfaces
at atomic and molecular levels.
As a result, macroscopic surface phenomena (adsorption,
bonding, catalysis, oxidation, and other surface reactions; diffusion, desorption, melting, and other phase transformations;
growth, nucleation, charge transport; atom, ion, and electron scattering; friction, hardness, lubrication) are being reexamined on the molecular scale. This has led to a remarkable growth of surface chemistry that has continued uninterrupted up to the present. The discipline has again become
one of the frontier areas of chemistry. The newly gained
knowledge of the molecular ingredients of surface phenomena
has given birth to a steady stream of high technology products, including new hard coatings that passivate surfaces;
chemically treated glass, semiconductor, metal, and polymer
surfaces to which the treatment imparts unique surface properties; newly designed catalysts, chemical sensors, and carbon
fiber composites; surface-space-charge-based copying; and
new methods of electrical, magnetic, and optical signal processing and storage. Molecular surface chemistry is being utilized increasingly in biological sciences.
External Surfaces

Surface Concentration
The concentration of atoms or molecules at the surface
of a solid or liquid can be estimated from the bulk density.
For a bulk density of 1 g/cm3 (such as ice or water), the molecular density , in units of molecules per cubic centimeter,
is 5 1022. The surface concentration of molecules (molecules/cm2) is proportional to 2/3, assuming cubelike packing, and is thus on the order of 1015 molecules/cm2. Because
the densities of most solids or liquids are all within a factor
of 10 or so of each other, 1015 molecules/cm2 is a good order-

Journal of Chemical Education Vol. 75 No. 2 February 1998 JChemEd.chem.wisc.edu

The Flexible Surface

of-magnitude estimate of the surface concentration of atoms


or molecules for most solids or liquids. Of course, surface
atom concentration of crystalline solids may vary by a factor
of two or three, depending on the type of packing of atoms
at a particular crystal face.

Clusters and Small Particles


All atoms in a three- or four-atom cluster are by necessity
surface atoms. As a cluster grows in size, some atoms may
become completely surrounded by neighboring atoms and
are thus no longer on the surface (Fig. 1). We frequently
describe a particle of finite size by its dispersion D, where D
is the ratio of the number of surface atoms to the total number
of atoms:
D = number of surface atoms
total number of atoms

For very small particles, D is unity. As the particle grows


and some atoms become surrounded by their neighbors, the
dispersion decreases (Fig. 1). Of course, D also depends somewhat on the shape of the particles and how the atoms are
packed. The dispersion is already as low as 103 for particles of
10-nm (100-) radius.
Many chemical reactions are facilitated by surface atoms
of heterogeneous catalysts. These catalysts increase the rate of
formation of product molecules and modify the relative dis-

Figure 1. Clusters of atoms with cubic packing having 8, 27, 64,


125, and 216 atoms. While in an 8-atom cluster all of the atoms
are on the surface, the dispersion rapidly declines with increasing
cluster size, as shown in the lower part of the figure.

tribution of products. Most catalysts are in small-particle


form, including those used to produce fuels and chemicals
ranging from high-octane gasoline to polyethylene.

Thin Films
Consider a monolayer of gold atoms (a layer of gold atoms
one atom thick) deposited on iron (Fig. 2). This film has a
dispersion of unity, since all the atoms are on the surface. About
50 layers of gold atoms (D = 1/50) are needed to obtain the
optical properties that impart the familiar yellow color characteristic of bulk gold.
Thin films are of great importance to many real-world
problems. Their material costs are very little compared to cost
of the bulk material, and they perform the same function
when it comes to surface processes. For example, a monolayer
of rhodium (a very expensive metal), which contains only
about 1015 metal atoms per square centimeter, can catalyze
the reduction of NO to N2 by its reaction with CO in the
catalytic converter of an automobile, or it can catalyze the
conversion of methanol to acetic acid by the insertion of a
CO molecule.
Thin ordered silicon layers optimize electron transport
in integrated electronic circuits and thin films of organic
molecules lubricate our skin or the moving parts of internal
combustion engines. A green leaf is a high-surface-area system
designed to maximize the absorption of sunlight in order to
carry out chlorophyll-catalyzed photosynthesis at optimum rates.
Often the surface of a thin film is roughened deliberately. Automobile brake pads are designed to optimize the
desired mechanical properties of surfaces in this way, as is
the corrugated design of rubber soles of tennis shoes. The
large number of folds of the human brain helps to maximize
the number of surface sites, which also facilitate charge transport and transport of molecules. These are some examples
that show how external surfaces are used in nature. External
surfaces are a key element of technology, ranging from catalysts and passivating coatings to computer-integrated circuitry
and the storage and retrieval of information.

Figure 2. An iron particle with one surface covered with a monolayer of gold atoms. When it comes to surface properties such as
adsorption or catalysis, one monolayer of atoms is all that is needed
to carry out the necessary chemistry.

JChemEd.chem.wisc.edu Vol. 75 No. 2 February 1998 Journal of Chemical Education

163

Viewpoints

Internal SurfacesMicroporous Solids


Microporous solids are materials that are full of pores of
molecular dimensions or larger. These materials have very
large internal surface areas. Many clays have layer structures
that can accommodate molecules between the layers by a process called intercalation. Graphite will swell with water vapor
to several times its original thickness as water molecules become incorporated between the graphitic layers. Crystalline
alumina silicates, often called zeolites, have ordered cages of
molecular dimensions (3), where molecules can adsorb or
undergo chemical reactions (Fig. 3). These materials are also
called molecular sieves, because they may preferentially adsorb
certain molecules according to their size or polarizability. This
property is of great commercial importance and may be used
to separate mixtures of gases (air) or liquids or to carry out
selective chemical reactions. Bones of mammals are made out
of calcium apatite, which has a highly porous structure, with
pores on the order of 10 nm (100 ) in diameter. Coal and
char have porous structures, with pore diameters on the order
of 102103 nm (103104 ). These materials have very large
internal surface areas, in the range of 100400 m2 per gram
of solid. As this short survey shows, nature has provided us
with many useful microporous materials; and many synthetic
microporous substances are used in technology, both to separate gas and liquid mixtures by selective adsorption and to
carry out surface reactions selectively in their pores, which
are often of molecular dimensions. Because surface reaction
rate (product molecules formed per second) is proportional
to surface area, materials with high internal surface areas carry
out surface reactions at very high rates.
Clean Surfaces
To study atomically clean surfaces, we must work under
so-called ultrahigh vacuum (UHV) conditions (4, 5), as the following rough calculation shows. We know that the concentration of atoms on the surface of a solid is on the order of 1015
cm2. To keep the surface clean for 1 s or for 1 h, the flux of
molecules incident on the initially clean surface must therefore
be less than 1015 molecules/cm2/s or 1012 molecules/cm2/s,
respectively. From the kinetic theory of gases (6), the flux, F, of
molecules striking the surface of unit area at a given ambient
pressure P is
N AP
F=
2MRT
or
20

F (atoms/ cm2 s) =

2.63 10

P (Pa)

M (g / mol) T

or
22

F (atoms/ cm s) =

3.51 10

P (torr)

M (g / mol) T

where M is the average molar weight of the gaseous species,


T is the temperature (in Kelvin), R is the gas constant, and
NA is Avogadros number. Substituting P = 4 104 Pa (3

164

Figure 3. Alumina silicates with pores of molecular dimensions


(zeolites) are used as selective absorbers of gases or liquids (molecular sieves) and as catalysts in chemical and petroleum technologies. The figure shows a synthetic zeolite, zeolite A. The
red spheres represent oxygen atoms and the yellow spheres represent either silicon or aluminum atoms. For each aluminum there
is a corresponding Na+ ion somewhere in one of the open channels. The molecular formula of this molecular sieve is
Na12(Al 12Si 12O 48) 27H 2O.

106 torr) and using the values M = 28 g/mol (for N2) and T
= 300 K, we obtain F 1015 molecules/cm2/s. Thus, at this
pressure the surface is covered with a monolayer of gas within
seconds, assuming that each incident gas molecule sticks.
For this reason the unit of gas exposure is 1.33 104 Pa-s
(106 torr-s), which is called the Langmuir (L). Thus, a 1-L
exposure will cover a surface with a monolayer amount of
gas molecules, assuming a sticking coefficient of unity. At
pressures on the order of 1.33 107 Pa (109 torr), it may
take 103 s before a surface is covered completely.
In practice, one usually wants to study a surface without worrying about contamination from ambient gases. Current surface science techniques can easily detect contamination on the order of 1% of a monolayer. This then will be
our operational definition of clean. Thus, ultrahigh vacuum
conditions (< 1.33 107 Pa = 109 torr) are required to maintain a clean surface for about 1 h, the time usually needed to
perform experiments on clean surfaces.
Interfaces
In most circumstances, however, and certainly in Earths
environment, surfaces are continually exposed to gases or liquids or placed in contact with other solids. As a result, we
end up investigating the properties of interfacesthat is, between solids and gases, between solids and liquids, between
solids and solids, and even between two immiscible liquids.
Thus, unless specifically prepared otherwise, surfaces are always covered with a layer of atoms or molecules from the
environment.

Journal of Chemical Education Vol. 75 No. 2 February 1998 JChemEd.chem.wisc.edu

The Flexible Surface

Adsorption
On approaching the surface, each atom or molecule encounters an attractive potential that ultimately will bind it
to the surface under proper circumstances. The process that
involves trapping of atoms or molecules incident on the surface is called adsorption. It is always an exothermic process.
For historical reasons, the heat of adsorption Hads is always
denoted as having a positive signunlike the enthalpy H,
which for an exothermic process would be negative according
to the usual thermodynamic convention. The residence time
of an adsorbed atom (7) is given by
= 0 exp(Hads/RT)

where 0 is correlated with the surface-atom vibration times


(it is frequently on the order of 1012 s), T is the temperature, and R is the gas constant. The value of can be 1 s or
longer at 300 K for Hads > 69 k J/mol (16.5 kcal/mol). The
surface concentration (in molecules/cm2) of adsorbed molecules on an initially clean surface is given by the product of
the incident flux F and the residence time :
= F

The surface of the material on which adsorption occurs


is often called the substrate. Substrate-adsorbate bonds are
usually stronger than the bonds between adsorbed molecules.
As a result, the monolayer of adsorbate bonded to the substrate is held most tenaciously and is difficult to remove.
Therefore, the properties of real surfaces are usually determined in the presence of an adsorbed monolayer.

Figure 4. Schematic diagram of a scanning tunneling microscope


capable of operating in a pressure range from UHV to 1 atmosphere. The STM is located inside a high-pressure chemical reactor, which is attached to a UHV surface characterization chamber.
The two sections are separated by a gate valve. For surface cleaning and analysis, a transfer system is used to move the sample
from the high-pressure cell to the UHV part of the apparatus (and
vice versa).

Techniques of Surface Science


Over the last three decades, a large number of techniques
have been developed to study various surface properties, including structure, composition, oxidation states, and changes
of chemical, electronic, and mechanical properties. The emphasis has been on surface probes that monitor properties
on the molecular level and are sensitive enough to detect eversmaller numbers of surface atoms. The frontiers of surface
instrumentation are constantly being pushed toward detection of finer detail: atomic spatial resolution, ever-smaller
energy resolution, and shorter time scales. Because no single
technique provides all necessary information about surface
atoms, the tendency is to use a combination of techniques.
The most commonly used techniques (Table 1) involve the
scattering, absorption, or emission of photons, electrons, atoms and ions, although some important surface-analysis techniques cannot be classified this way.
Most surface probes require high vacuum during their
application, which prevents their use during high-pressure
studies. To circumvent this restriction, UHV-compatible
high-pressure cells (environmental cells) were developed (8).
The sample to be analyzed is first subjected to the usual highpressure and/or high-temperature conditions encountered
during reactions in the environmental cell. Afterwards the
sample is transferred, without exposure to air, into the evacuated UHV chamber where the surface probe is located for
subsequent surface analysis (generally the sample surface
should be characterized before and after any treatment).
During the past five years, two new surface science techniques in particular proved capable of obtaining molecularlevel surface information during chemical change at both low
and high ambient pressures (9, 10): scanning tunneling microscopy (STM) and infraredvisible sum frequency generation
(SFG) surface vibrational spectroscopy. Both of these techniques
can operate within a 14-order-of-magnitude pressure range
(1010104 torr) without significant change in signal quality
in terms of spatial or energy resolution. Using these two techniques, we can monitor both substrate and adsorbate structures
during reactions at high pressures. One such apparatus,
designed for in situ STM, is shown in Figure 4.
Sample preparation is always an important part of surface
studies. Single crystals are oriented by X-ray back-diffraction,
cut, and polished. They are then ion-bombarded or chemically treated to remove undesirable impurities from their surfaces. Thin films are deposited from vapor by sublimation,
sputtering, or the use of plasma-assisted chemical vapor deposition. Materials of high internal surface area are prepared
from a sol-gel or by calcination at high temperatures. The
genesis and environmental history of the surface is primarily
responsible for its structure and composition and must always be carefully monitored.
Table 1 lists a selection of the surface-science techniques
used most frequently in recent years to learn about the interface on the atomic scale. The names of the techniques, their
acronyms, and brief descriptions are provided (along with
some references [1139], if a more detailed study of the capabilities and limitations of a particular technique is desired).
We also indicate the primary surface information that can
be obtained by the application of each technique.

JChemEd.chem.wisc.edu Vol. 75 No. 2 February 1998 Journal of Chemical Education

165

Viewpoints

Table 1. Selection of Most Frequently Used Surface Science Techniques


Primary Surface
Information

Acronym

Name (References)

Description

Adsorption or selective
chemisorption (11)

Atoms or molecules are physisorbed; the amount of gas adsorbed is a


measure of total surface area. Chemisorption of atoms or molecules yields
surface concentration of selected elements or adsorption sites.

Surface area, adsorption


site concentration

AD

Atom or helium diffraction


(12 )

Monoenergetic beams of atoms are scattered from ordered surfaces and


detected as a function of scattering angle, giving structural information on
the outermost layer of the surface. Method is extremely sensitive to surface
ordering and defects.

Surface structure

AES

Auger electron
spectroscopy (12 14 , 23,
24, 26 )

Core-hole excitations are created, usually by 1 10-keV incident electrons,


and Auger electrons of characteristic energies are emitted through a
2-electron process as excited atoms decay to their ground state. AES gives
information on the near-surface chemical composition.

Chemical composition

AFM

Atomic force microscopy


(15, 16 )

Similar to STM. An extremely delicate mechanical probe is used to scan


the topography of a surface by measuring forces exerted by surface atoms.
Light interference is used to measure the deflection of the mechanical
surface probe. This is designed to provide STM-type images of insulating
surfaces or to detect mechanical properties at the molecular level.

Surface structure

ELS or
EELS

Electron energy loss


spectroscopy (14, 17 19 )

Monoenergetic electrons ~5 50 eV are scattered off a surface and the


energy losses are measured. This gives information on the electronic
excitations of the surface and the adsorbed molecules (see HREELS).

Electronic structure,
surface structure

ESCA

Electron spectroscopy for


chemical analysis (12, 20 )

Now generally called XPS (see XPS).

Composition, oxidation
state

EXAFS

Extended X-ray absorption


fine structure (21, 22 )

Monoenergetic photons excite a core hole. Modulation of the absorption


cross-section with energy at 100 500 eV above the excitation threshold
yields information on radial distances to neighboring atoms. The cross
section can be monitored by fluorescence as core holes decay or by the
attenuation of the transmitted photon beam. EXAFS is one of the many "finestructure" techniques.

Local surface structure


and coordination
numbers

FEM

Field emission microscopy


(12, 20, 21 )

A strong electric field (on the order of V/ ) is applied to the tip of a sharp,
single-crystal wire. Electrons tunnel into the vacuum and are accelerated
along radial trajectories by Coulomb repulsion. When the electrons
impinge on a fluorescent screen, variations of the electric field strength
across the surface of the tip are displayed.

Surface structure

FIM

Field ionization microscopy


(12, 20, 21, 23 )

A strong electric field (on the order of V/ ) is created at the tip of a sharp,
single-crystal wire. Gas atoms, usually He, are polarized and attracted to
the tip by the strong electric field, then ionized by electrons tunneling from
the gas atoms into the tip. These ions, accelerated along radial trajectories
by Coulomb repulsion, map variations in the electric field strength across
the surface, showing the surface topography with atomic resolution.

Surface structure and


surface diffusion

FTIR

Fourier transform infrared


spectroscopy (13, 24 )

Broad-band IRAS experiments are performed, and the IR absorption


spectrum is deconvoluted by using a Doppler-shifted source and Fourier
analysis of the data. This technique is not restricted to surfaces.

Bonding geometry and


strength

HREELS

High-resolution electron
energy loss spectroscopy
(12, 21, 23 )

A monoenergetic electron beam, usually ~2 10 eV, is scattered off a


surface, and the energy losses below ~0.5 eV to bulk and surface phonons
and vibrational excitations of adsorbates are measured as a function of
angle and energy (also called EELS).

Bonding geometry,
surface
atom vibrations

IRAS

Infrared reflection
absorption spectroscopy
(21, 25 )

IR photons are reflected off a surface and the attenuation of IR intensity is


measured as a function of frequency yielding a spectrum of the vibrational
excitations of adsorbed molecules. Recent improvements in sensitivity allow
IRAS measurements to be made on single-crystal surfaces.

Molecular structure

ISS

Ion scattering spectroscopy


(12, 13, 21, 23, 24, 26 )

Ions are inelastically scattered from a surface, and the chemical


composition of the surface is determined from the momentum transfer to
surface atoms. The energy range is ~1 keV to 10 MeV, and the lower
energies are more surface sensitive. At higher energies this technique is
also known as Rutherford back-scattering (RBS).

Surface structure,
composition

LEED

Low-energy electron
diffraction (12 14, 20, 21,
23, 24, 26 )

Monoenergetic electrons below ~500 eV are elastically back-scattered


from a surface and detected as a function of energy and angle. This gives
information on the structure of the near surface region.

Atomic and molecular


surface structure
Continued on next page

166

Journal of Chemical Education Vol. 75 No. 2 February 1998 JChemEd.chem.wisc.edu

The Flexible Surface

Table 1. Selection of Most Frequently Used Surface Science Techniques Continued


Primary Surface
Information

Acronym

Name (References)

Description

Neutron diffraction (27 )

Neutron diffraction is not an explicitly surface-sensitive technique, but


experiments on large surface-area samples have provided important
structural information on adsorbed molecules and surface phase
transitions.

NMR

Nuclear magnetic
resonance (28 )

NMR is not an explicitly surface-sensitive technique, but NMR data on large Chemical state
s u r f a c e - a r e a s a m p l e s ( 1 m 2) h a v e p r o v i d e d u s e f u l i n f o r m a t i o n o n
molecular adsorption geometries. The nucleus magnetic moment interacts
with an externally applied magnetic field and provides spectra highly
dependent on the nuclear environment of the sample. The signal intensity
is directly proportional to the concentration of the active species. This
method is limited to the analysis of magnetically active nuclei.

RHEED

Reflection high-energy
electron diffraction
(13, 20, 21, 23, 24 )

Monoenergetic electrons of ~1 20 keV are elastically scattered from a


surface at glancing incidence and detected as a function of angle and
energy for small forward-scattering angles. Backscattering is less important
at high energies and glancing incidence is used to enhance surface
sensitivity.

Surface structure,
structure of thin films

SFG

Sum frequency generation


(29, 30 )

Similar to SHG, but the laser output is split into a visible laser beam and a
tunable IR beam. The two beams meet at the surface. The sum frequency
signal is monitored as a function of IR frequency. In this way the vibrational
spectrum of adsorbed molecules is obtained.

Molecular structure

SHG

Second harmonic
generation (29, 31 )

A surface is illuminated with a high-intensity laser, and photons are


generated at the second harmonic frequency through nonlinear optical
processes. For many materials only the surface region has the appropriate
symmetry to produce an SHG signal. The nonlinear polarizability tensor
depends on the nature and geometry of adsorbed atoms and molecules.

Electronic structure,
molecular orientation

SIMS

Secondary ion mass


spectrometry (12, 13, 21,
23, 24, 32 )

Ions and ionized clusters ejected from a surface during ion bombardment
are detected with a mass spectrometer. Surface chemical composition and
some information on bonding can be extracted from SIMS ion fragment
distributions.

Surface composition

STM

Scanning tunneling
microscopy (12, 13, 16,
24, 33 )

The topography of a surface is measured by mechanically scanning a


probe over a surface. The distance from the probe to the surface is
measured by the probe-surface tunneling current. Angstrom resolution of
surface features is routinely obtained.

Atomic surface structure

TEM

Transmission electron
microscopy (13, 24, 34,
35 )

TEM can provide surface information for carefully prepared and oriented
bulk samples. Real images have been formed of the edges of crystals where
surface planes and surface diffusions have been observed. Diffraction
patterns of reconstructed surfaces, superimposed on the bulk diffraction
pattern, have also provided surface structural information.

Surface structure

TDS

Thermal desorption
spectroscopy
(21, 23, 36, 37 )

An adsorbate-covered surface is heated, usually at a linear rate, and the


desorbing atoms or molecules are detected with a mass spectrometer. This
gives information on the nature of adsorbate species and some information
on adsorption energies and the surface structure.

Composition, heat of
adsorption, surface
structure

TPD

Temperature programmed
desorption (21, 36, 37 )

Similar to TDS, except the surface may be heated at a nonuniform rate to


obtain more selective information on adsorption energies.

Composition, heat of
adsorption, surface
structure

UPS

Ultraviolet photoemission
spectroscopy
(12, 14, 20, 21, 23 )

Electrons photoemitted from valence and conduction bands are detected


as a function of energy to measure the electronic density of states near the
surface. This gives information on the bonding of adsorbates to the surface.

Valence band structure

XPS

X-ray photoemission
spectroscopy (12, 13, 20,
21, 23, 24, 26, 38 )

Electrons photoemitted from atomic core levels are detected as a function


of energy. Shifts of core-level energies give information on the chemical
environment of the atoms.

Composition, oxidation
state

XRD

X-ray diffraction (39 )

X-ray diffraction has been carried out at extreme glancing angles of


incidence where total reflection ensures surface sensitivity. This provides
structural information that can be interpreted by well-known methods. An
extremely high X-ray flux is required to obtain useful data from single-crystal
surfaces. Bulk X-ray diffraction is used to determine the structure of organometallic clusters, which provide comparisons to molecules adsorbed on
surfaces. X-ray diffraction has also given structural information on large
surface-area samples.

Surface structure

Molecular structure

JChemEd.chem.wisc.edu Vol. 75 No. 2 February 1998 Journal of Chemical Education

167

Viewpoints

Figure 5. Restructuring at a step site on a clean surface. Each atom


attempts to optimize its coordination and cracks open to close
the step edges.

Figure 6. Side (a) and top (b) views on the Fe3O4(111) surface
structure with the spacing relaxations shown. The corresponding
bulk values are = 0.63 , and d12 = d23 = 1.19. The A and B
layers are strongly expanded by ~ 0.46.

Phenomena Discovered by Molecular Surface Chemistry

move in such a way that the positive and negative ions are
almost coplanar. Presumably because of the necessary condition
of charge neutrality, this type of surface structure is thermodynamically more stable than having alternating oxygen-ion
and iron-ion layers. Such an expansion at the surface is clearly
a property of ionic solids, and future studies will prove how
general this type of relaxation is.
Because of directionality of bonding in most solids, such
contraction or relaxation at the surface moves atoms from their
position of optimum bonding. As a result, the atoms not only
move perpendicular to the surface, but also parallel to the surface. This leads to the formation of new surface unit cells, a
phenomenon called surface reconstruction. Perhaps the most celebrated example is the (77) surface structure that forms on the
Si(111) crystal face. Figure 7a shows the low-energy electron diffraction (LEED) pattern from this surface. The complex unit
cell has 49 different locations of surface atoms that are distinguishable. The Si(100) surface shows a (21) reconstruction (Fig.
7b). It shows the formation of staggered dimers differing from
the arrangement of Si atoms in the bulk near the surface.

Surface Structure Is Different from Bulk Structure


Two dominant phenomena occurring at clean surfaces
of materials distinguish their atomic structure from that in
the bulk: relaxation and reconstruction (40). Upon relaxation
of metal surfaces, the first layer of atoms moves inward, and
this contraction leads to a much shortened interlayer spacing between the first and second layer of the surface. The
more open (rougher) the surface, the larger the relaxation.
Often, but not always, the contraction in the first layer is
followed by a small expansion in the second layer. At rough
edges, such as at stepped surfaces, the atoms at the step relax
by a large amount in order to smooth the surface irregularity. This is shown schematically in Figure 5.
At ionic surfaces, the nature of surface relaxation is very
different. Figure 6 shows what happens at iron oxide surfaces
(41). Iron oxide, in its bulk structure, shows alternating layers
of oxygen ions and iron ions, where the iron ions are in tetrahedral or octahedral positions. At the surface, the two ions

Figure 7. Left: Low-energy electron diffraction (LEED) pattern of the reconstructed Si(111) crystal face, exhibiting a (77) surface structure.
Right: The reconstructed Si(100) crystal face as obtained by LEED surface crystallography. Note that surface relaxation extends to three
atomic layers into the bulk.

168

Journal of Chemical Education Vol. 75 No. 2 February 1998 JChemEd.chem.wisc.edu

The Flexible Surface

Pt, Au, and Ir (100) surfaces that should have square


unit cells reconstruct to form hexagonal surface unit cells (40,
42). This is shown in Figure 8. STM studies have imaged
both reconstructed and unreconstructed surfaces (some domains
may be unreconstructed because of contamination by adsorbates of various types).
The water molecules at the ice surface vibrate with a much
larger amplitude (0.24 ) than molecules in bulk ice (0.1 )
(43). This motion causes surface softness, which is likely
responsible for the anomalously low friction coefficient of ice
(i.e., it makes ice so slippery).
Molecules at the liquidvapor interface also show restructuring, as shown in Figure 9 for the arrangement of molecules
at liquid alcohol surfaces (44). The alcohol molecules are oriented with their OH bonds pointing inward, presumably
for optimum hydrogen bonding. The alkane chains stick out
from the surface. This orientation has been readily detectable
by nonlinear laser optics sum frequency generation (SFG).

Adsorbate-Induced Restructuring of Surfaces


When the clean surfaces are covered with a near monolayer of chemisorbed molecules, the structure of the surface
undergoes profound alterations. This is perhaps best shown
in the field ion microscopy (FIM) studies carried out by Kruse
and coworkers (45) with rhodium field emission tips (Fig. 10).
When carbon monoxide is chemisorbed on these tips, every
crystal face restructures as shown by the figure. This massive
restructuring is reversible if CO is removed when the surface
is heated in vacuum.
Our new STM system (Fig. 4) is placed in an environmental cell that can be pressurized and heated to elevated
temperatures. It shows surface restructuring of the Pt(110)
surface when this surface is exposed to atmospheric pressures
of hydrogen, then oxygen, and then carbon monoxide (46).
When these surfaces are heated, the surface restructures from
a reconstructed ordered structure (exhibiting the so-called
missing row reconstruction) in the presence of hydrogen,

Figure 8. Top and side view of the Ir(100)-(15) surface reconstruction. The more open square (100) lattice is reconstructed into a
close-packed hexagonal overlayer, with a slight buckling as shown
in the side view.

to large (111) orientation facets in the presence of oxygen,


and then again to smooth (110) unreconstructed surfaces in
the presence of carbon monoxide (Fig. 11).
Low-energy electron diffraction surface crystallography
studies indicate the detailed atomic level nature of such reconstructions (42). When carbon is adsorbed on the Ni(100)
surface, it occupies fourfold hollow sites (Fig. 12). As a result of the formation of the carbonmetal chemisorption
bonds, the surface metal atoms move away from the adsorption
site, presumably to give more space to the carbon atom so it
can sink deeper into the surface, thereby forming bonds with
second-layer nickel atoms underneath (47). This expansion
around the chemisorption site induces strain, which is relieved
by rotation of the surface unit cell as shown in Figure 12.
When sulfur is chemisorbed on the Fe(110) crystal face
(48), the S atom pulls the neighboring Fe atoms into equal
distances from the chemisorption site to form four equal FeS
bonds. The strength of these bonds pays for the weakening of
the metalmetal bonds as a result of the restructuring. When
NO is adsorbed on the Ni(111) surface, the molecule occupies
a threefold hollow site, a so-called hcp (hexagonal close-packed)
hollow site. This means there is a metal atom directly underneath the chemisorption site in the second metal layer.
Chemisorption induces an upward movement of this metal
atom in the second layer, and rumpling of the metal surface.
When ethylene adsorbs on the Rh(111) surface (49), it rearranges and occupies a hollow site (in this case, again, an
hcp hollow site). The rearranged ethylene (which has lost a
hydrogen) is called ethylidyne. This is shown in Figure 13.
The metal atoms move away from the carbon atom bound
to the hollow site to allow the carbon to bond to the Rh atom
directly underneath the carbon in the second layer. On the
Pt(111) surface (50, 51), ethylene also forms an ethylidyne
moleculeagain in a threefold hollow site, but in this case
it is an fcc hollow site. That is, there is no metal atom directly underneath the carbon in the second metal layer. In
this circumstance, the surface metal atoms move inward to
presumably provide as strong a bond as possible to the carbon,
and metalmetal distances are altered on the surface as well.
The second metal atom next to the chemisorption bond
moves downward to produce a corrugated surface. It appears
that surface bonding is clusterlike, where nearest-neighbor
metal atoms that surround the adsorbate move to optimize
the surface chemical bond. The heat of adsorption, which is
always exothermic, pays for the weakening of the next nearest
neighbor metalmetal bonds, which are altered as a result of
the movement of the metal atoms nearest to the chemisorption site.

Figure 9. The normal alcohols show substantial ordering at the liquidvapor interface. The OH groups of the alcohols extend into
the liquid forming a hydrogen-bonded network, while the alkyl
chains are oriented away from the liquid.

JChemEd.chem.wisc.edu Vol. 75 No. 2 February 1998 Journal of Chemical Education

169

Viewpoints

When two molecules are coadsorbed on the surface,


adsorbate-induced reconstruction may be very different from
when only one or the other molecule is chemisorbed. This is
shown for benzene and CO coadsorption on the Rh(111)
surface (Fig. 14). When these molecules form a mixed unit
cell, under the benzene the metal atoms are closer to their
bulklike configuration than under the CO molecule (52).
There is rumpling of the surface that occurs because of the
differences in chemical bonding of the coadsorbed species to
the substrate metal atoms. When benzene adsorbs alone it is
bent. Four of the C atoms are in one type of surface site while
two of the others are in different types of surface sites. When
coadsorption occurs, the benzene molecule is flattened out.

Rough Surfaces Do Chemistry


Surface irregularities, steps, and kinks are very effective
for breaking adsorbate chemical bonds and in catalysis as well.
This is best known by the temperature-programmed desorption
(TPD) of H2 from flat, stepped, and kinked surfaces of Pt
(Fig. 15). H2 desorbs at maximum rates at the highest temperature from kink sites, then at somewhat lower temperatures from
step sites, and at even lower temperatures from flat (111) terraces (53). This indicates higher heats of adsorption of the H

Table 2. Structure Sensitivity of H2/D2 Exchange at Low


Pressures (~10 6 torr)
Surface

Reaction Probability

Stepped Pt(332)

0.9

Flat Pt(111)

~10 1

Defect-free Pt(111)

10 3

atom at these defect sites. Thus, the thermodynamic driving


force for dissociation is certainly greater at these sites, which
can explain their enhanced bond-breaking activity.
It is difficult to understand, however, that these same
strongly adsorbing sites are also very active sites for catalysis.
This is shown in Table 2. The reaction probability of H2/D2
exchange on stepped surfaces is near unity at low pressures
on a single scattering event, whereas it is below the detection
limit (< 103) on the flat (111) crystal face, as shown by molecular beam surface scattering studies (54). How is it possible
that the strongly adsorbing step sites, where H has a long
residence time because of its high binding energy, are also
the sites of rapid reaction turnover?
One possible explanation is that the strongly adsorbed
hydrogen restructures the surface near the step, thereby creat-

Figure 10. Field ion micrographs (image gas: Ne; T


= 85 K) of a (001)-oriented
Rh tip (top left ) before and
(bottom left ) after reaction
with 104 Pa CO for 30 min
at 420 K . The stereographic projections at the
right demonstrate the
change in morphology from
nearly hemispherical to polygonal. The scheme at the
bottom right indicates the
coarsening of the crystal
and the dissolution of a
number of crystallographic
planes due to the reaction
with CO.

170

Journal of Chemical Education Vol. 75 No. 2 February 1998 JChemEd.chem.wisc.edu

The Flexible Surface

Figure 11. In situ high-pressure STM


pictures showing adsorbate-induced
surface reconstructions of Pt(110) under atmospheric pressures: Top: Topographic image of the surface in hydrogen after heating to 425 K for 5
hours, showing (n1) missing-row reconstruction randomly nested. Vertical
range: z = 10 . Center: Topographic image of the surface in oxygen after heating to 425 K for 5
hours; z = 25 . Bottom: Topographic image of the surface in carbon monoxide after heating to 425 K
for 4 hours; z = 42 .

ing the active site for the catalytic exchange process. At the
low pressures of these molecular beam scattering experiments,
the low coverages keep the structure of the flat part of the
surface unaltered.
Similarly, stepped Ni surfaces dehydrogenate C2H4 at
much lower temperatures (< 150 K) than the (111) face of
Ni (~230 K). The more open (111) and (211) crystal faces
of Fe (Fig. 16) are several orders of magnitude more active
for NH3 synthesis than the close-packed Fe (110) crystal face,
which showed no detectable reaction rate ( 55).

Clusterlike Bonding of Adsorbed Molecules


When ethylene chemisorbs at ~ 300 K on the (111) crystal
faces of various transition metals (Pt, Rh, Pd), it chemically
rearranges to form the molecule-surface compound shown
in Figure 13. Its structure as determined by LEED-surface
crystallography is very similar to multinuclear organometallic complexes such as Os3 CCH3 or Co3 (CO)9 CCH3. The
rearranged ethylene, which has also lost a hydrogen, is called
ethylidyne and belongs to the alkylidyne group (species of the
formula CnH2n1), a common substituent in surface chemistry

Figure 12. Carbon-chemisorption-induced restructuring of the


Ni(100) surface.

and organometallic chemistry. The vibrational spectrum of


chemisorbed ethylidyne is nearly identical to that of the organometallic cluster Os3 CCH3 , which contains three metal atoms.
The CC bond distance (1.45 ) is slightly less than the
single carboncarbon bond length of 1.54 (0.154 nm), as
in cluster compounds. Thus, the surface chemical bond of
chemisorbed ethylene can, as a first approximation, be viewed
as a clusterlike bond that contains at least three metal atoms
(56). The CC bond order present in gaseous ethylene is
reduced from two to nearly one upon chemisorption. This
reduction in bond order of alkenes and alkynes upon chemisorption on metal surfaces is commonly observed, indicating
charge transfer from the molecules into the metal.
Benzene usually chemisorbs on metals with its ring parallel
to the surface (although it may adsorb in a different configuration when it loses hydrogen). Because of charge transfer to the
metal, CC bond elongations occur with respect to the symmetry of the adsorption site. The ring may even bend, with
two of the opposing carbon atoms closer to the metal surface
than the other four carbon atoms. Distortions and elongations of CC bonds are also found when benzene is bound
to clusters of metal atoms in organometallic complexes. Thus
the clusterlike bonding model appears to be valid for chemisorbed benzene as well.
The bonding picture of adsorbed molecules becomes
more complicated if there are more bonding sites available
on the same molecule. For example, pyridine (C5H5N) may
bind through the lone electron pair of its nitrogen or through
the -electrons of the carbon ring. Thus, depending on the
metal, the binding geometry of the substrate, the temperature,
or the adsorbate coverage, the molecule may be tilted with
respect to the substrate surface, its ring may be parallel with it,
or it may be upright with bonding solely through the nitrogen.
It is too simplistic to consider that only the nearestneighbor metal atoms of the substrate participate in the bonding. There is evidence that the atoms at next-nearest-neighbor sites change their location when chemisorption occurs,
moving either closer or further away from the chemisorption bonds.

JChemEd.chem.wisc.edu Vol. 75 No. 2 February 1998 Journal of Chemical Education

171

Viewpoints

Figure 13. The structure of


ethylidyne on Rh(111). Ethylidyne
is bonded on the hcp 3-fold hollow
site. This site has a metal atom right
underneath the carbon bonding site
in the second layer. The adsorptioninduced distortion in the top metal
layers pulls the nearest neighbor
metal atoms up out of the surface
plane.

The Flexible Surface


As a result of all these studies that indicate adsorbateinduced restructuring and clusterlike bonding, the new model
of the surface which has been adopted is the so-called flexible surface (53). In the past it was assumed that the metal
atoms at the surface are rigid and occupy equilibrium sites
dictated by the bulk unit cell. On adsorption, their location
would not be altered. Instead, the flexible surface is one where
the metal atoms move into new sites, dictated by the chemisorption bond so as to optimize that bond: upon adsorption
the surface restructures, thereby creating the active sites for
surface chemical processes.

The flexible-surface model explains why rough surfaces


or defect sites at surfaces are so active in surface chemistry.
Bond breaking and catalysis most frequently occur at low
coordination sites such as steps and kinks or at defect sites
such as oxygen vacancies in oxide surfaces. The lower the coordination of metal atoms (the fewer nearest neighbors), the
more easily they restructure to optimize the surface adsorption bond. Thus, rough surfaces or atoms at steps move more
readily, and of course small clusters of atoms where the coordination is much reduced are the most flexible. It is not
surprising therefore that we use nanoclusters in the field of
catalysis (and in many instances chemisorption) to optimize
chemical effects, such as chemical reactions or adsorption.
The extraordinary diversity of surface chemistry is due
to the chameleon-like change of surface structure and bonding as the chemical environment of the surface is altered.
Platinum is an excellent combustion catalyst operating in an
oxidizing atmosphere and is a primary ingredient of the automobile catalytic converter. Platinum is also an excellent catalyst
for hydrocarbon conversion under reducing conditions to
produce high-octane gasoline (aromatic molecules and
branched isomers) from straight chain alkanes (for example
from n-hexane and n-heptane, which have octane numbers
near zero). The platinum surface structure and thus its bonding
behavior is completely different under the different reaction
conditions, thereby mediating dramatically different catalytic
surface chemistry.
Technological Impact of Molecular Surface Chemistry:
Selected Examples

Figure 14. The coadsorbed surface structure of benzene and carbon monoxide on the Rh(111) crystal face as obtained by lowenergy electron diffraction surface crystallography.

172

Molecular surface chemistry contributed to the development and improvement of a wide range of technologies
(57). A complete description is far beyond the limits of this
article. In this section we will discuss only its present and
future impact on catalysis and semiconductor devices.
Since the early 1970s, molecular surface chemistry has
made significant contributions to our understanding of catalyst-

Journal of Chemical Education Vol. 75 No. 2 February 1998 JChemEd.chem.wisc.edu

The Flexible Surface

based chemical and petroleum technologies. The hydrogenation


of carbon monoxide yields methane or methanol exclusively,
oxygenated molecules containing several carbon atoms or liquid,
and high-molecular-weight hydrocarbon products (depending
on the type of catalyst employed) (40). The dissociation of
CO was found to be the dominant reaction step in producing
methane, followed by stepwise hydrogenation of the surface
carbon over several transition metal surfaces. Potassium was
found to be an outstanding promoter of CO dissociation
through weakening of the CO bond by charge transfer.
Methanol production was found to occur through the hydrogenation of undissociated CO2 or CO (which reaction dominates depends on catalyst formulation). Higher-molecular-weight
hydrocarbons are produced by secondary polymerization
reactions.
The reforming of naphtha over platinum was found to
be a structure-sensitive reaction. By altering the surface structure of platinum particles [(111) or (100) orientation] the
product distribution could be altered. Bimetallic platinumbased catalysts (PtRe, PtIr, and PtSn) have also been investigated by surface-science studies. These studies have contributed greatly to their optimization in this important highoctane fuel producing technology.
The same is true for the iron-based catalyst that produces
ammonia from N2 and H2. The structure sensitivity of this
reaction was uncovered, implicating the (111) and (211) crystal faces of iron as the most active (58, 59) (Fig. 16). The
surface structures that are more open and contain sites that
are surrounded by seven iron neighbor atoms (C7 sites) are
the most active. These are the (111) and (211) crystal faces.
The structure of the (110) crystal face does not allow the
adsorbed nitrogen species to bind with second- and third-layer
atoms (55), and probably for this reason the rate for the synthesis of ammonia is some 500 times lower than on the (111)

Figure 15. Thermal desorption spectra of hydrogen from a flat


(111), a stepped (557), and a kinked (12,9,8) surface of platinum.

Figure 16. Schematic representations of the idealized surface structures of the (111), (211), (100), (210), and (110) orientation of
iron single crystals (the coordination of each surface atom is indicated). The bar diagram shows the corresponding activity of the
single crystal surfaces in ammonia synthesis.

JChemEd.chem.wisc.edu Vol. 75 No. 2 February 1998 Journal of Chemical Education

173

Viewpoints

surface structure. The role of potassium promoters as bonding modifiers in aiding the dissociation of N2, as well as weakening the bonding of ammonia to the iron surface to inhibit
product poisoning, has been uncovered. The role of alumina
as a structure modifier, aiding the restructuring of iron particles to possess crystal faces most active for ammonia synthesis [(111) and (211)], has been proven by surface-science studies. As a result, a new generation of catalysts with superior activity could be prepared for this important industrial process.
Environmentally important catalytic processes have become the focus of rapid development in recent years. None
of them is more important than the 3-way catalytic converter
utilized to clean automobile exhaust. It utilizes Pt, Pd, Rh,
and cerium oxide as an important promoter (60). The challenge is to fully oxidize unburned hydrocarbons and CO to
CO2 while reducing NOx to N2 under all conditions of engine use: cold start, steady-state operation, and using a broad
range of air and fuel mixtures. This is achieved with the help
of one of the most successful chemical sensors, the oxygen
detector ( -probe) on automobiles. It helps to adjust the airto-fuel ratio of the mixture entering the internal combustion
engine and to optimize the efficiency of the 3-way catalytic
converter. This technology works well on the present-day
automobile. The development of lean-burning, more fuelefficient cars presents new challenges to surface chemistry and
the technology used to clean automobile exhaust.
Semiconductor-based technologies are at the heart of computer manufacturing. The fabrication of microelectronic circuits
often involves layer-by-layer deposition of semiconductor (Si,
GaAs, etc.), metal (Al, Cu, etc.), and insulator (SiO2, polymer)
thin films, in various configurations. The film thickness of
each of these materials is presently in the 103104- range,
and these layers alternate in both two and three dimensions.
Fabrication of these layers is carried out by surface processes
using chemical vapor deposition, sublimation, or sputter
deposition from a radio-frequency plasma. Nucleation and
growth mechanisms are monitored by surface-science techniques
such as reflection high-energy electron diffraction (RHEED)
and electron microscopy. We shall look at two problems of
semiconductor device technology that are currently the focus of intense surface-science studies.
The first is that insulating gate oxides for metal oxide
semiconductor field effect transistors (MOSFET) are produced
by oxidizing silicon to SiOx. Both the oxygen-to-silicon ratio
and the thickness of the oxide are important process variables, as they control the devices performance. The gate oxides
must become thinner, their surfaces or interfaces must be
atomically smooth, and their impurity concentrations must
be minimized in order to increase the speed of electron transport and device reliability.
The second problem is that the chemical and mechanical
integrity of the metal-insulator interfaces can be compromised
by water vapor or by the chemical attack of impurities segregating at the interface (e.g., alkali atoms, carbon, oxygen).
When this happens, the adhesion of the insulator oxide to
the metal is altered and delamination occurs. Trap states
that arise at the SiSiO2 interface from Si atoms with coordination numbers other than 4 are another problem because they
can trap charge at the interface. All these changes of chemical
and mechanical properties at the interface can have very deleterious effects on the electrical properties. It is essential that
we learn how to fabricate chemically stable insulatormetal
174

interfaces that maintain adhesion under changing ambient


conditions (temperature, humidity, etc.). As the insulating
oxide is replaced by a polymer with a smaller dielectric constant, the study of metalpolymer interfaces becomes a frontier area of the science of semiconductor surface technology.
Other important surface technologies developed in recent
years with the help of molecular surface chemistry should
also be mentioned. Air separation to oxygen and nitrogen was
accomplished with the help of molecular sieves because of
their higher heat of adsorption for N2 (~ 7 kcal/mol) than
for O2 (~ 3 kcal/mol), thereby preferentially releasing oxygen.
Conversely, microporous carbon engineered to have bimodal
distribution of pores adsorbs the smaller O2 (3.46 ) in its
small pores in preference to N2 (3.64 ) and preferentially
releases nitrogen. The magnetic disc drive provides information
storage in computers by nanoscale tracking of a magnetic thin
film that is coated by an atomically smooth carbon deposit
and lubricated by a drop of fluoro-ether lubricant. The optical
fiber operates on total internal reflection in glass by the
cladding that is made by another glass of different composition
and with a smaller refractive index. Such glass structures are
produced using chemical vapor deposition of oxides in an
appropriate sequence. Diamond films are produced by application of a high-energy plasma of methane and hydrogen or
by chemical vapor deposition to provide a chemically inert,
extremely hard coating with high thermal conductivity, that
withstands chemical and mechanical attacks superbly (61).
Adhesives and the contact lens are important surface technologies using polymers that provide controlled adhesion and
some degree of biocompatibility, respectively.
Future Directions in Surface Chemistry

Coadsorption on Surfaces
Most studies of modern surface chemistry focus on single
adsorbate systems, atoms, or molecules and investigate structure
and bonding, adsorption, diffusion, and desorption dynamics
as a function of temperature. In most chemical reactions, however, two or more reactants are utilized. Oxygen, hydrogen,
and water are the coadsorbates present most frequently, although
mixtures of organic molecules are adsorbed during hydrocarbon conversion reactions. Catalytic systems always use
additives that accelerate the reaction rate or improve selectivity. Potassium, which is an electron donor to iron, weakens
the bonding of the product molecule, ammonia (NH3 ),
thereby accelerating its desorption from the surface. Conversely, potassium increases the bonding of carbon monoxide
(an electron acceptor) to iron and other transition metals,
leading to the dissociation of the molecule (62). As a result,
the rate of production of hydrocarbons by the hydrogenation
of CO is greatly accelerated. In the future, more studies involving coadsorbed atoms and molecules must be pursued to uncover their influence on surface chemistry as a consequence of
their interactions, adsorbateadsorbate and adsorbate-substrate.

High-Pressure Surface Science


Increasing the coverage of adsorbates often enhances restructuring of the substrate and changes the surface chemistry. The
easiest way to increase coverage is to increase the reactant pressure, since pressure is proportional to coverage (adsorption isotherm). When CO adsorbs on platinum surfaces it occupies
mostly top and bridge sites, with its C=O bond perpendicular

Journal of Chemical Education Vol. 75 No. 2 February 1998 JChemEd.chem.wisc.edu

The Flexible Surface

to the metal surface (63). As the pressure is increased to above


10 torr a high coverage incommensurate overlayer forms, and
at even higher pressures ( 100 torr) the formation of platinum
carbonyl clusters with CO-to-Pt-ratio greater than one can be
detected. This pressure-dependent change in surface chemistry
is driven by the formation of strong COPt bonds that induce the weakening and then the breaking of metalmetal
bonds to produce a thermodynamically more stable surface
structure. Newly available techniques permit atomic and molecular studies of external surfaces in the presence of highpressure gas or a liquid at the interface. These include the
scanning tunneling and atomic force microscopes (STM and
AFM) and sum frequency generation (SFG)surface vibrational spectroscopy (10, 46). In the future these techniques
and others will be used to carry out in situ studies of molecular surface chemistry at high reactant pressures and at high
temperatures.

Monitoring Surface Chemistry at Ever-Improving


Spatial Resolution and Time Resolution
STM and AFM provide spatial resolution of surfaces and
surface species on the nanometer scale. There is a continuing
need to develop spectroscopic techniques that have the same
resolution because they will provide means to study and manipulate surfaces on that spatial scale. Increased time resolution will permit us to monitor the motion of surface atoms and
molecules, their diffusion, rotation, vibrational and electronic
excitation, and their reaction dynamics.

Studies of the Buried Interfaces, SolidLiquid and


SolidSolid
Techniques that open up high-pressure surface chemistry also permit molecular studies of the buried interfaces. This
will result in rapid developments in molecular phenomena
at solidliquid (64) and solidsolid interfaces, including electrochemistry, biology, and tribology (friction, lubrication,
wear) (65).

Achieving 100% Selectivity in Surface Reactions


the Environmental Imperative
In most surface catalyzed reactions we desire to obtain
only one product, although the formation of other chemicals is
also thermodynamically allowed. We need to understand catalytic selectivity, how to obtain 100% selectivity to avoid the
formation of undesirable molecules that often lead to separation problems, pollution, or catalytic deactivation.

Nanoparticles: Surfaces in Three Dimensions


Particles with dispersions between 1 and 0.1 represent
transition between single atoms and molecules and the bulk
solid. They have many surprising properties as their electronic
structure, atomic structure, and phase diagramand as a consequence, their surface chemistrychange with particle size.
The fabrication of nanoparticles of uniform size (66, 67) and
their study is one of the intellectual frontiers of modern surface chemistry.
Molecular surface chemistry is one of the most rapidly
developing branches of chemistry. It is an intellectual frontier
of the discipline with enormous potential to develop surface
technologies that improve our quality of life, create employment, and create wealth. It will remain a rapidly advancing
frontier area for many years to come.

Acknowledgment
This work was supported by the Director, Office of Energy Research, Office of Basic Energy Sciences, Materials Sciences Division, of the U.S. Department of Energy under
Contract No. DE-AC03-76SF00098.
Literature Cited
1. Berzelius, J. Jahres-Bericht ber die Fortschritte der Physischen
Wissenschaften (Tbingen), 1836, 15.
2. Thomas, J. M. Michael Faraday and the Royal Institution; IOP: Bristol,
1991.
3. Csicsery, S. M. In Zeolite Chemistry and Catalysis; Rabo, J. A., Ed.;
ACS Monographs, Vol. 171; American Chemical Society: Washington, DC, 1976; Chapter 1.
4. OHanlon, J. F. A Users Guide to Vacuum Technique, 2nd ed.; Wiley:
New York, 1989.
5. Klauber, C. In Surface Analysis Methods in Materials Science; Springer Series in Surface Sciences, Vol. 23; OConnor, D. J.; Sexton, B. A.; Smart,
R. S. C., Eds.; Springer: Berlin, 1992; pp 6776.
6. de Boer, J. H. The Dynamical Character of Adsorption; Oxford University Press: New York, 1968.
7. Tompkins, F. C. Chemisorption of Gases on Metals; Academic: New
York, 1978.
8. Cabrera, A. L.; Spencer, N. D.; Kozak, E.; Davies, P. W.; Somorjai,
G. A. Rev. Sci. Instrum. 1982, 53, 1888.
9. Somorjai, G. A.; Rupprechter, G. In Dynamics of Surfaces and Reaction Kinetics in Heterogeneous Catalysis; Studies in Surface Science and
Catalysis Series, Vol. 109; Froment, G. F.; Waugh, K. C., Eds.;
Elsevier: Amsterdam, 1997; p 35.
10. Su, X., et al. Faraday Discuss. 1996, 105, 263.
11. Greg, S. J.; Sing, K. S. W. Adsorption, Surface Area, and Porosity; Academic: New York, 1967.
12. Hudson, J. B. Surface Science: An Introduction; ButterworthHeinemann: Boston, 1992.
13. MacDonald, R. J.; King, B. V. In Surface Analysis Methods in Materials Science; Springer Series in Surface Sciences, Vol. 23; OConnor,
D. J.; Sexton, B. A.; Smart, R. S. C., Eds.; Springer: Berlin, 1992;
Chapter 5. Also see Chapters 3, 6, 1014 in this volume.
14. Ertl, G.; Kppers, J. Low Energy Electrons and Surface Chemistry; VCR:
Weinheim, 1985.
15. Binnig, G.; Quate, C. F.; Gerber, C. Phys. Rev. Lett. 1986, 56, 930.
16. Hansma, P. K.; Elings, V. B.; Marti, O.; Bracker, C. E. Science 1988,
242, 157.
17. Gasser, R. P. H. An Introduction to Chemisorption and Catalysis by
Metals; Oxford University Press: New York, 1985.
18. Richardson, N. V.; Bradshaw, A. M. In Electron Spectroscopy: Theory,
Techniques and Applications, Vol. 4; Brundle, C. R.; Baker, A. D., Eds.;
Academic: New York, 1981; pp 153193. Also see Chapters 1, 2, and
5.
19. Ibach, H.; Mills, D. L. Electron Energy Loss Spectroscopy and Surface
Vibration; Academic: New York, 1982.
20. Prutton, M. Surface Physics; Oxford University Press: New York, 1975.
21. Woodruff, D. P.; Delchar, T. A. Modern Techniques of Surface Science;
Cambridge Solid State Science Series; Cambridge University Press:
New York, 1986.
22. Heald, S. M. In X-ray Absorption; Chemical Analysis, Vol. 92;
Koningsberger, D. C.; Prins, R., Eds.; Wiley: New York, 1988.
23. Somorjai, G. A.; Van Hove, M. A. In Absorbed Monolayers on Solid
Surfaces ; Structure and Bonding Series, Vol. 38; Dunitz, J. D.;
Goodenough, J. B.; Hemmerich, P.; Ibers, J. A.; Jorgensen,
C. K.;
/
Neilands, J. B.; Reinen, D.; Williams, R. J. P., Eds.; Springer: Berlin,
1979; Chapter 4.
24. Roberts, N. K. In Surface Analysis Methods in Materials Science;
Springer Series in Surface Sciences, Vol. 23; OConnor, D. J.; Sexton, B. A.; Smart, R. S. C., Eds.; Springer: Berlin, 1992; Chapter
8, pp 187201.
25. Willis, R. F.; Lucas, A. A.; Mahan, G. D. In Adsorption at Solid State
Surfaces. The Chemical Physics of Solid Surfaces and Heterogeneous
Catalysis, Vol. 2; King, D. A.; Woodruff, D. P., Eds.; Elsevier: New
York, 1983.
26. Somorjai, G. A.; Van Hove, M. A. In Investigations of Interfaces and Surfaces and Interfaces, Part B, Vol. IXB; Rossiter, B. W.; Baetzold, R. C.,
Eds.; Wiley-Interscience: New York, 1993; Chapter 1.

JChemEd.chem.wisc.edu Vol. 75 No. 2 February 1998 Journal of Chemical Education

175

Viewpoints
27. Lechner, R. E.; Riekel, C. In Neutron Scattering and Muon Spin Rotation; Springer Tracts in Modern Physics, Vol. 101; G. Hhler, Ed.;
Springer: Berlin, 1983; pp 184.
28. Barrie, P. J.; Klinowski, J. Prog. Nucl. Magn. Reson. 1992, 24, 91.
29. Shen, Y. R. Nature 1989, 337, 519.
30. Shen, Y. R. Annu. Rev. Phys. Chem. 1989, 40, 327.
31. Richmond, G. L.; Robinson, J. M.; Shannon, V. L. Prog. Surf. Sci.
1988, 28, 1.
32. MacDonald, R. J.; King, B. V. In Surface Analysis Methods in Materials
Science; Springer Series in Surface Sciences, Vol. 23; OConnor, D. J.;
Sexton, B. A.; Smart, R. S. C., Eds.; Springer: Berlin, 1992.
33. Binnig, G.; Rohrer, H. Rev. Mod. Phys. 1987, 59, 615.
34. Cowley, J. M. Prog. Surf. Sci. 1986, 21, 209.
35. Thomas, G. Ultramicroscopy 1986, 20, 239.
36. Yates, J. T. In Solid State Physics: Surfaces; Methods of Experimental
Physics Series, Vol. 22; Park, R. L.; Lagally, M. G., Eds.; Academic: New
York, 1985; Chapter 8, pp 425464.
37. Madix, R. J. In Chemistry and Physics of Solid Surfaces, Vol. 2;
Vanselow, R., Ed.; CRC: Boca Raton, FL, 1979; pp 6372.
38. Moulder, J. F.; Stickle, W. F.; Sobol, P. E.; Bomben, K. D. Handbook of
X-Ray Photoelectron Spectroscopy; Perkin Elmer: Eden Prairie, MN, 1992.
39. Robinson, I. K.; Tweet, D. J. Rep. Prog. Phys. 1992, 55, 599.
40. Somorjai, G. A. Introduction to Surface Chemistry and Catalysis; Wiley:
New York, 1994.
41. Barbieri, A.; Weiss, W.; Van Hove, M. A.; Somorjai, G. A. Surf. Sci.
1994, 302, 259.
42. Starke, U.; Van Hove, M. A.; Somorjai, G. A. Prog. Surf. Sci. 1994,
46, 305.
43. Materer, N.; Starke, U.; Barbieri, A.; Hove, M. A. V.; Somorjai, G.
A. Surf. Sci. 1997, 381, 190.
44. Stanners, C. D.; Du, Q.; Chin, R. P.; Cremer, P.; Somorjai, G. A.;
Shen, Y.-R. Chem. Phys. Lett. 1995, 232, 407.
45. Kruse, N.; Gaussman, H. Surf. Sci. 1992, 266, 51.
46. McIntyre, B. J.; Salmeron, M. B.; Somorjai, G. A. Rev. Sci. Intrum.
1993, 64, 687.

Gabor A. Somorjai

Gnther Rupprechter

University of California at Berkeley


Department of Chemistry and
Material Sciences Division of
Berkeley National Laboratory

University of California at Berkeley


Department of Chemistry and
Material Sciences Division of
Berkeley National Laboratory

Ph.D., Chemistry, 1960, University


of California at Berkeley
B.S., Chemical Engineering,
1956, Technical University,
Budapest, Hungary

Ph.D., Chemistry, 1996, Leopold


Franzens University, Innsbruck
B.S., Chemistry, 1992, Leopold
Franzens University, Innsbruck

Gabor A. Somorjai is one of the most influential scientists


in the area of surface chemistry today. He has received numerous awards, including the 1997 Von Hippel Award which
he was awarded on December 3, 1997. His influence on surface chemistry can be seen in more than 750 papers published on surface science, heterogeneous catalysis, and solid
state chemistry. Somorjai has educated more than 90 Ph.D.
students and collaborated more than 110 postdoctoral scientist of whom Gnther Rupprechter is one. The research
interests of his group include molecular studies of the structure and bonding of surfaces, surface science of heterogenous
catalysis, and molecular studies of polymer surfaces and polymerization. During the past 30 years Somorjai has made
significant contributions in surface science, new surface instrumentation, and catalysis.

176

47. Gauthier, Y.; Baudoing-Savois, R.; Heinz, K.; Landskron, H. Surf.


Sci. 1991, 251, 493.
48. Shih, H. E.; Jona, F.; Marcus, P. M. Phys. Rev. Lett. 1981, 46, 731.
49. Wander, A.; Van Hove, M. A.; Somorjai, G. A. Phys. Rev. Lett. 1991,
67, 626.
50. Cremer, P. S.; Stanners, C.; Niemantsverdriet, J. W.; Shen, Y. R.;
Somorjai, G. A. Surf. Sci. 1995, 328, 111.
51. Cremer, P. S.; Su, X.; Shen, Y. R.; Somorjai, G. A. J. Am. Chem. Soc.
1996, 118, 2942.
52. Van Hove, M. A.; Lin, R. F.; Somorjai, G. A. J. Am. Chem. Soc. 1986,
108, 2532.
53. Somorjai, G. A. Langmuir 1991, 7, 3176.
54. Salmeron, M.; Gale, R. J.; Somerjai, G. A. J. Chem. Phys. 1979, 70,
2807.
55. Somorjai, G. A.; Materer, N. Top. Catal. 1994, 1, 215.
56. Van Hove, M. A.; Bent, B.; Somorjai, G. A. J. Phys. Chem. 1988, 92,
973.
57. Somorjai, G. A. Chem. Rev. 1996, 96, 1223.
58. Strongin, D. R.; Carrazza, J.; Bare, S. R.; Somorjai, G. A. J. Catal.
1987, 103, 213.
59. Spencer, N. D.; Schoonmaker, R. C.; Somorjai, G. A. J. Catal. 1982,
74, 129.
60. Taylor, K. C. Catal. Rev.Sci. Eng. 1993, 35, 457.
61. Perry, S. S.; Somorjai, G. A. J. Vac. Sci. Technol. A 1994, 12, 1513.
62. Crowell, J. E.; Tysoe, W. T.; Somorjai, G. A. J. Phys. Chem. 1986,
89, 1598.
63. Su, X.; Cremer, P. S.; Shen, Y. R.; Somorjai, G. A. Phys. Rev. Lett.
1996, 77, 3858.
64. Somorjai, G. A. Surf. Sci. 1995, 335, 10.
65. Perry, S. S.; Somorjai, G. A.; Mate, C. M.; White, R. L. Tribol. Lett.
1995, 1, 233.
66. Jacobs, P. W.; Wind, S. J.; Ribeiro, F. H. Somorjai, G. A. Surf. Sci.
1997, 372, L249.
67. Eppler, A.; Rupprechter, G.; Guczi, L.; Somorjai, G. A. J. Phys. Chem.
B 1997, 101, 99739977.

Gnther Rupprechter received his Ph.D. in 1996 under the


supervision of Konrad Hayek on Microstructural and morphological changes on epitaxially grown noble metal catalyst
particles upon oxidation and reductive activation. At present
he is a postdoctoral fellow at the University of California at
Berkeley working with Gabor A. Somorjai. Rupprechters
current research interests include structureactivity correlations in heterogeneous catalysis, fabrication of well-faceted
polyhedral nanocrystals on oxidic supports by epitaxial
growth and electron beam lithography, and analysis of surface structure and surface composition. He also has received
a number of awards and fellowships.

Journal of Chemical Education Vol. 75 No. 2 February 1998 JChemEd.chem.wisc.edu

Вам также может понравиться