Wei
Wang†
a,
Yu
Cheng†
b and
Xiaojiang
Xie
*a
aDepartment of Chemistry, The Hong Kong University of Science and Technology, Clear Water Bay, Kowloon, Hong Kong, China. E-mail: xiexj@ust.hk
bDepartment of Chemistry, Southern University of Science and Technology, Shenzhen, 510085, China
First published on 13th May 2025
Photochromic compounds, capable of reversibly switching between distinct molecular states upon light irradiation, have emerged as powerful tools for quantitative chemical analysis and sensing. This feature reviews recent advancements in this developing field, focusing on the design principles and applications of photoswitchable sensors. We begin with a concise overview of the fundamental photophysics and photochromism of key compound classes, and then discuss the mechanisms of analyte recognition and signal transduction, showcasing how light-induced isomerization modulates analyte binding and enhances signal contrast compared to conventional optical sensors. The unique sensitivity of the photoswitching process to the microenvironment is also explored. Finally, we outline future research directions and challenges for realizing the full potential of photochromic compounds in analytical chemistry related fields, including diagnostics, environmental monitoring, and materials science.
For instance, in optical data storage and switching, their light-induced changes in absorption characteristics enable the encoding of binary information.10 Photochromic materials are also crucial in the development of light-controlled photoswitching devices, which are essential for advancements in photonic circuits and quantum information processing.11,12 Additionally, these compounds are being explored in the design of molecular machines and nanorobotics systems, where light-driven isomerization can induce mechanical motion or alter material properties, with potential applications in drug delivery, environmental monitoring, and nanomanipulation.13–17 Furthermore, their capacity to modulate color or fluorescence in response to external stimuli, such as pH, temperature, or specific ions, makes them powerful tools for chemical sensing.18–20 Combining photochromic compounds with advanced optical techniques like fluorescence microscopy and spectroscopy further enhances their versatility in sensing applications.
This invited feature article focuses on recent advances in the application of photochromic compounds for quantitative chemical analysis and sensing. Other classical applications of photochromic compounds in color-changing materials, anti-counterfeiting technologies, and super-resolution microscopy have been discussed and highlighted elsewhere.21–24 We begin with an introduction to representative photochromic systems, explaining the fundamental principles of photoswitching from a thermodynamic perspective. Subsequently, we discuss emerging applications, focusing on the use of photochromic compounds for high-performance sensing and imaging of ions, molecules, and biomolecules, as well as their integration into more complex sensing platforms. Particular emphasis is placed on exploiting reaction dynamics to amplify detection signals and on developing multiplexed and high-contrast optical sensing and imaging strategies. Finally, we provide an outlook on the current state and future directions of photochromic compounds in quantitative chemical sensing and analysis.
Fig. 1 shows the structures and photoisomerization reactions of representative photochromic compounds. These compounds are crucial in fluorescence modulation, chemical sensing, and bioimaging, offering versatile, light-controllable properties for advanced materials and devices.
Boron difluoride (BF2) coordinated azobenzene derivatives represent a unique class of visible light responsive azo compounds. Through the formation of BF2 complex, the azo group undergoes strong electronic interactions with the boron center, significantly stabilizing the electronic states involved in the isomerization. This coordination leads to a red shift in the π–π* transition, making photoisomerization possible under visible light irradiation.30
The structurally similar dimethylcethrene is also an intriguing type of compounds, particularly due to the chiroptical properties. As a diradical, dimethylacetylene contains two unpaired electrons in an excited state, making it highly reactive. This unique reactivity is especially evident in its photoinduced form, where the excited-state electronic structure enables reversible switching between two distinct states: a more stable closed form and a more reactive open form.53
The structural arrangement is crucial in determining the photochromic behaviour of these dimers. The ability of the dimer to undergo reversible transformations depends on the rigidity of its structure, also the proximity and orientation of the imidazole units which influences the thermal stability of each isomer and the rate of back transformation.72,76 As such, the structural design of imidazole dimers, whether non-covalent or covalent, directly impacts their photochromic properties, determining critical factors such as response time, stability, and their potential applications in optical or sensing technologies.
![]() | ||
Fig. 2 Schematic representation of the potential energy surface during the isomerization process of photoswitchable compounds between the thermodynamically stable isomer and the light induced metastable isomer.58 |
A key concept is the conical intersection (CI) during the photoisomerization process.79 This is a region on the potential energy surface where two electronic states become degenerate, providing an efficient pathway for non-radiative transition between the excited state and the ground state of the new isomer. The molecule does not lose all its excitation energy through vibrational relaxation before reaching the CI. Instead, the transition through the CI is a non-adiabatic process that leads directly to the ground state of the isomer. Vibrational relaxation continues to occur after the transition to the new ground state, dissipating the remaining excess energy.80
The newly formed isomer may exhibit thermodynamic stability or exist in a metastable state relative to the original isomer. The difference in Gibbs free energy (ΔG) between the two isomers determines the relative stability and the equilibrium distribution in the dark (thermal back-reaction). The energy barrier for the thermal back-reaction (from the metastable isomer to the original one) determines the thermal stability of the metastable isomer, preventing immediate thermal equilibration.81 It is important to distinguish between the energy barrier for thermal isomerization in the ground state and the pathway through the CI in the excited state.
The photostationary state (PSS) describes the dynamic equilibrium achieved under continuous irradiation.82 At PSS, the rates of forward and reverse photoisomerization are equal, resulting in a specific ratio of the isomers. The PSS composition depends on the absorption spectra of all isomers and the wavelength of the light used for irradiation.
The quantum yield of photoisomerization is the number of isomerization events per photon absorbed. It is a measure of the efficiency of the photochemical process. Factors influencing the quantum yield include molecular structure, excitation wavelength, solvent, competing deactivation pathways like fluorescence, internal conversion to the ground state of the same isomer or intersystem crossing.83
In summary, photoswitching involves light absorption, excitation to a higher electronic state, isomerization through CI to a new ground state isomer, and establishment of a photostationary state. The thermodynamics of the process are governed by Gibbs free energy difference between the isomers and the energy barrier for back-isomerization, while the efficiency of the photochemical process is described by the quantum yield.84
Light intensity: the rate of photoisomerization is directly proportional to the intensity of incident light. A higher photon flux increases the number of excitation events per unit time, accelerating the reaction. This relationship is typically linear unless the intensity reaches levels where multi-photon absorption or other nonlinear effects occur.
Absorption cross-section: the efficiency of photon absorption depends on the chromophore's absorption cross-section at the excitation wavelength. A larger cross-section enhances the probability of photon capture, thereby increasing the reaction rate.
Quantum yield of isomerization: the quantum yield quantifies the efficiency of conversion from the excited state to the isomerized product. A higher quantum yield means a greater fraction of excited molecules undergo isomerization, leading to a faster overall reaction.
Excited-state dynamics: the isomerization rate is also influenced by the excited-state lifetime and the speed at which the molecule crosses the conical intersection. Faster transitions through the conical intersection promote quicker isomerization. These dynamics are affected by molecular structure, the surrounding environment (such as solvent or matrix), and temperature.
Assuming that photoisomerization follows a single-photon absorption process and that external factors such as competitive relaxation pathways and solvent effects are negligible, the rate of the forward photoisomerization reaction (kf) can be expressed as:85
kf = Φ × σ × I × [A] | (1) |
Thermal back-isomerization: this is the spontaneous return of the metastable isomer to the thermodynamically more stable form in the dark. The rate of this process is governed by:
Energy barrier (ΔG‡): the activation energy barrier for the thermal back-reaction is a critical factor in determining the thermal stability of the metastable isomer. A higher energy barrier results in a slower thermal back-reaction, thereby enhancing stability. This barrier is quantified by the Gibbs free energy of activation and the rate constant of the thermal-back reaction (kΔ) is described by the Eyring equation:87,88
![]() | (2) |
Temperature dependence: as indicated by the Eyring equation, the rate of thermal back-isomerization increases exponentially with temperature. This means that higher temperatures significantly accelerate the back-reaction, reducing the lifetime of the metastable isomer.89
Photo-induced back-isomerization: in some photochromic compounds which the newly formed isomer is relatively stable compared to the original isomer. As a result, the backward reaction cannot occur spontaneously and need to be induced by light irradiation of a specific wavelength. In this case, the kinetics are similar to the forward reaction, with the light intensity, absorption cross-section of the new isomer, and the quantum yield of the reverse isomerization determining the rate.90
For example, spiropyran serves as a versatile photo-switchable platform for sensing of metal ions.33,92Fig. 3a shows a spiropyran-based probe for reversible Zn2+ sensing in cells, offering improved signal-to-background ratios and faster photoswitching.93 The structure of this probe not only incorporates a bis(2-pyridylmethyl)amine substituent into the spiropyran scaffold to facilitate Zn binding but also introduces an aryl carboxylic acid unit to enhance hydrophilicity and cellular permeability. Consequently, this probe detects exogenous Zn2+ in HEK 293 and endothelial cells without affecting proliferation, and identifies endogenous Zn2+ efflux during apoptosis. Biocompatibility and photoswitching were demonstrated in endothelial cells, highlighting its potential for intracellular Zn2+ efflux studies.
![]() | ||
Fig. 3 Photoswitchable fluorescent materials for cations identification. (a) The 6′-fluoro spiropyran enables reversible Zn2+ sensing, faster photoswitching, biocompatibility, and detects Zn2+ efflux, supporting apoptosis studies in cells.93 (b) Photoswitchable spiropyran-based Ca2+ sensors detect nanoliter-level changes by transitioning between non-fluorescent and fluorescent states, effectively monitoring Ca2+ in HEK 293 cells.94 (c) A photoswitchable diarylethene with an azacrown receptor enhances fluorescence upon metal-ion binding, functioning as a chemosensor for metal-ion detection.95 (d) A hemithioindigo fluorescent probe selectively detects Cu2+ and facilitates its monitoring in environmental and biological systems.96 |
Additionally, crown ether structures feature cavities that correspond to the ionic radii of specific metal ions, enabling them to chelate metal ions through coordination and electrostatic interactions. Ca2+ plays a critical role in numerous cellular functions, including enzyme activity, cell attachment, migration, tissue morphology, metabolic regulation, signal transduction, replication, and the electrochemical processes of specialized cells such as muscle and nerve cells.97,98 As shown in Fig. 3b, Sabrina et al. reported a novel,94 rationally designed, and photoswitchable spiropyran-based Ca2+ sensors. They incorporated the ion-chelating domain (1-aza-18-crown-6) into the benzopyran ring, where Ca2+ forms a complex with the photoinduced merocyanine (MC) structure, resulting in enhanced absorbance and fluorescence. The probe achieved an excellent detection limit of 100 nM for Ca2+ sensing. In practical applications, the sensors were successfully used to monitor changes in Ca2+ levels in HEK 293 cells. Furthermore, the authors employed density functional theory (DFT) calculations to elucidate the photophysical processes involving Ca2+ ions and the photochromic compounds.
Besides spiropyran, Morimoto et al. reported a photoswitchable fluorescent diarylethene with an azacrown ether receptor (Fig. 3c), demonstrating metal-ion-gated enhancement of photoreactivity and fluorescence.95 Without metal ions, photoisomerization and fluorescence are suppressed by PET. Metal ion binding blocks PET, significantly enhancing photoreactivity and fluorescence. This molecule offers potential as a metal-ion chemosensor and probe for metal-ion-sensitive super-resolution fluorescence microscopy.
Visible light responsive systems such as hemithioindigo has also been exploited. In Fig. 3d, Xie et al. developed a hemithioindigo-type fluorescent probe capable of highly selective Cu2+ detection.96 Cu2+ is a vital element in cellular processes, making the monitoring of copper ions in biological systems critically important.99 This probe integrates hemithioindigo with 8-hydroxyquinoline, the latter serving as an excellent scaffold for recognizing and coordinating metal cations. Upon binding to Cu2+, the probe exhibits a significant red shift of 102 nm and ratio fluorescence changes. Under optimized conditions, it achieves an impressive detection limit of 0.02 μM. In practical applications, the Cu2+-selective nanoprobes is internalized by macrophage RAW 264.7 cells to track changes in Cu2+ levels under external stimulation. This probe holds promise for chemical sensing of Cu2+ in both environmental and biological samples.
The cyanide anion (CN−) is highly toxic to organisms, as it strongly binds to cytochrome C, disrupting the mitochondrial electron transport chain.100,101 This interference impairs oxidative metabolism and oxygen utilization, resulting in severe cellular dysfunction. As shown in Fig. 4a, Shiraishi et al. reported that a spiropyran derivative acts as a selective CN− receptor in water.102 UV-induced nucleophilic addition forms a detectable 1-CN− adduct (>1.7 μM), regenerable by visible light without additives.
![]() | ||
Fig. 4 Photoswitchable materials for anions identification. (a) A spiropyran derivative serves as a selective CN− receptor, forming a UV-detectable adduct that is regenerable by visible light.102 (b) A novel probe combining spiropyran and mesoporous silica nanoparticles adsorbs/desorbs F− reversibly under dark and visible light irradiation, respectively.103 (c) Novel fluorescent probes based on naphthalimide and spiropyran structures image SO2 in living cells, demonstrating potential for biological sensing.104 |
Fluorine pollution poses significant health risks, including neurodegenerative conditions such as Alzheimer's disease and thyroid disorders.105 In Fig. 4b, Guan et al. developed a novel fluorescent probe by combining carboxyl spiropyran (SP-COOH) with mesoporous amino silica (NH2–SiO2) nanoparticles through amidation.103 The mesoporous SiO2 particles serve as a solid surface to anchor the spiropyran, utilizing their unique characteristics: nanometer size, porous structure, rough surface, and amorphous nature. These properties allow the spiropyran to undergo a SP/MC photoisomerization in a conformationally free manner. Upon the addition of F− to the SP/MC mixed solution in the dark the positively charged nitrogen in the indolenine of MC form electrostatically absorbed F− from the solution. The formation of the MC–F− complex promoted the isomerization of SP to MC form, resulting in an increase in absorbance. Meanwhile, visible light induced the reverse reaction, leading to the desorption of F− and a subsequent decrease in absorbance. The innovative approach offers a promising method for reversibly detecting fluorine pollution in biological and environmental samples, potentially aiding in the prevention and monitoring of fluorine-related health issues.
Sulfur dioxide (SO2) plays a crucial role in physiological processes as a key gas signaling molecule with significant implications for human health. Endogenous SO2 in cells is primarily produced from thiol-containing amino acids through enzymatic catalysis and is subsequently converted into sulfite and bisulfite in body fluids.106 In Fig. 4c, Yong et al. developed a novel fluorescent probes based on naphthalimide and spiropyran structures denoted as HSP.104 Upon HSP binding to human serum albumin (HSA), HSP/HSA complex was formed which exhibits green fluorescence. Upon UV irradiation, the spiropyran moiety undergoes isomerization to generate the merocyanine form (HMr/HSA), displaying fluorescence in both the red and green channels. The introduction of SO2 conjugated addition reaction with α,β-unsaturated imine cation in the merocyanine structure, resulting in a decrease in red fluorescence. These fluorescent probes not only could simultaneously have identified HSA and SO2, but also successfully applied to image SO2 in living cells, demonstrating their potential for biological sensing and imaging applications.
![]() | ||
Fig. 5 Photoswitches for biologically relevant species. (a) A diarylethene-based probe, D-HBT-NBD, developed for photochromic and fluorescent detection of cysteine, homocysteine, and glutathione.107 (b) A β-galactosidase-responsive probe, NpG, forming NpG@HSA for fluorescence visualization and STORM imaging of cellular β-Gal activity detection.109 (c) NBD1, a photoswitchable norbornadiene-based probe, revealing amyloid beta plaque heterogeneity through fluorescence emission changes and reversible photoisomerization.110 (d) A photoswitchable excitonic switch developed using double-methylated 2’-deoxyuridine and tricyclic cytidine oligonucleotides.111 |
As shown in Fig. 5b, Li et al. designed a β-galactosidase (β-Gal)-responsive photochromic fluorescent probe, NpG, that forms a complex with human serum albumin (HSA), known as NpG@HSA.109 This complex exhibited enhanced fluorescence at 520 nm, allowing for visualization in aqueous media. NpG was synthesized by reacting a 2,3,3-trimethyl-3H-indole-functionalized naphthalimide with 5-nitrosalicylaldehyde, which decorated by an acetyl-protected β-galactose group. When introduced β-Gal, the galactose group was cleaved, leading to the phenolic hydroxyl oxygen released, producing NpM@HSA with red fluorescence at 620 nm. Additionally, the photoisomerization between spiropyran and merocyanine was activated, enabling STORM (stochastic optical reconstruction microscopy) through ON/OFF photo-blinking for precise nanoscale tracking of β-Gal in cells at super resolution level.
Amyloid beta (Aβ) plaques exhibit structural polymorphism, and their maturation from diffuse to compact forms may play a key role in Alzheimer's disease pathogenesis, as revealed by fluorescence imaging.112 As shown in Fig. 5c, Hanrieder et al. introduced NBD1, a photoswitchable fluorescent probe based on a norbornadiene derivative, for imaging Aβ plaques.110 NBD1 features turn-off photoswitching, and after incorporating long arms with a π-conjugated system, the NBD1 can interact with β-sheets of amyloid plaques through hydrophobic interactions. Additionally, the V-shaped overall molecular structure formed by the two long arms is a crucial feature for recognizing Aβ aggregates. In transgenic AD mouse models, it revealed broader emission at plaque peripheries and narrower emission in dense cores. The reversible photoisomerization to quadricyclane enhances its functionality, providing new insights into Aβ plaque heterogeneity and AD pathology.
In Fig. 5d, Jäschke et al. developed doubly methylated diarylethenes derived from 2′-deoxyuridine with exceptional thermal stability and fatigue resistance.111 An all-optical excitonic switch was designed using two oligonucleotides: one strand contained a fluorogenic double-methylated 2′-deoxyuridine as the fluorescence donor, while the other strand included tricyclic cytidine as the acceptor. These oligonucleotides formed a highly efficient photochromic FRET pair, exhibiting strong ON/OFF contrast and operating for 100 cycles without detectable fatigue in both liquid and solid phases.
In Fig. 6a, Xie at al. introduced a cellular temperature sensing method based on chemical kinetics and photoswitchable naphthopyran-mediated luminescence quenching.19 The rapid ring-closing thermal-back reaction rate of naphthopyrans is temperature dependent. The authors first synthesized a nanosensors composed of nonfluorescent naphthopyran molecules and a semiconducting luminescent polymer (F8BT). The emission intensity at 540 nm was continuously monitored over time under 470 nm excitation to evaluate its kinetic behaviour in Fig. 6b. Under UV irradiation, the ring-opening of naphthopyran leads to fluorescence quenching in F8BT, resulting in a rapid decrease in emission intensity. Once the UV light is removed, the ring-opened structure gradually reverts to the ring-closed form, allowing the fluorescence to recover. Fitting the recovery curves to decay function to generate an apparent time constant τF, which enabled temperature detection via a linear relationship between the logarithmic decay time constant (lnτ) and reciprocal temperature (T−1) through the variation of Arrhenius’ formula. To further enhance the nanosensors sensitivity and response range, structural modifications of naphthopyrans were investigated (Fig. 6c). This cost-effective approach enabled cellular temperature imaging and successfully tracked lysosomal heating induced by gold nanorods, highlighting potential applications in biological research.
![]() | ||
Fig. 6 Chemical kinetics of photoswitchable fluorescent materials. (a) Schematic illustration of cellular temperature sensing and imaging relying on the thermal relaxation of photoswitchable naphthopyran-nanosensors. (b) Normalized fluorescence intensity kinetics of the nanosensors at 30 °C upon 365 nm UV light irradiation. (c) Temperature calibration and influence of naphthopyran with different structures, represented with the reciprocal temperature (T−1) plotted against the natural logarithm of the fluorescence lifetime (ln![]() |
As shown in Fig. 6d, Xie et al. also presented an innovative chemical sensing approach with photoswitchable fluorescent hemithioindigo (HTI) for subcellular pH imaging.18 The fluorescent probes contained nitrogen-containing heterocycles, such as pyridine (denoted as HTI-Py) and ethylimidazole (denoted as HTI-EI). These heterocycles facilitate the formation of stable intramolecular hydrogen bonds following the photoinduced Z–E isomerization of the HTI molecule, thereby slowing down the thermal back-isomerization rate. Moreover, the higher the proton concentration, the slower the rate. The time constant of the thermal back-isomerization can be used to quantitatively assess the environmental pH. In addition, the pyridine-containing HTI exhibited lysosomal accumulation, and its kinetic fluorescence evolution during photoswitching was promising for the differentiation of subcellular compartments with varying pH levels. Notably, in HeLa cells, the fluorescence of HTI-Py was repeatedly switched on and off using a 488 nm laser (Fig. 6e).
The photochromic naphthopyrans were also utilized to develop a fluorescence encoding strategy a single-color channel.113 As shown in Fig. 7a, Xie et al. encapsulated naphthopyran and the fluorescent dye BODIPY in polystyrene microspheres. They encoded the microspheres based on the extent of fluorescence quenching, using the initial intensity (F0) and the fluorescence decrease (F1/F0). Fig. 7b illustrates the fluorescence quenching. Since the donor–acceptor ratio significantly affects the extent of fluorescence quenching, Fig. 7c displays the encoding of microspheres with nine different donor–acceptor ratios.
![]() | ||
Fig. 7 Multiplexed optical barcoding and sensing with photoswitchable naphthopyrans. (a) Schematic diagram of polystyrene microspheres containing naphthopyran and dye for two-dimensional encoding. (b) Fluorescent quenching between the ring-opened naphthopyran and dye. (c) Encoding by initial intensity (F0) and the fluorescence decrease (F1/F0) of microspheres C1–C9, composed of different donor–acceptor ratio. (d) Classification of microspheres based on fluorescence intensity ratios. (e) Fluorescence image of signal DNA bound to microspheres under 625 nm excitation. (f) Quantification of the four target DNAs in the assay (n = 10) based on fluorescence intensity measurements.113 |
This technique, demonstrated for COVID-19 DNA detection, is promising for multiplexed, high-throughput applications with simple encoding/decoding. As shown in Fig. 7d, pre-coded microspheres functionalized with DNA probes were uniformly mixed in a sample tube. Signal DNA and three different target DNAs were introduced into the tube, with the signal DNA labeled with Cy5 as a universal fluorescent marker. The post-reaction results, presented in Fig. 7e, show that the four codes can be easily distinguished. The variations in fluorescence intensity, depicted in Fig. 7f and g, correspond to the initial concentrations of the four DNA targets, with a detection limit reaching the nanomolar level.
Besides, complementary strategies have been devised to exploit the dynamic optical behaviour of photoswitches for advanced signal processing. Since photoswitch is capable of inducing fluorescence intensity change, it is quite obvious that a non-photoswitchable optical background could be removed during fluorescence detection to obtain better signal-to-noise ratio. This aspect was first demonstrated using nanospheres containing naphthopyran and a fluorescent donor (BODIPY) for DNA quantification in serum samples. The fluorescence of BOIDPY was effectively reduced when naphthopyran was switched to ring-opened forms under UV, and by comparing the fluorescence intensity before and after UV illumination, a better contrast was obtained.114
Recently, phase-sensitive detection mode was introduced using photoswitchable probes with naphthopyran and fluorescent donors.115 As shown in Fig. 8a, this approach leveraged reaction kinetics in the millisecond-to-second range, allowing frequency domain detection with inexpensive equipment. A phase shift (Δφ) in the fluorescence oscillation caused by modulated excitation was derived using fast Fourier transform, which was dependent on the donor-to-acceptor ratio. Δφ acted as a self-referencing signal for selectively highlighting fluorescent probes and providing dynamic readouts in chemical sensing.
![]() | ||
Fig. 8 Sensing and imaging of photoswitchable nanoprobes in phase-sensitive mode. (a) Schematic illustration of phase-sensitive photoswitchable probes (b) Phase-sensitive imaging of capillaries containing varying molar ratios of fluorescent donor molecules to naphthopyrans (ranging from 1![]() ![]() ![]() ![]() |
As shown in Fig. 8b, the Δφ of the nanoprobes gradually increased from 0 to π as the molar ratio of fluorescent molecules to nanoprobes increased. The spatial distribution of Δφ across all image pixels, as shown in Fig. 8c, was relatively narrow, which is promising for microscopic imaging and encoding. A preliminary application of the photoswitchable nanoprobes was demonstrated with protamine sensing. Protamine is a positively charged polycationic protein (Fig. 8d). It interacted with the fluorescent donor in the photoswitchable nanoprobes due to electrostatic attraction on the particle surface. This interaction in turn changed the average distance between the fluorescent donor and the photoswitchable naphthopyran, ultimately resulting in a change in Δφ (Fig. 8e).
A key distinction between photoswitchable optical sensors and conventional optical probes lies in their mechanism of interaction with analytes and their response to environmental stimuli. Conventional probes typically exist in a single state that interacts with the target analyte, and quantification relies on the thermodynamic equilibrium of this interaction, reflecting the binding affinity. In contrast, photoswitchable sensors possess multiple distinct molecular configurations (isomers) that can exhibit different affinities for the analyte. This multi-state behavior allows for enhanced signal contrast and controlled interaction; as molecular structures are modulated by light-induced isomerization.
Furthermore, the photoswitching process itself, encompassing both forward (light-driven) and backward (thermal or light-driven) reactions, providing a unique sensitivity to the probe's microenvironment. The dynamic isomerization process allows such probes to explore a range of configurations and interactions with its surroundings, offering a rich platform for sensing subtle environmental changes. This dynamic control also confers a “smart” character to photoswitchable sensors; they can be actively “switched” between ON–OFF states using light, unlike conventional probes that passively respond to the presence of an analyte. This active control enables more sophisticated sensing strategies, such as background reduction, ratiometric measurements, phase-sensitive detection, and spatiotemporal control, opening up possibilities beyond the capabilities of traditional optical sensors.
Despite the significant potential of photochromic compounds for quantitative chemical sensing, several challenges remain. Enhancing the quantum yield of photoisomerization is crucial, as low quantum yields in some compounds limit their sensitivity and overall performance. Selectivity is another key area for improvement; achieving specific responses to target analytes while minimizing interference from other chemical species requires the development of carefully designed, functionalized photochromic compounds. Long-term stability under repeated light exposure is also a concern, as some photochromic compounds can degrade over time. Improving their photostability is essential for practical applications and commercial viability. Embedding the compounds in polymeric matrix might be advantageous. Finally, integration with detection systems is vital for broader accessibility and widespread use. Developing strategies to seamlessly integrate photoswitchable sensors with advanced detection platforms, such as portable devices or lab-on-chip systems, will translate their unique advantages into practical, field-deployable sensing technologies.
Footnote |
† W. W. and Y. C. contributed equally to this work. |
This journal is © The Royal Society of Chemistry 2025 |