+ All Categories
Home > Documents > Structuration and Integration of Magnetic Nanoparticles on Surfaces and Devices

Structuration and Integration of Magnetic Nanoparticles on Surfaces and Devices

Date post: 22-Apr-2023
Category:
Upload: independent
View: 0 times
Download: 0 times
Share this document with a friend
27
1 © 2012 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim wileyonlinelibrary.com Magnetic Nanoparticles Structuration and Integration of Magnetic Nanoparticles on Surfaces and Devices Elena Bellido, Neus Domingo, Isaac Ojea-Jiménez, and Daniel Ruiz-Molina* Different experimental approaches used for structuration of magnetic nanoparticles on surfaces are reviewed. Nanoparticles tend to organize on surfaces through self-assembly mechanisms controlled by non-covalent interactions which are modulated by their shape, size and morphology as well as by other external parameters such as the nature of the solvent or the capping layer. Further control on the structuration can be achieved by the use of external magnetic fields or other structuring techniques, mainly lithographic or atomic force microscopy (AFM)- based techniques. Moreover, results can be improved by chemical functionalization or the use of biological templates. Chemical functionalization of the nanoparticles and/or the surface ensures a proper stability as well as control of the formation of a (sub)monolayer. On the other hand, the use of biological templates facilitates the structuration of several families of nanoparticles, which otherwise may be difficult to form, simply by establishing the experimental conditions required for the structuration of the organic capsule. All these experimental efforts are directed ultimately to the integration of magnetic nanoparticles in sensors which constitute the future generation of hybrid magnetic devices. Introduction 1. .............................................. 2 Un-assisted Self-Assembly 2. ..........................2 Assisted Self-Assembly and 3. Organization ...............................................6 Use of Biomolecular 4. Templates................................................. 15 Integration into Hybrid 5. Devices..................................................... 19 Summary and Perspectives 6. ...................... 23 From the Contents small 2012, DOI: 10.1002/smll.201101456
Transcript

Magnetic Nanoparticles

Structuration and Integration of Magnetic Nanoparticles on Surfaces and Devices Elena Bellido , Neus Domingo , Isaac Ojea-Jiménez , and Daniel Ruiz-Molina *

1© 2012 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim wileyonlinelibrary.com

Different experimental approaches used for structuration of magnetic nanoparticles on surfaces are reviewed. Nanoparticles tend to organize on surfaces through self-assembly mechanisms controlled by non-covalent interactions which are modulated by their shape, size and morphology as well as by other external parameters such as the nature of the solvent or the capping layer. Further control on the structuration can be achieved by the use of external magnetic fi elds or other structuring techniques, mainly lithographic or atomic force microscopy (AFM)-based techniques. Moreover, results can be improved by chemical functionalization or the use of biological templates. Chemical functionalization of the nanoparticles and/or the surface ensures a proper stability as well as control of the formation of a (sub)monolayer. On the other hand, the use of biological templates facilitates the structuration of several families of nanoparticles, which otherwise may be diffi cult to form, simply by establishing the experimental conditions required for the structuration of the organic capsule. All these experimental efforts are directed ultimately to the integration of magnetic nanoparticles in sensors which constitute the future generation of hybrid magnetic devices.

Introduction 1. .............................................. 2

Un-assisted Self-Assembly2. ..........................2

Assisted Self-Assembly and 3. Organization ...............................................6

Use of Biomolecular 4. Templates ................................................. 15

Integration into Hybrid 5. Devices ..................................................... 19

Summary and Perspectives 6. ......................23

From the Contents

small 2012, DOI: 10.1002/smll.201101456

E. Bellido et al.reviews

DOI: 10.1002/smll.201101456

Dr. E. Bellido , Dr. N. Domingo , Dr. I. Ojea-Jiménez , Dr. D. Ruiz-Molina Centro de Investigación en Nanociencia y Nanotecnología, (CIN2, ICN-CSIC)Esfera UAB. Campus UAB, Edifi cio CM7Cerdanyola del Vallès, 08193, Spain E-mail: [email protected]

Figure 1 . Schematic illustration of the experimental approaches used for the deposition of MNPs as self-assembled mono- and multi-layers on surfaces: a) drop-casting, b) substrate immersion into a colloidal solution, c) dip-coating, d) Langmuir–Blodgett deposition, e) spin-coating, and (f) tape-casting (also known as doctor-blade-casting). Images not drawn to scale.

1. Introduction

Due to their unique electrical, [ 1 ] optical, [ 2 , 3 ] and magnetic [ 4–6 ]

properties arising from their fi nite size and high surface area-

to-volume ratio, nanoparticles (NPs) have recently attracted

great interest for applications in different fi elds of nanotech-

nology. In particular, magnetic nanoparticles (MNPs) repre-

sent an important class of nanostructured materials [ 7 ] whose

exclusive properties have been extensively studied due to

their potential use in many applications such as magnetic

storage media, [ 8–10 ] spin transport devices, [ 11 ] medical drug

delivery, [ 12–14 ] and magnetic resonance imaging (MRI), [ 15 , 16 ]

among others. However, for such applications to become

a reality, the development of proper techniques for their

structuration on surfaces with control over their positioning,

geometry, and organization is a crucial issue. Such advances

can fuel the detailed understanding of their individual and

collective properties by having the possibility to study them

at the individual (or small number) level. [ 17–21 ] This would

allow, for instance, a detailed study of the shape- and size-

dependence of their magnetic properties as well as the mag-

netic interaction between MNPs at a desired interparticle

distance and organization. Moreover, from an applications

point-of-view, the continuous miniaturization of MNP-based

structures on surfaces has advantages in the development

of magnetic devices with higher storage densities and faster

speeds, [ 10 ] and the fabrication of new types of magnetoelec-

tronic devices. [ 22 , 23 ]

With the introduction of several micro- and nano-fabri-

cation techniques such as microcontact printing ( μ CP), [ 24 , 25 ]

electron beam lithography (EBL), [ 26 ] and atomic force micro-

scopy (AFM)-based lithographies, [ 27–30 ] scientists started the

fi rst steps to organize MNPs into ordered structures at the

nanometer scale. Since then, the number of methodologies

and approaches followed has been continuously increasing

so that, to date, one of the major challenges is the accurately

placing of a few or even a single MNP on the desired region

of a given surface. While there have been other comprehen-

sive reviews on various aspects of the nanoscale organization

of different building blocks on surfaces and interfaces [ 31–34 ]

as well as the nanofabrication of magnetic nanostructures for

their application in data storage media, [ 35 , 36 ] there is a need

for a comprehensive and detailed description of the different

strategies employed or developed for assembling MNPs into

ordered nanostructures. In this review, fi rst the different

approaches and factors infl uencing the self-assembly of

MNPs on surfaces are considered. Following this, examples of

self-assembly methods assisted by lithographic techniques, or

even external-fi eld- or biological template-assisted methods

are reviewed. Finally, in the last section, an overview of the

most representative experimental approaches so far, followed

for the integration of the MNPs on sensors and devices, is

given.

2. Un-assisted Self-Assembly

Self-assembly is the autonomous organization of pre-

existing entities through non-covalent interactions leading

2 www.small-journal.com © 2012 Wiley-VCH Ve

to spontaneously formed hierarchical and complex architec-

tures. [ 37 , 38 ] While being already common in Nature for the

creation of multi-dimensional and complex biological struc-

tures such as cells, [ 38 ] the concept of self-assembly has been

more recently used as a powerful strategy for controlling the

structure and properties of man-made nanostructures. MNPs

have not been indifferent to this approach. Self-assembly has

been undoubtedly one of the most fruitful approaches for

their organization on surfaces infl uenced by three main fac-

tors: i) deposition techniques, ii) category of interactions, and

iii) geometry and environmental parameters.

2.1. Deposition Techniques

The different deposition methodologies that can be used to

induce the self-assembly of MNPs on surfaces are shown in

Figure 1 . Among them, the simplest technique consists of

the deposition of a MNP colloidal solution mainly by drop-

casting (Figure 1 a) or directly immersing the substrate into

the solution (Figure 1 b). In both cases, the slow evaporation

of the solvent while in contact with the surface facilitates the

diffusion of the particles to the most favorable energetic and

ordered state. For this reason it is necessary to have control

rlag GmbH & Co. KGaA, Weinheim small 2012, DOI: 10.1002/smll.201101456

Magnetic Nanoparticles on Surfaces and Devices

Elena Bellido was born in Barcelona (Spain),

in 1982. She received a B.Sc. degree in Chem-

istry in 2006 at Autonomous University of

Barcelona (UAB). In 2011 she got her Ph.D.

at the Research Center on Nanoscience and

Nanotechnology (CIN2) under the supervi-

sion of Dr. Daniel Ruiz Molina. Her research

focuses on the development of novel strate-

gies for the structuration and integration of

functional materials on surfaces and devices

by means of AFM-based nanolithographic

techniques.

Daniel Ruiz-Molina received his Ph.D. on

polyradical dendrimers at the Institute of

Materials Science of Barcelona (CSIC)

with Prof. Jaume Veciana. He then took a

postdoctoral position at the UC San Diego

working with Prof. David N. Hendrickson

on single-molecule magnets and molecular

switches for three years. Since 2001 he has

held a permanent position at the CSIC and

more recently at the new Research Center

on Nanoscience and Nanotechnology, where

he is heading the Nanostructured Functional

Materials group. His main research areas are

fabrication of hybrid colloids and surfaces,

biomimetic functional nanostructures, and micro-/nano-particles for smart applica-

tions and encapsulation/delivery systems.

Isaac Ojea-Jiménez was born in Hospitalet

de Llobregat (Spain) in 1977. He got his PhD

in the UK working on asymmetric organic

chemistry at the University of Sheffi eld in

2005. After spending 1 year in Germany as a

post-doc in Roche Diagnostics, he returned to

Barcelona to start working in nanotechnology

for the industry. After 2 years he moved back

to the academic fi eld in a position in the Inor-

ganic Nanoparticles Laboratory at the Institut

Català de Nanotecnologia (ICN) working for

Prof. Victor Puntes. His current research in-

terests include synthesis and functionalization

of metallic and semiconductor nanoparticles,

as well as their applications in biomedicine.

over the temperature and/or the evaporation rate by the use

of high-boiling-temperature solvents (such as dichloroben-

zene, toluene, or octyl ether) and/or increasing the degree

of solvent saturation of the surrounding atmosphere. Alter-

natively, self-assembly can take place upon dip-coating the

substrate using a solution containing the particles (Figure 1 c).

Under these conditions, the MNPs tend to organize leading

to local, well-ordered, 2D and/or 3D supperlattices, strongly

infl uenced by the surfactant layer nature [ 9 ] and the particle

shape. [ 39 ] The organization of Co NPs ( ∼ 7 nm) into 3D super-

lattices with this technique has been extensively studied by

Pileni and co-workers. [ 40 ] They observed a clear dependence

of the structural organization on the temperature of the sub-

strate, which relied on the diffusion coeffi cient of the MNPs,

their solubility in the solvent, and the fl uidity of the coating

layer.

The three methodologies previously described are well-

suited to obtain local ordered domains of MNPs, but are

diffi cult to reproduce over extended areas. Particle distri-

bution and nucleation/growth of the assembly process are

not completely controlled with these techniques. This lack

of control results in inhomogeneities, different layer thick-

ness and undesired gaps among the regular domains, all

of these being major drawbacks for applications where

MNPs need to be arranged tightly and uniformly packed

over large areas such as ultrahigh-density magnetic storage

media. [ 10 ] Therefore, other methods to uniformly assemble

MNPs at high-scale areas have been employed, including

Langmuir–Blodgett (LB), spin-coating, and tape-casting

techniques (see Figure 1 d–f). With the LB method, MNPs

are forced to fl oat over a liquid surface (e.g., using as liquid

subphase water or less polar solvents such as glycol), which

results in the arrangement of the MNPs as monolayers at the

air–liquid interface. Then, the fl oating fi lm is transferred to a

solid surface (e.g., highly oriented pyrolitic graphite (HOPG)

or silicon) by dipping a vertically oriented substrate into the

liquid subphase (see Figure 1 d). [ 41 , 42 ] Alternatively the sub-

strate can also be dipped in a horizontal orientation, which

gives rise to the Langmuir–Shaefer technique. [ 43 , 44 ] Both

techniques have the potential to produce uniform 2D and 3D

layers over vast areas. For instance, well-ordered fi lms ranging

from the micrometer up to the millimeter range of MNPs

with different sizes (8-10 nm), shapes, and composition (Co

to Pt ratio) have been obtained by Weller and co-workers. [ 42 ]

These fi lms were successfully transferred, not only on silicon

surfaces, but also on samples structured with Au electrodes to

perform direct current (DC) measurements ( Figure 2 a–c).

With the spin-coating technique the solution is rapidly

spun at high revolutions over the substrate achieving solvent

evaporation and a good spreading of the fi lm in fractions of

seconds (see Figure 1 e). By varying the MNP concentration,

it is possible to easily control the MNP density of the layer at

a given rotational speed. [ 46 ] Implicit to the technique, spin-

coating is not feasible in delivering MNPs on non-planar sur-

faces, though it can successfully generate 2D fi lms of MNPs

on fl at substrates like Si/Si 3 N 4 substrates. [ 47 ]

Finally the tape-casting method (also known as doctor-

blade-casting) is a large-area coating technique for materials

processable in solution. First a solution containing the MNPs is

© 2012 Wiley-VCH Verlag GmbHsmall 2012, DOI: 10.1002/smll.201101456

drop-cast on one end of the substrate and then uniformly spread

over the whole surface by moving the blades (see Figure 1 f).

This method has been employed to self-assemble ∼ 11 nm parti-

cles with a wüstite (Fe x O) core and a cobalt ferrite (CoFe 2 O 4 )

shell into thin superlattice fi lms, on both hydrophobized Si sub-

strates and Pt-coated Si substrates (Figure 2 d–g). [ 45 ]

2.2. Category of Interaction

In this section the different families of non-covalent interac-

tions that favor a proper interplay between the MNPs and

the substrate as well as between MNPs are reviewed. One

of the interactions that has been reported to infl uence the

self-assembly of MNPs is dipolar interaction. In some cases,

as claimed by the authors, these inter actions can dominate

3www.small-journal.com & Co. KGaA, Weinheim

E. Bellido et al.

4 www.small-journal.com © 2012 Wiley-VCH Verlag GmbH & Co. KGaA, We

reviews

Figure 2 . a) Scanning electron microscopy (SEM) image of a cobalt-platinum NP fi lm deposited by the Langmuir–Blodgett method on electron-beam lithographically patterned Au electrodes, b) magnifi ed view of the NP fi lm, c) schematic illustration of the device. Reproduced with permission. [ 42 ] Copyright 2008, American Chemical Society. d) Background SEM image of ∼ 11 nm Fe x O/CoFe 2 O 4 NPs on a Pt-covered Si substrate by tape-casting method. Cross-sectional transmission electron microscopy (TEM) images of a thin NP superlattice made of 11 nm spherical-shaped (e), and ∼ 11 nm cubic-shaped (f,g) Fe x O/CoFe 2 O 4 NPs. Reproduced with permission. [ 45 ] Copyright 2009, American Chemical Society.

over the isotropic ones, such as van der

Waals, steric repulsions, or temperature

diffusion effects.

The magnetic interaction energy for

two dipoles in units of thermal energy

k B T is:

VkBT = µo µ2

4πkBTr 2[̂S1 · ̂S2 − 3(̂S1·)(̂S2·)]

where ̂Sl are the orientation unit vectors

of the magnetic dipoles, separated by a

center-to-center distance of ̂r , and μ is

the magnetic moment of a particle. Con-

sidering a particle with fully saturated

magnetization, that is a single-domain

particle, of volume v p , and saturation

magnetization of the bulk m s , the mag-

netic moment can be approximated by:

μ = v p m s . Thus, the magnetic dipolar

interactions increase quadratically with

the particle magnetic moment, which

is also proportional to its volume. In

this sense, dipole–dipole interactions

become effective for big enough par-

ticles with high magnetic moments.

Still, the thermally fl uctuating mag-

netic moment of a superparamagnetic

nanoparticle can also be stabilized

when the particle is in the vicinity of

another particle and under the action

of dipole–dipole interaction. Puntes

et al., [ 48 ] for example, estimated that the

maximum dipolar fi eld created along the

axis between two Co NPs (with diameter

of ∼ 6 nm, magnetic moment of the order of

μ ≈ 10 4 μ B , and blocking temperature of

T B = 130 K) in contact, is of the order

of 7000 Oe. Thus, the energy of a Co NP

in such a fi eld is one order of magnitude

higher than its thermal energy at room

temperature and can also play an impor-

tant role in the self-assembly process of

superparamagnetic nanoparticles during

the evaporation process.

Dipolar interactions have been

stated as responsible for the chain-like

assembly of ferrite NPs, both in their

reduced (Fe 3 O 4 ) and oxidized ( γ -Fe 2 O 3 )

forms, as well as for the assembly of Fe 3 S 4

and Fe 7 S 8 NPs inside magnetotactic

bacteria, which gives them the ability

to align themselves and swim along the

earth's magnetic fi eld. [ 49–51 ] Dipolar

interactions are thus at the origin of

magnetic anisotropy of chains of nano-

particles. In the case of an assembly of

aligned magnetosomes, it has been stated

that the origin of the magnetic anisot-

ropy of this system of aligned magnetic

inheim small 2012, DOI: 10.1002/smll.201101456

Magnetic Nanoparticles on Surfaces and Devices

Figure 3 . Cryo-TEM images of Fe NPs with different diameters (measured by TEM): a) ∼ 4 nm, b) ∼ 14 nm, and c) ∼ 16 nm. Mixture of Fe NPs with diameters of ∼ 4 and ∼ 14 nm in zero fi eld (d), and under external magnetic fi eld infl uence (e), with magnifi ed 4x image (inset). Reproduced with permission. [ 53 ] Copyright 2003, Macmillan Publishers Ltd: Nature Materials.

nanoparticles is the dipolar interactions (up

to 70%), while the contribution of the align-

ment of their individual easy axis is much

smaller (only of a 30%). [ 51 ]

Dipolar interactions have been fur-

ther investigated for Fe, Fe 3 O 4 , and Co

NPs. [ 48 , 49 , 51–53 ] While van der Waals inter-

actions and steric forces can be modulated

using different type of surfactants, dipolar

magnetic interactions can be modifi ed by

changing the size of the magnetic NPs.

By systematically increasing the MNP

diameter, and hence the magnetic dipole

moment (which is proportional to the

third power of the radius), Philipse et al.

tuned the dipole–dipole inter action of Fe

NPs coated with a thick surfactant layer. [ 53 ]

The self-assembly of such particles in the

liquid-phase, and in the absence of a mag-

netic fi eld, was studied with cryogenic

TEM (cryo-TEM), revealing the evolution

from isolated to randomly oriented chain-

like MNPs structures on going from ∼ 4 to

∼ 20 nm in diameter, respectively ( Figure 3 ).

To demonstrate that the chain-like struc-

tures formed arise from dominant dipolar

forces, an external magnetic fi eld (1.6 T)

was applied to the colloidal solution. By

contraposition with the previous results,

an alignment of the assemblies along the

fi eld direction in the form of parallel bun-

dles (Figure 3 d-e) was observed. This is the

characteristic assembly of weakly dipolar

MNPs that do not spontaneously organize

when orienting their magnetic moment

along an applied fi eld. [ 54 ] Similar results

were observed by Puntes et al. who observed disordered

magnetic aggregation and precipitation of 14 nm Co NP. [ 48 ] In

these cases, formation of ordered structures is prevented due

to too strong magnetic interactions.

In the absence of dipolar interactions, that is, for small

enough nanoparticles with low magnetic moments, isotropic

van der Waals interactions dominate the self-assembly

process and 2D or 3D arrays of NPs are formed. For instance,

Pileni et al. investigated the self-assembly of ∼ 10 nm γ -Fe 2 O 3

where the length of the capping layer is modifi ed by using

either octanoic (C 8 ) or dodecanoic acid (C 12 ). [ 55 ] The experi-

ments showed that the use of shorter alkyl chains favor the

presence of stronger van der Waals interactions, so directing

their self-assembly into spherical aggregates.

2.3. Geometry and Environmental Parameters

In addition to the type of interaction and/or deposition

methodology, there are other factors such as particle con-

centration, nature of the surfactant layer, and/or solvent, or

the morphology of the MNPs that have also been shown to

strongly infl uence the self-assembly.

© 2012 Wiley-VCH Verlag Gmsmall 2012, DOI: 10.1002/smll.201101456

Particle shape and morphology . These parameters clearly

infl uence an optimum packed with minimized energies. A nice

example of the morphology infl uence on the self-assembly proc-

esses was obtained by drop-casting a diluted solution of (Fe x O)

NPs onto a carbon-coated TEM grid. Well-defi ned cubic-

ordered superlattices were obtained when cubic-shaped wüstite

(Fe x O) NPs (11–13 nm in size) were used. [ 39 ] This result is in

contrast with the use of spherical Fe x O analogues (8–10 nm) [ 39 ]

and Co NPs ( ∼ 10 nm) [ 56 ] that result in hexagonal superlattices.

Particle concentration. The infl uence of the particle con-

centration was investigated by the deposition of ∼ 10 nm γ -Fe 2 O 3 NPs on different surfaces by means of the dip-coating

process. [ 57 ] The adsorption of MNPs on a surface was shown to

be strongly dependent on the particle concentration in solu-

tion, which typically leads to an increase of particle coverage

with concentration until a maximum is reached. At low con-

centrations and after repeating several times the dip and dry

process, the particle coverage is signifi cantly increased though

no multiple layers are formed because the surfactant pre-

vents their formation. In contrast, multilayered assemblies of

MNPs (organized into 3D superlattices) have been obtained

using an initial higher concentration with the drop-casting

method [ 56 , 58 ] or, otherwise, immersing the substrate into a

5www.small-journal.combH & Co. KGaA, Weinheim

E. Bellido et al.reviews

Figure 4 . TEM images of multilayered assemblies of MNPs. a) fcc multilayers of ∼ 7 nm Co NPs. Reproduced with permission. [ 58 ] b) hcp of ∼ 9 nm Co NPs. Reproduced with permission. [ 4 ] Copyright 1999, American Institute of Physics.

solution containing the MNPs and allowing a slow evapora-

tion. [ 40 , 54 , 55 , 58 ] This is the case of Co NPs self-assembled into

face-centered cubic (fcc) ( ∼ 7 nm) ( Figure 4 a) [ 58 ] and hexa-

gonal close-packed (hcp) ( ∼ 9 nm) (Figure 4 b) [ 4 ] multi layers.

Nature of the solvent. The solvents are also found to

be important in the uniformity and morphology of the

MNP layer. In a self-explaining experiment, Wang et al.

demonstrated the importance of selecting a solvent with a

suitable evaporation rate in order to allow MNPs to diffuse

on the surface to form a monolayer. [ 59 ]

Effect of the temperature. The temperature is another

important external parameter to consider, which can be used

as a powerful tool for directing self-assembly of NPs toward

desired structures. From a thermodynamic point of view, the

fi nal ordering of large self-assembled patterns of MNPs is given

by minimization of the Gibbs free energy G = H – TS . The tem-

perature can be used as the weighting factor for the enthalpy

( H ) and entropy ( S ), so that the role of entropy can effectively

be increased by raising the temperature, while keeping all other

parameters constant. The entropy changes during the self-

assembly process are generally related to the vibrational and

confi gurational degrees of freedom. A study of the self-assembly

of Fe x O/CoFe 2 O 4 NPs in chloroform onto carbon-coated

TEM grids at different temperatures (between –40 and 45 ° C)

revealed that different superlattice structures were possible. [ 60 ]

Although the formation of fcc packing density was favored over

hcp at 45 ° C due to the higher entropy, the hcp structures were

progressively dominant with decreasing temperature (down

to –20 ° C). Finally, an hexagonal ordering forming non-close-

packed structures was reproducibly observed at –40 ° C. Similar

experiments with Co NPs deposited on silicon wafers, by means

of drop-casting or spin-coating techniques, have been per-

formed at different substrate temperature. [ 61 ] It was found that,

at low temperatures (25 ° C), instabilities such as capillary fl ows

causing ring-stains frequently appear, which destroy the self-

assembly. This contrasts with the great homogeneities observed

for large areas at high temperatures (150 ° C), which are due to

the large mobility of MNPs. Vanmaekelbergh's group recently

reported a systematic study of the thermodynamic driving

force behind the formation of binary nanocrystal superlattices,

carrie d out by inducing by the evaporation of the solvent at

6 www.small-journal.com © 2012 Wiley-VCH Verlag GmbH & Co. KGaA

a controlled, and then variable, temperature

(from 20 to 70 ° C). [ 62 ] The authors showed

that at high temperature the formation of

these superlattices was driven by an increase

in entropy, but for low temperatures particle

interactions such as van der Waals became

more important.

Nature of the surfactant capping layer .

Transitions from hexagonal to a cubic

packing of ∼ 6 nm FePt NPs colloids

have been reported simply by modifying

the ligand capping layer of the NP. [ 9 ]

The exchange of the surfactant capping

layer can also improve the coverage of

particles by using a polymer-assisted

layer-by-layer assembly, which consists

on repeating a dip-coating process to

alternately adsorb polymer and MNP

layers. [ 63 , 64 ] By dipping fi rst a polymer-coated surface into a

solution containing FePt (4–6 nm) NPs the surface molecules

on the MNPs are replaced by the polymer, which forms a

strong monolayer on the substrate surface. Then, the thick-

ness of the multilayered assembly polymer/MNP was con-

trolled by adjusting the number of alternating coating steps.

Excess of surfactant molecules in the colloidal solution. Interestingly, the self-assembly of MNPs on surfaces has been

shown to be strongly dependent on the excess of surfactant

molecules freely dispersed in the colloidal solution (i.e., not

bound to the MNP surface). Kim et al. studied the infl uence

on the adsorption behavior of γ -Fe 2 O 3 ( ∼ 10 nm) NPs when

varying the concentration of surfactant molecules in the

colloidal solution. [ 57 ] In addition, to affect the viscosity of

the medium, a large excess of surfactant molecules can also

adsorb on the substrate surface and enter into competition

with the adsorption of MNPs. In contrast, a slight excess of

surfactant molecules may favor more uniform and long-range

ordering of MNPs with a relatively wide particle size distribu-

tion. For example, long-range self-assembly of ∼ 10 nm Fe 3 O 4

NPs coated with phospholipid molecules has been assisted

by the presence of free phospholipid molecules in the col-

loidal solution. [ 65 ] The authors suggested that due to the slight

excess of the lipid molecules, a phospholipid layer is formed

underneath the MNP monolayer, providing the necessary

binding force to hold them tied to the surface forming hexa-

gonal superlattices over extended areas of up to 2 cm × 2 cm.

The optimization of all these parameters (i.e., effects

of MNP shape and concentration, type of solvent, and the

presence of surfactant molecules) led to the successful depo-

sition, by the spin-coating process, of highly uniform mono-

layers of ∼ 6 nm FePt NPs with different morphology and

over large areas (1.5 cm × 1.5 cm) of a Si/SiO 2 surface. [ 59 ]

3. Assisted Self-Assembly and Organization

3.1. Chemically Assisted Assembly

Chemical assistance is an attractive approach to fabricate

highly stable and well-ordered monolayer fi lms of MNPs

, Weinheim small 2012, DOI: 10.1002/smll.201101456

Magnetic Nanoparticles on Surfaces and Devices

Figure 5 . Schematic represenation of the three main approaches used in the chemically-assisted assembly of MNPs on a substrate: a) functionalization of the NPs with linkers bearing terminal groups able to react with the surface, b) functionalization of the surface with linkers bearing terminal groups able to react with the NPs, and c) a combination of both these methods. (Images not drawn to scale).

over large substrate areas. With this approach the stability

of the fi lms is strongly improved by introducing appropriate

linkers with suitable terminal functionalities that anchor the

MNPs to the surface. Beyond the stability gain, this approach

clearly favors the monolayer distribution of the MNPs over

the substrate instead of the formation of multilayers. [ 66 , 67 ]

The use of the linkers can be done in three different ways:

(i) functionalization of the MNPs with linkers bearing ter-

minal groups able to react with the surface, (ii) functionaliza-

tion of the surface with linkers bearing terminal groups able

to react with the MNPs, and (iii) a combination of both of

these methods ( Figure 5 ).

3.1.1. Pre-functionalization of the MNPs

Usually, the coating layer of the as-synthesized MNPs is

not really suitable to react with bare substrates. For this

reason, one common strategy is to replace the coating lig-

ands by bifunctional molecules able to react both with the

surface and the MNPs (see Figure 5 a). For example, par-

tial exchange of the oleic acid coating of ∼ 7 nm Fe 3 O 4 NPs

with trimethoxy-7-octen-1-yl-silane, bearing trimethoxysi-

lane groups that interact with the Fe 3 O 4 surface, has been

achieved ( Figure 6 a). [ 66 ] The new ligand also bears free vinyl

groups that can be anchored to H-terminated Si surfaces

through a stable covalent C–Si bond (Figure 6 b), resulting

in homogeneous coverage of the substrate (Figure 6 c). Fur-

ther magnetization studies using a SQUID magnetometer

showed that the MNPs effectively bound to the silicon

surface without any signifi cant alteration of their magnetic

properties (i.e., zero-fi eld-cooled (ZFC) and fi eld-cooled

(FC) magnetizations, and hysteresis loop). Parallel studies

made use of 10-undecynoic acid as a cross-linker between

∼ 5 nm Fe 3 O 4 NPs and the H-terminated Si surface. [ 68 ] An

extensive characterization of the MNP layer showed a

© 2012 Wiley-VCH Verlag GmbHsmall 2012, DOI: 10.1002/smll.201101456

close-packing arrangement of the MNPs on the surface

with a non-altered morphology. Moreover, regions where

the MNPs formed a second layer were also observed due

to the terminal alkyne dimerization occurring between two

10-undecynoic acid molecules involving the fi rst and second

MNP layer.

Alternatively, an UV-light-induced hydrosilylation

route has also been effectively used for the immobilization

of ∼ 5 nm γ -Fe 2 O 3 NPs coated with a mixed capping layer

made of 10-undecylenic acid and oleic acid molecules. [ 70 ] In

contrast to the previous examples that are thermally acti-

vated, this strategy takes place at room temperature pre-

venting the MNPs from undesired side phenomena such as

size and morphological changes and crystalline-phase trans-

formations. Finally, anchoring of the MNPs on H-terminated

Si substrates can also be addressed with a more straightfor-

ward procedure by directly synthesizing the MNPs covered

with bifunctional molecules containing free vinyl terminal

groups. Following this strategy, vinyl-terminated CoFe 2 O 4

( ∼ 3 nm) and MnFe 2 O 4 (6–11 nm) NPs were directly synthe-

sized in non-polar solvents and then covalently assembled in

the form of well-ordered, tightly packed monolayer arrays on

H-terminated silicon surfaces (Figure 6 d,e). [ 69 , 71,72 ]

3.1.2. Pre-functionalization of the Surface

This goal has been successfully pursued by chemically mod-

ifying the surfaces with self-assembled monolayers (SAMs)

providing terminal head groups ready to interact with the

MNPs (see Figure 5 b). [ 67 , 73 ] For this purpose, the surface-

modifi ed substrate is incubated in a solution containing

the MNPs for a given time while the surfactant molecules

around the NPs are exchanged by the new functional

groups, followed by an extensive cleaning with solvent

that removes those NPs not directly attached to the head

groups. In this way, FePt NPs ( ∼ 3 nm) have been assembled

on a pre-modifi ed Si substrate using an amino-functional

silane, (3-(2-aminoethylamino)propyl) trimethoxysilane

(APTS), as a coupling layer. [ 74 ] The resulting FePt NP

fi lm was extremely robust and preserved its integrity after

in-situ annealing (up to 800 ° C and 10 − 8 Torr for 30 min)

with no apparent aggregation ( Figure 7 a,b). The annealing

process is needed to transform the as-synthesized NPs from

the chemically disordered fcc structure to an ordered face-

centered tetragonal (fct) structure (L1 0 phase) that exhibits

improved magnetic properties, such as larger coercivity,

required in applications such as high-density data storage

media, permanent magnetic nanocomposites, and biomag-

netic applications. [ 75 ]

This approach can also allow the formation of less closely

packed NP-decorated surfaces, if required. Mallah et al.

prepared a Si surface bearing cationic N 3 Ni(H 2 O) 3 groups

oriented out of the substrate, which interacted with the nega-

tively charged ∼ 6 nm CsNiCr NPs by means of electrostatic/

coordination bonds (Figure 7 c,d). [ 67 ] To reduce the surface

coverage, unreactive spacers with –CH 3 terminal groups were

intercalated among the anchoring functionalities. Therefore, a

rather large NP–NP distance was expected as a result of the

low concentration of anchoring groups on the surface. SQUID

7www.small-journal.com & Co. KGaA, Weinheim

E. Bellido et al.reviews

Figure 6 . Deposition of pre-functionalized MNPs with surface-reactive functional terminal groups. a) Schematic representation for the replacement of the oleic acid coating of ∼ 7 nm Fe 3 O 4 NPs by trimethoxy-7-octen-1-yl-silane, b) posterior immobilization of the NPs on an H-terminated Si substrate through reaction with the vinyl groups, and c) 3D AFM image of the resulting NP monolayer. Reproduced with permission. [ 66 ] d) Schematic illustration of the synthesis of ∼ 3 nm CoFe 2 O 4 NPs covered with vinyl-terminated groups ready to interact with H-terminated Si-substrates, and e) cross-section TEM image of the resulting NP monolayer. Reproduced with permission. [ 69 ] Copyright 2007, American Chemical Society.

characterization indicated that the structural and magnetic

integrity of the NPs were maintained upon grafting, being

consistent with average interparticle dipolar interactions sim-

ilar to those of the same MNPs diluted into a poly mer matrix.

Pichon et al. used a similar strategy to control the immobiliza-

tion of oleic-acid-coated ∼ 12 nm iron oxide NPs (composed of

Fe 3 O 4 / γ -Fe 2 O 3 ) on top of SAM-functionalized Au surfaces dis-

playing different functional groups (such as –COOH, –OH and

–CH 3 ). [ 73 ] They observed that SAMs decorated with –COOH

groups have the highest ability to interact with MNPs, while

non-polar terminal–CH 3 groups prevented the assembly.

3.1.3. Combined Functionalization of the MNP and the Substrate Surface

The chemical modifi cation of both features simultaneously

results in a more stable and specifi c MNP–substrate inter-

action (see Figure 5 c). Kinge et al. have recently employed

this strategy combined with click-chemistry. [ 76 ] The original

coating layer of as-synthesized ∼ 5 nm FePt NPs was replaced

by 5-hexanoic acid and 6-amino-1-hexyne groups to form

8 www.small-journal.com © 2012 Wiley-VCH Ve

alkyne-substituted FePt NPs ( Figure 8 a). Subsequently, a

SiO 2 surface is functionalized with azide-terminated SAMs

that react with the alkyne-functionalized NPs through the

formation of a triazole group. SEM analysis showed the suc-

cessful assembly of the MNPs in monolayers covering the

substrate (Figure 8 b). The nanoparticulated monolayer was

further annealed to transform the NPs from their as-synthe-

sized disordered fcc phase into the ordered L1 0 phase.

A different approach consists of the use of a functional-

ized polymer surface for the covalent immobilization of MNPs.

For instance, a polystyrene polymer matrix containing –COOH

functional groups was used for the carbodiimide-mediated

anchoring of γ -Fe 2 O 3 particles ( ∼ 200 nm) coated with an

amino-decorated polymeric shell. [ 77 ] This procedure led to a

single-layer of randomly distributed particles on top of the poly-

meric fi lm. Further arrangement of the magnetic particles into

ordered assemblies over large areas could be performed by the

application of an external magnetic fi eld parallel to the polymer

surface during the reaction process. Under these conditions, the

magnetic dipole of the particles aligned in the direction of the

applied fi eld and led to the formation of 1D chain-like ordered

rlag GmbH & Co. KGaA, Weinheim small 2012, DOI: 10.1002/smll.201101456

Magnetic Nanoparticles on Surfaces and Devices

Figure 7 . a) High-resolution SEM image of a ∼ 3 nm FePt NP monolayer formed over a surface functionalized with an SAM displaying terminal amino groups (bright dots correspond to the MNPs), and b) typical high-resolution SEM image of an FePt NP monolayer after thermal annealing process up to 800 ° C. Reproduced with permission. [ 74 ] Copyright 2003, American Institute of Physics. c) AFM image of a monolayer of ∼ 6 nm CsNiCr NPs formed on a surface functionalized with a SAM bearing cationic N 3 Ni(H 2 O) 3 and CH 3 terminal groups, and d) height profi le along the blue line drawn on (c). Reproduced with permission. [ 67 ] Copyright 2008, American Chemical Society.

Figure 8 . Assembly of ∼ 5 nm alkyne-substituted FePt NPs on an azide-terminated SAM using click-chemistry: a) schematic illustration of the procedure, and b) SEM image of the resulting FePt NP monolayer. Reproduced with permission. [ 76 ] Copyright 2011, American Chemical Society.

structures. The infl uence of an external magnetic fi eld on the

MNP assembly is further discussed in the next section.

3.2. Magnetic-Field-Assisted Assembly

The presence of an applied external magnetic fi eld allows

nanostructures to be grown in a specifi c directional manner,

either perpendicular or parallel to the substrate. [ 54 , 78 , 79 ] The

MNPs tend to align according to the direction of the applied

magnetic fi eld to minimize their magnetic energy, which is

strongly affected by the strength of the applied fi eld. [ 55 , 80 , 81 ]

The intensity of this force is directly proportional to a particle’s

magnetic moment and thus, to its magnetization properties.

On the other hand, magnetic moments feel a force under

magnetic fi eld gradients that tend to move the moment to

the region with higher magnetic fi eld density. So, in the pres-

ence of an external magnetic fi eld, the infl uence of the dipolar

© 2012 Wiley-VCH Verlag Gmbsmall 2012, DOI: 10.1002/smll.201101456

interactions in generating anisotropic chain-like structures on

the assembly process as presented in section 2.2 is perturbed,

since mainly, the external applied magnetic fi eld dominates

over the intrinsic anisotropic nature of the dipolar interactions

and allows for the formation of long-range ordered structures

due to the dominating interaction between each MNP and the

external magnetic fi eld. Such infl uence was carefully investi-

gated by Pileni et al., who immersed a HOPG substrate into a

solution containing ∼ 8 nm Co NPs (14% size distribution). [ 81 ]

Upon a slow solvent evaporation under an applied magnetic

fi eld perpendicular to the surface, well-defi ned 3D superlattices

were obtained ( Figure 9 ). At low magnetic fi elds (0.27 T), the

Co NPs assembled in homogeneous micrometer dots organized

in well-defi ned hexagonal networks. A slight enhancement of

the magnetic fi eld strength up to 0.45 T resulted in a reduction

of the size and interdistance between dots whereupon a fur-

ther increase up to 0.78 T induces a phase transition from dots

to worm-like and labyrinth structures. Interestingly, such struc-

tures could be observed over large areas of up to 0.02 mm 2 ,

with a fully covered thick fi lm of Co NPs lying underneath in

9www.small-journal.comH & Co. KGaA, Weinheim

E. Bellido et al.

10 www.small-journal.com © 2012 Wiley-VCH Verlag GmbH & Co

reviews

Figure 9 . SEM images of patterns obtained by immersing a HOPG substrate into an ∼ 8 nm Co NP solution under an applied magnetic fi eld perpendicular to the substrate surface until complete evaporation of the solvent. The strength of the applied fi eld was: a) 0 T, b–d) 0.27 T, e) 0.45 T, and f) 0.78 T. The magnifi cation of the hexagonal network obtained at 0.27 T is viewed: c) from above, and d) tilted. Reproduced with permission. [ 81 ]

Figure 10 . Assembly of MNPs under the infl uence of an applied magnetic fi eld parallel to the surface. TEM images of linear patterns made of ∼ 8 nm Co NPs obtained under: a) 0.27 T, and b) 0.56 T. Reproduced with permission. [ 18 ] c) Optical microscope image of needle-like structures made of cubic-shaped Fe x O NPs (11–13 nm in size) in the presence of a weak parallel magnetic fi eld. Reproduced with permission. [ 39 ] Copyright 2004, American Chemical Society. d) TEM images of cigar-shaped γ -Fe 2 O 3 NPs (length and width of ∼ 325 and ∼ 49 nm, respectively) deposited in the absence (d), and presence (e), of an external parallel magnetic fi eld (0.14 T). Reproduced with permission. [ 83 ] Copyright 2002, IOP Publishing Ltd.

all the cases. Pileni et al. later demonstrated that

the size distribution of the MNPs is key to deter-

mine the morphology of the 3D superlattices

obtained under both the absence and presence of

an applied magnetic fi eld. [ 82 ] Small distributions

(≤13%) of ∼ 6 nm Co NPs assembled into well-

defi ned columns (3.05 μ m in height and 1.7 μ m

in width) with a fcc structure when applying a

magnetic fi eld (0.25 T) perpendicular to the sur-

face during the evaporation process. By contrast,

an increase of the size distribution from 13% to

18% induced the formation of longer (3.8 μ m)

and thinner (0.7 μ m) columns with lower com-

pactness, which tended to aggregate in the form

of worm-like and labyrinth structures.

Application of magnetic fi elds parallel to

the substrate leads to longer linear patterns. For

Co NPs ( ∼ 8 nm), an increase of the strength of

the in-plane applied fi eld from 0.27 to 0.56 T

also infl uenced the 3D assembly, increasing the

order and compactness of the stripe structures

( Figure 10 a,b). [ 18 , 81 ] A similar behavior was

observed for ∼ 10 nm γ -Fe 2 O 3 NPs, [ 18 ] which

evolve from needle-like structures to homoge-

neous tube-like structures when increasing the

strength of a parallel external magnetic fi eld. The

shape of the MNP also plays an important role in

the 3D ordering. For instance, cubic-shaped wüs-

tite Fe x O NPs (11–13 nm in size) align parallel to

the substrate in the direction of the applied mag-

netic fi eld, forming needle-like structures made of

rectangular-shaped superlattices (Figure 10 c). [ 39 ]

Cigar-shaped γ -Fe 2 O 3 NPs (length and width of

∼ 325 and ∼ 49 nm, respectively) align with their

long axis along the applied fi eld direction and

form chain-like superlattices (Figure 10 d,e). [ 83 ]

In all these cases, the direction of the magnetic

fi eld plays an important role in the balance of the

forces. When it is applied perpendicular to the sub-

strate, there is a strong competition between the

magnetic energy F m = ∇( μ · H ) versus the gravita-

tional, F g = ( ρ MNP – ρ sol ) · V MNP · a , and the Brownian

motion energy, E B = k B · T , during the evaporation

process. On the other hand, when it is applied

parallel to the substrate, the main contribution

interactions are the torque of the MNP magnetic

moment due to the action of the external magnetic

fi eld, the dipole–dipole interactions modulated by

the external magnetic fi eld, and the MNP-substrate

interaction that can be reduced by modulating the

substrate temperature.

3.3. Lithography-Assisted Assembly

MNPs can be organized into nanostructured

motives on a substrate with lithographic tech-

niques following two main techniques, stamp-

based (soft-lithography) and lithographic

. KGaA, Weinheim small 2012, DOI: 10.1002/smll.201101456

Magnetic Nanoparticles on Surfaces and Devices

techniques. Moreover, each can be used following two dif-

ferent approaches: (i) direct-patterning of the MNPs, and

(ii) indirect-patterning deposition, where templates are used

to fi rst pattern a chemically modifi ed substrate that sub-

sequently guides the site-selective assembly of the MNPs (in

a similar way as described in section 3.1).

3.3.1. Stamp-Assisted Structuration (Soft-Lithographies)

The technique most widely used for the stamp-assisted struc-

turation of MNPs on surfaces is that of microcontact printing

( μ CP). This technique can be used for the direct patterning of

NPs on surfaces though most of the examples so far reported

rely on the indirect-patterning approach.

Indirect patterning . In general, in the indirect-patterning

approach fi rst the elastomeric stamp is wetted on a solu-

tion containing functional molecules and brought in contact

with the substrate to fabricate SAM patterns. Afterwards,

the MNPs are attached upon dipping the structured surfaces

( Figure 11 a). [ 24 , 25 , 84 ] For instance, Whitesides et al. applied

this process to the selective deposition of 5–12 nm Fe 3 O 4

NPs on Au substrates, which were previously patterned with

both hydrophobic and hydrophilic SAMs by μ CP. [ 24 ] By dip-

ping the patterned Au substrates into an aqueous solution

containing either positively or negatively charged Fe 3 O 4 NPs,

the selective wetting properties of the template directed the

deposition of particles only onto the hydrophilic patterned

areas with features of micrometer size and thickness between

50 and 100 nm. Further organization of MNPs into ring-like

© 2012 Wiley-VCH Verlag Gmb

Figure 11 . a) Schematic illustration of the indirect-patterning approach by means of μ CP. b) Optical image of a polystyrene-coated ∼ 10 nm Fe 3 O 4 NPs into ring-shaped structures. Reproduced with permission. [ 25 ] Copyright 2005, Elsevier. c) Schematic illustration of the direct-patterning approach by means of μ CP.

small 2012, DOI: 10.1002/smll.201101456

structures was also feasible by the fabrication of patterned

hydrophobic and hydrophilic SAMs using μ CP, followed by

the dewetting of a polystyrene-coated ∼ 10 nm Fe 3 O 4 NPs

solution on top of the pattern (Figure 11 b). [ 25 ] The patterned

surface was fi rst used to direct the condensation of water

droplets onto the hydrophilic areas after cooling below the

dew point. Then, a liquid layer of chloroform containing the

NPs was deposited on top of the water-patterned surface

by dip-coating, which after the complete evaporation of the

solvents resulted in an ordered pattern of magnetic rings of

approximately 10 μ m in diameter and 450 nm in height.

μ CP technique is also suitable for the fabrication of SAM

templates that induce the selective attachment of the MNPs

by means of ligand-place-exchange reactions, leaving highly

stable monolayers of MNPs. For example, Al 2 O 3 substrates

were patterned using aminobutylphosphonic acid (ABP)

SAMs, whose NH 2 -terminated groups are known to effectively

interact with FePt NPs by ligand-place-exchange reactions. [ 84 ]

After dip-coating the template into the colloidal solution,

SEM images revealed the selective formation of monolayers

of ∼ 10 nm FePt NPs on top of NH 2 -terminated micrometric

areas while leaving all the non-patterned areas uncovered.

Direct patterning . By combining both μ CP and click-

chemistry, Kinge et al. directly generated patterns of highly

packed ∼ 5 nm FePt NPs monolayers with an elastomeric

stamp that was directly immersed into a solution containing

the MNPs (Figure 11 c). [ 76 ] The stamp can also been func-

tionalized with a LB fi lm, technique known as patterned

Langmuir– Blodgett (pLB) ( Figure 12 a). [ 85–87 ] Following this

11www.small-journal.comH & Co. KGaA, Weinheim

Figure 12 . a) Schematic illustration of pLB procedure followed by additional thermal treatment. SEM images of the obtained FePt microstructures after thermal annealing: b) lines (1 μ m wide), c) squares (2 μ m wide), and d) meshes (10 μ m dot diameter). Reproduced with permission. [ 86 ]

E. Bellido et al.

1

reviews

Figure 13 . Pattern fabrication by means of MIMIC. a) Schematic illustration of the procedure. Details of the last steps of the process when using: b) high, or c) low colloidal solution concentration. AFM image of patterns of ∼ 13 nm Fe 3 O 4 NPs monolayers showing the variations in the morphology depending on the solution volume inside the micrometer-channels, which are: d) full (100%), and e) ∼ 25% of the volume. Reproduced with permission. [ 89 ] Copyright 2008, IOP Publishing Ltd.

approach Yang et al. prepared LB fi lms on top of a stamp of

γ -Fe 2 O 3 NPs ( ∼ 13 nm) [ 85 ] or core–shell ∼ 4 nm Pt@Fe 2 O 3 NPs

patterns [ 86 ] by pLB technique that were later transferred

to silicon or other substrates (Figure 12 b). Uniform MNP

patterns with different shapes could be printed including

micrometer size lines, squares and meshes, with heights of

40 ± 3 nm corresponding to six MNPs (considering the length

of the coating layer and the NP diameter). These MNP pat-

terns were then converted into L1 0 phase FePt fi lms upon

thermal annealing, after which the original pattern design

was left unmodifi ed.

An important advantage of the pLB technique is that it

ensures a thickness control of the fi lm due to the formation

of single-layered MNP structures. [ 87 , 88 ] For instance, the for-

mation of patterned monolayers of ∼ 7 nm FePt and ∼ 25 nm

FePt@SiO 2 NPs organized in highly ordered hexagonal close-

packed assemblies was reported by Korgel and co-workers. [ 88 ]

For this, LB fi lms made of MNP monolayers were transferred

to elastomeric stamps patterned with dots, squares and lines

of 1.5 to 20 μ m in size and stamped onto a silicon substrate.

Moreover, the fabrication of strictly controlled multilayer

2 www.small-journal.com © 2012 Wiley-VCH Verlag GmbH & Co. KGaA,

structures of NPs was possible by either

repeated wetting of the stamp with the LB

layer or by repeated printing of the stamp

on the substrate.

Finally, a variation of the stamp-

assisted methodology named as micro-

injection molding in capillaries (MIMIC)

was used by Cavallini et al. to pattern

∼ 13 nm Fe 3 O 4 NPs in micrometric stripes

and dots on mica ( Figure 13 ). [ 89 ] First, an

elastomeric stamp is put in contact with a

solid surface forming a sub-micrometric

channel network at the interface. Then, a

solution containing the MNPs is depos-

ited at the open end of the stamp spon-

taneously fi lling the micro meter-channels

by capillary forces. The MIMIC method

allowed the formation of spatially well-

organized microstripes whose thickness

corresponded to a single layer of Fe 3 O 4

NPs. Moreover the morphology of the

fi nal patterns could be controlled by the

MNP concentration and the type of sol-

vent (Figure 13 b–d).

3.3.2. Lithography-Assisted Structuration

This approach involves mostly the use

of electron beam lithography (EBL) and

photolitho graphy for the fabrication of

physical templates that can be used to

structure the MNPs. The topographi-

cally structured surfaces physically guide

the assembly of MNPs to specifi c areas,

such as pores on a substrate, without

the need of specifi c chemistry. [ 26 , 90, 91 ]

For instance, Chen et al. have recently

reported the use of the Damascene

process for the controlled positioning of spherical-shape

∼ 18 nm core–shell FeO@CoFe 2 O 4 and cubic-shaped ∼ 50 nm

Fe 3 O 4 NPs ( Figure 14 ). [ 26 ] The Si surface was pre-patterned

by means of either EBL or via charged-particle nanopat-

terning (CHARPAN) to create round- or square-shaped

nano metric pits, respectively. The MNPs were then depos-

ited on the template by drop-casting and, after complete

evaporation of the solvent, organized along the surface in

a close-packed assembly. A mechanical polishing process

was subsequently applied in order to remove the MNPs

from the fl at parts of the substrate, leaving only those

MNPs already placed at the lithographically fabricated

motives. Interestingly, this method successfully controlled

the number of MNPs down to a single NP per unit pit by

accurately controlling the pit and the MNP size, as well as

the MNP concentration. Therefore, pit diameters of 90 nm

yielded a hexagonal arrangement of 7 closely packed NPs,

while for pit diameters of 25 nm, single NP occupation was

successfully obtained (Figure 14 b–d).

An alternative approach is the use of masks that guide the

self-assembly of the MNPs exclusively onto the bare areas of

Weinheim small 2012, DOI: 10.1002/smll.201101456

Magnetic Nanoparticles on Surfaces and Devices

Figure 14 . Patterns of ∼ 18 nm FeO/CoFe 2 O 4 core–shell and ∼ 50 nm Fe 3 O 4 NPs fabricated by means of the Damascene process. a) Large-area SEM image of an array with 110 nm pit diameter fi lled with NPs. b–d) SEM images showing the control over the number of MNPs per pit down to the single NP level, simply by adjusting the pit diameter from 90 to 25 nm. Reproduced with permission. [ 26 ]

the substrate, while the lift-off of the mask removes the rest of

the MNPs. The masks are prepared by spin-coating a polymer

on a surface and posterior pattern with the desired motive by

means of photolithography. [ 92 , 93 ] Using this strategy, 3–4 nm

FePt NPs were covalently anchored on a H-terminated sil-

icon surface and patterned into ordered arrays of 1 μ m 2 dots

with heights of 15–30 nm, which corresponded to 2–4 layers

of NPs (considering the surfactant coating of the MNP was

∼ 2 nm). [ 92 ]

3.4. AFM-Based Lithographies

AFM-based lithography techniques have attracted much

interest because of their exceptional capabilities for

© 2012 Wiley-VCH Verlag Gmbsmall 2012, DOI: 10.1002/smll.201101456

patterning MNPs onto a given substrate with nanometric

resolution. Interesting developments have been implemented

in this direction including the use of local oxidation nano-

lithography (LON), nanoembedding (n-EM), dip-pen nano-

lithography (DPN), and thermal dip-pen nanolithography

(t-DPN). Moreover, as previously described for the litho-

graphic techniques, each of them can be used following two

different approaches: i) direct-patterning of the MNPs, and

ii) indirect-patterning deposition, where templates are used

to fi rst pattern a chemically modifi ed substrate that subse-

quently guides the site-selective assembly of the MNPs. The

examples are reviewed next.

3.4.1. Local Oxidation Nanolithography (LON)

LON, also known as local anodic oxidation, [ 94 ] consists

of a bias-assisted AFM nanolithographic technique used

for the fabrication of nanopatterned SAM templates that

afterwards are used to induce the assembly of MNPs. By

applying a suitable voltage to a SAM through a conduc-

tive tip, an electrochemical reaction can be induced exclu-

sively at the terminal functionalities of the SAM while fully

preserving the structural integrity of the whole layer. For

example, terminal CH 3 -groups of N -octadecyltrichlorosi-

lane (OTS) monolayers self-assembled on a Si substrate

were fi rst locally converted into –COOH groups by elec-

trochemical AFM nanolithography and then used as a tem-

plate for the selective attachment of monolayers of MNPs

such as γ -Fe 2 O 3 , Fe 3 O 4 , and FePt ( Figure 15 ). [ 95–97 ] High-

resolution MNP patterns consisting of stripes or arrays of

dots with feature sizes of sub-100 nm were achieved by this

technique. [ 97 ]

3.4.2. Nanoembedding (n-EM)

The n-EM approach reported by Cavallini et al., based

on the bias-assisted nanolithography technique described

above, allows the spatially controlled integration of NPs

inside silicon surfaces. In a fi rst step the deposition of the

particles takes place by techniques such as spin-coating or

lithographically controlled wetting (LCW). Subsequently

the MNPs are embedded by a conductive tip that applies a

bias voltage to the silicon surface, inducing the growth of a

local SiO 2 coating layer while preserving their morpholog-

ical and magnetic properties. The use of this technique has

been restricted so far for the nanoembedding of CoFe 2 O 4

NPs ( ∼ 6 nm) on silicon substrates. [ 98 ] Moreover, replace-

ment of the tip by a conductive stamp permitted the struc-

turation of MNPs into parallel stripes or grids (width of 200

nm) applied along vast areas of 1 cm × 1 cm. The number of

embedded NPs depended on the MNP concentration of the

initial solution.

3.4.3. Dip-Pen Nanolithography (DPN)

The two techniques previously described follow the indirect-

patterning method, that relies fi rst on the fabrication of tem-

plates to later on assist the assembly of MNPs. In addition

to this, DPN is an AFM-based nanolithographic technique

13www.small-journal.comH & Co. KGaA, Weinheim

E. Bellido et al.

14

reviews

Figure 15 . Pattern of ∼ 9 nm γ -Fe 2 O 3 fabricated by means of bias-assisted AFM nanolithography. a) Schematic illustration of the procedure. b) AFM image of the magnetic pattern, whose height scale is 100 nm, and c) height profi le along the white dashed line in (b). Reproduced with permission. [ 95 ] Copyright 2007, American Chemical Society.

that can also allow direct-writing to directly deliver materials

(ink) to nanoscopic regions of a substrate simply by using the

tip as a nanoscale pen ( Figure 16 a). [ 99–101 ] The possibility to

directly write MNPs at specifi c positions of a substrate in a

www.small-journal.com © 2012 Wiley-VCH Verlag GmbH & Co. KGaA

single-step, without the need of pre-func-

tionalization steps of either the substrate

or the NP surface, simplifi es enormously

the structuration process. It is also impor-

tant to remark that this technique can

virtually be used on any type of substrate

allowing control of the feature shape, dot

size, and line width by simply adjusting

the tip movement over the surface, the tip-

substrate contact time, and the scan speed,

respectively.

For example, after coating a tip by

immersion into a colloidal dispersion of

citrate-capped ∼ 11 nm γ -Fe 2 O 3 NPs and

drying, Gundiah et al. fabricated MNPs

nanostructures by scanning, in contact

mode, a mica or Si substrate (with and

without the native layer of oxide). [ 103 ]

Although this method achieved lines

ranging from 140 to 200 nm and lengths

extending to 10 μ m, the major drawback

was its extremely slow patterning process,

which needed scan speeds of ∼ 1 μ m · s − 1 for

a period of 30 min. By contrast, Li et al.

reported a DPN process for MNP nano-

structuration over large areas on a Si/SiO 2

or quartz substrate at shorter time periods

(e.g., dots of 381 nm in diameter could be

generated by holding the coated tip in con-

tact with a surface for 0.8 s). [ 104 ] Co NPs

( ∼ 5 nm) capped with oleic acid and tri-

octyl phosphine oxide (TOPO) were kept

on the tip in a fl uid-like state in order to

ensure an effi cient material transfer from

the tip to a substrate through a meniscus

composed of the ink material itself. By

controlling the contact time of the tip over

the substrate, lines of 880 nm in width and

dots down to 68 nm in diameter composed

of 2–3 layers of MNPs were generated by

DPN. Additionally, the patterned Co NPs

served as catalysts to guide the site-selec-

tive growth of well-aligned single-walled

carbon nanotube (SWCNTs) arrays on

quartz substrates by chemical vapor depo-

sition (CVD).

Though the use of DPN for the

direct-writing has the advantages previ-

ously described, DPN can also be used to

assist the indirect writing and assembly of

MNPs on surfaces. [ 27 , 28 , 105 ] One possibility

is to fabricate templates/motives made of

SAMs bearing COOH-terminal groups

on a Au surface while the rest of the sur-

face is passivated with an unreactive layer

of CH 3 -terminated SAM. [ 27 ] For this, alkanethiol molecules

were fi rst coated and dried onto the tip and then brought

in contact with the Au substrate. The resulting template was

then dip-coated into a basic solution containing either Fe 3 O 4

, Weinheim small 2012, DOI: 10.1002/smll.201101456

Magnetic Nanoparticles on Surfaces and Devices

Figure 16 . a) Schematic illustration of the DPN process. Reproduced with permission. [ 102 ] Copyright 2004, Elsevier. b) AFM image of an array made of lines accommodating individual Fe@C NPs and, c) a higher magnifi cation of the same array. Reproduced with permission. [ 28 ] Copyright 2008, American Chemical Society.

or MnFe 2 O 4 NPs (both of ∼ 10 nm) coated with positively

charged tetramethylammonium ligands in order to favor the

electrostatic interaction with the negatively charged COO − -

pattern. Single-layer MNPs structures of dots and stripes—

with high control of the feature size (from few micrometers

to 45 nm)—and homogeneous arrays of 21 × 22 dots with a

fi xed dot diameter (85 ± 5 nm) were achieved. Further studies

with the DPN deposition method were directed towards

accommodating single MNPs per feature of a nanoarray. [ 28 ]

This was accomplished by using a solvent that selectively

wets the carboxylic pattern but not the methyl-passivated

regions. The effect of the template feature size on the MNP

assembly was studied with dots of sizes in the range of 50 to

120 nm, and it was found that feature sizes of 60 nm were

optimal to generate single Fe@C NP ( ∼ 5 nm) per feature,

which demonstrates the high control on the deposition at a

single particle level (Figure 16 b,c).

3.4.4. Thermal Dip-Pen Nanolithography (t-DPN)

Finally, t-DPN is a variation of the DPN technique that

makes use of a cantilever with integrated resistive heaters.

Initially the ink deposited on the tip is solid at room tem-

perature, but upon heating the melted ink fl ows from the

tip to the surface. [ 106 ] Depending on the scan speed and the

tip temperature, the deposition can be switched on and off

© 2012 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheimsmall 2012, DOI: 10.1002/smll.201101456

and the feature sizes controlled to sub-

100 nm resolution. Interestingly, this

technique allows both the deposition of

nanocomposites composed by a wide

range of polymers containing either NPs

or molecules as well as non-specifi c sub-

strate deposition. In particular, Sheehan

et al. reported the use of t-DPN for the

nanostructuration of MNPs by coating a

tip with a composite ink of ∼ 7 nm Fe 3 O 4

NPs and poly(methyl methacrylate)

(PMMA). [ 107 ] In this way lines between

78 and 400 nm in width, by controlling

the tip temperature and scan speed, were

fabricated ( Figure 17 ). Furthermore,

the poly mer matrix can be successfully

removed from the pattern upon exposure

to oxygen plasma treatment, which drasti-

cally reduced the width of the lines down

to just 10 nm well-defi ned single NP rows.

4. Use of Biomolecular Templates

The use of biomimetic templates pro-

vides a new avenue for the synthesis

and assembly of NPs. Special attention

has been paid to the use of protein cage

supramolecular structures, such as fer-

ritin and viral capsids. [ 108 ] Their specifi c

size and highly reproducible biosynthesis

make them ideal templates for accommo-

dating a variety of inorganic NPs with controlled size, shape,

and crystal structure in their inner cavities, a process com-

monly known as biomineralization. [ 50 , 109 , 110 ] For example,

tobacco mosaic virus (TMV) hollow cylinders have been

used as biotemplates for the controlled formation and

organization of Pt, Au, Ag, FePt, CoPt, and FePt 3 NPs. [ 111–113 ]

However, among the great variety of cage-shaped protein

molecules, the most extensively studied and used as a tem-

plate is ferritin, mainly from horse spleen (HsAFr), [ 109 , 112 ]

though other ferritin-like protein such as that of the bacteria

Listeria innocua are also ideally suited for the synthesis of

NPs, like γ -Fe 2 O 3 and Co 3 O 4 NPs. [ 114 , 115 ] Ferritin is a roughly

spherical protein consisting of 24 protein subunits that self-

assemble into a cage-like architecture with a diameter of

∼ 12 nm and an internal cavity of ∼ 7 nm. Inside the cavity,

this protein naturally stores up to 4500 Fe atoms predomi-

nantly as a nanomineral of ferrihydrite ( Figure 18 ). [ 116 ] By

suitable chemical processes, it is possible to remove the

internal inorganic core and use the empty protein shell,

named apoferritin, as a confi ned environment in which dif-

ferent inorganic NPs possessing potential for magnetic,

catalytic, and biomedical applications can be artifi cially syn-

thesized. [ 117 ] This approach has been successfully used in

the synthesis of a wide variety of metal oxide NPs, such as

Fe 3 O 4 / γ -Fe 2 O 3 (known as magnetoferritin), [ 118 , 119 ] Co 3 O 4 , [ 120 ]

as well as Co, Mn, Eu, Fe, and Ti oxyhydroxide NPs. [ 121–123 ]

15www.small-journal.com

E. Bellido et al.

16 www.small-journal.com © 2012 Wiley-VCH V

reviews

Figure 17 . a) Schematic illustration of the fabrication of MNP-based motives by means t-DPN. After deposition of the composite ink (made of ∼ 7 nm Fe 3 O 4 NPs and PMMA); the polymer can be removed by oxygen plasma treatment. AFM and SEM images of the fabricated pattern before (b), and after (c, d) the elimination of the polymer, resulting in single MNP lines. Reproduced with permission. [ 107 ] Copyright 2010, American Chemical Society.

Figure 18 . Schematic illustration of the structure of ferritin molecule. Reproduced with permission. [ 130 ] Copyright 2010, Elsevier.

Metallic and semiconductor NPs have also been synthesized,

such as Cu, [ 124 ] Pd, [ 125 ] Ni, [ 126 ] Co, [ 126 ] CoPt, [ 127 ] CdSe, [ 128 ] and

ZnSe. [ 129 ]

Interestingly, the external protein shell can be modi-

fied to guide the assembly of ferritin on a given substrate

while preserving all the inner characteristics. This prop-

erty allows the use of ferritin as a vehicle (template) for

guiding and positioning a large variety of internal inor-

ganic NPs to desired positions on a surface by using in

all the cases the same experimental conditions. Indeed,

since proteins are more easily decomposed than inor-

ganic materials, once structured on a surface the organic

protein shell can be selectively eliminated by means of

heat or UV/ozone treatment, while the inorganic mate-

rial preserves its original integrity and location on the

surface. [ 131 , 132 ] In view of these unique properties, ferritin

emerges as a fascinating scaffold for the nanostructu-

ration of MNPs on a substrate. Representative examples

for the structuration of these materials both assisted and

un-assisted are considered next. The representative exam-

ples have been divided following the same experimental

sections as those previously described without biological

templates.

4.1. Un-assisted Self-Assembly

To date, several strategies to organize ferritin mol-

ecules into well-ordered mono- and multi-layers,

including the transference of a ferritin fi lm from an air–

liquid interface onto a solid support [ 131 , 132 ] as well as the use

of scratching [ 133 ] and spread-coating [ 134 ] methods, have been

described. Yamashita et al. reported a method called Bio

Nano Process for the fabrication of 2D FeO MNP mono-

layers on substrates. [ 131 , 132 ] In this process, the ferritin mole-

cules fi rst self-assemble into well-ordered 2D hexagonal

close-packed arrays by means of a charged polypeptide

layer spread at an air–liquid interface. Then, the MNP fi lm

was transferred onto a hydrophobic Si surface (pre-treated

with 1,1,1,3,3,3-hexamethyldisilazane) by laying the sub-

strate on the air–liquid interface ( Figure 19 a). Finally, the

erlag GmbH & Co. KGaA, Weinheim

Figure 19 . High-resolution SEM images of a 2D array of ferritin on a Si substrate before (a) and after (b) heat treatment at 450 ° C under nitrogen fabricated by means of the Bio Nano Process method. White spots correspond to the iron oxide cores of ferritin molecules. b) A tilted high-resolution SEM image where secondary electron intensity has been transformed into a height profi le to form a pseudo 3D image, showing the achieved high-order of the 2D array of FeO NPs. Reproduced with permission. [ 132 ] Copyright 2001, Elsevier.

small 2012, DOI: 10.1002/smll.201101456

Magnetic Nanoparticles on Surfaces and Devices

Figure 20 . a) Schematic illustration of the fabrication of 2D arrays of ferritin molecules which were genetically modifi ed in order to selectively promote their adsorption onto the droplet liquid surface. The nucleation takes place in a controlled way along the contact line at the liquid surface. b) SEM image of the single-domain 2D array where the white spots represent the ferritin cores and the dark spots are considered to be apoferritin molecules, invisible in SEM. Inset corresponds to an FFT image of the SEM image confi rming the hexagonally close-packed arrangement of the ferritin molecules. Reproduced with permission. [ 140 ] Copyright 2008, American Chemical Society.

protein shell was selectively removed by subjecting the

sample to either heat treatment (450 ° C under nitrogen

for 1 h) or ozone treatment, while preserving the order of

the array; thus leaving inorganic cores with diameters less

than 6 nm (Figure 19 b). XPS analysis after heat treatment

indicated that the original core was transformed into FeO

mixed with a small proportion of Fe 2 N.

4.2. Assisted Self-Assembly and Organization

4.2.1. Chemically Assisted Self-Assembly

The interaction of the ferritin with the substrate can be tai-

lored through specifi c covalent or electrostatic bonds to

induce the formation of more stable ferritin networks. Sev-

eral strategies have been developed in this direction, such

as modifi cation of the substrate surface with SAMs bearing

terminal functionalities (i.e., succinimidyl groups) suitable

to covalently interact with the protein shell. [ 135–137 ] Alterna-

tively, functional groups such as thiol can be introduced to

the protein shell to facilitate the attachment of the protein

on bare Au substrates, [ 135 , 138 ] or the shell can be genetically

modifi ed with target-specifi c affi nity peptides to induce a

specifi c interaction of the modifi ed ferritin with the target

inorganic surface. [ 139 , 140 ] For example, genetically modifi ed

ferritin molecules, displaying at the outer surface peptide

aptamers against carbonaceous materials (named NHBP-1)

(with artifi cially synthesized Fe and indium oxide cores with

sizes of ∼ 6 nm and ∼ 7 nm, respectively), were used to fab-

ricate 2D hexagonally close-packed assemblies of ferritin

with a typical domain size of 100 nm 2 directly on a carbon

fi lm surface. [ 139 ] Furthermore, the same NHBP-1 peptide

was used for inducing low affi nity between ferritin and a

SiO 2 hydrophilic surface in order to promote the adsorp-

tion of ferritin mole cules on the liquid surface of a droplet

instead of adsorbing on the substrate surface. [ 140 ] The method

reported by Ikezoe et al., and schematically depicted in

Figure 20 a, succeeded in fabricating well-ordered dense

2D assembly domains with large sizes of approximately

100 μ m 2 , where the ferritins encapsulated indium oxide NPs

inside their cavities (Figure 20 b).

A completely different strategy takes advantage of the

electrostatic interaction between ferritin and a given sub-

strate. [ 141 , 142 ] In particular, the charge of the protein shell can

be easily manipulated by changing the pH of the solution

media. Since the isoelectric point of ferritin is ∼ 4.5, at lower

pH values it becomes positively charged, while at larger pH

values the ferritin is negatively charged. For example, fer-

ritin was immobilized onto a 3-aminopropyltrimethoxysilane

(3-APTMS)-modifi ed Si substrate by electrostatic interactions

between the negatively charged aminoacids of the ferritin

and the amino-terminal functional groups of 3-APTMS. [ 143 ]

The protein shell was then selectively removed by heat treat-

ment (400 ° C for 1 h) leading to 2D NP assemblies derived

from ferritin cores on the Si substrate. The MNP array was

composed of FeOOH iron oxide NPs of ∼ 5 nm, as deter-

mined by AFM measurements. Interestingly, in situ reaction

of iron chelators with the ferritin molecules immobilized on

© 2012 Wiley-VCH Verlag Gmbsmall 2012, DOI: 10.1002/smll.201101456

the substrate before heat treatment allowed tuning of the size

of the NPs in the range of 0–5 nm.

Yamashita and co-workers also realized that to organize

the ferritin molecules into high-density monolayers, it was

necessary to control, not only the sign of the charge, but

also the quantitative balance of the surface charge on both

the protein and substrate. [ 144 ] To this end, they made use of a

genetically modifi ed ferritin with a moderate negative charge

at neutral pH, which was able to densely adsorb as a mono-

layer onto a highly positively charged layer of 3-aminopropyl-

triethoxysilane (APTES) layer supported on a Si substrate.

4.2.2. Lithography and Stamp-Assisted Self-Assembly

Several patterning techniques have been implemented,

such as nanosphere lithography (NSL), [ 145 ] microlens array

(MA) patterning, [ 146 ] block copolymer (BCP) micelle litho-

graphy, [ 147 ] electron beam litho graphy (EBL), [ 148 ] biomi-

metic layer-by-layer assembly (BioLBL), [ 149 ] and AFM-based

lithographies, [ 150 , 151 ] among others. The great majority of

the developed approaches are based on indirect-patterning

17www.small-journal.comH & Co. KGaA, Weinheim

E. Bellido et al.

18

reviews

Figure 21 . a) SEM image of Ti pattern on SiO 2 fabricated by means of NSL and EBL technique. Inset shows an AFM image of the pattern. b) SEM image after site-selective adsorption of ferritin (displaying Ti-binding peptides on their surface). The maximum number of adsorbed ferritins per Ti pattern was four. Three imaginary polystyrene particles used for the Ti patterning by NSL technique were drawn. Reproduced with permission. [ 153 ]

deposition techniques, which fi rst fabricate pre-modifi ed

chemical patterns as templates that guide the posterior site-

selective immobilization of the protein onto pre-defi ned areas.

Following this strategy, well-ordered ferritin arrays have been

successfully fabricated bearing few or even a single ferritin

molecule immobilized on each feature of the array.

A genetically modifi ed ferritin with target-specifi c affi nity

peptides was used to guide the immobilization of the modi-

fi ed protein selectively onto lithographically pre-patterned

areas (i.e., Au [ 152 ] or Ti [ 149 , 153 , 154 ] ). For example, Yamashita

et al. made use of a modifi ed ferritin displaying a peptide

with a high affi nity for Ti, named TBP-1, which selectively

adsorbs onto a hexagonal pattern made of triangular 40 nm

Ti dots. [ 153 , 154 ] The Ti template was fabricated on a SiO 2 sur-

face using NSL and EBL techniques, using a 2D hexa gonally

close-packed array of poly styrene (150 nm) particles as a

mask for the deposition of the Ti fi lm ( Figure 21 a). This

was then immersed in a ferritin solution for a given time

at room temperature followed by intensive washing and

drying (Figure 21 b). After ferritin immobilization, the pro-

tein shell could be selectively removed by UV/ozone treat-

ment (at 110 ° C for 40 min) while preserving the hexagonal

periodic array, leading to well-ordered NP arrays on the Ti

pattern. The specifi c binding of the modifi ed protein was not

affected by the inner NPs, and therefore, the procedure could

be reproduced for ferritin displaying two different NPs syn-

thesized inside its cavity, ferrihydrite and Co 3 O 4 cores. More

recently, Sano et al. made use of the peptide (TBP-1) to

genetically modify the ferritin molecules and fabricate mul-

tilayered structures selectively adsorbed on nanopatterned Ti

stripes fabricated by EBL. [ 149 ] This method, called BioLBL,

employed the target-specifi c affi nity peptide not only as

a binding molecule to immobilize the ferritins but also as

a mediator for the biomineralization of a thin fi lm of the

target material (i.e. TiO 2 ) from a precursor reagent present

in solution. Moreover, the TiO 2 fi lm could be used in turn

as a target for the adsorption of a second ferritin layer. By

repeating the process, multilayered structures with different

NP-containing ferritins (fi lled with Fe and Co oxides) depos-

ited between TiO 2 layers were selectively built on top of the

www.small-journal.com © 2012 Wiley-VCH Ve

original strip pattern and characterized by means of SEM and

energy dispersive spectro scopy (EDS) mapping.

In further studies, Yamashita and co-workers used elec-

trostatic interactions as a driving force to make ferritins

selectively adsorb individually onto specifi c regions of a lith-

ographically pre-patterned substrate. [ 155 , 156 ] They fabricated

a pattern made of circular APTES disks with two different

sizes ( ∼ 15 and ∼ 45 nm) on a SiO 2 substrate by means of EBL,

vapor-phase APTES deposition and lift-off processes. Ferritin

was fi rst genetically modifi ed to increase its surface charge

density and, therefore, enhance the electrostatic interaction

with APTES disks while avoiding secondary ferritin adsorp-

tion ( Figure 22 a). The patterned substrate was then incubated

in the ferritin solution for a given time at room temperature

and studied by means of SEM. The suitable tailoring of the

electrostatic interactions (supported by numerical calcula-

tion) allowed the placement of a single ferritin molecule on

top of each APTES disk of the array, whose disk diameter

(up to 45 nm) was much larger than the size of a ferritin

( ∼ 12 nm) (Figure 22 b). Additionally, the obtained single fer-

ritin array on 15 nm APTES disks was subjected to rapid

thermal annealing (RTA) under O 2 gas at 500 ° C for 10 min

in order to selectively eliminate not only the protein shell but

also the APTES pattern; thus leaving independent iron oxide

NPs located on the designated positions of the array. Interest-

ingly, the obtained periodic single NP array has recently been

used as a catalyst for the position-controlled vertical growth

of individual carbon nanotubes (CNTs). [ 157 ]

4.2.3. AFM-Assisted Self-Assembly

Most of the examples described up to know are based on the

indirect writing of ferritins on tip-prefunctionalized areas.

LON [ 94 ] has been used for the fabrication of oxide dots and

stripes of only few nanometers that can serve as templates

for the selective attachment of ferritin molecules. [ 30 , 158 , 159 ]

This method applies a suitable bias voltage to a conduc-

tive tip, which makes the water meniscus formed in the tip-

substrate gap dissociate and the oxidative ions to react with

the substrate, leaving localized oxide nanostructures with a

rlag GmbH & Co. KGaA, Weinheim small 2012, DOI: 10.1002/smll.201101456

Magnetic Nanoparticles on Surfaces and Devices

Figure 22 . a) Schematic illustration of the adsorption of a single ferritin molecule onto an APTES disk, where the repulsive fi elds between the substrate surface and the fi rst adsorbed molecule avoids the approach of a second ferritin. b) SEM image of an APTES array showing the adsorption of single ferritin molecules on each disk. Reproduced with permission. [ 156 ] Copyright 2009, American Institute of Physics.

resolution down to 10 nm. Yoshinobu et al., explored two

different protein patterning routes based on the LON tech-

nique. [ 158 ] In the case of the negative patterning, a Si substrate

was modifi ed with positively charged APTES molecules

ready to electrostatically interact with the negatively charged

ferritins. Then, oxide patterns were formed on this surface by

LON and subsequently the ferritins were selectively immo-

bilized outside the oxide patterns. In the case of the positive

patterning, the Si substrate was fi rst protected with OTS to

introduce unreactive long alkyl groups on the surface. Then,

LON was applied to fabricate oxide patterns which in turn

were modifi ed with APTES to promote the immobilization

of the ferritin molecules. Interestingly, the use of electrostatic

interactions presents additional advantages such as control-

ling the attachment/removal of this protein on the surface by

changing the pH of the solution. [ 30 , 159 ] The deposition at the

single particle level reported by Coronado and co-workers

was possible due to the high spatial resolution of LON, which

allows accurate matching of the size of the oxide pattern to

the protein size ( ∼ 12 nm). [ 30 ] Thus, individual ferritin mol-

ecules could be positioned into chain-like structures over

sub-micrometer areas ( Figure 23 a). By means of print-based

methods such as lithography-controlled dewetting instead

of LON, this approach could be successfully extended to

pattern ferritin particles over macroscopic regions (1 cm 2 )

(Figure 23 b).

Recently, our group has reported the deposition of fer-

ritin-based NPs using DPN technique via the direct-pat-

terning approach, which avoids previous functionalization

of either the substrate or the molecule, and does not require

the application of external stimuli. [ 160 ] Ferritin nanostruc-

tures were fabricated by coating a tip with an aqueous fer-

ritin solution (ink). Afterwards the tip was brought in contact

with the substrate surface to dispensie femtoliter volumes of

the protein solution with control of the shape and size, mini-

aturizing the dot features down to 100 nm. Moreover a wide

variety of substrates of technological interest such as Au, Si/

SiO 2 , Nb, and Al have been successfully used ( Figure 24 a).

More interesting is the fact that ferritin molecules deposited

on each dot were organized as sub-monolayers. After ferritin

deposition, the ferritin patterns were subjected to heat treat-

ment (450 ° C for 1 h in an inert atmosphere) to selectively

© 2012 Wiley-VCH Verlag GmbHsmall 2012, DOI: 10.1002/smll.201101456

remove the organic protein shell. SEM characterization of the

heated array revealed that the order, dimensions, and spacing

of the DPN-generated dot-like nanostructures, as well as the

NP location, were maintained after heating. The presence of

iron oxide NPs on these dot-like features and the complete

elimination of the protein portion were confi rmed by time-of-

fl ight secondary-ion mass spectrometry (TOF-SIMS) analysis.

In addition, the direct-write capabilities of DPN provides the

potential not only to fabricate miniaturized arrays but also to

have fi ne control of the number of ferritins deposited at spe-

cifi c locations on a surface. We have demonstrated by TEM

analyses that this could be afforded by calculating the contact

angle between the ink ferritin solution and the substrate, and

controlling the initial protein concentration and the dot-like

feature diameter (Figure 24 b). [ 29 ]

5. Integration into Hybrid Devices

Up to now, we have reviewed the structuration of MNPs on

surfaces from 3D down to 1D organization. A new aspect of

the structuration fi eld must be addressed when site-selectivity

with control at the nano scale level is required. This is the case

for the integration of single or a few nanoparticles into nano-

devices whose objectives rely on either, using the MNPs as a

subject of study to understand the magnetization relaxation

at the individual NP level, like in nanoHall sensors or nano-

SQUIDs, or by using the single MNPs as analysers, polarizers

or fi lters, as for example in spin valve nanodevices. In this

section we will review the different approaches followed to

integrate MNPs on different sensors and briefl y overview the

magnetic and transport measurements that can be derived

from them.

5.1. Randomized Integration into Devices

The simplest method initially used to integrate MNPs into

nanoscale devices consists of the random deposition of the

sample by using some of the deposition techniques previ-

ously described in section 2.1. Representative examples are

drop-casting, by placing a drop of the solution with dispersed

19www.small-journal.com & Co. KGaA, Weinheim

E. Bellido et al.

20

reviews

Figure 23 . a) Schematic illustration of the ferritin patterning by combining LON and surface functionalization with OTS and APTES. Local oxide pattern made of parallel lines with widths of 10–15 nm before (b) and after (c) adsorption of ferritin molecules into single molecule chains. d,e) Parallel patterning of ferritin molecules by means of lithography-controlled dewetting. Insets correspond to phase images indicating the lines widths. Reproduced with permission. [ 30 ]

target particles on a chip containing hundreds of devices, [ 161 ]

or by spray pyrolisis of nanodroplets of a solution containing

the sample nanoparticles. [ 162 ] After drying, the particles are

physisorbed onto the chip, but only the ones falling in the

sensitive zones of the sensors such as the nanobridge of the

SQUID loop, have enough fl ux coupling with the sensor to

be detected. Those are selected for the magnetic measure-

ments. In all these cases, the position and morphology of the

www.small-journal.com © 2012 Wiley-VCH V

particles is determined by SEM only after magnetization

measurements are carried out. Typically, parameters such

as concentration of MNPs in the solution or evaporation

time are adjusted to ensure a dilute submonolayer distribu-

tion of MNPs on the sensors array, so that MNPs are more

or less isolated on the surface and only one or few of them

can statistically reach the target position on the nano devices.

The advantages of random integration are mainly the

erlag GmbH & Co. KGaA, Weinheim small 2012, DOI: 10.1002/smll.201101456

Magnetic Nanoparticles on Surfaces and Devices

Figure 24 . a) 3D AFM image of a ferritin array generated by direct-write DPN on Au substrate with spacing between dot-like features of 500 nm and dot size of 100 nm. Reproduced with permission. [ 160 ] b) TEM images of a ferritin array fabricated on a carbon-coated TEM grid. The number of ferritin molecules deposited on each dot of the array could be clearly determined and studied due to the high contrast of the inorganic core of the protein. Reproduced with permission. [ 29 ]

straightforward application of 2D structuration techniques

already known without any modifi cations. The structura-

tion step is simple and many nanodevices can be function-

alized in a single experiment. The major drawback of this

methodology relies on the implicit poor control for the

exact particle position that hinders the experimental work

for the nanodevice operation characterization. Tens or even

hundreds of experiments must be done before an optimal

nanodevice with one or very few MNPs placed in the right

position is found. While this fact favors the obtaining of sta-

tistics over the operational properties, it can also be very

expensive and time-consuming.

The development of magnetic sensor devices such as

SQUIDs has been closely linked with the aim of measuring

the magnetization reversal of an individual magnetic par-

ticle. [ 163 ] For this, the SQUIDs not only have to increase their

sensitivity up to a single particle level, diminishing its char-

acteristic size down to few nanometers, but also MNPs must

be integrated on the areas of the device that show highest

sensitivity. To this end, Wernsdorfer and co-workers used a

randomized structuration methodology for the successful

© 2012 Wiley-VCH Verlag Gmbsmall 2012, DOI: 10.1002/smll.201101456

integration of individual nanosized nickel wires, as well as

MNPs of different types and sizes in the range of 10 nm to

50 nm, onto microSQUID detectors. [ 161 , 164 ]

Integration of the magnetic nanomaterials into CNT-

based devices [ 165 ] is itself another active fi eld of research.

For most CNT-based devices, the structuration process is less

demanding in terms of nanoscale location, but more chal-

lenging since it requires grafting strategies, such as the use

of non-covalent bonds, that minimize the presence of scat-

tering sites. [ 166 ] Sequential addition of a small and controlled

number of NPs can be achieved by reiterative repetition of

the integration process. For example, making use of van der

Waals interactions between the CNTs and the surfactant

alkylic chains that cover the MNPs, Bogani et al. have self-

assembled MNPs of CoFe 2 O 4 of ∼ 6 nm in diameter onto

metallic and semiconductor CNTs and studied the resulting

transport properties on electronic devices. [ 167 ] The hybrids

were produced by repeating immersions of CNT-based elec-

tronic devices into a dispersion of the MNPs. While the semi-

conductor CNT is sensitive to the attachment, the transport

properties of a metallic CNT remain unchanged, which is

desirable for the use of this grafting technique for the pro-

duction of CNT–SQUID devices. Moreover, it has been dem-

onstrated that the use of hydrophobic tails that promote van

der Waals interactions with the CNTs is well suited for a con-

trolled and selective grafting process since it allows a sequen-

tial grafting of a controlled number of NPs. This is of great

importance as spintronic devices require the sequential addi-

tion of a small, but very controlled, number of nanomagnets.

Finally, in some cases, CNT-based devices are built with

pre-derivatized CNTs. For instance, Kazakova et al. fabri-

cated magnetic and electromechanical nanodevices for single

particle detection and characterization with a CNT bridging

metallic electrodes previously decorated with (magnetic

iron)–(iron oxide) core–shell nanocubes of ∼ 18 nm. [ 168 ] The

decoration of the CNT with iron oxide nanocubes was per-

formed, fi rst in solution, and then the hybrid CNTs were

deposited and aligned into the devices by nanomanipulation

(i.e., a sharpened carbon fi ber was used to lift and position

the CNT), wet deposition (i.e., drop-casting), or a technique

based on AC-dielectrophoresis ( Figure 25 ).

5.2. Site-Selective Integration into Devices

By contrast with the random distribution based on statistics

methods, this approach requires the use of structuration and

functionalization methods with nanometric control and accu-

racy over the localization. Site-selective integration arises in

specifi c cases as, for example, when it is needed to optimize the

coupling factor between the MNP and the nanodevice (as is the

case for micro- or nano-SQUID sensors), or when some parts

of the nanodevice need to be kept in pristine conditions and

cannot be randomly functionalized, as happens with gradient

nanoHall magnetometer devices.

The challenge of integrating a controlled number of enti-

ties into magnetic sensor devices such as microSQUIDs with

accuracy and high-fl ux coupling factor has been addressed by

the use of tips. [ 169–173 ] In this sense, DPN has been shown to

21www.small-journal.comH & Co. KGaA, Weinheim

E. Bellido et al.

22 www.small-journal.com © 2012 Wiley-VCH Verlag GmbH

reviews

Figure 25 . a) TEM image of (iron)–(iron oxide) core–shell nanocubes ( ∼ 18 nm in size) attached to a single multiwall carbon nanotube (MWCNT). b) SEM image of an Au electrode array. c) SEM image of a single MWCNT decorated with (iron)-(iron oxide) nanocubes (light spots) and connected to two Au electrodes through wet deposition method. Reproduced by permission. [ 168 ] Copyright 2009, IOP Publishing Ltd.

Figure 26 . a) SEM image of the SQUID sensor showing the four rectangular shaped pick-up coils. b) Schematic illustration of the ferritin nanoparticle deposition process by DPN on the most active areas of the SQUID sensor. c,d) SEM images of a sensor directly after generating ferritin nanostructures on the desired positions of the sensor. Some of the obtained nanostructures are highlighted in image (d) with red circles. Reproduced with permission. [ 174 ] Copyright 2011, American Institute of Physics.

be appropriate to deposit a controlled number of

ferritin (with artifi cially synthesized ∼ 2 nm CoO

NPs inside) as submonolayers on the most sensitive

areas of a microSQUID, so maximizing the mag-

netic fl ux coupling of the sample with the sensor

( Figure 26 ). [ 174 ] Such areas were found, using theo-

retical calculations, to be located on the top of the

conducting wires or as close as possible to them. The

precision reached with this integration process was

also demonstrated by SEM images (Figure 26 c,d)

while the maximization of the fl ux coupling was

probed with the magnetic measurements performed

with the hybrid sensor. Furthermore, it was dem-

onstrated that the single monolayer arrangement

of magnetic ferritin nanoparticles allowed a better

thermalization of the sample with respect to tradi-

tional bulk measurements.

Another approximation is the use of litho-

graphic-assisted structuration techniques. Lam and

co-workers selectively integrated ferritin nanoparti-

cles, with a magnetic moment of μ < 300 μ B , into a

nanoSQUID (a step further on the miniaturization)

and detected its magnetization reversal. [ 175 ] In this

case, to optimize the fl ux coupling, a small number

of ferritin nanoparticles were directly attached over

the nanojunction with a method combining mono-

layer self-assembly with EBL ( Figure 27 ). [ 176 ] For

this, PMMA is deposited onto the Nb nanoSQUID

that has been previously covered with gold. After-

wards a 200 nm × 200 nm window in the PMMA

resistor is opened directly over the nanojunction by

EBL. Finally, organic linker molecules are deposited

directly over the exposed part of the gold and used

to attach the ferritin magnetic molecules to the Nd

nanoSQUID device. Even this approximation has

high accuracy, it shows two main disadvantatges:

i) the presence of the linker between the magnetic

material and the sensor surface diminishes their

effective coupling, and ii) surface functionalization

can modify the sensor characteristics.

Apart from these two previous methodologies, a

range of other techniques based on the use of tips can

also be applied. [ 173 , 177–179 ] Pakes et al. demonstrated

great control of the positioning of ferritin by mechan-

ical pushing of a single particle with the tip of an

AFM. [ 173 ] Garno et al. reviewed other methods based

on nanoshaving processes of SAM to deposit Au

nano particles onto the exposed areas. [ 178 ] Hao et al.

used a SEM together with a sharpened probe (carbon

fi bre) to lift and position a FePt nanobead directly

onto a nanoSQUID loop in a reversible way. [ 179 ]

Finally, one of the main candidates of interest

for the future in the fi eld of magnetic sensors is

graphene. It has shown itself to be a promising

material for use in different types of improved

devices based on its fascinating structural, elec-

trical and magnetic features, namely high crystal-

line order, mass-less Dirac fermion-like charge

carriers, or long spin coherence length. [ 180–182 ] It

& Co. KGaA, Weinheim small 2012, DOI: 10.1002/smll.201101456

Magnetic Nanoparticles on Surfaces and Devices

Figure 27 . a) Schematic diagram of the covalent attachment of a ferritin particle to the Au surface through the use of organic ligands [EDC refers to 1-ethyl-3-(3-dimethylaminopropyl) carbodiimide and NHS to N -hydroxysuccinimide]. b) AFM image of a 700 nm × 700 nm area of the SQUID sensor. The positions of the ferritin entities are indicated with arrows. The square on the image indicates the approximate location of the PMMA window used for attachment. Reproduced by permission. [ 176 ] Copyright 2008, IOP Publishing Ltd.

has recently been shown that Hall sensors made on graphene

are very promising for nanoscale detection of small mag-

netic signals. [ 183 ] The intrinsic high mobility of graphene

results in a high magnetic fi eld sensitivity while the structure

of graphene also provides an atomically thick sensor layer

located right at the surface that reduces the loss of magnetic

fl ux throughout the sensor thickness; therefore increasing the

resolution and the sensitivity. [ 184 ] However, a delicate step in

the integration on this device is the proper manipulation and

the controlled deposition of a magnetic material while pre-

venting damage of the graphene fl akes. Moreover, nanoHall

gradiometry technique for example, requires that the mate-

rial is positioned exclusively over one of the cross-sections of

the device, that can be of the order of a few square nano-

meters, while the other is kept empty. This situation obviously

requires the use of an accurate and nanometer site-selective

structuration technique. In this case, direct-write AFM nano-

structuration technique has been shown to be appropriate

since the probe tip can be used to both visualize the sensor

with nanometric resolution without damaging the surface to

locate the nanometric target region, and afterwards directly

deposit nanomaterials on it with site-specifi city, using the tip

© 2012 Wiley-VCH Verlag Gmbsmall 2012, DOI: 10.1002/smll.201101456

to deliver controlled quantities of an ink by capillary trans-

port. Domingo and co-workers have successfully used this

selective and direct integration method to deposit some

tens of Co MNPs of ∼ 8 nm diameter on the 1 μ m 2 cross-

section of a graphene-based nanoHall sensor with accurate

resolution and without damage of the graphene fl ake (see

Figure 28 ). [ 185 ]

6. Summary and Perspectives

A detailed consideration of the main scientifi c approaches

developed over the last years for the deposition and structur-

ation of magnetic nanoparticles on surfaces has been given in

this review. For this, the manuscript has been divided into four

main areas we consider crucial and which, therefore, require

special attention: (i) un-assisted self-assembly, (ii) assisted

self-assembly and organization, (iii) use of biomolecular tem-

plates, and (iv) integration into hybrid devices.

As expected, the fi rst two sections involving the devel-

opment of surface structuration techniques is where sci-

entists started to develop their work, and consequently,

where most examples have been described. Such numerous

examples clearly reveal the large number of experimental

techniques used for: i) the structuration of MNPs on sur-

faces that range from a few nanometers up to several mil-

limeters, as well as ii) controling the number of MNPs

deposited from complete layers down to the deposition

of a single nanoparticle. A high degree of control can be

achieved simply by considering structural parameters such

as shape, size, and morphology of the nanoparticles. Special

attention has also been given to the different non-covalent

interactions that may infl uence the self-assembly proc-

esses, where dipolar interactions play an important role.

Whereas van der Waals interactions and steric forces can

be modulated by using different type of surfactants, dipolar

magnetic interactions can be modifi ed by changing the

size of the magnetic NPs since the magnetic dipolar inter-

actions increase quadratically with the particle magnetic

moment, which is also proportional to its volume. In this

sense, dipole–dipole interactions become the most effective

supramolecular interaction for large enough particles with

high magnetic moments.

The deposition methods, from the most simple drop-

casting or spin-coating techniques, to lithographically assisted

methodologies, have also been shown to have a crucial infl u-

ence. For instance, if the deposition is assisted by lithographic,

AFM, or similar techniques, structured arrays of diverse

motives can be achieved with control of the number, as well

as the morphology, size, and pitch of the dots. Moreover,

improvement of the resulting arrays is achieved by chemical

functionalization of either the nanoparticle and/or the sur-

face, allowing in most cases excellent control of the formation

of a (sub)monolayer. To improve the effectiveness of the

techniques previously described it is possible to use bio-

logical templates such as viruses or ferritins. The advantages

are twofold. First, once the conditions for the structuration

of the organic capsule are established, these can be used for

the structuration of a broad range of inorganic nanoparticles

23www.small-journal.comH & Co. KGaA, Weinheim

E. Bellido et al.reviews

Figure 28 . a) 3D AFM image of a graphene-based nanoHall sensor with cross-sections of 1 μ m × 1 μ m in size. b) Hall cross indicated in image (a) was functionalized with Co NPs ( ∼ 8 nm) by means of AFM-based nanolithography. c) Height profi le along the white dashed line in image (b). [ 185 ]

that can be held inside, enormously simplifying the process.

Second, the organic capsule can be modifi ed by means of a

biochemical approach to ensure and/or improve the resulting

arrays.

In spite of the large number of techniques already

described, new examples of techniques appear daily, trying to

fulfi l the high demands to build magnetic nanoscale devices.

Although still necessary, it would be of rather higher benefi t

to the scientifi c community to focus the development of these

techniques towards integrating the nanoparticles on devices

rather than simply describing new structuration procedures.

From the different techniques reported, those that have been

shown to have more potential are self-assembly or dip-pen

nanotechnology (DPN). Remarkable advances have been

obtained especially with DPN, which has allowed the inte-

gration of the MNPs on the most sensitive areas of 1 μ m ×

1 μ m graphene-based sensors. These results are very impor-

tant since integration techniques must, in the future, fulfi ll the

requirements presented by the continuous miniaturization

of devices, such as ensuring high precision of the deposition

process and careful manipulation, in order to minimize the

effect over the sensor at any time. Unfortunately, there is

a long way to go before this technique can be applied, due

to the main limitations of reproducibility and scalability.

Improving both limitations will require full control over the

self-assembly processes of the molecules, which is not an easy

task. A good solution to overcome these limitations could be

the combination of self-assembly with AFM-assisted lithog-

raphies in solution. While the integration of the molecular

materials on the sensors and devices could represent a real

challenge, the maintenance of the magnetic properties upon

integration is a must. Perturbations arising from surface

24 www.small-journal.com © 2012 Wiley-VCH Ve

interactions or other chemical modifi cations (i.e., redox proc-

esses, chemical decomposition, etc.) that alter the magnetic

properties must also be strongly taken into consideration.

These factors, which have been shown to be crucial for softer

materials such as molecular magnetic materials, are at the

present moment not always considered.

Acknowledgements

We thank the Spanish government for the project MAT2009-13977-C03-03. N.D and E.B. thank the Spanish government for a Ramon y Cajal and an F.P.I. grant, respectively.

[ 1 ] C. Burda , X. Chen , R. Narayanan , M. A. El-Sayed , Chem. Rev. 2005 , 105 , 1025 – 1102 .

[ 2 ] A. P. Alivisatos , Science 1996 , 271 , 933 – 937 . [ 3 ] C. B. Murray , C. R. Kagan , M. G. Bawendi , Annu. Rev. Mater. Sci.

2000 , 30 , 545 – 610 . [ 4 ] S. Sun , C. B. Murray , J. Appl. Phys. 1999 , 85 , 4325 – 4330 . [ 5 ] T. Hyeon , Y. Chung , J. Park , S. S. Lee , Y.-W. Kim , B. H. Park , J.

Phys. Chem. B 2002 , 106 , 6831 – 6833 . [ 6 ] T. Hyeon , Chem. Commun. 2003 , 927 – 934 . [ 7 ] C. M. Sorensen in Nanoscale Materials in Chemistry , 6 (Ed:

K. J. Klabunde ), John Wiley & Sons, Inc. , New York , 2002 ; pp 169 – 221 .

[ 8 ] D. E. Speliotis , J. Magn. Magn. Mater. 1999 , 193 , 29 – 35 . [ 9 ] S. Sun , C. B. Murray , D. Weller , L. Folks , A. Moser , Science 2000 ,

287 , 1989 – 1992 . [ 10 ] C. A. Ross , Annu. Rev. Mater. Res. 2001 , 31 , 203 – 235 .

rlag GmbH & Co. KGaA, Weinheim small 2012, DOI: 10.1002/smll.201101456

Magnetic Nanoparticles on Surfaces and Devices

[ 11 ] C. Chappert , A. Fert , F. N. Van Dau , Nat. Mater. 2007 , 6 , 813 – 823 .

[ 12 ] P. K. Gupta , C. T. Hung , F. C. Lam , D. G. Perrier , Int. J. Pharm. 1988 , 43 , 167 – 177 .

[ 13 ] V. P. Torchilin , Eur. J. Pharm. Sci. 2000 , 11 , S81 – S91 . [ 14 ] H.-T. Song , J.-s. Choi , Y.-M. Huh , S. Kim , Y.-w. Jun , J.-S. Suh ,

J. Cheon , J. Am. Chem. Soc. 2005 , 127 , 9992 – 9993 . [ 15 ] B. Bonnemain , J. Drug Target. 1998 , 6 , 167 – 174 . [ 16 ] U. I. Tromsdorf , N. C. Bigall , M. G. Kaul , O. T. Bruns ,

M. S. Nikolic , B. Mollwitz , R. A. Sperling , R. Reimer , H. Hohenberg , W. J. Parak , S. Förster , U. Beisiegel , G. Adam , H. Weller , Nano Lett. 2007 , 7 , 2422 – 2427 .

[ 17 ] V. Russier , C. Petit , J. Legrand , M. P. Pileni , Phys. Rev. B 2000 , 62 , 3910 .

[ 18 ] M. P. Pileni , Adv. Funct. Mater. 2001 , 11 , 323 – 336 . [ 19 ] M. P. Pileni , J. Phys. Chem. B 2001 , 105 , 3358 – 3371 . [ 20 ] C. Petit , V. Russier , M. P. Pileni , J. Phys. Chem. B 2003 , 107 ,

10333 – 10336 . [ 21 ] M. P. Pileni , Acc. Chem. Res. 2007 , 40 , 685 – 693 . [ 22 ] R. P. Cowburn , M. E. Welland , Science 2000 , 287 , 1466 – 1468 . [ 23 ] H. Zeng , C. T. Black , R. L. Sandstrom , P. M. Rice , C. B. Murray ,

S. Sun , Phys. Rev. B 2006 , 73 , 020402 . [ 24 ] S. Palacin , P. C. Hidber , J.-P. Bourgoin , C. Miramond , C. Fermon ,

G. M. Whitesides , Chem. Mater. 1996 , 8 , 1316 – 1325 . [ 25 ] L. An , W. Li , Y. Nie , B. Xie , Z. Li , J. Zhang , B. Yang , J. Colloid Inter-

face Sci. 2005 , 288 , 503 – 507 . [ 26 ] G. Chen , M. I. Bodnarchuk , M. V. Kovalenko , G. Springholz ,

W. Heiss , W. Jantsch , E. Platzgummer , H. Loeschner , J. Schotter , Adv. Mater. 2010 , 22 , 1364 – 1368 .

[ 27 ] X. Liu , L. Fu , S. Hong , V. P. Dravid , C. A. Mirkin , Adv. Mater. 2002 , 14 , 231 – 234 .

[ 28 ] Y. Wang , W. Wei , D. Maspoch , J. Wu , V. P. Dravid , C. A. Mirkin , Nano Lett. 2008 , 8 , 3761 – 3765 .

[ 29 ] E. Bellido , R. de Miguel , D. Ruiz-Molina , A. Lostao , D. Maspoch , Adv. Mater. 2010 , 22 , 352 – 355 .

[ 30 ] R. V. Martínez , J. Martínez , M. Chiesa , R. Garcia , E. Coronado , E. Pinilla-Cienfuegos , S. Tatay , Adv. Mater. 2010 , 22 , 588 – 591 .

[ 31 ] Y. Xia , J. A. Rogers , K. E. Paul , G. M. Whitesides , Chem. Rev. 1999 , 99 , 1823 – 1848 .

[ 32 ] S. Koh , Nanoscale Res. Lett. 2007 , 2 , 519 – 545 . [ 33 ] S. Kinge , M. Crego-Calama , D. N. Reinhoudt , ChemPhysChem

2008 , 9 , 20 – 42 . [ 34 ] Z. Nie , A. Petukhova , E. Kumacheva , Nat. Nanotechnol. 2010 , 5 ,

15 – 25 . [ 35 ] C. Ross , Annu. Rev. Mater. Res. 2001 , 31 , 203 – 235 . [ 36 ] B. D. Terris , T. Thomson , J. Phys. D: Appl. Phys. 2005 , 38 , R199 . [ 37 ] J. H. Fendler , Chem. Mater. 1996 , 8 , 1616 – 1624 . [ 38 ] G. M. Whitesides , B. Grzybowski , Science 2002 , 295 ,

2418 – 2421 . [ 39 ] F. X. Redl , C. T. Black , G. C. Papaefthymiou , R. L. Sandstrom ,

M. Yin , H. Zeng , C. B. Murray , S. P. O’Brien , J. Am. Chem. Soc. 2004 , 126 , 14583 – 14599 .

[ 40 ] I. Lisiecki , P.-A. Albouy , M.-P. Pileni , J. Phys. Chem. B 2004 , 108 , 20050 – 20055 .

[ 41 ] S. A. Iakovenko , A. S. Trifonov , M. Giersig , A. Mamedov , D. K. Nagesha , V. V. Hanin , E. C. Soldatov , N. A. Kotov , Adv. Mater. 1999 , 11 , 388 – 392 .

[ 42 ] V. Aleksandrovic , D. Greshnykh , I. Randjelovic , A. Frömsdorf , A. Kornowski , S. V. Roth , C. Klinke , H. Weller , ACS Nano 2008 , 2 , 1123 – 1130 .

[ 43 ] T. Fried , G. Shemer , G. Markovich , Adv. Mater. 2001 , 13 , 1158 – 1161 .

[ 44 ] E. Pohjalainen , M. Pohjakallio , C. Johans , K. s. Kontturi , J. V. I. Timonen , O. Ikkala , R. H. A. Ras , T. Viitala , M. T. Heino , E. T. Seppälä , Langmuir 2010 , 26 , 13937 – 13943 .

[ 45 ] M. I. Bodnarchuk , M. V. Kovalenko , S. Pichler , G. Fritz-Popovski , G. n. Hesser , W. Heiss , ACS Nano 2009 , 4 , 423 – 431 .

© 2012 Wiley-VCH Verlag Gmbsmall 2012, DOI: 10.1002/smll.201101456

[ 46 ] Y.-K. Hong , H. Kim , G. Lee , W. Kim , J.-I. Park , J. Cheon , J.-Y. Koo , Appl. Phys. Lett. 2002 , 80 , 844 – 846 .

[ 47 ] L. Chitu , Y. Chushkin , S. Luby , E. Majkova , A. Satka , J. Ivan , L. Smrcok , A. Buchal , M. Giersig , M. Hilgendorff , Mater. Sci. Eng.C 2007 , 27 , 23 – 28 .

[ 48 ] M. Varón , L. Peña , L. Balcells , V. Skumryev , B. Martinez , V. Puntes , Langmuir 2009 , 26 , 109 – 116 .

[ 49 ] A. P. Philipse , D. Maas , Langmuir 2002 , 18 , 9977 – 9984 . [ 50 ] M. T. Klem , M. Young , T. Douglas , Materials Today 2005 , 8 ,

28 – 37 . [ 51 ] M. P. Pileni , Phys. Chem. Chem. Phys. 2010 , 12 , 11821 – 11835 . [ 52 ] S. L. Tripp , S. V. Pusztay , A. E. Ribbe , A. Wei , J. Am. Chem. Soc.

2002 , 124 , 7914 – 7915 . [ 53 ] K. Butter , P. H. H. Bomans , P. M. Frederik , G. J. Vroege ,

A. P. Philipse , Nat. Mater. 2003 , 2 , 88 – 91 . [ 54 ] L. Motte , A. Courty , A.-T. Ngo , I. Lisiecki , M.-P. Pileni in Nano-

crystals Forming Mesoscopic Structures , (Ed.: M.-P. Pileni ), Wiley-VCH , Weinheim , 2006 ; pp 1 – 47 .

[ 55 ] Y. Lalatonne , J. Richardi , M. P. Pileni , Nat. Mater. 2004 , 3 , 121 – 125 .

[ 56 ] V. F. Puntes , K. M. Krishnan , A. P. Alivisatos , Science 2001 , 291 , 2115 – 2117 .

[ 57 ] T. S. Yoon , J. Oh , S. H. Park , V. Kim , B. G. Jung , S. H. Min , J. Park , T. Hyeon , K. B. Kim , Adv. Funct. Mater. 2004 , 14 , 1062 – 1068 .

[ 58 ] I. Lisiecki , P. A. Albouy , M. P. Pileni , Adv. Mater. 2003 , 15 , 712 – 716 .

[ 59 ] H. Wang , H. Wang , F. Yang , J. Zhang , Q. Li , M. Zhou , Y. Jiang , Surf. Coat. Technol. 2010 , 204 , 1509 – 1513 .

[ 60 ] M. I. Bodnarchuk , M. V. Kovalenko , W. Heiss , D. V. Talapin , J. Am. Chem. Soc. 2010 , 132 , 11967 – 11977 .

[ 61 ] L. Pena , M. Varon , Z. Konstantinovic , L. Balcells , B. Martinez , V. Puntes , J. Mater. Chem. 2011 , 21 , 16973 – 16977 .

[ 62 ] W. H. Evers , B. D. Nijs , L. Filion , S. Castillo , M. Dijkstra , D. Vanmaekelbergh , Nano Lett. 2010 , 10 , 4235 – 4241 .

[ 63 ] S. Sun , S. Anders , H. F. Hamann , J.-U. Thiele , J. E. E. Baglin , T. Thomson , E. E. Fullerton , C. B. Murray , B. D. Terris , J. Am. Chem. Soc. 2002 , 124 , 2884 – 2885 .

[ 64 ] S. Sun , S. Anders , T. Thomson , J. E. E. Baglin , M. F. Toney , H. F. Hamann , C. B. Murray , B. D. Terris , J. Phys. Chem. B 2003 , 107 , 5419 – 5425 .

[ 65 ] J. H. Kim , A.-Y. Song , D. H. Kwon , H. H. An , H. S. Ahnn , Y.-K. Kim , C. S. Yoon , Appl. Surf. Sci. 2011 , 257 , 3128 – 3134 .

[ 66 ] C. Altavilla , E. Ciliberto , D. Gatteschi , C. Sangregorio , Adv. Mater. 2005 , 17 , 1084 – 1087 .

[ 67 ] B. Fleury , F. Volatron , L. Catala , D. Brinzei , E. Rivière , V. Huc , C. David , F. Miserque , G. Rogez , L. Baraton , S. Palacin , T. Mallah , Inorg. Chem. 2008 , 47 , 1898 – 1900 .

[ 68 ] F. Cattaruzza , D. Fiorani , A. Flamini , P. Imperatori , G. Scavia , L. Suber , A. M. Testa , A. Mezzi , G. Ausanio , W. R. Plunkett , Chem. Mater. 2005 , 17 , 3311 – 3316 .

[ 69 ] C. Altavilla , E. Ciliberto , A. Aiello , C. Sangregorio , D. Gatteschi , Chem. Mater. 2007 , 19 , 5980 – 5985 .

[ 70 ] E. Fanizza , P. D. Cozzoli , M. L. Curri , M. Striccoli , E. Sardella , A. Agostiano , Adv. Funct. Mater. 2007 , 17 , 201 – 211 .

[ 71 ] G. Leem , A. C. Jamison , S. Zhang , D. Litvinov , T. R. Lee , Chem. Commun. 2008 , 4989 – 4991 .

[ 72 ] G. Leem , S. Zhang , A. C. Jamison , E. Galstyan , I. Rusakova , B. Lorenz , D. Litvinov , T. R. Lee , ACS Appl. Mater. Interfaces 2010 , 2 , 2789 – 2796 .

[ 73 ] B. t. P. Pichon , A. Demortiére , M. Pauly , K. Mougin , A. Derory , S. Bégin-Colin , J. Phys. Chem. C 2010 , 114 , 9041 – 9048 .

[ 74 ] A. C. C. Yu , M. Mizuno , Y. Sasaki , M. Inoue , H. Kondo , I. Ohta , D. Djayaprawira , M. Takahashi , Appl. Phys. Lett. 2003 , 82 , 4352 – 4354 .

[ 75 ] S. Sun , Adv. Mater. 2006 , 18 , 393 – 403 .

25www.small-journal.comH & Co. KGaA, Weinheim

E. Bellido et al.reviews

[ 76 ] S. Kinge , T. Gang , W. J. M. Naber , W. G. van der Wiel ,

D. N. Reinhoudt , Langmuir 2010 , 27 , 570 – 574 . [ 77 ] A. Bousquet , E. Ibarboure , C. Labrugere , E. Papon ,

J. Rodríguez-Hernández , Langmuir 2007 , 23 , 6879 – 6882 . [ 78 ] G. Leo , Y. Chushkin , S. Luby , E. Majkova , I. Kostic , M. Ulmeanu ,

A. Luches , M. Giersig , M. Hilgendorff , Mater. Sci. Eng. C 2003 , 23 , 949 – 952 .

[ 79 ] J.-I. Park , Y.-w. Jun , J.-s. Choi , J. Cheon , Chem. Commun. 2007 , 5001 – 5003 .

[ 80 ] C. Petit , J. Legrand , V. Russier , M. P. Pileni , J. Appl. Phys. 2002 , 91 , 1502 – 1508 .

[ 81 ] J. Legrand , A. T. Ngo , C. Petit , M. P. Pileni , Adv. Mater. 2001 , 13 , 58 – 62 .

[ 82 ] V. Germain , M. P. Pileni , Adv. Mater. 2005 , 17 , 1424 – 1429 .

[ 83 ] A. T. Ngo , M. P. Pileni , New J. Phys. 2002 , 4 , 871 – 878 . [ 84 ] O. Yildirim , T. Gang , S. Kinge , D. Reinhoudt , D. Blank ,

W. van der Wiel , G. Rijnders , J. Huskens , Int. J. Mol. Sci. 2010 , 11 , 1162 – 1179 .

[ 85 ] Q. Guo , X. Teng , S. Rahman , H. Yang , J. Am. Chem. Soc. 2002 , 125 , 630 – 631 .

[ 86 ] Q. Guo , X. Teng , H. Yang , Adv. Mater. 2004 , 16 , 1337 – 1341 . [ 87 ] J.-I. Park , W.-R. Lee , S.-S. Bae , Y. J. Kim , K.-H. Yoo , J. Cheon ,

S. Kim , J. Phys. Chem. B 2005 , 109 , 13119 – 13123 . [ 88 ] R. N. Patel , A. T. Heitsch , C. Hyun , D.-M. Smilgies ,

A. de Lozanne , Y.-L. Loo , B. A. Korgel , ACS Appl. Mater. Inter-faces 2009 , 1 , 1339 – 1346 .

[ 89 ] M. Cavallini , E. Bystrenova , M. Timko , M. Koneracka , V. Zavisova , P. Kopcansky , J. Phys.: Condens. Matter. 2008 , 20 , 204144 .

[ 90 ] X. Yang , C. Liu , J. Ahner , J. Yu , T. Klemmer , E. Johns , D. Weller , J. Vac. Sci. Technol. B 2004 , 22 , 31–34 .

[ 91 ] I. Seo , C. W. Kwon , H. H. Lee , Y. S. Kim , K. B. Kim , T. S. Yoon , Electrochem. Solid State Lett. 2009 , 12 , K59 – K62 .

[ 92 ] A. Mari , E. Agostinelli , D. Fiorani , A. Flamini , S. Laureti , D. Peddis , A. M. Testa , G. Varvaro , M. V. Mansilla , A. Mezzi , S. Kaciulis , Superlattices Microstruct. 46 , 95 – 100 .

[ 93 ] M. Chen , D. E. Nikles , H. Yin , S. Wang , J. W. Harrell , S. A. Majetich , J. Magn. Magn. Mater. 2003 , 266 , 8 – 11 .

[ 94 ] R. Garcia , R. V. Martinez , J. Martinez , Chem. Soc. Rev. 2006 , 35 , 29 – 38 .

[ 95 ] B. Basnar , J. Xu , D. Li , I. Willner , Langmuir 2007 , 23 , 2293 – 2296 .

[ 96 ] S. Hoeppener , A. S. Susha , A. L. Rogach , J. Feldmann , U. S. Schubert , Curr. Nanosci. 2006 , 2 , 135 – 141 .

[ 97 ] Y. Sasaki , M. Mizuno , A. C. C. Yu , T. Miyauchi , D. Hasegawa , T. Ogawa , M. Takahashi , B. Jeyadevan , K. Tohji , K. Sato , S. Hisano , IEEE Trans. Magn. 2005 , 41 , 660 – 664 .

[ 98 ] M. Cavallini , F. C. Simeone , F. Borgatti , C. Albonetti , V. Morandi , C. Sangregorio , C. Innocenti , F. Pineider , E. Annese , G. Panaccione , L. Pasquali , Nanoscale 2010 , 2 , 2069 – 2072 .

[ 99 ] R. D. Piner , J. Zhu , F. Xu , S. Hong , C. A. Mirkin , Science 1999 , 283 , 661 – 663 .

[ 100 ] D. S. Ginger , H. Zhang , C. A. Mirkin , Angew. Chem., Int. Ed. 2004 , 43 , 30 – 45 .

[ 101 ] B. Basnar , I. Willner , Small 2009 , 5 , 28 – 44 . [ 102 ] J. Haaheim , R. Eby , M. Nelson , J. Fragala , B. Rosner , H. Zhang ,

G. Athas , Ultramicroscopy 2005 , 103 , 117 – 132 . [ 103 ] G. Gundiah , N. S. John , P. J. Thomas , G. U. Kulkarni , C. N. R. Rao ,

S. Heun , Appl. Phys. Lett. 2004 , 84 , 5341 – 5343 . [ 104 ] B. Li , C. F. Goh , X. Zhou , G. Lu , H. Tantang , Y. Chen , C. Xue ,

F. Y. C. Boey , H. Zhang , Adv. Mater. 2008 , 20 , 4873 – 4878 . [ 105 ] D. Nyamjav , A. Ivanisevic , Chem. Mater. 2004 , 16 , 5216 – 5219 . [ 106 ] P. E. Sheehan , L. J. Whitman , W. P. King , B. A. Nelson , Appl.

Phys. Lett. 2004 , 85 , 1589 – 1591 . [ 107 ] W. K. Lee , Z. Dai , W. P. King , P. E. Sheehan , Nano Lett. 2009 , 10 ,

129 – 133 . [ 108 ] G. Jutz , A. Böker , Polymer 2011 , 52 , 211 – 232 .

26 www.small-journal.com © 2012 Wiley-VCH V

[ 109 ] I. Yamashita , J. Mater. Chem. 2008 , 18 , 3813 – 3820 . [ 110 ] M. Young , W. Debbie , M. Uchida , T. Douglas , Annu. Rev. Phy-

topathol. 2008 , 46 , 361 – 384 . [ 111 ] E. Dujardin , C. Peet , G. Stubbs , J. N. Culver , S. Mann , Nano Lett.

2003 , 3 , 413 – 417 . [ 112 ] M. Kobayashi , M. Seki , H. Tabata , Y. Watanabe , I. Yamashita ,

Nano Lett. 2010 , 10 , 773 – 776 . [ 113 ] R. Tsukamoto , M. Muraoka , M. Seki , H. Tabata , I. Yamashita ,

Chem. Mater. 2007 , 19 , 2389 – 2391 . [ 114 ] M. Allen , D. Willits , J. Mosolf , M. Young , T. Douglas , Adv. Mater.

2002 , 14 , 1562 – 1565 . [ 115 ] M. Allen , D. Willits , M. Young , T. Douglas , Inorg. Chem. 2003 ,

42 , 6300 – 6305 . [ 116 ] P. M. Harrison , P. Arosio , Biochim. Biophys. Acta, Bioenerg.

1996 , 1275 , 161 – 203 . [ 117 ] K. Iwahori , I. Yamashita , J. Cluster Sci. 2007 , 18 , 358 – 370 . [ 118 ] F. Meldrum , B. Heywood , S. Mann , Science 1992 , 257 ,

522 – 523 . [ 119 ] M. J. Martinez-Perez , R. de Miguel , C. Carbonera ,

M. Martinez-Julvez , A. Lostao , C. Piquer , C. Gomez-Moreno , J. Bartolome , F. Luis , Nanotechnology 2010 , 21 , 465707 .

[ 120 ] R. Tsukamoto , K. Iwahori , M. Muraoka , I. Yamashita , Bull. Chem. Soc. Jpn. 2005 , 78 , 2075 – 2081 .

[ 121 ] T. Douglas , V. T. Stark , Inorg. Chem. 2000 , 39 , 1828 – 1830 . [ 122 ] F. C. Meldrum , T. Douglas , S. Levi , P. Arosio , S. Mann , J. Inorg.

Biochem. 1995 , 58 , 59 – 68 . [ 123 ] M. T. Klem , J. Mosolf , M. Young , T. Douglas , Inorg. Chem. 2008 ,

47 , 2237 – 2239 . [ 124 ] N. Galvez , P. Sanchez , J. M. Dominguez-Vera , Dalton Trans.

2005 , 2492 – 2494 . [ 125 ] T. Ueno , M. Suzuki , T. Goto , T. Matsumoto , K. Nagayama ,

Y. Watanabe , Angew. Chem., Int. Ed. 2004 , 43 , 2527 – 2530 . [ 126 ] N. Galvez , P. Sanchez , J. M. Dominguez-Vera , A. Soriano-Portillo ,

M. Clemente-Leon , E. Coronado , J. Mater. Chem. 2006 , 16 , 2757 – 2761 .

[ 127 ] B. Warne , O. I. Kasyutich , E. L. Mayes , J. A. L. Wiggins , K. K. W. Wong , IEEE Trans. Magn. 2000 , 36 , 3009 – 3011 .

[ 128 ] I. Yamashita , J. Hayashi , M. Hara , Chem. Lett. 2004 , 33 , 1158 – 1159 .

[ 129 ] K. Iwahori , K. Yoshizawa , M. Muraoka , I. Yamashita , Inorg. Chem. 2005 , 44 , 6393 – 6400 .

[ 130 ] I. Yamashita , K. Iwahori , S. Kumagai , Biochim. Biophys. Acta, Gen. Subj. 2010 , 1800 , 846 – 857 .

[ 131 ] T. Hikono , Y. Uraoka , T. Fuyuki , I. Yamashita , Jpn. J. Appl. Phys. 2003 , 42 , L398 – L399 .

[ 132 ] I. Yamashita , Thin Solid Films 2001 , 393 , 12 – 18 . [ 133 ] E. Adachi , K. Nagayama , Langmuir 1996 , 12 , 1836 –

1839 . [ 134 ] Z. Yuan , D. N. Petsev , B. G. Prevo , O. D. Velev , P. Atanassov ,

Langmuir 2007 , 23 , 5498 – 5504 . [ 135 ] K. Won , M. J. Park , H. H. Yoon , J. H. Kim , Ultramicroscopy 2008 ,

108 , 1342 – 1347 . [ 136 ] B. L. Scott , D. C. Zapien , Electroanalysis 2010 , 22 , 379 – 383 . [ 137 ] J. M. Domínguez-Vera , L. Welte , N. Gálvez , B. Fernández ,

J. Gómez-Herrero , F. Zamora , Nanotechnology 2008 , 19 , 025302 .

[ 138 ] J.-W. Kim , S. H. Choi , P. T. Lillehei , S.-H. Chu , G. C. King , G. D. Watt , J. Electroanal. Chem. 2007 , 601 , 8 – 16 .

[ 139 ] T. Matsui , N. Matsukawa , K. Iwahori , K.-I. Sano , K. Shiba , I. Yamashita , Langmuir 2007 , 23 , 1615 – 1618 .

[ 140 ] Y. Ikezoe , Y. Kumashiro , K. Tamada , T. Matsui , I. Yamashita , K. Shiba , M. Hara , Langmuir 2008 , 24 , 12836 – 12841 .

[ 141 ] A. Miura , Y. Uraoka , T. Fuyuki , S. Kumagai , S. Yoshii , N. Matsukawa , I. Yamashita , Surf. Sci. 2007 , 601 , L81 – L85 .

[ 142 ] K. Uto , K. Yamamoto , N. Kishimoto , M. Muraoka , T. Aoyagi , I. Yamashita , J. Mater. Chem. 2008 , 18 , 3876 – 3884 .

[ 143 ] M. Tominaga , M. Matsumoto , K. Soejima , I. Taniguchi , J. Colloid Interface Sci. 2006 , 299 , 761 – 765 .

erlag GmbH & Co. KGaA, Weinheim small 2012, DOI: 10.1002/smll.201101456

Magnetic Nanoparticles on Surfaces and Devices

[ 144 ] K. Yamada , S. Yoshii , S. Kumagai , I. Fujiwara , K. Nishio , M. Okuda , N. Matsukawa , I. Yamashita , Jpn. J. Appl. Phys. 2006 , 45 , 4259 – 4264 .

[ 145 ] S. Daniels , J. Ngunjiri , J. Garno , Anal. Bioanal. Chem. 2009 , 394 , 215 – 223 .

[ 146 ] F. Zhang , R. J. Gates , V. S. Smentkowski , S. Natarajan , B. K. Gale , R. K. Watt , M. C. Asplund , M. R. Linford , J. Am. Chem. Soc. 2007 , 129 , 9252 – 9253 .

[ 147 ] Y. X. Hu , D. A. Chen , S. Park , T. Emrick , T. P. Russell , Adv. Mater. 2010 , 22 , 2583 – 2587 .

[ 148 ] J.-K. Chen , Z.-Y. Chen , H.-C. Lin , P.-D. Hong , F.-C. Chang , ACS Appl. Mater. Interfaces 2009 , 1 , 1525 – 1532 .

[ 149 ] K.-I. Sano , S. Yoshii , I. Yamashita , K. Shiba , Nano Lett. 2007 , 7 , 3200 – 3202 .

[ 150 ] E. Bellido , R. de Miguel , D. Ruiz-Molina , A. Lostao , D. Maspoch , Adv. Mater. 2010 , 22 , 352 .

[ 151 ] R. V. Martinez , J. Martinez , M. Chiesa , R. Garcia , E. Coronado , E. Pinilla-Cienfuegos , S. Tatay , Adv. Mater. 2010 , 22 , 588 .

[ 152 ] K. Ishikawa , K. Yamada , S. Kumagai , K. I. Sano , K. Shiba , I. Yamashita , M. Kobayashi , Appl. Phys. Express 2008 , 1 , 034006 .

[ 153 ] K. Yamashita , H. Kirimura , M. Okuda , K. Nishio , K. I. Sano , K. Shiba , T. Hayashi , M. Hara , Y. Mishima , Small 2006 , 2 , 1148 – 1152 .

[ 154 ] N. Matsukawa , K. Nishio , K. Sano , K. Shiba , I. Yamashita , Lang-muir 2009 , 25 , 3327 – 3330 .

[ 155 ] S. Kumagai , S. Yoshii , K. Yamada , N. Matsukawa , I. Fujiwara , K. Iwahori , I. Yamashita , Appl. Phys. Lett. 2006 , 88 , 153103 .

[ 156 ] S. Yoshii , S. Kumagai , K. Nishio , A. Kadotani , I. Yamashita , Appl. Phys. Lett. 2009 , 95 , 133702 .

[ 157 ] S. Kumagai , T. Ono , S. Yoshii , A. Kadotani , R. Tsukamoto , K. Nishio , M. Okuda , I. Yamashita , Appl. Phys. Express 2010 , 3 , 015101 .

[ 158 ] T. Yoshinobu , J. Suzuki , H. Kurooka , W. C. Moon , H. Iwasaki , Electrochim. Acta 2003 , 48 , 3131 – 3135 .

[ 159 ] E. Manning , S.-T. Yau , J. Vac. Sci. Technol., B 2005 , 23 , 2309 – 2313 .

[ 160 ] E. Bellido , R. d. Miguel , J. Sesé , D. Ruiz-Molina , A. Lostao , D. Maspoch , Scanning 2010 , 32 , 35 – 41 .

[ 161 ] W. Wernsdorfer , E. B. Orozco , K. Hasselbach , A. Benoit , B. Barbara , N. Demoncy , A. Loiseau , H. Pascard , D. Mailly , Phys. Rev. Lett. 1997 , 78 , 1791 – 1794 .

[ 162 ] S. Bertram , T. Unto , I. K. Esko , M. Michel , R. Leif , P. Mikko , W. Wolfgang , B. Alain , Appl. Organomet. Chem. 1998 , 12 , 315 – 320 .

[ 163 ] W. Wernsdorfer , Supercond. Sci. Technol. 2009 , 22 , 064013 .

[ 164 ] W. Wernsdorfer , B. Doudin , D. Mailly , K. Hasselbach , A. Benoit , J. Meier , J. P. Ansermet , B. Barbara , Phys. Rev. Lett. 1996 , 77 , 1873 .

© 2012 Wiley-VCH Verlag Gmbsmall 2012, DOI: 10.1002/smll.201101456

[ 165 ] L. Bogani , W. Wernsdorfer , Inorg. Chim. Acta 2008 , 361 , 3807 – 3819 .

[ 166 ] Y.-L. Zhao , J. F. Stoddart , Acc. Chem. Res. 2009 , 42 , 1161 – 1171 .

[ 167 ] L. Bogani , R. Maurand , L. Marty , C. Sangregorio , C. Altavilla , W. Wernsdorfer , J. Mater. Chem. 2010 , 20 , 2099 – 2107 .

[ 168 ] I. Rod , O. Kazakova , D. C. Cox , M. Spasova , M. Farle , Nanotech-nology 2009 , 20 , 335301 .

[ 169 ] P. H. Beton , A. W. Dunn , P. Moriarty , Appl. Phys. Lett. 1995 , 67 , 1075 – 1077 .

[ 170 ] T. R. Ramachandrany , C. Baur , A. Bugacov , A. Madhukar , B. E. Koel , A. Requicha , C. Gazen , Nanotechnology 1998 , 9 , 237 .

[ 171 ] M. Martin , L. Roschier , P. Hakonen , U. Parts , M. Paalanen , B. Schleicher , E. I. Kauppinen , Appl. Phys. Lett. 1998 , 73 , 1505 – 1507 .

[ 172 ] J. L. O’Brien , S. R. Schofi eld , M. Y. Simmons , R. G. Clark , A. S. Dzurak , N. J. Curson , B. E. Kane , N. S. McAlpine , M. E. Hawley , G. W. Brown , Phys. Rev. B 2001 , 64 , 161401 .

[ 173 ] C. I. Pakes , D. P. George , S. Ramelow , A. Cimmino , D. N. D. N. Jamieson , S. Prawer , J. Magn. Magn. Mater. 2004 , 272–276 , E1231 – E1233 .

[ 174 ] M. J. Martínez-Pérez , E. Bellido , R. d. Miguel , J. Sesé , A. Lostao , C. Gómez-Moreno , D. Drung , T. Schurig , D. Ruiz-Molina , F. Luis , Appl. Phys. Lett. 2011 , 99 , 032504 .

[ 175 ] P. F. Vohralik , S. K. H. Lam , Supercond. Sci. Technol. 2009 , 22 , 064007 .

[ 176 ] S. K. H. Lam , W. R. Yang , H. T. R. Wiogo , C. P. Foley , Nano-technology 2008 , 19 , 285303 .

[ 177 ] M. Faucher , P. O. Jubert , O. Fruchart , W. Wernsdorfer , V. Bouchiat , Supercond. Sci. Technol. 2009 , 22 , 064010 .

[ 178 ] J. C. Garno , Y. Yang , N. A. Amro , S. Cruchon-Dupeyrat , S. Chen , G.-Y. Liu , Nano Lett. 2003 , 3 , 389 – 395 .

[ 179 ] L. Hao , C. Assmann , J. C. Gallop , D. Cox , F. Ruede , O. Kazakova , P. Josephs-Franks , D. Drung , T. Schurig , Appl. Phys. Lett. 2011 , 98 , 092504 .

[ 180 ] K. S. Novoselov , A. K. Geim , S. V. Morozov , D. Jiang , Y. Zhang , S. V. Dubonos , I. V. Grigorieva , A. A. Firsov , Science 2004 , 306 , 666 – 669 .

[ 181 ] A. K. Geim , Science 2009 , 324 , 1530 – 1534 . [ 182 ] S. Guo , S. Dong , Chem. Soc. Rev. 2011 , 40 , 2644 – 2672 . [ 183 ] S. Pisana , P. M. Braganca , E. E. Marinero , B. A. Gurney , Nano

Lett. 2010 , 10 , 341 – 346 . [ 184 ] A. K. Geim , K. S. Novoselov , Nat. Mater. 2007 , 6 , 183 – 191 . [ 185 ] E. Bellido , I. Ojea-Jiménez , A. Ghirri , C. Alvino , A. Candini ,

V. Puntes , M. Affronte , D. Ruiz-Molina , N. Domingo , unpublished.

Received: July 18, 2011 Revised: November 7, 2011Published online:

27www.small-journal.comH & Co. KGaA, Weinheim


Recommended