You are on page 1of 16

Viewpoints: Chemists on Chemistry

The Flexible Surface


Gabor A. Somorjai and Günther Rupprechter
The Flexible Surface: 162
Molecular Studies Explain the
Extraordinary Diversity of Surface
Chemical Properties
About Viewpoints

Outline
Historical Perspective
V iewpoints is a major feature of the celebration of the Journal
of Chemical Education’s 75th year. It is being supported by The
External Surfaces Camille and Henry Dreyfus Foundation, Inc., which recently cel-
Surface Concentration ebrated its own 50th anniversary. Each paper in the Viewpoints se-
Clusters and Small Particles ries will be written by a chemist or group of chemists with special
Thin Films
expertise in a particular field, with the aim of providing an overview
Internal Surfaces—Microporous Solids of that field’s accomplishments, importance, and prospects. The goal
Clean Surfaces 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 perspec-
Interfaces
Adsorption
tive encompassed by Viewpoints corresponds with the 75 years of
this Journal’s lifetime and reflects its comprehensive interest in all of
Techniques of Surface Science chemistry. The 50-year retrospective view of each field corresponds
Phenomena Discovered by Molecular with the period during which the Camille and Henry Dreyfus Foun-
Surface Chemistry dation has been supporting the chemical sciences.
Surface Structure Is Different from Bulk Structure Authors of Viewpoints papers will provide perspectives on what
Adsorbate-Induced Restructuring of Surfaces chemistry has done during the lifetime of the Dreyfus Foundation
Rough Surfaces Do Chemistry and to set the stage for what chemistry will become well into the
Clusterlike Bonding of Adsorbed Molecules
The Flexible Surface
next century. The papers will be written at a level appropriate for
upper-division undergraduate chemistry students and will extend and
Technological Impact of Molecular Surface enhance the Journal’s role as, in the words of an early editor, “a liv-
Chemistry ing textbook of chemistry”. In addition, they will be published in
Future Directions in Surface Chemistry electronic format via JCE Online (whose founding was also supported
Coadsorption on Surfaces by the Dreyfus Foundation). In the Journal, Viewpoints papers will
High-Pressure Surface Science take full advantage of color graphics, which will also appear in the
Monitoring Surface Chemistry at Ever-Improving electronic version. In JCE Online there also will be links to the au-
Spatial Resolution and Time Resolution
Studies of the Buried interfaces, Solid–Liquid
thors’ and other related Web sites, and video and animations when
and Solid–Solid relevant.
Achieving 100% Selectivity in Surface The Viewpoints series begins this month with “The Flexible Sur-
Reactions—the Environmental Imperative face”, a paper from the laboratories of Gabor A. Somorjai of the Uni-
Nanoparticles: Surfaces in Three Dimensions versity of California, Berkeley. Somorjai and his co-author, Günther
Rupprechter, discuss concepts related to external and internal sur-
faces and interfaces and the dynamic nature of surfaces during chemi-
cal processes. A broad overview of the most frequently used surface
Other Material on science techniques is also included, along with references for each
Surface Chemistry in This Issue technique. Somorjai and Rupprechter close with a discussion of the
present and future technological impact of surface chemistry on ca-
On the Surface 176A talysis and semiconductor devices, and the chemist’s role in surface
Mini-Activities Exploring Surface science in the coming years.
Phenomena: JCE Classroom Activity #6
Glenn T. Seaborg, Chair of the Viewpoints Editorial Board
Flying over Atoms CD-ROM 247
JCE Software Special Issue 19,
by John R. Markham

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 Günther 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 demic and industrial laboratories focusing on surface studies


have been formed in Germany (Haber, Polanyi, Farkas,
Surface science in general and surface chemistry in par- Bonhoefer), the United Kingdom (Rideal, Roberts, Bowden),
ticular have a long and distinguished history. The spontane- the United States (Langmuir, Emmet, Harkins, Taylor, Ipatief,
ous spreading of oil on water was described in ancient times Adams), and many other countries. They have helped to bring
and was studied by Benjamin Franklin. The use of surface- surface chemistry into the center of development of chemistry,
chemical processes on a large industrial scale began in the both because of the intellectual challenge to understand the
early part of the 19th century. The application of catalysis rich diversity of surface phenomena and because of its im-
started with the discovery of the platinum-surface-catalyzed portance in chemical and energy conversion technologies.
reaction of H2 and O2 in 1823 by Döbereiner. Döbereiner Up to the 1950s, studies of surfaces were mostly on the
used this reaction in his portable “flame source”, of which he macroscopic scale. Then the rise of the solid-state-device-
sold a large number. By 1835 the discovery of heterogeneous based electronics industry and the availability of economical
catalysis was complete, thanks to the studies of Kirchhoff, ultra-high-vacuum systems—developed by research in space
Davy, Henry, Philips, Faraday, and Berzelius (1). It was about sciences—provided surface chemistry with new challenges
this time that the Daguerre process was introduced for photog- and opportunities and resulted in explosive growth of the dis-
raphy. The study of tribology, which includes friction, lubrica- cipline. Clean surfaces of single crystals could be studied for
tion, and adhesion, also started around this time; this coin- the first time and the development of a large number of new
cides with the industrial revolution, as machinery with mov- techniques (cf. section on Techniques of Surface Science) from
ing parts became prevalent (although some level of under- the 1960s onwards made possible the investigation of surfaces
standing of friction appears in the work of Leonardo da Vinci). at atomic and molecular levels.
Surface-catalyzed-chemistry-based technologies first appeared As a result, macroscopic surface phenomena (adsorption,
in the period of 1860 to 1912, starting with the Deacon pro- bonding, catalysis, oxidation, and other surface reactions; dif-
cess (2HCl + 1/2 O2 → H2O + Cl2), SO2 oxidation to SO3 fusion, desorption, melting, and other phase transformations;
(Messel, 1875), the reaction of methane with steam to produce growth, nucleation, charge transport; atom, ion, and elec-
CO and H2 (Mond, 1888), ammonia oxidation (Ostwald, tron scattering; friction, hardness, lubrication) are being re-
1901), ethylene hydrogenation (Sabatier, 1902), and ammonia examined on the molecular scale. This has led to a remark-
synthesis (Haber, Mittasch, 1905–1912). Surface tension able growth of surface chemistry that has continued unin-
measurements and recognition of equilibrium constraints on terrupted up to the present. The discipline has again become
surface chemical processes led to the development of the thermo- one of the frontier areas of chemistry. The newly gained
dynamics of surface phases by Gibbs (1877). knowledge of the molecular ingredients of surface phenomena
The existence of polyatomic or polymolecular aggregates has given birth to a steady stream of high technology prod-
that lack crystallinity and diffuse slowly (gelatine and albu- ucts, including new hard coatings that passivate surfaces;
min, for example) was described in 1861 by Graham, who chemically treated glass, semiconductor, metal, and polymer
called these systems “colloids”. Polymolecular aggregates that surfaces to which the treatment imparts unique surface proper-
exhibit internal structure were called “micelles” by Nageli, ties; newly designed catalysts, chemical sensors, and carbon
and stable metal colloids were prepared by Faraday. The colloid fiber composites; surface-space-charge-based copying; and
subfield of surface chemistry gained prominence in the be- new methods of electrical, magnetic, and optical signal pro-
ginning of the 20th century with the rise of the paint industry cessing and storage. Molecular surface chemistry is being uti-
and the preparation of artificial rubbers. Studies of high-surface- lized increasingly in biological sciences.
area gas-absorber materials for gas masks and other gas-separation
technologies and investigations of the lifetime of the light bulb External Surfaces
filament led to the determination of the dissociation prob-
ability and adsorption probability of many diatomic molecules Surface Concentration
on surfaces as a function of gas pressure (adsorption isotherm) The concentration of atoms or molecules at the surface
and temperature (Langmuir 1915). The properties of chemi- of a solid or liquid can be estimated from the bulk density.
sorbed and physisorbed monolayers, adsorption isotherms, For a bulk density of 1 g/cm3 (such as ice or water), the mo-
dissociative adsorption, energy exchange, and sticking upon lecular density ρ, in units of molecules per cubic centimeter,
gas–surface collisions were studied. Studies of electrode sur- is ≈ 5 × 1022. The surface concentration of molecules σ (mol-
faces in electrochemistry led to detection of the surface space ecules/cm2) is proportional to ρ2/3, assuming cubelike pack-
charge (2). Surface diffraction of low-energy electrons was ing, and is thus on the order of 1015 molecules/cm2. Because
discovered by Davisson and Germer in 1927, and atom dif- the densities of most solids or liquids are all within a factor
fraction (helium) from surfaces, somewhat later. Major aca- of 10 or so of each other, 1015 molecules/cm2 is a good order-

162 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 tribution of products. Most catalysts are in small-particle
or molecules for most solids or liquids. Of course, surface form, including those used to produce fuels and chemicals
atom concentration of crystalline solids may vary by a factor ranging from high-octane gasoline to polyethylene.
of two or three, depending on the type of packing of atoms
Thin Films
at a particular crystal face.
Consider a monolayer of gold atoms (a layer of gold atoms
Clusters and Small Particles one atom thick) deposited on iron (Fig. 2). This film has a
All atoms in a three- or four-atom cluster are by necessity dispersion of unity, since all the atoms are on the surface. About
“surface atoms”. As a cluster grows in size, some atoms may 50 layers of gold atoms (D = 1/50) are needed to obtain the
become completely surrounded by neighboring atoms and optical properties that impart the familiar yellow color char-
are thus no longer on the “surface” (Fig. 1). We frequently acteristic of bulk gold.
describe a particle of finite size by its dispersion D, where D Thin films are of great importance to many real-world
is the ratio of the number of surface atoms to the total number problems. Their material costs are very little compared to cost
of atoms: of the bulk material, and they perform the same function
when it comes to surface processes. For example, a monolayer
D = number of surface atoms of rhodium (a very expensive metal), which contains only
total number of atoms about 1015 metal atoms per square centimeter, can catalyze
For very small particles, D is unity. As the particle grows the reduction of NO to N2 by its reaction with CO in the
and some atoms become surrounded by their neighbors, the catalytic converter of an automobile, or it can catalyze the
dispersion decreases (Fig. 1). Of course, D also depends some- conversion of methanol to acetic acid by the insertion of a
what on the shape of the particles and how the atoms are CO molecule.
packed. The dispersion is already as low as 10᎑3 for particles of Thin ordered silicon layers optimize electron transport
10-nm (100-Å) radius. in integrated electronic circuits and thin films of organic
Many chemical reactions are facilitated by surface atoms molecules lubricate our skin or the moving parts of internal
of heterogeneous catalysts. These catalysts increase the rate of combustion engines. A green leaf is a high-surface-area system
formation of product molecules and modify the relative dis- 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 deliber-
ately. 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 trans-
port 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 cata-
lysts and passivating coatings to computer-integrated circuitry
and the storage and retrieval of information.

Figure 1. Clusters of atoms with cubic packing having 8, 27, 64, Figure 2. An iron particle with one surface covered with a mono-
125, and 216 atoms. While in an 8-atom cluster all of the atoms layer of gold atoms. When it comes to surface properties such as
are on the surface, the dispersion rapidly declines with increasing adsorption or catalysis, one monolayer of atoms is all that is needed
cluster size, as shown in the lower part of the figure. to carry out the necessary chemistry.

JChemEd.chem.wisc.edu • Vol. 75 No. 2 February 1998 • Journal of Chemical Education 163


Viewpoints

Internal Surfaces—Microporous 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 pro-
cess called intercalation. Graphite will swell with water vapor
to several times its original thickness as water molecules be-
come 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 Å) in diameter. Coal and
char have porous structures, with pore diameters on the order
of 102–103 nm (103–104 Å). These materials have very large Figure 3. Alumina silicates with pores of molecular dimensions
internal surface areas, in the range of 100–400 m2 per gram (zeolites) are used as selective absorbers of gases or liquids (mo-
of solid. As this short survey shows, nature has provided us lecular sieves) and as catalysts in chemical and petroleum tech-
with many useful microporous materials; and many synthetic nologies. The figure shows a synthetic zeolite, zeolite A. The
microporous substances are used in technology, both to sepa- red spheres represent oxygen atoms and the yellow spheres rep-
resent either silicon or aluminum atoms. For each aluminum there
rate gas and liquid mixtures by selective adsorption and to
is a corresponding Na+ ion somewhere in one of the open chan-
carry out surface reactions selectively in their pores, which nels. The molecular formula of this molecular sieve is
are often of molecular dimensions. Because surface reaction Na12(Al 12Si 12O 48) ⴢ27H 2O.
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.
10᎑6 torr) and using the values M = 28 g/mol (for N2) and T
Clean Surfaces = 300 K, we obtain F ≈ 1015 molecules/cm2/s. Thus, at this
pressure the surface is covered with a monolayer of gas within
To study atomically clean surfaces, we must work under seconds, assuming that each incident gas molecule “sticks”.
so-called ultrahigh vacuum (UHV) conditions (4, 5), as the fol- For this reason the unit of gas exposure is 1.33 × 10᎑4 Pa-s
lowing rough calculation shows. We know that the concentra- (10᎑6 torr-s), which is called the Langmuir (L). Thus, a 1-L
tion of atoms on the surface of a solid is on the order of 1015 exposure will cover a surface with a monolayer amount of
cm᎑2. To keep the surface clean for 1 s or for 1 h, the flux of gas molecules, assuming a sticking coefficient of unity. At
molecules incident on the initially clean surface must therefore pressures on the order of 1.33 × 10᎑7 Pa (10᎑9 torr), it may
be less than ≈ 1015 molecules/cm2/s or ≈ 1012 molecules/cm2/s, take 103 s before a surface is covered completely.
respectively. From the kinetic theory of gases (6), the flux, F, of In practice, one usually wants to study a surface with-
molecules striking the surface of unit area at a given ambient out worrying about contamination from ambient gases. Cur-
pressure P is rent surface science techniques can easily detect contamina-
N AP tion on the order of 1% of a monolayer. This then will be
F= our operational definition of “clean”. Thus, ultrahigh vacuum
2πMRT conditions (< 1.33 × 10᎑7 Pa = 10᎑9 torr) are required to main-
or tain a clean surface for about 1 h, the time usually needed to
20
perform experiments on clean surfaces.
2.63 × 10 ⋅ P (Pa)
F (atoms/ cm2 ⋅ s) = Interfaces
M (g / mol) T
or In most circumstances, however, and certainly in Earth’s
environment, surfaces are continually exposed to gases or liq-
22 uids or placed in contact with other solids. As a result, we
2
3.51 × 10 ⋅ P (torr)
F (atoms/ cm ⋅ s) = end up investigating the properties of interfaces—that is, be-
M (g / mol) T tween solids and gases, between solids and liquids, between
solids and solids, and even between two immiscible liquids.
where M is the average molar weight of the gaseous species, Thus, unless specifically prepared otherwise, surfaces are al-
T is the temperature (in Kelvin), R is the gas constant, and ways covered with a layer of atoms or molecules from the
NA is Avogadro’s number. Substituting P = 4 × 10᎑4 Pa (3 × environment.

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


The Flexible Surface

Adsorption Techniques of Surface Science


On approaching the surface, each atom or molecule en- Over the last three decades, a large number of techniques
counters an attractive potential that ultimately will bind it have been developed to study various surface properties, in-
to the surface under proper circumstances. The process that cluding structure, composition, oxidation states, and changes
involves trapping of atoms or molecules incident on the sur- of chemical, electronic, and mechanical properties. The em-
face is called adsorption. It is always an exothermic process. phasis has been on surface probes that monitor properties
For historical reasons, the heat of adsorption ∆Hads is always on the molecular level and are sensitive enough to detect ever-
denoted as having a positive sign—unlike the enthalpy ∆H, smaller numbers of surface atoms. The frontiers of surface
which for an exothermic process would be negative according instrumentation are constantly being pushed toward detec-
to the usual thermodynamic convention. The residence time tion of finer detail: atomic spatial resolution, ever-smaller
τ of an adsorbed atom (7) is given by energy resolution, and shorter time scales. Because no single
τ = τ0 exp(∆Hads/RT) technique provides all necessary information about surface
atoms, the tendency is to use a combination of techniques.
where τ0 is correlated with the surface-atom vibration times The most commonly used techniques (Table 1) involve the
(it is frequently on the order of 10᎑12 s), T is the tempera- scattering, absorption, or emission of photons, electrons, at-
ture, and R is the gas constant. The value of τ can be 1 s or oms and ions, although some important surface-analysis tech-
longer at 300 K for ∆Hads > 69 k J/mol (16.5 kcal/mol). The niques cannot be classified this way.
surface concentration σ (in molecules/cm2) of adsorbed mol- Most surface probes require high vacuum during their
ecules on an initially clean surface is given by the product of application, which prevents their use during high-pressure
the incident flux F and the residence time τ: studies. To circumvent this restriction, UHV-compatible
σ = Fτ high-pressure cells (“environmental cells”) were developed (8).
The sample to be analyzed is first subjected to the usual high-
The surface of the material on which adsorption occurs
pressure and/or high-temperature conditions encountered
is often called the substrate. Substrate-adsorbate bonds are
during reactions in the environmental cell. Afterwards the
usually stronger than the bonds between adsorbed molecules.
sample is transferred, without exposure to air, into the evacu-
As a result, the monolayer of adsorbate bonded to the sub-
ated UHV chamber where the surface probe is located for
strate is held most tenaciously and is difficult to remove.
subsequent surface analysis (generally the sample surface
Therefore, the properties of real surfaces are usually deter-
should be characterized before and after any treatment).
mined in the presence of an adsorbed monolayer.
During the past five years, two new surface science tech-
niques in particular proved capable of obtaining molecular-
level surface information during chemical change at both low
and high ambient pressures (9, 10): scanning tunneling micros-
copy (STM) and infrared–visible sum frequency generation
(SFG) surface vibrational spectroscopy. Both of these techniques
can operate within a 14-order-of-magnitude pressure range
(10᎑10–104 torr) without significant change in signal quality
in terms of spatial or energy resolution. Using these two tech-
niques, 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 chemi-
cally treated to remove undesirable impurities from their sur-
faces. Thin films are deposited from vapor by sublimation,
sputtering, or the use of plasma-assisted chemical vapor depo-
sition. 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 al-
ways be carefully monitored.
Table 1 lists a selection of the surface-science techniques
Figure 4. Schematic diagram of a scanning tunneling microscope used most frequently in recent years to learn about the inter-
capable of operating in a pressure range from UHV to 1 atmo-
face on the atomic scale. The names of the techniques, their
sphere. The STM is located inside a high-pressure chemical reac-
tor, which is attached to a UHV surface characterization chamber.
acronyms, and brief descriptions are provided (along with
The two sections are separated by a gate valve. For surface clean- some references [11–39], if a more detailed study of the ca-
ing and analysis, a transfer system is used to move the sample pabilities and limitations of a particular technique is desired).
from the high-pressure cell to the UHV part of the apparatus (and We also indicate the primary surface information that can
vice versa). 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
Acronym Name (References) Description
Information

— Adsorption or selective Atoms or molecules are physisorbed; the amount of gas adsorbed is a Surface area, adsorption
chemisorption (11) measure of total surface area. Chemisorption of atoms or molecules yields site concentration
surface concentration of selected elements or adsorption sites.

AD Atom or helium diffraction Monoenergetic beams of atoms are scattered from ordered surfaces and Surface structure
(12 ) 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.

AES Auger electron Core-hole excitations are created, usually by 1– 10-keV incident electrons, Chemical composition
spectroscopy (12– 14 , 23, and Auger electrons of characteristic energies are emitted through a
24, 26 ) 2-electron process as excited atoms decay to their ground state. AES gives
information on the near-surface chemical composition.

AFM Atomic force microscopy Similar to STM. An extremely delicate mechanical probe is used to scan Surface structure
(15, 16 ) 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.

ELS or Electron energy loss Monoenergetic electrons ~5– 50 eV are scattered off a surface and the Electronic structure,
EELS spectroscopy (14, 17– 19 ) energy losses are measured. This gives information on the electronic surface structure
excitations of the surface and the adsorbed molecules (see HREELS).

ESCA Electron spectroscopy for Now generally called XPS (see XPS). Composition, oxidation
chemical analysis (12, 20 ) state

EXAFS Extended X-ray absorption Monoenergetic photons excite a core hole. Modulation of the absorption Local surface structure
fine structure (21, 22 ) cross-section with energy at 100– 500 eV above the excitation threshold and coordination
yields information on radial distances to neighboring atoms. The cross numbers
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 "fine-
structure" techniques.

FEM Field emission microscopy A strong electric field (on the order of V/Å ) is applied to the tip of a sharp, Surface structure
(12, 20, 21 ) 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.

FIM Field ionization microscopy A strong electric field (on the order of V/Å ) is created at the tip of a sharp, Surface structure and
(12, 20, 21, 23 ) single-crystal wire. Gas atoms, usually He, are polarized and attracted to surface diffusion
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.

FTIR Fourier transform infrared Broad-band IRAS experiments are performed, and the IR absorption Bonding geometry and
spectroscopy (13, 24 ) spectrum is deconvoluted by using a Doppler-shifted source and Fourier strength
analysis of the data. This technique is not restricted to surfaces.

HREELS High-resolution electron A monoenergetic electron beam, usually ~2– 10 eV, is scattered off a Bonding geometry,
energy loss spectroscopy surface, and the energy losses below ~0.5 eV to bulk and surface phonons surface
(12, 21, 23 ) and vibrational excitations of adsorbates are measured as a function of atom vibrations
angle and energy (also called EELS).

IRAS Infrared reflection IR photons are reflected off a surface and the attenuation of IR intensity is Molecular structure
absorption spectroscopy measured as a function of frequency yielding a spectrum of the vibrational
(21, 25 ) excitations of adsorbed molecules. Recent improvements in sensitivity allow
IRAS measurements to be made on single-crystal surfaces.

ISS Ion scattering spectroscopy Ions are inelastically scattered from a surface, and the chemical Surface structure,
(12, 13, 21, 23, 24, 26 ) composition of the surface is determined from the momentum transfer to composition
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).

LEED Low-energy electron Monoenergetic electrons below ~500 eV are elastically back-scattered Atomic and molecular
diffraction (12– 14, 20, 21, from a surface and detected as a function of energy and angle. This gives surface structure
23, 24, 26 ) information on the structure of the near surface region.

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
Acronym Name (References) Description
Information

— Neutron diffraction (27 ) Neutron diffraction is not an explicitly surface-sensitive technique, but Molecular structure
experiments on large surface-area samples have provided important
structural information on adsorbed molecules and surface phase
transitions.

NMR Nuclear magnetic NMR is not an explicitly surface-sensitive technique, but NMR data on large Chemical state
resonance (28 ) s u r f a c e - a r e a s a m p l e s ( ⱖ 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 Monoenergetic electrons of ~1– 20 keV are elastically scattered from a Surface structure,
electron diffraction surface at glancing incidence and detected as a function of angle and structure of thin films
(13, 20, 21, 23, 24 ) energy for small forward-scattering angles. Backscattering is less important
at high energies and glancing incidence is used to enhance surface
sensitivity.

SFG Sum frequency generation Similar to SHG, but the laser output is split into a visible laser beam and a Molecular structure
(29, 30 ) 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.

SHG Second harmonic A surface is illuminated with a high-intensity laser, and photons are Electronic structure,
generation (29, 31 ) generated at the second harmonic frequency through nonlinear optical molecular orientation
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.

SIMS Secondary ion mass Ions and ionized clusters ejected from a surface during ion bombardment Surface composition
spectrometry (12, 13, 21, are detected with a mass spectrometer. Surface chemical composition and
23, 24, 32 ) some information on bonding can be extracted from SIMS ion fragment
distributions.

STM Scanning tunneling The topography of a surface is measured by mechanically scanning a Atomic surface structure
microscopy (12, 13, 16, probe over a surface. The distance from the probe to the surface is
24, 33 ) measured by the probe-surface tunneling current. Angstrom resolution of
surface features is routinely obtained.

TEM Transmission electron TEM can provide surface information for carefully prepared and oriented Surface structure
microscopy (13, 24, 34, bulk samples. Real images have been formed of the edges of crystals where
35 ) 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.

TDS Thermal desorption An adsorbate-covered surface is heated, usually at a linear rate, and the Composition, heat of
spectroscopy desorbing atoms or molecules are detected with a mass spectrometer. This adsorption, surface
(21, 23, 36, 37 ) gives information on the nature of adsorbate species and some information structure
on adsorption energies and the surface structure.

TPD Temperature programmed Similar to TDS, except the surface may be heated at a nonuniform rate to Composition, heat of
desorption (21, 36, 37 ) obtain more selective information on adsorption energies. adsorption, surface
structure

UPS Ultraviolet photoemission Electrons photoemitted from valence and conduction bands are detected Valence band structure
spectroscopy as a function of energy to measure the electronic density of states near the
(12, 14, 20, 21, 23 ) surface. This gives information on the bonding of adsorbates to the surface.

XPS X-ray photoemission Electrons photoemitted from atomic core levels are detected as a function Composition, oxidation
spectroscopy (12, 13, 20, of energy. Shifts of core-level energies give information on the chemical state
21, 23, 24, 26, 38 ) environment of the atoms.

XRD X-ray diffraction (39 ) X-ray diffraction has been carried out at extreme glancing angles of Surface structure
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 organo-
metallic clusters, which provide comparisons to molecules adsorbed on
surfaces. X-ray diffraction has also given structural information on large
surface-area samples.

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 Figure 6. Side (a) and top (b) views on the Fe3O4(111) surface
attempts to optimize its coordination and “cracks” open to close structure with the spacing relaxations shown. The corresponding
the step edges. bulk values are ∆ = 0.63 Å, and d12 = d23 = 1.19Å. The A and B
layers are strongly expanded by ~ 0.46Å.

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
Surface Structure Is Different from Bulk Structure of charge neutrality, this type of surface structure is thermo-
Two dominant phenomena occurring at clean surfaces dynamically more stable than having alternating oxygen-ion
of materials distinguish their atomic structure from that in and iron-ion layers. Such an expansion at the surface is clearly
the bulk: relaxation and reconstruction (40). Upon relaxation a property of ionic solids, and future studies will prove how
of metal surfaces, the first layer of atoms moves inward, and general this type of relaxation is.
this contraction leads to a much shortened interlayer spac- Because of directionality of bonding in most solids, such
ing between the first and second layer of the surface. The contraction or relaxation at the surface moves atoms from their
more open (“rougher”) the surface, the larger the relaxation. position of optimum bonding. As a result, the atoms not only
Often, but not always, the contraction in the first layer is move perpendicular to the surface, but also parallel to the sur-
followed by a small expansion in the second layer. At rough face. This leads to the formation of new surface unit cells, a
edges, such as at stepped surfaces, the atoms at the step relax phenomenon called surface reconstruction. Perhaps the most cel-
by a large amount in order to smooth the surface irregular- ebrated example is the (7×7) surface structure that forms on the
ity. This is shown schematically in Figure 5. Si(111) crystal face. Figure 7a shows the low-energy electron dif-
At ionic surfaces, the nature of surface relaxation is very fraction (LEED) pattern from this surface. The complex unit
different. Figure 6 shows what happens at iron oxide surfaces cell has 49 different locations of surface atoms that are distin-
(41). Iron oxide, in its bulk structure, shows alternating layers guishable. The Si(100) surface shows a (2×1) reconstruction (Fig.
of oxygen ions and iron ions, where the iron ions are in tetra- 7b). It shows the formation of staggered dimers differing from
hedral or octahedral positions. At the surface, the two ions the arrangement of Si atoms in the bulk near the surface.

Figure 7. Left: Low-energy electron diffraction (LEED) pattern of the reconstructed Si(111) crystal face, exhibiting a (7×7) 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 to large (111) orientation facets in the presence of oxygen,
unit cells reconstruct to form hexagonal surface unit cells (40, and then again to smooth (110) unreconstructed surfaces in
42). This is shown in Figure 8. STM studies have imaged the presence of carbon monoxide (Fig. 11).
both reconstructed and unreconstructed surfaces (some domains Low-energy electron diffraction surface crystallography
may be unreconstructed because of contamination by adsor- studies indicate the detailed atomic level nature of such re-
bates of various types). constructions (42). When carbon is adsorbed on the Ni(100)
The water molecules at the ice surface vibrate with a much surface, it occupies fourfold hollow sites (Fig. 12). As a re-
larger amplitude (0.24 Å) than molecules in bulk ice (0.1 Å) sult of the formation of the carbon–metal chemisorption
(43). This motion causes surface “softness”, which is likely bonds, the surface metal atoms move away from the adsorption
responsible for the anomalously low friction coefficient of ice site, presumably to give more space to the carbon atom so it
(i.e., it makes ice so slippery). can sink deeper into the surface, thereby forming bonds with
Molecules at the liquid–vapor interface also show restruc- second-layer nickel atoms underneath (47). This expansion
turing, as shown in Figure 9 for the arrangement of molecules around the chemisorption site induces strain, which is relieved
at liquid alcohol surfaces (44). The alcohol molecules are ori- by rotation of the surface unit cell as shown in Figure 12.
ented with their O–H bonds pointing inward, presumably When sulfur is chemisorbed on the Fe(110) crystal face
for optimum hydrogen bonding. The alkane chains stick out (48), the S atom pulls the neighboring Fe atoms into equal
from the surface. This orientation has been readily detectable distances from the chemisorption site to form four equal Fe–S
by nonlinear laser optics sum frequency generation (SFG). bonds. The strength of these bonds pays for the weakening of
the metal–metal bonds as a result of the restructuring. When
Adsorbate-Induced Restructuring of Surfaces
NO is adsorbed on the Ni(111) surface, the molecule occupies
When the clean surfaces are covered with a near mono- a threefold hollow site, a so-called hcp (hexagonal close-packed)
layer of chemisorbed molecules, the structure of the surface hollow site. This means there is a metal atom directly under-
undergoes profound alterations. This is perhaps best shown neath the chemisorption site in the second metal layer.
in the field ion microscopy (FIM) studies carried out by Kruse Chemisorption induces an upward movement of this metal
and coworkers (45) with rhodium field emission tips (Fig. 10). atom in the second layer, and rumpling of the metal surface.
When carbon monoxide is chemisorbed on these tips, every When ethylene adsorbs on the Rh(111) surface (49), it re-
crystal face restructures as shown by the figure. This massive arranges and occupies a hollow site (in this case, again, an
restructuring is reversible if CO is removed when the surface hcp hollow site). The rearranged ethylene (which has lost a
is heated in vacuum. hydrogen) is called ethylidyne. This is shown in Figure 13.
Our new STM system (Fig. 4) is placed in an environ- The metal atoms move away from the carbon atom bound
mental cell that can be pressurized and heated to elevated to the hollow site to allow the carbon to bond to the Rh atom
temperatures. It shows surface restructuring of the Pt(110) directly underneath the carbon in the second layer. On the
surface when this surface is exposed to atmospheric pressures Pt(111) surface (50, 51), ethylene also forms an ethylidyne
of hydrogen, then oxygen, and then carbon monoxide (46). molecule—again in a threefold hollow site, but in this case
When these surfaces are heated, the surface restructures from it is an fcc hollow site. That is, there is no metal atom di-
a reconstructed ordered structure (exhibiting the so-called rectly underneath the carbon in the second metal layer. In
“missing row” reconstruction) in the presence of hydrogen, this circumstance, the surface metal atoms move inward to
presumably provide as strong a bond as possible to the carbon,
and metal–metal 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 metal–metal bonds, which are altered as a result of
the movement of the metal atoms nearest to the chemisorp-
tion site.

Figure 8. Top and side view of the Ir(100)-(1×5) surface reconstruc- Figure 9. The normal alcohols show substantial ordering at the liq-
tion. The more open square (100) lattice is reconstructed into a uid–vapor interface. The OH groups of the alcohols extend into
close-packed hexagonal overlayer, with a slight buckling as shown the liquid forming a hydrogen-bonded network, while the alkyl
in the side view. 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, Table 2. Structure Sensitivity of H2/D2 Exchange at Low
adsorbate-induced reconstruction may be very different from Pressures (~10 ᎑ 6 torr)
when only one or the other molecule is chemisorbed. This is Surface Reaction Probability
shown for benzene and CO coadsorption on the Rh(111) Stepped Pt(332) 0.9
surface (Fig. 14). When these molecules form a mixed unit
Flat Pt(111) ~10 ᎑1
cell, under the benzene the metal atoms are closer to their
bulklike configuration than under the CO molecule (52). Defect-free Pt(111) ⱕ10 ᎑3
There is rumpling of the surface that occurs because of the
differences in chemical bonding of the coadsorbed species to atom at these defect sites. Thus, the thermodynamic driving
the substrate metal atoms. When benzene adsorbs alone it is force for dissociation is certainly greater at these sites, which
bent. Four of the C atoms are in one type of surface site while can explain their enhanced bond-breaking activity.
two of the others are in different types of surface sites. When It is difficult to understand, however, that these same
coadsorption occurs, the benzene molecule is flattened out. strongly adsorbing sites are also very active sites for catalysis.
This is shown in Table 2. The reaction probability of H2/D2
Rough Surfaces Do Chemistry exchange on stepped surfaces is near unity at low pressures
Surface irregularities, steps, and kinks are very effective on a single scattering event, whereas it is below the detection
for breaking adsorbate chemical bonds and in catalysis as well. limit (< 10᎑3) on the flat (111) crystal face, as shown by mo-
This is best known by the temperature-programmed desorption lecular beam surface scattering studies (54). How is it possible
(TPD) of H2 from flat, stepped, and kinked surfaces of Pt that the strongly adsorbing step sites, where H has a long
(Fig. 15). H2 desorbs at maximum rates at the highest tempera- residence time because of its high binding energy, are also
ture from kink sites, then at somewhat lower temperatures from the sites of rapid reaction turnover?
step sites, and at even lower temperatures from flat (111) ter- One possible explanation is that the strongly adsorbed
races (53). This indicates higher heats of adsorption of the H hydrogen restructures the surface near the step, thereby creat-

Figure 10. Field ion micro-


graphs (image gas: Ne; T
= 85 K) of a (001)-oriented
Rh tip (top left ) before and
(bottom left ) after reaction
with 10᎑4 Pa CO for 30 min
at 420 K . The stereo-
graphic projections at the
right demonstrate the
change in morphology from
nearly hemispherical to po-
lygonal. 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) un-
der atmospheric pressures: Top: Topo-
graphic image of the surface in hydro-
gen after heating to 425 K for 5
hours, showing (n×1) missing-row re-
construction randomly nested. Vertical
range: ∆ z = 10 Å . Center: Topo-
graphic image of the surface in oxy-
gen after heating to 425 K for 5
hours; ∆ z = 25 Å . Bottom: Topo-
graphic image of the surface in car-
bon monoxide after heating to 425 K
for 4 hours; ∆z = 42 Å.

ing the active site for the catalytic exchange process. At the and organometallic chemistry. The vibrational spectrum of
low pressures of these molecular beam scattering experiments, chemisorbed ethylidyne is nearly identical to that of the organo-
the low coverages keep the structure of the flat part of the metallic cluster Os3 CCH3 , which contains three metal atoms.
surface unaltered. The C–C bond distance (1.45 Å) is slightly less than the
Similarly, stepped Ni surfaces dehydrogenate C2H4 at single carbon–carbon bond length of 1.54 Å (0.154 nm), as
much lower temperatures (< 150 K) than the (111) face of in cluster compounds. Thus, the surface chemical bond of
Ni (~230 K). The more open (111) and (211) crystal faces chemisorbed ethylene can, as a first approximation, be viewed
of Fe (Fig. 16) are several orders of magnitude more active as a clusterlike bond that contains at least three metal atoms
for NH3 synthesis than the close-packed Fe (110) crystal face, (56). The C–C bond order present in gaseous ethylene is
which showed no detectable reaction rate ( 55). reduced from two to nearly one upon chemisorption. This
reduction in bond order of alkenes and alkynes upon chemi-
Clusterlike Bonding of Adsorbed Molecules
sorption on metal surfaces is commonly observed, indicating
When ethylene chemisorbs at ~ 300 K on the (111) crystal charge transfer from the molecules into the metal.
faces of various transition metals (Pt, Rh, Pd), it chemically Benzene usually chemisorbs on metals with its ring parallel
rearranges to form the molecule-surface compound shown to the surface (although it may adsorb in a different configura-
in Figure 13. Its structure as determined by LEED-surface tion when it loses hydrogen). Because of charge transfer to the
crystallography is very similar to multinuclear organometal- metal, C–C bond elongations occur with respect to the sym-
lic complexes such as Os3 CCH3 or Co3 (CO)9 CCH3. The metry of the adsorption site. The ring may even bend, with
rearranged ethylene, which has also lost a hydrogen, is called two of the opposing carbon atoms closer to the metal surface
ethylidyne and belongs to the alkylidyne group (species of the than the other four carbon atoms. Distortions and elonga-
formula CnH2n᎑1), a common substituent in surface chemistry tions of C–C 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 chemi-
sorbed 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 π-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 nearest-
neighbor metal atoms of the substrate participate in the bond-
ing. There is evidence that the atoms at next-nearest-neigh-
bor sites change their location when chemisorption occurs,
Figure 12. Carbon-chemisorption-induced restructuring of the moving either closer or further away from the chemisorp-
Ni(100) surface. tion 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 adsorption-
induced distortion in the top metal
layers pulls the nearest neighbor
metal atoms up out of the surface
plane.

The Flexible Surface The flexible-surface model explains why rough surfaces
As a result of all these studies that indicate adsorbate- or defect sites at surfaces are so active in surface chemistry.
induced restructuring and clusterlike bonding, the new model Bond breaking and catalysis most frequently occur at low
of the surface which has been adopted is the so-called “flex- coordination sites such as steps and kinks or at defect sites
ible surface” (53). In the past it was assumed that the metal such as oxygen vacancies in oxide surfaces. The lower the co-
atoms at the surface are rigid and occupy equilibrium sites ordination of metal atoms (the fewer nearest neighbors), the
dictated by the bulk unit cell. On adsorption, their location more easily they restructure to optimize the surface adsorp-
would not be altered. Instead, the flexible surface is one where tion bond. Thus, rough surfaces or atoms at steps move more
the metal atoms move into new sites, dictated by the chemi- readily, and of course small clusters of atoms where the co-
sorption bond so as to optimize that bond: upon adsorption ordination is much reduced are the most flexible. It is not
the surface restructures, thereby creating the active sites for surprising therefore that we use nanoclusters in the field of
surface chemical processes. 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 bond-
ing 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 auto-
mobile 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
Molecular surface chemistry contributed to the devel-
opment 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
Figure 14. The coadsorbed surface structure of benzene and car- future impact on catalysis and semiconductor devices.
bon monoxide on the Rh(111) crystal face as obtained by low- Since the early 1970s, molecular surface chemistry has
energy electron diffraction surface crystallography. made significant contributions to our understanding of catalyst-

172 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 C–O bond by charge transfer.
Methanol production was found to occur through the hydro-
genation of undissociated CO2 or CO (which reaction domi-
nates 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 struc-
ture of platinum particles [(111) or (100) orientation] the
product distribution could be altered. Bimetallic platinum-
based catalysts (Pt–Re, Pt–Ir, and Pt–Sn) have also been in-
vestigated by surface-science studies. These studies have con-
tributed greatly to their optimization in this important high-
octane 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) crys-
tal 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 syn-
thesis of ammonia is some 500 times lower than on the (111)

Figure 16. Schematic representations of the idealized surface struc-


tures of the (111), (211), (100), (210), and (110) orientation of
Figure 15. Thermal desorption spectra of hydrogen from a flat iron single crystals (the coordination of each surface atom is indi-
(111), a stepped (557), and a kinked (12,9,8) surface of plati- cated). The bar diagram shows the corresponding activity of the
num. 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 bond- interfaces that maintain adhesion under changing ambient
ing modifiers in aiding the dissociation of N2, as well as weak- conditions (temperature, humidity, etc.). As the insulating
ening the bonding of ammonia to the iron surface to inhibit oxide is replaced by a polymer with a smaller dielectric con-
product poisoning, has been uncovered. The role of alumina stant, the study of metal–polymer interfaces becomes a fron-
as a structure modifier, aiding the restructuring of iron par- tier area of the science of semiconductor surface technology.
ticles to possess crystal faces most active for ammonia synthe- Other important surface technologies developed in recent
sis [(111) and (211)], has been proven by surface-science stud- years with the help of molecular surface chemistry should
ies. As a result, a new generation of catalysts with superior ac- also be mentioned. Air separation to oxygen and nitrogen was
tivity could be prepared for this important industrial process. accomplished with the help of molecular sieves because of
Environmentally important catalytic processes have be- their higher heat of adsorption for N2 (~ 7 kcal/mol) than
come the focus of rapid development in recent years. None for O2 (~ 3 kcal/mol), thereby preferentially releasing oxygen.
of them is more important than the 3-way catalytic converter Conversely, microporous carbon engineered to have bimodal
utilized to clean automobile exhaust. It utilizes Pt, Pd, Rh, distribution of pores adsorbs the smaller O2 (3.46 Å) in its
and cerium oxide as an important promoter (60). The chal- small pores in preference to N2 (3.64 Å) and preferentially
lenge is to fully oxidize unburned hydrocarbons and CO to releases nitrogen. The magnetic disc drive provides information
CO2 while reducing NOx to N2 under all conditions of en- storage in computers by nanoscale tracking of a magnetic thin
gine use: cold start, steady-state operation, and using a broad film that is coated by an atomically smooth carbon deposit
range of air and fuel mixtures. This is achieved with the help and lubricated by a drop of fluoro-ether lubricant. The optical
of one of the most successful chemical sensors, the oxygen fiber operates on total internal reflection in glass by the
detector ( λ-probe) on automobiles. It helps to adjust the air- cladding that is made by another glass of different composition
to-fuel ratio of the mixture entering the internal combustion and with a smaller refractive index. Such glass structures are
engine and to optimize the efficiency of the 3-way catalytic produced using chemical vapor deposition of oxides in an
converter. This technology works well on the present-day appropriate sequence. Diamond films are produced by appli-
automobile. The development of lean-burning, more fuel- cation of a high-energy plasma of methane and hydrogen or
efficient cars presents new challenges to surface chemistry and by chemical vapor deposition to provide a chemically inert,
the technology used to clean automobile exhaust. extremely hard coating with high thermal conductivity, that
Semiconductor-based technologies are at the heart of com- withstands chemical and mechanical attacks superbly (61).
puter manufacturing. The fabrication of microelectronic circuits Adhesives and the contact lens are important surface technolo-
often involves layer-by-layer deposition of semiconductor (Si, gies using polymers that provide controlled adhesion and
GaAs, etc.), metal (Al, Cu, etc.), and insulator (SiO2, polymer) some degree of biocompatibility, respectively.
thin films, in various configurations. The film thickness of
each of these materials is presently in the 103–104-Å range, Future Directions in Surface Chemistry
and these layers alternate in both two and three dimensions.
Fabrication of these layers is carried out by surface processes Coadsorption on Surfaces
using chemical vapor deposition, sublimation, or sputter Most studies of modern surface chemistry focus on single
deposition from a radio-frequency plasma. Nucleation and adsorbate systems, atoms, or molecules and investigate structure
growth mechanisms are monitored by surface-science techniques and bonding, adsorption, diffusion, and desorption dynamics
such as reflection high-energy electron diffraction (RHEED) as a function of temperature. In most chemical reactions, how-
and electron microscopy. We shall look at two problems of ever, two or more reactants are utilized. Oxygen, hydrogen,
semiconductor device technology that are currently the fo- and water are the coadsorbates present most frequently, although
cus of intense surface-science studies. mixtures of organic molecules are adsorbed during hydro-
The first is that insulating gate oxides for “metal oxide carbon conversion reactions. Catalytic systems always use
semiconductor field effect transistors” (MOSFET) are produced additives that accelerate the reaction rate or improve selec-
by oxidizing silicon to SiOx. Both the oxygen-to-silicon ratio tivity. Potassium, which is an electron donor to iron, weakens
and the thickness of the oxide are important process vari- the bonding of the product molecule, ammonia (NH3 ),
ables, as they control the device’s performance. The gate oxides thereby accelerating its desorption from the surface. Con-
must become thinner, their surfaces or interfaces must be versely, potassium increases the bonding of carbon monoxide
atomically smooth, and their impurity concentrations must (an electron acceptor) to iron and other transition metals,
be minimized in order to increase the speed of electron trans- leading to the dissociation of the molecule (62). As a result,
port and device reliability. the rate of production of hydrocarbons by the hydrogenation
The second problem is that the chemical and mechanical of CO is greatly accelerated. In the future, more studies involv-
integrity of the metal-insulator interfaces can be compromised ing coadsorbed atoms and molecules must be pursued to un-
by water vapor or by the chemical attack of impurities segre- cover their influence on surface chemistry as a consequence of
gating at the interface (e.g., alkali atoms, carbon, oxygen). their interactions, adsorbate–adsorbate and adsorbate-substrate.
When this happens, the adhesion of the insulator oxide to
the metal is altered and delamination occurs. “Trap” states High-Pressure Surface Science
that arise at the Si–SiO2 interface from Si atoms with coordina- Increasing the coverage of adsorbates often enhances restruc-
tion numbers other than 4 are another problem because they turing of the substrate and changes the surface chemistry. The
can trap charge at the interface. All these changes of chemical easiest way to increase coverage is to increase the reactant pres-
and mechanical properties at the interface can have very del- sure, since pressure is proportional to coverage (adsorption iso-
eterious effects on the electrical properties. It is essential that therm). When CO adsorbs on platinum surfaces it occupies
we learn how to fabricate chemically stable insulator–metal mostly top and bridge sites, with its C=O bond perpendicular

174 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 Acknowledgment


10 torr a high coverage incommensurate overlayer forms, and
at even higher pressures (≥ 100 torr) the formation of platinum This work was supported by the Director, Office of En-
carbonyl clusters with CO-to-Pt-ratio greater than one can be ergy Research, Office of Basic Energy Sciences, Materials Sci-
detected. This pressure-dependent change in surface chemistry ences Division, of the U.S. Department of Energy under
is driven by the formation of strong CO–Pt bonds that in- Contract No. DE-AC03-76SF00098.
duce the weakening and then the breaking of metal–metal
bonds to produce a thermodynamically more stable surface Literature Cited
structure. Newly available techniques permit atomic and mo- 1. Berzelius, J. Jahres-Bericht über die Fortschritte der Physischen
lecular studies of external surfaces in the presence of high- Wissenschaften (Tübingen), 1836, 15.
2. Thomas, J. M. Michael Faraday and the Royal Institution; IOP: Bristol,
pressure gas or a liquid at the interface. These include the 1991.
scanning tunneling and atomic force microscopes (STM and 3. Csicsery, S. M. In Zeolite Chemistry and Catalysis; Rabo, J. A., Ed.;
AFM) and sum frequency generation (SFG)–surface vibra- ACS Monographs, Vol. 171; American Chemical Society: Washing-
tional spectroscopy (10, 46). In the future these techniques ton, DC, 1976; Chapter 1.
4. O’Hanlon, J. F. A User’s Guide to Vacuum Technique, 2nd ed.; Wiley:
and others will be used to carry out in situ studies of mo- New York, 1989.
lecular surface chemistry at high reactant pressures and at high 5. Klauber, C. In Surface Analysis Methods in Materials Science; Springer Se-
temperatures. ries in Surface Sciences, Vol. 23; O’Connor, D. J.; Sexton, B. A.; Smart,
R. S. C., Eds.; Springer: Berlin, 1992; pp 67–76.
Monitoring Surface Chemistry at Ever-Improving 6. de Boer, J. H. The Dynamical Character of Adsorption; Oxford Uni-
Spatial Resolution and Time Resolution versity Press: New York, 1968.
7. Tompkins, F. C. Chemisorption of Gases on Metals; Academic: New
STM and AFM provide spatial resolution of surfaces and York, 1978.
surface species on the nanometer scale. There is a continuing 8. Cabrera, A. L.; Spencer, N. D.; Kozak, E.; Davies, P. W.; Somorjai,
need to develop spectroscopic techniques that have the same G. A. Rev. Sci. Instrum. 1982, 53, 1888.
9. Somorjai, G. A.; Rupprechter, G. In Dynamics of Surfaces and Reac-
resolution because they will provide means to study and ma- tion Kinetics in Heterogeneous Catalysis; Studies in Surface Science and
nipulate surfaces on that spatial scale. Increased time resolu- Catalysis Series, Vol. 109; Froment, G. F.; Waugh, K. C., Eds.;
tion will permit us to monitor the motion of surface atoms and Elsevier: Amsterdam, 1997; p 35.
molecules, their diffusion, rotation, vibrational and electronic 10. Su, X., et al. Faraday Discuss. 1996, 105, 263.
excitation, and their reaction dynamics. 11. Greg, S. J.; Sing, K. S. W. Adsorption, Surface Area, and Porosity; Aca-
demic: New York, 1967.
Studies of the Buried Interfaces, Solid–Liquid and 12. Hudson, J. B. Surface Science: An Introduction; Butterworth-
Heinemann: Boston, 1992.
Solid–Solid 13. MacDonald, R. J.; King, B. V. In Surface Analysis Methods in Materi-
Techniques that open up high-pressure surface chemis- als Science; Springer Series in Surface Sciences, Vol. 23; O’Connor,
D. J.; Sexton, B. A.; Smart, R. S. C., Eds.; Springer: Berlin, 1992;
try also permit molecular studies of the buried interfaces. This Chapter 5. Also see Chapters 3, 6, 10–14 in this volume.
will result in rapid developments in molecular phenomena 14. Ertl, G.; Küppers, J. Low Energy Electrons and Surface Chemistry; VCR:
at solid–liquid (64) and solid–solid interfaces, including elec- Weinheim, 1985.
trochemistry, biology, and tribology (friction, lubrication, 15. Binnig, G.; Quate, C. F.; Gerber, C. Phys. Rev. Lett. 1986, 56, 930.
wear) (65). 16. Hansma, P. K.; Elings, V. B.; Marti, O.; Bracker, C. E. Science 1988,
242, 157.
Achieving 100% Selectivity in Surface Reactions— 17. Gasser, R. P. H. An Introduction to Chemisorption and Catalysis by
Metals; Oxford University Press: New York, 1985.
the Environmental Imperative 18. Richardson, N. V.; Bradshaw, A. M. In Electron Spectroscopy: Theory,
In most surface catalyzed reactions we desire to obtain Techniques and Applications, Vol. 4; Brundle, C. R.; Baker, A. D., Eds.;
Academic: New York, 1981; pp 153–193. Also see Chapters 1, 2, and
only one product, although the formation of other chemicals is 5.
also thermodynamically allowed. We need to understand cata- 19. Ibach, H.; Mills, D. L. Electron Energy Loss Spectroscopy and Surface
lytic selectivity, how to obtain 100% selectivity to avoid the Vibration; Academic: New York, 1982.
formation of undesirable molecules that often lead to sepa- 20. Prutton, M. Surface Physics; Oxford University Press: New York, 1975.
ration problems, pollution, or catalytic deactivation. 21. Woodruff, D. P.; Delchar, T. A. Modern Techniques of Surface Science;
Cambridge Solid State Science Series; Cambridge University Press:
Nanoparticles: Surfaces in Three Dimensions New York, 1986.
22. Heald, S. M. In X-ray Absorption; Chemical Analysis, Vol. 92;
Particles with dispersions between 1 and 0.1 represent Koningsberger, D. C.; Prins, R., Eds.; Wiley: New York, 1988.
transition between single atoms and molecules and the bulk 23. Somorjai, G. A.; Van Hove, M. A. In Absorbed Monolayers on Solid
Surfaces ; Structure and Bonding Series, Vol. 38; Dunitz, J. D.;
solid. They have many surprising properties as their electronic Goodenough, J. B.; Hemmerich, P.; Ibers, J. A.; Jorgensen, C. K.;
/
structure, atomic structure, and phase diagram—and as a con- Neilands, J. B.; Reinen, D.; Williams, R. J. P., Eds.; Springer: Berlin,
sequence, their surface chemistry—change with particle size. 1979; Chapter 4.
The fabrication of nanoparticles of uniform size (66, 67) and 24. Roberts, N. K. In Surface Analysis Methods in Materials Science;
their study is one of the intellectual frontiers of modern sur- Springer Series in Surface Sciences, Vol. 23; O’Connor, D. J.; Sex-
ton, B. A.; Smart, R. S. C., Eds.; Springer: Berlin, 1992; Chapter
face chemistry. 8, pp 187–201.
Molecular surface chemistry is one of the most rapidly 25. Willis, R. F.; Lucas, A. A.; Mahan, G. D. In Adsorption at Solid State
developing branches of chemistry. It is an intellectual frontier Surfaces. The Chemical Physics of Solid Surfaces and Heterogeneous
of the discipline with enormous potential to develop surface Catalysis, Vol. 2; King, D. A.; Woodruff, D. P., Eds.; Elsevier: New
York, 1983.
technologies that improve our quality of life, create employ- 26. Somorjai, G. A.; Van Hove, M. A. In Investigations of Interfaces and Sur-
ment, and create wealth. It will remain a rapidly advancing faces and Interfaces, Part B, Vol. IXB; Rossiter, B. W.; Baetzold, R. C.,
frontier area for many years to come. 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 Ro- 47. Gauthier, Y.; Baudoing-Savois, R.; Heinz, K.; Landskron, H. Surf.
tation; Springer Tracts in Modern Physics, Vol. 101; G. Höhler, Ed.; Sci. 1991, 251, 493.
Springer: Berlin, 1983; pp 1–84. 48. Shih, H. E.; Jona, F.; Marcus, P. M. Phys. Rev. Lett. 1981, 46, 731.
28. Barrie, P. J.; Klinowski, J. Prog. Nucl. Magn. Reson. 1992, 24, 91. 49. Wander, A.; Van Hove, M. A.; Somorjai, G. A. Phys. Rev. Lett. 1991,
29. Shen, Y. R. Nature 1989, 337, 519. 67, 626.
30. Shen, Y. R. Annu. Rev. Phys. Chem. 1989, 40, 327. 50. Cremer, P. S.; Stanners, C.; Niemantsverdriet, J. W.; Shen, Y. R.;
31. Richmond, G. L.; Robinson, J. M.; Shannon, V. L. Prog. Surf. Sci. Somorjai, G. A. Surf. Sci. 1995, 328, 111.
1988, 28, 1. 51. Cremer, P. S.; Su, X.; Shen, Y. R.; Somorjai, G. A. J. Am. Chem. Soc.
32. MacDonald, R. J.; King, B. V. In Surface Analysis Methods in Materials 1996, 118, 2942.
Science; Springer Series in Surface Sciences, Vol. 23; O’Connor, D. J.; 52. Van Hove, M. A.; Lin, R. F.; Somorjai, G. A. J. Am. Chem. Soc. 1986,
Sexton, B. A.; Smart, R. S. C., Eds.; Springer: Berlin, 1992. 108, 2532.
33. Binnig, G.; Rohrer, H. Rev. Mod. Phys. 1987, 59, 615. 53. Somorjai, G. A. Langmuir 1991, 7, 3176.
34. Cowley, J. M. Prog. Surf. Sci. 1986, 21, 209. 54. Salmeron, M.; Gale, R. J.; Somerjai, G. A. J. Chem. Phys. 1979, 70,
35. Thomas, G. Ultramicroscopy 1986, 20, 239. 2807.
36. Yates, J. T. In Solid State Physics: Surfaces; Methods of Experimental 55. Somorjai, G. A.; Materer, N. Top. Catal. 1994, 1, 215.
Physics Series, Vol. 22; Park, R. L.; Lagally, M. G., Eds.; Academic: New 56. Van Hove, M. A.; Bent, B.; Somorjai, G. A. J. Phys. Chem. 1988, 92,
York, 1985; Chapter 8, pp 425–464. 973.
37. Madix, R. J. In Chemistry and Physics of Solid Surfaces, Vol. 2; 57. Somorjai, G. A. Chem. Rev. 1996, 96, 1223.
Vanselow, R., Ed.; CRC: Boca Raton, FL, 1979; pp 63–72. 58. Strongin, D. R.; Carrazza, J.; Bare, S. R.; Somorjai, G. A. J. Catal.
38. Moulder, J. F.; Stickle, W. F.; Sobol, P. E.; Bomben, K. D. Handbook of 1987, 103, 213.
X-Ray Photoelectron Spectroscopy; Perkin Elmer: Eden Prairie, MN, 1992. 59. Spencer, N. D.; Schoonmaker, R. C.; Somorjai, G. A. J. Catal. 1982,
39. Robinson, I. K.; Tweet, D. J. Rep. Prog. Phys. 1992, 55, 599. 74, 129.
40. Somorjai, G. A. Introduction to Surface Chemistry and Catalysis; Wiley: 60. Taylor, K. C. Catal. Rev.—Sci. Eng. 1993, 35, 457.
New York, 1994. 61. Perry, S. S.; Somorjai, G. A. J. Vac. Sci. Technol. A 1994, 12, 1513.
41. Barbieri, A.; Weiss, W.; Van Hove, M. A.; Somorjai, G. A. Surf. Sci. 62. Crowell, J. E.; Tysoe, W. T.; Somorjai, G. A. J. Phys. Chem. 1986,
1994, 302, 259. 89, 1598.
42. Starke, U.; Van Hove, M. A.; Somorjai, G. A. Prog. Surf. Sci. 1994, 63. Su, X.; Cremer, P. S.; Shen, Y. R.; Somorjai, G. A. Phys. Rev. Lett.
46, 305. 1996, 77, 3858.
43. Materer, N.; Starke, U.; Barbieri, A.; Hove, M. A. V.; Somorjai, G. 64. Somorjai, G. A. Surf. Sci. 1995, 335, 10.
A. Surf. Sci. 1997, 381, 190. 65. Perry, S. S.; Somorjai, G. A.; Mate, C. M.; White, R. L. Tribol. Lett.
44. Stanners, C. D.; Du, Q.; Chin, R. P.; Cremer, P.; Somorjai, G. A.; 1995, 1, 233.
Shen, Y.-R. Chem. Phys. Lett. 1995, 232, 407. 66. Jacobs, P. W.; Wind, S. J.; Ribeiro, F. H. Somorjai, G. A. Surf. Sci.
45. Kruse, N.; Gaussman, H. Surf. Sci. 1992, 266, 51. 1997, 372, L249.
46. McIntyre, B. J.; Salmeron, M. B.; Somorjai, G. A. Rev. Sci. Intrum. 67. Eppler, A.; Rupprechter, G.; Guczi, L.; Somorjai, G. A. J. Phys. Chem.
1993, 64, 687. B 1997, 101, 9973–9977.

Gabor A. Somorjai Günther Rupprechter


University of California at Berkeley University of California at Berkeley
Department of Chemistry and Department of Chemistry and
Material Sciences Division of Material Sciences Division of
Berkeley National Laboratory Berkeley National Laboratory

Ph.D., Chemistry, 1960, University Ph.D., Chemistry, 1996, Leopold


of California at Berkeley Franzens University, Innsbruck
B.S., Chemical Engineering, B.S., Chemistry, 1992, Leopold
1956, Technical University, Franzens University, Innsbruck
Budapest, Hungary

Gabor A. Somorjai is one of the most influential scientists Günther Rupprechter received his Ph.D. in 1996 under the
in the area of surface chemistry today. He has received nu- supervision of Konrad Hayek on “Microstructural and mor-
merous awards, including the 1997 Von Hippel Award which phological changes on epitaxially grown noble metal catalyst
he was awarded on December 3, 1997. His influence on sur- particles upon oxidation and reductive activation.” At present
face chemistry can be seen in more than 750 papers pub- he is a postdoctoral fellow at the University of California at
lished on surface science, heterogeneous catalysis, and solid Berkeley working with Gabor A. Somorjai. Rupprechter’s
state chemistry. Somorjai has educated more than 90 Ph.D. current research interests include structure–activity correla-
students and collaborated more than 110 postdoctoral sci- tions in heterogeneous catalysis, fabrication of well-faceted
entist of whom Günther Rupprechter is one. The research polyhedral nanocrystals on oxidic supports by epitaxial
interests of his group include molecular studies of the struc- growth and electron beam lithography, and analysis of sur-
ture and bonding of surfaces, surface science of heterogenous face structure and surface composition. He also has received
catalysis, and molecular studies of polymer surfaces and po- a number of awards and fellowships.
lymerization. During the past 30 years Somorjai has made
significant contributions in surface science, new surface in-
strumentation, and catalysis.

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

You might also like