Open Access Article
Marcin
Szalkowski
ab,
Agata
Kotulska
a,
Magdalena
Dudek
a,
Zuzanna
Korczak
a,
Martyna
Majak
a,
Lukasz
Marciniak
a,
Malgorzata
Misiak
a,
Katarzyna
Prorok
a,
Artiom
Skripka
ce,
P. James
Schuck
d,
Emory M.
Chan
*e and
Artur
Bednarkiewicz
*a
aInstitute of Low Temperature and Structure Research, Polish Academy of Sciences, ul. Okolna 2, 50-422 Wroclaw, Poland. E-mail: a.bednarkiewicz@intibs.pl
bNanophotonics Group, Institute of Physics, Faculty of Physics, Astronomy and Informatics, Nicolaus Copernicus University in Toruń, 87-100 Toruń, ul. Grudziądzka 5, Poland
cDepartment of Chemistry, Oregon State University, Corvallis, Oregon 97331, USA
dDepartment of Mechanical Engineering, Columbia University, New York, NY, USA
eThe Molecular Foundry, Lawrence Berkeley National Laboratory, Berkeley, California 94720, USA. E-mail: emchan@lbl.gov
First published on 11th December 2024
Photon avalanche (PA)—where the absorption of a single photon initiates a ‘chain reaction’ of additional absorption and energy transfer events within a material—is a highly nonlinear optical process that results in upconverted light emission with an exceptionally steep dependence on the illumination intensity. Over 40 years following the first demonstration of photon avalanche emission in lanthanide-doped bulk crystals, PA emission has been achieved in nanometer-scale colloidal particles. The scaling of PA to nanomaterials has resulted in significant and rapid advances, such as luminescence imaging beyond the diffraction limit of light, optical thermometry and force sensing with (sub)micron spatial resolution, and all-optical data storage and processing. In this review, we discuss the fundamental principles underpinning PA and survey the studies leading to the development of nanoscale PA. Finally, we offer a perspective on how this knowledge can be used for the development of next-generation PA nanomaterials optimized for a broad range of applications, including mid-IR imaging, luminescence thermometry, (bio)sensing, optical data processing and nanophotonics.
A resurgence of interest in PA has been driven by the recent observation of room-temperature PA in nanoscale materials2,3 and the corresponding advances in the synthesis,2,4–6 heterostructuring,3,5,7–9 modelling,2,3,6,10–13 sensitization,3,7–9,11,14 and biofunctionalization15,16 of lanthanide-doped colloidal nanoparticles that exhibit PA. Scaling PA below 100 nm allows its extreme nonlinearity to be leveraged for applications in sub-diffraction imaging and sensing of temperature, forces, and analytes. In these applications, performance metrics such as spatial resolution and sensitivity are determined by the small size and high nonlinearity of the probe. The colloidal nature of PA nanoparticles also facilitates incorporation of these materials into biological environments and the solution-phase processing of high-density devices for optical storage, computing, and detection. A key advantage of these avalanching nanoparticles (ANPs), which are built upon decades of research into lanthanide-doped upconverting nanoparticles, is that they exhibit excellent photostability, narrowband absorption and emission, high biocompatibility, relatively low critical pump threshold, and extreme susceptibility to signals such as temperature, infrared light, and chemical environment.
This review surveys the recent developments in photon avalanching nanomaterials, with emphasis on their design, synthesis, characterization, and application. This review complements the canonical reviews on PA in bulk materials17–19 to cover the surge in reports on PA nanomaterials within the past 5 years. Initially, we introduce the general principles, mechanism and key features of PA (Section 2) and its brief history in bulk materials (Section 3). Section 4 and Section 5 survey the recent developments in nanomaterials that exhibit PA and discuss their unique optical properties. Section 6 reviews the applications enabled by these PA nanomaterials. Finally, we offer a perspective on the future opportunities and challenges for research in the field of PA nanomaterials in Section 7. By providing a deep understanding of PA and its current state-of-the-art, we hope that this review will stimulate the further development of PA nanomaterials for photonics, biomedicine, sensing, computing, and other applications that would benefit from extreme nonlinearity.
. However, besides a few recently published reports,20 the probability of this process is rather low and can only be observed in non-centrosymmetric crystals under precise phase matching conditions between the laser beam and crystallographic axes of the crystal.21 Another type of anti-Stokes emission can occur following a 2- or 3-photon absorption process (Fig. 2b), in which the excited state of the emitter is reached via the simultaneous absorption of 2 or 3 photons, respectively. While 2- or 3-photon microscopy is feasible and used to enhance the imaging depth and spatial confinement, similar to the previous processes, these mechanisms require excitation with high-power, ultrafast pulsed lasers.
Most upconversion applications utilizing lanthanide ions follow the two-photon upconversion processes shown schematically in Fig. 2d and e. The upconversion process in Fig. 2d involves the sequential absorption of two photons by the same ion, i.e., the GSA between the ground energy level and a metastable intermediate level, followed by ESA from the intermediate level to a higher emitting state. A variant of this upconversion scheme, known as ET upconversion (ETU) or APTE (addition de photon par transferts d’energie), utilizes ET to sensitize absorption with a second lanthanide ion such as Yb3+ (Fig. 2e). In the case of both ESA and ETU, it is preferable if the energy of the photon absorption transitions is equal to the energy of the excitation radiation. However, these phenomena may still be observed with a limited amount of energy mismatch with the assistance of phonons, albeit with lower transition probabilities and efficiencies. Thus, as a general rule, resonant photon absorption is desirable for most upconversion applications. The other mechanisms include the much weaker cooperative sensitization (CS) and cooperative luminescence (CL). By exploiting core–shell heterostructured nanomaterials, where the core and the shell are individually optimized and doped with different Ln3+ ions, the combination of EM and ETU has opened new possibilities, such as UC under the more biocompatible 800 nm excitation and multicolour emission.23–26
Stemming from the basic ET processes found in Ln3+ ions (Fig. 2c) and conventional upconversion mechanisms (Fig. 2d–h) in homogenously and heterogeneously doped nanomaterials, various PA mechanisms can be derived, such as the conventional single-dopant PA (Fig. 2i – PA) and sensitized PA (Fig. 2j – SPA), which are the underlying mechanisms for interfacial ET-based (Fig. 2k – IFPA) and energy migration-based (Fig. 2l – EMPA) photon avalanche emission. The major difference between the conventional non-linear (SHG/THG, 2/3PhAbs) or upconversion processes (i.e., ETU, ESA, CS, and CL) and photon avalanche (i.e., PA, SPA, IFPA, and EMPA) is that PA exploits a combination of inefficient GSA with efficient ESA and purposefully augmented CR processes, whose coexistence enables unique PA features to be achieved, but requires a paradigm change in materials design.
![]() | (2-1) |
000 is typically known as energy looping rather than PA. The primary distinction between energy looping and PA is that PA exhibits a much more nonlinear optical response at the excitation power threshold at which the populations and luminescence suddenly increase. Furthermore, although no clear definitions of avalanche and looping processes exist, based on the present literature, S ∼ 10 seems to be a commonly observed non-linearity that separates the looping (3 < S < 10) regime from the avalanching (S ≥ 10) regime. A β parameter larger than 10
000 is commonly assumed to be required, but not a sufficient condition to achieve PA. The second condition is an efficient CR process to multiply the intermediate manifold population, which then enhances the efficiency of ESA. Due to the increased concentration of doping ions, the cross relaxation process between neighbouring ions is competitive with the luminescence in the PA regime, ultimately leading to the avalanche of photons, i.e., the release of energy accumulated in the system during the ESA+CR looping. Above the PA threshold (ITH), rapid luminescence intensity (IL) growth is observed in response to a minute increase in the pump intensity (IP). This pump power-dependent relationship is described by a simple power law,27 as follows:| IL ∼ (IP/ITH)S | (2-2) |
(i) Although measuring ESA directly is challenging, the ESA/GSA ratio should exceed β > 104. Given that the ESA and GSA are wavelength dependent, the ability to tune the laser wavelength to maximize the β value is critical and a non-optimal β may be prohibitive to get PA emission or reduce its PA character.
(ii) There are no rigid rules about nonlinearity and PA gain, but the values of PA non-linearity of S ≥ 10 and PA gain of ΔAV ≥ 100 (eqn (2-7)), respectively, are typically considered to indicate PA. Materials that do not meet these criteria but exhibit the same mechanism are considered to be undergoing energy looping. Measuring these values experimentally in a reliable and operator-to-operator variation free manner is often difficult, and thus some hints are presented in Section 2.7. Also, some computer algorithms have been developed to automatically derive S, ITH and ΔAV.6
(iii) A significant slowing down of the rise times at the PA threshold, and their shortening as excitation intensities are increased above the threshold.
![]() | (2-3) |
![]() | (2-4) |
![]() | (2-5) |
and IGSA, σGSA, denote the pump intensity and the absorption cross section under GSA, while IESA, σESA denote the pump intensity and absorption cross section under ESA; s31,WNR denote the CR rate between level 3 and 1, bai denotes the branching ratio between states
and Aa is the radiative rate from level a (a = 2 and 3), respectively. In the steady state,
occurs (for all the levels), which enables the derivation of the n3 population the luminescence intensity (A3n3) as functions of the excitation power density. In the case of PA, analytically solving the set of rate equations may be too complex or even impossible, and thus numerical methods are used.
Due to the inherent nature of PA emission, pump power-dependent luminescence rise times of up to hundreds of milliseconds (τR ∼ 100–1000 ms) have been observed (Fig. 4). Under sufficiently long excitation pulses (t > τR), these kinetic profiles of the PA(t) luminescence intensities enable the extraction of the steady-state luminescence intensity (ISS) and rise time of the luminescence intensity. The inverse of the experimental luminescence decay (τexp) equals the sum of all radiative (kR) and nonradiative (kNR) rates, as follows:
![]() | (2-6) |
Differential rate equation (DRE) modelling can be useful to both quantitatively and qualitatively understand the particular PA mechanism (e.g. the role and mechanism of Yb3+ sensitization of Pr3+ avalanching),3,11 and likewise understand some trends in the PA thresholds, slopes or emission quantum yields that occur in response to a variation in the critical PA parameters (e.g. the rates of CR, radiative, non-radiative or absorption cross-section).2,6,11 Moreover, DRE's can be equipped with terms corresponding to additional physical phenomena such as additional ET and relaxations, and has been proven to be an important tool, beside experimental evidence, helping to understand way PA becomes modified in response to a variation in temperature10 or the underlying quenching mechanism affecting the looping and emitting levels.12
![]() | ||
| Fig. 5 Characteristic properties of the PA phenomenon: excitation power-dependent (a) steady-state – IL(IP), (b) nonlinearities (S), (c) risetime – t0.5(IP) and (d) PA gain. The dotted line on (a) represents the experimental values for 8%Tm3+-doped NaYF4 avalanching NPs,2 while the blue lines are generated with the DRE model for an increasing Tm3+ concentration developed in ref. 2 (b)–(d) data have been derived from (a) using algorithms from ref. 6 orange dashed lines show correspondence between the (a)–(d) plots and account for the light blue data line (15%Tm3+). | ||
The rise times (Fig. 4 and 5c) are initially in the micro-second range (very short), then significantly prolong to tens or hundreds of milliseconds (close to the PA threshold), and finally shorten again due to the saturation process with an increase in the pumping intensity.2,29
By plotting the pump intensity (IP)-dependent steady-state PA emission (ISS), one may derive further parameters that characterize the PA emission, such as the pump intensity at which PA occurs (PA threshold, ITH) or saturates (ISAT) and the slope (S) of the power dependence curve (Fig. 5a and b). One may also calculate a pump power-dependent PA gain, which is defined by Lee et al.,2 as follows:
![]() | (2-7) |
One of the most distinctive and unintuitive features of PA is that near the PA pump power threshold, the PA rise time t0.5 gets significantly longer, as long as milliseconds or even seconds. These long rise times are useful both for confirming the presence of PA, and also as critical parameters for evaluating PA materials. These long rise times can be disadvantageous, e.g., typically being viewed as detrimental for achieving fast frame rates during imaging.3 Alternatively, these long rise times allows one to extract ‘pure avalanche photons’ in the time domain,29i.e., the photons that have their origin in the photon avalanche phenomenon and not from the linear absorption and emission (IP ≤ ITH) or from saturated emission (ISAT ≤ IP). These “pure” avalanche photons, IPA, may be quantified by the IPA = I(t2) − I(t1) difference (t1 < t2), where t1 and t2 (0 < t1 < t2) are arbitrarily selected to discriminate long rise times photons, corresponding to PA, from the faster ones, which can be ascribed to the conventional luminescence.
More precisely, t1 denotes the time point when steady-state emission is obtained using an excitation intensity close to the saturation region, significantly exceeding the PA pumping threshold. Alternatively, t2 is the shortest time required to observe steady-state emission from the system excited with the power close to the PA threshold. The “pure” PA photons is a term that may be important for applications where only most nonlinear behaviour is critical, such as in photon avalanche single-beam super-resolution imaging (PASSI).29
However, certain domains of ions in the emitting level are surrounded by neighbours also excited to the emitting level, making the emission the only possible way to relax the energy. Importantly, after the emission of photons, this ground state ion may be immediately promoted to the intermediate level by the CR with one of the neighbours in the emitting energy level, and then both of them can turn back to the emitting level after capturing the ESA-resonant excitation light. However, a further increase in the illumination power density gradually limits the number of the ground-state ions in the host material, limiting further CRs (Fig. 6d). The dominant process is now the release of accumulated energy in the form of photons (Fig. 6, bottom row). In the power dependence measurements, a linear dependence between the excitation and emission intensities is observed again at this stage (Fig. 6a) because a balance among ESA, CR, kMPR and luminescence processes is reached. Based on the DRE modelling of the PA Tm3+ emission, Fig. 6b–d show the importance of various processes (shown schematically by arrow thickness in energy levels (EL) insets), the average population (AP), as well as the population kinetics (PK) of the ground (n1), intermediate (n2) and emitting (n3) levels, and the contribution of the CR process kinetics (CRK) to the whole PA process. A mechanistic and time-lap like explanation for PA process is schematically presented in Fig. 7.
Fig. 8 semi-quantitatively demonstrates the impact of the radiative and non-radiative rates of the 2nd and 3rd levels in the simplified scheme of an exemplary Tm3+ PA material on its power dependence characteristics. It presents the significantly different impact of the dynamics of both levels on the PA performance and its main parameters. In particular, the increasing radiative and non-radiative rates of the 2nd level (the intermediate level during PA) clearly shift the power dependence curve (and the values of the key parameters ITH and ISAT) toward higher powers (Fig. 8a and c).6 A much smaller trend can be observed in the case of the corresponding rates of the 3rd level (the emitting level), as shown in Fig. 8b and d.
![]() | ||
| Fig. 8 Simulations of pump power-dependent PA emission intensity of NaYF4:8%Tm3+: materials in response to the radiative rates of level 2 (a) and 3 (b) (WR2 and WR3, correspondingly), as well as the impact of non-radiative rates for levels 2 (c) and 3 (d) (WNR2 and WNR3 rates, respectively). Simulations include the pristine rates found for the original NaYF4: 8%Tm3+ material (experimental data are shown as black dots), while varying respective parameters around original values.6 | ||
The discovery of PA stimulated researchers to study different materials from this unconventional point of view, inducing further interest and development of infrared quantum counter detectors1,31 and becoming an impulse for the advancement of upconversion lasers.32–36 In addition to Pr3+-doped LaCl3, PA was obtained with other quantum counters such as LaBr3:Sm3+,37 and CeCl3:Nd3+.38 Other early observations of the PA phenomenon were also reported for bulk materials such as LiYF4:Nd3+,19,32,39 YAG:Tm3+,40 ZrF4–BaF2–LaF3–AlF3–NaF (ZBLAN):Er3+,41 YAG:Ho3+ (ref. 42) and ZBLAN co-doped with Yb3+ and Pr3+.43 PA was also investigated in optical fiber materials, especially in ZBLAN cores doped with Tm3+ (ref. 44) as well as with Er3+ (ref. 45) or co-doped with Yb3+ and Pr3+.46 ZBLAN glasses and YAlO3 doped with Ho3+ were also studied as PA materials.42,47 PA has been reported for YAG waveguides doped with Tm3+ (ref. 17 and 48) and BIGaZYTZr glass doped with Tm3+,49 and for many other materials doped with thulium ions, such as YAlO3,50 Cs2NaGdCl6,51 LaF3,52 Y2O3,53 Y2SiO5,54 KYF4,55 and CdF2.56 Moreover, LiYF4 was identified as a favourable material for hosting PA due to its low phonon energy. Doping this host with Tm3+,57 as well as with Nd3+,32 Er3+ (ref. 58) or Ho3+ (ref. 42 and 59) has been shown to provide suitable conditions to obtain PA. Co-doping materials with a second ion such as Yb3+ was beneficial to obtain the PA phenomenon. In these instances, the Yb3+ ions act as a sensitizer,46 enabling the pumping of the emitting states of the activators via ETU, and take a part in energy cross-relaxation processes, which is known to be crucial for the energy looping and PA mechanism.
An overview of the early work on PA, which was practically limited to bulk crystals, glasses, fibers and ceramics, can be found in a few early review articles.17–19,48 However, despite these successful demonstrations, PA remained a scientific curiosity rather than a mainstream research area, as can be concluded based on several publications from a given year reporting investigations of PA in hosts doped with various Ln3+ ions (Fig. 9). After the first decade since the groundbreaking experiments revealing this new phenomenon, only a few new reports were published. In the next few decades, due to trials on utilizing PA mechanisms to develop new possibilities in the field of laser materials, broadened studies of this subject were performed in several research groups, resulting in over ten new papers per year and leading to a much better understanding of the details of this phenomenon, mainly based on Tm3+-, Er3+- or Nd3+-doped host materials. Nevertheless, the rather specific conditions required to observe PA, often limited to low temperatures, even in bulk materials, have limited the development of this topic, especially compared to the number of publications related to the more general issue of upconversion in lanthanides.
Changing the host lattice and doping ion concentration allows the design of nanoparticles with predetermined properties. Many types of materials have been studied as hosts for the upconversion process between lanthanide ions including phosphates,60 vanadates,61,62 sulfides,63 borates,64–66 oxides67 and fluorides. Among the available types of materials, fluorides, ALnFx (A = alkali metal, Ln = lanthanide), such as NaLnF4 and KLnF4, and LnF3, CaF2, and KMnF3 are considered as ideal host candidates. They are characterized by low phonon energies, high chemical stability and good optical transparency over a wide wavelength range; therefore, they are often used as the host materials for upconversion. Additionally, the host lattice based on Na+, Ca2+, and Y3+ cations with an ionic radius close to the lanthanide dopant ions prevents the formation of crystal defects. Among the fluorides, hexagonal phase sodium yttrium fluoride (β-NaYF4) is regarded as the most efficient host materials for upconversion due to the very low characteristic phonon energy for this crystal lattice (350 cm−1).68 The crystal structure of the hexagonal phase NaYF4 has been considered beneficial for upconversion efficiency because two types of relatively low-symmetry cation sites are occupied by Na+ and RE3+ ions.69 Numerous studies have shown that the hexagonal NaYF4 is a much better host lattice for the upconversion emission than its cubic counterpart. Compared to cubic NaYF4:Yb3+,Er3+, the green emission in hexagonal-phase NaYF4:Yb3+,Er3+ is approximately 10 times more efficient. The CaF2 host lattice shows high thermal and chemical stability, wide transmission range and low phonon energy (328 cm−1). However, the introduction of trivalent lanthanide ions in the divalent alkaline earth fluoride MF2 (M = Ca, Sr, Ba), in which the Ln3+ ion substitutes the M2+ ion, potentially leads to the formation of crystal defects and lattice stress due to the difference in charge between the ions.70 Only lanthanum fluoride (LaF3) has intense low-energy phonons (Fig. 11), which are advantageous for reducing the multiphonon quenching of the upconversion emission. These modes are located at 227 and 390 cm−1.71 Other heavy halides (e.g., chlorides, iodides and bromides) exhibit phonon energies less than 300 cm−1, but because of their low chemical stability and hygroscopicity, their potential for application is somewhat limited. The next group of materials that is used as upconversion lattices are oxides (such as Y2O3, Gd2O3, and Lu2O3). Oxide particles are stable in a wide temperature range; moreover, they have good chemical stability and can be doped with a wide variety of lanthanide ions due to the relatively small difference in the ionic radius of the dopant and the same charge as the RE host cations. However, their phonon energy is relatively high (larger than 500 cm−1) due to the stretching vibration of the host lattice. The emission intensity of α-NaYF4 co-doped with Er3+ and Yb3+ was 20 times higher in comparison to Y2O3 co-doped with the same ions.
The choice of low-phonon energy host lattice is crucial for obtaining efficient upconversion or PA. In the high-energy phonon hosts, the luminescence lifetimes, and thus the populations of the intermediate states are reduced by multiphonon relaxation, resulting in low efficiency for upconversion processes. Recently, spectacular PA results have been achieved with ultra phonon energy nanoparticles (KPb2Cl5 and KPb2Br54) and a more conventional tetrafluoride host (NaLuF428), which clearly indicate that low cut-off energy phonons are important, but they are not the only parameters determining the possibility to get highly non-linear PA emission.
The Stark levels belonging to one J manifold of a given lanthanide ion (4fN: 2S+1LJ) in a given host material are usually split by no more than the available maximum host phonon energy. Because the electron–phonon interaction for the 4f optically active electrons (being screened by electrons in the filled 5s and 5p orbitals) in trivalent lanthanides is low, the relaxation rates of the transitions within a single J manifold are usually much faster than the inter-manifold (J ⇝ J', with energy gap ΔEJJ') relaxation rates. A population quasi-equilibrium is established within a single manifold, while inter-manifold transitions are realized by either photon emission (Ar) or sum of all non-radiative transitions (Wnr) accompanied by the generation of matrix phonons.
![]() | (3-1) |
Based on the Miyakawa–Dexter theory,74 the MPR rates WJJ′(ΔEJJ′) were phenomenologically found to exhibit exponential dependence on the energy gap ΔEJJ′ between the J and J′ manifolds:
| WJJ′(ΔEJJ′) = WJJ′(0)·exp(−α·ΔEJJ′) | (3-2) |
A similar relationship describes PAET processes (e.g. CR, ET between two lanthanide ions, etc.), as follows:
| WPAT(ΔEJJ′) = WPAT(0)·exp(−β·ΔEJJ′) | (3-3) |
Based on theory,10 the multi-phonon-assisted ET (either relaxation or inter ionic ET) can be expressed as follows:
![]() | (3-4) |
Research in PA utilization is also limited by several practical considerations, such as the need to (i) precisely control the excitation wavelength to ensure resonance or lack of resonance with narrow excitation bands, (ii) pump strongly and continuously or (iii) utilize cryogenic temperatures to diminish MPR. Finally, experimentally confirming the presence of PA is challenging; although there are rigorous criteria for avalanching (Section 2.5), these criteria are debated by some in the community, and ambiguous to others. Consequently, numerous scientific reports on PA luminescence do not definitively confirm all the PA features, and many could be ascribed to pre-PA cases (i.e. energy looping)77 or explained as a combination of ESA and ETU.17 All these facts have contributed to PA remaining a scientific curiosity with niche applications for over 40 years since its discovery.
Compared to conventional Ln3+-doped materials, the PA phenomenon requires a radical change in the paradigms in materials design. For example, in Ln3+-doped bulk materials photoexcited through GSA (Fig. 10a), the effective concentration of doping ions has been typically limited to below 1% due to the belief that higher concentrations result in “concentration quenching” of the luminescence due to CR or energy migration to defects.78 Alternatively, in PA luminophores, the absorption occurs from the first excited level, which originally is almost empty, and thus the sample is initially transparent at the pump wavelength. The population of this level and the resulting absorption coefficient may be increased, but the CR process must be engaged to multiply (double) the number of ions in this state (Fig. 10b). This is achieved by increasing the interaction between neighbouring ions by increasing their concentration. Because the luminescence intensity is proportional to the concentration of the activator (in some range), these two facts make the concentration effects a concentration enhancement rather than concentration quenching. However, it should be noted that by further increasing the dopant concentration, the quenching effects start to outweigh the concentration enhancement. Simultaneously, great care must be devoted to optimizing the selection of the hosts and dopants to reduce the non-radiative depopulation of the starting level and stream the pumping energy to enhance the population of the first excited level or the luminescence. The former is achieved through the selection of the appropriate (typically low) phonon host materials, optimization of the temperature, elimination of the sensitizer ions used typically for upconversion, and selection of the lanthanide activators with a supportive energy level scheme that promotes efficient ESA and CR and reduces non-radiative de-excitation.
:
V) ratio is high. The large SA
:
V ratio of nanoscale materials means that a large portion of these materials is exposed to surface quenching from the environment and defects. The high dopant concentrations required for PA can foster rapid energy migration to the surface of nanomaterials, which is more prone to defects and has a higher concentration of quenching species such as ligands and solvent molecules. Colloidal nanoparticles must be stabilized by ligands to maintain the advantages of colloidal dispersion (discussed in Section 4.4). To date, only a few theoretical6,79 and experimental12,80 studies have discussed the impact of surface quenching on PA. Thus, the lack of understanding how to mitigate quenching in nanoscale PA materials has limited their discovery.
Another challenge in achieving PA in nanoscale materials has been the need to develop reliable methods for synthesizing colloidal nanoparticles with robust and reproducible properties. Fortunately, the field of lanthanide-doped nanoparticles has been maturing since 2004, when Haase et al.68 and soon after Gue et al.,81 Yan et al.82 and Capobianco et al.83 managed to develop low-temperature methods for the synthesis of monodisperse NaYF4 nanoparticles. In the 20 years since these seminal reports, researchers have developed robust methods for synthesizing high-quality lanthanide upconverting nanoparticles, e.g., using thermal decomposition.83 These methods allow control of the crystal size, shape, and crystal phases, e.g., low-phonon-energy matrices such as β-NaYF4,84,85 which is well-known to host higher upconversion efficiencies than its α-NaYF4 analogue. Additional upconversion enhancement is realized through the synthesis24,25,85–87 of core–shell nanoparticle heterostructures.88–91 Undoped shells passivate the doped cores by preventing deleterious ET to surfaces, while the doped “active” shells can be used to sensitize absorption,25 segregate incompatible dopants, and mediate ET between domains,23 thereby allowing fine control of the complex ET networks among lanthanide ions. The ability to grow controlled shells has enabled new doping compositions and paradigms, particularly highly doped and alloyed nanoparticles.2,77,92–95 In addition to methods for minimizing surface quenching,93,96–98 biofunctionalization protocols have been developed. All these advancements contributed significantly to the current state of research into lanthanide-doped nanoparticles. In principle, ANPs offer similar advantages together with higher upconversion efficiencies and potential use in applications that exploit the extremely nonlinear response, although several challenges remain. Below, we discuss important considerations when adapting PA for nanoscale materials.
:
V. Moreover, the distribution of luminescent ions is non-homogenous within the volume of NPs, which may impact the upconversion rates.99 A significant amount of active ions in nanosized materials is located close to the surface,100 and thus their excited states are susceptible to quenching by crystal structure defects, the local chemical environment of the active ions, ligands or the solvent molecules.101 For example, for spherical particles with a diameter of 3 nm, nearly 50% of their atoms are present on the surface. In the case of 10 nm particles, this decreases to 20%, while for a particle with a diameter of 1 μm, only ∼0.15% of its atoms are found on its surface. The increasing amount of superficial lanthanide ions makes them susceptible to non-radiative relaxation of both emitting and intermediate levels, occurring in the course of interaction with solvent and ligands molecules. The higher SA
:
V also increases the relative contribution of surface defects, which cause the additional luminescence quenching. As a consequence, the quantum yield for NaYF4:2%Er3+, 20%Yb3+ nanoparticles with a size in the range of 10 to 100 nm is in the range of 0.005–0.3%, while for bulk materials, this parameter can be considerably higher (3%).100 Direct evidence of the impact of SA
:
V and its possibly thermal effects on PA was shown by Deng, et al.102 In their work, isolated nanoparticles did not exhibit PA, but clear avalanching phenomena emerged when microscale aggregates of their nanocrystals were formed, which may be a sign of thermal issues. Approaches to circumvent surface quenching have been extensively studied for conventional UCNPs, and thus some fundamental knowledge exists. There are several methods to reduce the detrimental role of MPR and augment the emission quantum yields of up-converting nanoparticles. These methods include optimization of the dopant and co-dopant concentrations,103,104 intentional altering of the local chemical and structural environment by adding passive co-dopants or changing the ‘structural’ cations, and managing the distribution of active ions in the host materials by the core–shell approach.
![]() | ||
| Fig. 11 Phonon properties of the crystalline host and ligands used to stabilize colloidal nanoparticles. Band absorption of vibration modes from solvents and surface ligands. The maximum vibrational phonon energies of commonly used inorganic lattice, based on digitized data for CaF2,124 NaYF4,125 Y2O3,126 YVO4,127 LaPO4,128 YAG,129 LaF3,129 LaCl3,130 KPb2Br5 and KPb2Cl5.4 | ||
This ESI† focuses on papers that either hypothesize or directly show photon avalanche behaviour in nanoparticles. There are many reports ascribing luminescence behaviour to PA (given that slopes are higher than 4 and simple ESA/ETU is not sufficient to explain the UC process), but most claims are unsupported and are not formally considered PA because they do not meet the strict criteria outlined in Section 2.5, i.e. (i) β > 104 and observation of (ii) clear threshold behavior, and (iii) elongation of rise times near the threshold. The term, “photon avalanche-like behaviour,” can be also found in the literature, which typically indicates a situation in which only part of the expected PA features is observed and the authors suspect other mechanisms than pure PA (e.g. thermal effects, multiple CR and ESA) play a role. Owing to the set of specific requirements necessary to observe PA, the efficiency of PA varies between different dopant types, host types and compositions, as well as the presence of external quenchers or temperature. These considerations, supplemented with a collection of the most up-to-date literature on PA emission in nanomaterials are presented in Table S10 (ESI†).
![]() | ||
Fig. 12 Overview graph of PA processes that have been observed or can be predicted in singly (a) and co-doped (b) lanthanide ions. The detailed studies on PA in Pr3+,135 Nd3+,29 Sm3+,136 Ho3+,137 Er3+,58 and Tm3+ 2 ions and ion-combinations for Yb3+–Pr3+,135,138 Yb3+–Ho3+,139 Yb3+–Er3+ and Yb3+–Tm3+ 140 are provided in more detail in the summary tables (available in ESI†). The (1), (2), (3), etc. symbols indicate the sequence of ET steps in the Yb co-doped photon avalanching pairs. The line named ‘GSA’ indicates the energy mismatch between the photoexcitation PA photons in relation to the energy levels. | ||
In these systems, the excitation wavelength of the sensitizer remains off-resonance with the wavelengths used to excite the emitter ions through resonant ESA. Owing to the high absorption cross-section and limited concentration quenching of Yb3+ ions, these ions accept energy from the looping emitter ions and (i) store it in their long-lived 2F5/2 levels or (ii) spread this energy efficiently in the host. Then, these excited Yb3+ ions are capable, as in conventional upconversion, of sequentially donating this energy back to neighbouring emitting ions, and thus build the population of intermediate excited states. This is in contrast to the conventional upconversion, in which Yb3+ ions are excited directly with 940–980 nm light. Thus, these sensitized PA schemes may be envisioned to occur (similar to conventional UC scheme) between Yb3+ sensitizers and 3H4 → 1G4 transitions in Pr3+; 5I8 → 5I5, 5I7 → 5F5 and 5I6 → 5S2 transitions in Ho3+; 4I15/2 → 4I11/2 and 4I11/2 → 4F7/2 transitions in Er3+ and 3F4 → 3F3, 3H4 → 1G4 and 1G4 → 1D2 transitions in Tm3+ ions. The preliminary results135,141 for the sensitized approach support the feasibility of achieving PA emission at a much wider selection of emission lines. For example, under 854 nm, Yb3+,Pr3+ PA behaviour was observed, showing high pump power dependence slopes and pump power-dependent rise times. The 854 nm wavelength is resonant with 1G4 → 3P0,1,2,1I6 ESA in Pr, and the two CR processes between the Pr–Pr pair (Pr: 3P0,1,2, 1I6; Pr: 3H4) → (Pr: 1G4, Pr: 1G4) and between Pr and Yb (Pr: 3P0,1,2, 1I6; Yb: 2F7/2) → (Pr3+: 1G4; Yb: 2F5/2) drive energy looping (Fig. 12b). Ultimately, Yb can enhance the Pr: 1G4 population through the Yb: 2F5/2 → Pr: 1G4 transition. Interestingly, the described Yb3+-sensitized Pr3+ PA emission was observed in the LiYF4 single-crystal matrix, but not in the YAlO3 crystal host. This is explained by the fact that the spectral overlap between ESA in Pr3+ and GSA in Yb3+ is non-negligible, and higher ET rates between Yb3+ and Pr3+ and within the Pr3+ ions were present as well as slower decays found for YAlO3 (i.e. 50/11/2000 μs for YLF and 12/4/650 ms for YAlO3 for the Pr3+: 3P0/Pr: 1G4 and Yb3+: 2F5/2). These facts show the sensitivity of PA to nuances related to the absorption cross section for ESA and GSA, the metastable character of the intermediate levels, and the host phonon energies. Similar PA schemes may be derived for other lanthanide pairs, in which Yb sensitizes photon avalanching in Er3+, Ho3+ or Tm3+.
Two of these schemes have been proposed recently for UCNPs, exploiting resonant energy migration through either Yb3+ or Gd3+ ions in UCNPs. Both schemes have been demonstrated experimentally for avalanching nanoparticles.3,5 In 2022, Liang et al.37 first reported this “migrating” PA (Fig. 13) with slopes as high as 28 in a four-layer nanoparticle heterostructure consisting of an NaYF4:Yb3+/Ho3+ (3/4%) active shell sandwiched between a core and 2nd shell doped with Yb3+/Pr3+ (15/0.5%), with an undoped outer shell of NaLuF4. When Ho3+ was replaced with Tm3+ in these 4-layer heterostructures, high slopes of 46 were observed in the 452 nm emission of Tm3+. In these heterostructures, the 852 nm excitation wavelength is resonant with the ESA of the Pr3+ ions, and is simultaneously off-resonance with Yb3+ and with GSA of Pr3+. The 852 nm pump is not absorbed by the Ho3+ or Tm3+-doped active layers; instead, PA occurs in the Yb3+/Pr3+-co-doped layers. The PA-enhanced population is transferred via resonant energy migration through Yb3+ ions to the Yb3+/Ho3+- or Yb3+/Tm3+-doped active layers. Notably, the emission from all of the doping ions, Pr3+, Tm3+ and Ho3+, exhibited features of avalanche emission at many lines in the visible spectrum (e.g. Tm3+: 452 nm; Ho3+: 541, 646 nm; and Pr3+: 484, 609 nm), providing avenues for multicolour PA emission at visible wavelengths and PASSI microscopy.
![]() | ||
| Fig. 13 Basic principles of migrating PA emission. The Yb–Pr pair undergoes sensitized PA mechanism, which makes the Yb3+ intermediate migrating ions transfer this behaviour to other lanthanide ions that they sensitize (e.g. Tm3+ and Ho3+).3,142 Some emission of emitting ions, which stems from the Yb–Pr3+ PA workhorse, spectrally overlap with the Pr3+ emission bands. Adapted with permission from ref. 142 Copyright 2024, SPIE and Chinese Laser Press (https://creativecommons.org/licenses/by/4.0/). | ||
One disadvantage of using the Yb3+/Pr3+ avalanching pair is the overlap of multiple Pr3+ emission lines with the multiple visible emission of other intermediate lanthanide dopants (Fig. 13), such as Tm3+, Ho3+, Tb3+ and Eu3+.3,9,142 This may pose some technical challenges in distinguishing these spectral fingerprints from various PA labels in complex biological samples. As an alternative to migration through the Yb3+ sublattice, Skripka et al.5 used migration through an intermediate Gd3+-doped 1st shell to transfer the excited-state populations generated by PA in the NaGdF4:Tm3+-doped core to Eu3+, Tb3+, Er3+, and Ho3+ emitters in a 2nd shell of NaGdF4, which was overcoated with a passivating NaYF4 shell (Fig. 14). Exciting the Tm3+ avalanching dopants at 1064 nm resulted in visible emission with nonlinearities of 14.6, 17.2, 10.7, and 11.5, respectively. Notably, the visible emission lines for these emitting dopants do not overlap with that of the Tm3+ avalanching dopants. Furthermore, the use of Eu3+ and Tb3+ is significant because these dopants are typically thought to be less suitable for anti-Stokes emitters.
![]() | ||
| Fig. 14 Mechanisms of multicolour PA emission and examples of PA systems capable of generating PA emission at various wavelengths. (1st and 2nd row) The base approach with a single dopant ion, i.e. Tm3+ (ref. 2, 6 and 13) or Nd3+.4 Adapted with permission from ref. 4 Copyright 2023, John Wiley and Sons. (3rd row) Sensitized photon avalanche emission obtained for Pr3+, Yb3+.3,10,14 (4th and 5th row) New emission lines enabled by multi-shell architecture powered by energy migration from avalanching core.3–5 Reprinted with permission from ref. 5 Copyright 2023, the American Chemical Society. | ||
In addition to imprinting PA behaviour onto other lanthanide dopants, Skripka et al.5 also demonstrated the first transfer of PA nonlinearity to non-lanthanide emitters. NaGdF4:20%Tm3+@NaGdF4 core shell ANPs were used to transfer high energy excitation to CdS/CdSe/CdS quantum dot (QD) heterostructures, which exhibited nonlinearities of 10.5 at the 630 nm QD emission wavelength under sub-band gap 1064 nm excitation. The experience gained with these examples is important because the stringent conditions for PA emission to occur are much more difficult to satisfy compared to simple ET upconversion. The ability to exploit the work-horse photon avalanche ‘engine’ to achieve PA characteristics broadens the selection of emission spectra to any fluorophore or luminophore, e.g. for multiplexing applications.
However, another contribution to multicolour PA emission was recently shown by C. Wang et al.,8 who employed a Tm3+ ‘PA engine’ and Yb3+ migrating layer to achieve >60 nonlinearities and Ho3+ (S up to 37), Tb3+ (S = 48), Eu3+ (S = 37), Dy3+ (S = 35) and Nd3+ emission bands in the PA mode in NaYF4:Tm (8%)@NaGdF4:Yb/Tm (10/1%)@NaGdF4:X (X = Tb, Eu, Dy, Nd)@NaYF4 materials. Although multicolour PA emission was shown, the emission spectra were obviously composed of strong Tm3+ emission bands at 450 and 800 nm, which are overlapped with additional weaker bands characteristic for the additional ions.
Beside engineering of MPR, the local crystal field affects the Stark level splitting and has been shown to have a profound importance in PA. For example, replacing the Y3+ structural ions with smaller Lu3+ ions led to the contraction of the unit cell, and a slight but negligible blue shift in the available phonons and modification of the local crystal field.28 The spectacularly high, monotonic rise in nonlinearity from S = 41 to over S = 150 was noticed for Lu3+ progressively (from 0 to 85%) replacing the Y3+ structural ions (at a constant 15%Tm doping of avalanching ions), and a single particle-to-particle variable of S > 400 (up to 525) at 500–600 kW cm−2 PA thresholds was explained by the modified magnetic–dipole to electric–dipole transition strength. Even if the results are early and their explanations debatable, this enormously high nonlinearity combined with much shorter rise times of around 9 ms enables the further prediction of new applications in imaging or all-optical data processing. This high nonlinearity enabled the study of S in the center and the edge of 172 nm diameter NPs, confirming the substantial role of surface quenching in PA emission, given that S dropped from 450 in the NP center down to 186 close to the edge.
![]() | ||
| Fig. 15 Novel applications of PA (nano)materials. (a) Highly nonlinear pump power dependence enables a reduction in the effective point spread function volume and allows for sub-diffraction imaging of individual ANPs with single excitation raster scanned beam.2,29 Reprinted with permission from ref. 29. Copyright 2019, the Royal Society of Chemistry. Reprinted with permission from ref. 2. Copyright 2021, Springer Nature. (b) Efficient energy looping and ESA may help reach high energy levels close to the conduction band of the host material and enable reversible, indefinite and bidirectional photo switching by changing the PA threshold.146 Reprinted with permission from ref. 146. Copyright 2023, Springer Nature. (c) Positive energy loop is also susceptible to disruption of the energy looping due to changes in temperature and possibly other physical and chemical factors.10,14 Reprinted with permission from ref. 14. Copyright 2023, AIP Publishing. (d) Positive energy looping within avalanching mechanism is extremely susceptible to quenching by either specific or unspecific quenchers, leading to disruption of both emitting and looping level, and consequently quenching of PA luminescence being significantly more efficient than in the conventional Stokes photoluminescence mode.12 Reprinted with permission from ref. 12. Copyright 2024, the Royal Society of Chemistry (https://creativecommons.org/licenses/by/3.0/). (e) By chemical functionalization of the surface of the UCNP, including also ANPs, with photosensitive ligands tending to agglomerate irradiated NCs, UV/NIR or e-beam lithography is possible.147 Reprinted with permission from ref. 147. Copyright 2024, the American Chemical Society. (f) All optical data processing that mimics functioning of biological synapses; the slow kinetics of the ground (n1), looping (n2) and emitting (n3) levels combined with ESA photoexcitation, lead to paired pulse facilitation and photoexcitation history dependence, which enables optical data analysis such as raster scanned image (e.g. of digits) feature extraction.13 Reprinted with permission from ref. 13. Copyright 2023, John Wiley and Sons. | ||
Compared to conventional non-linear (i.e., SHG/THG, 2/3PhAbs) or upconversion (i.e., ETU, ESA, CS, and CL) processes, PA shows some features that are critically important and key to enable numerous applications, as follows:
• PA brightness is gained from significantly increased concentration of luminescence centers, which enhance both the absorption and emission intensity.
• PA non-linearity is significantly higher and easier to access (in terms of pump power densities) than other anti-Stokes processes.
• PA nanomaterials stem from UCNP technology, which has made advances in many areas including biofunctionalization, surface passivation, and the development of models that yield a deep understanding of internal ET processes that occur between lanthanide dopants.
Simultaneously, the high non-linearity of the PA luminescence intensity makes it extremely vulnerable to the pump source stability and infinitesimal quenching, which put higher demands on the pump sources and purity of PA materials compared to UC materials, respectively. Although the PA pump power thresholds are relatively moderate (10–100 of kW cm−2) considering sensitive biomedical applications, they are still higher than in conventional upconverting materials (single W cm−2). Moreover, PA materials are much more sensitive to the excitation wavelength selection compared to UC, and suitable compact and cheap laser sources are not always available on the market. Therefore, it is important to explore new applications concurrently with the continuing design of new and improved PA materials.
![]() | ||
| Fig. 16 Sub-diffraction imaging with avalanching nanoparticles. (a) Comparison of instrumentation required for conventional raster scanning STED (left) and (b) sub-diffraction imaging using highly nonlinear avalanching nanomaterials (right).9 Reprinted with permission from ref. 9. Copyright 2024, Elsevier. (c) Resolution enhancement originating from highly non-linear emission ref. 28. (d) Photodarkening-based imaging of ANPs demonstrated 2 nm localization precision146 (scale bar: 20 nm). Reprinted with permission from ref. 146. Copyright 2023, Springer Nature. (e) First demonstration of in vitro 3D sub-diffraction imaging with super-linear LnNP labels compared to confocal Stokes and UCNP-based imaging.160 Adapted with permission from ref. 160. Copyright 2019, Springer Nature (https://creativecommons.org/licenses/by/4.0/). (f) Two colour photon avalanche single beam sub-diffraction imaging.9 Reprinted with permission from ref. 9. Copyright 2024, Elsevier. | ||
The need to finely control the optical properties of super-resolution fluorophores has led to the use of nanoparticles for super-resolution imaging.150–152 In particular, lanthanide-doped nanoparticles have been successfully used as luminescent probes in several super-resolution microscopy techniques, including STED,95 nonlinear structured illumination microscopy,161 and near-infrared emission saturation nanoscopy.162 These excellent demonstrations exploit the unique features of lanthanides, such as long luminescence lifetimes (facilitating depletion), perfect photostability and narrowband excitation end visible/NIR emission bands, with relatively low excitation/depletion power density thresholds and cheap and easily available CW light sources.
One of the first STED-inspired approaches for the application of lanthanide-doped nano labels was realized by Kolesov et al.,163 where Pr3+-doped YAG NPs were imaged with a system composed of two beams, a 609 nm pumping pulse and 532 nm depleting one, giving a spatial resolution of 50 nm for the image collected for UV emission. A different operating scheme for application of lanthanide-doped NPs for STED imaging was proposed by Liu et al.95 In this realization, NaYF4 nanocrystals co-doped with Yb3+ and Tm3+ were imaged with pumping with a 980 nm laser beam to observe an emission from the 1D2 state, while an 808 nm donut-shaped beam was utilized for inducing stimulated emission (3H4 → 3H6 transition), and therefore depleting the emission from higher energetic levels. As a result, fluorescence imaging with a spatial resolution of 28 nm was achieved. A similar mechanism in Tm3+-doped hosts was shown to give imaging resolutions of 50 nm for laser scanning through tens of micrometers thick tissue.162
Although STED-based techniques result in an impressive quality of direct super-resolution imaging, these methods still require rather complex dedicated optical systems operating with two concentric beams, one with a Gaussian profile and the other (depleting beam) characterized by a cylindrical donut shape. These two beams must also be properly synchronized in time to most effectively confine the point spread function of the STED microscope, which is not trivial when femtosecond pulses are used (Fig. 16a, left). In contrast, stochastic methods such as PALM and STORM are not necessarily demanding in terms of complex optical setup, but the sequence of excitation, acquisition, and analysis of fluorescence images at single-photon sensitivity is demanding for the photodetectors and requires significant computational power. From a user perspective, the need to computationally reconstruct images means that these images cannot be viewed and interpreted in real time during acquisition.
Extending the concept of using nonlinear materials for improving the optical resolution, Caillat et al.167 proposed the use of lanthanide-doped upconversion, and over a 2-fold resolution improvement was found for Yb3+,Tm3+ upconverting nanocrystals characterized with a nonlinearity order equal to S = 4. However, this value is the ultimate limit, given that the higher the upconversion order, the dimmer and more blue-shifted (into UV region) the lanthanide luminescence. Extending this concept for enhanced optical resolution was possible with photon avalanche, where high nonlinearity orders are enabled (Fig. 16a, right). Moreover, it is achieved without any blueshift in the emission band, in contrast to other multiphoton processes.
In recent years, several imaging methods have validated the theory of PA single-beam super-resolution imaging (PASSI).2,3,6,9,142 Table S11 (ESI†) compares the previously demonstrated super-resolution imaging using lanthanide-doped nanoparticles with currently presented examples of PASSI imaging and its modifications. The first experimental realization of PASSI was the 2021 report by Lee et al.,2 who used Tm3+-doped ANPs with nonlinearities as high as 31 to achieve resolutions finer than 70 nm at 800 nm emission. Notably, these resolutions, which were 7.5-fold finer than the diffraction limit of the 1064 nm excitation, precisely match that calculated ones using the revised Abbe diffraction equation.29 Later, in 2022, multicolour core–shell-shell migrating PA heterostructures with a reported 46th-order nonlinearity were used to label HeLa cells and image them with resolutions reported as high as 62 nm (Fig. 16d–f) at 550 nm (NaGdF4:Yb/Pr (25/0.5%)@NaGdF4:Yb/Tm (8/2%)@NaGdF4:Tb (20%)@NaYF4 ANPs) and 452 nm (NaYF4:Yb/Pr (25/0.5%)@Yb, Tm(10/4%)@Yb/Pr (25/0.5%)@NaYF4).9 Recently, due to the extreme non-linearity of S > 500, FWHM = 33 nm and 36 nm localization precision were achieved for a single, 27-nm diameter ANP.28
These experimental realizations of PASSI and subsequent reports are directly analogous to multi-photon microscopy methods, with the major difference being the ability of ANPs to realize the resolution enhancement of >30–500 photon processes with minimal anti-Stokes shift. Most importantly, these exquisite <30–100 nm resolutions were achieved with inexpensive, commonly available continuous wave lasers and standard confocal microscopes.152,168 These advantages suggest that PASSI has the potential to democratize sub-diffraction-resolution imaging, although some challenges remain to be solved, such as the size of the ANP label compared to biological species, bio-functionalization, and corona formation around inorganic particles entering living tissues.
The ability to switch the bright and dark states of ANPs on demand is reminiscent of photoswitchable fluorophores used for stochastic localization microscopy techniques such as PALM and STORM. Lee et al.146 identified power densities that stochastically photobrightened a sub-population of ANPs such that the centroid positions of individual ANPs could be localized without obfuscation from neighbouring ANPs (which were most likely dark). Due to the robustness of the photobrightening/photodarkening cycle of ANPs, the ANPs could be cycled indefinitely without permanent photobleaching. This indefinite cyclability of ANPs is a critical feature given that the accuracy of photon localization methods such as PALM scales inversely with the square root of the number of collected photons, which is the current limiting factor for the resolution of PALM and STORM. Lee et al. used repeated cycling and localization of ANPs to demonstrate indefinite photon avalanche localization microscopy (INPALM), exhibiting <1 Å localization accuracy even with ANPs that were touching (Fig. 16b).146
The ability to use the same Tm3+-doped ANPs to achieve two different modes of sub-diffraction imaging (<70 nm resolution with PASSI, and <1 Å accuracy with INPALM) demonstrates the unique and powerful applications of PA nanomaterials.
Lanthanide emission is perfectly photostable, which is considered one of the most important advantages of these labels and accounts for either Stokes, upconverting or avalanching nanoparticles. This is critical for biosensing capabilities, given that it not only enables the signal to be reliably acquired over longer dwell times to improve the photon budget at a significantly lower background, but most of all it enables long-term time lapse imaging to be performed over hours of continuous illumination. Although this possibility has not answered any serious biological questions to date (spreading of NPs in the blood stream, small organism development, etc.), these types of studies will definitely gain interest in the near future. Combining cellular component targeting with multiplexed labels and bio-sensing or super-resolution imaging, these novel luminescence labels may offer unprecedented tools for developmental biology or in situ continuous sensing.
In conventional Ln3+ luminescence, short-wavelength excitation beams are required, which simultaneously photoexcite organic fluorescent molecules, and thus increase the background.180 The anti-Stokes emission in UCNPs diminish these problems, but to sensitize their upconversion, Yb3+ ions are used with an excitation wavelength at 980 nm, which matches the absorption spectrum of water, and thus may cause undesired sample heating. With the development of core–shell inorganic UCNP nanomaterials, the possibility to sensitize their upconversion with Nd3+ ions at ca. 800 nm excitation was demonstrated with efficiency and brightness comparable to excitation with Yb3+, and less overheating.181 Thus far, PA has been observed under a few excitation wavelengths in the visible and NIR (e.g., 852 and 1064 nm) regions, which is promising for biomedical sensing. For example, PA was observed in Tm3+-doped ANPs under biocompatible excitation at 1064 nm using power densities typical or smaller than in conventional confocal microscopes.2 These high pump power densities are too high for wide-field imaging, but are acceptable for raster scanning and should not harm biological samples.
In contrast, PA emission should be susceptible to the presence of acceptor molecules in a spectacular manner79 because a small perturbation to the energy looping should be evidenced by a significant perturbation to the photon avalanche emission. In this context, the PA does not change the physics of the dipole–dipole interaction between the donor and the acceptor but enables the quantification of minute FRET signals occurring at distances beyond 100 nm, where the FRET effectiveness drops down below 1% in conventional approaches. Moreover, the volume of the ANPs that interacts with acceptors would increase or the surface passivation (to protect ANPs from unspecific quenching from ligands) would favour the bio-specific response of the ANPs to the presence of appropriate acceptors. This is in contrast to conventional UCNPs used for FRET sensing, where only superficial lanthanide dopants have a chance to interact with the acceptors (Fig. 17) and undergo surface passivation, although to improve the brightness, the acceptors need to be moved further away from the surface of the UCNPs and the donor lanthanides ions.
![]() | ||
| Fig. 17 Issues related to UC-FRET.79 In molecular FRET, the resonant ET between the donor (D) and acceptor (A) occurs owing to the presence of an antigen. The FRET becomes enabled in carefully selected D and A moieties by a D–A distance (rDA) of a few nanometers and spectral overlap of the D emission with A absorption. In upconverting nanoparticles, not only a single UCNP contains 100–1000 of donor lanthanide ions that participate in RET to multiple A molecules, but these individual D ions remain in statistically variable distance from the UCNP surface, where the A molecules can attach to. Obviously, this complicates the analysis. PA emission may become an alternative type of D nanoparticles, which will augment the FRET efficiency, enable surface passivation and biofunctionalization without compromising the FRET sensitivity. Adapted with permission from ref. 79. Copyright 2020, the Royal Society of Chemistry (https://creativecommons.org/licenses/by/3.0/). | ||
Although no PA-based biosensors have been experimentally demonstrated beyond simulations,79 some progress has been made recently in understanding the susceptibility of the PA emission to the presence and concentration of PA quenchers (see Section 7.1.2).12,80 These studies pave a way to develop PA transducers that can detect specific D–A interactions at increased length scales.
PA performance is expected to be strongly affected by the temperature, resulting in a shift in the PA threshold and in variations of the nonlinearity of the emission (similarly to Fig. 15c).10 These macroscale effects are a result of the complex interplay between opposite microscopic effects and especially in PA materials may modify the luminescence in a non-trivial way. Interestingly, in the case of PA materials, high temperature relative sensitivity, SR, was predicted between 4% K−1 and over 35% K−1, with a useful temperature range (where SR > 1% K−1) of ca. 150 K. Moreover, the range with the highest sensitivity may be tuned by the selection of the pump power used for probing. Optimizing the system for higher sensitivity ranges results in limited temperature useful regions. As a result, two opposite effects can be observed. On the one hand, the capability to select a particular calibration curve (T = IL−1 (T)) tailored for the expected sensitivity and broadest temperature sensitivity range provides the unique ability to dynamically adjust the optical thermometry performance for the given utilization and for its specific requests. Nevertheless, this effect should most probably limit the practical application of PA luminescence thermometry of these materials due to the necessary use of a matrix of pump power and temperature-dependent calibration curves. The other limitation in the utilizing of these types of materials in thermometry is their thermal stability and thermal decomposition at >580 K.199 Other effects, such as nanoscale effects, presence of ligands and solvents (discussed in Section 4.4), as well the presence of quenchers (discussed in Section 6.2.2), will further make the PA susceptible to the external chemical environment. Thus, the understanding and proper management of these factors are critical for the further development of new PA nanomaterials and their applications.
Therefore, alternative luminescent thermometers and reliable methods for temperature-dependent spectral properties quantification are sought, and the PA concept may offer new possibilities (Fig. 3b). In this case, the so-called single-band ratiometric (SBR) approach,200–203 the luminescence intensity of a single emission band can be monitored under two, ground and excited-state absorption photoexcitation lines. This approach is advantageous given that it is technically less challenging and cheaper to all-electronically switch between 2 laser/LED light sources and record two (GSA and ESA based) images within a single spectral channel than the opposite of recording two spectral images under a single photoexcitation line. Additionally, by using the same spectral range of emission, the risk of disturbing the temperature readout due to the influence of the medium will be eliminated. The unique configuration of the 4f energy levels of Ln3+ opens a wide variety of different ions for which the ESA process can be observed. The former, temperature-dependent ESA-based thermometry has been recently proposed and studied in Nd3+, Eu3+, Tb3+- and Pr3+-doped nanoparticles204–207 In the case of materials with thermal coupling between the ground and excited states, they enable an enhancement in the ESA-excited luminescence by an increase in the population of the upper energetic state according to the Boltzmann distribution. However, versatile studies enabled the understanding that the interionic interactions, usually via the CR process, lead to boosting of the thermal response of the luminescence signal. Therefore, the optimization of the dopant concentration has led to the development of luminescent thermometers with sensitivity to temperature changes as high as >10% K−1. Although the probability of the CR process is temperature independent, the splitting energy of the particular 4f multiplets involved in this process may shift the CR process out of resonance. Therefore, the assistance of phonons will lead to the strong thermal dependence of the CR process (see Section 3.2.2). Given that the splitting energy of the 4f levels is dependent on the crystal field strength of the host material, by deliberate selection of the host material stoichiometry, the probability of this process can be modified. Considering the essential dependence of the PA process probability on the CR performance, the sensitivity of luminescent thermometers using photon avalanching nanoparticles will be unprecedently high and should definitely exceed the values reported thus far for thermometers based on the ESA/GSA SBR approach.208,209 The relative sensitivity of PA-based luminescent thermometers reaching tens of % K−1 in a broad and tuneable (by tuning the excitation power) temperature range was predicted recently with the numerical simulations employing the phenomenological model of an avalanching system.10
Beside technical issues, metrology factors are also important and the two emission wavelengths (in traditional ratiometric imaging) or the two excitation wavelengths (in ESA/GSA thermometry concept) may be affected by the sample in different ways, which is strongly related to the sample spectral properties (e.g. dispersive tissue absorption and scattering coefficients).209 These aspects are difficult to predict, given that the sample-to-sample and patient-to-patient variability may hinder simple external temperature calibration.203,209 A solution can be offered by the temperature-dependent kinetics of the luminescence rise or decay times. In the case of luminescence decay, various luminescent compounds have been proposed, whose lifetimes are typically shortened with an increase in temperature due to thermal quenching. However, in the case of thermal imaging, the acquisition time needs to be shorter than the decay itself. Therefore, the shortening of the decay observed at elevated temperatures may affect the number of photons absorbed, and thus the reliability of the temperature estimation. On the other hand, in case of the PA process, for the threshold conditions (excitation density and temperature), the rise time is extended. Therefore, kinetic-based thermometry may beneficially influence the signal-to-noise ratio by facilitating a reliable temperature measurement. Although this approach has not been confirmed experimentally to date, its applicative potential for remote temperature determination from the theoretical perspective is indubitably appealing.
As can be noted in Fig. 3b, temperature may contribute to the PA emission by enhancing the population of the starting level. An educated guess leads to the conclusion that an increase in temperature should reduce the PA threshold, but due to the rich energy level scheme of most lanthanides, temperature may also negatively impact the PA emission, leading to thermal quenching by enhancing multiphonon relaxation rates.10 Moreover, the rise times of luminescence under PA photoexcitation and looping conditions is not only affected by temperature, but also dependent on the excitation intensity, IEXC, and due to the sample properties, IEXC is actually difficult to be determined. Therefore, the latter features may hinder the practical implementation of PA-based thermometry at the nanoscale in ‘difficult’ environments, such as absorbing and scattering samples (e.g. tissue and cells).
The fact that a slight change in temperature can affect both the spectral broadening of the absorption bands and the activation of energy bridging the CR processes means that even a small change in temperature can cause rapid (up to a dozen or even orders of magnitude) changes in the intensity of luminescence. In this case, the expected rapid changes in thermometric parameters should take place in a narrow temperature range. However, a wide range of possibilities of selecting the appropriate energy levels and activating various physical processes through the appropriate selection of the luminescent host material and Ln3+ admixture ions offered by the unique configuration of the 4f energy levels can enable the design of a wide range of super-sensitive luminescence thermometers/manometers based on the PA process, whose operating range can cover a wide range of temperatures/pressures. A necessary but not sufficient condition for the occurrence of the PA phenomenon is an efficient ESA process. The first attempts to create luminescence thermometers using the ESA process in nanocrystals, glasses and microcrystalline materials doped with lanthanide ions such as Nd3+, Pr3+, Tb3+, and Eu3+ confirmed the high potential of this type of materials for remote temperature reading, which was confirmed by the sensitivity of nearly 15% K−1. Moreover, the favourable role of CR processes on the variability in temperature parameters was confirmed, which indicates that for the conditions required for the occurrence of PA, high relative sensitivities will be obtained. A limitation of this type of thermometer may be the fact that it will be necessary to provide a reference signal enabling a ratiometric temperature reading. This can be done using the SBR approach, which due to the need to change the wavelength, may cause some experimental difficulties. Another possibility is the use of the emission band of the co-dopant ion, for which the excitation wavelength used to initiate the PA process, allowing luminescence to be obtained by GSA. This concept has been proposed for ESA/GSA processes but is unpopulated for the PA process thus far. This is because interionic interactions can disturb the equilibrium conditions for PA occurrence. Nevertheless, the appropriate selection of dopant ions and the optimization of their concentration may enable the creation of this type of thermometer in the future. The undoubtful profits coming from the utilization of PA-based luminescence thermometry resulting in the achievements of the nano-spatial and 10−3 K thermal resolution of thermal imaging are sufficient motivation, initiating research studies devoted to this in the near future.
Thus far, beside homogenous temperature-dependent PA studies (Fig. 15c)14,203,209,210 and theoretical modelling of temperature dependence of PA emission,10 no experimental evidence of temperature quantification at the single ANP level has been demonstrated.
![]() | ||
| Fig. 18 Photon avalanche-based force sensing. (a) Scheme of the experiment, measurement of the impact of the mechanic force applied by an atomic force microscope tip on avalanche emission of single ANPs; (b) force-induced shifting of the power dependence observed for ANPs (doped with 4.5–8% of Tm3+); (c) force-induced avalanching of pre-ANPs (doped with 4% of Tm3+); and (d) mechano-chromic effect observed for piezo-chromic ANPs (doped with 15% of Tm3+).211 Adapted with permission from ref. 211 (https://creativecommons.org/licenses/by/4.0/). | ||
The application of PA to force sensing was demonstrated in experiments where Tm3+-doped NCs were studied spectroscopically with forces applied in situ using an atomic force microscope tip. Importantly, the impact of the force on the emission varied with the doping concertation of the NPs under study. In the case of ANPs doped with 4.5–8% of Tm3+ ions, for which the full-performance PA is observed in a single ANP (ϕ ∼ 25 nm), applying force (from the hundreds of pN to few μN range) resulted in a shift in the power dependence curve towards higher values (Fig. 18b). This enables force sensing based on the luminescence intensity from a single ANP. Alternatively, reducing the doping concentration (to 4%) results in ‘pre-avalanching’ nanoparticles that exhibit energy looping but not the steep emission intensity rise characteristic for PA. In such pre-ANPs (Fig. 18c), applied mechanical force and reduction in the average distances between the photoactive ions in the NCs pushes these nanoparticles into the avalanching regime, resulting in much brighter luminescence. In a third force-sensing modality, much heavier doping (15%) enables remote force sensing based on the modulation of emission peak ratios with force, known as mechano- or piezochromism. This approach has been demonstrated with the emission from two Tm3+ energy levels (3F2,3 and 3H4), whose relative intensities respond differently to applied mechanical force (Fig. 18d). It is believed that a force-induced increase in the nonradiative rates results in the relaxation of energy from the higher level (3F2,3, emission at 700 nm) to the lower one (3H4, emission at 800 nm). This enables sensing of forces with single digit nN resolution.
Besides non-volatile memory, the photoswitching properties of Nd3+-doped KPb2Cl5 ANPs have been leveraged as prototype optical transistors. Skripka et al. demonstrated the ability of these bistable ANPs to move between their dark and bright states without changing the intensity of the 1064 nm pump laser.212 A low-power (<100 nW) 808 nm laser pulse was used to cross the instability region that separates the bright and dark states of these ANPs at intermediate powers (thereby creating the hysteresis). This instability crossing switched the ANP emission from a dark to bright state, with the latter maintained by the continuously applied 1064 nm bias. In this configuration, many ANPs may be excited with a diffusely spread 1064 nm laser and flipped from a dark to a bright state by raster scanning with an 808 nm laser in any arbitrary pattern. This low-power, fast, and flexible photoswitching is particularly attractive in developing all-optical processors, memory, and interconnects, although advances in these IOB nanomaterials are needed to increase their operating temperatures to ambient conditions.
This application leverages the photoexcitation history dependence, which can be especially pronounced in PA given that it is severely dependent on the initial population of the looping level. In a simplified perspective, to observe PA emission, a certain seed population of the looping level must be established through processes of GSA and ESA. A larger level population is facilitated by significantly longer lifetimes, typically in the order of tens of milliseconds,6 in contrast to other levels with decay times measured in the tens to hundreds of microsecond range. Although the emitting level exhibits short electron lifetimes and can be depopulated easily through both radiative and non-radiative relaxation, as well as other non-radiative processes such as ET, the looping level may retain a residual population even after a few tens of ms. This long-lived population of the looping level can be likened to a form of system short-term memory. By exciting the PA material with a pulsed light source, with a time gap between the pulses smaller than the lifetime of the looping level, one can observe that the emission is dependent on the train of photo-excitation subsequent pulses, and the first pulse facilitates the emission achieved in the course of the second and subsequent pulses (Fig. 19). The facilitation effect stems from the fact that with the first few pulses, the looping level electron seed population is established, and the subsequent pulses start with non-zero population and contribute to further population growth, which enhances the emission intensity with each subsequent pulse (Fig. 19). The intensity of this emission, which is dependent on the photoexcitation history, can be controlled in situ by adjusting the pulse frequency, pulse and gap duration, and pulse power.13
![]() | ||
| Fig. 19 Complex nature of luminescence originating from the multiple level and multiple dopant interactions in PA mode, where ESA absorption depends on the residual population of the looping level (n2), and thus, two subsequent excitation pulses (a) induce a pulse-gap dependent emission and the so-called (b) paired-pulse signal facilitation (PPF).13 Adapted with permission from.13 Copyright 2023, John Wiley and Sons (https://creativecommons.org/licenses/by/4.0/). | ||
This memory-like behaviour of PA resembles the operational principles of biological synapses, given that it demonstrates in situ plasticity, inherent short-term memory, and signal facilitation across multiple subsequent pulses. This similarity of the observed behaviour of PA upon pulsed excitation to neural circuits holds particular interest from the perspective of all-optical, training-free, reservoir computing, an alternative computational paradigm aiming to mimic brain function, to reduce the computation time and energy consumption in analysing temporal signals. Additionally, owing to their short memory and high nonlinearity, PA materials can function as optical, analogue AND logic gates with multiple optical inputs.13
There are also other characteristic features of the synaptic operation that have been evidenced, i.e., the threshold intensity of the stimulus below which it is neglected, as well as the paired pulse facilitation effect. Both of these features match the characteristics of avalanching materials. One of the required features thereof is a metastable long-living intermediate energy level, given that the 3H4 level in Tm3+ is characterized with a lifetime of around 16 ms,6,13 which is much longer compared to the lifetimes of the levels from where avalanche emission occurs. One of the imminent features of the avalanching is slowing down of the intensity rise dynamics for excitation powers close to the PA threshold. However, the presence of the oddment population in the metastable level at the beginning of illumination leads to a significant reduction in the rise time required to achieve a steady-state emission intensity. Thus, during pulsed excitation with time gaps between the pulses comparable in duration to the lifetime of the metastable level, at the beginning of each of the consecutive pulses, the starting population in the metastable level is a bit higher compared to the previous one, facilitating faster build-up of the emission intensity and providing an increased population of metastable level for the next pulse. This effect can be understood as the short-term (on the timescale of the lifetime of the metastable level, up to tens of milliseconds, typically) memory required for the realization of the paired pulse facilitation effect, in which the response to the stimulus appearing shortly after former stimulus will be enhanced.
This effect, with operation at some points similar to that found in natural synaptic systems, allows all-optical data processing.13 This includes pulse counting enabled by the short-memory effect caused by maintaining the oddment population in the intermediate state of the avalanching system between the pulses. Thus, for each next pulse, the response from the system (emission intensity) depends on the number of these pulses. However, this effect may be further developed by adding additional parameters modulating the rise dynamics of the avalanche emission. In general, this dynamic not only depends on the frequency of the pulses and their duration and width, but also on the sequence of these parameter changes. This sensitivity to the excitation history opens other fields of application, including pattern recognition, features amplification or extraction, and phase-sensitive detection.13
Although the discovery of the PA phenomenon was the result of serendipity, actual progress in the intentional development of PA materials was triggered by understanding the fundamental differences between UC and PA and the shift in paradigms such as an increase in the concentration of the dopants (against conventional thinking about concentration quenching in lanthanide-doped phosphors) and the positive role of energy CR. Therefore, despite their many similarities, these new PA (nano)materials require de novo optimization of their composition. Obviously, the developments in PA materials stem from the last two decades of intensive progress and optimization of upconverted materials and technology. The most prominent contributions of UCNP to ANP technology are as follows: (i) the existence of reliable protocols for the synthesis of colloidal nanoparticles, (ii) ligand stabilization of individual nanoparticles, (iii) understanding the surface quenching mechanisms and (iv) clustering of dopants, resulting in great interest in single nanoparticle studies and further enabling the development of core-multiple-shell designs. Concerning core–shell nanoparticles, surface passivation with undoped homo- or heterogeneous shells was the key technology enabling the practical implementation of UCNPs in biosensing and bioimaging. Another great step forward was the studies dedicated to a compositional architecture, i.e. intentionally doped core and individual shell layers with various compositions of co-dopants, which enabled the engineering of the luminescent properties of these nanoparticles on demand. These fundamental achievements will further stimulate work and discoveries related to interfacial energy migration and lead to improved control of the luminescence colour and lifetimes of PA nanomaterials.
Because of the slow rise time of PA emission, some technological aspects must be also developed, such as pre-pumping (to speed up the raster scanning) and parallelization of multiple excitation spots (to image and localize ANPs in many areas simultaneously) similar to localization microscopy or Nipkow-disk approaches.217 Optimization, i.e. shortening of the PA emission rise time, is another way to circumvent the slow frame rates of current PASSI-type imaging.3,28
Moreover, PA sensitization with Yb3+ co-doping can potentially be beneficial to improve the brightness and reduce the PA threshold and rise times of ANPs. The use of Yb3+ ions enables connotation with core–shell nanomaterials, where the Yb3+ ions are responsible for the energy migration between the absorbing core and emissive shell, which can potentially increase the number of available PA colours, aiming to perform multiplexed bio-detection or bio-imaging.
In the case of conventional luminescence, the presence of Q reduces the emission output of the donor (D) (IDQ) and shortens the luminescence lifetime of the donor (tDQ) compared to bare donor (ID and tD, respectively) proportionally to the D–Q distance, rD–Q, which can be quantified using the well-known Förster relations and theory.182 In contrast to RET occurring in the conventional photoexcitation scheme (Fig. 20a), for the PA process, the presence of Q may affect the emitting (n3), looping (n2), or both levels, while both levels participate in the energy looping and ESA. In mechanism I (QI), the quenching occurs to the emitting level, which cannot participate in the CR with neighbouring ions in the ground state. In mechanism II (QII), the quenching occurs to the looping level, which hinders ESA. These two mechanisms should affect the PA features in different manners.12
The impact of the quencher on the emitting level has been theoretically considered by Bednarkiewicz et al.79 The impact of the acceptor was defined as the relative ratio of WRET to the CR rate, WCR, existing in an unperturbed PA system. Next, the steady-state luminescence ratio ΩSS of D accompanied by the acceptor, which is related to the luminescence intensity of the donor alone, was defined as
enabling the quantification of how the PA emission reacts to an increase in the D–Q effectiveness when assuming that the FRET mechanism (R0 = 5 nm) is responsible for this ET distance. These simulations indicated the significant susceptibility of the PA emission intensity to perturbations introduced by the quenching species, and consequently it was hypothesized that the interaction distance between D and Q can be extended by a few-fold (2–6) beyond the conventional Förster distance. To some extent, these simulations have been recently confirmed experimentally by studies devoted to unspecific quenching80 and by cooping the Nd3+ quencher in the avalanching Tm3+ ions.12
The unspecific surface quenching has been demonstrated and modelled at the single-nanocrystal level to uncover the design-dependent heterogeneity in the ANP threshold intensity.80 The 8%Tm3+-doped NaYF4 core of the Rc radius was passivated by the undoped NaYF4 shell with a variable thickness (from ca. 3 to ca. 9 nm). As expected, a correlation was found between the mean threshold pump power and the ANP shell thickness (d ± δ, where δ is the standard deviation of the shell thickness) up to ca. 9 nm, beyond which the ANP behaviour did not depend on the ‘unspecific’ ligand field anymore. It was found that when the CR rate is large, the total depopulation rate, A2 + k2, of the first excited 3F4 level in Tm3+ is proportional to the product of the PA pump threshold and ESA cross-section. Interestingly, the non-radiative quenching through the surface (k2) was well described by the surface-quenching model, i.e. k2 was found to be dependent on
where κ [nm−1] is the exponential passivation improvement with shell thickness (Fig. 21a and b).218 FRET theory was considered in this work; however, it provided less accurate results compared to the surface-quenching model, which most probably originates from the differences between the specific (FRET) and non-specific quenching.
![]() | ||
| Fig. 21 Impact of quenching on PA emission. (a) Unspecific quenching of Tm3+–Tm3+ avalanching pair by surface effects in Tm3+-doped NaYF4 ANPs, (b) mean ± variance pump power threshold versus the shell thickness for 2.6, 5.6 and 8.6 nm shell thickness. (b) Correlation between shell thickness and the mean ± variance of single-ANP Ith distributions.80 Reprinted with permission from ref. 80. Copyright 2021, the American Chemical Society. (c) Specific quenching of the Tm3+–Tm3+ avalanching pair by volumetric Nd3+ ions. (d) Pump power-dependent luminescence of Tm3+ ions for an increasing concentration of the quencher in LiYF4 microcrystals and intensity profiles (inset) at fixed pump powers.12 Adapted with permission from ref. 12. Copyright 2024, the Royal Society of Chemistry (https://creativecommons.org/licenses/by/3.0/). | ||
The specific RET quenching of the emitting and looping levels of the looping Tm3+ ions has been recently tackled by incorporating Nd3+ acceptor ions within the LiYF4 host material12 (Fig. 21c and d). The microcrystals were proposed to exclude surface effects and unspecific quenching, which in principle enabled control of RET in a more precise way by varying the average donor–acceptor distance, which was controlled by the acceptor concentration. The 2H9/2,4S5/2 and 4I13/2 levels of Nd3+ remained resonant to the 3H4 (emitting) and 3F4 (looping) levels in Tm3+ ions, respectively. The GSA and ESA-based photoexcitation were directly compared, which evidenced many important aspects of PA behaviour. Firstly, a gradual increase in the acceptor concentration from 0.1% to 1% led to gradual quenching of the Tm3+ emission was observed both in the steady-state and kinetic domains, but while the GSA only led to a 4 times decrease in the Stokes emission intensity, in the ESA mode, the PA emission dropped by 4 orders of magnitude. Simultaneously, the luminescence decay dropped from 240 μs to ca. 80 μs (for 1%Nd3+ co-doping), which confirmed the non-radiative RET mechanism. By pump–probe measurement, the luminescence lifetime of the looping level at ca. 6000 cm−1 was dramatically shortened from 13.6 ms for the pristine Tm3+ only-doped crystals to 5.2 ms for as little as 0.1%Nd3+ co-doping, which obviously disrupted the energy looping, resulting in a chance to achieve PA emission. Higher acceptor concentrations prohibited the further kinetic characterization of the looping levels. These experimental studies were supplemented with further theoretical modelling based on the phenomenological model of Tm3+ PA in NaYF4 nanoparticles,2 which remains a very useful method to evaluate the impact of the phenomenological parameters on the PA performance.6,10 Here, this model and the rate equations describing both the looping and emitting levels were extended with the RET mechanism. These simulations enabled the impact of the acceptor individually on the emitting and the looping levels to be distinguished. These theoretical considerations are extremely valuable because it is difficult to find and quantitatively compare the impact of the specific quenching of the looping level because it is typically located in the range of 2000–6000 cm−1. Moreover, the Nd3+ levels are suitable to quench both the looping and emitting levels of Tm3+ ions, and because within single ANP there are hundreds to thousands of avalanching Tm3+ ions, the output luminescence is the sum of luminescence from all the ions exposed to quenchers staying at the average distance. There, the simulations clearly showed distinct effects for the two quenching mechanisms, i.e. through quenching of the emitting (QI) and the looping (QII) levels. The QII mechanism increased the PA threshold but in principle did not affect the nonlinear behaviour and brightness in the saturation regime. The QI mechanism reduced the nonlinearity but additionally reduced the brightness. The combination of both processes should be actually observed in real systems.
These results suggest that PA emission is significantly susceptible to the presence of acceptor/quencher molecules at distances, which are a few-fold larger than that expected from the Förster distance calculated for the same D–A pair. This susceptibility is not related to any change in the properties of the D or A alone or the fundamental mechanisms behind RET, but originates from the extreme susceptibility of the donor excitation to minute perturbations in photoexcitation and energy distribution on the looping and emitting levels. The predicted and demonstrated susceptibility is of critical importance because of the practical design of in situ bio-sensors.
Another reason to minimize the excitation power is to avoid inadvertent photodarkening. Although photodarkening has been leveraged for photo switching and sub-diffraction imaging,169 in other applications, photodarkening can be a major disadvantage of ANPs compared to conventional UCNPs, which are generally stable under most excitation powers. Further research into the mechanism of photodarkening in different PA nanosystems is needed, given that there are studies on the mitigation of photodarkening, e.g., through the development of synthetic methods and structures that suppress the defect states in the host matrix.
It is also critical to mention that the optical setup used to characterize the PA properties should be extremely well evaluated and optimized because any unconscious variation in the pump power in the PA regime (originating from the laser diode intensity/laser optical mode switching/positioning of the diffraction limited beam against the sample, short and long term instabilities, switching optical components in the light path) can be a source of significant errors. In view of the possible variation in the derived parameters depending on the pump power intensities (due to photodarkening,214 sample density and way of preparation, single NPs versus a monolayer of NPs versus aggregates of ANPs), the experimental conditions should be precisely reported. Moreover, for single ANP studies, a change in the ANP position against the laser beam should be verified before and after the experimental evaluation, given that tiny sample shifts in XYZ space occurring at the time of the studies may significantly contribute to the outcome and interpretation of the PA results. Moreover, thermal issues should also be avoided and evaluated in the analysis of the results. Typically, large aggregates of particles and a high-power laser with a large beam size will make the samples generate excessive heat, which will be evidenced with broadened emission lines up to white/yellow broadband emission generation and irreversible sample burning and damage.
In any biosensing system, the bio-specific interaction between the optode and the analyte should be translated to a variation in some spectral properties capable of being quantified. In luminescence, they are typically luminescence spectra (or emission intensities at specific wavelength(s)) or luminescence kinetics. In the photon upconversion mode, the typical short wavelength photoexcitation from the commonly applied light sources (lamps with filters, LEDs or sometimes lasers) must be exchanged with stronger NIR lasers matching the sensitizing Yb3+ (at 980 nm) or Nd3+ (at 800 nm) ion absorption. This imposes further modifications in the spectral filtering of the detection system. Firstly, anti-Stokes emission requires short-pass filters to replace long-pass ones. The large Stokes shift simplifies filtering out the laser excitation, but the small QY translates to a large discrepancy between the excitation and emission line, and thus the dichroic filters must be effective in reflecting the NIR excitation and transmitting the weak visible luminescence. Beneficially NIR photoexcitation undergoes scattering to a lesser extent compared to short wavelength photoexcitation. Alternatively, the tens of microseconds to milliseconds long kinetics are technically much easier to be measured compared to the very demanding time-correlated single-photon counting techniques required for (sub)-nanosecond-long kinetics. However, these slower kinetics are related to dimmer luminescent labels, and ultimately slower fluorescence imaging, e.g. during raster scanned imaging. In summary, these properties are distinct from conventional fluorophores and require customization of the detection instruments (e.g. well plate readers, and microscopes), but it is clear that the benefits of UCNPs justifies this work. Furthermore, the requirements needed to generate and utilize photon avalanche are not very different than that valid for the conventional UCNPs. With current developments in semiconductor lasers, available photodetectors and appropriate filters, all these modifications are easily within reach. PA requires relatively high pump power (in the order of tens to hundreds of kW cm−2), which precludes wide-field imaging. Nevertheless, ANPs offer other unprecedented features such as single-beam photon avalanche super-resolution imaging (Section 6.1.3) at no additional cost. This means that converting any confocal microscope to the anti-Stokes luminescence mode is not only beneficial for an improved signal-to-background ratio, but with no charge, enables imaging below the diffraction limit with resolutions of ca. 60 nm. Combining these features with the expected bio-sensitivity of ANPs should enable functional super-resolution imaging in the near future.
Despite the engineering of materials with different visual colors,221 a limited number of spectral fingerprints can be effectively considered when lanthanide-doped nanoparticles are not overlapping spatially. Otherwise, two labels that differ only in proportion between, e.g. the green and red emission band ratio, will not be distinguished if they co-localize in space (e.g. within the point spread function of the confocal photoexcitation beam). A number of attempts tried to provide spectrally pure emission by precise optimization of the co-doping with various lanthanides223 or by co-doping the upconverting lanthanides with additional ions (e.g. Mn2+ ions) that can quench some of the energy levels, e.g. green Er3+ emission.224 Moreover, although lanthanides produce characteristic (lanthanide and host dependent) emission bands, recognizing these labels is not trivial and requires high-resolution spectrally resolvable photodetectors, which make these detection systems more complex, costly, slower and computationally (due to spectral deconvolution) demanding.
The multiple band emissions from lanthanides have been demonstrated to be suitable for the so-called ratiometric sensing, which exploits the inner filter effect. For example, Er3+-doped nanoparticles emit simultaneously at 520–540 nm (green) and 650 nm (red). When only green emission intensity is modulated by external factors, for example by pH-sensitive dye, whose absorption spectrum varies in this green spectral region, the erbium red emission may serve as a reference for the pH (indirectly)-dependent green emission, and after a simple calibration procedure this green-to-red ratio pH probe can be easily detected or imaged. It should be noted there is no direct interaction between the emitting erbium ions and the pH-sensitive dye, but the erbium green emission from the nanoparticle is spectrally filtered (modulated) by the dye anchored at the UCNP surface. PA emission should show similar features, but the very steep nonlinearity of ANPs to the pump power may complicate the technical side of the measurements. Alternatively, PA offers an unprecedented referencing feature because the PA emission under ESA photoexcitation may be used as a FRET sensor reporter, while the emission at the same wavelength but excited in GSA mode may provide the reference signal. Alternatively, the very slow and pump power-dependent PA rise times have not been explored at all thus far. All these possibilities will surely be studied in the future.
In contrast to multicolour spectral fingerprints, fingerprints that enable various luminescence labels to be distinguished in the time domain have been also proposed and demonstrated. For example, by varying the concentration quenching labels that emit photons of the same energy (colour), but whose luminescence lifetimes are designed such that the luminescence kinetics may be analysed and ascribed. PA emission is quite specific and different from the conventional upconversion. All the problems with creating multilabel (multicolour or time tags) emission are also valid for ANPs, but the specific features of PA emission may enable the generation of additional emission lines or the capability to emit only at a specific photoexcitation wavelength. There is some expectancy in sensitized avalanche emission and core–shell ANP designs, which under a single excitation wavelength may generate numerous new emission lines. However, experimental demonstrations are missing thus far, which will gather interest in the near future.
Multiplexing, i.e. the ability to target multiple analytes in liquid or complex samples such as cells of tissues, requires unique and easily quantifiable spectral codes. Considering the recent emergence of PA nanomaterials, the range of emission wavelengths and excitation wavelengths for ANPs is still limited, presenting an opportunity for the continued development of methods to expand the library of ANPs. A good example of tunability for multiplexing is quantum dots, where the quantum confinement effect produces a spectrally distinct shift in the emission bands. Despite their efficient absorption and emission, improved photostability and small size, QDs demonstrate relatively broad and overlapping emission bands, and ubiquitous short wavelength excitation, which overlap with endogenous chromophores in the absorption, scattering and excitation regions.
To realize the multiplexed imaging of multiple labels, a wider range of spectrally distinct ANP labels must be developed. There are four viable strategies for using lanthanide-doped ANPs as luminescence labels, as follows:
1. The first is displacing these labels in space or sample volume, aiming to address them one-by-one by raster scanning confocal imaging. However, due to light diffraction, the best spatial resolution offered does not go below a few hundreds of nanometers. Therefore, despite the multicolour emission, achieving many unique fingerprints with UCNP LnNP labels remains a challenge.
2. The second approach relies on creating pure, single emission band labels, which enable the detection of various LnNP labels in the spectral domain even when the spatial resolution is limited. However, due to the nature of the Ln3+ emission, this approach can only generate a limited number of spectral codes, often reducing the brightness.223
3. The third approach relies on time domain labels. However, although the lifetimes of LnNPs are long (μs–ms) and they can be engineered to design up to a few time tags,225 the possibility to distinguish them in complex samples containing multiple labels within the diffraction limited volumes is in practice difficult for numerous reasons, e.g., brightness under short pulses becomes weak and results in a poor signal to noise ratio.
4. The last possibility is using various excitation wavelengths suitable to excite a particular LnNP. However, although short wavelength excitation is often suitable to excite multiple Ln3+ nano-labels, it is typically undesirable in bio-applications because it stimulates bio-sample autofluorescence and enhances the background signal. In the case of upconverting labels, NIR excitation (at 920, 976, 800 or 1530 nm) can initiate visible emission, but this remains a technical challenge to switch between light sources, avoid crosstalk between these excitation lines and optically align multispectral images under various excitations.
Therefore, it remains highly desirable to develop ANPs as nanoscale labels, exhibiting multiple pure colour emission spectral fingerprints under a single excitation laser line. However, thus far, only a limited number of singular lanthanide ions in a limited number of host materials under well-adjusted excitation wavelength of sufficient excitation pump power density has been capable of generating PA emission at the nanoscale.
There is great demand for the detection of MIR photons due to the importance of this radiation in many applications, such as spectroscopy, metrology, (bio)imaging, molecular analysis of gases and astronomy.238–242 Unfortunately, the detection of MIR radiation in a sensitive and cost-effective way is challenging, mostly due to the insufficient sensitivity of the available detectors in this wavelength range.242 Moreover, its detection is hampered by the high noise signal due to the natural emission above absolute zero in the infrared range. Thus, there is a substantial need to develop new MIR photon counters allowing to bypass the limitations resulting from the low energy of the analysed radiation, and hence the high thermal background.243
One of the proposed methods for the detection of the MIR signal is its conversion to visible radiation, where the detectors are much more sensitive, cheap and widely available. An example is the upconversion process, which in the most well-known case shifts the NIR photons to a higher energy in the visible, and thus it can be detected with the generally available detectors.243 In 1959, Bloembergen proposed an IR quantum counter (IRQC), the operation of which is based on a two-step excitation process.244 In the Bloembergen IRQC, the operation scheme includes the first step when infra-red radiation excites the atoms to a metastable level, followed by the next step in which the intense pump excites the atoms to a higher emission level. Therefore, the IR signal is incoherently upconverted to emission in the spectral range where very sensitive and much cheaper detectors are commonly available.
Based on the Bloembergen approach, in 1968, Esterowitz et al. proposed an ideal five-level IRQC system (Fig. 22a), where the IR signal transition (1 → 3) is followed by rapid depopulation via a nonradiative transition to the lower lying metastable level (2).145 Subsequently, the pump transition (2 → 5) occurs, which is followed by a fast nonradiative transition to the fluorescent lower lying level (4). The proposed system allows the problems occurring in the simplest IRQC system with three energy levels to be bypassed, in particular to eliminate the reabsorption of the resulting radiation by the active ions and pump saturation. Moreover, additional levels allow the electron to be recycled, and thus for each absorbed IR photon, more than one visible photon may be emitted. In their work, the authors reported 165 schemes of IRQC found in trivalent rare earth ions in various host materials, the amount of which, as indicated, can be greatly increased by the addition of another co-dopant ion (e.g. Yb3+). However, most of the reported IRQC schemes were established for IR radiation shorter than 3 μm. Recently, Liang et al.238 proposed an approach with an inverted sequence, with continuous visible pumping responsible for promoting the ions from the ground to the excited level, which is followed by fast energy relaxation to the lower energy levels, and then NIR emission. However, presence of the MIR photons results in the ions being excited back from this emitting level to a higher one, from where also higher energy emission occurs. Thus, the intensity of the MIR radiation affects the luminescence intensity of these two emitting bands.
![]() | ||
| Fig. 22 MIR photon counting (a) five-level quantum counter scheme (hvp – pump, hvs – signal, hvf – fluorescence. Wavy arrows indicate non-radiative transitions).145 (b) IR detection scheme based on Pr3+.245 Dashed arrows are possible cross relaxations (CR) responsible for photon avalanche phenomenon. Transparent lines are additional possible CR and fluorescence transitions not shown in ref. 245. | ||
PA medium infrared quantum counter detectors (MIRQC) were first reported in 1979 by Chivian et al., who while studying materials for IR quantum counters, discovered the PA phenomenon in Pr3+-doped LaCl3 and LaBr3 hosts.1 The underlying concept was simple, i.e., the materials were pre-pumped with wavelength matching ESA (3H5 → 3PJ) instead of GSA (3H4 → 3PJ) excitation. Above some pump power density threshold (ca. 1.2–12.2 W cm−2), a sharp 100-fold increase in the red luminescence intensity at 600–640 nm (3P0 → 3H6, 3P1 → 3F2, 3P0 → 3F2) was observed. When the system was close to the PA threshold, the 4.5 μm photons (matching ground 3H4 to 3H5 level of Pr3+ ions) at 1–10 mW cm−2 intensities were sufficient to trigger PA emission, and consequently a 4-fold increase in the sensitivity of MIR photon detection was demonstrated.1 The interpretation of the obtained results was based on the sequence of ion interactions, in which the absorption of one photon in the infrared region led to the generation of multiple excitations.
In 1983, Krasutsky presented the first crystal quantum counter working at a wavelength longer than 5 μm. This system, based on the LaBr3:Sm3+ crystal, was sensitive in the spectral range of 2 to 10 μm due to the utilization of a low phonon material.37 The authors described the pump-dependent background noise as the main limitation of the studied system detectivity. The PA, observed above pump power densities of 3 kW cm−2, was considered a limitation mainly due to difficulties in performing experiments and ensuring pump stability and intensity. Consequently, the PA phenomenon set an upper limit on the maximum pump power density that could be used for IRQC experiments.
Although benefits arising from the utilization of PA were not fully recognized initially, in 2003, D. B. Gatch et al. showed that due to PA, the signal-to-noise ratio for the IRQC can be improved.245 This was shown by both experimental results and numerical modelling. Moreover, this experiment was carried out at room temperature. The authors used an LaCl3:8%Pr3+ crystal pumped by a focused dye laser beam (560 nm), while the IR signal was provided by a tungsten-halogen lamp. They demonstrated an enhancement in the detection of the IR signal corresponding to the three transitions, i.e., 3H4 → 3H6/3FJ/1G4, in the range of 1.0–2.3 μm for monitored PA emission. Below the critical threshold of pump power (30 mW) and with the addition of an 11 mW IR signal, an increase in the visible emission was observed with a signal-to-noise ratio greater than one. In contrast, for the same IR signal but for a pump power above the threshold, an increase in the visible emission was observed with a signal-to-noise ratio of less than unity. Moreover, the signal-to-noise ratio decreased to 0.5 at the threshold power of 30 mW when the IR signal was reduced to 5 mW.
Fig. 22b presents a schematic illustration of the IRQC based on Pr3+ ions. When the ion concentration is increased, cross-relaxation (CR) processes loop the energy between the excited and ground states of two neighbouring Pr3+ ions, leading to doubling of the population of intermediate excited level in every iteration of the loop and an increase in resonant photon absorption of pump photons, hνP. Consequently, a visible emission was observed from the 3P0 level, and the intensity thereof is dependent on the initial population of the 3H5 level. The population of the latter level depends on the flux of MIR photons, and thus the visible emission is nonlinearly dependent on the MIR light detected by this quantum converter.
Among the methods for detecting single photons in MIR, one can mention the technique based on a crystal detector utilizing the sum frequency generation phenomenon. Photons are upconverted to near infrared radiation, and then detected by avalanche photodiodes.246 Another method for MIR detection is the use of the nonlinear optics process of three-wave mixing, but this technique requires precise phase matching and high pulse powers, which are only achievable with femto- and picosecond lasers.246 InGaAs semiconductor materials provide technical feasibility to detect or image MIR photons, but these detectors are not very sensitive and are also very expensive.247 The mid-infrared light detection by optomechanical frequency upconversion is a novel advanced method and was first theoretically studied by P. Roelli et al.248 and further developed by W. Chen et al.242 and A. Xomalis et al.249 In this method the frequency upconversion was achieved by the utilization of the IR absorption and Raman activity of molecular vibrations. The coherent interaction of the electronic state of a molecule with its vibrational states through IR light, together with specially designed antennas and gold nanospheres confining the MIR energy to a narrow slot allowed the detection of the MIR signal with a power of less than a microwatt. However, despite their enormous potential, the presented devices showed limited stability due to the use of organic molecules. The lifetime of these device can exceed one month only if proper encapsulation that excludes the presence of oxygen is provided.243
The broad application field of MIR detectors is constantly motivating researchers to develop new sensitive, cost-effective detection schemes. IRQC based on the PA phenomenon, as detailed in this review, can bring new possibilities of measurements at room temperature with a reduced signal-to-noise ratio and, through ladder-like energy levels, prevent rapid saturation of the level from which emission occurs. The development of new PA materials based on various lanthanide ions and nanoscale host materials together with advancements in measuring systems may also open new possibilities of MIR detection at the nano-microscale in biology through optical microscopes that traditionally work in the visible spectral range.1,37,145,243,245,246,250
Extending the concept of photon counting with PA nanomaterials, a straightforward extrapolation is IR imaging instead of simple intensity measurements counting from one single point or total intensity emitted from the bulk volume of a sample. The involvement of PA as a workhorse enables the reaping of the essential benefits connected to it, such as great sensitivity and signal amplification accompanied with translation of the detected MIR photons into visible or NIR photons, which are more suitable to detect even with simple and cheap photodetectors. Two valuable architectures of the PA-based optically active element can be proposed. The first, utilizing a bulk monocrystal, acting as a screen or photographic film, inside which the image constructed by the incoming MIR photons is transferred into the visible projection, making it possible to map with optical microscopy techniques. An analogous operating blueprint can be also found for the second approach, where the active element is composed of a film of nanocrystals, and each of them can operate as an independent pixel (or voxel), providing better resolution due to the limited cross-talk (owing to energy migration) between neighbouring regions. Moreover, this design opens the possibility of sensitivity at multiple spectral bands owing to the use of mixtures of several ANP. Nevertheless, in these approaches for narrowband sensitivity, MIR detection is inherently connected with several technical difficulties. Firstly, a wide area of the avalanching material should be pumped homogeneously with a stable power density slightly below the threshold value. This can be challenging because tens or hundreds of kW cm−2 are required for the most well-known materials featuring efficient PA. Furthermore, the MIR spectral sensitivity can require the application of various PA ions or PA schemes, and thus different pumping wavelengths or thresholds may apply. These issues can be at least partially overcome, similarly as some challenges arising from the point of view of sub diffraction imaging, thermometry or FRET sensors, by the development of materials featuring efficient PA and characterized by a low power threshold value.
PA pumping of lasers has been observed since 1990 in Pr3+ (7%)-doped LaCl3 material,252 where 677 nm input photons were converted into a 644 nm laser output, giving an efficiency of 25%. Other early reports presented lasing in bulk materials and fibers46 doped with Nd3+ (ref. 32) and Tm3+ (ref. 253) and co-doped with combinations such as Pr3+/Yb3+.254 A broad range of visible emissions is covered by PA lasers, with most of the laser systems (and all the ones exhibiting efficiencies above 10%) produced based on Pr3+ or Pr3+/Yb3+ co-doped materials. However, the great majority of the known reports on this topic were published in the 90s and early 2000s, which had no chance to gain from the current knowledge and technologies developed and provided by the nanotechnology field in the last decade.
The use of upconverting nanomaterials as gain media for lasers has opened new applications for lasers, such as microscale lasers as ultrasharp emitters for bioimaging. Unlike bulk materials, UCNPs can be readily coated onto or assembled into resonators and cavities (including the whispering gallery modes),255–257 microcavities or plasmonic array nanocavities,258,259 Promising results have been reported using NaYF4 core–shell nanoparticles doped with Tm3+ ions, which are known for hosting energy looping77 and PA. These nonlinear processes promote population inversion through the avalanche-like building of the population of the 3H4 metastable state.95 These nanoparticles, assembled on 5 μm polystyrene microspheres, emit at the room temperature whispery gallery mode-enhanced continuous wave laser light at 800 nm and 450 nm.260 The avalanche-like pumping (with a slope of ∼7 or ∼4 for emissions at 800 nm and 450 nm, respectively, and the average threshold value of 44 ± 23 kW cm−2) was presented with a 1064 nm ESA-resonant excitation beam. Moreover, controlled assembling of the UCNP layer on the microcavities enabled the production of systems with a down-shifted threshold power density to 0.7–0.9 kW cm−2, with demonstrated potential for biosensing.261 Other types of lasing with avalanche-like characteristics (slope of 4.4 and threshold of emission intensity at 70 W cm−2) have been reported for NaYF4 core–shell nanoparticles, with their core co-doped with Er3+ and Yb3+ ions, assembled on an array of plasmonic Ag nanopillars.262 The studies of the highly nonlinear response of materials and possibility of laser emission in micro or nano laser cavities can not only provide insights into laser light and interaction at the molecule level but also for photonics devices or mechanisms on the nanoscale.
Although PA-based lasers preserve all the key advantages of conventional lanthanide ion-based upconverting lasers, they offer new possibilities. Firstly, they significantly extend the selection of available combinations of pumping wavelengths, which are not resonant with GSA transitions (like in most common pumping schemes) and have to be well fitted to the desired ESA transition. The resulting laser lines often provide emission and amplification in the visible range, similar to the case of Pr3+-doped systems. Also important, in conventional UC mechanisms, efficient energy CR is a parasitic effect hindering population inversion and attenuating the laser emission, and therefore higher dopant concentrations are neither recommended nor assure the optimal performance. However, an increase in the concentration of dopants in the case of the PA scheme is highly desirable because energy looping, which is governed by the CR mechanism, is responsible for fuelling the multiplication of ions in the intermediate metastable state. Not only a ca. 10-fold larger absorption cross section (for excitation powers exceeding the threshold value) can be obtained compared to laser materials operating in other schemes, but the brightness is enhanced as well with an increase in the number of emitting ions. The nature of the PA phenomenon exhibits a highly nonlinear relation between pumping power and resulting output efficiency. Consequently, PA lasers provide a high in-out efficiency performance; however, they require pumping power densities exceeding the threshold value for their optimal operation.251 Finally, the fact that the PA has become available at the nanoscale level opens new possibilities in the fields of nano-bio-technology, biology or environment sensing, nano-photonics, telecommunication and optical computing.251
| AF | Autofluorescence |
| ANPs | Photon avalanche nanoparticles |
| APD | Avalanche photodiodes |
| Bkg | Background signal (e.g. autofluorescence) |
| CET | Cooperative energy transfer |
| CL | Cooperative luminescence |
| CR | Cross relaxation |
| CS | Cooperative sensitization |
| CW | Continuous wave |
| C@S | General naming strategy for core@shell nanoparticles |
| EM | Energy migration |
| EMU | Energy migration-based energy transfer upconversion |
| ESA | Excited state absorption |
| ET | Energy transfer |
| ETU | Energy transfer upconversion or APTE addition de photon par transferts d’energie |
| FED | Fluorescence emission difference |
| FRET | Förster resonant energy transfer |
| GSA | Ground state absorption |
| IRQC | The infrared quantum counter |
| PMT | Photomultiplier |
| LIR | Luminescence intensity ratio |
| LN | Nonlinear luminescence |
| LnNP | Lanthanide doped nanoparticle |
| LT | Luminescence lifetime |
| MIR | Medium infrared |
| MPA | Migrating photon avalanche |
| MPR | Multiphoton relaxation |
| MRI | Magnetic resonance imaging |
| NGRER | Non-GSA-resonant ESA-resonant |
| NIR | Near infrared |
| NIRES | Near-infrared emission saturation |
| NR | Nonradiative relaxation |
| NSI | Nonlinear structured illumination microscopy |
| PA | Photon avalanche |
| PAA | Poly(acrylic acid) |
| PAET | Photon assisted energy transfer |
| PALM | Photoactivated localization microscopy |
| PANP or ANP | Photon avalanching nanoparticles, |
| PA-RET | Photon avalanche resonant energy transfer |
| PASSI | Photon avalanche single beam super resolution imaging |
| PAT | Photon avalanche thermometry |
| PEG | Poly(ethylene glycol) |
| PEI | Poly(ethyleneimine) |
| PVP | Polyvinylpyrrolidone |
| qCW | Quasi continuous wave |
| QD | Quantum dot |
| RS | Raster scanning |
| RE ions | Rare earth ions |
| RET | Resonant energy transfer |
| SAT | Saturation nanoscopy |
| SBR | Single band ratiometric |
| SFG | Sum frequency generation |
| SHG | Second harmonic generation, |
| SIM | Structured illumination microscopy |
| SPA | Sensitized photon avalanche |
| STED | Super-resolution stimulated emission depletion microscopy |
| STORM | Stochastic optical reconstruction microscopy |
| THG | Third harmonic generation |
| UC | Up-conversion |
| UCNP | Upconverting nanoparticle |
| UC RET | Upconversion RET |
| uSEE | Super-linear excitation–emission |
| WNR | Non-radiative rates |
| VIS | Visible radiation |
| ZBLAN | ZrF4–BaF2–LaF3–AlF3–NaF |
Footnote |
| † Electronic supplementary information (ESI) available: Overview of PA in various lanthanide ions, comparison table of PA In nano- and micromaterials, comparison table of sub-diffraction imaging with PA nanoparticles. See DOI: https://doi.org/10.1039/d4cs00177j |
| This journal is © The Royal Society of Chemistry 2025 |