Sie sind auf Seite 1von 26

View Article Online / Journal Homepage / Table of Contents for this issue

Nanoscale

Dynamic Article Links <

Cite this: Nanoscale, 2012, 4, 4301

REVIEW

www.rsc.org/nanoscale

Published on 16 May 2012. Downloaded by Universidade Federal de Alagoas on 15/09/2015 15:05:40.

Luminescence nanothermometry
Daniel Jaque*a and Fiorenzo Vetrone*b
Received 30th March 2012, Accepted 14th May 2012
DOI: 10.1039/c2nr30764b
The current status of luminescence nanothermometry is reviewed in detail. Based on the main
parameters of luminescence including intensity, bandwidth, bandshape, polarization, spectral shift and
lifetime, we initially describe and compare the different classes of luminescence nanothermometry.
Subsequently, the various luminescent materials used in each case are discussed and the mechanisms at
the root of the luminescence thermal sensitivity are described. The most important results obtained in
each case are summarized and the advantages and disadvantages of these approaches are discussed.

A. Introduction
A.1

Fluorescence Imaging Group, Departamento de Fsica de Materiales C-04,


Insitituto Nicol
as Cabrera, Facultad de Ciencias, Universidad Aut
onoma
de Madrid, Madrid 28049, Spain. E-mail: daniel.jaque@uam.es
b

Institut National de la Recherche Scientifique-Energie,
Mat
eriaux et
T
el
ecommunications, Universit
e du Quebec, Varennes, QC, J3X 1S2,
Canada. E-mail: vetrone@emt.inrs.ca; Fax: +1-450-929-8102; Tel: +1514-228-6847
This work was supported by the Universidad Aut
onoma de Madrid
and Comunidad Aut
onoma de Madrid (Project S2009/MAT-1756), by
the Spanish Ministerio de Educacion y Ciencia (MAT2010-16161) and
by Caja Madrid Foundation.

Daniel Jaque obtained his Ph.D.


in Physics at Universidad
Aut
onoma de Madrid (Madrid,
Spain) working on Multifunctional Solid State Lasers. After
working for several years on the
optical and conducting properties of nanostructured superconducting and metallic thin
films, he co-founded the Fluorescence Imaging Group at the
Universidad
Aut
onoma
de
Madrid. His current research
Daniel Jaque
activity is focused on the development of fluorescent nanoparticles capable of thermal
sensing at the nanoscale. He has been Distinguished Invited
Professor at Heriot Watt University (UK) and Swinburne
University of Technology (Australia). He has published over 225
publications and attended as invited speaker to 16 international
conferences.

This journal is The Royal Society of Chemistry 2012

Why nanothermometry?

Nanothermometry aims to extract knowledge of the local


temperature of a given system with sub-micrometric spatial
resolution. Such knowledge is required for the complete understanding of micrometric and nanostructured systems whose
dynamics and performance are strongly determined by temperature. The recent development of nanotechnology and nanomedicine brought about the appearance of a large number of
such systems. It is difficult (if not even impossible) to enumerate
all the systems (fields) that presently exploit nanothermometry.
Nevertheless, we can initially identify three areas that can clearly

Fiorenzo Vetrone received his


Ph.D. in Chemistry at Concordia
University (Montreal, Canada)
followed by postdoctoral fellowships funded by the Natural
Science
and
Engineering
Research Council (NSERC) of
Canada and the Royal Society
(UK). Fiorenzo Vetrone is
currently an Assistant Professor
of Nanobiotechnology at Universit
e du Qu
ebec, Institut
National de la Recherche

Fiorenzo Vetrone
Scientifique-Energie,
Mat
eriaux
et T
el
ecommunications (INRSEMT) in Varennes, Canada
where his research activities are focused around multiphoton excited
luminescent nanoparticles for use in the development of multifunctional nanoplatforms for diagnostic and therapeutics of various
diseases. His scientific contributions, including over 50 publications
and 20 invited presentations, have been recognized by various
national and international awards.
Nanoscale, 2012, 4, 43014326 | 4301

Published on 16 May 2012. Downloaded by Universidade Federal de Alagoas on 15/09/2015 15:05:40.

View Article Online

make use of nanothermometry and can benefit significantly by


the further development of nanothermometry towards higher
sensitivities and resolutions. The three areas covered in great
detail in this review are micro/nano-electronics, integrated
photonics and biomedicine.
In micro/nano-electronics, the reduced dimensions of electrical
conduction channels lead to relevant resistances in such a way
that the Joule heating effect becomes non-negligible. Therefore,
any small variation in the local resistance of the conduction
channels (caused by either a local modification of the micro/
nano-conductor geometry or by the existence of scattering
defects within it) could cause the appearance of a hot-spot, i.e.
a well-localized temperature increment. These hot spots can be
the catalyst for performance deterioration or ultimately, an
irreversible failure.1,2 The knowledge of the exact location of such
thermal singularities is crucial to avoid catastrophic damage,
which can be averted by either design modifications or by the
controlled incorporation of micro-cooling devices.3 Unfortunately, the magnitude and location of these temperature singularities are difficult to be predicted a priori, since they depend not
only on the device design but also on the quality of the integrated
circuits. Thus in micro/nano-electronics, nanothermometry
could become a pivotal tool for the detection of hot-spots in
integrated circuits during real operating conditions.
In integrated photonic devices, temperature is also a critical
parameter where the performance of such systems depends on
the optical properties of the constituting materials. For example,
light control and manipulation in integrated photonic devices are
achieved by the sub-micrometric controlled variation of the local
refractive index, which is a physical property that is strongly
temperature dependent.4 In fact, temperature not only affects the
refractive index but also the optical gain of the integrated laser
sources, spectral properties of the band-gap structures and
finally, the optical conversion efficiency of nonlinear converters
such as optical super-lattices.5 In this context, any variation in
the local temperature within an integrated photonic device could
strongly affect its performance or in the worst case, could even
lead to irreversible damage.6 Again, using nanothermometry to
glean knowledge of temperature singularities induced during
device operation is paramount for the development of reliable,
efficient and robust integrated photonic devices, making nanothermometry an indispensable tool. This can be extended to the
case of new generation optofluidic devices in which light is used
to manipulate micron sized objects (including living specimens),
which are propagating within micro-fluidics through the use of
optical forces.7 In this case, residual absorption of fluids at laser
radiation wavelengths can cause heat loading and can ultimately
lead to malfunctions of the device resulting from the onset of
convection currents.8
Another area that could benefit greatly from the high-resolution and high-sensitivity of nanothermometry is biomedicine. It
is well-known that in any biosystem, temperature is known to
play a crucial role in determining its dynamics and properties.9,10
For example, temperature is one of the critical parameters
determining cell division rates and hence, determining the rate of
tissue growth.11 Moreover, it also drastically affects the
mechanical, optical and structural properties of fundamental
biomolecules such as proteins where they can undergo a denaturation process when their temperature deviates by even a few
4302 | Nanoscale, 2012, 4, 43014326

degrees above 37  C.12 Thus, to understand the dynamics of


biosystems, the simultaneous monitoring of their temperature is
crucial to elucidate the origin of the observed behavior. Aside
from this fundamental interest, thermal sensing of biosystems is
also vital for the early detection and treatment of many diseases.
It is commonly accepted that one of the first signatures of any
given illness (such as inflammation, cancer or cardiac problems)
is the appearance of thermal singularities.13 In the particular case
of cancer, it has been postulated that the thermal singularity
associated with incipient tumors becomes detectable when they
reach a size consisting of thousands of cancer cells, i.e. when the
tumor size is well below 1 mm. Exploiting this phenomenon for
cancer detection would constitute a major leap forward when
compared with traditional imaging techniques such as tomography, for which the minimum detectable tumor size is on the
order of several millimeters.14 This is especially important since
the early detection of tumors would allow for significantly earlier
medical intervention. High-resolution thermal sensing is also
required in cancer therapeutic processes such as hyperthermia,
which is based on externally inducing an increment in the tumors
temperature up to cytotoxic levels (4345  C).15 This temperature
increase should be performed in a controlled manner such that
the thermally induced damage of surrounding tissues would be
minimized. This can only be accomplished if the temperature of
the tumor can be continuously monitored during the hyperthermia treatment such that it could be immediately terminated
when cytotoxic levels are reached, thereby avoiding excessive
(and unnecessary) heating. This thermal sensing platform is
quickly becoming a reality with the recent development of
nanothermometers, which due to their reduced size can be
incorporated into tumors and cancer cells and subsequently
provides the requisite thermal information.
Finally, it should be noted that a universal thermal sensor does
not exist since different applications require different properties.
For example, the requirements that must be satisfied by nanothermometers used in biological applications are very different
from those necessary for the thermal imaging of opto-fluidic
devices. In bio-medicine, the nanothermometers should be nontoxic, water soluble and very stable under light irradiation so that
toxic components are not delivered to the cells. In addition, the
final temperature resolution achievable should be on the order of
0.2  C since temperature variations as low as 1  C are already
relevant in biological dynamics. In the case of thermal imaging of
opto-fluidic devices the typical temperature changes caused by
laser radiation are on the order of several degrees so that the final
temperature resolution of the nanothermometers is not required
to be below 1  C. In opto-fluidics the toxicity is not a critical issue
but, due to the high laser intensities achieved within microchannels, the physical and chemical stability with respect to laser
irradiation becomes an essential aspect.
A.2

Nanothermometry techniques: a brief summary

Motivated by the numerous fundamental and practical applications, many research groups have presently focused their efforts
on the development of diverse techniques capable of achieving
thermal sensing with sub-micrometric resolution.16 This has led
to the development of several prospective avenues for nanothermometry and consequently, to the elaboration of several
This journal is The Royal Society of Chemistry 2012

Published on 16 May 2012. Downloaded by Universidade Federal de Alagoas on 15/09/2015 15:05:40.

View Article Online

nanothermometry techniques. These techniques have been


traditionally classified into three main groups: electrical,
mechanical, and optical.
The working principles of electrical techniques are the same as
those governing traditional thermocouples and thermistors, in
which the temperature reading is achieved from variation of the
resistance, voltage, conductivity or electrical capacity of a given
conducting probe.17 These traditional principles have also been
applied in nanothermometry due to the rapid development of
micro/nano-fabrication techniques, which have driven the
dimensions of the electrical thermal sensor down to the submicrometric scale.18 By using the same principles of atomic force
microscopes, these sub-micrometric thermal sensors were scanned (in pseudo-contact) over the surface to be thermally imaged
(known as scanning thermal microscopy).19 The spatial resolution reached using this technique can, in principle, be as good as
that of atomic force microscopes. Indeed, micro-thermocouples
based on PtCr thin films have been demonstrated to be capable
of thermal sensing with sub-100 nm resolution over areas of
hundreds of mm2.18 The scanning area could be increased up to
a few mm2 by utilizing two-dimensional arrays of micro-thermocouples although in this case serious difficulties in the
microfabrication procedure could be problematic.20 Despite
these promising results, there are several potential drawbacks in
using electrical nanothermometry. First, it is a surface technique
and thus can be only used to obtain two-dimensional thermal
images. In addition, the technique requires a physical contact
between the thermal probe and the system under study, which
can lead to complex heat transfer fluxes ultimately resulting in
a modification of the system temperature due to the interactive
measurement procedure. Finally, the advanced technology
required for the fabrication of the thermal probes increases the
costs of the required experimental set-up.
Mechanical nanothermometry is similar to micro-thermocouple based scanning thermal microscopy but replaces the
micro-thermocouple with a bi-material cantilever. Due to the
different mechanical properties of the two materials that make
up the cantilever, any temperature change in its surroundings
results in cantilever deflection. Based on this simple principle,
thermal sensitivities reaching the mK range have been demonstrated while maintaining the spatial resolutions typical of
atomic force microscopes.21 As in the case of electrical nanothermometry, mechanical nanothermometry can be only used to
obtain thermal images of surfaces and does not offer the possibility of reconstructing three-dimensional thermal images. In
addition, due to the requirement of physical contact between the
thermal probe and the specimen, the thermal measurement can
also be affected by artificially induced heat fluxes.22
Optical nanothermometry is based on the analysis of temperature-induced changes in the inherent passive optical properties
of materials. Optical nanothermometry includes a great variety
of techniques. Some of them are based on temperature-induced
changes in the optical path length of a transparent system due to
the temperature dependence of both refractive index and thickness.17,23 Interferometric techniques exploit these thermally
induced changes to reconstruct two-dimensional images of
transparent materials with spatial resolutions close to the micron
and with thermal sensitivities that in the best of the cases can
surpass the mK limit.23 The thermal dependence of the refractive
This journal is The Royal Society of Chemistry 2012

index consequently makes the reflectance of any interface sensitive to temperature variations. This is exploited by reflectance
based thermometry techniques that relate temperature to the
polarization and intensity of light reflected from a given surface.
In combination with high numerical aperture optics and
adequate choice of illumination wavelengths, real time thermal
images of micro-resistors and micro-coolers with sub-micrometric spatial resolutions and with thermal sensitivities of few
tens of mK have already been obtained.24 One of the limitations
of this technique is that it requires the existence of an optical
interface (surface) and consequently, can be only used for twodimensional thermal imaging. Traditionally, scanning Raman
thermal microscopy has been also cataloged as an optical
method. It is based on the analysis of the properties of Raman
(vibration) modes that are affected by temperature variations
through thermally induced structural changes. Raman thermal
microscopy, therefore, requires the scanning of an excitation
laser beam over the system to be thermally imaged. A subsequent
spectral analysis provides the Raman images of the system in
terms of the intensity, position and width variations of the
Raman modes, which are in turn translated into temperature
units. This procedure has been successfully applied to acquire
thermal images of micro-heating devices and high power integrated laser devices achieving thermal and spatial resolutions no
better than 10  C and 1 mm, respectively.17,25,26 Raman thermal
imaging appears to be a universal technique since any material
shows a Raman spectrum under laser excitation. In practice,
however, this technique is only applicable to those materials
showing large Raman efficiencies and that present high thermal
sensitivity of the Raman modes restricting it to a few number of
systems.
Finally, the Pyrometric and Infrared (IR) Thermometry
techniques should be also mentioned. Both are based on the
measurement of the electromagnetic radiation emitted from
a given surface at a specific spectral range (typically long
wavelengths). The temperature of the surface can then be
obtained by analyzing its emissivity within the Plancks blackbody theory. Again, this is a universal technique that has
already been applied to obtain thermal images of living specimens, optically pumped laser sources, integrated microelectronics and fluids. However, it too has several drawbacks such
as the impossibility of obtaining three-dimensional images and
the requirement of using long wavelengths (>1 mm).2729 This
last point necessitates that expensive infrared detectors are
required and obviously limits the final spatial resolution that
can be achieved (ranging from 5 to 30 mm).
A.3

Luminescence nanothermometry

Luminescence is the emission of light from a given substance,


occurring from electronically excited states that have been
populated by an external excitation source (optical radiation, in
the case of photoluminescence). The properties of the emitted
photons depend on the properties of the electronic states
involved in photon emission.30 These, in turn, depend on the
local temperature and thus luminescence nanothermometry
exploits the relationship between temperature and luminescence
properties to achieve thermal sensing from the spatial and
spectral analysis of the light generated from the object to be
Nanoscale, 2012, 4, 43014326 | 4303

Published on 16 May 2012. Downloaded by Universidade Federal de Alagoas on 15/09/2015 15:05:40.

View Article Online

thermally imaged. The grouping of luminescence nanothermometry into different classes is based on the particular
parameter of luminescence which is analyzed and from which the
thermal reading is ultimately extracted. Fig. 1 schematically
depicts the six parameters that define the luminescence emission
of a given material: intensity, band-shape, spectral position,
polarization, lifetime and bandwidth. Fig. 1 also qualitatively
demonstrates how the luminescence emission spectrum is modified when each of these parameters is varied. Thus, based on
these variations it is possible to define the following luminescence
nanothermometry sub-classes:
Intensity Luminescence Nanothermometry (ILNth). In this
case, thermal sensing is achieved through the analysis of the
luminescence intensity. When temperature changes, there is an
overall change in the number of emitted photons per second such
that the emission spectrum becomes less (or more) intense.
Temperature induced changes in the luminescence intensity are
generally caused by the thermal activation of luminescence
quenching mechanisms and/or increases in the non-radiative
decay probabilities.
Band-Shape Luminescence Nanothermometry (BSLNth). The
term band-shape refers to the relative intensity between the
different spectral lines that make up the luminescence spectrum.
Thermally induced variations in the band-shape usually take
place when the electronic states from which emission is generated
are very close in energy such that they are thermally coupled. It
can be also present in mixed systems, i.e. systems containing
more than one class of emitting centers.

Fig. 1 Schematic representation of the possible effects caused by


a temperature increment on the luminescence. Red lines correspond to
higher temperatures.

4304 | Nanoscale, 2012, 4, 43014326

Spectral Luminescence Nanothermometry (SLNth). It is based


on the analysis of the spectral positions of the emission lines,
which are unequivocally determined by the energy separation
between the two electronic levels involved in the emission. In
turn, this depends on a large variety of temperature dependent
parameters of the emitting material including refractive index
and inter-atomic distances (density). Thus, in any emitting
material the spectral positions of the luminescence lines are
expected to be temperature dependent, and this is exploited by
spectral luminescence nanothermometry to translate spectral
shifts into temperature.
Polarization Luminescence Nanothermometry (PLNth). In
anisotropic media, the emitted radiation is generally non-isotropically polarized and consequently, the shape and intensity of
emitted radiation are strongly dependent on its polarization. This
allows for the definition of the polarization anisotropy
parameter, which is the ratio between the luminescence intensities emitted at two orthogonal polarization states. As a result,
polarization luminescence nanothermometry is based on the
influence of temperature on this polarization anisotropy.
Bandwidth Luminescence Nanothermometry (BLNth). The
width of the various emission lines that make up any luminescence spectrum is determined by the properties of the material
(such as the degree of disorder) and temperature. It is wellknown that as the temperature of a luminescent material is
increased, a corresponding increase in the density of phonons
occurs resulting from the spectral contribution of homogeneous
line broadening. Generally, in the vicinity of room temperature,
homogeneous line broadening leads to a linear relationship
between bandwidth and temperature. The change in the bandwidth of the luminescence spectra is exploited in bandwidth
luminescence nanothermometry to achieve a thermal reading.
Lifetime Luminescence Nanothermometry (LLNth). Luminescence lifetime, sf, is defined as the time that the emitted luminescence intensity decays down to 1/e of its initial value after
a pulsed excitation. This is an indication of the total decay
probability of the emitted intensity (indeed this probability is
defined as the inverse of the luminescence lifetime). Decay
probabilities from electronic levels depend on a great variety of
factors and many of them are related to temperature (such as
phonon assisted energy transfer processes and multiphonon
decays). This temperature dependence makes it possible to
extract temperature readings from the determination of the
luminescence lifetime.
Thus, luminescence nanothermometry provides several
options to achieve thermal sensing from the analysis of the
emission spectrum generated by the system under study. It
should be noted that the thermal sensitivity would vary from
system to system and would obviously depend on the magnitude
of the thermally induced spectral variations. Clearly, systems
that show remarkable changes in luminescence properties for
small temperature changes will provide the largest thermal
sensitivities. Similarly, spatial resolution in luminescence nanothermometry is difficult to be estimated as it is determined by the
spatial resolution at which the emission spectrum is recorded.
High spatial resolution would require the active luminescence
volume to be as small as possible, i.e. it requires acquisition of the
luminescence emission spectrum with high spectral resolution.
For example, if the luminescent system to be thermally imaged
This journal is The Royal Society of Chemistry 2012

Published on 16 May 2012. Downloaded by Universidade Federal de Alagoas on 15/09/2015 15:05:40.

View Article Online

possesses luminescent centers that are homogeneously distributed, achieving high spatial resolution would require the use of
confocal luminescence microscopy (based on the combination of
high numerical optics and spatial filters) or of near-field optical
microscopes (based on the spatial scanning of low aperture
optical fibers). On the other hand, luminescence nanothermometry of non-luminescent systems (such as fluids, electrical circuits or living cells) is also possible yet it requires the
introduction of luminescent micro- and nanoparticles with
a temperature dependent luminescence within the system to be
imaged. Ultimately, the spatial resolution of the thermal
measurement would be limited by the spatial dimension of the
luminescent particle (provided it has been discretely incorporated
into the system). Recent achievements in nanoparticle synthesis
have made it possible to prepare highly efficient nanoparticles
(sizes down to few nanometers), which after proper surface
functionalization, can be dispersed in solutions allowing for their
facile incorporation in different systems.
Resulting from this great diversity in luminescence nanothermometry, numerous approaches and experimental demonstrations of sub-micrometric thermal images of a great variety of
systems have appeared. In the next section we review the most
relevant ones where we classify them in terms of the luminescence
nanothermometry sub-classes. Moreover, the different thermal
and spatial resolutions achieved by the different approaches are
compared and discussed.

B. Luminescence thermal images


B.1 Intensity luminescence nanothermometry
As described in Section A.3, intensity luminescence nanothermometry is based on the thermal variation of the luminescence intensity to achieve a temperature reading. The basic
elements of this technique are luminescence systems with
a hypersensitive thermal response, i.e. whose luminescence
intensity is strongly affected because of small temperature variations. Previous work reporting on thermal imaging and sensing
based on luminescence intensity analysis can be grouped differently depending on the nature of the luminescence centers used as
thermal probes. ILNth has been previously demonstrated in
different systems including semiconductor nanocrystals (hereafter quantum dots, QDs), luminescent organic dyes, rare earth
ions, transition metal ions, luminescent nanogels and luminescent polymers. In the following sections we will summarize the
most relevant results obtained for each case.
B.1.1 QD-based intensity luminescence nanothermometry.
Semiconductor QD nanocrystals are perhaps one of the most
ubiquitous optical nanoprobes due to their excellent photostability, large luminescence quantum yield and also because of
their size-tunable absorption and emission wavelengths.31
Recombination of electronhole pairs within the QD volume
gives rise to a typical near Gaussian emission band whose peak
wavelength depends on the QD material and the QD size.32 Most
of the QDs used for imaging applications have their averaged
emission wavelength in the visible range and as can be observed
in Fig. 2(a), this emission band is strongly modified by temperature variations.31 One of the most remarkable changes is
This journal is The Royal Society of Chemistry 2012

Fig. 2 (a) Luminescence of CdSe QDs as obtained at different temperatures. The inset shows the integrated luminescence intensity as a function of temperature. Dots are experimental data and the solid line is
a linear fit. Data were extracted from ref. 33. (b) Scanning electron
microscope image of a 200 nm sized silica sphere covered with CdSe/ZnS
QDs. Figure 2(a) is reprinted with permission from G. W. Walker, V. C.
Sundar, C. M. Rudzinski, A. W. Wun, M. G. Bawendi and D. G. Nocera,
Appl. Phys. Lett., 2003, 83, 3555. Copyright 2003, American Institute of
Physics. Figure 2(b) is reprinted with permission from L. Aiguoy, B.
Samson, G. Julie, V. Mathet, N. Lequeux, C. N. Allen, H. Diaf and B.
Dubertret, Rev. Sci. Instrum., 2006, 77, 063702. Copyright 2006, American Institute of Physics.

produced in the luminescence intensity where as the temperature


increases the QD emission becomes weaker.33 This seems to be
a universal phenomenon in QDs and is attributed to the activation of phonon assisted processes as well as to the presence of
thermally assisted energy transfer processes from bulk to surface
(non-radiative) states. A very interesting feature of thermally
induced luminescence quenching in QDs is its linearity. As can be
observed in the inset of Fig. 2(a), the emitted intensity of the QD
decreases almost linearly in the vicinity of room temperature.
Although the results included in Fig. 2 correspond to CdSe/ZnS
QDs, similar linearities have been also observed in other systems
such as CdTe QDs.34,35 Linearity is an outstanding feature for
thermal sensing since it ensures a constant thermal sensitivity in
the whole working temperature range of the thermal probe.
Indeed a very simple calibration procedure is required to obtain
thermal sensing from the analysis of the QD emission intensity.
There are two different approaches to obtain luminescence
images based on the analysis of QD luminescence.
The first approach relies on the controlled scanning of a single
QD over the system to be thermally imaged while recording its
luminescence intensity. This approach was first demonstrated by
Aigouy et al. who modified an AFM tip with a subwavelengthsized silica sphere covered with CdSe/ZnS QDs (see Fig. 2(b)).36
The continuous monitoring of QD luminescence and the
Nanoscale, 2012, 4, 43014326 | 4305

Published on 16 May 2012. Downloaded by Universidade Federal de Alagoas on 15/09/2015 15:05:40.

View Article Online

subsequent translation into temperature units could generate the


thermal image of the area under scanning. This technique has
been previously used for optical transmission of sub-micrometric
objects and its application in thermal sensing and imaging is
imminent. The main disadvantage of this approach is that it is
a pseudo-contact method so that it is not clear how the vicinity of
the tip can modify the actual temperature distribution. In addition, this technique is restricted to surface sensing.
The second approach involves a reduced technological
complexity and involves the artificial incorporation of the QDs
into the studied system. The spatial variation of the QD luminescence intensity (that is recorded by a conventional or by
a scanning confocal microscope) would allow for thermal
imaging to be achieved. A representative example of this
approach is the pioneering work of Han et al. who used CdSe
QDs to monitor the temperature of cancer cells incubated in
a mixed solution of QDs and metallic nanoparticles.34 The
authors designed a double beam experiment as that shown in
Fig. 3(a). One near-infrared (NIR) laser beam (820 nm) was used
as the heating source since it promoted the excitation of surface
plasmons of the metallic nanoparticles. A second light source
(halogen lamp) illuminated the whole incubation dish. The
subsequent QD luminescence intensity was recorded by a CCD
camera and was used to obtain thermal images of the cell population during NIR illumination. An example is given in Fig. 3(b)
where it is evident that at the central region (where the NIR laser
spot is focused) the QD intensity is decreased. This is a clear
indication that the metallic nanoparticle mediated laser-induced
local heating occurs (see the darker region indicated by a black

Fig. 3 (a) Schematic of the experimental set-up used for QD based


ILNth of a cell culture during gold nanoparticle mediated heating by an
infrared laser beam, (b) QD intensity luminescence image obtained
during the heating process (20 s after starting). The black arrow indicates
the position of the heating laser spot. Note that the intensity reduces
radially with respect to the focus. (c) Micrographs of darkfield (left) and
luminescent field (right) indicating that the cells are still attached to the
dish after heat treatment. The luminescence image reveals cell death in the
surroundings of the laser heating focus. Reprinted with permission from
ref. 34. Copyright 2009 Springer.

4306 | Nanoscale, 2012, 4, 43014326

arrow in Fig. 3(b)). After a proper calibration, the authors


concluded that the maximum temperature at the focal point was
as large as 50  C. This temperature increment was high enough to
produce cell ablation at the focal point as can be observed in
Fig. 3(c). This was pioneering work since it demonstrated the first
application of QD based luminescence intensity nanothermometry in a biosystem. The final temperature resolution
achievable by using QDs is determined by the temperatureinduced luminescence quenching rate. In the case of CdSe QDs,
this quenching rate is close to 1% per  C such that sub-degree
thermal resolutions are easily achievable.
An alternative approach that, because of its novelty, deserves
to be mentioned in detail was proposed by Lee et al.37 They
described a reversible nanothermometer built from two types of
nanoparticles connected by a polymer that was acting as
a molecular spring. Fig. 4(a) shows a schematic representation of
the system proposed by Lee et al. The nanoscale superstructure
consisted of an Au nanosphere covered by a poly(ethyleneglycol)
(PEG) film with a thickness of few nanometers. The outer side of
the PEG film was conjugated with 3.7 nm CdTe QDs and they
were excited through plasmonexciton interactions. The nanothermometer displayed the characteristic exciton luminescence of
CdTe QDs (at approximately 550 nm) following optical excitation of the surface plasmon resonance of the Au nanoparticle (at
around 633 nm). The efficiency of this plasmonexciton energy
transfer is strongly dependent on the distance between the Au
surface and CdTe QDs, i.e. on the PEG film thickness. Since
temperature variation from 20 to 50  C results in an expansion of
the PEG molecule, the PEG thickness varies substantially within
this temperature range, producing a concomitant change in the
luminescence output (it becomes temperature dependent). Lee
et al. demonstrated the ability of the superstructure depicted in
Fig. 4(a) for ILNth by analyzing the luminescence intensity of the
AuPEG QD system when it was subjected to a time dependent
temperature variation as schematically indicated in Fig. 4(b). As
can be observed in Fig. 4(c), the CdTe QD luminescence intensity

Fig. 4 (a) Schematic representation of the superstructure proposed by


Lee et al. for nanoscale thermometry that consists of a metallic nanoparticle covered with a PEG film and CdTe QDs. (b) Time evolution of
the superstructure temperature and (c) corresponding time evolution of
the CdTe luminescence intensity. Data were extracted with permission
from ref. 37.

This journal is The Royal Society of Chemistry 2012

View Article Online

Published on 16 May 2012. Downloaded by Universidade Federal de Alagoas on 15/09/2015 15:05:40.

is shown as a function of time and the AuPEG QD responds


instantaneously to the temperature variations. From Fig. 4(b)
and (c) it is possible to estimate a thermally induced luminescence
quenching rate of 0.6% per  C for the AuPEG QD system that is
quite close to that of QDs under direct exciton excitation.
B.1.2 Dye-based intensity luminescence nanothermometry.
The term dye refers to organic compounds that show a strong
luminescence (usually in the visible) when optically excited with
short wavelength radiation (usually in the ultraviolet or blue).
Light emission is generated by the electronic transitions taking
place between vibrational energy levels of different electronic
states. The emission and absorption bands of dyes are both
strongly dependent on the particular composition of the
compound. Furthermore, for a particular dye, the spectral
properties of the luminescence band depend on many factors,
such as the solvent, concentration, pH and temperature. As
a general rule, the luminescence intensity generated by organic
dyes decreases as the temperature increases. This can be
explained in terms of the activation of multiphonon de-excitations as well as the thermal promotion of electrons up to vibrational levels of higher energies, from which radiative deexcitation probability is reduced.38 The magnitude of the thermal
dependence varies from dye to dye where some dyes have a much
stronger sensitivity to temperature. This is the case of Rhodamine B, which belongs to the xanthene group and possesses
a luminescence band centered around 550 nm (see Fig. 5(a)).
Rhodamine is probably the most popular dye as a result of its
superior chemical stability as well as its relatively high luminescence efficiency. Fig. 5(b) shows the luminescence intensity
generated from a UV excited Rhodamine B solution as a function of the solution temperature. As can be observed, temperature induces a remarkable luminescence quenching. Close to
room temperature (25  C), the luminescence intensity reduces
linearly with temperature at a rate close to 2% per  C, which is
similar to that found for QDs. There are several reports in the

Fig. 5 (a) Room temperature luminescence spectrum of Rhodamine B.


(b) Integrated emission of Rhodamine B as a function of temperature.
Dots are experimental data and the solid line is a guide for the eyes.

This journal is The Royal Society of Chemistry 2012

literature dealing with thermal imaging from the analysis of the


luminescence intensity variations of Rhodamine B. They can be
classified into two types, depending on how the Rhodamine was
introduced in the imaged system: liquid solution or thin film.
In the first case an aqueous solution containing Rhodamine B
was introduced into a microfluidic device that was optically
excited at 525 nm. The spatial variation of the dye emission
intensity was recorded by placing the microfluidic device in
a conventional luminescence microscope. Ross et al. demonstrated that using this facile technique they were able to measure
the temperature of the microfluidic device with sub-micrometric
spatial resolution.39 In fact, the authors were able to measure
two-dimensional temperature distributions resulting from Joule
heating in a variety of electrokinetically pumped microfluidic
systems. Fig. 6(a) shows the thermal image of a multi-branched
microfluidic circuit that is schematically shown in Fig. 6(b) where
a voltage difference of 1000 V was applied between the two fluid
reservoirs (light gray circles). The area enclosed by the dashed
line in Fig. 6(b) is the field of view of the thermal image where
Ross et al. stated that the maximum temperature resolution
achievable with their technique could surpass 0.1  C.39 These
outstanding thermal sensitivities allowed the observation of
complex phenomena, in which each branch of the fluid circuit
carries a different fraction of electric current and, therefore,
suffers from different thermal loading.
In the second approach, a thin film containing the luminescent
Rhodamine dye was deposited onto the surface to be thermally
imaged. The physical contact between the surface and film
creates a temperature distribution on the film that reproduces the

Fig. 6 (a) Color coded thermal image of a multi-branched microfluidic


circuit. (b) Schematic showing a layout of the multi-branched circuit.
Reproduced from ref. 39 with permission from The Royal Society of
Chemistry.

Nanoscale, 2012, 4, 43014326 | 4307

Published on 16 May 2012. Downloaded by Universidade Federal de Alagoas on 15/09/2015 15:05:40.

View Article Online

surface of interest. Thus, surface thermal imaging was achieved


from the visualization of the luminescence intensity of the thin
film. This approach has been applied to obtain thermal images of
microfluidic systems subjected to Joule heating as well as microand nanowires under real operation conditions.40,41 The latter
case is of special relevance because of the low dimensions of the
heating source. Fig. 7(a) shows the basic experimental set-up
used by L
ow et al. where a nickel nanowire is covered by
a Rhodamine containing thin film that is optically excited in the
530550 nm range with a mercury lamp.40 The film luminescence
(containing thermal information through its intensity variations)
was collected by a microscope objective and analyzed by a CCD
detector. In the absence of applied voltage, the surface temperature is homogeneous and the luminescence intensity of the thin
film is constant over the entire surface. Conversely, when an
electrical current flows thorough the nanowire (Fig. 7(b)), Joule
heating appears and the thin film surface temperature is no
longer homogeneous, creating a reduction in the luminescence at
the nanowire location (see the luminescence image included in
Fig. 7(b)). From the analysis of the decrease in luminescence
intensity, it is possible to determine the local temperature of the
nanowire (as it is schematically illustrated in Fig. 7(c)). In this
case the thermal image corresponds to that of an 80 mm-long, and
2 mm-wide nickel wire when a current of 20 mA is flowing. The
thermal image clearly demonstrates that this technique allows for
the measurement of local temperatures in excess of 90  C while
keeping sub-micrometric spatial resolution. One of the clear
drawbacks of using luminescent thin films is that they allow only

Fig. 7 (a) Depiction of a nanowire on a Rhodamine B thin film. (b)


Spatial variation of the Rhodamine B emission intensity when a 20 mA
current is applied along the nanowire (at the center of the image). (c)
Thermal image of the nanowire as obtained from the luminescence of the
image shown in (b). Reproduced from ref. 40 with permission.

4308 | Nanoscale, 2012, 4, 43014326

for the thermal imaging of surfaces, i.e. 2D thermal imaging.


Moreover, it should be noted that the existence of a physical
contact between the sensing film and the system under investigation could modify the actual surface temperature such that the
thermal image could actually differ from that of the device in
normal working conditions, i.e. in the absence of the sensing film.
B.1.3 Rare earth-based intensity luminescence nanothermometry. Rare earth atoms lie within the sixth row of the
periodic table, from cerium to ytterbium, inclusive. The electronic configuration of trivalent ions is considered as a singularity since it shows an unfilled 4f shell, partially screened from
the environment by the electrons in the 5s2 and 5p6 shells.30
Optical transitions involving different levels within the 4f shell
are normally forbidden. However, they may become partially
allowed when the ions are located within a matrix, by virtue of
the crystal field of the local environment. Due to the partial
electronic screening from the environment, the luminescence
lines of rare earth ions are typically very narrow and the corresponding lifetimes are relatively long, from a few ms to a few ms.
Due to these properties, rare earth ions have been used in low
threshold, high gain solid state lasers. The intensity of the
luminescence lines of rare earth ions depends on several
parameters among which, temperature is one of the most critical
ones.30 There are a large number of mechanisms that could link
the intensity of rare earth luminescence with temperature
including activation of the multiphonon decay probability,
activation of energy transfer between rare earth ions or
quenching centers, population re-distribution due to Boltzmann
statistics, appearance of phonon assisted Auger conversion
processes and thermal enhancement of energy transfer processes
between rare earth ions and the host levels or charge transfer
states.30
Due to the simultaneous interaction of those different
phenomena, it is very difficult to predict and understand how the
luminescence intensity of rare earths is influenced by temperature. Indeed the thermal behavior of different lines could differ
by a great amount to that observed for other lines. This is
illustrated in Fig. 8 that includes the change in intensity observed
for the various lines of the 5Dj / 7Fi transitions of Eu3+ doped in

Fig. 8 Schematic representation of the energy level diagram of trivalent


europium (Eu3+) ions in La2O2S (left). The figure on the right shows the
temperature dependence of the luminescence intensity associated with the
different transitions indicated by the arrows on the energy level diagram.
Data extracted from ref. 44 with permission from the American Institute
of Physics.

This journal is The Royal Society of Chemistry 2012

Published on 16 May 2012. Downloaded by Universidade Federal de Alagoas on 15/09/2015 15:05:40.

View Article Online

La2O2S.4244 This demonstrates that each line is affected in


a different temperature range, so that this particular compound
can be used to detect temperature changes in different ranges
from 100 to 500 K.
The inherent ability of Eu3+ ions for temperature sensing in the
vicinity of room temperature has been previously exploited for
the microscopic detection of thermogenesis in single living cells.
The idea proposed by Suzuki et al. was to establish a physical
contact between a HeLa cancer cell and a glass micropipette filled
with an Eu3+ containing solution (see Fig. 9(a)).9 This compound
suffers from a relevant temperature induced luminescence
quenching in the biophysical range 2050  C following optical
excitation in the UV, as can be observed in Fig. 9(b). The
intensity reduction rate was found to be close to 3% per  C,
superior to that previously described for luminescent dyes, such
as Rhodamine. The intensity of the visible emission of the Eu3+
compound is registered in real time so that the authors were able
to follow the time evolution of the cells temperature. This system
was used to detect, at the single cell level, the intracellular
temperature changes caused by a modification in the intracellular
free Ca2+ concentration. Fig. 9(c) shows the time variation of the
EuTTA luminescence intensity. The arrow indicates the
moment at which the Ca2+ concentration was externally varied

by adding a 1 mM Ca2+ solution and the arrowhead indicates the


moment at which the initiation of positive thermogenesis was
detected. The system proposed by Suzuki et al. was able to detect
real-time intracellular temperature variations as small as 1  C in
the biophysical range.9

Fig. 9 (a) Schematic illustration of the set-up used by Suzuki et al. for
the measurement of the temperature of a single cell by using pipettes filled
with a luminescent Eu3+ solution. (b) Temperature variation of the
luminescence generated by the Eu3+ containing solution. (c) Time variation of the Eu3+ luminescence intensity and cell temperature variation.
The arrow indicates the moment at which the Ca2+ concentration was
externally varied. The arrowhead indicates the moment at which the
initiation of positive thermogenesis was detected. Reprinted with
permission from ref. 9. Copyright 2007 Elsevier.

Fig. 10 (a) Depiction of the luminescent nanogel proposed for ILNth by


Gota et al. indicating the temperature induced shape change in the
nanogel. (b) Luminescence intensity generated by the nanogel schematically shown in (a) as a function of temperature. (c) Phase contrast and
luminescence images of living COS7 cells containing the nanogel of (a) at
three different culture medium temperatures. (d) Time evolution of
intracellular luminescence intensity and temperature after the addition of
camptothecin to the culture medium at t 0. Reprinted with permission
from ref. 45. Copyright 2009 American Chemical Society.

This journal is The Royal Society of Chemistry 2012

B.1.4 Luminescent nanogel-based intensity luminescence


nanothermometry. The potential use of luminescent nanogels
(system typically composed of several synthetic polymers or
biopolymers which are chemically or physically cross-linked) as
nanothermometers was initially proposed and demonstrated by
C. Gota et al. in 2009.45 The base of their temperature sensitive
nanogel was the thermoresponsive poly-NIPAM combined with
the DBD-AA water sensitive fluorophore. The resulting
compound was treated by an emulsion polymerization technique
using a cross-linker (MBAM). The working principles of the
nanogel based thermometer are schematically shown in
Fig. 10(a). At low temperature (below 30  C) the nanogel swells
by absorbing water into its interior and this inner water causes
a luminescence quenching of the DBD-AA units. When the
nanogel temperature is raised above 30  C the nanogel shrinks,
causing the release of water molecules. As a consequence, the
DBD-AA units are no longer quenched and the nanogel luminescence increases remarkably. This is illustrated in Fig. 10(b) in
which the 515550 nm luminescence intensity generated from the
nanogel under 488 nm excitation is shown in the 2040  C range.
At first glance, the luminescent nanogel has the outstanding
characteristic of a luminescence enhancement with temperature.
This is at variance with the previously described cases (such as
QDs and rare earth ions) allowing for the measurement of
heating processes without a relevant loss of luminescence intensity and, hence, of thermal sensing accuracy. A particular
disadvantage is that the nanogel intensity becomes highly
temperature sensitive only in a reduced temperature range (27
35  C). Although this limits its application to a very reduced
temperature range, it allows for the investigation, for example, of

Nanoscale, 2012, 4, 43014326 | 4309

Published on 16 May 2012. Downloaded by Universidade Federal de Alagoas on 15/09/2015 15:05:40.

View Article Online

intracellular temperature variations. This possibility was


demonstrated by Gota et al. and their results are summarized in
Fig. 10(c) that shows the phase contrast and luminescence images
(at three different culture medium temperatures) of living COS7
cells containing the nanogel.45 It is observed that the nanogel
emission was strong enough to provide luminescence images only
at medium and high temperatures (29 and 35  C, respectively), in
accordance with the calibration curve shown in Fig. 10(b).
Fig. 10(d) shows the time evolution of intracellular luminescence
intensity and temperature after the addition of camptothecin to
the culture medium at t 0, which is expected to cause intracellular heating due to the early response of apoptosis such as
DNA cleavage. Gota et al. were able to detect this intracellular
heating and found that the pattern of temperature variation was
different among the cells, a behavior that they explained in terms
of the cell-division-dependent effects of camptothecin. Moreover, they estimated the temperature resolution achievable using
their nanogel-based nanothermometer and in the best of cases,
they approached 0.3  C. Data included in Fig. 10 clearly reveal
the excellent potential of nanogel-based luminescence nanothermometry for the future study of intracellular dynamics and
single cell manipulation and therapy.
B.1.5 Luminescent polymer-based intensity luminescence
nanothermometry. A luminescent polymer is a large molecule
(macromolecule) composed of repeating structural units (usually
connected by covalent chemical bonds) that typically shows
visible luminescence (500600 nm) when optically excited by UV
radiation. Although luminescent polymers show relatively low
luminescence efficiencies (quantum yield values well below 0.4),46
they are becoming quite popular for luminescence imaging of
fluids and biological systems due to their very good solubility in
water. In polymers, luminescence is explained in terms of the
existence of luminescent monomers within the macromolecule.
The luminescence intensity of such monomers is affected by
a great variety of parameters such as microenvironmental
polarity and symmetry, strength as well as the number of
chemical bonds in their surroundings.4650 As a consequence, any
change in the structural properties of the luminescent polymer
would cause a large variation in the resulting luminescence
intensity.
Some of the most relevant examples of polymer luminescence
nanothermometers are those based on N-isopropylacrylamide
(NIPAM). As reported by Uchiyama et al., water dissolved
NIPAM based luminescent polymers prepared under certain
conditions undergo a temperature-induced phase transition at
around 32  C.51 Therefore, the luminescence intensity generated
by the luminescent monomers increases drastically, as can be
observed in Fig. 11(a), in which we also included the polymer
emission band obtained at different temperatures (inset). Data
included in Fig. 11(a) correspond to the so-called poly(DBD-AEco-NIPMAM) compound for which the sharp increment on the
luminescence intensity occurs as a consequence of the change in
the microenvironmental polarity during the temperature-induced
phase transition.52
The optical contrast of thermal images based on such polymers
is expected to be great since the temperature-induced phase
transition produces an intensity increment larger than one order
of magnitude. Unfortunately, such an amazing intensity
4310 | Nanoscale, 2012, 4, 43014326

Fig. 11 (a) Temperature dependence of the luminescence intensity


generated from a water solution of poly(DBD-AE-co-NIPMAM). The
inset shows the luminescence spectra at different temperatures. (b)
Temperature dependence of the luminescence intensity generated from
luminescent polymers based on N-alkylacrylamide and fluorophore units.
The inset shows the luminescence spectra at different temperatures. (c)
Digital photos demonstrating the remarkable temperature increment in
the luminescence intensity of an aqueous solution of the N-alkylacrylamide based polymers. Reprinted with permission from ref. 51 and 52.
Copyright 2004 and 2003 American Chemical Society.

enhancement is produced only in a very small temperature range


(2124  C), which limits its application for thermal imaging to
systems with larger temperature variations. However, Uchiyama
et al. overcame this limitation the following year and reported on
the temperature-induced luminescence intensity increment in
polymers based on N-alkylacrylamide and fluorophore units.51
The UV-excited emission spectra are shown in the inset of
Fig. 11(b) and similar to results from Fig. 11(a), the luminescent
intensity increases with temperature by more than one order of
magnitude. Although the magnitude of the luminescence intensity increment observed in NIPMAM and N-alkylacrylamide
based polymers is very similar, the former shows that the
intensity variation is produced in a larger temperature range (10
40  C), thus it can be used for thermal sensing in a wider
temperature range. Digital photos included in Fig. 11(c) show the
remarkable temperature increment in the luminescence intensity
of an aqueous solution of the N-alkylacrylamide based polymers.
These pictures also reveal that even for temperatures close to
40  C, the polymer solution remains stable without any evidence
of precipitation.
Despite the excellent results and perspectives described in this
section, it should be noted at this point that the use of luminescence intensity as the thermal indicator has several drawbacks.
Perhaps the main shortcoming of the systems described above is
that the luminescence intensity depends on temperature but also
on the local concentration of emitting centers. Thus, when performing thermal imaging based on luminescence intensity, it is
a likely possibility that the observed luminescence intensity
This journal is The Royal Society of Chemistry 2012

View Article Online

Published on 16 May 2012. Downloaded by Universidade Federal de Alagoas on 15/09/2015 15:05:40.

variations could not only be caused by temperature variations


but also by fluctuations in the local concentration of emitting
centers. This, obviously, would cause a false thermal image. The
thermal imaging methods described in the forthcoming section
have been developed in an attempt to overcome this problem by
extracting temperature information from a spectral magnitude
that is independent of the local concentration of emitting centers.
B.2 Band-shape luminescence nanothermometry
Band-shape luminescence nanothermometry (BSLNth) is
generally based on luminescent systems whose luminescence
spectra consist of several lines/bands with a relative intensity that
is strongly dependent on temperature. There are two main
operation schemes for BSLNth. In the first approach, the luminescence lines/bands are generated by different emitting centers
so that the temperature induced band-shape change arises from
the different thermal quenching of each center or from the
thermally induced changes in the energy transfer rates among
these centers. In the second approach, the different lines/bands
that contribute to the luminescence spectra are generated by
a unique luminescent center. In this case, the modification in the
luminescence band-shape is generally caused by a thermally
induced population re-distribution among the different energy
levels of the emitting center. Of note, in both cases the relative
intensity of the luminescence lines/bands depends only on
temperature but not on the local concentration of emitting
centers. This represents a significant advancement over intensity
luminescence nanothermometry since the temperature reading is
not affected by fluctuations in the concentration of luminescent
centers. There are a great variety of examples in the literature
that report on luminescent systems capable of band-shape
dependent thermal sensing based on either of these two schemes.
These systems can be composed of rare earth ions, quantum dots
and organic dyes. Below we summarize the most relevant results
in each of these cases.
B.2.1 QD-based band-shape luminescence nanothermometry.
Conventional QDs show a simple luminescence spectrum consisting of a single broad luminescence band correlated to the
lowest energy excitonic recombination. The spectral position and
intensity of this luminescence band are temperature dependent
and they are, in fact, widely used for both ILNth and SLNth.
Nevertheless, the presence of this single band does not allow the
use of simple QDs for BSLNth since, as explained previously,
BSLNth requires the simultaneous presence of two or more
luminescence lines/bands to extract a temperature reading from
their relative intensities. Very recently, Vlaskin et al. demonstrated the potential use of QDs in BSLNth, in particular, CdSe
QDs that were doped with Mn2+ ions.53 The presence of Mn2+
ions leads to the appearance of energy states within the QD
band-gap as can be observed in Fig. 12(a). When the QD system
is optically excited to its high-energy excitonic states, efficient
energy transfer (kET) quenches excitonic emission and sensitizes
the Mn2+ luminescence. As a consequence, the emission spectrum
becomes more complicated and now consists of two luminescence bands: one corresponding to the exciton re-combination
(characteristic broad emission band of QDs) and another broad
band (at longer wavelengths, i.e. at lower energies)
This journal is The Royal Society of Chemistry 2012

Fig. 12 (a) Schematic representation of the energy levels and bands of


a QD + Mn2+ system. (b) Luminescence emission spectra at different
temperatures as obtained from colloidal Zn0.99Mn0.01Se/ZnCdSe nanocrystals under UV excitation. (c) Temperature dependence of the intensity ratio between the excitonic related emission and overall total
emission of Zn0.99Mn0.01Se/ZnCdSe nanocrystals under UV excitation.
(d) Digital picture of the colloidal suspension of Zn0.99Mn0.01Se/ZnCdSe
nanocrystals under UV excitation, as obtained at 210 and 400 K.
Reprinted with permission from ref. 53. Copyright 2010 American
Chemical Society.

corresponding to optical-de-excitations from the inter-gap


energy levels of Mn2+.54 The relative intensities between these two
bands simultaneously depend on the luminescence efficiency of
the excitonic and inter-gap states as well as on the relative population of these two states. In turn, this depends on several
parameters including temperature, energy separation, DE,
between the Mn2+ excited state and the lowest energy excitonic
level, as well as on the energy transfer rate (also temperature
dependent).5355 Consequently, the intensity ratio of these two
emission bands is expected to be temperature dependent and in
fact, this is exactly what has been observed. Fig. 12(b) shows the
Nanoscale, 2012, 4, 43014326 | 4311

Published on 16 May 2012. Downloaded by Universidade Federal de Alagoas on 15/09/2015 15:05:40.

View Article Online

normalized emission spectra of ZnMnSe/ZnCdSe nanocrystals as


obtained for different temperatures. It is clear that the temperature induces a remarkable change in the intensities of the Mn2+
and excitonic bands (centered at around 570 and 600 nm,
respectively). Fig. 12(c) shows the relative contribution of the
excitonic emission to the total (integrated) luminescence intensity
and as can be observed the relative contribution of the excitonic
emission increases with temperature with an almost linear
dependence near room temperature. Thus, the temperature
reading can be achieved through the simple measurement of the
relative contribution of the excitonic emission with respect to the
overall emission or by the ratio between the excitonic and Mn2+
emissions. Indeed, the temperature induced change in this ratio is
strong enough to produce a clear change in the color of the
colloidal solution as can be observed from the pictures included
in Fig. 12(d). The temperature range in which thermal sensing is
achieved can be tailored by a rational design of the energy level
scheme during QD synthesis. In particular, Vlaskin et al.
demonstrated that the temperature at which the change in the
intensity ratio occurs depends on the energy separation between
the first excitonic state and the inter-gap Mn2+ state (DE in
Fig. 12(a)).53 This energy separation depends on the QD size in
such a way that large QDs lead to a reduced energy separation
between these states so that the temperature sensing can occur at
lower temperatures. In particular, the thermal sensing rate
reduces from 300 K down to 150 K when the QD size is reduced
from 4.7 to 3.5 nm.53 In the pioneering work of Vlaskin et al. it
was also demonstrated that the QD + Mn2+ luminescence system
can provide thermal reading with a resolution as good as 0.2
 54
C. Nevertheless its application in real thermal imaging has not
yet been demonstrated although great advances towards the
incorporation of this system into biological structures have been
made through the synthesis of QD + Mn2+ in water, which has
been recently reported.55
B.2.2 Rare earth-based band-shape luminescence nanothermometry. There are numerous examples in the literature of
rare earth (RE) based BSLNth (RE-BSLNth). The fundamental
mechanism of RE-BSLNth depends on whether the luminescence
lines under analysis are generated by a single type of rare earth or
by a combination of different rare earth ions. Based on this
criterion, it is possible to classify the RE-BSLNth into two
groups: single-center RE-BSLNth and multi-center RE-BSLNth.
The basic working principles, materials and applications of each
class are briefly described below.
B.2.2.1 Single-center RE-BSLNth. Rare earth ions (also
known as lanthanide ions), with a few exceptions, show a rich
energy level scheme.30 Due to the partial screening of the f electrons, the energy level scheme of a particular rare earth ion
remains practically unchanged from host to host. The emission
probability for each energy level depends on a great variety of
factors but one of the most critical parameters is the energy
separation between the RE energy levels. If this energy separation is small (i.e. comparable to the thermal energy kBT where kB
is the Boltzmann constant and T is the temperature), then it is not
possible to populate a single level since, due to the Boltzmann
statistics, the population will be re-distributed among energy
levels with similar energy. Lets suppose that a RE energy level
4312 | Nanoscale, 2012, 4, 43014326

(labeled as 1) is populated as a consequence of optical excitation


from the ground state. Due to the proximity of a second level of
higher energy (labeled as 2), the population initially excited into
level 1 is thermally re-distributed among levels 1 and 2. In
a steady state condition, the population of the high-energy state,
N2, will be given by:
N2 N1exp(DE/kBT)

(1)

where N1 is the population of level 1 and DE is the energy


separation between states 1 and 2. Since both states are populated, both contribute to the overall luminescence spectrum with
the presence of two luminescence lines at different energies. The
intensity of the luminescence line corresponding to the de-excitations from state 1 down to the ground state, I1, is given by:
I1 41N1

(2)

where 41 is a constant whose value depends on a great variety of


geometrical factors as well as on intrinsic properties of the
emitting level (such as branching ratios and luminescence
quantum efficiency). Similarly, the intensity of the luminescence
line corresponding to the de-excitations from state 2 is given by:
I2 42N2

(3)

Thus, the ratio between both intensities is given by:


I2/I1 (42/41)(N2/N1) (42/41)exp(DE/kBT)

(4)

such that from the experimental determination of the intensity


ratio, a temperature reading can be reached. Of course, appreciable thermal sensitivity would be achieved only if the energy
separation between emitting levels is small, so that small
temperature changes can induce large population re-distributions. As a consequence, not all the rare earth ions can be used
for BSFNt but only those possessing radiative states with an
energy separation of the order of few hundreds of wavenumbers,
i.e. with an energy separation comparable to the room temperature thermal energy (the so-called thermally coupled states).
Fig. 13 includes some of the energy level diagrams of the rare
earth ions that have been previously used for BSLNth:
neodymium (Nd3+), thulium (Tm3+), europium (Eu3+), erbium
(Er3+) and dysprosium (Dy3+).5663 In each case the luminescence
lines used for BSLNth are indicated. The bottom graph of Fig. 13
shows the temperature variation of the normalized luminescence
intensity ratios determined experimentally for a series of Er3+,
Dy3+, Eu3+ and Nd3+ doped systems.5759 As can be observed,
a remarkable change in the luminescence ratio is observed in all
the cases. The temperature-induced variation of the luminescence
ratio has been found to be very similar in the cases of Er3+, Eu3+
and Dy3+. The luminescence ratio has been found to be much
more temperature dependent for Eu3+-doped La2O2S system. In
this particular case, the luminescence intensity ratio between the
emission lines generated from the 5D0 and 5D1 levels varies by
almost one order of magnitude when the temperature is only
increased few tens of degrees.56,59 This surprising result has been
tentatively explained in the past by involving a third state (charge
transfer state corresponding to the La2O2S host) in the thermal
coupling between the two emitting states. Therefore, in this case
This journal is The Royal Society of Chemistry 2012

Published on 16 May 2012. Downloaded by Universidade Federal de Alagoas on 15/09/2015 15:05:40.

View Article Online

studies, a thermally sensitive luminescent nanoparticle was


mechanically attached to an AFM tip (see Fig. 14(a)). Based on
the same principles of conventional AFM, the luminescent
nanoparticle was scanned in pseudo-contact mode over the
surface to be thermally imaged. The visible emission generated
from the nanoparticle after 975 nm multiphoton excitation was
simultaneously collected and analyzed by a fixed microscope
objective, as is schematically shown in Fig. 14(a). The spatial
variation of the intensity ratio between the 520 and 540 nm
luminescence bands provides the thermal image of the surface
after proper translation into temperature units. The reduced size
of the luminescence probe, in combination with the accurate
control of both position and height of the AFM tip, allowed the
achievement of high spatial resolutions. This is illustrated in
Fig. 14(b) in which the thermal image of an integrated circuit,
where hot spots have been artificially introduced, is shown.73
This approach has been also applied for the thermal imaging of
fluids, however, it still has the same limitations as any other
thermal scanning microscopy technique.75 It is restricted to bidimensional thermal imaging of surfaces and since it is a pseudocontact method, can alter the original temperature distribution.
Since these first works demonstrating the ability of Er3+ ions as
ratiometric temperature sensors, they have been widely used in
a great variety of applications ranging from the determination of
local temperatures in the surroundings of optically excited
metallic nanoparticles, to the thermal imaging of microfluidics in
the presence of external heating sources.60 Very recently, Er3+
ions have emerged as versatile biocompatible luminescent
nanothermometers. This possibility has been demonstrated by
using Er3+ and Yb3+ co-doped NaYF4 (NaYF4:Er3+, Yb3+)

Fig. 13 Energy level diagrams of the trivalent rare earth ions used for
BSLNth: neodymium, dysprosium, erbium and europium. The transitions used in BSLNth are also indicated. Bottom graph: temperature
variation of the intensity ratio between the thermally coupled luminescence transitions in different systems. Data were extracted with permissions from ref. 5759. Dots are experimental data and dashed lines are
guides for the eyes.

the hyperthermal sensitivity is caused not by a Boltzmann


redistribution but by a strong thermal dependence of the energy
transfer rate between the charge transfer states and the Eu3+ 4f
emitting levels.56,59
Among the different rare earth ions capable of single-center
BSLNth, erbium is probably the most used one. Er3+ ions typically show a very intense green emission that consists of two
luminescence bands centered at 520 and 540 nm, whose relative
intensity is strongly temperature dependent.57,61,62,6467 Er3+ ions
have been widely used as BSLNth since the two bands, which
provide the thermal reading, lie within the green spectral range
where highly efficient detection is possible. Moreover, these two
bands can be easily populated through multiphoton absorption
of NIR radiation with the assistance of sensitizer ions such as
ytterbium (allowing for spatial resolution enhancement).68 Due
to these two facts, it is possible to find numerous examples in the
literature of thermal sensing and imaging using Er3+ ions.6070
One of the most attractive approaches is the use of Er3+-doped
nanoparticles in scanning thermometry.36,60,7174 In pioneer
This journal is The Royal Society of Chemistry 2012

Fig. 14 (a) Experimental set-up used for scanning thermal microscopy


based on an erbium doped luminescent particle. (b) Electron microscopy
image of an integrated conducting wire in which a series of hot spots were
artificially created in the form of wire width reductions (left). The corresponding thermal image, as obtained by BSLNth, is shown at the right.
Reproduced from ref. 73 with permission.

Nanoscale, 2012, 4, 43014326 | 4313

Published on 16 May 2012. Downloaded by Universidade Federal de Alagoas on 15/09/2015 15:05:40.

View Article Online

biocompatible nanoparticles.65 NaYF4:Er3+, Yb3+ nanoparticles


are widely used because of their outstanding luminescence
brightness, their low toxicity and their good chemical and
physical stability.7679 The main goal of the Yb3+ ions is to
provide efficient excitation of the Er3+ ions under NIR optical
excitation (980 nm) via a two-photon absorption process (such
that autoluminescence is avoided while achieving high spatial
resolutions). The two-photon excited emission of NaYF4:Er3+,
Yb3+ nanoparticles (resembling the typical Er3+ emission)
obtained at two different temperatures is shown in Fig. 15(a).
The presence of two thermally coupled levels is evidenced by the
temperature induced change in the ratio between the luminescence bands centered at 525 and 545 nm.80 The temperature
variation of the luminescence intensity ratio follows an almost
linear relationship up to 40  C allowing for the straightforward
determination of temperature.65 Once the NaYF4:Er3+, Yb3+
nanothermometers were calibrated they were introduced to
HeLa cancer cells by a simple incubation procedure. Once
incorporated within the HeLa cancer cell, they were used to
monitor cell apoptosis caused by an external heating source (in

Fig. 15 (a) Luminescence emission spectra generated from Er3+/Yb3+ codoped NaYF4 nanoparticles under 920 nm excitation as obtained at two
different temperatures. (b) Intracellular temperature (as obtained from
the spectral analysis of the intracellular Er3+ luminescence) as a function
of the applied voltage to a resistance attached to the micro-chamber
containing the cell. The optical transmission images of the cell at different
applied voltages are also shown. Reprinted with permission from ref. 65.
Copyright 2010 American Chemical Society.

4314 | Nanoscale, 2012, 4, 43014326

this case an ohmic resistance attached to the micro-chamber


containing the HeLa cancer cells). A 920 nm laser beam was
focused inside the cell so that the Er3+ luminescence was induced
through a two-photon absorption process involving the excited
states of Yb3+ ions and was collected by the same microscope
objective that was used for focusing the excitation laser beam
inside the cell. A subsequent spectral analysis of the intracellular
luminescence non-invasively and accurately determined the
intracellular temperature. This simple approach has been used to
determine the intracellular temperature of a single cell as
a function of the voltage applied to the ohmic resistance in
physical contact with the micro-chamber containing the cells.
Results are summarized in Fig. 15(b). The intracellular temperature was found to follow a quadratic behavior with respect to
the applied voltage, in accordance with a Joule heating process.
Fig. 15(b) also shows the transmission optical images of the
investigated cells at different temperatures where at room
temperature (25  C), the cancer cells show their typical irregular
shape (green dotted outline). A subsequent temperature increase
to 35  C does not produce any relevant changes to its
morphology. However, increasing the temperature to 45  C,
a small membrane fragment was observed, which is indicative of
cell death. Results included in Fig. 15 demonstrated the potential
use of Er3+ doped nanoparticles for the investigation of intracellular thermal dynamics.
B.2.2.2 Multi-center RE-BSLNth. Multi-center RE-BSLNth
is based on the incorporation of a luminescent compound containing two rare earth ions (or two different species of the same
rare earth ion), within the system under thermal investigation.
The luminescence intensities of each luminescent center follow
very different thermal behaviors, in such a way that the luminescence intensity ratio between their emissions would be
strongly temperature dependent.
Among the most relevant examples of RE-BSLNth is the work
of Brites et al., who reported on the synthesis of luminescent
molecular thermometers consisting of terbium (Tb3+) and Eu3+
co-doped g-Fe2O3 nanoparticles.70,81 The luminescent system
was specially designed in such a way that it shows an excited
triplet state with energy slightly above that of the Tb3+ 5D4
emitting state. The small energy difference between the Tb3+ and
host states caused the occurrence of a Tb3+ to host energy
transfer process, which is strongly temperature dependent as it is
a phonon-assisted process.70,81
Fig. 16(a) shows the temperature dependence of the luminescence generated by the Eu3+, Tb3+ co-doped g-Fe2O3 nanoparticles after UV optical excitation. It should be noted that in
this system coordination compounds of Tb3+ and Eu3+ ions are
embedded and, thus, the triplet state refers to the organic part of
the complex. The two luminescence lines used for temperature
sensing are indicated and properly labeled in terms of the emitting ion. As can be observed, the luminescence line at around 547
nm (generated by Tb3+ ions) suffers from a relevant quenching
near room temperature due to the thermal activation of energy
transfer processes from the 5D4 (Tb3+) state to the host triplet
state. On the other hand, the luminescence line at around 610 nm
(corresponding to Eu3+ ions) remains almost constant over the
entire temperature range due to the large energy difference
between the host band and the 5D0 excited state of the Eu3+ ions.
This journal is The Royal Society of Chemistry 2012

Published on 16 May 2012. Downloaded by Universidade Federal de Alagoas on 15/09/2015 15:05:40.

View Article Online

Fig. 16 (a) Temperature dependence of the luminescence generated by


the Eu3+, Tb3+ co-doped g-Fe2O3 nanoparticles after UV optical excitation. (b) Temperature dependence of the ratio between the Eu3+ and Tb3+
luminescence intensities. Reproduced from ref. 81 with permission.

Fig. 16(b) shows the temperature dependence of the ratio


between the Eu3+ and Tb3+ luminescence intensities. It is clear
that this luminescence ratio guarantees a precise and absolute
temperature measurement with an accuracy, estimated by Brites
et al., to be close to 0.5  C.81 The relevance of the results
summarized in Fig. 16 is even more notable given the flexibility in
the nanoparticle synthesis allowing, for example, for thin film
production. This, in turn, would allow for the opportunity of real
time temperature sensing during magnetic hyperthermia possible
by taking advantage of the superparamagnetism character of the
temperature sensitive nanoparticles. Despite these good
perspectives, the potential application of Eu3+, Tb3+ co-doped gFe2O3 based luminescent nanothermometers for biological
applications has not yet been demonstrated. At this point it
should be noted that temperature sensing based on Eu3+ and Tb3+
co-doped systems has been recently proposed by Cui et al. who
also reported on the remarkable temperature induced changes in
the Eu3+ to Tb3+ luminescence intensities in double doped metal
organic luminescent compounds.82
A different approach was proposed by Ishiwada et al. where
the phosphor thermometer consisted of a Tb3+/Tm3+ co-doped
Y2O3 system.83 BSLNth was achieved by recording the ratio
between the Tm3+ emission in the blue (at around 466 nm) and
the dominant green emission of Tb3+ ions (at around 540 nm).
The intensity ratio between these two emissions is strongly
This journal is The Royal Society of Chemistry 2012

temperature dependent, varying almost a 100% in the 300


1100  C range as can be observed in Fig. 17. In this case, the
temperature induced ratio variation is explained on the basis of
the different thermal quenching of the individual ions: the blue
emission of Tm3+ does not show any appreciable thermal
quenching in the 3001100  C, whereas the 540 nm Tb3+ emission
is fully quenched at 1100  C. Despite the large ratio contrast
shown by the Tb3+/Tm3+ co-doped Y2O3 system, it should be
noted that it is achieved over a large temperature range (over
almost 800  C) so that the net thermal sensitivity is low. In
addition the luminescence ratio is linear only in the 500900  C
range and is nearly flat close to room temperature. Therefore, the
system recently proposed by Ishiwada et al. would find practical
application in thermal imaging of high temperature systems
(such as combustion cells) rather than in biosystems and
microelectronic devices.
As a final example of multi-center RE-BSLNth we will briefly
describe the original idea of Seaver and Peele84 They realized that
when the Eu3+EDTA complex is dissolved in water, two
different species appear: Eu3+(EDTA)(H2O)2 and Eu3+
(EDTA)(H2O)3. The presence of these two species leads to the
existence of two slightly different crystal fields (local environments) for Eu3+ ions, i.e. two different Eu3+ centers. This causes
the appearance of two 7F0 / 5D0 emission lines close in energy
(separated by less than 1 nm). The ratio between these luminescence lines was found to be strongly temperature dependent as
a consequence of the different thermal quenching rates of both
Eu3+ centers as well as because of the temperature induced
chemical equilibrium between them. The luminescence ratio
varied at a rate of 8% per  C, allowing for thermal sensing with
a thermal resolution of 1  C. Seaver and Peele applied this novel
thermal sensing approach to determine the temperature of
acoustically levitated water drops as small as 200 mm, being
capable of determining single drop temperature rises of 7  C.84
B.2.3 Dye-based band-shape luminescence nanothermometry.
As described in previous sections, luminescent dyes (such as
Rhodamine B) have been widely used for ILNth. The sole use of
a luminescent dye for temperature sensing by ILNth can result in
problems due to local fluctuations in both excitation light
intensity and dye concentration. This is especially true in
dynamical systems such as micro-fluidics and living cells. Both
fluctuations make the calibration of ILNth very difficult if not

Fig. 17 Images of the temperature-sensitive visible luminescence of


Y2O3:Tb3+, Tm3+ phosphors. Rows (a), (b) and (c) correspond to different
Tb and Tm concentrations as described in ref. 83. Data reproduced from
ref. 83 with permission from Optical Society of America.

Nanoscale, 2012, 4, 43014326 | 4315

Published on 16 May 2012. Downloaded by Universidade Federal de Alagoas on 15/09/2015 15:05:40.

View Article Online

even impossible. These problems could be avoided by using


a second luminescent dye with temperature independent luminescence intensity, such as Rhodamine 110, that would act as
a luminescence reference. This was the approach used by Ebert
et al. who were able to determine the laser-induced thermal
loading of an optofluidic device that consists (see Fig. 18(a)) of
a micro-channel (capillary) side-pumped by two optical fibers
both coupled to a 1064 nm laser.8 An aqueous solution containing both Rhodamine B and Rhodamine 110 was injected into
the capillary and the laser induced temperature increment,
created by the non-vanishing absorption of water at 1064 nm,
was determined from the confocal luminescence image obtained
from the ratio between the luminescence intensities generated by
the two dye species (found to decrease at a rate of 1.3% per  C).
Fig. 18(b) shows a typical thermal image of the optofluidic device
as obtained when a 1064 nm laser with a total power of 2 W was
incident onto the microfluidic device. The thermal images
obtained by the dual dye system were compared to theoretical
predictions and were in excellent agreement, revealing the feasibility and applicability of this approach.8,85
B.3 Spectral luminescence nanothermometry
As previously discussed, spectral luminescence nanothermometry (SLNth) provides thermal reading from the

temperature induced spectral shift of luminescence lines. At


variance with ILNth and BLNth, SLNth is not based on the
analysis of absolute or relative intensities but on the determination of the spectral position of the luminescence lines. Consequently, temperature reading is not affected by luminescence
intensity fluctuations caused by variations in the local concentration of emitting centers. Obviously, high-resolution SLNth
implies the use of luminescence systems with large temperature
induced spectral shifts in such a way that small temperature
variations would cause remarkable shifts in the luminescence
lines. Although the presence of temperature induced spectral
shifts has been observed in many luminescent systems, only in the
case of semiconductor QDs has it been successfully used for highresolution SLNth.
Fig. 2(a) demonstrates that temperature increments cause an
overall intensity reduction of the QD luminescence as well as
a remarkable spectral shift. This is a general phenomenon in QD
luminescence as it has been observed and reported in many
systems.8690 The physical origin of the temperature induced
spectral shift of QDs is far from being simple. It has been stated
that it occurs as a result of the combination of different
phenomena. The thermal spectral coefficient of QDs (dl/dT,
where l denotes the spectral position of the luminescence line)
can be written as:8688
dl/dTf (dEg0/dT) + (dEconf/dT) + (dJeph/dT)

Fig. 18 Top. Temperature dependence of the ratio between luminescence intensities generated by Rhodamine B and Rhodamine 110 in an
aqueous solution containing both dyes. The optical transmission image in
the middle corresponds to an opto-fluidic device consisting of a microchannel side pumped by two optical fibers both coupled to a 1064 laser.
The graph at the bottom is the thermal image of the micro-fluidic when
the total 1064 nm power was set to 2 W. Data reproduced from ref. 83
with permission from Optical Society of America.

4316 | Nanoscale, 2012, 4, 43014326

(5)

where the first term (dEg0/dT) corresponds to the temperature


coefficient of the bandgap energy of the bulk material (Eg0),
a term that only depends on the QD constituent elements. The
second term (dEconf/dT) accounts for the temperature coefficient
of the quantum confined energy (Econf). According to previous
works, the magnitude of this second term depends on both the
thermal expansion coefficient of the QDs and on the radius of the
QD (dEconf/dT increases as the QD size decreases).88 Finally,
the third term (dJeph/dT) accounts for the temperature-induced
change in the electronphonon coupling energy (Jeph), caused by
the temperature-dependence of both the phonon energy and
density of states.87,88 Due to the presence of these different terms,
the a priori estimation of dl/dT of a given QD system is difficult
to undertake, since it depends on the intrinsic material properties
as well as on the geometrical properties (size) of the QDs. As
a matter of fact, it is possible to find some systems (such as PbSe
QDs) in which the sign of the dl/dT coefficient changes with the
QD size.88 This fact is illustrated in Fig. 19(a) and (b) in which we
have shown the temperature variation of the spectral shifts of the
3.9 and 6.9 nm diameter PbSe QDs, respectively. Note that for
small PbSe QDs the dl/dT coefficient is positive whereas for
larger PbSe QDs it is negative. More recent studies have shown
that QD size can be used to optimize the SLNth sensitivity. In the
particular case of CdTe QDs, Maestro et al. demonstrated that
the spectral thermal coefficient grows monotonously as the QD
size is reduced, as can be observed in Fig. 19(c).90 Note that for
the particular case of CdTe, ultra-small QDs lead to dl/dT
coefficients close to 0.8 nm per  C. For such large coefficients,
thermal sensitivities as good as 0.2  C are, in principle, expected.
The first demonstration (to the best of our knowledge) of QD
based SLNth was provided by Li et al. who used the luminescence of CdSe QDs to determine the temperature of an electrical
This journal is The Royal Society of Chemistry 2012

Published on 16 May 2012. Downloaded by Universidade Federal de Alagoas on 15/09/2015 15:05:40.

View Article Online

Fig. 19 Temperature dependence of the luminescence wavelength of (a)


3.9 nm and (b) 6.9 nm PbSe QDs, respectively. (c) Temperature coefficient of luminescence emission wavelength as obtained from CdTe QDs
of different sizes. In all the cases solid dots are experimental data and
solid lines are guides for the eyes. Reprinted with permission from ref. 88
and 89. Copyright 2010 American Chemical Society.

micro-heater on an aluminum microwire and their experimental


approach to achieve non-contact thermal sensing is illustrated in
Fig. 20(a).91 Individual CdSe QDs were placed on top of the
aluminum micro-heater and were optically excited by a tightly
focused green (532 nm) laser beam. The CdSe QD luminescence
was collected by the same objective used for focusing the 532 nm
radiation and then spectrally analyzed by a CCD camera
attached to a high-resolution monochromator. Any local change
in the micro-heater temperature was then detected by the
appearance of a red shift in the CdSe QD luminescence shifting at
a rate of 0.1 nm per  C.91 By scanning the focusing/collecting
microscope objective authors were able to measure the temperature profiles along the micro-heater with sub-micrometric
resolution. Representative results are shown in Fig. 20(b), where
the temperature profiles obtained for different applied voltages
are shown. It is clear that the experimentally obtained profiles
account well for the predictions made based on Joules law, i.e.
the temperature increment rises with the applied voltage. From
the temperature profiles the temperature uncertainty can be
estimated to be close to 1  C.
The results described above demonstrated the ability of QDs
to be used as highly sensitive optical probes for SLNth. This fact,
in combination with the possibility of incorporating QDs in
biological systems (such as cells and tissues) makes them unique
This journal is The Royal Society of Chemistry 2012

Fig. 20 (a) Experimental set-up used by Li et al. in ref. 91 to measure the


thermal images of a micro-heater using CdSe QDs. (b) Temperature
profiles obtained along the micro-heater for different applied voltages.
Reprinted with permission from ref. 91. Copyright 2007 American
Chemical Society.

probes for high-resolution SLNth of biological systems.92,93 It


should be noted that for the purpose of thermal imaging of
biological systems, QDs offer an additional advantage: they can
be optically excited within the biological window through
a multiphoton excitation process.94 This last possibility simultaneously allows for the achievement of an improved spatial
resolution as well as large penetration depths into tissues.95
During the last few years several groups have experimentally
demonstrated QD based SLNth of biological systems. The first
work reporting on the use of QDs for SLNth in a biological
system was published by Maestro et al.89 In this case, CdSe QDs
were incorporated into HeLa cancer cells by a simple incubation
procedure. These cancer cells were subsequently subjected to an
external heating process using an air-assisted micro-heater as is
schematically depicted in Fig. 21(a). At the same time, 800 nm
femtosecond laser pulses were focused into the cell in order to
excite the CdSe QD luminescence via a multiphoton excited
process. The intracellular QD luminescence was spectrally
analyzed allowing the authors to obtain, in real time, the time
evolution of the intracellular temperature during the heating
procedure. This is presented in Fig. 21(b), which shows the
intracellular QD emission spectra as obtained at two different
heating times. As can be observed, the QD intracellular emission
is partially quenched but concomitantly undergoes a relevant red
shift. From the analysis of this spectral red shift and based on the
Nanoscale, 2012, 4, 43014326 | 4317

View Article Online

Published on 16 May 2012. Downloaded by Universidade Federal de Alagoas on 15/09/2015 15:05:40.

analysis was used to monitor the local temperature response


inside single living cells upon external chemical stimuli. In
particular, they managed to map the intracellular heat generation in living cells following Ca2+ stress and cold shock. Their
results demonstrated the presence of an inhomogeneous intracellular temperature and are currently boosting the use of QDs as
high-resolution bio-nanothermometers.

B.4

Polarization luminescence nanothermometry

Polarization luminescence nanothermometry (PLNth) is based


on the temperature dependence of the polarization state of the
luminescence generated by luminescent molecules under Brownian dynamics. In general, when a luminescent molecule is illuminated by a linearly polarized laser beam it re-emits partially
polarized luminescence due to the random orientation of the
molecules. The polarization anisotropy factor of the luminescence, r, is defined as:97,98
r (Ik  It)/(Ik + 2It)

(6)

where Ik and It are the intensities of the luminescence polarized


parallel and perpendicular to the incident polarization. In the
absence of any molecular motion, the polarization anisotropy
reaches its maximum value of 0.4. On the other hand, in the
presence of molecular rotation (induced by, for example, its
Brownian dynamics) the polarization anisotropy decreases due
to an average-effect caused by the molecular rotations taking
place during the luminescence emission (during the lifetime
temporal window). When a luminescent molecule with a hydrodynamic molecular volume of V is incorporated in a fluid with
a h viscosity, its polarization anisotropy can be written as:97
r 0.4(1 + (Vh/kBTsf))1

Fig. 21 (a) Schematic drawing of the experimental set-up used by


Maestro et al. in ref. 89 for intracellular luminescence thermal sensing
based on CdSe QDs. (b) Intracellular luminescence spectra generated by
CdSe QDs as obtained at two different heating times (0 and 3 min). (c)
Intracellular temperature determined by SLNth using CdSe QDs.
Reprinted with permission from ref. 89. Copyright 2010 American
Chemical Society.

temperature spectral coefficient of CdSe QDs (close to 0.15 nm


per  C in this case), the authors were able to determine the cell
temperature during the different stages of the heating procedure
(see Fig. 21(c)).
A similar approach as that used by Maestro et al. was adopted
by Yang et al.96 CdSe QDs were also incorporated into HeLa
cancer cells by an incubation procedure, although in this case the
CdSe QD intracellular temperature obtained from the spectral
4318 | Nanoscale, 2012, 4, 43014326

(7)

where T is the molecule temperature and sf its luminescence


lifetime. Thus, the polarization of the luminescence is unequivocally related to the molecule temperature, so that a thermal
reading can be obtained from the analysis of the polarization
dependence of the luminescence.
One of the molecules used in the past for PLNth is fluorescein,
whose calibration curve obtained in aqueous solution (temperature dependence of its polarization anisotropy) is shown in
Fig. 22(a). It should be noted that in the biological range (20
50  C) the polarization anisotropy is modified by more than
a 100% making it a highly sensitive temperature sensor. Moreover, the temperature is determined from the analysis of an
optical parameter that is independent of the local concentration
of luminescence molecules so that neither normalization nor
reference procedures are necessary. Fluorescein based PLNth has
been recently applied to determine the heating potential of
diverse metallic nanostructures acting as nanometer-size heat
sources such as gold nanorods (GNRs).99 The procedure used for
PLNth is schematically illustrated in Fig. 22(b). The GNRs were
dispersed onto a fluorescein-containing film and placed in
a confocal luminescence microscope. The fluorescein molecules
were excited by a 473 nm beam and by an NIR laser beam tuned
to the surface plasmon resonance of GNRs. The parallel and
orthogonal components of the fluorescein luminescence were
This journal is The Royal Society of Chemistry 2012

Published on 16 May 2012. Downloaded by Universidade Federal de Alagoas on 15/09/2015 15:05:40.

View Article Online

Fig. 22 (a) Luminescence polarization anisotropy as a function of


temperature for fluorescein dissolved in water. Dots are experimental
data and the solid line is a guide for the eyes. (b) Schematic of the
experimental set-up and data analysis procedure adopted by Baffou et al.
to obtain a thermal image of optically excited GNRs by PLNth. Data
reproduced from ref. 99 with permission from Optical Society of
America.

separated by a polarizing cube and sent to two independent


detectors. The blue and infrared beams were scanned across the
sample plane for simultaneous local heating and temperature
measurement. From the confocal luminescence images obtained
in terms of the luminescence intensities at both polarizations, the
polarization anisotropy image was obtained. Based on the calibration curve of Fig. 22(a), this polarization anisotropy image
was translated into temperature units as it is shown in Fig. 22(b).
Based on this operation scheme, Baffou et al. were able to achieve temperature resolutions as good as 0.1  C while keeping
spatial resolutions below 0.5 mm.99

B.5 Bandwidth luminescence nanothermometry


In general, when a luminescent center is heated, homogeneous line
broadening causes spectral broadening of the emission lines as
a consequence of the local environment fluctuations due to the
thermal vibration of the luminescent center and also that of its
neighboring atoms/molecules.30 The magnitude of the temperature induced luminescence line broadening is in general small and
thus, can be only observed in systems showing inherent narrow
emission lines and in which homogeneous line broadening
This journal is The Royal Society of Chemistry 2012

dominates over the inhomogeneous one. This is the case of rare


earth ions incorporated in some crystalline hosts. One of the most
pertinent examples is the case of Nd3+ ions in the well-known
YAG host (Nd:YAG).100 Nd:YAG shows two hypersensitive
luminescence lines at around 940 nm that correspond to the low
energy emissions within the 4F3/2 / 4I9/2 transition (see Fig. 23(a))
and the bandwidth of these two transitions is strongly temperature
dependent as can be observed in Fig. 23(a). In the proximity of
room temperature, the luminescence linewidth increases almost
linearly with temperature in a wide temperature range (100  C)
with a rate close to 0.04 cm1 per  C.101 This linear dependence
allows for temperature reading from the determination of the
Nd3+ luminescence linewidth with a constant sensitivity in the
whole temperature range. Bandwidth luminescence nanothermometry (BLNth) based on the hypersensitive luminescence
lines of Nd:YAG has been recently demonstrated by Benayas et al.
who managed to obtain thermal images of a Nd:YAG ceramic
microchip laser component in the presence of a tightly focused
pump beam at 808 nm (where the Nd:YAG absorption is as large
as tens of cm1).101 For this purpose, the authors proposed the
experimental set-up schematically shown in Fig. 23(b). In this setup, the Nd:YAG microchip device that is pumped by a tightly
focused 808 nm laser beam, was placed in a modified confocal
microscope. The luminescence generated by the Nd3+ ions was
collected by a scanning microscope and, after passing several
filters and pinholes, was analyzed by a high-resolution spectrometer. This set-up allowed the authors to obtain the two
dimensional spatial distribution of the Nd3+ bandwidth in the
surroundings of the 808 nm tightly focused beam. This luminescence image can be easily translated into temperature units by
using the calibration curve presented in Fig. 23(a).
From the experimental set-up schematically shown in
Fig. 23(b), the authors were able to get thermal images (with
a sub-micrometric resolution) of a microchip laser element in the
presence of high intensity pump radiations. A typical example of
the thermal images obtained by BLNth is shown in Fig. 23(c). As
expected, the maximum temperature increment is produced at
the 808 nm focus position (center of the thermal image) being on
the order of tens of degrees for a total pump power of 1 W. The
large thermal conductivity of the YAG host produces a relevant
spreading of the pump-induced thermal loading as was observed
in the thermal image. Note that the thermal image is obtained
from the confocal luminescence image that in turn can be
expanded to three dimensions since the typical confocal axial
resolutions (few microns) are well below the microchip element
thickness (hundreds of microns). By acquiring the thermal
images at different pump powers, the authors found that the laser
induced thermal gradients at the focus position (determining the
laser performance and stability of the microchip device) follow
a non-linear behavior at variance to the linear one expected from
theory. This information can be critical in understanding the real
world laser performance of microchip elements.
B.6

Lifetime luminescence nanothermometry

The different methods reviewed up to this point require the


acquisition of the luminescence spectrum of the thermally
sensitive optical nanoprobe. In general, the requirement of high
spatial resolution leads to low signal levels because of the use of
Nanoscale, 2012, 4, 43014326 | 4319

View Article Online

Published on 16 May 2012. Downloaded by Universidade Federal de Alagoas on 15/09/2015 15:05:40.

noise ratio in the luminescence spectrum so that an accurate


temperature reading could be achieved. This, in turn, requires
long illumination times leading to the possibility of laser-induced
local heating of the system under investigation due to the activation of multiphonon decay processes. In order to avoid any
externally induced heating of the system to be thermally imaged,
acquisition times should be reduced. This can be accomplished
by extracting the thermal information not from the luminescence
spectrum but from the luminescence decay curve. In this case, the
signal can be acquired at a time interval on the order of the
luminescence lifetime (typically in the range of nanoseconds to
milliseconds). Furthermore, for the acquisition of luminescence
decay curves the use of high-resolution spectrometers is not
necessary since spectral analysis is no longer required. Moreover,
the time evolution of luminescence is usually recorded by using
high gain fast detectors (such as avalanche photodiodes or
photomultipliers) such that high signal-to-noise ratios are
usually achieved. Of course, it goes without saying that for lifetime luminescence nanothermometry (LLNth), systems showing
decay curves that are strongly dependent on temperature are
obviously required. In general, temperature is one of the most
critical parameters determining the luminescence decay time
(lifetime). This is because temperature increments lead to the
activation of phonon-assisted processes and of multiphonon
decays. Both effects cause an increase in the net de-excitation
probability of the emitting level and, hence, a decrease in its
luminescence lifetime. Up to now the systems that have been
proposed, based on the hyperthermal sensitivity of their luminescence lifetime, can be classified into four different groups
depending on whether they were developed from rare earth ions,
luminescent dyes, quantum dots, polymers or membranes. The
main results achieved in each case are summarized below.

Fig. 23 (a) Temperature variation of the luminescence bandwidth of the


Nd:YAG emission line at approximately 940 nm. The inset shows the
luminescence spectrum of Nd:YAG at circa 940 nm showing the two
hypersensitive lines of Nd:YAG. (b) Experimental set-up used for
thermal imaging of tightly pumped Nd:YAG ceramic microchip devices.
(c) Thermal image of an 808 nm tightly pumped Nd:YAG ceramic based
microchip. The laser spot is at the center of the thermal image, where
temperature is maximum. Reprinted with permission from ref. 101.
Copyright 2011 Springer.

high numerical optics, high-resolution spectrometers and of


luminescence particles with reduced dimensions (with a relatively
low absorption efficiency).80 In these conditions, large acquisition times have to be employed to ensure an adequate signal-to4320 | Nanoscale, 2012, 4, 43014326

B.6.1 Rare earth based lifetime luminescence nanothermometry. Several works have been published proposing rare
earth doped crystals and complexes as good candidates for
LLNth. One of the most interesting cases is that of Ce3+ doped
YAG nanophosphors. Allison et al. reported that when the size
of Ce3+:YAG micro-crystals is reduced to the sub-micrometric
range (down to the nanoscale) the temperature dependence of the
Ce3+ luminescence lifetime changes dramatically.102 Allison et al.
found that while in micrometric Ce3+:YAG crystals the luminescence lifetime was almost temperature independent near room
temperature, the lifetime of Ce3+:YAG nanocrystals becomes
strongly temperature dependent, following an almost linear
decay near room temperature (see Fig. 24). The ability of any
given luminescent system for LFTNth is usually evaluated from
its normalized lifetime thermal coefficient, as, which is defined
as as dsnor(T)/dT where snor(T) is the luminescence decay time
at temperature T normalized to the room temperature value (i.e.
snor(T) sf(T)/sf(25  C)). For nanosized Ce3+:YAG crystals, as is
close to 0.01 per  C, in the 1070  C temperature range. Additionally, the luminescence lifetime of Ce3+:YAG nanocrystals is
in the tens of nanoseconds range and can be considered as a short
lifetime value when compared with other rare earth ions. This is
a very favorable property for LLNth as it increases the
measuring rate. In addition to Ce3+, Eu3+ and Tb3+ have been
recently proposed as good luminescent probes for LLNth in the
070  C temperature range. Yu et al. demonstrated that Eu3+ and
This journal is The Royal Society of Chemistry 2012

Published on 16 May 2012. Downloaded by Universidade Federal de Alagoas on 15/09/2015 15:05:40.

View Article Online

Fig. 24 Temperature induced luminescence lifetime variation in


different rare earth based systems. Dots are experimental data and solid
lines are guides for the eyes. Data were extracted with permissions from
ref. 102104.

Tb3+ complexes, with large luminescence quantum yields, showed


a drastic shortening of the luminescence lifetimes when the
temperature was increased from 0 to 70  C (see Fig. 24).103 The
normalized lifetime thermal coefficients found for Eu3+ and Tb3+
complexes were determined to be close to 0.02 and 0.012 per  C,
respectively. According to Jiangbo et al., these values of as
ensure a temperature resolution of approximately 0.1  C (this
being constant over the entire temperature range due to the linear
relation between lifetime and temperature).103 At this point, the
work of Cai et al. should be mentioned as they report on the
thermal sensitivity of the luminescence lifetime of Er3+ ions
incorporated in microspheres composed of a heavy-metal fluoride glass (ZBLALiP).104 The authors determined that the lifetime decreases monotonously (but not linearly) with temperature
in the range of 170 to 50  C. When compared with the other
rare earth based LLNth systems, Er3+ has the disadvantage of
a non-linear relation between lifetime and temperature and of an
almost temperature independent lifetime in the vicinity of room
temperature. Nevertheless, the interest of the work proposed by
Cai et al. resides in the fact that the luminescence thermal sensor
is incorporated in a glass microsphere that can be potentially
controlled by optical tweezers.104 For example, this possibility
could lead to the achievement of three-dimensional thermal
images by the three-dimensional scanning of the thermal sensor.
Despite their vast potential, to the best of our knowledge, there
are no reported thermal images obtained by rare earth based
LLNth.
B.6.2 Dye-based lifetime luminescence nanothermometry. The
temperature-induced reduction of the luminescence lifetime of
luminescent organic dyes is a well-known phenomenon and is
especially noticeable in the case of the ubiquitous Rhodamine B
dye. Fig. 25(a) shows the Rhodamine B luminescence lifetime as
a function of temperature in the 1070  C range. The clear
temperaturelifetime relation, obeying the Arrhenius equation,
has been tentatively explained in the past in terms of the intramolecular activated non-radiative decay from the first excited
This journal is The Royal Society of Chemistry 2012

singlet of the Rhodamine B molecule.105107 A luminescence


lifetime reduction close to 75% is observed. This yields an averaged normalized lifetime thermal coefficient of approximately
0.0125 per  C, although the nonlinear relation between lifetime
and temperature leads to larger lifetime based sensitivities at
lower temperatures. In the biological temperature range (25
45 mine B is estimated to be as y 0.03 per  C, allowing for
thermal reading with resolutions better than 1  C. Rhodamine B
has been extensively used for LLNth specifically of fluids in the
presence of external heating sources. The additional possibility of
multiphoton excitation of the characteristic red luminescence of
Rhodamine B has opened the possibility of true 3D thermal
imaging of fluids with sub-micrometric spatial resolution.108,109
This has been, indeed, already demonstrated by Benninger et al.
who were able to obtain a 3D thermal image of a 130  40  100
mm3 micro-channel device specifically designed for polymerase
chain reaction (PCR) and is shown in Fig. 25(b).110 In this case,
one of the micro-channel walls (upper one) was held at a constant
temperature of 73  C by using an integrated heating source. It is
clear from this thermal image that the intra-channel temperature
in a PCR device is far from being constant. Indeed, variation as
large as 5  C has been measured, this being a non-negligible
temperature variation. As a matter of fact, such intra-channel

Fig. 25 (a) Temperature dependence of the Rhodamine B luminescence


lifetime. Dots are experimental data and the solid line is a guide for the
eyes. (b) Thermal image obtained by Rhodamine B based LLNth of a 130
 40  100 mm3 micro-channel device specially designed for polymerase
chain reaction. Reprinted with permission from ref. 110. Copyright 2006
American Chemical Society.

Nanoscale, 2012, 4, 43014326 | 4321

Published on 16 May 2012. Downloaded by Universidade Federal de Alagoas on 15/09/2015 15:05:40.

View Article Online

temperature variations could cause important variations in the


PCR reaction efficiency.111
Finally, the work recently published by Bennet et al. should be
mentioned due to its original approach for thermal imaging and
sensing of a microfluidic device.112 In this case, the authors did
not fill the microchannel with a Rhodamine B solution, rather
they encapsulated the temperature-sensitive Rhodamine B within
a micro-droplet, which could be held and manipulated in the
microfluidic flow by using optical tweezers. This novel approach
retains the capability of LLNth to deliver quantitative mapping
of microfluidic temperatures without the disadvantageous need
to introduce a luminescent dye that pervades the entire microfluidic system. As a disadvantage, it should be mentioned that in
this case the spatial resolution of the thermal imaging technique
is limited by the size of the micro-droplet containing the
Rhodamine B. For instance, the size of the micro-droplets used
by Bennet et al. were close to 10 mm in diameter,112 so that the
final spatial resolution of their technique was far from that
achieved by a direct incorporation of Rhodamine B into the
micro-channel, given by the spatial resolution of the microscopy
system (typically close to 1 mm).
Although Rhodamine B is the most popular dye for lifetime
thermal sensing, it has some serious drawbacks such as its low
solubility in water. This makes it unsatisfactory for measuring
aqueous microfluidics temperatures because of problems such as
undesired deposition on surfaces. In this respect it should be
mentioned that Mendels et al. have used Kiton Red, a watersoluble derivative of Rhodamine B, that shows the same lifetime
temperature response as the parent fluorophore, for LLNth of
aqueous microfluidic systems.113 The ability of this alternative
lifetime based luminescent thermal probe was investigated by
Mendels et al. by comparing the thermal Kiton Red LLNth
based thermal images of a microfluidic T-mixer when two fluids
of very different temperatures (25 and 60  C) were injected, as it
is schematically drawn in Fig. 26. On the left hand side of Fig. 26,
a series of LLNth thermal images are shown as obtained at
different locations of the T-mixer device. These images reveal
how the local temperature of the microfluidic becomes homogenous as the two fluids are mixed along the common branch. The
accuracy of their experimental measurements was checked by
comparing these images with theoretical simulations (right image
in Fig. 26) showing an excellent agreement.113
B.6.3 Polymer-based lifetime luminescence nanothermometry.
The use of luminescent polymers for LLNth is not as widely
known as luminescent dyes despite the recently published results
where they have emerged as highly sensitive lifetime thermal
probes. As it has been previously described in Section B.1.6, there
are luminescent polymers (such as the so-called poly(DBD-AEco-NIPAM)) that, in an aqueous solution, undergo a thermally
induced phase transition that is accompanied by a drastic
decrease in the microenvironmental polarity in the vicinity of the
main polymer chain.52 Since the luminescence properties of the
polymer are affected by the solvent polarity, the phase transition
causes a drastic change in the luminescence properties of the
polymer. This fact has been used for thermal sensing based on the
luminescence intensity variations during phase transition. Graham et al. demonstrated that the phase transition does not only
cause a drastic change in the luminescence efficiency of the
4322 | Nanoscale, 2012, 4, 43014326

Fig. 26 Thermal images (left column) of a microfluidic T-mixer at


different positions. Two fluids at different temperatures (60 and 25  C)
are injected into the microfluidic device. The thermal images were
obtained from the spatial variation of the luminescence lifetime of the
Kiton Red dye. The figure at the right is a schematic representation of the
device and includes the theoretical temperature distribution. Reprinted
with permission from ref. 113. Copyright 2008 Springer.

polymer but also in its luminescence lifetime.114 This fact is


shown in Fig. 27(a) that includes the temperature dependence of
the luminescence lifetime of the poly(DBD-AE-co-NIPAM)
polymer in the vicinity of the phase transition temperature (close
to 32.5  C). As can be observed the lifetime based thermal
sensitivity achieved is large. In fact, the normalized lifetime
thermal coefficient is as large as 0.06 per  C, i.e. one order of
magnitude larger than those typically achieved with rare earth
ions or luminescent dyes. These large values of as are only achieved during the phase transition so that the temperature operation range of the polymer based LLNth is very limited. Despite
this narrow operation range, luminescent polymers have been
already used in thermal imaging of PCR microfluidic devices. A
typical example is given in Fig. 27(b), showing the thermal image
of a microfluidic chamber positioned between a heating source
(at the left-hand of the chamber) and a cold sink (at the righthand of the chamber). The LLNth image clearly shows the
temperature gradient across the chamber. The temperature
This journal is The Royal Society of Chemistry 2012

View Article Online

Published on 16 May 2012. Downloaded by Universidade Federal de Alagoas on 15/09/2015 15:05:40.

cytoplasm was found to be strongly dependent on the particular


cell cycle. In addition, authors also found that the majority of
cells showed a well-localized temperature singularity (indicated
by the arrowheads in Fig. 28(b)) that was tentatively associated
with a centrosome-specific thermogenesis.116,117 Results obtained
by Okabe et al. demonstrated that polymeric LLNth could be
used to identify the relationships between the temperature and
organelle functions.
B.6.4 Quantum dot-based lifetime luminescence nanothermometry. As described in Section B.3, the energy level
diagram of QDs is temperature dependent due to a combination
of different factors, including temperature-induced changes in
QD size, band-gap, electronphonon interaction and confinement energy. This temperature induced variation in the QD
energy level scheme is known to cause a luminescence shift that
has been already used for thermal sensing and imaging based on
SLNth. As it has been also described, this temperature induced

Fig. 27 (a) Temperature dependence of the luminescence lifetime of the


poly(DBD-AE-co-NIPAM) polymer in the vicinity of the phase transition temperature (close to 32.5  C). (b) Thermal image of a microfluidic
chamber positioned between a heating source (left-hand of the chamber)
and a cold sink (right-hand of the chamber). Data were obtained from the
analysis of the spatial variation of the luminescence lifetime of the
poly(DBD-AE-co-NIPAM) polymer that is introduced into the microchamber in an aqueous solution. Reproduced from ref. 114 with
permission from The Royal Society of Chemistry.

resolution of the thermal image included in Fig. 27(b) has been


estimated by the authors to be less than 0.1  C.114 The achievement of these outstanding thermal sensitivities within the physiologically relevant temperature range makes luminescent
polymers, such as poly(DBD-AE-co-NIPAM), valuable lifetime
based thermal probes for biomedical applications, in which
temperature control is critical.
Very recently, Okabe et al. have taken advantage of the
outstanding properties of luminescent polymers for LLNth for
the achievement of high-resolution thermal images of single cells
with spatial and temperature resolutions as good as 200 nm and
0.18  C, respectively.115 Okabe et al. incubated COS7 cells in
a solution containing a NNPAM based luminescent polymer
with a phase transition at around 35  C. During phase transition
the luminescence lifetime of the polymer was found to increase
linearly with temperature with a normalized lifetime thermal
coefficient as large as 0.065 per  C.115 Once the luminescent
polymer was incorporated into the cells Okabe et al. obtained the
corresponding thermal image from the analysis of the spatial
variation of the luminescence lifetime of the polymer. Representative data are included in Fig. 28. Okabe et al. found
experimental evidence of the existence of a highly inhomogeneous intracellular temperature distribution. Indeed, as can be
observed in Fig. 28(b), the nucleus of the COS7 cells showed
significantly higher temperatures than the cytoplasm. The
observed temperature gap between the nucleus and the
This journal is The Royal Society of Chemistry 2012

Fig. 28 (a) Confocal luminescence image of living COS7 cells incubated


with a luminescent polymeric thermometer. (b) Thermal image of the
same cells as those shown in (a) as obtained by LLNth. Arrowheads
indicate the location of the well localized hot-spots observed in the
majority of cells examined by Okabe et al. in ref. 115. Reprinted with
permission from ref. 115. Copyright 2012 Nature Publishing Group.

Nanoscale, 2012, 4, 43014326 | 4323

Published on 16 May 2012. Downloaded by Universidade Federal de Alagoas on 15/09/2015 15:05:40.

View Article Online

spectral shift is accompanied by a relevant luminescence


quenching. The observed temperature induced luminescence
quenching is attributed to several causes. On the one hand, the
luminescence efficiency of the excitonic level decreases as
a consequence of the thermal activation of multiphonon assisted
electronhole recombination. On the other hand, temperature
enhances the energy transfer probability from bulk (excitons) to
non-radiative surface (trap) states (since thermally induced QD
expansion reduces the energy separation between excitonic
states) and favors energy migration among them until a nonradiative state is reached.118121 Both effects lead to an increment
in the non-radiative de-excitation probability and, thus, in the
total de-excitation probability. Therefore, it is expected that
temperature increments in the surroundings of ambient temperature would also cause a decrease in the luminescence lifetime.
This is indeed what has been recently demonstrated by HaroGonzalez et al. who examined in detail how the CdTe QD
luminescence lifetime varies within the biophysical thermal range
(2550  C).122 The authors found that the luminescence lifetime
decreases almost linearly for all the QD sizes, however, the
smaller the CdTe QD, the larger the temperature induced lifetime
reduction. Fig. 29(a) shows, as an example, the luminescence
decay curves of 1.2 nm CdTe QDs obtained at both 30 and 50  C
and reveal a drastic temperature induced luminescence decay

Fig. 29 (a) The luminescence decay curves of 1.2 nm CdTe QDs


obtained at 30 and 50  C. (b) Luminescence decay time of 1.2 nm CdTe
QDs as a function of temperature. Dots are experimental data and the
solid line is a guide for the eyes.

4324 | Nanoscale, 2012, 4, 43014326

time reduction. This is even more noticeable in Fig. 29(b) that


shows the luminescence decay time of 1.2 nm CdTe QDs as
a function of temperature. As can be observed, luminescence
lifetime decreases linearly with temperature. From Fig. 29(b),
Haro-Gonz
alez et al. estimated, for CdTe QDs, a luminescence
lifetime thermal coefficient as large as 0.017 per  C, this being one
of the largest as values ever reported (superior than that reported
for Rhodamine B).122 This fact makes CdTe QDs very promising
luminescent probes for high (spatial and thermal) resolution
luminescence lifetime thermal imaging. Haro-Gonzalez et al.
already demonstrated the ability of CdTe QDs for lifetime
thermal sensing in micro-fluidics and the obvious next step is
their application for thermal imaging of biosystems (such as
single cells and tissues).

C. Conclusions
In summary, we have presented a detailed review of the diverse
methods proposed to date for the achievement of high-resolution
thermal sensing from the analysis of luminescence. It has been
shown that many luminescent systems (including polymers,
organic dyes, rare earth doped crystals, rare earth doped nanocrystals, semiconductor nanocrystals, rare earth doped
complexes, phosphorous and quantum dots) can be used as basic
light-emitting materials for nanothermometry. It is not possible
to highlight one material over the rest since the most suitable one
would depend on the actual system to be thermally imaged.
This review describes in detail how these luminescence systems
have been previously used for high-resolution thermal imaging of
a great variety of systems including living cells, microfluidics,
electronic nano-devices and solid state lasers.
The continuous development of new microscopy techniques
coupled with novel and cutting edge synthesis processes (allowing for the rational design of novel luminescent materials) will
ensure the speedy development of luminescence nanothermometry, although working principles will be probably the
same as those summarized in this review.
A plethora of examples have been described and explained in
detail. The achievement of temperature resolutions well below
the  C limit has fostered the use of luminescent nanothermometers in biological applications. This is by far the most
challenging of all possible applications described in this review,
due to the complex nature of the biological milieu. To overcome
some of the hurdles associated with working in a biological
environment, employing luminescent nanothermometers optically excited in the near-infrared (NIR) region is particularly
promising. In combination with multiphoton microscopy, this
would allow for high-resolution, three-dimensional thermal
imaging of living specimens. Although a great deal of effort has
been invested in the development of NIR excited luminescent
nanothermometers, their real-world application in three-dimensional thermal imaging of bio-systems is far from being a reality.
For the most part, this is because these nanothermometers
typically emit in the visible range, where tissue transparency is
low. In our opinion the forthcoming advances in the field of
luminescence nanothermometry will be propelled by the development of luminescent nanothermometers, which operate within
the biological window (700900 nm). That is, both their excitation and emission wavelengths lie within this optimal window.
This journal is The Royal Society of Chemistry 2012

Published on 16 May 2012. Downloaded by Universidade Federal de Alagoas on 15/09/2015 15:05:40.

View Article Online

Furthermore, we are also firmly convinced that the most


groundbreaking results in the field of luminescent nanothermometry will be obtained by using ratiometric thermal
sensors, such as those described throughout this review. The
main reasons, of course, are that exploiting the luminescence
ratios of particular emission bands to obtain the thermal reading
is not affected by local intensity fluctuations (i.e. concentration
of emitting centers) but moreover, would require simple experimental instrumentation.
Finally, we also believe that luminescent nanothermometry
will be widely used in the future not only for disease detection but
also for monitoring and control of laser induced thermal treatments (i.e. hyperthermia). In this sense, the development of
multi-functional nanoparticles capable of simultaneous laser
induced heating and luminescent nanothermometry will yield
significant advancements in disease diagnoses and therapeutics in
the near future.

Notes and references


1 B. A. Shakouri, Proc. IEEE, 2006, 94, 16131638.
2 J. I. M. Colvin, Microelectron. Reliab., 1998, 38, 17051714.
3 A. Bar-Cohen and P. Wang, Microgravity Sci. Technol., 2009, 21,
351359.
4 D. H. Jundt, Opt. Lett., 1997, 22, 15531555.
5 J. J. Romero, C. Arag
o, J. A. Gonzalo, D. Jaque and J. Garca Sole,
J. Appl. Phys., 2003, 93, 3111.
6 G. Nenna, G. Flaminio, T. Fasolino, C. Minarini, R. Miscioscia,
D. Palumbo and M. Pellegrino, Macromol. Symp., 2007, 247, 326
332.
7 A. Ashkin, J. M. Dziedzic and T. Yamane, Nature, 1987, 330, 771.
8 S. Ebert, K. Travis, B. Lincoln and J. Guck, Opt. Express, 2007, 15,
1549315499.
9 M. Suzuki, V. Tseeb, K. Oyama and S. Ishiwata, Biophys. J., 2007,
92, L46L48.
10 O. Zohar, M. Ikeda, H. Shinagawa, H. Inoue, H. Nakamura,
D. Elbaum, D. L. Alkon and T. Yoshioka, Biophys. J., 1998, 74,
8289.
11 F. Tardieu, M. Reymond, P. Hamard, C. Granier and B. Muller, J.
Exp. Bot., 2000, 51, 15051514.
12 G. N. Somero, Annu. Rev. Physiol., 1995, 57, 4368.
13 M. Stark, Cancer, 1974, 33, 16641670.
14 J. S. Michaelson, S. Satija, D. Kopans, R. Moore, M. Silverstein,
A. Comegno, K. Hughes, A. Taghian, S. Powell and B. Smith,
Cancer, 2003, 98, 21142124.
15 T. B. Huff, L. Tong, Y. Zhao, M. N. Hansen, J.-X. Cheng and
A. Wei, Nanomedicine, 2007, 2, 125132.
16 J. Lee and N. A. Kotov, Nano Today, 2007, 2, 4851.
17 P. R. N. Childs, J. R. Greenwood and C. A. Long, Rev. Sci. Instrum.,
2000, 71, 29592978.
18 L. Shi, O. Kwon, A. C. Miner and A. Majumdar, Design, 2001, 10,
370378.
19 A. Majumdar, Annu. Rev. Mater. Sci., 1999, 29, 505585.
20 J. Park and M. Taya, Electron. Lett., 2004, 40, 45.
21 O. Nakabeppu, M. Chandrachood, Y. Wu, J. Lai and A. Majumdar,
Appl. Phys. Lett., 1995, 66, 694696.
22 K. Kim, J. Chung, G. Hwang, O. Kwon and J. S. Lee, ACS Nano,
2011, 5, 87008709.
23 G. Baffou, P. Bon, J. Savatier, J. Polleux, M. Zhu, M. Merlin,
H. Rigneult and S. Monneret, ACS Nano, 2012, 6, 24522458.
24 J. Christofferson and A. Shakouri, Rev. Sci. Instrum., 2005, 76,
024903.
25 T. Beechem, S. Graham, S. P. Kearney, L. M. Phinney and
J. R. Serrano, Rev. Sci. Instrum., 2007, 78, 061301.
26 M. Kuball, J. M. Hayes, M. J. Uren, T. Martin, J. C. H. Birbeck,
R. S. Balmer and B. T. Hughes, IEEE Electron Device Lett., 2002,
23, 712.
27 S. Chenais, S. Forget, F. Druon, F. Balembois and P. Georges, Appl.
Phys. B: Lasers Opt., 2004, 79, 221224.

This journal is The Royal Society of Chemistry 2012

28 K. S. Schwartzkopf-Genswein and J. M. Stookey, Can. J. Anim. Sci.,


1997, 77, 577583.
29 S. Huth and O. Breitenstein, J. Appl. Phys., 2000, 88, 40004003.
30 G. F. Imbusch and B. Henderson, Optical Spectroscopy of Inorganic
Solids, Oxford Science Publications, London, 2006.
31 X. Michalet, F. F. Pinaud, L. A. Bentolila, J. M. Tsay, S. Doose,
J. J. Li, G. Sundaresan, A. M. Wu, S. S. Gambhir and S. Weiss,
Science, 2012, 538, 538544.
32 O. I. Micic, H. M. Cheong, H. Fu, A. Zunger, J. R. Sprague,
A. Mascarenhas and A. J. Nozik, J. Phys. Chem. B, 1997, 101,
49044912.
33 G. W. Walker, V. C. Sundar, C. M. Rudzinski, A. W. Wun,
M. G. Bawendi and D. G. Nocera, Appl. Phys. Lett., 2003, 83,
35553557.
34 B. Han, W. L. Hanson, K. Bensalah, A. Tuncel, J. M. Stern and
J. A. Cadeddu, Ann. Biomed. Eng., 2009, 37, 12301239.
35 S. Wang, S. Westcott and W. Chen, J. Phys. Chem. B, 2002, 106,
1120311209.
36 L. Aigouy, B. Samson, G. Julie, V. Mathet, N. Lequeux, C. N.
Allen, H. Diaf and B. Dubertret, Rev. Sci. Instrum., 2006, 77,
063702.
37 J. Lee, A. O. Govorov and N. A. Kotov, Angew. Chem., Int. Ed.,
2005, 44, 74397442.
38 J. R. Lakowicz, I. Gryczynski, V. Bogdanov and J. Kusba, J. Phys.
Chem., 1994, 98, 334342.
39 D. Ross, M. Gaitan and L. E. Locascio, Anal. Chem., 2001, 73,
41174123.
40 P. L
ow, B. Kim, N. Takama and C. Bergaud, Small, 2008, 4, 908
914.
41 R. Samy, T. Glawdel and C. L. Ren, Anal. Chem., 2008, 80, 369375.
42 S. W. Allison, Rev. Sci. Instrum., 1997, 68, 26152650.
43 S. V. Yap, R. M. Ranson, W. M. Cranton, D. C. Koutsogeorgis and
G. B. Hix, J. Lumin., 2009, 129, 416422.
44 H. Fonger and C. W. Struck, J. Chem. Phys., 1970, 52, 63646366.
45 C. Gota, K. Okabe, T. Funatsu, Y. Harada and S. Uchiyama, J. Am.
Chem. Soc., 2009, 131, 27662767.
46 A. Rassamesard, Y.-F. Huang, H.-Y. Lee, T.-S. Lim, M. C. Li,
J. D. White, J. H. Hodak, T. Osotchan, K. Y. Peng, S. A. Chen,
J.-H. Hsu, O. M. Hayashi and W. Fann, J. Phys. Chem. C, 2009,
113, 1868118688.
47 S. Uchiyama, Y. Matsumura, A. P. de Silva and K. Iwai, Anal.
Chem., 2004, 76, 17931798.
48 I. F. Pierola, D. Radic and L. Gargallo, Chem. Mater., 2006, 18,
20812085.
49 C. David, M. Piens and G. Geukens, Eur. Polym. J., 1972, 8, 1019
1023.
50 G. Kwak, S. Fukao, M. Fujiki and T. Sakaguchi, Chem. Mater.,
2006, 18, 20812085.
51 S. Uchiyama, N. Kawai, A. P. de Silva and K. Iwai, J. Am. Chem.
Soc., 2004, 126, 30323033.
52 S. Uchiyama, Y. Matsumura, A. P. de Silva and K. Iwai, Anal.
Chem., 2003, 75, 59265935.
53 V. A. Vlaskin, N. Janssen, J. van Rijssel, R. Beaulac and
D. R. Gamelin, Nano Lett., 2010, 10, 36703674.
54 E. J. McLaurin, V. A. Vlaskin and D. R. Gamelin, J. Am. Chem.
Soc., 2011, 133, 1497814980.
55 C.-H. Hsia, A. Wuttig and H. Yang, ACS Nano, 2011, 5, 95119522.
56 D. J. Bizzak, Int. J. Heat Mass Transfer, 1995, 38, 267274.
57 P. Haro-Gonzalez, I. R. Martn, L. L. Martn, S. F. Le
on-Luis,
C. Perez-Rodrguez and V. Lavn, Opt. Mater., 2011, 33, 742745.
58 D. J. Bizzak and M. K. Chyu, Rev. Sci. Instrum., 1994, 65, 102107.
59 A. L. Heyes, S. Seefeldt and J. P. Feist, Opt. Laser Technol., 2006, 38,
257265.
60 M. T. Carlson, A. Khan and H. H. Richardson, Nano Lett., 2011, 11,
10611069.
61 Z. P. Cai and H. Y. Xu, Sens. Actuators, A, 2003, 108, 187192.
62 B. S. Cao, Y. Y. He, Z. Q. Feng, Y. S. Li and B. Dong, Sens.
Actuators, B, 2011, 159, 811.
63 D. Li, Y. Wang, X. Zhang, K. Yang, L. Liu and Y. Song, Opt.
Commun., 2012, 285, 19251928.
64 J. Petit, B. Viana and P. Goldner, Opt. Express, 2011, 19, 11381146.
65 F. Vetrone, R. Naccache, A. Zamarr
on, A. Juarranz de la Fuente,
F. Sanz-Rodrguez, L. Martinez Maestro, E. Martn Rodriguez,
D. Jaque, J. Garca Sole and J. A. Capobianco, ACS Nano, 2010,
4, 32543258.

Nanoscale, 2012, 4, 43014326 | 4325

Published on 16 May 2012. Downloaded by Universidade Federal de Alagoas on 15/09/2015 15:05:40.

View Article Online

66 M. Quintanilla, E. Cantelar, F. Cuss


o, M. Villegas and
A. C. Caballero, Appl. Phys. Express, 2011, 4, 022601.
67 M. A. R. C. Alencar, G. S. Maciel, C. B. de Ara
ujo and A. Patra,
Appl. Phys. Lett., 2004, 84, 47534755.
68 E. Cantelar and F. Cusso, Appl. Phys. B: Lasers Opt., 1999, 69, 29
33.
69 M. T. Carlson, A. J. Green and H. H. Richardson, Nano Lett., 2012,
12, 15341537.
70 C. D. S. Brites, P. P. Lima, N. J. O. Silva, A. Millan, V. S. Amaral,
F. Palacio and L. D. Carlos, New J. Chem., 2011, 35, 1177
1183.
71 L. Aigouy, B. Samson, E. Sa, L. Peter and C. Bergaud, Thermal
Nanosystems and Nanomaterials, Springer Berlin Heidelberg,
Berlin, Heidelberg, 2009, vol. 118.
72 E. Sadi, B. Samson, L. Aigouy, S. Volz, P. L
ow, C. Bergaud and
M. Mortier, Nanotechnology, 2009, 20, 115703.
73 E. Sadi, N. Babinet, L. Lalouat, J. Lesueur, L. Aigouy, S. Volz,
J. Labeguerie-Egea and M. Mortier, Small, 2011, 7, 259264.
74 B. Samson, L. Aigouy, P. Low, C. Bergaud, B. J. Kim and
M. Mortier, Appl. Phys. Lett., 2008, 92, 023101.
75 L. Aigouy, L. Lalouat, M. Mortier, P. L
ow and C. Bergaud, Rev.
Sci. Instrum., 2011, 82, 36106.
76 F. Vetrone, R. Naccache, A. Juarranz de la Fuente, F. SanzRodrguez, A. Blazquez-Castro, E. M. Rodriguez, D. Jaque,
J. Garca Sole and J. A. Capobianco, Nanoscale, 2010, 2, 495498.
77 L. M. Maestro, E. M. Rodriguez, F. Vetrone, R. Naccache,
H. L. Ramirez, D. Jaque, J. A. Capobianco and J. Garca Sole,
Opt. Express, 2010, 18, 2354423553.
78 D. K. Chatterjee, A. J. Rufaihah and Y. Zhang, Biomaterials, 2008,
29, 937943.
79 M. Wang, C.-C. Mi, W.-X. Wang, C.-H. Liu, Y.-F. Wu, Z.-R. Xu,
C.-B. Mao and S.-K. Xu, ACS Nano, 2009, 3, 15801586.
80 D. Jaque, L. M. Maestro, E. Escudero, E. M. Rodrguez,
J. A. Capobianco, F. Vetrone, A. Juarranz de la Fuente, F. SanzRodrguez, M. C. Iglesias-de la Cruz, C. Jacinto, U. Rocha and
J. Garca Sole, J. Lumin., 2011, DOI: 10.1016/j.jlumin.2011.12.022,
in press.
81 C. D. S. Brites, P. P. Lima, N. J. O. Silva, A. Millan, V. S. Amaral,
F. Palacio and L. D. Carlos, Adv. Mater., 2010, 22, 44994504.
82 Y. Cui, H. Xu, Y. Yue, Z. Guo, J. Yu, Z. Chen, J. Gao, Y. Yang,
G. Qian and B. Chen, J. Am. Chem. Soc., 2012, 134, 39793982.
83 N. Ishiwada, S. Fujioka, T. Ueda and T. Yokomori, Opt. Lett., 2011,
36, 760762.
84 M. Seaver and J. R. Peele, Appl. Opt., 1990, 29, 49564961.
85 E. J. G. Peterman, F. Gittes and C. F. Schmidt, Biophys. J., 2003, 84,
13081316.
86 A. Narayanaswamy, L. F. Feiner, A. Meijerink and P. J. van der
Zaag, ACSNano, 2009, 3, 25392546.
87 A. Olkhovets, R.-C. Hsu, A. Lipovskii and F. Wise, Phys. Rev. Lett.,
1998, 81, 35393542.
88 Q. Dai, Y. Zhang, Y. Wang, M. Z. Hu, B. Zou, Y. Wang and
W. W. Yu, Langmuir, 2010, 26, 1143511440.
89 L. Martinez Maestro, E. Martn Rodrguez, F. Sanz Rodrguez,
M. C. Iglesias-de la Cruz, A. Juarranz, R. Naccache, F. Vetrone,
D. Jaque, J. A. Capobianco and J. Garca Sole, Nano Lett., 2010,
10, 51095115.
90 L. M. Maestro, C. Jacinto, U. R. Silva, F. Vetrone,
J. A. Capobianco, D. Jaque and J. Garca Sole, Small, 2011, 7,
17741778.
91 S. Li, K. Zhang, J.-M. Yang, L. Lin and H. Yang, Nano Lett., 2007,
7, 31023105.

4326 | Nanoscale, 2012, 4, 43014326

92 R. S. Yang, L. W. Chang, J.-P. Wu, M.-H. Tsai, H.-J. Wang,


Y.-C. Kuo, T.-K. Yeh, C. S. Yang and P. Lin, Environ. Health
Perspect., 2007, 115, 13391343.
93 W. J. Parak, T. Pellegrino and C. Plank, Nanotechnology, 2005, 16,
R9R25.
94 D. R. Larson, W. R. Zipfel, R. M. Williams, S. W. Clark,
M. P. Bruchez, F. W. Wise and W. W. Webb, Science, 2003, 300,
14341436.
95 L. M. Maestro, J. E. Ramrez-Hernandez, N. Bogdan,
J. A. Capobianco, F. Vetrone, J. Garca Sole and D. Jaque,
Nanoscale, 2012, 4, 298302.
96 J.-M. Yang, H. Yang and L. Lin, ACS Nano, 2011, 5, 50675071.
97 A. Kawski, Crit. Rev. Anal. Chem., 1993, 23, 459529.
98 B. Valeur, Molecular Fluorescence: Principles and Applications,
Wiley-VCH, 2002.
99 G. Baffou, M. P. Kreuzer, F. Kulzer and R. Quidant, Opt. Express,
2009, 17, 32913298.
100 A. Kaminskii, Laser Photonics Rev., 2007, 1, 93177.
101 A. Benayas, E. Escuder and D. Jaque, Appl. Phys. B: Lasers Opt.,
2012, 107, 697701.
102 S. W. Allison, G. T. Gillies, A. J. Rondinone and M. R. Cates,
Nanotechnology, 2003, 14, 859863.
103 J. Yu, L. Sun, H. Peng and M. I. J. Stich, J. Mater. Chem., 2010, 20,
69756981.
104 Z. P. Cai, L. Xiao, H. Y. Xu and M. Mortier, J. Lumin., 2009, 129,
19941996.
105 D.-A. Mendels, E. M. Graham, S. W. Magennis, A. C. Jones and
F. Mendels, Microfluid. Nanofluid., 2008, 5, 603617.
106 T. Karstens and K. Kobs, J. Phys. Chem., 1980, 84, 18711872.
107 J. J. Shah, M. Gaitan and J. Geist, Anal. Chem., 2009, 81, 82608263.
108 A. Nag and D. Goswami, J. Photochem. Photobiol.,A, 2009, 206,
188197.
109 P. C. Jha, Y. Wsang and H. Agren, ChemPhysChem, 2008, 9, 111
116.
110 R. K. P. Benninger, Y. Koc, O. Hofmann, J. Requejo-Isidro,
M. A. A. Neil, P. M. W. French and A. J. deMello, Anal. Chem.,
2006, 78, 22722278.
111 D. Y. Wu, L. Ugozzoli, B. K. Pal, J. Qian and R. B. Wallace, DNA
Cell Biol., 1991, 10, 233238.
112 M. A. Bennet, P. R. Richardson, J. Arlt, A. McCarthy, G. S. Buller
and A. C. Jones, Lab Chip, 2011, 11, 38213828.
113 D.-A. Mendels, E. M. Graham, S. W. Magennis, A. C. Jones and
F. Mendels, Microfluid. Nanofluid., 2008, 5, 603617.
114 E. M. Graham, K. Iwai, S. Uchiyama, A. P. de Silva,
S. W. Magennis and A. C. Jones, Lab Chip, 2010, 10, 12671273.
115 K. Okabe, N. Inada, C. Gota, Y. Harada, T. Funatsu and
S. Uchiyama, Nat. Commun., 2012, 3, 705.
116 S. Doxsey, Nat. Rev. Mol. Cell Biol., 2001, 2, 688698.
117 J. S. Andersen, C. J. Wilkinson, T. Mayor, P. Mortensen, E. A. Nigg
and M. Mann, Nature, 2003, 426, 570574.
118 A. M. Kapitonov, A. P. Stupak, S. V. Gaponenko, E. P. Petrov,
A. L. Rogach and A. Eychm
uller, J. Phys. Chem. B, 1999, 103,
1010910113.
119 M. Leistikow, J. Johansen, A. Kettelarij, P. Lodahl and W. Vos,
Phys. Rev. B: Condens. Matter Mater. Phys., 2009, 79, 19.
120 G. Schlegel, J. Bohnenberger, I. Potapova and A. Mews, Phys. Rev.
Lett., 2002, 88, 137401.
121 Y. Ebenstein, T. Mokari and U. Banin, J. Phys. Chem. B, 2004, 108,
9399.
122 P. Haro-Gonzalez, L. Martinez Maestro, I. R. Martin, J. Garca Sole
and D. Jaque, Small, DOI: 10.1002/sml.201102736, submitted.

This journal is The Royal Society of Chemistry 2012

Das könnte Ihnen auch gefallen