Nanolithography and Nanochemistry: Probe ... - Multiple Choices

Nanolithography and Nanochemistry: Probe ... - Multiple Choices Nanolithography and Nanochemistry: Probe ... - Multiple Choices

schubert.group.de
from schubert.group.de More from this publisher
26.06.2014 Views

Reviews U. S. Schubert and D. Wouters Nanolithography Nanolithography and Nanochemistry: Probe-Related Patterning Techniques and Chemical Modification for Nanometer-Sized Devices Daan Wouters and Ulrich S. Schubert* Keywords: atomic force microscopy · lithography · nanostructures · oxidation · scanning tunneling microscopy Dedicated to Professor Dieter Schubert on the occasion of his 65th birthday Angewandte Chemie 2480 2004 Wiley-VCH Verlag GmbH &Co. KGaA, Weinheim DOI: 10.1002/anie.200300609 Angew. Chem. Int. Ed. 2004, 43, 2480 – 2495

Reviews<br />

U. S. Schubert <strong>and</strong> D. Wouters<br />

<strong>Nanolithography</strong><br />

<strong>Nanolithography</strong> <strong>and</strong> <strong>Nanochemistry</strong>: <strong>Probe</strong>-Related<br />

Patterning Techniques <strong>and</strong> Chemical Modification for<br />

Nanometer-Sized Devices<br />

Daan Wouters <strong>and</strong> Ulrich S. Schubert*<br />

Keywords:<br />

atomic force microscopy · lithography ·<br />

nanostructures · oxidation ·<br />

scanning tunneling microscopy<br />

Dedicated to Professor Dieter Schubert<br />

on the occasion of his 65th birthday<br />

Angew<strong>and</strong>te<br />

Chemie<br />

2480 2004 Wiley-VCH Verlag GmbH &Co. KGaA, Weinheim DOI: 10.1002/anie.200300609 Angew. Chem. Int. Ed. 2004, 43, 2480 – 2495


<strong>Nanolithography</strong><br />

Angew<strong>and</strong>te<br />

Chemie<br />

The size regime for devices produced by photolithographic techniques<br />

is limited. Therefore, other patterning techniques have been intensively<br />

studied to create smaller structures. Scanning-probe-based patterning<br />

techniques, such as dip-pen lithography, local force-induced<br />

patterning, <strong>and</strong> local-probe oxidation-based techniques are highly<br />

promising because of their relative ease <strong>and</strong> widespread availability.<br />

The latter of these is especially interesting because of the possibility of<br />

producing nanopatterns for a broad range of chemical <strong>and</strong> physical<br />

modification <strong>and</strong> functionalization processes; both the production of<br />

nanometer-sized electronic devices <strong>and</strong> the formation of devices<br />

involving (bio)molecular recognition <strong>and</strong> sensor applications is<br />

possible. This Review highlights the development of various scanning<br />

probe systems <strong>and</strong> the possibilities of local oxidation methods, as well<br />

as giving an overview of state-of-the-art nanometer-sized devices, <strong>and</strong><br />

a view of future development.<br />

1. Introduction<br />

Nanotechnology is currently at the center of attention in<br />

both the media <strong>and</strong> in public opinion as well as in<br />

governmental <strong>and</strong> industrial research programs. Large funding<br />

programs related to nanotechnology were recently<br />

initiated in Europe, Japan, <strong>and</strong> the USA. [1,2] This development<br />

is reflected by a shift of attention in academic <strong>and</strong><br />

industrial research centers. The envisioned fields <strong>and</strong> their<br />

potential technological impact on society range from “smart”<br />

drugs or sensitive DNA/protein sensors to nanostructured<br />

electronic devices with applications in information storage,<br />

optical computers, solar cells, <strong>and</strong> energy storage, as well as<br />

novel materials with improved properties such as “intelligent”<br />

<strong>and</strong> self-cleaning or color-changing coatings. [3–6] It should be<br />

noted that many of these envisioned improved properties<br />

relate to changes <strong>and</strong>/or control of surface properties. This<br />

directly implies the need for well-defined, controllable,<br />

reliable, <strong>and</strong> affordable surface modification <strong>and</strong> characterization<br />

techniques. In general, two approaches for the<br />

creation of nanostructured surfaces are available. The first,<br />

the so-called “top-down” approach, epitomizes efforts by<br />

physicists <strong>and</strong> engineers to reduce the size of structures with<br />

techniques that are currently available. Most results using this<br />

strategy originate from the downscaling of lithographic<br />

techniques in semiconductor research. The second “bottomup”<br />

approach has been developed by (bio)chemists, who have<br />

considered the self-assembling processes in nature when<br />

creating similar patterns that utilize noncovalent molecular<br />

interactions such as hydrogen bonding, p–p stacking, <strong>and</strong>/or<br />

metal–lig<strong>and</strong> coordination. [7–11] Although the bottom-up<br />

approach is clearly interesting, <strong>and</strong> can lead to unique <strong>and</strong><br />

interesting structures, it also has some clear disadvantages:<br />

The ability of patterning alone is insufficient for successful<br />

application in advanced (nano)devices, since besides size <strong>and</strong><br />

structure, control over both direction <strong>and</strong> position is also<br />

essential. Moreover, the ability to characterize the created<br />

From the Contents<br />

1. Introduction 2481<br />

2. Physical <strong>Probe</strong> Lithography<br />

Techniques 2483<br />

3. Oxidative <strong>Probe</strong> Lithography<br />

Techniques 2486<br />

4. Conclusions <strong>and</strong> Outlook 2493<br />

structures <strong>and</strong> the feasibility of producing<br />

these structures on large areas<br />

is of central interest.<br />

As a top-down approach, lithography<br />

represents an important surfacepatterning<br />

technique. Lithography was<br />

introduced in 1798 as a printing tool by<br />

Alois Senefelder, who found that ink adsorbed to an image<br />

drawn with a greasy fluid onto limestone (lithographic stones,<br />

Figure 1) could be used to transfer the image to a piece of<br />

Figure 1. Lithographic stones were introduced by Alois Senefelder. [12]<br />

paper pressed against the stone. [12–15] Since then, this printing<br />

technique has evolved dramatically <strong>and</strong> is nowadays not only<br />

used for printing but also for the (chemical) modification of<br />

substrates on a submicrometer scale, <strong>and</strong> is often referred to<br />

as microcontact printing. [16–18] The most prominent application<br />

of lithography is that of UV-mask lithography, which is<br />

widely used in the semiconductor industry. With the development<br />

of (computer) chips <strong>and</strong> their increasing complexity, the<br />

requirement for techniques able to create smaller structures is<br />

ongoing. Therefore, industry <strong>and</strong> academia invest intensive<br />

[*] D. Wouters, Prof. Dr. U. S. Schubert<br />

Laboratory of Macromolecular Chemistry <strong>and</strong> Nanoscience<br />

Eindhoven University of Technology <strong>and</strong><br />

the Dutch Polymer Institute (DPI)<br />

P.O. Box 513, 5600 MB, Eindhoven (The Netherl<strong>and</strong>s)<br />

Fax: (+ 31)40-247-4186<br />

E-mail: u.s.schubert@tue.nl<br />

Prof. Dr. U. S. Schubert<br />

Center for NanoScience, Ludwig-Maximilians-Universität München<br />

Geschwister-Scholl Platz 1, 80333 München (Germany)<br />

Angew. Chem. Int. Ed. 2004, 43, 2480 – 2495 DOI: 10.1002/anie.200300609 2004 Wiley-VCH Verlag GmbH &Co. KGaA, Weinheim<br />

2481


Reviews<br />

U. S. Schubert <strong>and</strong> D. Wouters<br />

effort in developing UV <strong>and</strong> deep-UV lithographic techniques,<br />

which can provide structural resolution down to<br />

80 nm. [19] At the same time, particle-, ion-, <strong>and</strong> electronbeam<br />

lithographic techniques have become available for<br />

structuring substrates at the 50-nm level. Smaller dimensions<br />

are possible, but the costs rise enormously <strong>and</strong> the addressable<br />

substrate sizes decrease enormously. Both UV, particle,<br />

<strong>and</strong> e-beam lithography are well documented, <strong>and</strong> a large<br />

number of recent publications are available. [20–26] A publication<br />

by Grunze et al. is of particular note because it introduces<br />

the possibility for chemical lithography using an electron<br />

beam. Focused electrons are used to chemically modify the<br />

terminal nitro groups of self-assembled monolayers (SAMs)<br />

to amines; the subsequent functionalization with carboxylic<br />

acid anhydrides was then demonstrated (see Figure 2). [27,28]<br />

Also of note is the preparation of ordered arrays of metallic<br />

nanoparticles <strong>and</strong> metallic rods using the self-assembly of<br />

metal-containing block copolymer micelles onto electronbeam-patterned<br />

substrates, effectively combining techniques<br />

[29, 30]<br />

from the bottom-up <strong>and</strong> the top-down approaches.<br />

As a result of decreasing pattern size, the dem<strong>and</strong> for<br />

microscopic characterization techniques has also increased.<br />

Since their invention in 1982, [31] scanning tunneling microscopy<br />

(STM) <strong>and</strong> atomic force microscopy (AFM) have<br />

proven to be effective techniques capable of identifying the<br />

structure <strong>and</strong> nature of surfaces <strong>and</strong> adsorbates on substrates<br />

down to molecular <strong>and</strong> atomic resolution. The application of<br />

AFM <strong>and</strong> STM has not been limited to inorganic semiconductor<br />

substrates: also in polymer <strong>and</strong> supramolecular<br />

research has their value also been demonstrated by the<br />

elucidation of, for example, polymer phase separation <strong>and</strong><br />

molecular ordering. [32–34] In biochemistry, the techniques have<br />

been applied for the determination of interaction forces<br />

between proteins <strong>and</strong> receptors. [35–38] Soon after the invention<br />

of AFM <strong>and</strong> the observation of its ability to modify substrates,<br />

the field of scanning-probe-based lithography (SPL) was<br />

developed. [39] The application of probe-based techniques for<br />

the modification of substrates has been widespread <strong>and</strong> a<br />

large variety of techniques have evolved, which range from<br />

the subtle movement of atoms (using STM), [40] the formation<br />

of local deformations in soft substrates using high-contact<br />

force AFM (see Section 2.1) to the local application of “inks”<br />

(dip-pen lithography) <strong>and</strong> the local oxidation of suitable<br />

substrates (probe oxidation, see Sections 2.2 <strong>and</strong> 2.3). With<br />

Figure 2. Electron-induced chemical lithography using self-assembled<br />

monolayers (SAMs) as a platform for the generation of patterns. Scanning<br />

force microscopy images show the selective binding of acetic acid<br />

anhydride (R 1 = CH 3 ) <strong>and</strong> perfluorobutyric acid (R 2 = C 3 F 7 ) onto a SAM<br />

platform. [27]<br />

the application of probe-based techniques, control over<br />

position <strong>and</strong> direction is evident. Moreover, due to the very<br />

recently developed multi-probe systems (see Section 2.3) <strong>and</strong><br />

Daan Wouters studied chemical engineering<br />

(1995–2000) at the Eindhoven University of<br />

Technology (The Netherl<strong>and</strong>s). During this<br />

period he spent 4 months at DSM Desotech<br />

(Elgin, USA) on a project regarding UV-curable<br />

coatings. He graduated in 2000 under<br />

the supervision of Dr. P. C. Thüne (the Surface<br />

Science group of Prof. J. W. Niemantsverdriet)<br />

working on a surface-science model<br />

of the Philips catalyst. Since 2001, he has<br />

studied for a PhD with Prof. U. S. Schubert<br />

at the same university. His research centers<br />

on scanning probe lithography <strong>and</strong> nanostructured<br />

surfaces.<br />

Ulrich S. Schubert was born in 1969 (Tübingen,<br />

Germany). He studied chemistry <strong>and</strong><br />

biochemistry in Frankfurt, Bayreuth, <strong>and</strong><br />

Richmond (USA). He obtained his PhD<br />

(1995) under the supervision of Profs. C. D.<br />

Eisenbach (Bayreuth) <strong>and</strong> G. R. Newkome<br />

(Tampa). After postdoctoral training with<br />

Prof. J.-M. Lehn (Strasbourg), he obtained<br />

his habilitation with Prof. O. Nuyken (TU<br />

München) in 1999, followed by a year at<br />

the Center for NanoScience (LMU, München).<br />

Since June 2000 he has been a Full<br />

Professor at the Eindhoven University of<br />

Technology (Chair for Macromolecular<br />

Chemistry <strong>and</strong> Nanoscience).<br />

2482 2004 Wiley-VCH Verlag GmbH &Co. KGaA, Weinheim www.angew<strong>and</strong>te.org Angew. Chem. Int. Ed. 2004, 43, 2480 – 2495


<strong>Nanolithography</strong><br />

Angew<strong>and</strong>te<br />

Chemie<br />

automated scanning probe equipment, [41] the patterning of<br />

large areas has become accessible. Finally, not only patterns<br />

on substrates can be created, but also chemical modification<br />

reactions on specific substrates are accessible, which allow the<br />

combination of techniques from both the top-down <strong>and</strong> the<br />

bottom-up approaches. Bearing this in mind, lithographic<br />

techniques based on scanning probe techniques are the<br />

method of choice for applications in nanotechnology. [42]<br />

This Review provides an overview of current probe<br />

lithographic techniques <strong>and</strong> their subsequent (chemical)<br />

modification processes, which result in the creation of functional<br />

structures. The field of probe lithography can be split<br />

into two main areas: chemical <strong>and</strong> physical surface modification.<br />

The latter includes the formation of patterns by the<br />

physical movement of material on a substrate (for example,<br />

by applying a high contact force or through modifications<br />

induced by locally applied heat). Moreover, dip-pen lithography,<br />

in which atoms <strong>and</strong> molecules attached by adhesion to<br />

a tip can be transferred to a substrate in a controlled fashion,<br />

is also described in this context, since in this process chemical<br />

bonds are neither created or broken. The category of<br />

chemical modification processes include experiments that<br />

use local oxidation processes. This process was first observed<br />

in STM experiments, but has also been transferred to AFM<br />

<strong>and</strong> can be applied to conducting (metallic) substrates as well<br />

as to thin layers of nonconducting (organic) resists.<br />

For many of the classes in the physical modification<br />

category, a number of publications <strong>and</strong> some applications<br />

have been presented, <strong>and</strong> many good reviews are already<br />

available. For this reason, only a short description will be<br />

presented here. Moreover, no attempt will be made to present<br />

a comprehensive overview; rather a number of typical<br />

examples <strong>and</strong> applications of subsequent chemical modification<br />

will be shown. The second part provides examples of the<br />

electrical modification of substrates <strong>and</strong> organic resists using<br />

both STM <strong>and</strong> AFM techniques. Since STM techniques are<br />

rather well documented, this Review focuses on the application<br />

of conductive AFM tips in the formation of functional<br />

nanostructured surfaces.<br />

2. Physical <strong>Probe</strong> Lithography Techniques<br />

2.1. Mechanical Surface Patterning<br />

Since its invention, the goal of AFM has been to image<br />

surfaces without causing surface damage. However, the<br />

imaging of soft (organic) materials in contact mode (with<br />

relatively hard tips <strong>and</strong> stiff cantilevers) leads to surface<br />

damage. Therefore contact-mode AFM for these kind of<br />

samples is usually displaced by noncontact or intermittent<br />

contact AFM to reduce tip-induced surface damage. The<br />

concept of the controlled mechanical deformation of substrates<br />

using st<strong>and</strong>ard AFM tips can be transferred to almost<br />

any substrate. For example, the controlled application of<br />

(soft) tip-crashes in STM led to the formation of patterned<br />

substrates. The limiting factor in creating reproducible<br />

patterns is the stability of the tip itself, which is prone to<br />

deformation <strong>and</strong> contamination. To prevent excessive deformation,<br />

several groups have used diamond or diamondcoated<br />

tips. Mechanical patterning (force lithography) has<br />

successfully been applied to substrates <strong>and</strong> films of soft metals<br />

such as copper, gold, nickel, <strong>and</strong> silver. [43–47] The structures<br />

that have been created include thin lines <strong>and</strong> pits. Figure 3a<br />

Figure 3. a) AFM image (2.5 ” 2.5 mm 2 , z-range = 10 nm) of a resist<br />

(Shipley SP25) patterned with an intermittent-contact SPM. The patterning<br />

was performed in vector mode; the dark lines are plowed<br />

grooves while the bright lines are embankments due to the displaced<br />

resist; [46] b) AFM image (2.5 ” 2.6 mm 2 ) of a resonance-SPM modification<br />

of a polycarbonate film on silicon in which the patterning force<br />

has been applied according to greyscale bitmap image. [48]<br />

shows an example in which a force-induced pattern was<br />

inscribed into a photoresist film by high-force intermittentcontact<br />

scanning probe microscopy (SPM). High-force AFM<br />

has also been applied to polymer substrates <strong>and</strong> polycarbonate<br />

films (see Figure 3b), both in contact as well as<br />

intermittent operation. [46–48] Besides the mechanically induced<br />

deformation of substrates <strong>and</strong> films, scanning probe techniques<br />

can also be used for the gentle movement of metallic [49–51]<br />

or latex [52] nanoparticles into dimers, trimers, linear structures,<br />

<strong>and</strong> letters. The same method has also been applied for the<br />

local removal of parts of SAMs <strong>and</strong> Langmuir–Blodgett (LB)<br />

films. [53] If these experiments are carried out in a solution<br />

containing a competing molecule or nanoparticle the free<br />

space created by the tip will be filled in situ. Therefore, stable,<br />

ordered, mixed monolayers or a special arrangement of<br />

nanoparticles can be produced. Liu et al. presented an<br />

example of the mixed-monolayer method, where monolayers<br />

of 1-octadecanethiol <strong>and</strong> 11-sulfanyl-1-undecanol on gold<br />

were utilized as a substrate <strong>and</strong> ocatdecyltrichlorosilane<br />

(OTS) as a reagent in subsequent positive <strong>and</strong> negative<br />

pattern-transfer reactions (Figure 4). The same group also<br />

demonstrated the possibility of creating holes in SAMs of<br />

long-chain alkane thiols on gold surfaces, into which gold<br />

nanoparticles with a shell containing free thiols could be<br />

deposited. [54]<br />

Atomic manipulation is perhaps the most appealing form<br />

of modification from a scientific viewpoint. In 1990, Eigler<br />

<strong>and</strong> co-workers exhibited the possibility of creating artificial<br />

structures by the direct movement of adsorbed atoms on a<br />

metal surface in an ultrahigh vacuum (UHV) at low temperatures.<br />

[55] A notable example of a working logic device<br />

consisting of molecular cascades developed at IBM was also<br />

recently presented by Eigler. [56] By arranging adsorbed CO<br />

molecules on a Cu(111) substrate <strong>and</strong> using the instability of<br />

bent trimer adsorbates, working AND <strong>and</strong> OR logic circuits<br />

Angew. Chem. Int. Ed. 2004, 43, 2480 – 2495 www.angew<strong>and</strong>te.org 2004 Wiley-VCH Verlag GmbH &Co. KGaA, Weinheim<br />

2483


Reviews<br />

U. S. Schubert <strong>and</strong> D. Wouters<br />

2.2. Dip-Pen Lithography<br />

Figure 4. Schematic diagrams <strong>and</strong> corresponding AFM images<br />

(600 ” 600 nm 2 ) of a series of experiments for the construction of 3D<br />

nanostructures. Nanopatterns of 1-octadecanethiol (b) are grafted into<br />

a matrix of 11-sulfanyl-1-undecanol/Au(111) SAMs (a). The hydroxyl<br />

end groups within the nanostructure react with octadecyltrichlorosilane.<br />

The nanopatterning results in a negative pattern transfer (c). [53]<br />

The group of Mirkin has pioneered a patterning technique<br />

that relies on material being transferred from an AFM tip to<br />

the surface, known as dip-pen nanolithography (DPN). This<br />

material can simply be the constituents of the tip itself (i.e.,<br />

gold), induced by force or current, or it can be physisorbed<br />

material. Material on the tip is transferred to the substrate by<br />

capillary forces. Using DPN, Dravid et al. [57] have patterned<br />

silicon substrates with various dyes, whereas Rayment et al. [58]<br />

applied up to fifth-generation dendrimers (G5 DAB, amino<br />

terminated) to Si(100) substrates to form several micrometerlong<br />

lines with diameters up to 100 nm. The effective<br />

obtainable line widths were highly dependent on the utilized<br />

writing speed <strong>and</strong> temperature. [59] Although studies suggested<br />

that the transport of material was dependent on a water<br />

meniscus between the tip <strong>and</strong> the substrate, [60–66] writing<br />

experiments have also been performed at zero relative<br />

humidity, which indicates surface diffusion as being a transport<br />

factor. [59,67] Mirkin et al. have used DPN extensively to<br />

deposit proteins on surfaces. Recently the formation of a<br />

nanoarray of proteins on a gold substrate has been<br />

reported. [68] In this study, proteins (IgG <strong>and</strong> Lysozome)<br />

were positioned in nanoarrays (isl<strong>and</strong>s with diameters of<br />

200 nm) <strong>and</strong> subsequently coupled with antibodies (anti-<br />

IgG). Figure 6 schematically demonstrates the procedure.<br />

were combined into a two-input sorter composed of a setup of<br />

198 molecules on a 9 ” 9 nm area (Figure 5).<br />

Figure 6. Schematic representation of protein nanoassembly: In the<br />

first step, dip-pen lithography was used to create 16-sulfanyl-hexadecanoic<br />

acid (HMA) isl<strong>and</strong>s on a gold substrate (left). Rabbit-IgG proteins<br />

were adsorbed onto these isl<strong>and</strong>s (center), followed by coupling with<br />

anti-IgG (right). The process could be followed by observing increases<br />

in the height profiles of AFM images after each step. [68]<br />

The same group have utilized oligonucleotides, polystyrene<br />

latex particles (with diameters of 190 nm <strong>and</strong> 1 mm), [66,69]<br />

magnetic Fe 3 O 4 nanoparticles, [70] <strong>and</strong> metal-precursor-based<br />

inks in order to create patterns down to 45 nm. Nanostructures<br />

of Al 2 O 3 , SiO 2 , <strong>and</strong> SnO 2 have successfully been<br />

prepared from dip-pen depositing solutions of the corresponding<br />

MCl n -salts followed by hydrolysis to the corresponding<br />

oxide. [71] Dip-pen lithography also allows the<br />

creation of structures with more than one ink. By switching<br />

to another tip, it is possible to functionalize an area next to a<br />

previously prepared structure with another ink. [72] A review<br />

by Mirkin et al. supplies a more comprehensive overview on<br />

this topic. [73]<br />

Figure 5. STM image (12 ” 17 nm 2 ) of CO molecules adsorbed on a<br />

Cu(111) crystal at 5 K arranged into a three-input sorter. By moving a<br />

CO molecule with the STM tip (at one or more of the inputs) into an<br />

unstable trimer configuration, a molecular cascade is triggered (see<br />

inset) that travels down to the output where the STM was used to<br />

read out the result of the operation. [56]<br />

2.3. Thermal Patterning with Multi-Tip Systems<br />

The idea of using thermomechanic patterning for data<br />

storage was developed by Mamin <strong>and</strong> Rugar at IBM<br />

2484 2004 Wiley-VCH Verlag GmbH &Co. KGaA, Weinheim www.angew<strong>and</strong>te.org Angew. Chem. Int. Ed. 2004, 43, 2480 – 2495


<strong>Nanolithography</strong><br />

Angew<strong>and</strong>te<br />

Chemie<br />

Almaden. [74] Using a laser diode to heat a tip that is in contact<br />

with a poly(methyl methacrylate) (PMMA) film to a temperature<br />

higher than the softening point of PMMA creates an<br />

indentation in the film that has the shape of the tip. In this<br />

way, densities of up to 30 Gbin 2 were achieved. Although the<br />

size of the features was initially 100 nm, writing <strong>and</strong> readout<br />

was reliable (a system with a single tip traveling a total of<br />

16 km in 145 h reading 200 nm marks with a 400 nm period<br />

has been reported). [75,76] Practical data storage applications<br />

were envisioned with the development of sharper tips, higher<br />

readout frequency cantilevers, integrated heating, <strong>and</strong><br />

smoother substrates. Initial tips that had the disadvantage of<br />

laser-diode heating were replaced by tips with integrated<br />

heating circuits. [75, 77,78] However, mechanical limitations in the<br />

resonant frequencies of the cantilevers restricted the obtainable<br />

data transfer rates to a few Mb s 1 . [76, 79] Patterning speed<br />

as well as read-out speed could be significantly increased by<br />

the development of multi-tip systems. [80–88] To the best of our<br />

knowledge, the group of Quate (Stanford) first reported the<br />

use of a multi-tip system for lithography experiments on<br />

amorphous silicon in 1995. [87] The system had five parallel tips<br />

(Figure 7a) <strong>and</strong> was based on an earlier development by<br />

Tortonese. The group fabricated piezoresistive cantilevers<br />

that could operate without external sensing requirements due<br />

to integrated force sensing. Because the tip arrays did not<br />

allow for individual z-movement, the arrays had to be aligned<br />

Figure 7. a) SEM image of an assembly combining five AFM tips for<br />

parallel operation, as developed by Quate et al; b) an AFM image<br />

(400 ” 100 mm) of a pattern obtained by the parallel operation of four<br />

AFM tips; c) Pattern created by the parallel operation of two AFM tips<br />

(by local-probe oxidation of Si, see Section 3). The result demonstrates<br />

the limitation of individual z-control: the AFM image on the right<br />

(deflection controlled by a laser/detector system) shows the expected<br />

pattern, whereas the left tip (assumed to be in contact) performed<br />

poorly. [87]<br />

manually to the substrate. Utilizing this setup, 400 ” 100 mm<br />

parallel images were obtained in constant-height mode<br />

(Figure 7b). In a multiple-tip-scanner setup either an increase<br />

in scan size or in scan speed proportional to the number of tips<br />

could be realized. The lack of an individual height control<br />

limited the applicability of the system for lithographic<br />

purposes to a total of two tips operated in parallel. In<br />

Figure 7c a lithographic experiment using two paralleloperated<br />

tips demonstrates the problems associated with the<br />

lack of individual z-control: only one of the two tips was<br />

constantly in contact during patterning, while the other tip<br />

produced an incomplete pattern.<br />

In 1998, Vettiger <strong>and</strong> co-workers (IBM, Zürich) reported<br />

a 5 ” 5-tip array with integrated force sensing <strong>and</strong> its<br />

successful application for imaging. For stable operation, the<br />

tip array <strong>and</strong> the substrate had to be precisely levelled. This<br />

was realized by detecting the deflection of three cantilevers<br />

on three edges of the chip <strong>and</strong> by control through a setup of<br />

actuators on the array <strong>and</strong> the sample. [85] Moreover, Quate<br />

et al. reported in 1998 a centimeter-scale AFM [84] that used up<br />

to 50 parallel cantilevers with integrated force sensors. Surface<br />

areas of 2 ” 2 mm were imaged at a 0.4 mm pixel size in<br />

30 min with 10 parallel tips. An array of 32 tips was used to<br />

study 1.28 mm 2 of a diffraction grating <strong>and</strong> an array of<br />

50 parallel tips was utilized for lithography on hydrogenpassivated<br />

Si(100) by electric-field-enhanced oxidation at<br />

15 V. The resulting thin oxide lines were subsequently used as<br />

an etching mask resulting in a 1 cm 2 patterned area with lines<br />

that were 1.1 mm wide.<br />

By combining the original idea of Rugar <strong>and</strong> that of a<br />

multi-tip system, an AFM-based data storage system capable<br />

of producing high data density <strong>and</strong> improved read-out speeds<br />

became reality. In 2000, <strong>and</strong> again in 2003, Vettiger, Binnig<br />

<strong>and</strong> co-workers reported the development of the so-called<br />

“Millipede” system (Figure 8). [81–83] The “Millipede” consists<br />

of an array of 1024 tips (32 ” 32) with integrated read/write<br />

capabilities using combined effects of contact force <strong>and</strong><br />

applied heat (Figure 8a). The process of writing bits into thin<br />

PMMA films was achieved by heating the tips to 4008C using<br />

a current traveling through a highly doped section at the end<br />

of the cantilever. This allowed the prestressed lever structure<br />

to flex into the polymer (Figure 8b). Erasing large areas of<br />

written data required heating the whole medium to 1508C,<br />

which causes the PMMA film to undergo thermal reflow.<br />

Individual bits could be erased by forming a second pit<br />

adjacent to the first pit. Readout of the data was achieved by<br />

scanning a warm tip (3008C) over the sample <strong>and</strong> measuring<br />

the heat resistance over the bits: when a tip entered a hole, the<br />

cantilever came closer to the PMMA film, thus increasing the<br />

heat conductivity.<br />

By combining 1024 tips on a 3 ” 3 mm chip, patterns with a<br />

12 mm pitch <strong>and</strong> 40 nm data bits corresponding to aerial<br />

densities of 400 Gbin 2 , were obtained. The first prototypes<br />

of 4096-tip arrays (64 ” 64; 6.4 mm 2 ) were already reported [83]<br />

<strong>and</strong> design parameters are now focused on commercializing<br />

the concept. The “Millipede” system will probably be the first<br />

real commercial launch of an AFM-based “nanostorage<br />

device” as long as issues of cost reduction <strong>and</strong> long-term<br />

operation can be resolved. [83]<br />

Angew. Chem. Int. Ed. 2004, 43, 2480 – 2495 www.angew<strong>and</strong>te.org 2004 Wiley-VCH Verlag GmbH &Co. KGaA, Weinheim<br />

2485


Reviews<br />

U. S. Schubert <strong>and</strong> D. Wouters<br />

In addition to the aforementioned<br />

techniques, probe oxidation is a promising<br />

approach to create structures that can<br />

be utilized in subsequent modification<br />

steps. In this case, local surface oxidation<br />

introduces the chemical (<strong>and</strong> physical)<br />

functionality of the corresponding substrate.<br />

Figure 8. a) The “Millipede” concept developed by IBM consists of an array of 32 ” 32 tips. Each tip<br />

can be individually heated by applying an electric current through a highly doped area in the cantilever;<br />

b) when the heated tips were pressed against a thin PMMA film, small holes were created. These holes<br />

can be used as bits for data-storage applications. [82]<br />

Research into multi-tip systems has not been limited to<br />

storage applications. Progress in parallel dip-pen lithography<br />

with an eight-pen nanoplotter was reported by Mirkin <strong>and</strong><br />

Hong in 2000. [80] The operating principle is similar to that of<br />

the “Millipede”, but in this case an eight-tip array has been<br />

used to deposit octadecylthiol on gold in eight identical<br />

patterns. [80] In an inverted approach, Lee demonstrated the<br />

application of a tipless cantilever <strong>and</strong> an array of inverted tips<br />

on a substrate for application in force distance measurements.<br />

[89] Difficulties in conventional force measurements<br />

involving the placement of addressable active individual<br />

molecules on a single tip could be reduced by this approach.<br />

All of the probe-based lithographic techniques described<br />

above do not chemically modify the substrate. Functionality is<br />

only obtained through the mechanical movement of atoms<br />

<strong>and</strong> molecules. Therefore no chemical modifications are<br />

introduced that may be used in subsequent steps. Surfacemodification<br />

processes using tip-induced force are accompanied<br />

by a significant problem in the form of tip wear. This is in<br />

contrast to the heat-induced phase changes in thin polymer<br />

films (like in the “Millipede” project), where tip wear on the<br />

medium-to-long-term time scale (months) has been shown to<br />

be small. In contrast to mechanical <strong>and</strong>/or thermal patterning,<br />

dip-pen lithography does not modify the substrate but adds<br />

functionality by depositing molecules. These molecules introduce<br />

the possibility of secondary modifications using, for<br />

example, proteins. Although modification with dip-pen techniques<br />

is rather slow, both due to limited ink capacity <strong>and</strong><br />

relatively slow writing speed, progress can be envisaged<br />

through the development of multi-tip approaches. [80–88]<br />

3. Oxidative <strong>Probe</strong> Lithography<br />

Techniques<br />

3.1. Scanning Tunneling Microscopy<br />

Patterning<br />

Soon after the invention of scanning<br />

tunneling microscopy (STM) by G.<br />

Binnig <strong>and</strong> H. Rohrer in 1982, [31]<br />

researchers observed the ability of STM<br />

to modify surfaces, in addition to the<br />

effects of pure surface imaging. By 1985<br />

Güntherodt et al. had reported the possibility<br />

of creating structures in Pd 81 Si 19<br />

substrates (Figure 9). [39] Structure formation<br />

was attributed to the polymerization<br />

of an adsorbed oil layer originating from<br />

the vacuum oil-pumps.<br />

In electrochemical patterning experiments,<br />

STM can be used in two distinct emission modes. In the<br />

first mode, electrons from the tip (negative bias on the tip) are<br />

used to locally oxidize the material below the tip; a type of<br />

low-energy (tunneling) electron beam induces local oxidation<br />

of the material. In the second method (the field emitting<br />

mode), the area below the tip is oxidized by the electric field,<br />

while material between the tip (e.g., oxygen or water)<br />

Figure 9. SEM image of freshly drawn lines on glassy Pd 81 Si 19 . The<br />

lines were drawn by applying a bias voltage to the STM tip (see text).<br />

The effect was ascribed to the local electron-induced polymerization of<br />

oil fragments, which originated from the vacuum pump system. [39]<br />

2486 2004 Wiley-VCH Verlag GmbH &Co. KGaA, Weinheim www.angew<strong>and</strong>te.org Angew. Chem. Int. Ed. 2004, 43, 2480 – 2495


<strong>Nanolithography</strong><br />

Angew<strong>and</strong>te<br />

Chemie<br />

decomposes <strong>and</strong> their reaction products oxidize the substrate.<br />

A general reaction scheme proposed by Sugimura <strong>and</strong><br />

Nakagiri describes the anodic reactions [Eq. (1)] <strong>and</strong><br />

[Eq. (2)] that occur at the substrate surface <strong>and</strong> the cathodic<br />

reaction [Eq. (3)] that occurs at the tip (see also Figure 10): [101]<br />

M þ x H 2 O ! MO x þ 2 x H þ þ 2 x e<br />

2H 2 O ! O 2 þ 4H þ þ 4e<br />

2H 2 O þ 2e ! H 2 þ 2OH ð3Þ<br />

ð1Þ<br />

ð2Þ<br />

Snow reported the application of this technique in order to<br />

create one of the first working devices, initially using a silicon<br />

tip <strong>and</strong> cantilever in an experiment oxidizing thin (7 nm)<br />

titanium films. [106] Using a poorly conducting tip no current<br />

could be detected, which implied a field-effect induced<br />

oxidation of titanium to titanium oxide with adsorbed<br />

water. By partly oxidizing thin titanium wires produced by<br />

conventional optical techniques, thin titanium wires (10 nm)<br />

<strong>and</strong> metal-oxide junctions were prepared (Figure 11). The<br />

I–V characteristics of a 10 nm constricted titanium wire at<br />

room temperature <strong>and</strong> at 4.2 K were compared to those of a<br />

premodified (thick) wire.<br />

Figure 10. Schematic representation of the tip-induced oxidation of<br />

surfaces. The substrate was oxidized by an anodic reaction with the<br />

surface water layer. At the tip a cathodic reaction generates gaseous<br />

hydrogen.<br />

An example of the field emission method is the direct<br />

oxidation of a hydrogen-terminated Si(100) substrate. Currently,<br />

this represents the most frequently studied system,<br />

together with the direct modification of layered substrates<br />

(i.e., HOPG or MoS 2 ) or metal surfaces (i.e., Ti or Cr);<br />

oxidation experiments of organic resists on substrates have<br />

been reported less often.<br />

Heinzmann et al. have reported patterning on organic<br />

SAMs. [90,91] In this case, hexadecanethiol <strong>and</strong> N-biphenylthiol<br />

on Au(111) <strong>and</strong> octadecyltrichlorosilane on Si(100) were<br />

utilized. A pattern could be etched into the silicon substrate<br />

using two wet-etch steps (5% HF, 30% KOH). Earlier<br />

experiments by McCord <strong>and</strong> Pease also described the<br />

patterning of PMMA <strong>and</strong> alkyl halide resists. [92,93] They also<br />

prepared the first working device (a thin-film resistor) by<br />

patterning a PMMA resist, followed by a lift-off technique.<br />

[93, 94] A comprehensive review covering the early work<br />

in STM-induced surface modification was published in 1990<br />

by Shedd <strong>and</strong> Russell. [95]<br />

3.2. AFM <strong>Probe</strong> Lithography<br />

3.2.1. Oxidation of Metallic <strong>and</strong> Semiconducting Substrates<br />

In the early 1990s, successful lithography experiments<br />

with STM led to several groups transferring oxidation<br />

techniques to AFM. Field-effect induced oxidation has been<br />

applied to Si, [96–98] Ti, [87,99] Ta, [100] <strong>and</strong> Cr. [101] Electron-induced<br />

oxidation has been applied to organic resists, [102,103] SAMs, [104]<br />

<strong>and</strong> LB films. [105] As early as 1995, the group of Campbell <strong>and</strong><br />

Figure 11. a) I--V characteristics of a thick Ti wire <strong>and</strong> a Ti wire with a<br />

constriction of < 10 nm, measured at 300 <strong>and</strong> 4.2 K; b) AFM image of<br />

a 120 nm Ti wire with a narrow constriction. [106]<br />

Also in 1995, a metal-oxide semiconductor field-effect<br />

transistor (MOSFET) was produced using AFM lithography<br />

to create electrodes at the gate level, with the remaining<br />

structure prepared by mask lithography. [107] The process was<br />

confirmed with a-Si films on silicon dioxide, silicon nitride,<br />

sapphire, <strong>and</strong> chromium. Using a tip bias of 12 V (negative at<br />

the tip) <strong>and</strong> patterning speeds of 0.55 mms 1 , a MOSFET with<br />

an effective channel length of 0.1 mm was formed (with a<br />

patterning width of 0.21 mm). With transconductances of<br />

279 msmm 1 , the electron-transport properties of the device<br />

were comparable to those produced by electron-beam<br />

techniques.<br />

In 1999, Avouris et al. reported the preparation of a<br />

single-electron transistor from constricted thin metal wires<br />

(Ti <strong>and</strong> Nb) by means of the local probe oxidation of a thin<br />

metal film <strong>and</strong> a process entitled current induced local<br />

Angew. Chem. Int. Ed. 2004, 43, 2480 – 2495 www.angew<strong>and</strong>te.org 2004 Wiley-VCH Verlag GmbH &Co. KGaA, Weinheim<br />

2487


Reviews<br />

U. S. Schubert <strong>and</strong> D. Wouters<br />

oxidation (CILO). [108] First a thin metal wire was locally<br />

oxidized to constrict the wire to 100–200-nm-wide gaps. Then<br />

a transistor was produced by applying a 2 V bias over the<br />

constricted metal nanowire while measuring the current over<br />

the wire. Using CILO, a 10–50 nm metal-oxide barrier was<br />

formed across the constriction of wire in a self-limiting<br />

fashion without the need of external control. The experiments<br />

suggested that the barrier formation involved current-induced<br />

atomic rearrangements <strong>and</strong> local heating. Using both the<br />

probe oxidation <strong>and</strong> the CILO process, T-shaped junctions of<br />

two tunneling barriers with differing widths <strong>and</strong> a metal isl<strong>and</strong><br />

were produced (Figure 12).<br />

a width of 20 nm at the base <strong>and</strong> 7 nm at the top were<br />

produced. Packed patterns of lines with periodicities of 13 nm<br />

were also prepared (Figure 13).<br />

Figure 13. a <strong>and</strong> b) Local-probe oxidation of Si to SiO 2 by applying a<br />

sequence of 21 V pulses for 1 ms to create a set of 23 interdigitated<br />

lines placed 43 nm apart; c) a set of 19 lines placed 13 nm apart was<br />

created by applying a potential of 24.7 V for 80 ms. [110]<br />

Figure 12. a <strong>and</strong> b) 1.5 mm thick Ti wire with a narrow gap at the<br />

center that was produced by local probe oxidation; c) I–V curves<br />

across the device before <strong>and</strong> after the creation of the barrier;<br />

d) current-induced local oxidation (CILO) created a nanometer-scale<br />

barrier. [108]<br />

In 1995, Quate <strong>and</strong> co-workers demonstrated probe<br />

lithography with 40 nm resolution at very high tip speed. [109]<br />

Speeds of 1 mms 1 were obtained on Si substrates coated with<br />

thin films of a siloxene, known as “spin-on glass” (SOG).<br />

SOG is used as a positive resist, <strong>and</strong> the etch speed of the<br />

underlying silicon depends on the amount of organic content<br />

in the SOG, which is reduced by the field-induced oxidation.<br />

The direct oxidation of hydrogen-passivated Si substrates<br />

can also be used in the formation of nanostructured functional<br />

patterns. In 1995, Sugimura <strong>and</strong> Nakagiri reported a method<br />

in which the created SiO 2 patterns were used as negative<br />

resists in an alkaline etching step that resulted in 30 nm<br />

steps. [101] The substrate with the same SiO 2 patterns could also<br />

be covered with a gold film. By using electroless plating, gold<br />

was deposited on the Si–H surface <strong>and</strong> not on the insulating<br />

silicon oxide patterns. Recently, García et al. determined the<br />

linewidth <strong>and</strong> shape of SiO 2 on silicon. [110] For this purpose,<br />

several high-resolution patterns were produced on silicon<br />

wafers with a 2 nm coating of a native oxide by applying 24 V<br />

pulses through an n-doped tip; trapezoidial-shaped lines with<br />

Lyuksyutov <strong>and</strong> co-workers have investigated tip–surface<br />

interactions during the oxidation of an Si–H surface. They<br />

reported an approach for monitoring tip–surface interactions<br />

based on power spectral analysis of tip oscillations during<br />

scanning probe oxidation. [111] Utilizing a single-mode harmonic<br />

oscillator model, the amplitude, frequency, <strong>and</strong> damping<br />

factor of oscillation were monitored as a function of bias<br />

voltage. The results obtained were consistent with the concept<br />

that a water meniscus <strong>and</strong> electrostatic tip–surface interactions<br />

dominate contact AFM lithography, that is, a high bias<br />

voltage destabilized the water meniscus <strong>and</strong> the oscillation<br />

amplitude was reduced.<br />

3.2.2. Oxidation of Substrates with Organic Resist Layers<br />

The previous examples have described the direct oxidation<br />

of Si to SiO 2 . This process can also be achieved with Si<br />

samples that are covered with an organic resist. Experiments<br />

utilizing resists consisting of (mixed) LB films [112] <strong>and</strong> (mixed)<br />

SAMs [113–117] have been reported. In the first case, the<br />

application of mixed LB films of hexadecylamine (HAD)<br />

<strong>and</strong> palmitic acid (PA) was reported by Lee et al. Different<br />

ratios of the LB components were used <strong>and</strong> it was concluded<br />

that the mixed layers led to significantly thinner pattern<br />

widths. Although lines that were approximately three times<br />

thinner than those prepared by alternative techniques were<br />

observed, there were no comments concerning the shape <strong>and</strong><br />

condition of the tip that was employed. Liu et al. reported a<br />

method to create assemblies of gold nanoparticles on<br />

patterned SAMs of OTS on SiO 2 /Si (Figure 14). [113] By locally<br />

oxidizing the silicon under the organic resist, they created<br />

2488 2004 Wiley-VCH Verlag GmbH &Co. KGaA, Weinheim www.angew<strong>and</strong>te.org Angew. Chem. Int. Ed. 2004, 43, 2480 – 2495


<strong>Nanolithography</strong><br />

Angew<strong>and</strong>te<br />

Chemie<br />

Figure 14. Functionalization method introduced by Liu et al.: In the<br />

first step a Si wafer with an OTS resist layer was locally oxidized to<br />

create protruding SiO 2 dots (top right). A second silane (APTMS) was<br />

then coupled to the dots to introduce amine functionalities exclusively<br />

to the oxidized areas. The amine-functionalized areas could be used to<br />

bind negatively charged gold nanoparticles (bottom right). If the dots<br />

were very close to each other, the amine pattern was not fully covered<br />

by gold particles, which was ascribed to mutual repulsion between the<br />

gold particles themselves. [113]<br />

arrays or SiO 2 dots with diameters of 15 nm <strong>and</strong> a 40 nm<br />

spacing. The protruding SiO 2 was used in a subsequent<br />

chemical modification step in order to introduce new<br />

functionalities. In this step a second silane (aminopropyltrimethoxysilane<br />

(APTMS)) reacted with the SiO 2 created by<br />

the probe oxidation, whereas the unoxidized OTS resist layer<br />

protected the SiO 2 below. The created SiO 2 pattern was<br />

transferred <strong>and</strong> terminal amino groups were introduced to<br />

adsorb gold nanoparticles. The successful preparation of a<br />

site-selective assembly of 15 nm gold nanoparticles on 100 nm<br />

AFM-defined dots with 300 nm spacing was reported. When<br />

an array of 30 nm dots with 70 nm spacing was utilized as a<br />

template, the site-selective adsorption of gold nanoparticles<br />

was incomplete: large portions of the template remained<br />

unoccupied. This was attributed to electrostatic repulsion<br />

between individual gold nanoparticles. Takai <strong>and</strong> co-workers<br />

also utilized SAMs of OTS on silicon wafers to create SiO 2<br />

nanostructures by local tip-induced oxidation processes. [114,115]<br />

In addition, they also reported a functionalization method by<br />

introducing a fluoroalkyl silane (FAS, heptadecafluoro-<br />

1,1,2,2-tetrahydro-decyl-1-trimethoxysilane). [115] Due to the<br />

large molecular dipole that is introduced, Kelvin-probe force<br />

microscopy could be used to study the created nanopatterns.<br />

In an earlier publication, Takai <strong>and</strong> Nakagiri reported the<br />

successful application of the same technique to create nanostructured<br />

gold patterns on surface. [116] First a silicon wafer<br />

with an alkyl silane was locally oxidized to create SiO 2<br />

nanopatterns. The SiO 2 was then removed in a 0.5% HF<br />

etching step. Finally, gold was introduced to the patterns in an<br />

electroless plating bath. Successful deposition of gold onto the<br />

structures with 100 nm spatial resolution was shown by SEM<br />

imaging.<br />

In 2002, Takai <strong>and</strong> co-workers explored parameter space<br />

in the oxidation of Si. [114] In this study, relationships between<br />

the lateral <strong>and</strong> vertical dimensions of oxide dots, formed on<br />

octadecyltrimethoxysilane(ODS)-passivated Si, <strong>and</strong> the<br />

applied bias voltage as well as bias duration time were<br />

reported. Based on the results, a stepwise oxidation reaction<br />

was proposed: In the first step the organic resist was locally<br />

oxidized <strong>and</strong>/or destroyed. As the SAM layer degrades,<br />

anodization of the underlying silicon becomes apparent, as<br />

indicated by an observed increase in the lateral force contrast.<br />

For the preparation of line structures, a minimum line width<br />

of around 20 nm has been found, which is independent of<br />

oxidation voltage <strong>and</strong> probe oxidation speed.<br />

The oxidative removal of SAMs has also been studied for<br />

octadecanethiol on gold by Uosaki et al., who reported that<br />

thiols could be removed from the gold substrate by applying a<br />

2–3 V positive or negative bias voltage, with both the tip <strong>and</strong><br />

the substrate placed in toluene containing varying concentrations<br />

of water. [118] The created structures did not consist of<br />

bare gold, because only a 1 nm height change could be<br />

observed. The remaining nanometer of material was ascribed<br />

to adsorbed reaction products of the thiol–gold cleavage,<br />

which is consistent with the measured resistivity (1.7 ” 10 9 W)<br />

for a 1 nm resist on gold.<br />

In addition to the oxidation of Si to SiO 2 <strong>and</strong> the<br />

subsequent functionalization reactions, the oxidation of<br />

molybdenum to molybdenum oxide by local probe-oxidation<br />

techniques was also reported by Dai <strong>and</strong> co-workers. This<br />

approach represents another successful route for the preparation<br />

of devices that then allows subsequent modification<br />

steps. [119] The principle of functionalization in this approach is<br />

based on the different solubility of Mo <strong>and</strong> MoO 3 . A thin<br />

layer of Mo on a Si substrate was locally oxidized. The<br />

produced MoO 3 could easily be removed by dissolving it in<br />

water <strong>and</strong> therefore exposing the bare Si substrate underneath.<br />

The exposed silicon was wet-etched by a 30 % aqueous<br />

solution of KOH at 708C for 30 s, which yielded V-shaped<br />

grooves in the underlying silicon.<br />

Titanium wires were prepared by using a more elaborate<br />

functionalization technique. [119] A thin film of Mo was applied<br />

on top of a PMMA film on Si. After the MoO 3 was removed,<br />

the PMMA film was locally removed by oxygen plasma<br />

etching, which opened a pathway to the silicon wafer for Ti<br />

deposition. The remaining Mo film protected the unexposed<br />

areas. By this technique, Ti wires (35 nm wide) were prepared<br />

(see Figure 15). The same group also suggested the application<br />

of chromium <strong>and</strong> germanium films based on the solubility<br />

of the corresponding metal oxides.<br />

3.2.3. Oxidation of Organic Resist Layers: Introducing Chemical<br />

Functionality<br />

All of the functionalization processes mentioned so far<br />

have been based on the direct oxidation of substrates or the<br />

destruction of a thin passive resist layer on an oxidizable<br />

substrate. Although complex structures have mainly been<br />

prepared using wet etching steps, <strong>and</strong> in some cases by silane<br />

Angew. Chem. Int. Ed. 2004, 43, 2480 – 2495 www.angew<strong>and</strong>te.org 2004 Wiley-VCH Verlag GmbH &Co. KGaA, Weinheim<br />

2489


Reviews<br />

U. S. Schubert <strong>and</strong> D. Wouters<br />

Figure 15. A Mo film (4 nm) was evaporated on top of a PMMA film<br />

that was spin-coated onto a p-doped Si substrate. a) The Mo layer was<br />

locally oxidized (7–11 V, 0.4–1 mms 1 ) by local-probe oxidation to<br />

create a slightly protruding soluble MoO 3 pattern with 35 nm wide<br />

lines; b) in the second step, the MoO 3 was dissolved in water to create<br />

25 nm wide gaps that exposed the underlying PMMA film; c) the<br />

exposed PMMA film was then etched by exposure to oxygen, to create<br />

deep channels; d) in the last step, Ti was evaporated onto the whole<br />

substrate. Removal of the PMMA film by a lift-off procedure (using<br />

acetone) yielded Ti nanowires on the remaining Si support. [119]<br />

chemistry, the range of applicable functionalization techniques<br />

is limited. A notable exception to destructive functionalization<br />

was reported by Sagiv et al., who found that the tipinduced<br />

oxidation of organic monolayers introduced surface<br />

functionality, which could then be used in a large number of<br />

subsequent functionalization reactions with both organic <strong>and</strong><br />

inorganic substances yielding a range of nanopatterned<br />

functional structures. [120–124] The work involved is summarized<br />

in Scheme 1. The procedure was first demonstrated on<br />

monolayers of 18-nonadecenyltrichlorosilane (NTS) on a p-<br />

doped silicon wafer (see Scheme 1, path 1). [121] The terminal<br />

vinylic end group could be locally oxidized by a conductive<br />

AFM tip with an applied bias voltage of approximately 9 V<br />

(tip negative). Oxidation was carried out at 5 mms 1 with a<br />

variety of conductive tips (boron doped, diamond coated,<br />

tungsten, tungsten carbide, copper- <strong>and</strong> silver coated silicon<br />

nitride, <strong>and</strong> highly doped silicon). Structures were obtained<br />

by repeatedly (up to 40 times) scanning the tip along a<br />

predefined vector during which no current (above 5 nA)<br />

could be detected. The successful selective oxidation of the<br />

terminal vinylic end groups to terminal carboxylic acid groups<br />

was confirmed by subsequently imaging the area in contact<br />

mode with the same probe (without applying a voltage),<br />

simultaneously recording height <strong>and</strong> lateral friction signals.<br />

For the unaffected areas the surface remained unchanged,<br />

whereas for the oxidized areas the height remained the same<br />

but an increase in friction was observed, which was attributed<br />

to changes in local surface polarity. Images of this patterning<br />

procedure are displayed in Figures 16 <strong>and</strong> 17 in which this<br />

procedure is used for subsequent modification reactions.<br />

Oxidation experiments were verified by partial wet-chemical<br />

oxidation: the substrate was partially exposed to a 5 mm<br />

solution of KMnO 4 /dicyclohexano-[18]crown-6 <strong>and</strong> also to<br />

electric oxidation of large surface areas by means of a copper<br />

grid (connected to a 13 V negative bias voltage). Brewster<br />

angle infrared spectroscopy indicated the presence of carboxylic<br />

acid groups.<br />

To prevent oxidation of the underlying silicon it is<br />

essential that the surface is covered by a densely packed<br />

closed monolayer. By controlling the pathway of the tip over<br />

the surface, complicated structures with high resolution (9 nm<br />

linewidth) could be created. For this procedure both vectorbased<br />

trajectories as well as bitmap (pixel per pixel) oxidation<br />

programs were reported. In their first publication, Sagiv et al.<br />

reported the possibility of functionalizing structures using<br />

chemical reactions that react exclusively on the oxidized<br />

patterns. [121] The concept was demonstrated by the formation<br />

of a second stable monolayer of OTS on top of the oxidized<br />

sections of the 18-nonadecenyl layer. The coupling of the<br />

second monolayer was followed by contact-mode AFM<br />

imaging. Through the coupling of OTS to the oxidized<br />

areas, the terminal carboxylic end groups were converted to<br />

terminal methyl end groups. This was observed in contactmode<br />

imaging by the disappearance of the friction signal<br />

along with a simultaneous increase in the height signal<br />

(Figure 16). [121,123]<br />

Later publications showed that not only NTS SAMs could<br />

be used but also inert SAMs of OTS could locally be oxidized<br />

to carboxylic acid end groups (Scheme 1, path 2), thus<br />

opening a route for chemical modification. [123] Noteworthy<br />

is the approach in which a monolayer of NTS was attached to<br />

locally oxidized OTS (Scheme 1). Thus functionalized patterns<br />

have been used as templates for the formation of<br />

metallic layers by using wet-oxidation <strong>and</strong> subsequent<br />

adsorption <strong>and</strong> reduction of metal ions (Scheme 1, path 2a).<br />

The terminal vinyl groups could also be converted to thiol<br />

groups by UV irradiation (Hg lamp, 254 nm) for 10 min in an<br />

H 2 S/Ar (1:1) atmosphere. As a result a stable pattern<br />

decorated with thiol end groups placed in well-defined<br />

positions was created, <strong>and</strong> could be used in subsequent<br />

functionalization steps. Stable layers of metals could be<br />

formed on the templates by reduction of suitable physisorbed<br />

metal ions (either tip-induced or by wet chemistry). Application<br />

of gold nanoparticles (Au 55 ) to the thiol groups resulted<br />

in the formation of gold isl<strong>and</strong>s connected by lines that were<br />

only two nanoparticles wide (Figure 17). [120] The method<br />

presented by Sagiv et al. for surface patterning <strong>and</strong> modification<br />

has many possibilities for the production of nanometersized<br />

devices as a result of the large number of available<br />

chemical <strong>and</strong> physical modification steps, for example using<br />

silane chemistry (silanes with terminal functionality), the<br />

coupling of amines <strong>and</strong> cationic species, thiol chemistry, <strong>and</strong><br />

conversion to anhydrides. [125]<br />

We have performed surface modifications on OTSfunctionalized<br />

Si wafers using a method similar to that<br />

2490 2004 Wiley-VCH Verlag GmbH &Co. KGaA, Weinheim www.angew<strong>and</strong>te.org Angew. Chem. Int. Ed. 2004, 43, 2480 – 2495


<strong>Nanolithography</strong><br />

Angew<strong>and</strong>te<br />

Chemie<br />

Scheme 1. Functionalization method developed by Sagiv et al. 1) A monolayer of 18-nonadecenyltrichlorosilane (NTS) on a Si wafer was locally<br />

oxidized. Chemical functionalities were introduced that were used as templates for subsequent modification steps, such as the attachment of a<br />

second (different) chlorosilane; 2) the oxidation procedure was also applied to octadecyltrichlorosilane (OTS) monolayers, introducing similar<br />

functionalities. Wafers were prepared with terminal methyl groups (from OTS) patterned with local terminal vinylic functionality end groups. NTS<br />

was coupled to local-probed oxidation-induced patterns. The local presence of vinyl groups was utilized in two different pathways: 2a) through<br />

chemical oxidiation to carboxylic acid functionalities, <strong>and</strong> 2b) in which the vinyl group was converted to a thiol group. Both functional patterns<br />

were used in various subsequent steps including the coupling of gold particles <strong>and</strong> the formation of metallic silver <strong>and</strong> cadmium sulfide isl<strong>and</strong>s.<br />

Figure 16. AFM images showing the formation of a NTS-functionalized<br />

template structure (see also Scheme 1a). a) After tip-induced inscription<br />

of a pattern of electrooxidized OTS (OTSeo); b) Images of an<br />

NTS/OTS bilayer formed by exposure of the patterned OTS monolayer<br />

to a 5 mm solution of NTS in biscyclohexane (BCH). [123]<br />

developed by Sagiv et al. By utilizing a number of conductive<br />

AFM tips with various coatings (Au, TiN, Pt, <strong>and</strong> W 2 C) <strong>and</strong> a<br />

broad range of stiffnesses (0.15–5 N m 1 ), an OTS monolayer<br />

was locally oxidized to create templates down to 20 nm<br />

resolution, which were then used in a number of different<br />

subsequent functionalization steps. [126,127] Figure 18 shows the<br />

Figure 17. Topography (left) <strong>and</strong> phase-contrast (right) AFM images<br />

showing portions of gold-nanoparticle-coated lines produced by localprobe<br />

oxidation (see also Scheme 1, path 2b). Using this method,<br />

both wires <strong>and</strong> contact pads were produced. [120]<br />

Angew. Chem. Int. Ed. 2004, 43, 2480 – 2495 www.angew<strong>and</strong>te.org 2004 Wiley-VCH Verlag GmbH &Co. KGaA, Weinheim<br />

2491


Reviews<br />

U. S. Schubert <strong>and</strong> D. Wouters<br />

results of the specific adsorption of trimethyloctadecylammonium<br />

bromide (TOA). In the first step, the OTS monolayer<br />

was locally oxidized to create carboxylic acid groups. The<br />

local height remained almost constant after oxidation<br />

(Dz max = 0.5 nm) <strong>and</strong> when measured in contact mode a<br />

clear friction signal originating from the different surface<br />

properties of the unmodified OTS <strong>and</strong> the carboxylic acid<br />

groups was observed (Figure 18a <strong>and</strong> b). Upon adsorption of<br />

Figure 19. Tapping-mode height image (1.1 ” 1.1 mm, z-range 18 nm)<br />

of cationic gold nanoparticles adsorbed onto an oxidized structure.<br />

The observed height of the particles ranges from 17 to 20 nm.<br />

Successful experiments have also been performed in which<br />

functional silanes were coupled to the oxidized patterns. Two<br />

silanes, one with a terminal vinyl functionality <strong>and</strong> one with a<br />

terminal methylmethacrylate group (both targets for free<br />

radical polymerization reactions) were coupled to the template.<br />

In both cases the polymer was grafted to the surface<br />

species; the method appears to be promising due to the wide<br />

availability of functional silanes.<br />

3.3. Surface Modification by Switching Tip Properties<br />

Figure 18. a) Contact-mode height image (8 ” 14 mm, z-range 0.5 nm)<br />

of an oxidized triangle; b) friction image of the same triangular shape,<br />

indicating that successful oxidation had occurred; c) height image<br />

(8 ” 14 mm, z-range 0.5 nm) after exposing the surface to a solution of<br />

trimethyloctadecylammonium bromide (TOA) in water, which causes<br />

the height of the triangular area to increase by 2.0 nm <strong>and</strong> indicates<br />

the adsorption of one layer of TOA; d) the corresponding friction<br />

image; e) contact-mode height image (30 ” 30 mm) of a structure created<br />

by a sequential (six-step) method using both oxidation <strong>and</strong> the<br />

adhesion of TOA.<br />

the TOA from an aqueous solution, a height step (D = 1.5 nm)<br />

was observed, which corresponds to the length of the<br />

octadecyl chain; simultaneously the friction signal disappeared<br />

due to the reformation of an apolar top surface<br />

(Figure 18 c <strong>and</strong> d). The created structures were stable when<br />

washed with water. The oxidation/adsorption procedure was<br />

repeated three times, each time introducing a structure next<br />

to the previous one, but never destroying the original pattern<br />

(Figure 18 e).<br />

Besides the adsorption of quaternary ammonium salts, the<br />

procedure has also been demonstrated for positively charged<br />

gold nanoparticles (d = 15 nm, coated with poly-l-lysine;<br />

Figure 19). The protein C<strong>and</strong>ida Antarctica Lipase B was<br />

selected for protein-coupling experiments as it contains thio<br />

functionalites on its outer surface. Initial results showed the<br />

successful coupling of proteins to a gold template. [128] Possible<br />

applications may be found in nanosensor array devices.<br />

Ober et al. [129] introduced a technique called redox probe<br />

microscopy (RPM), in which an AFM tip was modified with<br />

redox-active materials. When a potential was applied, the tipsurface<br />

interaction could be switched.<br />

Two methods of lithographic operation were presented:<br />

The first is based on controlling the adhesion between the tip<br />

<strong>and</strong> the particles on the surface, effectively picking them up<br />

<strong>and</strong> positioning them in a pattern (see Figure 20). For this<br />

procedure the tip was coated with poly(vinylferrocene)<br />

(PVF), which is neutral 0 V, but is oxidized to the ferricenium<br />

state when the bias is increased (+ 1.0 V), thus becoming<br />

highly positively charged. The positively charged tip strongly<br />

interacted with negatively charged beads (15 mm, sulfonic<br />

acid terminated chromatography beads), which could be<br />

Figure 20. A poly(vinylferrocene)-coated tip in the reduced state is neutral.<br />

However, when oxidized (tip bias + 1.0 V) the positively charged<br />

tip could be used to pick up negatively charged particles. [129]<br />

2492 2004 Wiley-VCH Verlag GmbH &Co. KGaA, Weinheim www.angew<strong>and</strong>te.org Angew. Chem. Int. Ed. 2004, 43, 2480 – 2495


<strong>Nanolithography</strong><br />

Angew<strong>and</strong>te<br />

Chemie<br />

picked up, <strong>and</strong> released again by reducing the PVF film. The<br />

described experiments were repeated with switchable tips<br />

consisting of a gold-coated tip modified with a 6-(ferrocenylcarbonyl)hexanethiol.<br />

In the second method, surface modification was performed<br />

by an AFM tip that caused a change in pH value when<br />

an electric bias was applied (Figure 21). In this approach, a<br />

gold-coated tip was modified with thiomethyl-2,5-hydroquinone<br />

(thioQH 2 ). At positive biases the thioQH 2 is oxidized to<br />

the quinone <strong>and</strong> two protons per molecule are generated. As<br />

a result, a film of a pH-sensitive triblock copolymer (MMA-<br />

TBMA-MAA) has been patterned. [129]<br />

Figure 21. a) The pH lithographic method: The oxidation of thiomethyl-2,5-hydroquine<br />

releases protons that can be used to locally pattern<br />

a pH-sensitive block copolymer film of methyl methacrylate<br />

(MMA), tert-butyl methyl methacrylate (TBMA) <strong>and</strong> methyl acrylic acid<br />

(MAA); b) AFM images of the block copolymer MMA-TBMA-MAA (A,<br />

left) showing the effects of tip-induced surface damage due to the<br />

repeated scanning (30 min) of a 500 ” 500 nm area. Image B (right)<br />

shows a similar approach where repeated scanning over an area was<br />

performed with the AFM tip cycled between oxidative <strong>and</strong> reductive<br />

modes. Note the enhancement in the generated pattern. [129]<br />

4. Conclusions <strong>and</strong> Outlook<br />

UV<strong>and</strong> UV-mask lithographic processes are currently the<br />

most widely employed <strong>and</strong> studied techniques for the<br />

production of sub-micrometer-sized devices. However, no<br />

dramatic improvements in the accessible sizes should be<br />

expected in the future. Although good progress has been<br />

made in field of electron-beam lithography, the conditions<br />

remain very dem<strong>and</strong>ing, which leads to excessive costs.<br />

With this in mind, scanning-probe-based lithographic<br />

techniques seem more feasible. Within this field, research<br />

efforts into dip-pen lithographic techniques have led to welldefined<br />

systems capable of depositing (biological) materials<br />

onto flat substrates at a relatively high speed. With the<br />

development of multi-ink systems, applications into biorelated<br />

sensors are envisioned in much the same way as<br />

devices already available that are based on microcontact<br />

printing. However, dip-pen lithography patterning is limited<br />

due to the low availability of suitable ink combinations, the<br />

limited stability of the patterning due to diffusion, <strong>and</strong> the low<br />

patterning speed. The patterning speed can be increased<br />

dramatically by the introduction of multiple-tip systems, such<br />

as those developed by IBM for AFM-based storage systems.<br />

The simultaneous operation of 1024 tips allows reasonable<br />

reading <strong>and</strong> writing speeds to be obtained. In combination<br />

with automated AFM systems <strong>and</strong> sample stages, the<br />

patterning of large substrates then becomes possible.<br />

<strong>Probe</strong>-induced oxidative techniques may also become<br />

more viable. Such techniques have the advantage of being<br />

relatively easy, with respect to the required conditions <strong>and</strong><br />

equipment. Even more importantly, the method is highly<br />

versatile both in applicable substrates as well as in the<br />

availability of functionalization routes. Local oxidation can be<br />

performed on almost any conducting sample, even on thin<br />

organic resists, at high resolution. The technique can be<br />

employed to locally oxidize the sample <strong>and</strong> resist layers, thus<br />

inducing changes in physical properties. The obtained systems<br />

can be employed in both positive as well as negative<br />

development of the pattern by various wet-etching techniques.<br />

By this technique, a number of small metallic nanowires<br />

<strong>and</strong> nanometer-sized electronic systems have been prepared.<br />

Besides development by etching procedures, local oxidation<br />

can also be used for chemical modification steps on organic<br />

resists. Chemical functionalities such as thiols, amines, or<br />

carboxylic acids can be introduced into defined patterns.<br />

Thus, a large number of modification routes are possible,<br />

which open up new avenues for the preparation of nanometer-sized<br />

functional architectures on surfaces. Applications<br />

can be envisaged in various areas, ranging from the defined<br />

introduction of biological material (for example, DNA or<br />

proteins) for sensor applications, to the coupling of polymerization<br />

initiators or catalyst species for mechanistic studies<br />

<strong>and</strong> the preparation of special smart coatings or intelligent<br />

polymer brushes.<br />

The Dutch Polymer Institute (DPI) <strong>and</strong> the Fonds der<br />

Chemischen Industrie are acknowledged for their financial<br />

support.<br />

Received: May 26, 2003 [A609]<br />

[1] C. Macilwain, Nature 1999, 400, 95 – 95.<br />

[2] European Commission IST Programme, R. Compaæó, 2000,<br />

www.cordis.lu/ist/fetnidqf.htm.<br />

[3] G. Y. Tseng, J. C. Ellenbogen, Science 2001, 294, 1293 – 1294.<br />

[4] T. Appenzeller, Science 1991, 254, 1300 – 1311.<br />

[5] K. J. Klabunde, Nanoscale Materials in Chemistry, Wiley-<br />

Interscience, New York, 2001, pp. 1 – 14.<br />

[6] C. Hui, Nanotechnology 1999, 10, 113 – 116.<br />

[7] G. M. Whitesides, J. P. Mathias, C. T. Seto, Science 1991, 254,<br />

1312 – 1326.<br />

[8] G. A. Ozin, Adv. Mater. 1992, 4, 612 – 648.<br />

Angew. Chem. Int. Ed. 2004, 43, 2480 – 2495 www.angew<strong>and</strong>te.org 2004 Wiley-VCH Verlag GmbH &Co. KGaA, Weinheim<br />

2493


Reviews<br />

U. S. Schubert <strong>and</strong> D. Wouters<br />

[9] D. Philp, J. F. Stoddart, Angew. Chem. 1996, 108, 1242 – 1286;<br />

Angew. Chem. Int. Ed. Engl. 1996, 35, 1155 – 1196.<br />

[10] R. F. Service, Science 2001, 293, 782 – 785.<br />

[11] U. S. Schubert, C. Eschbaumer, Angew. Chem. 2002, 114, 3016 –<br />

3050; Angew. Chem. Int. Ed. 2002, 41, 2892 – 2926.<br />

[12] Einführung in die Lithographie, Hanns Eggen KG, Hannover,<br />

1968.<br />

[13] J. Kohler, Everything you wanted to know about Printing,<br />

International Paper, http://www.ippaper.com/gettips_pp_l_history.html.<br />

[14] W. Chappell, A Short History of the Printed Word, Dorset Press,<br />

New York, 1970.<br />

[15] L. Febvre, H.-J. Martin, The Coming of the Book, Verso,<br />

London, 1997.<br />

[16] S. M. Miller, S. M. Troian, S. Wagner, J. Vac. Sci. Technol. B<br />

2002, 20, 2320 – 2327.<br />

[17] Y. Xia, G. M. Whitesides, Annu. Rev. Mater. Sci. 1998, 28, 153 –<br />

184.<br />

[18] H. A. Biebuyck, N. B. Larsen, E. Delamarche, B. Michel, IBM<br />

J. Res. Dev. 1997, 41, 159 – 170.<br />

[19] R. F. Service, Science 2001, 293, 785 – 786.<br />

[20] K. A. Bates, M. Rothschild, T. M. Bloomstein, T. H. Fedynyshyn,<br />

R. R. Kunz, V. Liberman, M. Switkes, IBM J. Res. Dev.<br />

2001, 45, 605 – 614.<br />

[21] C. Brodsky, J. Byers, W. Conley, R. Hung, S. Yamada, K.<br />

Patterson, M. Somervell, B. Trinque, H. V. Tran, S. Cho, T.<br />

Chiba, S.-H. Lin, A. Jamieson, H. Johnson, T. V<strong>and</strong>er Heyden,<br />

C. Grant Willson, J. Vac. Sci. Technol. B 2000, 18, 3396 – 3401.<br />

[22] B. C. Trinque, T. Chiba, R. J. Hung, C. R. Chambers, M. J.<br />

Pinnow, B. P. Osburn, H .V. Tran, J. Wunderlich, Y.-T. Hsieh,<br />

B. H. Thomas, G. Shafer, D. D. DesMarteau, W. Conley, C.<br />

Grant Willson, J. Vac. Sci. Technol. B 2002, 20, 531 – 536.<br />

[23] L. Karapiperis, C. A. Lee, Appl. Phys. Lett. 1979, 35, 395 – 397.<br />

[24] J. Melngailis, A. A. Mondelli, I. L. Berry, R. Mohondro, J. Vac.<br />

Sci. Technol. B 1998, 16, 927 – 957.<br />

[25] F. J. Hohn, H<strong>and</strong>book of Surface Imaging <strong>and</strong> Visualization,<br />

CRC press, London, 1995, pp. 115 – 129.<br />

[26] T. R. Groves, D. Pickard, B. Rafferty, N. Crosl<strong>and</strong>, D. Adam, G.<br />

Schubert, Microelectron. Eng. 2002, 61–62, 285 – 293.<br />

[27] J. P. Spatz, Angew. Chem. 2002, 114, 3507 – 3510; Angew. Chem.<br />

Int. Ed. 2002, 41, 3359 – 3362.<br />

[28] A. Gölzhäuser, W. Eck, W. Geyer, V. Stadler, T. Weimann, P.<br />

Hinze, M. Grunze, Adv. Mater. 2001, 13, 806 – 809.<br />

[29] J. A. Massey, M. A. Winnik, I. Manners, V. Z.-H. Chan, J. M.<br />

Ostermann, R. Enchelmaier, J. P. Spatz, M. Möller, J. Am.<br />

Chem. Soc. 2001, 123, 3147 – 3148.<br />

[30] J. P. Spatz, V. Z.-H. Chan, S. Mößmer, F.-M. Kamm, A. Plettl, P.<br />

Ziemann, M. Möller, Adv. Mater. 2002, 14, 1827 – 1832.<br />

[31] G. Binnig, H. Rohrer, C. Gerber, E. Weibel, Phys. Rev. Lett.<br />

1982, 49, 57 – 60.<br />

[32] T. Inoue, Polym. Blends H<strong>and</strong>b. 2002, 1, 547 – 576.<br />

[33] D. Snetivy, G. J. Vancso, Macromolecules 1992, 25, 3320 – 3322.<br />

[34] Y. Thomann, J. Suhm, R. Thomann, G. Bar, R.-D. Maier, R.<br />

Mülhaupt, Macromolecules 1998, 31, 5441 – 5449.<br />

[35] E.-L. Florin, V. T. Moy, H. E. Gaub, Science 1994, 264, 415 –<br />

417.<br />

[36] V. T. Moy, E.-L. Florin, H. E. Gaub, Science 1994, 266, 257 –<br />

259.<br />

[37] G. U. Lee, L. A. Chrisey, R. J. Colton, Science 1994, 266, 771 –<br />

773.<br />

[38] H. Skulason, C. D. Frisbie, Anal. Chem. 2002, 74, 3096 – 3104.<br />

[39] M. Ringger, H. R. Hidber, R. Schlögl, P. Oelhafen, H.-J.<br />

Güntherodt, Appl. Phys. Lett. 1985, 46, 832 – 834.<br />

[40] D. M. Eigler, E. K. Schweizer, Nature 1990, 344, 524 – 526.<br />

[41] R. Neffati, A. Alexeev, S. Saunin, J. C. M. Brokken-Zijp, D.<br />

Wouters, S. Schmatloch, U. S. Schubert, J. Loos, Macromol.<br />

Rapid Commun. 2003, 24, 113 – 117.<br />

[42] T.-C. Shen, Surf. Rev. Lett. 2000, 7, 683 – 688.<br />

[43] R. M. Nyffenegger, R. M. Penner, Chem. Rev. 1997, 97, 1195 –<br />

1230.<br />

[44] R. W. Carpick, Chem. Rev. 1997, 97, 1163 – 1194.<br />

[45] J. C. Rosa, M. Wendel, H. Lorenz, J. P. Kotthaus, M. Thomas, H.<br />

Kroemer, Appl. Phys. Lett. 1998, 73, 2684 – 2686.<br />

[46] B. Klehn U. Kunze, J. Appl. Phys. 1999, 85, 3897 – 3903.<br />

[47] M. Heyde, K. Rademann, B. Cappella, M. Geuss, H. Sturm, T.<br />

Spangenberg, H. Niehus, Rev. Sci. Instrum. 2001, 72, 136 – 141.<br />

[48] Image courtesy of S. N. Saunin, State Research Institute of<br />

Physical Problems & NT-MDT, Moscow, 2002.<br />

[49] L. T. Hansen, A. Kuhle, A. H. Sorensen, J. Bohr, P. E. Lindelof,<br />

Nanotechnology 1998, 9, 337 – 342.<br />

[50] T. R. Ramach<strong>and</strong>ran, C. Baur, A. Bugacov, A. Madhukar, B. E.<br />

Koel, A. Requicha, C. Gazen, Nanotechnology 1998, 9, 237 –<br />

245.<br />

[51] T. Junno, K. Deppert, L. Montelius, L. Samuelson, Appl. Phys.<br />

Lett. 1995, 66, 3627 – 3629.<br />

[52] C. Ritter, M. Heyde, U. D. Schwarz, K. Rademann, Langmuir<br />

2002, 18, 7798 – 7803.<br />

[53] J.-F. Liu, S. Cruchon-Dupeyrat, J. C. Garno, J. Frommer, G.-Y.<br />

Liu, Nano Lett. 2002, 2, 937 – 940.<br />

[54] J. C. Garno, Y. Yang, N. A. Amro, S. Cruchon-Dupeyrat, S.<br />

Chen, G.-Y. Liu, Nano Lett. 2003, 3, 389 – 395.<br />

[55] Y. Nakamura, Y. Mera, K. Maeda, Phys. Rev. Lett. 2002, 89,<br />

2668051 – 2668054.<br />

[56] A. J. Heinrich, C. P. Lutz, J. A. Gupta, D. M. Eigler, Science<br />

2002, 298, 1381 – 1387.<br />

[57] M. Su, V. P. Dravid, Appl. Phys. Lett. 2002, 80, 4434 – 4436.<br />

[58] R. McKendry, W. T. S. Huck, B. Weeks, M. Fiorini, C. Abell, T.<br />

Rayment, Nano Lett. 2002, 2, 713 – 716.<br />

[59] P. V. Schwartz, Langmuir 2002, 18, 4041 – 4046.<br />

[60] R. D. Piner, J. Zhu, S. Hong, C. A. Mirkin, Science 1999, 283,<br />

661 – 663.<br />

[61] S. Hong, J. Zhu, C. A. Mirkin, Science 1999, 286, 523 – 525.<br />

[62] C. A. Mirkin, MRS Bull. 2000, 25, 43 – 54.<br />

[63] C. A. Mirkin, Inorg. Chem. 2000, 39, 2258 – 2272.<br />

[64] S. Hong, J. Zhu, C. A. Mirkin, Langmuir 1999, 15, 7897 – 7900.<br />

[65] B. W. Maynor, Y. Li, J. Liu, Langmuir 2001, 17, 2575 – 2578.<br />

[66] S. Rozhok, R. Piner, C. A. Mirkin, J. Phys. Chem. B 2003, 107,<br />

751 – 757.<br />

[67] P. E. Sheehan, L. J. Whitman, Phys. Rev. Lett. 2002, 88,<br />

1561041 – 1561044.<br />

[68] K.-B. Lee, S.-J. Park, C. A. Mirkin, J. C. Smith, M. Mrksich,<br />

Science 2002, 295, 1702 – 1705.<br />

[69] L. M. Demers, C. A. Mirkin, Angew. Chem. 2001, 113, 3159 –<br />

3161; Angew. Chem. Int. Ed. 2001, 40, 3069 – 3071.<br />

[70] X. Liu, L. Fu, S. Hong, V. P. Dravid, C. A. Mirkin, Adv. Mater.<br />

2002, 14, 231 – 234.<br />

[71] M. Su, X. Liu, S.-Y. Li, V. P. Dravid, C. A. Mirkin, J. Am. Chem.<br />

Soc. 2002, 124, 1560 – 1561.<br />

[72] See Ref. [61].<br />

[73] D. S. Ginger, H. Zhang, C. A. Mirkin, Angew. Chem. 2004, 116,<br />

30 – 46; Angew. Chem. Int. Ed. 2004, 43, 30 – 45.<br />

[74] H. J. Mamin, D. Rugar, Appl. Phys. Lett. 1992, 61, 1003 – 1005.<br />

[75] B. D. Terris, S. A. Rishton, H. J. Mamin, R. P. Ried, D. Rugar,<br />

Appl. Phys. 1998, 66, S809 – S813.<br />

[76] H. J. Mamin, B. D. Terris, L.-S. Fan, S. Hoen, R. C. Barrett, D.<br />

Rugar, IBM J. Res. Dev. 1995, 39, 681 – 700.<br />

[77] B. W. Chui, T. D. Stone, T. W. Kenny, H. J. Mamin, B. D. Terris,<br />

D. Rugar, Appl. Phys. Lett. 1996, 69, 2767 – 2769.<br />

[78] R. P. Ried, H. J. Mamin, B. D. Terris, L.-S. Fan, D. Rugar, J.<br />

Microelectromech. Syst. 1997, 6, 294 – 302.<br />

[79] H. J. Mamin, R. P. Ried, B. D. Terris, D. Rugar, Proc. IEEE<br />

1999, 87, 1014 – 1027.<br />

[80] S. Hong, C. A. Mirkin, Science 2000, 288, 1808 – 1811.<br />

2494 2004 Wiley-VCH Verlag GmbH &Co. KGaA, Weinheim www.angew<strong>and</strong>te.org Angew. Chem. Int. Ed. 2004, 43, 2480 – 2495


<strong>Nanolithography</strong><br />

Angew<strong>and</strong>te<br />

Chemie<br />

[81] U. Dürig, G. Cross, M. Despont, U. Drechsler, W. Häberle, M. I.<br />

Lutwyche, H. Rothuizen, R. Stutz, R. Widmer, P. Vettiger,<br />

G. K. Binnig, W. P. King, K. E. Goodson, Tribol. Lett. 2000, 9,<br />

25 – 32.<br />

[82] P. Vettiger, M. Despont, U. Drechsler, U. Dürig, W. Häberle,<br />

M. I. Lutwyche, H. Rothuizen, R. Stutz, R. Widmer, G. K.<br />

Binnig, IBM J. Res. Dev. 2000, 44, 323 – 340.<br />

[83] P. Vettiger, G. Binnig, Sci. Am. 2003, Jan., 35 – 40.<br />

[84] S. C. Minne, J. D. Adams, G. Yaralioglu, S. R. Manalis, A.<br />

Atalar, C. F. Quate, Appl. Phys. Lett. 1998, 73, 1742 – 1744.<br />

[85] M. Lutwyche, C. Andreoli, G. Binnig, J. Brugger, U. Drechsler,<br />

W. Haeberle, H. Rohrer, H. Rothuizen, P. Vettiger, “Microfabrication<br />

<strong>and</strong> Parallel Operation of 5 ” 5 2D AFM Cantilever<br />

Array for Data Storage <strong>and</strong> Imaging”, Proceedings of the IEEE<br />

11th International Workshop on Micro Electro Mechanical<br />

Systems, Heidelberg, 1998, 8–11.<br />

[86] M. I. Lutwyche, G. Cross, M. Despont, U. Drechsler, U. Dürig,<br />

W. Häberle, H. Rothuizen, R. Stutz, R. Widmer, G. K. Binnig,<br />

P. Vettiger, IEEE International Solid-State Circuits Conference,<br />

San Francisco, 2000..<br />

[87] S. C. Minne, P. Flueckiger, H. T. Soh, C. F. Quate, J. Vac. Sci.<br />

Technol. B 1995, 13, 1380 – 1385.<br />

[88] T. Akiyama, U. Staufer, N. F. de Rooij, D. Lange, C. Hagleitner,<br />

O. Br<strong>and</strong>, H. Baltes, A. Tonin, H. R. Hidber, J. Vac. Sci.<br />

Technol. B 2000, 18, 2669 – 2675.<br />

[89] J.-B. D. Green, A. Novoradovsky, D. Park, G. U. Lee, Appl.<br />

Phys. Lett. 1999, 74, 1489 – 1491.<br />

[90] U. Kleineberg, A. Brechling, M. Sundermann, U. Heinzmann,<br />

Adv. Funct. Mater. 2001, 11, 208 – 212.<br />

[91] J. Hartwich, M. Sundermann, U. Kleineberg, U. Heinzmann,<br />

Appl. Surf. Sci. 1999, 144–145, 538 – 542.<br />

[92] M. A. McCord, R. F. W. Pease, J. Vac. Sci. Technol. B 1987, 5,<br />

430 – 433.<br />

[93] M. A. McCord, R. F. W. Pease, J. Vac. Sci. Technol. B 1988, 6,<br />

293 – 296.<br />

[94] R. Wiesendanger, Appl. Surf. Sci. 1992, 54, 271 – 277.<br />

[95] G. M. Shedd, P. Russell, Nanotechnology 1990, 1, 67 – 80.<br />

[96] J. A. Dagata, J. Schneir, H. H. Harary, C. J. Evans, M. T. Postek,<br />

J. Bennet, Appl. Phys. Lett. 1990, 56, 2001 – 2003.<br />

[97] P. M. Campbell, E. S. Snow, P. J. McMarr, Appl. Phys. Lett.<br />

1993, 63, 749 – 751.<br />

[98] M. Yasutake, Y. Ejiri, T. Hattori, Jpn. J. Appl. Phys. 1993, 32,<br />

L1021 – L1023.<br />

[99] N. Kramer, H. Birk, J. Jorritsma, C. Schönenberger, Appl. Phys.<br />

Lett. 1995, 66, 1325 – 1327.<br />

[100] H. Sugimura, T. Uchida, N. Kitamura, H. Masuhara, Appl.<br />

Phys. Lett. 1993, 63, 1288 – 1290.<br />

[101] H. Sugimura, N. Nakagiri, Jpn. J. Appl. Phys. 1995, 34, 3406 –<br />

3411.<br />

[102] C. R. K. Marrian, E. A. Dobsisz, J. Vac. Sci. Technol. B 1992, 10,<br />

2877 – 2881.<br />

[103] A. Majudmar, P. I. Oden, J. P. Carrejo, L. A. Nagahara, J. J.<br />

Graham, J. Alex<strong>and</strong>er, Appl. Phys. Lett. 1992, 61, 2293 – 2295.<br />

[104] C. R. K. Marrian, F. K. Perkins, S. L. Br<strong>and</strong>ow, T. S. Koloski,<br />

E. A. Dobisz, J. M. Calvert, Appl. Phys. Lett. 1994, 64, 390 –<br />

392.<br />

[105] L. Stockman, G. Neuttiens, C. van Haesendonck, Y. Bruynseraede,<br />

Appl. Phys. Lett. 1993, 62, 2935 – 2937.<br />

[106] E. S. Snow, P. M. Campbell, Science 1995, 270, 1639 – 1641.<br />

[107] S. C. Minne, H. T. Soh, P. Flueckiger, C. F. Quate, Appl. Phys.<br />

Lett. 1995, 66, 703 – 705.<br />

[108] R. Martel, T. Schmidt, R. L. S<strong>and</strong>strom, P. Avouris, J. Vac. Sci.<br />

Technol. A 1999, 17, 1451 – 1456.<br />

[109] S. W. Park, H. T. Soh, C. F. Quate, S.-I. Park, Appl. Phys. Lett.<br />

1995, 67, 2415 – 2417.<br />

[110] M. Tello, F. García, R. García, J. Appl. Phys. 2002, 92, 4075 –<br />

4079.<br />

[111] R. D. Ramsier, R. M. Ralich, S. F. Lyuksytov, Appl. Phys. Lett.<br />

2001, 79, 2820 – 2822.<br />

[112] S. J. Ahn, Y. K. Jang, S. A. Kim, H. Lee, Ultramicroscopy 2002,<br />

91, 171 – 176.<br />

[113] Q. Li, J. Zheng, Z. Liu, Langmuir 2003, 19, 166 – 171.<br />

[114] H. Sugimura, T. Hanji, K. Hayashi, O. Takai, Ultramicroscopy<br />

2002, 91, 221 – 226.<br />

[115] H. Sugimura, T. Hanji, K. Hayashi, O. Takai, Adv. Mater. 2002,<br />

14, 524 – 526.<br />

[116] H. Sugimura, O. Takai, N. Nakagiri, J. Electroanal. Chem. 1999,<br />

473, 230 – 234.<br />

[117] W. Lee, E. R. Kim, H. Lee, Langmuir 2002, 18. 8375 – 8380.<br />

[118] J. Zhao, K. Uosaki, Langmuir 2001, 17, 7784 – 7788.<br />

[119] M. Rol<strong>and</strong>i, C. F. Quate, H. Dai, Adv. Mater. 2002, 14, 191 – 194.<br />

[120] S. Liu, R. Maoz, G. Schmid, J. Sagiv, Nano Lett. 2002, 2, 1055 –<br />

1060.<br />

[121] R. Maoz, S. R. Cohen, J. Sagiv, Adv. Mater. 1999, 11, 55 – 61.<br />

[122] R. Maoz, E. Frydman, S. R. Cohen, J. Sagiv, Adv. Mater. 2000,<br />

12, 424 – 429.<br />

[123] R. Maoz, E. Frydman, S. R. Cohen, J. Sagiv, Adv. Mater. 2000,<br />

12, 725 – 731.<br />

[124] S. Hoeppener, R. Maoz, S. R. Cohen, L. Chi, H. Fuchs, J. Sagiv,<br />

Adv. Mater. 2002, 14, 1036 – 1041.<br />

[125] V. Checknik, R. M. Crooks, C. J. M. Stirling, Adv. Mater. 2000,<br />

12, 1161 – 1171.<br />

[126] D. Wouters, U. S. Schubert, Langmuir 2003, 19, 9033 – 9038.<br />

[127] D. Wouters, B. Kösters, U. S. Schubert, Polym. Prepr. Am.<br />

Chem. Soc. Div. Polym. Chem. 2003, 44, 169 – 170.<br />

[128] D. Wouters, B. Kösters, U. S. Schubert, 2003, unpublished<br />

results.<br />

[129] D. J. Díaz, J. E. Hudson, G. D. Storrier, H. D. Abruna, N.<br />

Sundararajan, C. K. Ober, Langmuir 2001, 17, 5932 – 5938.<br />

Angew. Chem. Int. Ed. 2004, 43, 2480 – 2495 www.angew<strong>and</strong>te.org 2004 Wiley-VCH Verlag GmbH &Co. KGaA, Weinheim<br />

2495

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!