You are on page 1of 25

Surface Science 500 (2002) 879–903

www.elsevier.com/locate/susc

Organic functionalization of group IV semiconductor


surfaces: principles, examples, applications, and prospects
Stacey F. Bent *
Department of Chemical Engineering, Stanford University, Stanford, CA 94305-5025, USA
Received 5 July 2000; accepted for publication 22 June 2001

Abstract
Organic functionalization is emerging as an important area in the development of new semiconductor-based materials
and devices. Direct, covalent attachment of organic layers to a semiconductor interface provides for the incorporation of
many new properties, including lubrication, optical response, chemical sensing, or biocompatibility. Methods by which
to incorporate organic functionality to the surfaces of semiconductors have seen immense progress in recent years, and in
this article several of these approaches are reviewed. Examples are included from both dry and wet processing envi-
ronments. The focus of the article is on attachment strategies that demonstrate the molecular nature of the semicon-
ductor surface. In many cases, the surfaces mimic the reactivity of their molecular carbon or organosilane counterparts,
and examples of functionalization reactions are described in which direct analogies to textbook organic and inorganic
chemistry can be applied. This article addresses the expected impact of these functionalization strategies on emerging
technologies in nanotechnology, sensing, and bioengineering. Ó 2001 Elsevier Science B.V. All rights reserved.

Keywords: Single crystal surfaces; Silicon; Germanium; Diamond; Chemisorption; Alkenes; Infrared absorption spectroscopy;
Scanning tunneling microscopy

1. Introduction ubiquitous. Semiconductor-based devices can now


be found in everything from automobiles and
The microelectronics revolution that was started home appliances, to means of communication and
fifty years ago has grown into an industry that equipment in our doctor’s office.
drives much of the world’s technology today. It The rapidity with which microelectronics tech-
is responsible for a nearly $1 trillion worldwide nologies have advanced is unprecedented. For
electronics market. Microelectronics use semicon- example, the number of electronic components on
ductor materials as their basic building block. a single silicon chip doubles every 18 months, and
These semiconducting solids––such as silicon, computer memory that used to cost thousands
gallium arsenide, and germanium––have become of dollars can now be fabricated for only pennies.
The widespread availability of microelectronic-
controlled machines depends upon these revolu-
*
Tel.: +1-650-723-0385; fax: +1-650-723-9780. tionary advances. Many of these advances
E-mail address: stacey.bent@stanford.edu (S.F. Bent). have resulted from miniaturization of the most

0039-6028/01/$ - see front matter Ó 2001 Elsevier Science B.V. All rights reserved.
PII: S 0 0 3 9 - 6 0 2 8 ( 0 1 ) 0 1 5 5 3 - 9
880 S.F. Bent / Surface Science 500 (2002) 879–903

fundamental building blocks of electronic devices, the semiconductor wafer. However, with the rapid
in which more and more active units are packed miniaturization of devices, understanding of the
into smaller and smaller areas on the semicon- atomic-level phenomena underlying these pro-
ductor chip. Integrated circuits now contain nearly cesses becomes even more critical. The sizes that
one billion transistors on a chip less than a square microelectronics are projected to reach in the near
inch in size [1,2]. future means that a given material in a device
As we enter a new century, the sizes of micro- might be just a few layers thick. As a result, much
electronic devices have gotten so small that the of the functionality of a device will be due to
active elements now approach molecular dimen- physical processes that occur within a few molec-
sions. For example, it is predicted by an industry ular layers of some interface, such as a buried in-
guideline known as the Technology Roadmap for terface, e.g. between Si/SiO2 or Si/metal, or an
semiconductors [3] that by the year 2014, lateral outer surface exposed to ambient. For this reason,
device dimensions will reach 35 nm (350 A ); ver- research on the interfacial chemistry at semicon-
tical dimensions will be even smaller. Because ductor surfaces has seen a tremendous amount of
molecules themselves are on the order of a few growth over the past decade, and promises to be
tenths of nanometers in size, it follows that future an increasingly important field in the future.
technologies increasingly will depend upon the The importance of atomic level surface chemis-
functionality of individual molecules and atoms. try is highlighted in the growing field of organic
Surface phenomena have always been a functionalization of semiconductors. An area that
cornerstone of the microelectronics industry. Pro- has been researched extensively over the past few
cesses such as epitaxy [4], chemical vapor deposi- years, surface functionalization or organic modi-
tion, etching, oxidation and passivation [5], which fication is the process of depositing layers of or-
are used routinely in the industry, involve chemical ganic molecules (i.e. those that contain carbon) at
or physical processes occurring at the surface of semiconductor surfaces. Fig. 1 illustrates a semi-

Fig. 1. Illustration of organic molecules attached to a silicon substrate in the initial stages of film deposition. Layered organic/
semiconductor materials are being explored for possible application in sensor development, molecular electronics, and biocompatible
implants.
S.F. Bent / Surface Science 500 (2002) 879–903 881

conductor surface upon which organic molecules ventional microelectronics processing, such as for
have been attached. The motivation to deposit new-generation dielectric materials for metal in-
organic layers on a semiconductor stems from a terconnect isolation, or for surface passivation and
desire to impart some property of the organic protection.
material to the semiconductor device. Over the Fortunately, the surfaces of most semiconduc-
same time period during which semiconductor tors of interest for microelectronics have charac-
(typically Si, Ge, GaAs) based devices were un- teristics that enable organic molecules to be
dergoing rapid size scaling, revolutionary advances attached by a number of different chemical reac-
also were made in the development of new solid tions. These attachment chemistries provide an
state organic materials. Organic molecules com- insightful view of the reactivity of semiconductor
prise more than 95% of all known chemical com- surfaces. They also demonstrate how the wealth of
pounds. Carbon can form a myriad of molecules chemical understanding that has been developed
differing in shape, size and composition [6]. For over the past century within the field of organic
example, the organic molecules attached to the chemistry can now be used to manipulate semi-
surface in Fig. 1 will exhibit certain properties conductor surfaces to impart properties useful for
based on the types and specific groupings of atoms a variety of applications. The focus of this article is
in the molecules. As a result, organic materials on several different approaches that have been
offer engineers great flexibility in designing and taken to functionalize semiconductor surfaces with
creating unique molecular properties, that can organic molecules. The list of attachment reactions
then be exploited to provide new capabilities in covered in this article is not intended to be com-
optical, electronic, and mechanical function as well prehensive, but rather a selected group of systems
as in chemical and biological activity. has been chosen for the level of insight they pro-
An especially promising approach is to combine vide into the molecular nature of semiconductor
organic materials with conventional semicon- surfaces. The article will use these examples to il-
ductor materials [7]. Microelectronics technology lustrate some general principles describing the
provides an opportunity to create hybrid devices chemistry at semiconductor surfaces. For more
exploiting the best properties and features of both complete technical reviews on the subject of at-
organic and inorganic materials. As an example, tachment reactions at semiconductor surfaces, the
consider a chemical or biological sensor built with reader is referred to Refs. [8–11].
a mixed organic/semiconductor design. Organic
molecules such as those shown in Fig. 1 can be
deposited on a silicon substrate. The organic layer 2. The silicon surface
may be terminated with a variety of end groups
that respond to different chemical or biological We will begin with an investigation into the
stimuli. A ‘‘sensing’’ response occurs if a species of structure of the silicon surface. Silicon is the pre-
interest binds to the end group, causing trans- dominant semiconductor material used in the
duction of signal within the organic layer. If this microelectronics industry, due in part to several
signal can be coupled into the silicon substrate, important properties. Silicon can be produced
then all of the capabilities of silicon-based micro- in single crystalline form at better than
electronic circuitry––such as signal amplification, 99.999999999% purity, it forms an excellent oxide
processing and storage––become available. The at its surface, and its electronic properties can be
end product is a chip-based chemical or biological tuned dramatically by substituting only a small
sensor. fraction of silicon atoms in the lattice with another
Hybrid organic/semiconductor materials also element in a process called ‘‘doping’’. Silicon is a
are being explored for use in molecular electronics covalent solid that crystallizes into a diamond
and computing and for imparting biocompatibility cubic lattice structure, illustrated in Fig. 2. In a
to semiconductor devices for implantation. Or- covalent solid, bonding is produced by overlap of
ganic materials are also finding use in more con- the highly directional electronic orbitals of the
882 S.F. Bent / Surface Science 500 (2002) 879–903

dices, which designate the particular orientation at


which the bulk is terminated. The top surface in
Fig. 2, e.g., has the (1 0 0) orientation. Both the
(1 0 0) and (1 1 1) surfaces undergo extensive re-
constructions, i.e. their surface atomic geometry
differs significantly from that of the bulk. More-
over, the two surfaces have markedly different
surface structures [14,15]. The Si(1 0 0) surface re-
constructs into a ð2  1Þ structure, where ð2  1Þ
designates the new periodicity of the surface atoms.
The Si(1 0 0)-2  1 surface, illustrated in Fig. 3(a),
consists of pairs of silicon atoms in adjacent rows
that have bonded to each other, thereby reducing
Fig. 2. The diamond cubic lattice is the crystal structure found
the number of dangling bonds. These pairs of sili-
for diamond, silicon and germanium. Each atom is bonded to con atoms are called dimers. Si(1 1 1), on the other
four neighboring atoms in a tetrahedral geometry. hand, reconstructs into a complex (7  7) structure
that contains 49 surface atoms in the new unit cell.
It is useful to consider in more detail the
atoms. Like carbon, its group IV homolog, silicon bonding of the dimers on the Si(1 0 0)-2  1 sur-
atoms hybridize into a tetrahedral bonding con- face, because they play such an important role in
figuration. In the bulk solid, the diamond cubic
lattice allows silicon atoms to achieve this tetra-
hedral configuration. At the surface of the mate-
rial, however, the bulk is truncated, so the stable
bulk tetrahedral bonding is disturbed.
The reactivity of the silicon surface is controlled
in part by the unsatisfied bonding orbitals, or so-
called dangling bonds, that remain upon truncat-
ing the bulk. A surface silicon atom that is bonded
to only two other silicon atoms instead of four, as
shown at the top surface of the crystal illustrated
in Fig. 2 for example, would have two dangling
bonds. Dangling bonds contain single electrons,
whereas normal covalent bonds contain two spin-
paired electrons. At the surface, atoms can read-
just to minimize the total free energy of the system
and eliminate the dangling bonds. This process is
referred to as surface ‘‘reconstruction’’. The energy
minimization is a trade-off between energy gained
by forming new local bonds (to eliminate the
dangling bonds) with energy lost because of bond Fig. 3. Two surfaces of silicon that are important in organic
strain that results from its new configuration. This functionalization chemistry. (a) The clean Si(1 0 0)-2  1 sur-
energy trade-off often leads to complex surface face, with rows of silicon dimers lining the surface. These di-
reconstructions [12–15]. mers play an important role in the chemistry of organic
molecules at this surface. (b) The hydride-terminated Si(1 1 1)
The crystallographic faces of silicon that are
surface. The golden spheres represent silicon atoms; hydrogen
most important industrially are the Si(1 0 0) and the atoms are shown in white. The hydrogen–silicon bonds are the
Si(1 1 1) surfaces. The numbers in parentheses refer important molecular group in hydrosilylation attachment
to a crystallographic notation known as Miller in- chemistry, discussed in the text.
S.F. Bent / Surface Science 500 (2002) 879–903 883

heating to temperatures near 900 °C. The resulting


surfaces are not stable in air, instead quickly oxi-
dizing to form an SiO2 layer. Hence, experiments
on the clean Si(1 0 0)-2  1 and Si(1 1 1)-7  7 sur-
faces must be performed in vacuum, and the
organic reactants for functionalization are intro-
duced in vapor form into the vacuum chamber.
On the other hand, silicon surfaces can be
rendered relatively stable in air (i.e. relatively re-
Fig. 4. Illustration of a r bond and a p bond in a double- sistant to oxidation) by coating the surface with
bonded molecule. The r bond (green) is cylindrically symmetric hydrogen. A number of recipes have been devel-
around the molecular axis, while the two lobes of the p bond oped to terminate the silicon surface with Si–H
(red and blue) are separated by a nodal plane. The molecule (silicon hydride) groups. An excellent overview of
shown is ethylene.
this process is given by Weldon et al. [5]. The
most common methods involve exposure to di-
lute, aqueous hydrofluoric acid solutions, and
the reactivity of this surface toward organic mole- these wet etch procedures are used both at
cules. The Si(1 0 0)-2  1 dimers are attached by bench-top research laboratories and large-scale
something akin to a double bond [16]. In the lan- microfabrication facilities alike. Not only does
guage of chemical bonding, a double bond consists terminating a silicon surface with hydrogen atoms
of two types of bonds: a r bond (one that has render it passivated toward many reactions, but it
symmetry around the axis connecting the two also changes the surface structure. Because the
atoms) and a p bond (one with a nodal plane along reconstructed surface contains some weak Si–Si
the axis) [17]. Fig. 4 provides an illustration of a r bonds, replacement of these with strong Si–H
and a p bond in a molecule; note that the p bond bonds can relax the surface reconstructions. One
has two lobes. important surface for organic functionalization
However, in contrast to the p bond in the mol- reactions is that of hydride-terminated Si(1 1 1).
ecule shown in Fig. 4 (ethylene), the p bond be- Due to the strong bonding and passivating nature
tween the silicon atoms is sufficiently weak that the of hydrogen at the interface, Si(1 1 1) that is hy-
dimer is not held in a symmetric configuration. dride terminated does not reconstruct, revealing
Instead, the energy of the Si(1 0 0)-2  1 surface instead a nearly bulk-like periodicity of the sur-
can be reduced even further if the dimer moves face silicon atoms [18]. The structure of the
from a symmetric configuration to one in which hydride-terminated Si(1 1 1) surface is shown in
the dimer is tilted. Hence, at low temperatures, the Fig. 3(b).
dimers of the Si(1 0 0)-2  1 surface are observed Both the bare and hydrogen-terminated silicon
to be tilted [14]. At higher temperatures, thermal surfaces have been used as substrates for organic
energy induces a rapid change in the direction of functionalization. Because of their different rela-
the tilt which causes the dimers to appear sym- tive reactivities, most of the reactivity studies of
metric, as represented in Fig. 3(a). The presence of bare silicon are performed under ultrahigh vacuum
the p bond in the silicon dimer and the dimer conditions while reactions with hydride-termi-
tilting will both be important in understanding the nated silicon involve wet chemical methods. De-
reactivity of the surface. spite the different approaches and surface
Even after reconstruction, both the Si(1 0 0)- structures, however, a unifying feature in all of
2  1 and Si(1 1 1)-7  7 surfaces are highly reac- these attachment chemistries is the molecular
tive. Both surfaces are typically cleaned and nature of the chemistry, which in turn allows
prepared in an ultrahigh vacuum environment some common principles to be formulated for
(p ¼ 1010 Torr) by bombarding the surface with functionalization chemistry at semiconductor sur-
argon ions to sputter away surface layers and/or faces.
884 S.F. Bent / Surface Science 500 (2002) 879–903

3. Principles of cycloaddition chemistry at semicon- p bonded molecules come together to form a new
ductor surfaces cyclic molecule, losing two p bonds and making
two new r bonds in the process. The reactions are
The covalent nature of the semiconductor sur- designated by how many p electrons of each re-
face permits its reactivity to be described within a actant molecule are involved in the reaction. Two
molecular framework. This covalency leads to a types of cycloaddition reactions are shown in Fig.
reactivity that is often quite distinct from that of 5. In the first, known as a ½2 þ 2 reaction (i.e. two
metals. On metals, bonding may not be strongly p electrons in the ethylene molecule þ two p
site specific. This characteristic is markedly differ- electrons in the other ethylene molecule), two
ent from that of a surface such as silicon, where alkenes come together to form a new four-mem-
bonding is both localized and directional. To a bered ring. The second type of cycloaddition in
significant degree, the reactivity of semiconductor Fig. 5 is the ½4 þ 2 reaction, also known as the
surfaces can be understood by drawing analogies Diels–Alder reaction [19]. This important reaction
to known molecular systems, and in particular to is named after Otto Diels and Kurt Alder, who
systems in organic chemistry. In this section, this received the Nobel Prize in Chemistry in 1950
molecular analogy is developed to understand for their studies of this reaction. Here, a ‘‘diene’’
and predict one class of organic reactions that molecule with two neighboring (conjugated) p
can be applied to surface functionalization of the
Si(1 0 0)-2  1 surface. These are known as cyclo-
addition reactions. In Section 4, examples of cy-
cloaddition chemistry will be described. Then in
Section 5, some of the same principles developed
for surface cycloaddition reactions will be ex-
tended to other classes of organic reactions used to
functionalize silicon.
The Si(1 0 0)-2  1 reconstructed surface, shown
in Fig. 3(a), to some extent mimics an organic
reagent. As described above, the dimers of the
Si(1 0 0)-2  1 surface are attached by two bonds: a
r bond and a p bond (Fig. 4) [16]. The p bond
across the dimer is quite weak, but nevertheless its
presence provides a nice way of describing the di-
mer [17]. Because of these two bonds, it is useful to
draw analogies between the Si dimer and a car-
bon–carbon double bond, since C and Si are both
in the same group (group IV) of the periodic table.
Carbon–carbon double bonds, known as alkenes,
are a very well studied functional group in organic
chemistry. The simplest alkene molecule, ethylene,
is the molecule displayed in Fig. 4.
A number of reactions are known which can
form bonds with an alkene group. One syntheti-
cally useful class of these reactions are ‘‘cycload-
dition’’ reactions. Cycloadditions are widely used Fig. 5. Cycloaddition reactions are an important class of re-
in organic synthesis as a means of forming new actions for forming new ring compounds in organic chemistry.
The ½2 þ 2 cycloaddition between two alkenes forms a four-
carbon–carbon bonds and new carbon rings be- membered ring, while the ½4 þ 2 cycloaddition between an
cause of their versatility and high stereoselectivity alkene and a diene forms a six-membered ring. The designations
[19–21]. Cycloadditions are reactions in which two refer to the number of p electrons involved in the reaction.
S.F. Bent / Surface Science 500 (2002) 879–903 885

bonds such as butadiene reacts with an alkene


(ethylene) to form a new six-membered ring. In
Fig. 5, the p bond and the r bond are each rep-
resented by a line, so that the double bond of
ethylene is shown with two lines while the product
cyclobutane, which contains only r bonds, is
shown with single lines.
The cycloaddition reactions in Fig. 5 are subject
to what are know as the Woodward–Hoffman se-
lection rules. These selection rules stem from an
analysis based on frontier orbital theory which
examines the symmetries of the highest occupied
and lowest unoccupied molecular orbitals (HOMO
and LUMO, respectively) of the reactants as they
come together to form the reaction product.
Frontier orbital theory was formulated by Wood-
ward and Hoffman [22]. Professor Hoffman was
awarded the Nobel Prize in Chemistry in 1981,
Fig. 6. Examination of two cycloaddition reactions by frontier
together with Professor Fukui of Kyoto Univer-
orbital analysis. The frontier orbitals under consideration are
sity, for his contributions in this area. The resulting the highest occupied molecular orbital (HOMO) and the lowest
Woodward–Hoffman selection rules are widely unoccupied molecular orbital (LUMO). The blue and red lobes
used for predicting how readily an organic reaction represent the mathematical sign of the molecular orbital, such
will occur. Such an analysis is illustrated in Fig. 6 that blue combines with blue and red combines with red.
Mixing the orbitals leads to a symmetry forbidden reaction, as
for ½4 þ 2 and ½2 þ 2 cycloaddition reactions. In
in the case of the ½2 þ 2 reaction.
cycloaddition reactions, the frontier p orbitals
(characterized by positive and negative lobes, surface. We can therefore formulate the following
shown in different colors in Fig. 6) must overlap ‘‘in guiding principle. To the extent that the simple
phase’’ for the reaction to be symmetry allowed. picture of surface Si@Si double bonds is valid,
Because of the particular properties of these p similar symmetry selection rules as developed for
orbitals in the ½2 þ 2 reaction, ½2 þ 2 cycloaddi- organic reactions should apply to the surface. On
tions are found to be ‘‘symmetry forbidden’’. That the other hand, as the analogy of the dimer as a
is, the Woodward–Hoffman symmetry rules dic- true double bond becomes less accurate, deviations
tate that this reaction should not occur without from the symmetry selection rules will be expected.
significant energy activation. In fact, in organic Indeed, this prediction has been borne out in a
chemistry this reaction is largely limited to number of recent studies of cycloaddition reac-
synthesis involving photochemical activation. tions on Si(1 0 0)-2  1 [23–51], as well as on the
The ½4 þ 2 reaction, in contrast, has frontier or- related surfaces of Ge(1 0 0)-2  1 [52–55] and
bital properties (shown in Fig. 6) that make it C(1 0 0)-2  1 [56–58]. These studies are described
‘‘allowed’’ by symmetry considerations. Indeed, in Section 4.
Diels–Alder reactions are commonly used in or-
ganic synthesis as a means of forming new C–C
bonds and ring structures. 4. Examples
If the analogy that we have drawn between the
Si@Si dimer on the Si(1 0 0)-2  1 surface and an 4.1. Examples of [2 þ 2] cycloaddition reactions on
alkene group is reasonable, then certain parallels silicon
might be expected to exist between cycloaddition
reactions in organic chemistry and reactions that The first examples of what can be categorized
occur between alkenes or dienes and the silicon as ½2 þ 2 type cycloaddition reactions occurring
886 S.F. Bent / Surface Science 500 (2002) 879–903

between an alkene and a silicon surface were re- Just as two ethylene molecules (CH2 @CH2 ) react
ported the late 1980’s. Alkenes such as ethylene, as to form a four-membered cyclobutane ring, eth-
well as the related alkyne (triply bonded) molecule ylene reacts with the silicon dimer (Si@Si) on the
acetylene, were reacted with the clean Si(1 0 0)- Si(1 0 0)-2  1 surface to form a four-membered
2  1 surface in vacuum [31,46–51,59–69]. The Si2 (CH2 Þ2 ring with the surface.
alkenes were found to chemisorb at room tem- Interestingly, the surface ½2 þ 2 cycloaddition is
perature, forming stable species like that shown in a relatively fast reaction, occurring readily with
Fig. 7(c) that ‘‘bridge-bonded’’ across the silicon most alkenes at room temperature [9]. This high
dimers on the surface. The reaction proceeded by reactivity would not be expected for a true, sym-
formation of two new r bonds between Si and C metric 2 þ 2 reaction, which formally is symmetry
atoms, hence the bonding was referred to as ‘‘di- forbidden and is slow for the analogous homoge-
sigma’’ bonding. In addition, it was shown that neous reaction [21,22]. The symmetry analysis
while the p bonds of the alkene and of the Si–Si shown in Fig. 6, however, applies to a system in
dimer are broken, the r bonds remain intact which the Si dimer is symmetric and closely mimics
[65,66,68]. With this kind of reaction product an alkene. This is the pathway illustrated in Fig.
formed, one can draw a direct analogy with the 7(a). However, the Si dimer is not symmetric in its
surface adsorption of the alkene (Fig. 7) and the lowest energy configuration. The lowest energy
½2 þ 2 cycloaddition reaction shown in Fig. 5(a). state of the Si(1 0 0)-2  1 surface is actually one in

Fig. 7. A ½2 þ 2 reaction at the silicon surface occurs between an alkene molecule and the silicon dimer. The product is bridge-bonded
across the dimer.
S.F. Bent / Surface Science 500 (2002) 879–903 887

which the silicon dimers are tilted. It has been The experiments by Hamers and coworkers
proposed that this low symmetry (tilted) geometry were performed on a Si surface that was inten-
allows the reaction to proceed through an asym- tionally cut slightly (4°) off-axis from the (1 0 0)
metric pathway in which the alkene approaches crystal face. This miscut has the effect of producing
the dimer from one side, as shown in Fig. 7(b) a stepped surface, where between each step resides
[10,31,64,69]. This asymmetric approach is of a a terrace
8 atoms wide. The orientation of the
lower symmetry and can occur with little or no dimer rows on the terraces is indicated schemati-
energetic barrier. This example serves as an inter- cally in Fig. 9(a). Fig. 9(b) shows a scanning
esting case of a solid state effect in the silicon tunneling microscope (STM) image of the clean 4°
(tilting of the dimers) controlling a surface reac- miscut Si(1 0 0) surface [35]. In the picture, each
tion (via relaxation of the symmetry constraints). oblong white feature is an image of a Si–Si dimer.
The result of the ½2 þ 2-like reaction is a tightly It is readily seen that on the flat terrace regions, the
bonded organic group, directly attached to the Si@Si dimers all maintain the same orientation
silicon surface through strong Si–C bonds ( 82 with respect to the step. If cyclopentene molecules
kcal/mol each [25]), as illustrated in Fig. 7(c). are now adsorbed on this surface, and if they bond
Furthermore, as reported above, the reaction across the dimer, as expected for a ½2 þ 2 cyclo-
occurs readily for many alkenes. Consequently, addition, we would expect to see rows of bound
½2 þ 2 cycloaddition has attracted much attention cyclopentene that follow the underlying silicon
as a means to functionalize the Si(1 0 0)-2  1 sur- rows. In fact, the STM images obtained after cy-
face. Many beautiful examples have been reported clopentene exposure clearly show ordered molec-
of silicon surfaces functionalized by means of ½2 þ ular adsorption of cyclopentene at this surface. In
2 chemistry using alkene or dialkene precursors. Fig. 9(c), the white features (larger than those in
One example is the work by Hamers and co- Fig. 9(b)) are images of individual adsorbed cy-
workers. In 1997, this group began a series of clopentene molecules. These images offer beautiful
studies showing that this type of attachment chem- evidence for the degree of molecular order that can
istry can be used to form well-ordered monolayer be achieved in this reactive system. Long range
organic films across the surface of silicon [31–36]. order on a length scale of millimeters was con-
By adsorbing cyclopentene (the molecule shown in firmed by measuring the retention of orientational
Fig. 8) at room temperature in vacuum, layers that anisotropy across the length of the silicon sample
are ordered in registry with the underlying silicon using polarized vibrational spectroscopy [32].
surface lattice could be generated [32,35]. These studies provide support for the use of cy-
cloaddition chemistry at the Si(1 0 0)-2  1 surface
as a method to prepare ordered monolayer organic
films.
The success in forming ordered overlayers is
especially significant in light of the following
characteristics of this system. As with ethylene,
cyclopentene was found to react at the Si(1 0 0)-
2  1 surface with near unity reactivity, i.e., each
gas molecule that collides with the surface has an
almost perfect chance of sticking and bonding at
the surface. Furthermore, diffusion is expected to
be very slow in these systems at room temperature,
suggesting that once the molecules have bonded to
the surface, they remain at or near the same site on
the surface [9]. The formation of ordered over-
layers in the absence of significant surface mobility
Fig. 8. The cyclopentene molecule. is in contrast to the behavior of other growth
888 S.F. Bent / Surface Science 500 (2002) 879–903

coverage. This is usually achieved by holding the


substrate at an elevated temperature. These cy-
clopentene films, therefore, do not grow like or-
dinary metal or semiconductor crystals. Rather,
the strong and directional bonding at the interface
is sufficient to direct the molecules into the neces-
sary sites, so that the semiconductor surface acts as
a template for the placement of the organic mol-
ecules. The effect is one in which the structure of
the underlying surface is propagated into ordered
overlayers of organic molecules.
Ordered overlayers are not achieved in all
systems, however, and it appears that a certain
rigidity and molecular size is necessary to form
well-ordered films, as evident in the following ex-
ample. Although Hamers et al. have shown that
other alkenes and dialkenes (such as 1,5-cyclo-
octadiene, displayed in Fig. 10) could be used to
produce ordered monolayer films through the
same ½2 þ 2 attachment [33], a number of other
alkenes investigated did not adsorb with a high
degree of order. For example, cyclohexadiene and
cyclooctatetraene, both illustrated in Fig. 10, while
reacting readily with the Si(1 0 0)-2  1 surface, did
not form ordered films on Si(1 0 0)-2  1 [34,36].
This reduced packing may stem from either disor-
der in the configurations of individual adsorbates
or from too low a final (saturation) coverage of
molecules at the surface. Rigidity of the organic
molecule, generally provided by a ring structure,
was found to help in the formation of ordered lay-
ers. In addition, having a single type of functional
group or a single means of bonding to the surface
was advantageous, because molecules that bind in
several different configurations do not maintain the
degree of order necessary for close packing [34].
Fig. 9. (a) Schematic illustration of the oriented rows of silicon
dimers present on a clean Si(1 0 0) surface that has been inten- In addition to forming densely packed organic
tionally miscut by 4°. (b) STM images of this surface, and (c) layers, the ½2 þ 2 cycloaddition reaction can be
the same surface after adsorption of cyclopentene. Figure re- used to generate a chiral surface, according to a
printed from Ref. [35] with permission from Elsevier Science. recent study by Lopinski et al. [39]. Chirality refers
to the handedness of a molecule, and chirality
systems. In most common methods of semicon- plays an enormously important role in the chem-
ductor film growth, such as molecular beam epit- istry of natural products. A chiral molecule is one
axy (MBE) or chemical vapor deposition (CVD), that cannot be superimposed on its mirror image
diffusion of the depositing species at the surface is similar to a human hand or a conch shell. For
a requisite for the formation of good overlayers example, in Fig. 11, there is no way to take the
[1]. Motion of atoms or molecules to find the op- image of either the molecule or the hand on the left
timal binding sites is necessary for conformal film side of the picture and superimpose it with the
S.F. Bent / Surface Science 500 (2002) 879–903 889

Fig. 10. The 1,3-cyclohexadiene, 1,4-cyclohexadiene, 1,5-cyclooctadiene, and cyclooctatetraene molecules.

corresponding image on the right (the mirror im- to bridge across two dimers by reacting also at the
age). Both objects are chiral. Chiral molecules that cyclopropane group. Calculated structures for ca-
share the same stoichiometry and bonding se- rene molecule adsorbed at a Si(1 0 0)-2  1 surface
quence but are non-superimposable are known as are indicated in Fig. 12. The structure in Fig. 12(b)
enantiomers. An example is the molecule pictured is of the ½2 þ 2 cycloaddition product; that in Fig.
in Fig. 11. Nature makes broad use of enan- 12(c) is of carene bridging across two dimers.
tiomeric species, and most biochemical processes The carene molecule was chosen in the study
select for one enantiomer over another. For because it contains a bulky dimethyl group that
example, biological enzymes distinguish between geometrically hinders adsorption at one face of the
enantiomers. molecule. This so-called steric hindrance forces the
Lopinski et al. have demonstrated the formation ½2 þ 2 reaction to take place with only one facial
of a synthetic chiral surface by adsorbing the orientation of the molecule. Lopinski et al. verified
molecule 1SðþÞ-3-carene on the Si(1 0 0)-2  1 that the reaction was enantiospecific, i.e. produced
surface [39]. Carene, illustrated in Fig. 12(a), is a only one of the possible chiral products, using
chiral, two-ringed molecule containing a C@C scanning tunneling microscopy. The STM image
double bond and a three-carbon-atom ring (cy- allows the determination of absolute chirality of
clopropane). Carene initially adsorbs by ½2 þ 2 adsorbed molecules, as shown in earlier work from
cycloaddition between the alkene group and the the same group [38]. The resulting carene adsor-
silicon dimer. The final adsorbed state is believed bate contains chiral centers and produces a chiral
890 S.F. Bent / Surface Science 500 (2002) 879–903

4.2. Examples of Diels–Alder reactions on silicon

In 1997, a series of papers was published in


which the ½4 þ 2 (Diels–Alder) cycloaddition re-
action product (Fig. 5(b)) was first predicted and
observed to occur on the Si(1 0 0)-2  1 surface.
Konecny and Doren performed a theoretical study
in which they predicted that Diels–Alder reactions
should occur across silicon dimers on a Si(1 0 0)-
2  1 surface, just as they do across C@C double
bonds in a molecule [23,24]. Their prediction
was based on ab initio calculations of the reaction
of a model conjugated diene (1,3-cyclohexadiene,
shown in Fig. 10) with a silicon surface.
In the calculations, the silicon surface was
modeled by a silicon cluster containing nine silicon
atoms and featuring the silicon dimer, as shown
in Fig. 13. This silicon cluster and related larger
clusters have been shown to provide adequate
Fig. 11. A chiral molecule or object (such as the human hand)
models of the silicon surface while remaining
is one that cannot be superimposed on its mirror image. computationally tractable. Doren et al. followed
the reaction of cyclohexadiene with the silicon
cluster, and examined two possible reaction path-
surface [39]. This result is significant because of its ways: a ½4 þ 2 reaction or a ½2 þ 2 reaction
implications for sensor development using organ- [23,24]. Fig. 13 illustrates the two possible reaction
ically modified surfaces. In molecular recognition, products for a related molecule, 1,3-butadiene (see
distinguishing between enantiomers of molecules Fig. 5). The calculations showed that there is a
requires a chiral probe. Hence, if organic layers on substantial thermodynamic driving force (54 kcal/
silicon are to be used to achieve molecular recog- mol) to form the product of a ½4 þ 2 reaction
nition, particularly in biological systems, chiral between cyclohexadiene and the silicon dimer.
surfaces such as this one must be utilized. This is significantly more energy than is gained in

Fig. 12. Structures of (a) the 1SðþÞ-3-carene molecule; (b) carene molecule bonded by ½2 þ 2 cycloaddition across the silicon dimer; (c)
carene molecule bonded in a bridging configuration across two silicon dimers. Both adsorbed structures are chiral. Figure adapted with
permission from Ref. [39]. Copyright 1999 American Chemical Society.
S.F. Bent / Surface Science 500 (2002) 879–903 891

Fig. 13. In ab initio theoretical studies of silicon surface reactivity, the Si(1 0 0)-2  1 surface is modeled using a nine silicon atom
cluster. Two possible outcomes for the reaction of the Si cluster with 1,3-butadiene are shown: the product of the Diels–Alder reaction
and the product of the ½2 þ 2 reaction.

the competing ½2 þ 2 reaction (39 kcal/mol). In


addition, the calculations predicted that the reac-
tion should occur with little or no energetic barrier
across a symmetric silicon dimer, consistent with
the Woodward–Hoffman rules for a ½4 þ 2 reac-
tion [23,24].
Studies by Teplyakov et al. provided the ex-
perimental evidence for the formation of the
Diels–Alder reaction product at the Si(1 0 0)-2  1
surface [26,27]. A combination of surface-sensitive
techniques was applied to make the assignment,
including surface vibrational spectroscopy, ther-
mal desorption studies, and synchrotron-based
X-ray absorption spectroscopy. Vibrational spec-
troscopy in particular provides a molecular ‘‘fin-
gerprint’’ and is useful in identifying bonding and
structure in the adsorbed molecules. Fig. 14 pro-
vides an example of a vibrational spectrum used to Fig. 14. Measured vibrational spectra giving information on
assign the bonding of the butadiene molecule at a the bonding of butadiene at the Si(1 0 0)-2  1 surface. The
peaks shown in the spectrum correspond to absorption of light
surface. The peaks shown in the spectrum corre- at the energies of vibration over the spectral region for carbon–
spond to absorption of light at the energies of vi- hydrogen bond stretches. The units of energy are in wave-
bration (in a unit of reciprocal centimeters, or number, or cm1 .
cm1 , called wavenumber) for carbon–hydrogen
bonds in the molecules. Because there are multiple
C–H bonds present in the molecule, the spectrum can be assigned. An analysis of the vibrational
contains multiple peaks. By comparing the spec- spectra of adsorbed butadiene on Si(1 0 0)-2  1 in
trum for the butadiene molecule (Fig. 14(a)) with which several isotopic forms of butadiene (i.e.,
the spectrum for reacted butadiene (Fig. 14(b)), some of the H atoms were substituted with D
the bonding configuration of the reacted molecule atoms) were compared showed that the majority of
892 S.F. Bent / Surface Science 500 (2002) 879–903

butadiene molecules formed the Diels–Alder re- ality of the surface cycloaddition reaction. In other
action product at the surface. Very good agree- words, one can ask whether cycloaddition reac-
ment was also found between the experimental tions also occur on the surfaces of related covalent
vibrational spectra obtained by Teplyakov et al. materials. The closest homologs to silicon for
[26,27] and frequencies calculated for the Diels– comparison are carbon and germanium, also in
Alder surface adduct by Konecny and Doren group IV of the periodic table.
[23,24]. Later studies by Hovis et al. using scan- Both carbon (in its diamond form) and germa-
ning tunneling microscopy confirmed the Diels– nium are of technological interest. Silicon germa-
Alder product in this system and quantified the nium alloys, for example, are used for high
formation of a small ½2 þ 2 side product [34]. frequency devices in applications such as wireless
Both the ½2 þ 2 and ½4 þ 2 products of reaction communications [70,71]. Diamond has received
between a diene and the Si(1 0 0)-2  1 surface much attention for its unique combination of
have a number of features that are desirable for material properties, such as high thermal conduc-
surface functionalization. As shown in Fig. 13, tivity, wide optical transparency, hardness and
both reactions produce organic molecules that are chemical inertness [72].
strongly bound to the silicon surface through two Like Si(1 0 0), the (1 0 0) crystallographic face of
covalent Si–C bonds. The high strength of these both Ge [73] and diamond [74] undergo a ð2  1Þ
bonds means that the organic layer at the surface reconstruction to form rows of surface dimers.
is typically stable to temperatures in excess of one These dimers each are believed to have some de-
to two hundred degrees celsius [30]. In addition, gree of double-bond character, with a strong r
both of the cycloaddition reaction products be- bond and a partial p bond [73,74]. However, the
tween the diene and the surface retain one double surfaces differ in the strength of the p bond and
bond in the organic layer. This remaining double the degree of tilt that is observed in the dimers.
bond (alkene group) can be used as a starting The dimers on both the Si and Ge surfaces tilt due
point for further derivatization of the surface. For to the weakness of the p bond. In the case of
example, the alkene group in the first layer can act Si(1 0 0)-2  1, the dimers are observed in their
as the attachment point for the second layer, a new tilted form only at low temperature, as discussed
functional group in the second layer as the at- previously, whereas on Ge(1 0 0)-2  1, the dimers
tachment point for the third layer, and so on. In are tilted at both low and moderate temperatures
this way, there is the potential to perform con- [14]. An appropriate picture of the dimers on these
trolled, layer-by-layer synthesis of organic films surfaces is that shown in Fig. 7(b). In contrast, the
on the Si(1 0 0)-2  1 surface. diamond dimer is essentially different. Due to good
p bond formation between C atoms, the diamond
4.3. Down the periodic table: solid state effects on surface forms symmetric dimers, best described by
cycloaddition chemistry the picture in Fig. 7(a), over a large temperature
range [74].
It is intriguing that the analogs of key reactions Over the past few years, researchers have begun
in organic chemistry––such as the Diels–Alder and to probe for detailed parallels between the surface
½2 þ 2 cycloadditions––occur across the silicon chemistry of Si(1 0 0) with that of Ge(1 0 0) and
dimer on the Si(1 0 0)-2  1 surface. These results C(1 0 0), with special consideration paid to the role
lend support to the analogy that was drawn be- of the surface dimers. The first published studies
tween a dimer on Si(1 0 0)-2  1 and the C@C focused on Ge surfaces. Adsorption studies of
double bond. The success of this analogy, in turn, alkene-containing molecules such as ethylene [52],
advances the concept of describing the reactivity of butadiene [53], cyclopentene [54], and cyclohex-
a covalent surface within a molecular framework. adiene [55] at the Ge(1 0 0)-2  1 surface in ul-
It has therefore been a challenge to both experi- trahigh vacuum revealed that like Si, Ge(1 0 0)
mentalists and theoreticians to test the limits of can participate in cycloaddition chemistry. Both
this molecular framework by probing the gener- ½2 þ 2 and ½4 þ 2 reaction products have been
S.F. Bent / Surface Science 500 (2002) 879–903 893

observed experimentally. Theoretical investiga- and coworkers [56,57] and by Hossain et al. [58]
tions by Musgrave and coworkers have also re- have shown that the carbon dimers on the dia-
vealed close parallels in energetics between mond(1 0 0) surface behave like C@C double
cycloaddition reactions on Si(1 0 0) vs Ge(1 0 0). bonds toward organic cycloaddition reactions. For
Namely, on both surfaces, the ½4 þ 2 cycloaddi- example, the reaction between butadiene and the
tion product is more stable by
15–20 kcal/mol C(1 0 0)-2  1 surface was found to form predom-
than the ½2 þ 2 cycloaddition product [25]. In- inantly the Diels–Alder product. The resulting six-
deed, experimental studies of Diels–Alder reac- membered ring tethered to the surface exhibited
tions on Ge(1 0 0)-2  1 show close similarities in remarkable similarities to cyclohexene, which is
bonding (based on the vibrational spectra) with the Diels–Alder product of the related reaction
the Si-based product [53]. Although the formation between butadiene and ethylene, shown in Fig. 5
of well-ordered monolayers, as in the case of sili- [56]. Recent theoretical calculations by Fitzgerald
con, has not yet been reported, it might be ex- and Doren have investigated the butadiene/dia-
pected that the proper choice of molecular mond(1 0 0) reaction system, and their results are
precursor will lead to similarly packed monolayers consistent with formation of a majority ½4 þ 2
on Ge. product with a minor ½2 þ 2 product [75].
Unlike on Si, however, the Diels–Alder reaction Studies of ½2 þ 2 reactions on diamond provide
was found to be reversible on Ge(1 0 0)-2  1, further insight into the reactivity of covalent solids
leading to loss (by desorption) of the organic layer [57] and further evidence for a solid state effect on
upon heating the surface to 300 °C [53]. The re- surface reactivity in these systems. Recall that
verse of the Diels–Alder reaction is called the ½2 þ 2 cycloadditions are not generally observed
retro-Diels–Alder reaction, and it also finds close in molecular systems because they are symmetry
parallels in molecular organic chemistry. This forbidden according to the Woodward–Hoffman
difference between Si and Ge surfaces in the be- rules. At the surfaces of Si and Ge solids, ½2 þ 2
havior of the organic layer upon heating is a sig- reactions are believed to occur because the surface
nificant one. It is the result of a fundamental dimers tilt, providing an asymmetric reaction ap-
difference in the strength of the bonds formed be- proach (transition state) that relaxes the symmetry
tween carbon and silicon atoms versus carbon and restrictions [31,69]. Because on diamond the di-
germanium atoms. A C–Ge bond is weaker than a mers do not tilt [74], it would be expected that the
C–Si bond by almost 10 kcal/mol [25] and bond diamond surface, of all the group IV surfaces,
for bond, an organic molecule bound to Ge will be would behave most like its C@C analog. Follow-
held less tightly than one bound to Si. Exploiting ing the Woodward–Hoffman selection rules, a
the reversibility of the attachment reaction on Ge ½2 þ 2 reaction along a symmetric reaction path-
due to its weaker bonding is being considered as a way would be unlikely to occur. Consistent with
way to spatially pattern organic layers on the this expectation, it was found that while cyclo-
surface [53]. pentene does undergo ½2 þ 2 addition on C(1 0 0),
Recently, there have been reports of both ½4 þ 2 the reaction probability is three orders of magni-
and ½2 þ 2 cycloaddition reactions occurring on tude lower than that on Si(1 0 0) or Ge(1 0 0) [57].
the diamond(1 0 0)-2  1 surface [56–58]. Unlike Hence the structure of the surface exerts a strong
germanium or silicon, the diamond surface allows influence on the surface reactivity.
direct comparison to be made between chemistry in
a molecular system (i.e. organic chemistry) versus
chemistry on the surface of an extended solid (i.e. 5. Further applications: functionalization of hydride-
surface chemistry). First, both reactions directly terminated Si surfaces
involve carbon atoms. Second, like the corre-
sponding alkene, the carbon–carbon dimer on Both ½2 þ 2 and ½4 þ 2 cycloaddition reactions
the diamond(1 0 0)-2  1 surface is symmetric. In have now been demonstrated as a means of
agreement with this analysis, studies by Russell functionalizing the surfaces of three covalent
894 S.F. Bent / Surface Science 500 (2002) 879–903

semiconductor solids: C, Si, and Ge. Although Solution-based functionalization chemistry ex-
work is ongoing, the results already seem to be hibits a number of advantages over the vacuum
quite general for alkenes and dienes reacting on approaches. These include a much simpler appa-
these surfaces. Cycloaddition chemistry is a good ratus and the potential for significantly higher re-
method for forming organic layers on semicon- action rates due to higher reactant fluxes present in
ductor substrates. By proper choice of organic the liquid. In addition, for certain applications, the
reagent, a variety of useful surface layers can be solution-based method may be easier to integrate
engineered, including molecular monolayers ex- into an existing processing protocol. On the other
hibiting good stability, chirality, and a high degree hand, the degree of control over surface cleanliness
of order. The understanding of cycloaddition re- and structure, and the ability to perform in situ
actions at semiconductor surfaces is now reaching diagnostics, may be reduced for this method
a level where scientists can begin to apply these compared to a vacuum approach. For some ap-
reactions to the rational design of organic/semi- plications, the surface control and range of at-
conductor interfaces. tachment configurations that are possible by the
Conceptually, the framework for describing cy- vacuum-based approach may prove critical. For
cloaddition reactions in organic chemistry appears example, in the development of molecular elec-
to apply moderately well to reactions at these tronics, the electrical properties will depend upon
surfaces. These parallels allow one to apply the the structural properties of the surface and the
broad knowledge base available from organic specific interfacial bonding, which can be well
chemistry to tailor the reactions at the surface in controlled in a vacuum environment. In other
order to obtain targeted products. Furthermore, applications, the substrate being modified may not
the general principles that have can be developed be able to withstand the harsher treatment used
for the cycloaddition systems can be extended to in solution processing. The choice of using wet
understand an even broader range of reactions at chemistry (solution based) vs dry chemistry (vac-
semiconductors. For example, it is evident that uum based) ultimately will depend upon the needs
the surface structure plays a role in influencing of the particular application, and both methods
surface selection rules and reaction pathways. It continue to be actively pursued.
has also been shown that at these covalent sur- Most solution based functionalization ap-
faces, the formation of strong directional bonds proaches use the hydride-terminated silicon surface
with rigid organic molecules allows the production as a starting point. Hydride-terminated silicon,
of ordered overlayers, even in the absence of dif- particularly the H:Si(1 1 1) surface, can be gener-
fusion. ated with a high degree of order and atomic flat-
There are many additional applications in which ness, and can be rendered quite passive against
these guiding principles and a molecular approach oxidation and contamination [5]. The structure of
to designing surface reactions can be used to the H:Si(1 1 1) surface is shown in Fig. 3(b). The
functionalize semiconductor surfaces. One area of hydride-terminated surface does not require vac-
application that exhibits great promise is the so- uum conditions either for its preparation or sub-
lution-based functionalization of hydrogen-termi- sequent reaction [5].
nated silicon surfaces. Even before vacuum-based Recall that for the Si(1 0 0)-2  1 surface, the
cycloaddition methods began to be exploited for important functional group was the Si@Si dimer,
attaching organic groups to semiconductor sur- for which analogies were drawn to alkene chem-
faces (see Section 4), remarkable advances were istry. On the H:Si(1 1 1) surface, the group of in-
made in solution-based functionalization chemis- terest for functionalization is typically the Si–H
try. In these so-called ‘‘wet-chemical’’ methods, bond. The literature on organosilicon chemistry
monolayers of organic films could be obtained on (i.e., the chemistry of molecules with C–Si bonds)
silicon that imparted outstanding resistance to includes many known reactions for Si–H bonds.
oxidation and provided superb protection and One major class of these reactions is ‘‘hydrosily-
passivation to the underlying substrate. lation’’, whereby the silicon-hydrogen bond adds
S.F. Bent / Surface Science 500 (2002) 879–903 895

across an alkene (C@C) or alkyne (CBC) group.


Examples in which this chemistry is used to func-
tionalize silicon surfaces are described in the fol-
lowing sections.

5.1. Hydrosilylation chemistry at the H:Si(1 1 1)


surface

The first report of a densely packed organic


monolayer bound directly to silicon through Si–C
bonds was based upon the hydrosilylation reaction
on the H:Si(1 1 1) surface [76,77]. In pioneering
work, Chidsey and coworkers demonstrated that
long-chain 1-alkenes (those with the double bond
at one end of the chain) could be attached to hy-
drogen-terminated silicon surfaces at 100 °C in the
presence of a molecule called a radical initiator
[76,77]. A radical initiator is a species used to
create free radicals in reactions. Chidsey et al.
proposed that the surface attachment mechanism
was analogous to the hydrosilylation reaction in
solution phase chemistry. In the proposed surface
mechanism (illustrated in Fig. 15), the radical ini- Fig. 15. Proposed mechanism for surface hydrosilylation. Ini-
tiator starts the reaction by abstracting H from a tial loss of silicon hydride generates a silicon dangling bond.
surface Si–H bond, producing a silicon dangling Reaction between the silicon and an alkene molecule leads to an
bond (which is also referred to as a silicon radical). attached alkyl radical, which may abstract a hydrogen atom
In Fig. 15, a surface dangling bond is designated from a neighboring silicon [84].
by an orbital with a dot, where the dot represents
a single, unpaired electron. The silicon dangling drocarbons. More importantly, the organic layers
bond then reacts with the alkene molecule, form- formed by hydrosilylation were stable in a number
ing a new Si–C bond at the surface and leading to of environments, including boiling water, organic
the attachment of the organic molecule, which it- solvents, acids and bases, and air. The silicon
self is left with a carbon radical as shown in Fig. substrate was found to be protected by the
15. This surface bound organic radical is thought monolayer against oxidation after many weeks of
to abstract a hydrogen atom either from an unre- exposure to air [76,77].
acted alkene molecule or from a neighboring Si–H After the reports of the radical-initiated hydro-
group on the surface. The net reaction sequence silylation process, a number of different ap-
produces a stable, closed-shell organic group proaches for functionalization based on this
bound directly to the Si(1 1 1) surface. reaction were developed [8,77–79]. It was shown
Studies by Chidsey and coworkers [76,77] that the reaction could proceed not just with a
showed a remarkable stability and robustness to radical initiator, but also with other methods for
the organic monolayers formed by this process. A generating the initial silicon radical (dangling
number of diagnostic techniques were used to bond). The use of heat (>150 °C) succeeded by
characterize the monolayers, including vibrational thermally breaking the Si–H bond, although the
spectroscopy, X-ray spectroscopy and reflectivity, monolayers produced by this method were of a
ellipsometry and wetting measurements. These lower quality [77]. Another method involved use of
showed that the monolayers were tightly packed, acidic molecules (called Lewis acids) to catalyze
with a density close to 90% that of crystalline hy- the reaction [78,79]. Related approaches using
896 S.F. Bent / Surface Science 500 (2002) 879–903

reactions of organometallic species (molecules


involving both organic groups and metal atoms)
to functionalize silicon surfaces through the Si–H
bonds were also shown to work [80–83].

5.2. Radical chain reactions

A recent revisit of the proposed reaction mech-


anism for radical-initiated hydrosilylation [76,77]
has led to two beautiful demonstrations of silicon
functionalization in which the reaction propagates
along the surface, leaving attached organic mole-
cules in its wake. This occurs because the reaction
is ‘‘self-propagating’’. In a self-propagating reac-
Fig. 16. The styrene molecule.
tion, one reaction causes a subsequent reaction to
occur, which starts the next reaction, and so
on. Recall that in the surface hydrosilylation reaction during which hydrogen atoms are
mechanism (Fig. 15), the process is initiated by abstracted from neighboring silicon atoms. The
generating a silicon dangling bond on the hydride- observation of bunched-up islands of styrene ra-
terminated surface. Reaction of this silicon dan- ther than one-dimensional lines is attributed to a
gling bond with an alkene molecule leads to a ‘‘random walk’’ propagation of the reaction. Cic-
surface-bound organic group containing a carbon ero et al. suggest that the reaction stops at a silicon
radical, which must abstract a hydrogen atom in atom that no longer has any neighboring hydrogen
order to become a stable, closed shell species. If the atoms [84]. Note that this system provides another
adsorbate abstracts a hydrogen atom from a clear example of the localization of the reaction
neighboring Si–H group, a new silicon dangling and limited mobility of the organic molecules at
bond is produced, which can begin the reaction a silicon surfaces. If the styrene molecules were
new. This process is also known as a radical chain mobile, the bunching effect would not be observed.
reaction. In another report, Lopinski and coworkers have
This proposed radical-chain reaction mecha- used the same concept of the radical-initiated hy-
nism was recently confirmed in hydrosilylation drosilylation reaction on the Si(1 0 0)-2  1 surface
experiments performed in vacuum. In one report, to induce self-directed growth of molecular
Cicero et al. prepared hydride-terminated Si(1 1 1) ‘‘wires’’ on the surface [87]. On the Si(1 0 0)-2  1
surfaces using standard wet-etch methods, then surface, the radical chain reaction propagates
introduced the sample into an ultrahigh vacuum primarily along the direction of the dimer row,
chamber [84]. In vacuum, the tip of a STM was leading to ‘‘lines’’ of organic adsorbates, as shown
used to generate isolated silicon dangling bond in Fig. 17(B). Molecular assemblies as long as 130
sites surrounded by silicon hydride groups [85,86]. A (corresponding to 34 styrene molecules) were
This surface was then exposed to styrene, a mol- observed. Lopinski et al. suggested that by using
ecule which consists of an alkene group attached different reactant alkenes and seed conditions at the
to a benzene ring, as shown in Fig. 16. In the STM surface, this reaction may be used to engineer more
image of Fig. 17(A), islands of adsorbed styrene complex molecular nanostructures on silicon.
are observed on the silicon surface in areas sur-
rounding the locations of the initial dangling 5.3. Photoinitiation and photopatterning
bonds, which are designated by black dots [84]. In
other words, one dangling bond on the surface led Another interesting extension of the hydrosily-
to the deposition of many styrene molecules. This lation reaction involves photoinitiation. Chidsey
result supports the presence of a radical-chain and coworkers have shown that shining ultraviolet
S.F. Bent / Surface Science 500 (2002) 879–903 897

Fig. 17. (A) STM image of hydrocarbon islands formed by radical-initiated hydrosilylation chemistry of styrene on H-terminated
Si(1 1 1) [84] (reprinted with permission from Ref. [84]. Copyright 2001 American Chemical Society). (B) STM image of hydrocarbon
lines formed by the same mechanisms on H-terminated Si(1 0 0)-2  1. The four images show the propagation of the styrene nano-
structures with increasing styrene exposure [87] (reprinted from Ref. [87] with permission from nature).

light on the H:Si(1 1 1) surface while exposing the con, though, porous silicon is a strong emitter of
surface to a 1-alkene also led to the formation of light (called ‘‘luminescence’’). In fact, it luminesces
organic monolayers [88,89]. Buriak and coworkers in response to electrical, optical, and chemical
have used the photoinitiation of hydrosilylation in stimuli, making it of great interest for use in light
an exciting application of this process [90]. They emitting devices.
have examined functionalization reactions on po- Buriak et al. demonstrated that the hydrosily-
rous silicon, which is a cousin of single-crystalline lation reaction of alkenes could be induced by
silicon. Porous silicon contains a complex array of visible white light on photoluminescent samples of
pores at the surface, generating a surface structure hydride-terminated porous silicon [8,90]. Using a
of silicon reminiscent of stalagmites [8]. Porous simple optical setup in which light was only shined
silicon is generated by specialized etching of crys- on certain regions of the silicon, they showed that
talline silicon, and like crystalline silicon, can be using the photoreaction, they could form patterns
hydride terminated. In contrast to crystalline sili- of organic layers on the surface. Regions that were
898 S.F. Bent / Surface Science 500 (2002) 879–903

exposed to white light in the presence of alkene to pattern these attached organics, as shown in the
molecules became covered with organic groups, work by Buriak and coworkers [90], only expands
whereas regions that were kept dark remained the possibilities for future device applications.
terminated with hydrogen. Subsequently, the po-
rous silicon was etched in basic solution. Areas of
the substrate that were functionalized remained 6. Prospects
intact because the organic layer acted to protect
the underlying silicon, but areas that were hydride The microelectronics industry is headed toward
terminated were destroyed during the etch process. increasingly miniaturized devices. Within the next
A sample that had been subjected to this treat- decade or two, the industry will be facing the ne-
ment is shown in Fig. 18 while being illuminated cessity to fabricate atomic-scale devices. As semi-
with ultraviolet light. Some regions of the sample conductor devices continue their downward trend
appear orange. These are the areas that were in size, the importance of interfacial phenomena
photochemically coated with the organic layer; the and atomic scale manipulation will continue to
orange light is the visible luminescence that occurs rise.
when alkyl-functionalized porous silicon is ex- The tools available to surface chemists and
posed to ultraviolet light. The dark regions of the surface physicists already enable an unprecedented
sample reflect areas that do not photolumines- level of control in the ability to modify surfaces at
cence. These are areas that were not hydrosilylated the atomic level. New capabilities in both the ex-
and were etched in the basic solution. This simple perimental and theoretical arenas will give scien-
example serves to demonstrate the potential of tists and engineers increasing power to manipulate
organic functionalization in manipulating the surfaces in order to design specific functionality.
properties of a silicon substrate. Adding the ability The benefits of this molecular-level surface engi-
neering will be felt in many areas of microelec-
tronic fabrication––in atomic layer deposition, in
surface passivation, in deposition of ultrathin di-
electrics, for example. It will then go beyond ex-
isting processing and into the realm of new
uncharted territory. For example, signs are already
appearing of molecular electronics and molecular
computing [91].
One of the emerging areas in interface science is
organic functionalization of semiconductors. This
technology provides the opportunity to create
hybrid devices exploiting the properties and fea-
tures of both organic and inorganic materials. The
addition of organic layers at the interface provides
for myriad possibilities, including the incorpora-
tion of lubrication, optical response, chemical
sensing, or biocompatibility properties to semi-
conductor surfaces. Thus, the field of organic
Fig. 18. Photograph of a porous silicon sample that was modification ties directly into a number of tech-
photopatterned using the light-promoted hydrosilylation reac- nological areas that will become of increasing im-
tion. The orange regions of the pattern are coated with an or- portance.
ganic layer that protected the surface against etching, allowing
One area of impact is nanotechnology. Nano-
the underlying silicon to luminescence. The dark blue regions
were unprotected, and after etching they no longer lumines- technology refers to the development of materials,
cence (reproduced with permission of Wiley-VCH from Ref. systems and devices with a characteristic length
[90]). scale of 1–100 nm. For comparison, a H atom is
S.F. Bent / Surface Science 500 (2002) 879–903 899

roughly 0.1 nm in diameter. Such molecular-scale cisive role in influencing cell growth [92,93]. For
materials are being explored for many applications example, interfacial properties can be used either
in which novel electronic, optical or mechanical to inhibit or to accelerate cell adhesion or cell
properties are desired. One example is a so-called growth. Strategies utilizing patterned organic lay-
molecular computer, in which switches and mem- ers that combine adhesion proteins, cell growth
ory components will be constructed out of mole- factors, and inhibiting molecules can be used to
cules instead of patterned out of silicon. direct two-dimensional cell growth. It has been
Nanotechnology in effect relies upon manipulating demonstrated, for example, that the direction of
atoms into specific locations in order to construct axon extension from nerve cells can be controlled
such novel devices. It is evident that methods being by printing lm-scale lines of proteins on glass
developed for organic functionalization can be surfaces [94,95]. Directed cell growth, in turn, can
used to produce films of molecules on semicon- be used for tissue engineering. Organic modifica-
ductor surfaces with nanometer scale thicknesses, tion can also be used to impart biocompatibility
i.e. one molecule high. Achieving nanometer scale to semiconductor-based devices designed to be
control in the lateral direction is also possible. One implanted into living organisms.
example is the use of radical-initiated hydrosily- There are clearly a large number of applications
lation of alkenes to form self-limiting islands or for which organic functionalization of semicon-
lines of organic material several approximately 10– ductor surfaces can have a great impact. Each of
100 nm in size [84,87]. Other possibilities include these applications requires a deep understanding
writing complex patterns in attached organic lay- of how to modify the semiconductor surfaces
ers with the tip of a STM [45], or utilizing poly- controllably. Methods to incorporate organic
merization in the organic layer. The opportunities functionality to the surfaces of semiconductors
to make contributions in the area of nanotech- have seen immense progress in recent years. We
nology by creative use of functionalization strate- have shown in this review how direct, covalent
gies are vast. attachment of organic molecules at the surface is
Another area closely tied to organic function- being used to bind organic layers onto semicon-
alization is molecular recognition and chemical ductors. Methods developed for both dry and wet
or biological sensing. Organic molecules are key environments can impart excellent control of the
players in molecular recognition. Coupling organic interfacial properties. For example, organic layers
molecules to semiconductor solids provides op- can be deposited which protect the underlying
portunity to fabricate integrated chip-based sen- substrate from attack by etchants, impart chiral
sors. In such a scheme, the organic layer provides properties to the surface, or add new optical
the molecular recognition function, while the properties to silicon.
semiconductor provides capabilities for signal A number of challenges remain, however, before
processing, data storage, logic, and even wireless many of the futuristic applications––in nanotech-
communication. Using this approach, the fabri- nology, sensing, and others areas––can be realized
cation of compact sensor that can be used out in using this technology. One future need is to in-
the field and communicate data remotely can be crease the degree of control over both the selec-
envisioned. Key to such a device is the attachment tivity (i.e., in producing one product over
of an organic layer on the semiconductor sub- competing products) and the degree of order in the
strate, where the organic layer can be engineered organic layer. This in turn requires a deeper un-
with a variety of end groups designed for molec- derstanding of the driving forces that control the
ular recognition. It remains a challenge to develop attachment reactions. Evidence points to the im-
strategies for incorporating molecular receptors of portance of both the reaction rates (kinetics) and
a general nature onto the semiconductor surface. the energetic stability of the products (thermody-
Even bioengineering stands to gain significantly namics) in controlling these reaction systems.
from advances made in organic surface modifica- Therefore, it is important to understand how to
tion. Surface modification is known to play a de- design the system (by choice of reaction conditions
900 S.F. Bent / Surface Science 500 (2002) 879–903

or molecular precursors) in order to manipulate figures. The National Science Foundation (Grants
the product distribution. More work is also needed nos. DMR-9896333 and CHE-9900041) and the
in determining the detailed reaction mechanisms. Beckman Foundation are gratefully acknowledged
One of the key challenges is the ability to deposit for financial support which made this review pos-
multiple organic layers in a controllable fashion. sible. The author also thanks the Camille and
So far, most of the work in organic functional- Henry Dreyfus Foundation for support through a
ization by covalent Si–C attachment has focused Camille Dreyfus Teacher–Scholar award.
on the initial monolayer. There are some nice ex-
ceptions, however, such as the work by Bitzer and
References
Richardson [96]. It is clear that many applications
require flexibility in the deposition of multiple [1] S.A. Campbell, The Science and Engineering of Micro-
layers. In order to extend the organic modification electronic Fabrication, Oxford, New York, 1996.
to allow for next-layer attachment, a bifunctional [2] C.Y. Chang, S.M. Sze (Eds.), ULSI Technology, McGraw-
or polyfunctional organic molecule must be used Hill, New York, 1996.
[3] International Technology Roadmap for Semiconductors,
for the first layer. This layer, in turn, will retain a
Semiconductor Industry Association, 1999.
reactive functional group for further attachment. [4] See, for example: (a) J.R. Arthur, Molecular beam epitaxy,
This review, and indeed most of the work on or- Surf. Sci. 500 (2002) 189;
ganic functionalization by a covalent Si–C linkage, (b) P. Finnie, Y. Homma, Epitaxy: the motion picture,
has focused on the alkene functional group. Re- Surf. Sci. 500 (2002) 437.
[5] M.K. Weldon, K.T. Queeney, J. Eng Jr., K. Raghavachari,
search on other functional groups is in its infancy.
Y.J. Chabal, The surface science of semiconductor pro-
There is clearly a great need not only for under- cessing, Surf. Sci. 500 (2002) 859.
standing how different functional groups react at [6] J.W. Hill, D.K. Kolb, Chemistry for Changing Times,
the semiconductor surface, but also for learning Prentice Hall, Englewood Cliffs, NJ, 1995.
how to direct the reaction of a polyfunctional [7] J.T. Yates Jr., A new opportunity in silicon-based micro-
electronics, Science 279 (1998) 335.
molecule in a stepwise fashion.
[8] J.M. Buriak, Organometallic chemistry on silicon surfaces:
The application of organic chemistry at semi- formation of functional monolayers bound through Si–C
conductor surfaces for forming organic monolay- bonds, Chem. Commun. 12 (1999) 1051.
ers and thin films is becoming an important and [9] R.A. Wolkow, Controlled molecular adsorption on silicon:
growing area. By focusing on the molecular nature laying a foundation for molecular devices, Ann. Rev. Phys.
Chem. 50 (1999) 413.
of the surface, the functionalization methods ulti-
[10] R.J. Hamers, S.K. Coulter, M.D. Ellison, J.S. Hovis, D.F.
mately will provide the potential for control over Padowitz, M.P. Schwartz, C.M. Greenlief, J.N. Russell Jr.,
the monolayer bonding and order at the surface. Cycloaddition chemistry of organic molecules with semi-
The impact of these studies on emerging technol- conductor surfaces, Acc. Chem. Res. 33 (2000) 617.
ogies in microelectronics, sensing, and photonics is [11] S.F. Bent, Attaching organic layers to semiconductor
surfaces, J. Phys. Chem. B, submitted.
expected to be significant in the coming decades.
[12] J.J. Boland, The importance of structure and bonding in
semiconductor surface chemistry: hydrogen on the Si(1 1 1)-
7  7 surface, Surf. Sci. 244 (1991) 1.
Acknowledgements [13] A. Zangwill, Physics at Surfaces, Cambridge University
Press, New York, 1988.
[14] C.B. Duke, Semiconductor surface reconstruction: the
The author expresses appreciation for the many
structural chemistry of two-dimensional surface com-
fruitful interactions with students and colleagues pounds, Chem. Rev. 96 (1996) 1237.
that contributed immensely to this review. Thanks [15] H.N. Waltenberg, J.T. Yates Jr., Surface chemistry of
are also due to Prof. Charles Musgrave, Collin silicon, Chem. Rev. 95 (1995) 1589.
Mui, George Wang, Michael Filler, and Michael [16] R.J. Hamers, R.M. Tromp, J.E. Demuth, Electronic and
geometric structure of Si(1 1 1)-(7  7) and Si(0 0 1) sur-
Manzo for their contribution to the figures, and to
faces, Surf. Sci. 181 (1987) 346.
Prof. Jillian Buriak, Prof. Christopher Chidsey, [17] J.J. Boland, Scanning tunneling microscopy of the inter-
Prof. Robert Hamers and Dr. Robert Wolkow for action of hydrogen with silicon surfaces, Adv. Phys. 42
graciously providing permission to reprint their (1993) 129.
S.F. Bent / Surface Science 500 (2002) 879–903 901

[18] G.S. Higashi, Y.J. Chabal, G.W. Trucks, K. Raghava- [36] J.S. Hovis, R.J. Hamers, Structure and bonding of ordered
chari, Ideal hydrogen termination of the Si(1 1 1) surface, organic monolayers of 1,3,5,7-cyclooctatetraene on the
Appl. Phys. Lett. 56 (1990) 656. Si(0 0 1) surface: surface cycloaddition chemistry of an
[19] A. Wassermann, Diels–Alder Reactions: Organic Back- antiaromatic molecule, J. Phys. Chem. 102 (1998) 687.
ground and Physico-Chemical Aspect, Elsevier, New York, [37] G.P. Lopinski, T.M. Fortier, D.J. Moffatt, R.A. Wolkow,
1965. Multiple bonding geometries and binding state conversion
[20] G.B. Gill, M.R. Willis, Pericyclic Reactions, Chapman and of benzene/Si(1 0 0), J. Vac. Sci. Technol. A 16 (1998) 1037.
Hall, London, 1974. [38] G.P. Lopinski, D.J. Moffatt, D.D.M. Wayner, R.A.
[21] W. Carruthers, Cycloaddition Reactions in Organic Syn- Wolkow, Determination of the absolute chirality of indi-
thesis, Pergamon Press, New York, 1990. vidual molecules using the scanning tunneling microscope,
[22] R.B. Woodward, R. Hoffmann, The Conservation of Nature 392 (1998) 909.
Orbital Symmetry, Academic Press, New York, 1970. [39] G.P. Lopinski, D.J. Moffatt, D.D.M. Wayner, M.Z.
[23] R. Konecny, D. Doren, Theoretical prediction of a facile Zgierski, R.A. Wolkow, Asymmetric induction at a silicon
Diels–Alder reaction on the Si(1 0 0)-2  1 surface, J. Am. surface, J. Am. Chem. Soc. 121 (1999) 4532.
Chem. Soc. 119 (1997) 11098. [40] S. Gokhale, P. Trischberger, D. Menzel, W. Widdra, H.
[24] R. Konecny, D.J. Doren, Cycloaddition reactions of Dr€oge, H.P. Steinr€ uck, U. Birkenheuer, U. Gutdeutsch, N.
unsaturated hydrocarbons on the Si(1 0 0)-(2  1) surface: R€osch, Electronic structure of benzene adsorbed on single-
theoretical predictions, Surf. Sci. 417 (1998) 169. domain Si(0 0 1)-2  1: a combined experimental and the-
[25] C. Mui, S.F. Bent, C.B. Musgrave, A theoretical study oretical study, J. Chem. Phys. 108 (1998) 5554.
of the structure and thermochemistry of 1,3-butadiene on [41] B. Borovsky, M. Krueger, E. Ganz, Metastable adsorption
the Ge/Si(1 0 0)-2  1 surface, J. Phys. Chem. 104 (2000) of benzene on the Si(0 0 1) surface, Phys. Rev. B 57 (1998)
2457. 4269.
[26] A.V. Teplyakov, M.J. Kong, S.F. Bent, Vibrational [42] U. Birkenheuer, U. Gutdeutsch, N. Rosch, Geometrical
spectroscopic studies of Diels–Alder reactions with the structure of benzene absorbed on Si(0 0 1), Surf. Sci. 409
Si(1 0 0)-2  1 surface as a dienophile, J. Am. Chem. Soc. (1998) 213.
119 (1997) 11100. [43] K.W. Self, R.I. Pelzel, J.H.G. Owen, C. Yan, W. Widdra,
[27] A.V. Teplyakov, M.J. Kong, S.F. Bent, Diels–Alder W.H. Weinberg, Scanning tunneling microscopy study of
reactions of butadienes with the Si(1 0 0)-2  1 surface as benzene adsorption on Si(1 0 0)-2  1, J. Vac. Sci. Technol.
a dienophile: vibrational spectroscopy, thermal desorption A 16 (1998) 1031.
and NEXAFS studies, J. Chem. Phys. 108 (1998) 4599. [44] J.W. Lyding, T.C. Shen, G.C. Abeln, C. Wang, P.A. Scott,
[28] M.J. Kong, A.V. Teplyakov, J.G. Lyubovitsky, S.F. Bent, J.R. Tucker, P. Avouris, R.E. Walkup, Ultrahigh-vacuum
NEXAFS studies of adsorption and reaction of benzene scanning tunneling microscope-based nanolithography and
on Si(1 0 0)-2  1, Surf. Sci. 411 (1998) 286. selective chemistry on silicon surfaces, Isr. J. Chem. 36
[29] G.T. Wang, C. Mui, C.B. Musgrave, S.F. Bent, Cyclo- (1996) 3.
addition of cyclopentadiene and dicyclopentadiene on [45] G.C. Abeln, S.Y. Lee, J.W. Lyding, D.S. Thompson, J.S.
Si(1 0 0)-2  1: comparison of monomer and dimer adsorp- Moore, Nanopatterning organic monolayers on Si(1 0 0) by
tion, J. Phys. Chem. 103 (1999) 6803. selective chemisorption of norbornadiene, Appl. Phys.
[30] M.J. Kong, A.V. Teplyakov, J. Lyubovitsky, J. Jagmohan, Lett. 70 (1997) 2747.
S.F. Bent, Interaction of C6 cyclic hydrocarbons with a [46] A.J. Mayne, A.R. Avery, J. Knall, T.S. Jones, G.A.D.
Si(1 0 0)-2  1 surface: adsorption and hydrogenation reac- Briggs, W.H. Weinberg, An STM study of the chemisorp-
tions, J. Phys. Chem. 104 (2000) 3000. tion of C2 H4 on Si(1 0 0)(2  1), Surf. Sci. 284 (1993)
[31] H. Liu, R.J. Hamers, Stereoselectivity in molecule–surface 247.
reactions: adsorption of ethylene on the Si(0 0 1) surface, [47] C. Huang, W. Widdra, W.H. Weinberg, Adsorption of
J. Am. Chem. Soc. 119 (1997) 7593. ethylene on the Si(1 0 0)-2  1 surface, Surf. Sci. 315 (1994)
[32] R.J. Hamers, J.S. Hovis, S. Lee, H. Liu, J. Shan, L953.
Formation of ordered, anisotropic organic monolayers on [48] L. Clemen, R.M. Wallace, P.A. Taylor, M.J. Dresser, W.J.
the Si(0 0 1) surface, J. Phys. Chem. B 101 (1997) 1489. Choyke, W.H. Weinberg, J.T. Yates Jr., Adsorption and
[33] J.S. Hovis, R.J. Hamers, Structure and bonding of ordered thermal behavior of ethylene on Si(1 0 0)-(2  1), Surf. Sci.
organic monolayers of 1,5-cyclooctadiene on the Si(1 0 0) 268 (1992) 205.
surface, J. Phys. Chem. 101 (1997) 9581. [49] W. Widdra, C. Huang, S.I. Yi, W.H. Weinberg, Coad-
[34] J.S. Hovis, H. Liu, R.J. Hamers, Cycloaddition chemistry sorption of hydrogen with ethylene and acetylene on
of 1,3-dienes on the silicon(0 0 1) surface: competition Si(1 0 0)-(2  1), J. Chem. Phys. 105 (1996) 5605.
between ½4 þ 2 and ½2 þ 2 reactions, J. Phys. Chem. 102 [50] S.H. Xu, M. Keeffe, Y. Yang, C. Chen, M. Yu, G.J.
(1998) 6873. Lapeyre, F. Rotenberg, J. Denlinger, J.T. Yates Jr.,
[35] J.S. Hovis, H. Liu, R.J. Hamers, Cycloaddition chemistry Photoelectron diffraction imaging for C2 H2 and C2 H4
and formation of ordered organic monolayers on sili- chemisorbed on Si(1 0 0) reveals a new bonding configura-
con(0 0 1) surfaces, Surf. Sci. 402 (1998) 1. tion, Phys. Rev. Lett. 84 (2000) 939.
902 S.F. Bent / Surface Science 500 (2002) 879–903

[51] D.C. Sorescu, K.D. Jordan, Theoretical study of the [66] A.J. Fisher, P.E. Blchl, G.A.D. Briggs, Hydrocarbon
adsorption of acetylene on the Si(1 0 0) surface, J. Phys. Adsorption on Si(0 0 1): when does the Si dimer bond
Chem. B 104 (2000) 8259. break?, Surf. Sci. 374 (1997) 298.
[52] P. Lal, A.V. Teplyakov, Y. Noah, M.J. Kong, G.T. Wang, [67] W. Pan, T. Zhu, W. Yang, First-principles study of the
S.F. Bent, Adsorption of ethylene on the Ge(1 0 0)-2  1 structural and electronic properties of ethylene adsorption
surface: coverage and time-dependent behavior, J. Chem. on Si(1 0 0)-(2  1) surface, J. Chem. Phys. 107 (1997) 3981.
Phys. 110 (1999) 10545. [68] B.Q. Meng, D. Maroudas, W.H. Weinberg, Structure of
[53] A.V. Teplyakov, P. Lal, Y.A. Noah, S.F. Bent, Evidence chemisorbed acetylene on the Si(0 0 1)-(2  1) surface and
for a retro-Diels–Alder reaction on a single crystalline the effect of coadsorbed atomic hydrogen, Chem. Phys.
surface: butadienes on Ge(1 0 0), J. Am. Chem. Soc. 120 Lett. 278 (1997) 97.
(1998) 7377. [69] G.P. Lopinski, D.J. Moffatt, D.D.M. Wayner, R.A.
[54] R.J. Hamer, J.S. Hovis, C.M. Greenlief, D.F. Padowitz, Wolkow, How stereoselective are alkene addition reactions
Scanning tunneling microscopy of organic molecules and on Si(1 0 0)?, J. Am. Chem. Soc. 122 (2000) 3548.
monolayers on silicon and germanium (0 0 1) surfaces, Jpn. [70] L.E. Larson, High-speed Si/SiGe technology for next
J. Appl. Phys. 38 (1999) 3879. generation wireless system applications, J. Vac. Sci. Tech-
[55] S.W. Lee, L.N. Nelen, H. Ihm, T. Scoggins, C.M. nol. B 16 (1998) 1541.
Greenlief, Reaction of 1,3-cyclohexadiene with the [71] B.S. Meyerson, Silicon:germanium-based mixed-signal
Ge(1 0 0) surface, Surf. Sci. 410 (1998) L773. technology for technology for optimization of wired and
[56] G.T. Wang, S.F. Bent, J.N. Russell Jr., J.E. Butler, wireless telecommunications, IBM J. Res. Develop. 44
M.P. D’Evelyn, Functionalization of diamond(1 0 0) (2000) 391.
by Diels–Alder chemistry, J. Am. Chem. Soc. 122 (2000) [72] J.E. Field, The Properties of Natural and Synthetic
744. Diamond, Academic Press, London, 1992.
[57] J.S. Hovis, S.K. Coulter, R.J. Hamers, M.P. D’Evelyn, [73] (a) J.A. Kubby, J.E. Griffith, R.S. Becker, J.S. Vickers,
J.N. Russell, J.E. Butler, Cycloaddition chemistry at Tunneling microscopy of Ge(0 0 1), Phys. Rev. B 36 (1987)
surfaces: reaction of alkenes with the diamond(0 0 1)-2  1 6079;
surface, J. Am. Chem. Soc. 122 (2000) 732. (b) Y.J. Chabal, High resolution infrared spectroscopy of
[58] M.Z. Hossain, T. Aruga, N. Takagi, T. Tsuno, N. adsorbates on semiconductor surfaces: hydrogen on
Fujimori, T. Ando, M. Nishijima, Diels–Alder reaction Si(1 0 0) and Ge(1 0 0), Surf. Sci. 168 (1986) 594;
on the clean diamond(1 0 0) 2  1 surface, Jpn. J. Appl. (c) X. Torrelles, H.A. van der Vegt, V.H. Etgens, P.
Phys. 38 (1999) L1496. Fajardo, J. Alvarez, S. Ferrer, The structure of the
[59] M.J. Bozack, W.J. Choyke, L. Muehlhoff, J.T. Yates Jr., Ge(0 0 1)-(2  1) reconstruction investigated with X-ray
Reaction chemistry at the Si(1 0 0) surface-control diffraction, Surf. Sci. 364 (1996) 242.
through active-site manipulation, J. Appl. Phys. 60 (1986) [74] (a) B.D. Thoms, J.E. Butler, HREELS and LEED of H/
3750. C(1 0 0): the 2  1 monohydride dimer row reconstruction,
[60] M. Nishijima, J. Yoshinobu, H. Tsuda, M. Onchi, The Surf. Sci. 328 (1995) 291;
adsorption and thermal-decomposition of acetylene on (b) T.W. Mercer, P.E. Pehrsson, Surface state transitions
Si(1 0 0) and vicinal Si(1 0 0)9-degrees, Surf. Sci. 192 (1987) on the reconstructed diamond C(1 0 0) surface, Surf. Sci.
383. 399 (1998) L327;
[61] J. Yoshinobu, H. Tsuda, M. Onchi, M. Nishijima, The (c) M.P. D’Evelyn, in: Handbook of Industrial Diamonds
adsorbed states of ethylene on Si(1 0 0)c(4  2), Si(1 0 0)- and Diamond Films, Dekker, New York, 1998.
(2  1), and vicinal Si(1 0 0): electron energy loss spectro- [75] D.R. Fitzgerald, D.J. Doren, Functionalization of dia-
scopy and low-energy electron diffraction studies, J. Chem. mond(1 0 0) by cycloaddition of butadiene: first-principles
Phys. 87 (1987) 7332. theory, J. Am. Chem. Soc. 122 (2000) 12334.
[62] C.C. Cheng, R.M. Wallace, P.A. Taylor, W.J. Choyke, J.T. [76] M.R. Linford, C.E.D. Chidsey, Alkyl monolayers cova-
Yates Jr., Direct determination of absolute monolayer lently bonded to silicon surfaces, J. Am. Chem. Soc. 115
coverages of chemisorbed C2 H2 and C2 H4 on Si(1 0 0), (1993) 12631.
J. Appl. Phys. 67 (1990) 3693. [77] M.R. Linford, P. Fenter, P.M. Eisenberger, C.E.D. Chid-
[63] P.A. Taylor, R.M. Wallace, C.C. Cheng, W.H. Weinberg, sey, Alkyl monolayers on silicon prepared from 1-alkenes
M.J. Dresser, W.J. Choyke, J.T. Yates Jr., Adsorption and and hydrogen-terminated silicon, J. Am. Chem. Soc. 117
decomposition of acetylene on Si(1 0 0)-(2  1), J. Am. (1995) 3145.
Chem. Soc. 114 (1992) 6754. [78] J.M. Buriak, M.L. Allen, Lewis acid mediated functional-
[64] Q. Liu, R. Hoffmann, The bare and acetylene chemisorbed ization of porous silicon with substituted alkenes and al-
Si(1 0 0) surface, and the mechanism of acetylene chemi- kynes, J. Am. Chem. Soc. 120 (1998) 1339.
sorption, J. Am. Chem. Soc. 117 (1995) 4082. [79] J.M. Buriak, M.P. Stewart, T.W. Geders, M.J. Allen, H.C.
[65] B.I. Craig, P.V. Smith, Structures of small hydrocarbons Choi, J. Smith, D. Raftery, L.T. Canham, Lewis acid
adsorbed on Si(0 0 1) and Si terminated Beta-SiC(0 0 1), mediated hydrosilylation on porous silicon surfaces, J. Am.
Surf. Sci. 276 (1992) 174. Chem. Soc. 121 (1999) 11491.
S.F. Bent / Surface Science 500 (2002) 879–903 903

[80] A. Bansal, X. Li, I. Lauermann, N.S. Lewis, S.I. Yi, W.H. monolayers to the Si(1 1 1) surface using chemical-shift,
Weinberg, Alkylation of Si surfaces using a two-step scanned-energy photoelectron diffraction, Appl. Phys. Lett.
halogenation/Grignard route, J. Am. Chem. Soc. 118 71 (1997) 1056.
(1996) 7225. [89] J. Terry, M.R. Linford, C. Wigren, R.Y. Cao, P. Pianetta,
[81] J.H. Song, M.J. Sailor, Functionalization of nanocrystal- C.E.D. Chidsey, Alkyl-terminated Si(1 1 1) surfaces: a high-
line porous silicon surfaces with aryllithium reagents: resolution core-level photoelectron spectroscopy study, J.
formation of silicon–carbon bonds by cleavage of silicon– Appl. Phys. 85 (1999) 213.
silicon bonds, J. Am. Chem. Soc. 120 (1998) 2376. [90] M.P. Stewart, J.M. Buriak, Photopatterned hydrosilyla-
[82] N.Y. Kim, P.E. Laibinis, Derivatization of porous silicon tion on porous silicon, Angew. Chem. Int. Ed. 37 (1998)
by grignard reagents at room temperature, J. Am. Chem. 3257.
Soc. 120 (1998) 4516. [91] C.P. Collier, E.W. Wong, M. Belohradsky, F.M. Raymo,
[83] N.Y. Kim, P.E. Laibinis, Improved polypyrrole/silicon J.F. Stoddart, P.J. Kuekes, R.S. Williams, J.R. Heath,
junctions by surfacial modification of hydrogen-terminated Electronically configurable molecular-based logic gates,
silicon using organolithium reagents, J. Am. Chem. Soc. Science 285 (1999) 391.
121 (1999) 7162. [92] S. Zhang, L. Yan, M. Altman, M. Lssle, H. Nugent, F.
[84] R.L. Cicero, C.E.D. Chidsey, G.P. Lopinski, D.D.M. Frankel, D.A. Lauffenburger, G.M. Whitesides, A. Rich,
Wayner, R.A. Wolkow, Olefin additions on H-Si(1 1 1): Biological surface engineering: a simple system for cell
evidence for a surface chain reaction initiated at isolated pattern formation, Biomaterials 20 (1999) 1213.
dangling bonds, Langmuir, in press. [93] Y. Ito, Surface micropatterning to regulate cell functions,
[85] T.C. Shen, C. Wang, G.C. Abeln, J.R. Tucker, J.W. Biomaterials 20 (1999) 2333.
Lyding, P. Avouris, R.E. Walkup, Atomic-scale desorption [94] P. Clark, S. Britland, P. Connolly, Growth cone guidance
through electronic and vibrational excitation mechanisms, and neuron morphology on micropatterned laminim sur-
Science 268 (1995) 1590. faces, J. Cell Sci. 105 (1993) 203.
[86] P. Avouris, R.E. Walkup, A.R. Rossi, H.C. Akpati, P. [95] C.D. James, R. Davis, M. Meyer, A. Turner, S. Turner, G.
Nordlander, T.C. Shen, G.C. Abeln, J.W. Lyding, Break- Withers, L. Kam, G. Banker, H. Craighead, M. Isaacson,
ing individual chemical bonds via STM-induced excita- J. Turner, W. Shain, Aligned microcontact printing of
tions, Surf. Sci. 363 (1996) 368. micrometer-scale poly-L -lysine structures for controlled
[87] G.P. Lopinski, D.D.M. Wayner, R.A. Wolkow, Self- growth of cultured neurons on planar microelectrode
directed growth of molecular nanostructures on Si(1 0 0), arrays, IEEE Trans. Biomed. Eng. 47 (2000) 17.
Nature 406 (2000) 48. [96] T. Bitzer, N.V. Richardson, Route for controlled growth of
[88] J. Terry, M.R. Linford, C. Wigren, R. Cao, P. Pianetta, ultrathin polyimide films with Si–C bonding to Si(1 0 0)-
C.E.D. Chidsey, Determination of the bonding of alkyl 2  1, Surf. Sci. 144 (1999) 339.

You might also like