You are on page 1of 24

Applied Spectroscopy Reviews

ISSN: 0570-4928 (Print) 1520-569X (Online) Journal homepage: https://www.tandfonline.com/loi/laps20

A practical review of shorter than excitation


wavelength light emission processes

J. E. Moffatt, G. Tsiminis, E. Klantsataya, T. J. de Prinse, D. Ottaway & N. A.


Spooner

To cite this article: J. E. Moffatt, G. Tsiminis, E. Klantsataya, T. J. de Prinse, D. Ottaway & N. A.


Spooner (2019): A practical review of shorter than excitation wavelength light emission processes,
Applied Spectroscopy Reviews, DOI: 10.1080/05704928.2019.1672712

To link to this article: https://doi.org/10.1080/05704928.2019.1672712

Published online: 12 Oct 2019.

Submit your article to this journal

View related articles

View Crossmark data

Full Terms & Conditions of access and use can be found at


https://www.tandfonline.com/action/journalInformation?journalCode=laps20
APPLIED SPECTROSCOPY REVIEWS
https://doi.org/10.1080/05704928.2019.1672712

A practical review of shorter than excitation wavelength


light emission processes
J. E. Moffatta,b, G. Tsiminisa,b,c , E. Klantsatayaa , T. J. de Prinsea ,
D. Ottawaya,b,d , and N. A. Spoonera,b,e
a
Institute for Photonics and Advanced Sensing, School of Physical Sciences, The University of Adelaide,
Adelaide, Australia; bCRC for Optimising Resource Extraction, Brisbane, Australia; cARC Centre of
Excellence for Nanoscale BioPhotonics, School of Medicine, The University of Adelaide, Adelaide,
Australia; dARC Centre of Excellence for Gravitational Wave Discovery, OzGrav, The University of
Adelaide, Adelaide, Australia; eDefence Science and Technology Group, Adelaide, Australia

ABSTRACT KEYWORDS
Shorter-than-excitation-wavelength (STEW) optical emissions, where Upconversion; excited state
photons originating in a material have higher energies than those absorption; fluorescence
that created them, have in the past few decades become important upconversion; anti-Stokes
luminescence; two-photon
in science and medicine, with applications ranging from improving absorption; nonlinear optics;
the efficiency of solar cells to creating new lasers and performing materials science
background-free microscopy of biological samples. Assigning and
predicting the origin of STEW emissions is critical for accelerating
development and applications of new processes and materials. In
this review, we examine the different processes underlying STEW
emissions and outline pathways to identify them using readily avail-
able experimental techniques.

Introduction
Fluorescence upconversion and other anti-Stokes luminescence processes have rapidly
grown in application to diverse fields, including chemical sensing [1, 2], medical imag-
ing [3–5], medical dose distribution [6, 7], solar cell development [8–10], anti-counter-
feiting [11], and novel wavelength generation [12, 13]. The push for wavelength-specific
and efficient upconversion materials has resulted in the synthesis of new complex mate-
rials with features such as intricate host structures [14–17] and doping of multiple spe-
cies [18–20]. The increasing complexity of these materials can make it difficult to
correctly identify the pathway through which the observed emission is generated, and
consequently hampers progress in development of applications.
Shorter-than-excitation wavelength (STEW) processes are light-emitting phenomena
in which the photons emitted by a material have higher energy (shorter wavelength)
than the incident photons that create them. Common STEW processes include second
harmonic generation (SHG) [21, 22] and anti-Stokes processes such as two-photon
absorption (TPA) [23], anti-Stokes Raman [24], and upconversion [25–27].

CONTACT G. Tsiminis georgios.tsiminis@adelaide.edu.au Institute for Photonics and Advanced Sensing, School of
Physical Sciences, The University of Adelaide, Adelaide, South Australia, Australia.
Color versions of one or more of the figures in the article can be found online at www.tandfonline.com/laps.
ß 2019 Taylor & Francis Group, LLC
2 J. E. MOFFATT ET AL.

In this article, we start by clarifying a nomenclature: STEW emission is often called


“upconversion” in many fields (e.g. material synthesis), but in other fields (e.g. optics)
“upconversion” refers to a subset of STEW processes. Here, we will only refer to
“upconversion” if all fields agree on the nomenclature to avoid confusion. In this review, we
provide a set of tools and experimental methods to aid correct identification of and differen-
tiation between the various STEW processes. We start by briefly outlining the theory of each
process. Then we show that the origin of an observed emission can be identified by varying
the excitation wavelength, power and time duration, and observing the corresponding
changes to the emitted light. This procedure leads to new insight into the properties of
the material.

Non-electronic transition-based pathways


Sum frequency generation
Sum frequency generation (SFG) occurs when light travels through certain transparent
media. Light, being an electromagnetic (EM) wave, affects the charged parts (e.g. elec-
trons, dipoles, ions) of the medium through which it passes. When an EM wave trav-
els through a transparent optical medium it induces polarization changes that
propagate alongside the EM wave as a polarization wave that, in turn, produces a
second EM wave. Normally, this EM wave has the same frequency as the original
wave but for high light intensities harmonic frequencies of the EM wave can also be
generated. This process accounts for SHG and third harmonic generation. A com-
monly used SHG example is frequency doubled Nd:YAG lasers that emit at 532 nm
and are used for medical purposes [28], Raman spectroscopy [29], and as high power
lasers [30]. For example, a recent development based on SFG enables the production
of tunable deep-UV light from optical parametric oscillators (OPOs) that use the 3rd
harmonic of a Nd:YAG laser to pump a nonlinear crystal, producing a set of tunable
outputs in the region of 410–2500 nm. While UV production to reach the UV range
can be performed by frequency-doubling the tunable output, in principle enabling
access in the UV region down to 210.5 nm, the limited efficiency of this process results
in low output energies, restricting the availability of high energy UV sources for mate-
rials studies [29]. By using SFG to mix residual 532 nm light from the pump laser
with the OPO output, a new generation of high-efficiency tunable UV laser sources
has been created [31].
Wave-mixing is another SFG method in which multiple frequencies of light enter a
material. The material response generates new frequencies of light, at the sum of the
input frequencies and the difference between them, depleting the energy of the original
frequencies. A STEW example is three-wave mixing, in which two frequencies, f1 and f2
mix in a material and a third frequency, f3, is generated such that f3 ¼ f1 þ f2. These
processes are often described using so-called virtual energy levels, which are temporary
superpositions of energy levels with instantaneous lifetimes, as illustrated in Figure 1a
and b. Three-wave mixing is used in some opto-electrical circuits [32], where pulses of
light can be selectively amplified and shaped to enable greater operation efficiencies for
pulsed lasers and the detection of spectroscopic signals [33].
APPLIED SPECTROSCOPY REVIEWS 3

Figure 1. Energy level diagrams depicting (a) second harmonic generation (SHG), (b) sum frequency
generation (SFG), (c) anti-Stokes Raman (ASR), (d) anti-Stokes Brillouin scattering (ASBS), (e) two-pho-
ton absorption (TPA), (f) electron transition upconversion through ground state absorption/excited
state absorption (GSA/ESA), (g) laser cooling (LC), and (h) optically stimulated luminescence (OSL). GS
– ground state, ES – excited state, VB – valence band, CB – conduction band.

Inelastic scattering
When light interacts with matter, photons can scatter off atoms, molecules, or structures
within a material. In the case of inelastic scattering, photons can exchange energy with
matter and, in the case of Raman scattering, the energy exchanged is equal to the
energy of specific vibrational modes of a molecule. Photons can either lose energy equal
to the vibrational mode energy (Stokes Raman scattering) or gain that same energy
(anti-Stokes Raman scattering, ASRS), as shown in Figure 1c [34]. Raman scattering has
found applications in spectroscopic analysis of materials due to specificity of the energy
exchange, as demonstrated in analysis of geological samples [35] and in explosives
detection [36].
One of the most successful applications of Raman scattering is known as coherent
anti-Stokes Raman scattering (CARS) [37]. In this phenomenon, two photons of the
same energy Epump are used to excite a molecule: the first photon raises the molecule to
4 J. E. MOFFATT ET AL.

a virtual state, from which it relaxes rapidly down to a vibrational level of the ground
state, losing an amount of energy equal to the energy of the Stokes shift EStokes; the
second photon then raises the molecule to a new virtual state, from which it relaxes
down to the ground state by emitting a new photon of energy Eem ¼ 2Epump  EStokes,
higher than each of the two pump photons [38]. This technique has been successfully
exploited for advanced functional biomedical imaging [38] and standoff detection of
explosives [39]. The selectivity of the CARS technique in identifying chemical species
has led to some fascinating applications such as performing CARS-based microscopy on
in vitro samples of rabbit aortas to establish the location and quantity of atherosclerotic
plaque buildup due to high-fat diets, as a model to early identification of cardiovascular
disease [40].
Another form of inelastic scattering is Brillouin scattering [41, 42] which occurs
when a photon interacts with an acoustic phonon. In anti-Stokes Brillouin scattering
(ASBS), the energy of photon and phonon combine to make a higher energy scattered
photon (Figure 1d). Brillouin scattering can be enhanced at high powers of excitation in
suitable materials (stimulated Brillouin scattering, SBS). The review by Bai et al. [43]
provides more detail on the theory and applications of SBS and we note that SBS
behaves much like Raman scattering apart from its non-isotropic emission. Brillouin
scattering is commonly used in distributed sensing with optical fibers [44] and charac-
terizing the elastic properties of materials [45]. The in-fiber sensing applications of SBS
have resulted in recent demonstrations of more than two million sensing points along a
17.5 km optical fiber, whereby the SBS phenomenon has been combined with careful
signal modulation and meticulous characterization of the fiber used, a great benefit for
large-scale distributed sensing [46].

Photonically induced thermal radiation


When high intensity light is incident on a sample it can cause significant local heating,
which generates thermal radiation from a material as temperatures rise [47]. In an ideal
case this emission can be represented as a black body emitter. The peak of this thermal
emission is temperature dependent and if a material gets sufficiently hot it can result in
emission that is shorter than the absorbed wavelength. STEW emission from this effect
is more readily produced when there is high laser power absorption density in a
material with low thermal conductivity, generating a local hot spot [48–51].
Photonically induced thermal radiation (PITR) does in theory hold information about
the local lattice structure around the absorbing ion, but as the wavelength output is
very broad, this information is hard to extract. PITR is studied for applications where
high anti-Stokes’ emission output is needed but the emission wavelength is not import-
ant. PITR is currently researched as a process to convert near-infrared energy for use in
solar cells because of its potential for large STEW efficiencies [48].

Direct energy transitions


Light-excited electronic transitions occur when one or more photons are absorbed by a
material, which then subsequently excites an electron, or the absence of an electron
APPLIED SPECTROSCOPY REVIEWS 5

(known as an electron hole), to a higher excited state available within the material (for
brevity, we will only refer to electrons). This higher excited state could be a higher
electron orbital within an iron, a state caused by the interaction between two or more
elements in a material’s structure, or the conduction band of a material.
For luminescence emission to be favored, the excited states (excluding the conduction
band) [52] must be relatively isolated from the host material which limits non-radiative
relaxation processes. In bulk materials, these states are usually formed via material
defects or dopant ions that are not part of the general material structure, though rare
earth ion materials can be exceptions to this. Individual molecules are in general more
isolated from each other, and so different energy states within a molecule can up con-
vert in the same way as a dopant or defect in a bulk material.

Multi-photon absorption
In multi-photon absorption (MPA), an electron is promoted to a higher energy level by
the simultaneous absorption of two or more photons [53, 54]. This process is enabled
via intermediate virtual energy levels that arise due to the temporal coexistence of mul-
tiple excitation photons. The difference between the original and final state is the sum
of the energies of the absorbed photons (Figure 1e). Once at the higher excited state,
the material relaxes through multiple pathways, resulting in light emission from transi-
tions when possible. This creates emission spectra that usually look identical to the light
emission caused by absorption of a single higher energy photon.
TPA, a common subset of MPA, is used in [54] two-photon microscopy. This has
become a widely popular technique within biology and medicine, often as an alternative
to confocal microscopy [55]. Multiphoton approaches to fluorescent imaging allow for
very small areas of the sample to be visualized at a time. Due to the high intensity
needed to drive the process, excitation only occurs either right at the point of focus or
at the intersection of two separate laser beams [3, 56]. This technology has driven the
development of organic TPA dyes [6, 57] and nanoparticles [58] that are used as tag-
gants or contrasting agents, where visible light emissions are detected after excitation in
the near-infrared.

Sequential absorption upconversion


Also known as ground state absorption  excited state absorption (GSA/ESA), the pro-
cess of sequential absorption upconversion relies on energy states an electron can dir-
ectly access through photon absorption, and is therefore almost always seen via dopant
ion energy levels, with some exceptions [52, 59, 60]. An initial photon excites an elec-
tron to a first excited state, as illustrated in Figure 1f. If a second photon arrives before
the electron decays to a lower energy state, the electron can be further excited to a
higher state. Light emission may occur either directly or through cascading relaxations
down to lower energy levels.
This process has been used to detect low levels of gases in the atmosphere such as
nitric oxide [61]. GSA/ESA is an important process in the operation of infrared lasers
6 J. E. MOFFATT ET AL.

based on rare earth ions, but STEW emission from the GSA/ESA state is usually
undesirable [62–64].

Laser cooling
Laser cooling [65] occurs when there is a population of thermally excited electrons in a
material. A photon excites an electron to a higher energy excited state, which then radi-
atively decays to a lower energy level than the initial excited state, thereby lowering the
overall energy of the material, see Figure 1g. The same process can be used in chemical
analysis [66]. One can distinguish laser cooling from other STEW processes as the fluor-
escence intensity associated with laser cooling has a linear power dependence [51, 67],
but otherwise appears as an ESA process (see Section 5.3).

Optically stimulated luminescence


Optically stimulated luminescence and infrared stimulated luminescence (IRSL) are the
only STEW processes that are commonly seen in natural materials [68, 69]. Ionizing
radiation excites ground state electrons to excited metastable states in a material. These
metastable states can have lifetimes up to millions of years, so spontaneous emission
from these states is not detected. In OSL, photon excitation removes the electron (or
hole) from the metastable state, allowing it to de-excite to its original state via photon
emission, as shown in Figure 1h [68]. In IRSL, a photon excites an electron to an inter-
mediate energy level, from which it is removed to a final excited state via phonon
assisted transfer and emits in a similar process to OSL [69]. OSL and IRSL can be iden-
tified by the way they are depleted by excitation at relatively low excitation powers if
the sample is not exposed to additional ionizing radiation.
OSL is used as a dosimetry technique when designing in situ and instantaneous radi-
ation monitoring for medical and environmental sensors [70, 71]. When the environ-
mental radiation dose rate can be calculated, optically stimulated luminescence can also
be used as a dating technique, dating from time of last optical exposure, or “time of
burial”. It is used extensively for dating archeological sites and geological formations
[68, 72, 73]. For more theory, refer to Lucovsky [74] and McKeever [75]; for more
applications, see a review by Aitken [76].

Energy transfer upconversion


Energy transfer upconversion (ET-UC) requires two or more semi-isolated energy level
sites to be in close physical and energetic proximity so they can interact via non-radia-
tive energy transfer. These sites could be multiple dopant ions or defect centers in bulk
materials, energy levels in different constituents of a complex molecule, or neighboring
molecules. Energy transfer usually occurs via dipole–dipole, or other resonant interac-
tions. Applications include improving solar energy capture [77–79] and medical imaging
[3] because this process is in general more efficient than sequential processes [80].
Upconversion temperature sensors also use energy transfer processes [81]. There are six
distinct types of ET pathways, as discussed below.
APPLIED SPECTROSCOPY REVIEWS 7

ET-UC is used extensively in upconversion nanoparticle research, often towards


applications in chemical and biological sensing. Common dopants are erbium, gadolin-
ium, and thulium codoped with ytterbium [82], with erbium/ytterbium the most com-
mon codopant pair. Ytterbium strongly absorbs at 980 nm, in a low absorbance region
of water. This excitation therefore passes through water-based suspensions and is less
likely to damage organic tissue [3, 83]. Erbium emits in the green and red regions, and
so is easily detectable by silicon photodetectors.
Upconversion nanoparticles (UCNPs) can be used to track chemical or drug delivery
in tissues and cells [7, 84, 85]. When attached to a drug or large molecule, UCNPs can
be used in the same way as markers in scanning fluorescence microscopy to observe the
movement within biological samples over time. An infrared laser is scanned across the
surface of the sample, and the fluorescent light is collected. In this way, one creates a
spatially resolved image of the fluorescence of the material. UCNPs have been tracked
through the skin of live samples [86] and, in an interesting example, even been used to
extend the vision of mice into the infrared [87]. An advantage of the UCNPs is their
high photostability, so little to no bleaching or degradation of the signal is seen during
measurement, a problem that often occurs with organic fluorescent markers [88].
Binding UCNPs to a fluorescent marker is an established chemical sensing technique
[1]. Upon excitation, the UCNPs release the right wavelength needed to cause the
bound markers to fluoresce only in the presence of the target. Interaction with the tar-
get molecule can also be designed to occlude upconversion emission, with increased tar-
get concentration corresponding to a decrease in detected STEW emission. Sensors have
also been developed in which the presence of a target molecule alters other molecules
surrounding the UCNP in order to allow STEW emission to be observed [5].

Ground state absorption–energy transfer


Ground state absorption–energy transfer (GSA/ET) involves two electrons, each from a
different energy level site, absorbing a photon each, and getting excited to higher energy
states (steps 1 and 2 in Figure 2a). Before the first electron can de-excite, the other
transfers its energy to it, raising it to a higher energy state [27] (step 3 in Figure 2a).
Auger upconversion is an example of this process [89, 90], with potential applications
in more efficient solar cells [91].

Sequential energy transfer


Sequential energy transfer, otherwise known as “addition de photons par transferts
d’energie” (APTE) [26], requires three or more energy level sites to be in close proxim-
ity. In two of these sites, an electron absorbs a photon, which raises it to a higher
energy level (steps 1 and 2 in Figure 2b). First one, and then the other excited state
transfer their energy to an electron at the third site, promoting it to a first and then
second excited state (steps 3 and 4 in Figure 2b). This is demonstrated through the
design of a Cr3þ/Er3þ/Cr3þ complex, in which 750 nm light is absorbed by the two
independent chromium ions, followed by STEW emission from the Er3þ center at
543 nm [92].
8 J. E. MOFFATT ET AL.

Figure 2. Energy level diagrams depicting two-photon examples of the different energy transfer proc-
esses: (a) ground state absorption/energy transfer (GSA/ET), (b) sequential ET, (c) donor–acceptor
energy transfer (D–A ET), and (d) simultaneous ET. GS – ground state, ES – excited state.

Donor–acceptor energy transfer


Sequential donor–acceptor (D–A) ET is a special case of APTE, which occurs when
photons are sequentially absorbed by an electron in a higher absorption cross section
site, and then the electron transfers its energy to an electron in the second site
(Figure 2c). STEW emission can occur when this process takes place twice resulting on
the electron in the second site being promoted to higher energy level. The term sensi-
tiser/activator is also used in addition to donor/acceptor.

Simultaneous energy transfer


Simultaneous ET occurs when two excited state electrons transfer energy to a stable
electron at the same time, bypassing any potential intermediate energy sites, as shown
in Figure 2d [80]. Cooperative energy transfer or cooperative energy pooling is other
name for this type of energy transfer [93] often used to refer molecular-based upconver-
sion. Simultaneous ET is seen in Yb3þ/Tb3þ and Yb3þ/Eu3þ doped glasses at wave-
lengths where the Yb3þ ion has a high absorption cross section but an energy level
structure that differs from the Tb3þ or Eu3þ ion enough to make single ion energy
transfer inefficient [94, 95].
Apart from the common ET processes described above, there are special cases of
energy transfer STEW emission that are discussed below.
APPLIED SPECTROSCOPY REVIEWS 9

Figure 3. Simplified energy level diagrams of the triplet–triplet annihilation (TTA) and photon ava-
lanche (PA) processes. GS – ground state, ES and SES – singlet excited state, TES – triplet
excited state.

Triplet–triplet annihilation
Triplet–triplet annihilation (TTA) upconversion is a special case of APTE energy
transfer, in which molecules labeled as donors and acceptors (though these can be the
same species) are present in the same solution [96] or solid material [97]. In its sim-
plest form, TTA involves two molecules being excited to their first excited singlet state
(step 1 in Figure 3a), in which electron spin states are paired, followed by that state
changing into a triplet state (in which we have two unpaired electrons, one in the
ground state, and one in the excited state, of the same spin), a process known as
intersystem crossing, shown as step 2 in Figure 3a. Triplet states have long lifetimes
as Pauli’s exclusion principle prevents two electrons in the same energy state from
having the same spin which suppresses decay to the ground state [98]. If the two mol-
ecules are sufficiently close, the triplet state energy of one of them (donor) can be
transferred to the other (acceptor) due to resonant energy transfer [97] or Dexter (col-
lision) processes [99] (step 3 in Figure 3a). The donor molecule relaxes to its ground
state, the two triplet states are annihilated, and the acceptor molecule is raised to a
higher singlet state [96] from which it can fluoresce efficiently. The process can be
efficient – currently up to 16% due to the long lifetime of the triplet states involved
[99]. Apart from this increased efficiency, TTA processes in general behave the same
as other energy transfer processes and are primarily used in increasing the efficiency
of solar cells [97, 100].

Photon avalanche
Photon avalanche (PA) is a series of upconversion mechanisms that exhibit threshold-
like behavior relative to incident excitation power [101, 102]. The phenomenon relies
on an initial population of thermally excited ions which can be excited to a higher state.
10 J. E. MOFFATT ET AL.

The STEW variant of the process relies on energy exchange with the ground state
before fluorescence – the excited ion can either give fluorescence output (step 3 in
Figure 3b) or undergo energy transfer (step 2 in Figure 3b), leading to two ions in
the initial excited state. As the population of initial excited state ions increases, the
likelihood of absorbing pump photons increases and the process runs away, leading to
a dramatic increase in observed fluorescence. The PA effect is generally found in
materials that are efficient at upconverting with multiple upconversion pathways
[103]. Its threshold behavior can be used to identify the process in power dependence
scans. PA has found its applications in novel optical laser gain media [103] such as
Er:ZBLAN [104].

Connecting theory to observations


Mathematical models such as rate equations and Judd–Olfelt theory [105, 106] are gen-
erally used to predict emission behavior when optimizing STEW materials. This
becomes increasingly difficult as materials become more complex to fit more specific
applications. Identifying STEW emission processes experimentally is therefore becoming
more important as the field broadens to encompass more varied applications. Below we
discuss how to identify STEW processes using experimental parameters usually within
the control of the experimenter: excitation wavelength, emission wavelength, material
absorption, power dependence, and lifetime.

Wavelength dependence
A common initial diagnostic procedure for understanding the origin of a STEW signal
is to vary the wavelength of the excitation source and monitor the wavelength and
intensity of the emission. A simplified flow chart diagram of the wavelength dependence
of common STEW processes is shown in Figure 4. For harmonic generation, the energy
of the emitted photons will be an integer multiple of the incident photon energy. For
wave mixing, the energy of the emitted photons will be an addition or subtraction of
the energy of the input photons. Anti-Stokes Raman displays a similar behavior, with
the energy of the emitted photon being shifted from the energy of the excitation photon
by a fixed amount that corresponds to the energy of the specific vibrations within
the material.
In contrast, STEW emission from electronic transition processes generally changes
in intensity and not in emission wavelength as the excitation wavelength is varied.
The emission spectrum can change slightly due to changes in transition probability,
but there is not a defined spectral shift as in harmonic generation or Raman. No
emission is seen at wavelengths in which the electronic transitions are not ini-
tially excited.
The wavelength of PITR depends on the local temperature which is altered by the
absorbed power density of the incident radiation. This will be affected by the absorption
intensity which is wavelength dependent. The PITR will thus shift to longer wavelengths
when the excitation is off-band to the absorption peak.
APPLIED SPECTROSCOPY REVIEWS 11

Figure 4. Flow chart diagram of the wavelength dependence of various STEW processes. kem – emis-
sion wavelength, kex – excitation wavelength, I @ kem – intensity at the emission wavelength, kab –
absorption wavelength. kGSA and kESA are wavelengths corresponding to ground state and excited
state absorptions, respectively. Note that in some organic molecules with very broad emission and
excitation, these comparisons may not be resolved within experimental bands. We show here har-
monic generation (HG), and not wave mixing in general.

Material absorption
Absorption measurements are conducted by monitoring the fraction of light transmitted
through a material at different wavelengths. Absorption measurements can identify the
origin of observed optical phenomena from a material, although absorption can be due
to a combination of factors that are not always easy to separate. The following observa-
tions can be made:

 SFG/wave mixing: These processes only occur when the material is transparent
at both the incident and harmonic wavelengths (i.e. no absorption). If the har-
monic wavelengths are heavily absorbed, then two photon-absorption is likely to
dominate instead of SFG.
 Anti-Stokes Raman: Raman signals in general do not correlate to absorption
peaks as they originate in vibrations rather than electronic transitions. We note
here however that absorption peaks can enhance Raman signals from the mater-
ial, a process known as resonant Raman [107].
 PITR: A material undergoing PITR is likely to have a large absorption profile at
the excitation wavelengths that produce it.

Upconversion and MPA in solid materials are strongest when dopants or


defects occur at higher densities but absorption from these sites may not be
observable if the densities are too low. If the absorption is observable, then the
following applies:
12 J. E. MOFFATT ET AL.

Figure 5. Flow chart diagram of the absorption dependence of various STEW processes. kem – emis-
sion wavelength, kex – excitation wavelength, kab – absorption wavelength. kGSA and kESA are wave-
lengths corresponding to ground state and excited state absorptions, respectively. Note that the
MPA case above assumes no decay transitions before emission, which is not always applicable.

 MPA: There will be an absorption peak at integer values of the photon energy of
the excitation wavelength, but not one at the wavelength itself unless the exciting
beam is very intense.
 Sequential absorption: There will be an absorption peak related to the GSA
transition, and possibly but not necessarily the final excited state before
radiative decay.
 Energy transfer: If the photon absorption before energy transfer is a GSA
(usually the case), there will be absorption peaks related to this state.

Conventional absorption measurements do not acquire ESA information because the


initial state of ESA is often empty. ESA information can be acquired through simultan-
eously absorbing wavelengths that populate the initial state via GSA in addition to the
target absorption range [108].
Absorption measurements can also be used to distinguish the main types of defects in a
material. For example, rare earths are often distinguished from transition metals by their
sharp absorption peaks that do not change appreciably between different host materials
[109, 110]. Absorbance at a particular wavelength does not mean that wavelength is
necessarily converted into luminescence, especially anti-Stokes luminescence. However,
absorption measurements can be used to provide useful diagnostic information on the
electronic transitions contributing to anti-Stokes luminescence. A flow chart diagram in
Figure 5 depicts generalized absorption dependence of some STEW processes.

Power dependence
The properties of STEW emission are generally power-dependent. Power dependence
measurements are conducted by varying the excitation power incident on the sample
and monitoring the resultant luminescence spectrum and intensity.
Most STEW processes are nonlinear in power dependence; the exception is anti-
Stokes Raman scattering, which is linear in the ranges at which it is usually measured
[30]. In SFG, the power dependence is a function of the number of photons used in the
APPLIED SPECTROSCOPY REVIEWS 13

Figure 6. Flow chart diagram of the power dependence of various STEW processes. Pem – optical
power at emission wavelength, Pex – optical power at excitation wavelength, k – absorption cross-
section at excitation wavelength. For simplicity, Pex is considered at medium levels to eliminate small
scale or saturation effects. HG and MPA will not be distinguishable from electronic transitions in
power dependence plots.

process; for example, SHG has a squared power dependence before saturation [21].
Gaining power dependence curves for PITR can be difficult, as increases in power
increase the temperature and thus the width of the radiation peak. By lowering the
power, however, one should see the peak wavelength shift until it reaches longer-than-
excitation wavelengths. Figure 6 summarizes the power dependence of several STEW
processes at medium power levels.
Electronic transition STEW processes involve real energy level transitions and hence
have a more complex power dependence. The excitation power versus emission inten-
sity behavior can yield significant information if graphed in log–log format. For upcon-
version fluorescence, the slope of the resulting line yields details on the process
involved. At low powers, the gradient of the slope is approximately equal to the number
of excitation photons absorbed in the process; at high powers, the material becomes
saturated, trending to a slope of zero [111]. See Pollnau et al. [112] for a more in-depth
discussion on the slopes at mid-range powers.

Lifetime
The lifetime of time-resolved emission data is usually defined as the time taken for the lumi-
nescence signal to decay to 1/e of its initial value [113]. Most time-dependent measurements
have three components: a “zero” level when the excitation is off, an “initial rise” as the popu-
lation of the emitting energy state begins to fill, and a “decay” when the population of the
emitting energy states empty [114], as shown in Figure 7 for the upconversion emission of
Erþ3 ions in a ZBLAN glass [63] matrix at two different excitation wavelengths.
14 J. E. MOFFATT ET AL.

Figure 7. Lifetime measurement comparison of a 5 mol % erbium doped ZBLAN glass sample excited
at 955 nm (sharp peak) and 995 nm (broad peak). The 955 nm excitation upconverts to 550 nm pri-
marily by a GSA/ESA process, and the 995 nm excitation upconverts to 550 nm primarily by a GSA/ET
process. Note the longer lifetime and larger turnover in the energy transfer lifetime measurement.

The initial rise can be used to identify the lifetime of the first excited state, and thus
can be used to distinguish between donor–acceptor and same-ion processes. GSA/ET
will have a shorter overall lifetime than simultaneous or sequential ET (APTE), as a
maximum of one ET process will occur after the excitation pulse, rather than a
maximum of two.
Some STEW emissions, such as SHG and anti-Stokes Raman are effectively instantan-
eous and only occur when the exciting source is on [21, 24]. Electronic transition-based
emission can occur long after the excitation pulse has finished [115]. The lifetime
dependence of different STEW processes is shown in Figure 8.
Energy transfer processes depend on the atomic population densities and not the
photonic population densities. Therefore, the population of the emission state can con-
tinue to fill after the initial excitation pulse has gone, creating an “initial rise” effect
which often points towards ET processes but can also be due to a decay to a STEW
upper state following MPA or GSA/ESA [111, 116, 117].
Even if the energy level structures present are not known, ESA and ET processes can
be distinguished by using a higher energy source to excite the emitting state directly. By
comparing the lifetime of this Stokes fluorescence to the lifetime of the anti-Stokes pro-
cess, one can distinguish between a direct excitation and energy transfer process.

Distinguishing the STEW pathways


Shorter-than-excitation emission pathways can be efficiently distinguished using the
experiments in the following discussion. A flow chart in Figure 9 summarizes common
pathways with which STEW processes can be more effortlessly recognized. Most STEW
pathways can be easily identified, with the difference between MPA and GSA/ESA being
the most subtle.
APPLIED SPECTROSCOPY REVIEWS 15

Figure 8. Flow chart diagram of the lifetime dependence of various STEW processes. t – lifetime of the
emission, kES – lifetime of the excited state. kES1 and kES2 denote lifetimes of the excited states in the first
and second energy level sites. Note that in the extreme case of the ET rate being much larger than the
rate of emission (i.e. the emission lifetime being comparatively very large), the lifetime will approach kES.

Figure 9. A flow diagram showing a simple way to identify each STEW emission process.

SHG vs. anti-Stokes Raman: In a power dependence scan, SHG will generally
have a squared dependence in the low power limit, while ASR will have a linear dependence.
Anti-Stokes Raman and SHG vs. electronic transitions and photon induced thermal radi-
ation: When changing the excitation wavelength, the resulting wavelength will be at an
16 J. E. MOFFATT ET AL.

energy twice that of the excitation for SHG. The energy gap between the anti-Stokes
Raman photons and the excitation photons is fixed. By contrast, electronic transitions
and PITR will have emission ranges dependent on the absorption profile of the material.
Electronic transitions vs. photon induced thermal radiation: When power incident on a
sample is varied, the wavelength of emission will remain constant, but the intensity will vary
for an electronic transition based STEW emission. For PITR, lowering the power will lower
the temperature of the sample, shifting the emission wavelength peak to a longer wavelength.
Sequential absorption upconversion and TPA vs. ET-UC: In most cases, a lifetime of
GSA/ESA or TPA will be shorter than the lifetime of ET-UC. If in doubt, attempt to
measure the lifetime of the final excited state by exciting it directly; ET luminescence
will produce a longer decay than this lifetime, while the lifetime of the GSA/ESA or
TPA emission decay will be the same.
Sequential absorption upconversion vs. TPA: There are three methods that can distin-
guish between GSA/ESA and TPA, but each depend on specialized equipment or the
material involved. Listed below, they involve:

 Absorption and excitation measurements: Absorption measurements will pick up


the GSA transition if the wavelength dependent absorption changes are signifi-
cant. If GSA/ESA is the process involved, when using one-wavelength excitation,
the highest intensity of luminescence will be at an excitation wavelength slightly
offset from the maximum GSA transition, as the ESA absorption peak will not
be at the same wavelength due to loss of energy through the vibrational levels of
the material. In general, GSA/ESA pathways are far more probable than TPA, so
if GSA/ESA can occur, it will be stronger than TPA.
 Two tunable light sources: A GSA/ESA emission will be brightest when one wave-
length is tuned to the GSA absorption peak and the other tuned to the ESA absorp-
tion peak [118]. TPA will not show this as each individual photon does not go into
a real absorption band in the material. This experiment would benefit from a low-
ered temperature, which serves to narrow the absorption bands of the transitions.
 A low-jitter controlled pulse system: This method requires one to be able to fire
two pulses of light in quick succession, within the lifetime of the first excited
state but without the two pulses overlapping. In the case of MPA emission,
resulting from two individual MPA events as the pulses do not overlap, both
excitation pulses will result in the same intensity of emission. In the GSA/ESA
case, emission intensity from the second pulse will be higher than that from the
first excitation pulse, as the first pulse has built up a GSA population from which
upconversion emission is produced through the ESA pathway.

GSA/ET vs. APTE UC and donor-acceptor ET: GSA/ET will be measured as a slightly
faster processes than APTE or donor–acceptor ET. For two-photon upconversion, when
the excitation pulse has been switched off, a maximum of one process (step 3 in
Figure 2a) occurs in the GSA/ET process, whereas a maximum of two processes (steps
3 and 4 in Figure 2b) occur in APTE. If the energy transfer is occurring between identi-
cal ions, it is likely that APTE and GSA/ET will be occurring at the same time in the
material while the excitation pulses are on.
APPLIED SPECTROSCOPY REVIEWS 17

Donor–acceptor ET vs. APTE: Donor–acceptor ET occurs when more than one type
of energy level site is present in the structure. D–A ET is often phonon-assisted due to
the difference in energies between the donor transition and the acceptor transition, and
thus will be less efficient at lower temperatures. If the emission defect origin can be
identified, D–A ET is characteristic in that the excitation peak will not match the emis-
sion site’s potential transitions.
Simultaneous ET vs. donor–acceptor ET: Simultaneous ET is a less probable process
than sequential ET, and thus only occurs in materials where sequential ET cannot occur
[80]. D–A ET has a narrower excitation spectrum than simultaneous ET [80]. D–A ET
will have an excitation spectrum in the overlap between the absorption spectrum of the
first excited state and the absorption spectrum of the second excited state. This overlap
will be smaller than the absorption spectrum of each individual state. Simultaneous ET,
however, only uses one excited state, and thus has a broader range of possible excitation
wavelengths.

Future directions
STEW materials will continue to rapidly develop, to the benefit of all fields. With promise
of new applications in important industries, such as photovoltaics, materials science, and
medicine, including proven results in medical imaging, the push for new STEW materials
appears certain to grow and bring disruptive change to science and engineering.
However, STEW materials, in particular electronic transition-based materials, are
becoming more complicated as designs for more specific purposes drive their develop-
ment. Understanding of the processes behind STEW emission is the key to successful
design and optimization as their complexity challenges purely mathematical analysis.
Despite the increasingly broad fields of application, with optical techniques and fluores-
cence detection becoming ubiquitous across science, optical analysis of novel materials
is an achievable goal for research laboratories, and essential to enable STEW materials
to continue to have an increasing impact.
This review across the diverse field of STEW emissions has categorized the broad
range of underlying phenomena and provided a guide and best practice pathways for
their identification. Creating a common language to describe, identify and categorize
the multitude of STEW processes will enable the different fields that already benefit
from STEW materials to increase their mutual understanding and propel these exciting
new families of materials into a new era of applications.

Acknowledgements
This work was performed in part at the Optofab node of the Australian National Fabrication
Facility utilizing Commonwealth and South Australia State Government funding. S. Grano,
Institute for Mineral and Energy Resources, University of Adelaide, is thanked for his support in
the establishment of this project.

Conflict of interests
The authors declare that they have no conflict of interest.
18 J. E. MOFFATT ET AL.

Authors’ contribution
All authors contributed to this work, including literature review, description of phenomena and
organizing the manuscript. EK prepared the figures.

Funding
This research was funded by CRC Optimizing Resource Extraction (ORE), and the
Commonwealth Government Department of Defense, Next Generation Technologies Fund,
Grand Challenge - Counter Improvised Threats (CIED) grants. This research is supported by an
Australian Government Research Training Program (RTP) scholarship.

ORCID
G. Tsiminis http://orcid.org/0000-0002-4321-3837
E. Klantsataya http://orcid.org/0000-0001-7029-3337
T. J. de Prinse http://orcid.org/0000-0002-4567-299X
D. Ottaway http://orcid.org/0000-0001-6794-1591

References
1. Zhang, P.; Rogelj, S.; Nguyen, K.; Wheeler, D. Design of a Highly Sensitive and Specific
Nucleotide Sensor Based on Photon Upconverting Particles. J. Am. Chem. Soc., 2006, 128,
12410–12411.
2. Sun, L.-N.; Peng, H.; Stich, M. I. J.; Achatz, D.; Wolfbeis, O. S. pH Sensor Based on
Upconverting Luminescent Lanthanide Nanorods. Chem. Commun., 2009, 5000–5002.
3. Denk, W.; Strickler, J. H.; Webb, W. W. Two-Photon Laser Scanning Fluorescence
Microscopy. Science, 1990, 248, 73–76.
4. Larson, D. R.; Zipfel, W. R.; Williams, R. M.; Clark, S. W.; Bruchez, M. P.; Wise, F. W.;
Webb, W. W. Water-Soluble Quantum Dots for Multiphoton Fluorescence Imaging in
Vivo. Science, 2003, 300, 1434–1436.
5. Wu, X.; Chen, G.; Shen, J.; Li, Z.; Zhang, Y.; Han, G. Upconversion Nanoparticles: A
Versatile Solution to Multiscale Biological Imaging. Bioconjugate Chem., 2015, 26, 166–175.
6. Wang, X.; Krebs, L. J.; Al-Nuri, M.; Pudavar, H. E.; Ghosal, S.; Liebow, C.; Nagy, A. A.;
Schally, A. V.; Prasad, P. N. A Chemically Labeled Cytotoxic Agent: Two-Photon
Fluorophore for Optical Tracking of Cellular Pathway in Chemotherapy. Proc. Nat. Acad.
Sci. USA., 1999, 96, 11081–11084.
7. Tian, G.; Gu, Z.; Zhou, L.; Yin, W.; Liu, X.; Yan, L.; Jin, S.; Ren, W.; Xing, G.; Li, S.; et al.
Mn2þ Dopant-Controlled Synthesis of NaYF4:Yb/Er Upconversion Nanoparticles for in
Vivo Imaging and Drug Delivery. Adv. Mater., 2012, 24, 1226–1231.
8. Trupke, T.; Green, M. A.; W€ urfel, P. Improving Solar Cell Efficiencies by up-Conversion
of Sub-Band-Gap Light. J. Appl. Phys., 2002, 92, 4117–4122.
9. Li, D.; Ågren, H.; Chen, G. Near Infrared Harvesting Dye-Sensitized Solar Cells Enabled
by Rare-Earth Upconversion Materials. Dalton Trans., 2018, 47, 8526–8537.
10. Qiao, Y.; Li, S.; Liu, W.; Ran, M.; Lu, H.; Yang, Y. Recent Advances of Rare-Earth Ion
Doped Luminescent Nanomaterials in Perovskite Solar Cells. Nanomaterials, 2018, 8,
1–11.
11. Yao, W.; Tian, Q.; Wu, W. Tunable Emissions of Upconversion Fluorescence for Security
Applications. Adv. Opt. Mater., 2019, 7, 1801171.
12. Johnson, L. F.; Guggenheim, H. J. Infrared-Pumped Visible Laser. Appl. Phys. Lett., 1971,
19, 44–47.
APPLIED SPECTROSCOPY REVIEWS 19

13. Fernandez-Bravo, A.; Yao, K.; Barnard, E. S.; Borys, N. J.; Levy, E. S.; Tian, B.; Tajon,
C. A.; Moretti, L.; Altoe, M. V.; Aloni, S.; et al. Continuous-Wave Upconverting
Nanoparticle Microlasers. Nature Nanotechnol., 2018, 13, 572–577.
14. Wang, X.; Liu, Y.; Zhao, P.; Guo, Z.; Li, Y.; Qu, S. Valence State Change and Defect
Centers Induced by Infrared Femtosecond Laser in Yb:YAG Crystals. J. Appl. Phys., 2015,
117, 153104.
15. Wang, G.; Qin, W.; Zhang, J.; Zhang, J.; Wangyan, C.; Wang, C.; Wei, L.; Zhu, G.; Kim,
P.; Synthesis, R.; Mechanism, G. And Tunable Upconversion Luminescence of Yb3þ/
Tm3þ-Codoped YF3 Nanobundles. J. Phys. Chem. C, 2008, 112, 12161–12167.
16. Turshatov, A.; Busko, D.; Kiseleva, N.; Grage, S. L.; Howard, I. A.; Richards, B. S. Room-
Temperature High-Efficiency Solid-State Triplet–Triplet Annihilation up-Conversion in
Amorphous Poly[Olefin Sulfone)s. ACS Appl. Mater. Interf., 2017, 9, 8280–8286.
17. Kim, J.-H.; Kim, J.-H. Encapsulated Triplet–Triplet Annihilation-Based Upconversion in
the Aqueous Phase for Sub-Band-Gap Semiconductor Photocatalysis. J. Am. Chem. Soc.,
2012, 134, 17478–17481.
18. Ren, G.; Zeng, S.; Hao, J. Tunable Multicolor Upconversion Emissions and Paramagnetic
Property of Monodispersed Bifunctional Lanthanide-Doped NaGdF4 Nanorods. J. Phys.
Chem. C, 2011, 115, 20141–20147.
19. Gao, W.; Zheng, H.; Han, Q.; He, E.; Gao, F.; Wang, R. Enhanced Red Upconversion
Luminescence by Codoping Ce3þ in b-NaY(Gd0.4)F4:Yb3þ/Ho3þ Nanocrystals. J.
Mater. Chem. C, 2014, 2, 5327–5334.
20. Lai, J.; Zhang, Y.; Pasquale, N.; Lee, K. B. An Upconversion Nanoparticle with Orthogonal
Emissions Using Dual NIR Excitations for Controlled Two-Way Photoswitching. Angew.
Chem. Int. Ed., 2014, 53, 14419–14423.
21. Kleinman, D. A. Theory of Second Harmonic Generation of Light. Phys. Rev., 1962, 128,
1761–1775.
22. Loudon, R. Theory of Non-Linear Optical Processes in Semiconductors and Insulators.
Proc. Phys. Soc., 1962, 80, 952–961.
23. Kleinman, D. A. Laser and Two-Photon Processes. Physical Review, 1962, 125, 88–89.
24. Shen, Y. R.; Bloembergen, N. Theory of Stimulated Brillouin and Raman Scattering. Phys.
Rev., 1965, 137, A1787–A1805.
25. Bloembergen, N. Solid State Infrared Quantum Counters. Phys. Rev. Lett., 1959, 2, 84–85.
26. Auzel, F. Compteur Quantique Par Transfert Denergie Entre Deux Ions De Terres Rares
Dans Un Tungstate Mixte Et Dans Un Verre. Comptes Rendus Hebdomadaires Des
Seances De L Academie Des Sciences Serie B, 1966, 262, 1016–1019.
27. Ovsyakin, V. V.; Feofilov, P. P. Cooperative Sensitization of Luminescence in Crystals
Activated with Rare Earth Ions. Jetp Letters-Ussr, 1966, 4, 317–318.
28. L’Esperance, F. A. Jr., Clinical Photocoagulation with the Frequency-Doubled Neodymium
Yttrium-Aluminum-Garnet Laser. Am. J. Ophthalmol., 1971, 71, 631–638.
29. Yaney, P. P. Reduction of Fluorescence Background in Raman Spectra by the Pulsed
Raman Technique. J. Opt. Soc. Am., 1972, 62, 1297–1303.
30. Decker, C. D. Excited State Absorption and Laser Emission from Infrared Laser Dyes
Optically Pumped at 532 nm. Appl. Phys. Lett., 1975, 27, 607–609.
31. Fix, A.; Ehret, G. Intracavity Frequency Mixing in Pulsed Optical Parametric Oscillators
for the Efficient Generation of Continuously Tunable Ultraviolet Radiation. Appl. Phys.
B-Lasers Opt., 1998, 67, 331–338.
32. B. Zorin, A.; Khabipov, M.; Dietel, J.; Dolata, R. Traveling-Wave Parametric Amplifier
Based on Three-Wave Mixing in a Josephson Metamaterial. Presented at the 16th
International Superconductive Electronics Conference, Naples, 2017, 1–3.
33. Popov, A.; Shalaev, M.; Myslivets, S.; Slabko, V. Unidirectional Amplification and Shaping
of Optical Pulses by Three-Wave Mixing with Negative Phonons. Appl. Phys. A, 2013,
115, 523–529.
34. Ferraro, J. R.; Nakamoto, K.; Brown, C. W. Introductory Raman Spectroscopy, 2nd ed.
Amsterdam: Academic Press, 2003.
20 J. E. MOFFATT ET AL.

35. Edwards, H. G.; Hutchinson, I. B.; Ingley, R.; Parnell, J.; Vitek, P.; Jehlicka, J. Raman
Spectroscopic Analysis of Geological and Biogeological Specimens of Relevance to the
ExoMars Mission. Astrobiology, 2013, 13, 543–549.
36. Tsiminis, G.; Chu, F.; Warren-Smith, S. C.; Spooner, N. A.; Monro, T. M. Identification
and Quantification of Explosives in Nanolitre Solution Volumes by Raman Spectroscopy
in Suspended Core Optical Fibers. Sensors, 2013, 13, 13163–13177.
37. Begley, R. F.; Harvey, A. B.; Byer, R. L. Coherent anti-Stokes Raman Spectroscopy. Appl.
Phys. Lett., 1974, 25, 387–390.
38. Evans, C. L.; Xie, X. S. Coherent anti-Stokes Raman Scattering Microscopy: Chemical
Imaging for Biology and Medicine. Ann. Rev. Anal. Chem., 2008, 1, 883–909.
39. Bremer, M. T.; Wrzesinski, P. J.; Butcher, N.; Lozovoy, V. V.; Dantus, M. Highly Selective
Standoff Detection and Imaging of Trace Chemicals in a Complex Background Using
Single-Beam Coherent anti-Stokes Raman Scattering. Appl. Phys. Lett., 2011, 99, 101109.
40. Matthaus, C.; Dochow, S.; Bergner, G.; Lattermann, A.; Romeike, B. F.; Marple, E. T.;
Krafft, C.; Dietzek, B.; Brehm, B. R.; Popp, J. In Vivo Characterization of Atherosclerotic
Plaque Depositions by Raman-Probe Spectroscopy and in Vitro Coherent anti-Stokes
Raman Scattering Microscopic Imaging on a Rabbit Model. Anal. Chem., 2012, 84,
7845–7851.
41. Brillouin, L. Diffusion de la Lumiere et Des Rayons X Par un Corps Transparent
Homogene. Ann. Phys., 1922, 9, 88–122.
42. Krishnan, R. S.; Chandrasekharan, V. Thermal Scattering of Light in Quartz. Nature,
1950, 165, 406.
43. Bai, Z.; Yuan, H.; Liu, Z.; Xu, P.; Gao, Q.; Williams, R. J.; Kitzler, O.; Mildren, R. P.;
Wang, Y.; Lu, Z. Stimulated Brillouin Scattering Materials, Experimental Design and
Applications: A Review. Opt. Mater., 2018, 75, 626–645.
44. Bao, X.; Chen, L. Recent Progress in Optical Fiber Sensors Based on Brillouin Scattering
at University of Ottawa. Photonic Sens., 2011, 1, 102–117.
45. Speziale, S.; Marquardt, H.; Duffy, T. S. Brillouin Scattering and Its Application in
Geosciences. Rev. Mineral. Geochem., 2014, 78, 543–603.
46. Denisov, A.; Soto, M. A.; Thevenaz, L. Going beyond 1000000 Resolved Points in a
Brillouin Distributed Fiber Sensor: Theoretical Analysis and Experimental Demonstration.
Light. Sci. Appl., 2016, 5, e16074–e16074.
47. Robinson, R. D.; Spanier, J. E.; Zhang, F.; Chan, S.-W.; Herman, I. P. Visible Thermal
Emission from Sub-Band-Gap Laser Excited Cerium Dioxide Particles. J. Appl. Phys.,
2002, 92, 1936–1941.
48. Wang, J.; Ming, T.; Jin, Z.; Wang, J.; Sun, L.-D.; Yan, C.-H. Photon Energy Upconversion
through Thermal Radiation with the Power Efficiency Reaching 16%. Nat. Commun.,
2014, 5, 5669.
49. Ye, H.; Bogdanov, V.; Liu, S.; Vajandar, S.; Osipowicz, T.; Hernandez, I.; Xiong, Q. Bright
Photon Upconversion on Composite Organic Lanthanide Molecules through Localized
Thermal Radiation. J. Phys. Chem. Lett., 2017, 8, 5695–5699.
50. Cinkaya, H.; Eryurek, G.; Di Bartolo, B. White Light Emission Based on Both
Upconversion and Thermal Processes from Nd3þ Doped Yttrium Silicate. Ceram. Int.,
2018, 44, 3541–3547.
51. Strek, W.; Tomala, R.; Lukaszewicz, M.; Cichy, B.; Gerasymchuk, Y.; Gluchowski, P.;
Marciniak, L.; Bednarkiewicz, A.; Hreniak, D. Laser Induced White Lighting of Graphene
Foam. Sci. Rep., 2017, 7, 1–9.
52. Wang, X.; Qiu, J.; Song, J.; Xu, J.; Liao, Y.; Sun, H.; Cheng, Y.; Xu, Z. Upconversion
Luminescence and Optical Power Limiting Effect Based on Two- and Three-Photon
Absorption Processes of ZnO Crystal. Opt. Commun., 2007, 280, 197–201.
53. Singh, S.; Bradley, L. T. Three-Photon Absorption in Napthalene Crystals by Laser
Excitation. Phys. Rev. Lett., 1964, 12, 612–614.
54. Chin, R. P.; Shen, Y. R.; Petrova-Koch, V. Photoluminescence from Porous Silicon by
Infrared Multiphoton Excitation. Science, 1995, 270, 776–778.
APPLIED SPECTROSCOPY REVIEWS 21

55. Lee, J. H.; Lee, S.; Gho, Y. S.; Song, I. S.; Tchah, H.; Kim, M. J.; Kim, K. H. Comparison
of Confocal Microscopy and Two-Photon Microscopy in Mouse Cornea in Vivo. Exp. Eye
Res., 2015, 132, 101–108.
56. Miller, D. R.; Jarrett, J. W.; Hassan, A. M.; Dunn, A. K. Deep Tissue Imaging with
Multiphoton Fluorescence Microscopy. Curr. Opin. Biomed. Eng., 2017, 4, 32–39.
57. Sun, C.-L.; Li, J.; Wang, X.-Z.; Shen, R.; Liu, S.; Jiang, J.-Q.; Li, T.; Song, Q.-W.; Liao, Q.;
Fu, H.-B.; et al. Rational Design of Organic Probes for Turn-On Two-Photon Excited
Fluorescence Imaging and Photodynamic Therapy. Chem, 2019, 5, 600–616.
58. Wang, H.; Huff, T. B.; Zweifel, D. A.; He, W.; Low, P. S.; Wei, A.; Cheng, J.-X. In Vitro
and in Vivo Two-Photon Luminescence Imaging of Single Gold Nanorods. Proc. Nat.
Acad. Sci. USA., 2005, 102, 15752–15756.
59. Chen, S. L.; Stehr, J.; Reddy, N. K.; Tu, C. W.; Chen, W. M.; Buyanova, I. A. Efficient
Upconversion of Photoluminescence via Two-Photon Absorption in Bulk and Nanorod
ZnO. Appl. Phys. B, 2012, 108, 919–924.
60. Johnson, E. J.; Kafalas, J.; Davies, R. W.; Dyes, W. A. Deep Center EL2 and anti-Stokes
Luminescence in Semi-Insulating GaAs. Appl. Phys. Lett., 1982, 40, 993–995.
61. Bradshaw, J.; Davis, D. D. Sequential Two-Photon-Laser-Induced Fluorescence: A New
Method for Detecting Atmospheric Trace Levels of NO. Opt. Lett., 1982, 7, 224–226.
62. Kerridge-Johns, W. R.; Damzen, M. J. Analysis of Pump Excited State Absorption and Its
Impact on Laser Efficiency. Laser Phys. Lett., 2015, 12, 1–6.
63. Henderson-Sapir, O.; Munch, J.; Ottaway, D. New Energy-Transfer Upconversion Process
in Er3þ:ZBLAN Mid-Infrared Fiber Lasers. Opt. Express., 2016, 24, 6869–6883.
64. Librantz, A. F. H.; Jackson, S. D.; Gomes, L.; Ribeiro, S. J. L.; Messaddeq, Y. Pump
Excited State Absorption in Holmium-Doped Fluoride Glass. J. Appl. Phys., 2008, 103,
023105.
65. Sheik-Bahae, M.; Epstein, R. I. Optical Refrigeration. Nature Photon., 2007, 1, 693–699.
66. Huiran, Y.; Chunmiao, H.; Xingjun, Z.; Yi, L.; Yin, Z. K.; Shujuan, L.; Qiang, Z.; Fuyou,
L.; Wei, H. Upconversion Luminescent Chemodosimeter Based on NIR Organic Dye for
Monitoring Methylmercury in Vivo. Adv. Function. Mater., 2016, 26, 1945–1953.
67. Wang, Q.; Zhang, Q.; Zhao, X.; Luo, X.; Wong, C.; Wang, J.; Wan, D.; Venkatesan, T.;
Pennycook, S.; Loh, K.; et al. Photoluminescence Upconversion by Defects in Hexagonal
Boron Nitride. Nano Lett., 2018, 18, 6898–6905.
68. Huntley, D. J.; Godfrey-Smith, D. I.; Thewalt, M. L. W. Optical Dating of Sediments.
Nature, 1985, 313, 105–107.
69. H€utt, G.; Jaek, I.; Tchonka, J. Optical Dating: K-Feldspars Optical Response Stimulation
Spectra. Quat. Sci. Rev., 1988, 7, 381–385.
70. Henniger, J.; Horlbeck, B.; H€ ubner, K.; Prokert, K. The Evaluation of CaF2:Mn-
Polyethylene Detectors with the Aid of the Optically Stimulated Luminescence (OSL).
Nucl. Instrum. Methods Phys. Res., 1982, 204, 209–212.
71. Kalnins, C. A. G.; Ebendorff-Heidepriem, H.; Spooner, N. A.; Monro, T. M. Radiation
Dosimetry Using Optically Stimulated Luminescence in Fluoride Phosphate Optical
Fibres. Opt. Mater. Express, 2012, 2, 62–70.
72. Hamm, G.; Mitchell, P.; Arnold, L. J.; Prideaux, G. J.; Questiaux, D.; Spooner, N. A.;
Levchenko, V. A.; Foley, E. C.; Worthy, T. H.; Stephenson, B.; et al. Cultural Innovation
and Megafauna Interaction in the Early Settlement of Arid Australia. Nature, 2016, 539,
280–283.
73. Wintle, A. G.; Lancastert, N.; Edwards, S. R. Infrared Stimulated Luminescence (IRSL]
Dating of late-Holocene Aeolian Sands in the Mojave Desert, California, USA. Holocene,
1994, 4, 74–78.
74. Lucovsky, G. On the Photoionization of Deep Impurity Centers in Semiconductors. Solid
State Commun., 1965, 3, 299–302.
75. McKeever, S. W. S. Thermoluminescence of Solids; Cambridge Solid State Science Series
Cambridge University Press: Cambridge, 1985.
76. Aitken, M. J. Optical Dating – A Nonspecialist Review. Quat. Sci. Rev., 1994, 13, 503–508.
22 J. E. MOFFATT ET AL.

77. Pierre, G.; François, A.; Jean-Claude, G.; Khaled, Z. Below Band-Gap IR Response of
Substrate-Free GaAs Solar Cells Using Two-Photon up-Conversion. Jpn. J. Appl. Phys.,
1996, 35, 4401–4402.
78. Chandrasekaran, S.; Ngo, Y.-L. T.; Sui, L.; Kim, E. J.; Dang, D. K.; Chung, J. S.; Hur, S. H.
Highly Enhanced Visible Light Water Splitting of CdS by Green to Blue Upconversion.
Dalton Trans., 2017, 46, 13912–13919.
79. Monguzzi, A.; Oertel, A.; Braga, D.; Riedinger, A.; Kim, D. K.; Kn€ usel, P. N.; Bianchi, A.;
Mauri, M.; Simonutti, R.; Norris, D. J.; et al. Photocatalytic Water-Splitting Enhancement
by Sub-Bandgap Photon Harvesting. ACS Appl. Mater. Interf., 2017, 9, 40180–40186.
80. Auzel, F. Upconversion Processes in Coupled Ion Systems. J. Lumin., 1990, 45, 341–345.
81. Berthou, H.; J€orgensen, C. K. Optical-Fiber Temperature Sensor Based on Upconversion-
Excited Fluorescence. Opt. Lett., 1990, 15, 1100–1102.
82. Xu, S.; Huang, S.; He, Q.; Wang, L. Upconversion Nanophosphores for Bioimaging. TrAC
Trend. Anal. Chem., 2015, 66, 72–79.
83. Costa, D. F.; Mendes, L. P.; Torchilin, V. P. The Effect of Low- and High-Penetration
Light on Localized Cancer Therapy. Adv. Drug Deliv. Rev., 2019, 138, 105–116.
84. Nam, S. H.; Bae, Y. M.; Park, Y. I.; Kim, J. H.; Kim, H. M.; Choi, J. S.; Lee, K. T.; Hyeon,
T.; Suh, Y. D. Long-Term Real-Time Tracking of Lanthanide Ion Doped Upconverting
Nanoparticles in Living Cells. Angew. Chem. Int. Ed., 2011, 50, 6093–6097.
85. Hinde, E.; Thammasiraphop, K.; Duong, H. T. T.; Yeow, J.; Karagoz, B.; Boyer, C.;
Gooding, J. J.; Gaus, K. Pair Correlation Microscopy Reveals the Role of Nanoparticle
Shape in Intracellular Transport and Site of Drug Release. Nature Nanotechnol., 2017, 12,
81–89.
86. Tian, B.; Fernandez-Bravo, A.; Najafiaghdam, H.; Torquato, N. A.; Altoe, M. V. P.;
Teitelboim, A.; Tajon, C. A.; Tian, Y.; Borys, N. J.; Barnard, E. S.; et al. Low Irradiance
Multiphoton Imaging with Alloyed Lanthanide Nanocrystals. Nat. Commun., 2018, 9,
3082.
87. Ma, Y.; Bao, J.; Zhang, Y.; Li, Z.; Zhou, X.; Wan, C.; Huang, L.; Zhao, Y.; Han, G.; Xue,
T. Mammalian near-Infrared Image Vision through Injectable and Self-Powered Retinal
Nanoantennae. Cell, 2019, 177, 243–255.e215.
88. Wu, S.; Han, G.; Milliron, D. J.; Aloni, S.; Altoe, V.; Talapin, D. V.; Cohen, B. E.; Schuck,
P. J. Non-Blinking and Photostable Upconverted Luminescence from Single Lanthanide-
Doped Nanocrystals. Proc. Nat. Acad. Sci., 2009, 106, 10917–10921.
89. Seidel, W.; Titkov, A.; Andre, J. P.; Voisin, P.; Voos, M. High-Efficiency Energy Up-
Conversion by An “Auger Fountain” at an InP-AlInAs Type-II Heterojunction. Phys. Rev.
Lett., 1994, 73, 2356–2359.
90. Driessen, F. A. J. M. High-Efficiency Energy up-Conversion at GaAs-GaInP2 Interfaces.
Appl. Phys. Lett., 1995, 67, 2813–2815.
91. Makarov, N. S.; Lin, Q.; Pietryga, J. M.; Robel, I.; Klimov, V. I. Auger up-Conversion of
Low-Intensity Infrared Light in Engineered Quantum Dots. ACS Nano., 2016, 10,
10829–10841.
92. Aboshyan-Sorgho, L.; Besnard, C.; Pattison, P.; Kittilstved, K. R.; Aebischer, A.; Bunzli,
J. C.; Hauser, A.; Piguet, C. Near-Infrared–>Visible Light Upconversion in a Molecular
Trinuclear d-f-d Complex. Angew. Chem. Int. Ed., 2011, 50, 4108–4112.
93. Weingarten, D. H.; LaCount, M. D.; van de Lagemaat, J.; Rumbles, G.; Lusk, M. T.;
Shaheen, S. E. Experimental Demonstration of Photon Upconversion via Cooperative
Energy Pooling. Nat. Commun., 2017, 8, 1–7.
94. Liang, H.; Chen, G.; Li, L.; Liu, Y.; Qin, F.; Zhang, Z. Upconversion Luminescence in
Yb3þ/Tb3þ-Codoped Monodisperse NaYF4 Nanocrystals. Opt. Commun., 2009, 282,
3028–3031.
95. Dwivedi, Y.; Thakur, S. N.; Rai, S. B. Study of Frequency Upconversion in Yb3þ/Eu3þ by
Cooperative Energy Transfer in Oxyfluoroborate Glass Matrix. Appl. Phys. B, 2007, 89,
45–51.
APPLIED SPECTROSCOPY REVIEWS 23

96. Parker, C. A.; Hatchard, C. G. Sensitised anti-Stokes Delayed Fluorescence. Proc. Chem.
Soc. Lond., 1962, 386–387.
97. Singh-Rachford, T. N.; Castellano, F. N. Photon Upconversion Based on Sensitized
Triplet–Triplet Annihilation. Coordinat. Chem. Rev., 2010, 254, 2560–2573.
98. Lower, S. K.; El-Sayed, M. A. The Triplet State and Molecular Electronic Processes in
Organic Molecules. Chem. Rev., 1966, 66, 199–241.
99. Zhao, J.; Ji, S.; Guo, H. Triplet-Triplet Annihilation Based Upconversion: From Triplet
Sensitizers and Triplet Acceptors to Upconversion Quantum Yields. RSC Adv., 2011, 1,
937–950.
100. Gray, V.; Dzebo, D.; Abrahamsson, M.; Albinsson, B.; Moth-Poulsen, K. Triplet–Triplet
Annihilation Photon-Upconversion: Towards Solar Energy Applications. Phys. Chem.
Chem. Phys., 2014, 16, 10345–10352.
101. Joubert, M. F.; Guy, S.; Jacquier, B. Model of the Photon-Avalanche Effect. Phys. Rev. B,
1993, 48, 10031–10037.
102. Chivian, J. S.; Case, W. E.; Eden, D. D. Photon Avalanche - New Phenomenon in Pr
3(þ)-Based Infrared Quantum Counters. Appl. Phys. Lett., 1979, 35, 124–125.
103. Joubert, M.-F. Photon Avalanche Upconversion in Rare Earth Laser Materials. Opt.
Mater., 1999, 11, 181–203.
104. Chen, Y. H.; Auzel, F. Photon Avalanche in Er:ZBLAN Fibre Pumped at 690 nm. Electron.
Lett., 1994, 30, 1602–1603.
105. Ofelt, G. S. Intensities of Crystal Spectra of Rare-Earth Ions. J. Chem. Phys., 1962, 37,
511–520.
106. Judd, B. R. Optical Absorption Intensities of Rare-Earth Ions. Phys. Rev., 1962, 127,
750–761.
107. Carey, P. R. Resonance Raman Spectroscopy in Biochemistry and Biology. Quart. Rev.
Biophys., 1978, 11, 309–370.
108. Bialkowski, S. E. On the Determination of Kinetic Rate and Mass Transport Coefficients
in Laser Pump-Probe Experiments. Chem. Phys. Lett., 1981, 83, 341–345.
109. Wenger, O. S.; G€ udel, H. U. Chemical Tuning of the Photon Upconversion Properties in
Ti2þ-Doped Chloride Host Lattices. Inorg. Chem., 2001, 40, 5747–5753.
110. Eichhorn, M. Quasi-Three-Level Solid-State Lasers in the Near and Mid Infrared Based
on Trivalent Rare Earth Ions. Appl. Phys. B, 2008, 93, 269–316.
111. Gamelin, D. R.; G€ udel, H. U. Spectroscopy and Dynamics of Re4þ near-IR-to-Visible
Luminescence Upconversion. Inorg. Chem., 1999, 38, 5154–5164.
112. Pollnau, M.; Gamelin, D. R.; L€ uthi, S. R.; G€
udel, H. U.; Hehlen, M. P. Power Dependence
of Upconversion Luminescence in Lanthanide and Transition-Metal-Ion Systems. Phys.
Rev. B, 2000, 61, 3337–3346.
113. Werts, M. H. V.; Jukes, R. T. F.; Verhoeven, J. W. The Emission Spectrum and the
Radiative Lifetime of Eu3þ in Luminescent Lanthanide Complexes. Phys. Chem. Chem.
Phys., 2002, 4, 1542–1548.
114. Hsu, C.-P.; Georgievskii, Y.; Marcus, R. A. Time-Dependent Fluorescence Spectra of Large
Molecules in Polar Solvents. J. Phys. Chem. A, 1998, 102, 2658–2666.
115. Murakami, Y.; Kikuchi, H.; Kawai, A. Kinetics of Photon Upconversion in Ionic Liquids:
Time-Resolved Analysis of Delayed Fluorescence. J. Phys. Chem. B, 2013, 117, 5180–5187.
116. Yin, M.; Joubert, M. F.; Krupa, J. C. Infrared to Green up-Conversion in LaCl3: U3þ. J.
Lumin., 1997, 75, 221–227.
117. Buisson, R.; Vial, J. C. Transfer inside Pairs of Pr3þ in LaF3 Studied by up-Conversion
Fluorescence. J. Phyique. Lett., 1981, 42, 115–118.
118. Gamelin, D. R.; G€ udel, H. U. Two-Photon Spectroscopy of d3 Transition Metals: Near-IR-
to-Visible Upconversion Luminescence by Re4þ and Mo3þ. J. Am. Chem. Soc., 1998, 120,
12143–12144.

You might also like